You are on page 1of 15

TECTONICS, VOL. 31, TC4002, doi:10.

1029/2011TC003068, 2012

Kinematics of Tertiary to Quaternary intracontinental


deformation of upper crust in the Eastern Cordillera,
southern Central Andes, NW Argentina
Tasca Santimano1,2 and Ulrich Riller1
Received 20 November 2011; revised 16 April 2012; accepted 22 May 2012; published 4 July 2012.
[1] To elucidate factors controlling the geometry of, and kinematics associated with,
prominent upper-crustal structures in the Eastern Cordillera of NW Argentina, structural
analyses of key areas were complemented by fault slip analysis, remote sensing and 3D
representation of prominent faults. The analyses revealed that deformation during Tertiary
to Quaternary times was accomplished mostly by prominent orogen-parallel reverse faults,
the geometry of which was significantly influenced by Paleozoic and Cretaceous planar
structures. Locally, this deformation was preceded by kilometer-scale doming of
upper-crustal rocks. Analysis of 767 brittle faults at 67 stations in the studied areas
indicates that local upper-crustal doming and orogen-parallel reverse faults formed chiefly
under NW-SE and E-W shortening. Shortening directions inferred from fault slip data
portray the kinematics of first-order faults and folds. More specifically, shortening
directions are mostly uniform with respect to first-order structures. Age and kinematics of
deformation inferred from brittle fault analysis in the study areas is consistent with
equivalent data compiled from other parts of the Eastern Cordillera, Puna and Pampean
Ranges. Collectively, the data is at variance with hypotheses relating late Tertiary to
Quaternary deformation in the Eastern Cordillera to plate (boundary) kinematics.
The kinematics of intracontinental deformation in the southern Central Andes points to
deformation partitioning of upper crust as a consequence of bulk E-W shortening.
We, therefore, caution the usage of brittle fault-kinematics in continental interiors as
an indication for plate-kinematic changes.
Citation: Santimano, T., and U. Riller (2012), Kinematics of Tertiary to Quaternary intracontinental deformation of upper crust
in the Eastern Cordillera, southern Central Andes, NW Argentina, Tectonics, 31, TC4002, doi:10.1029/2011TC003068.

1. Introduction has been attributed chiefly to subduction zone mechanical


properties including changes in plate motion and plate con-
[2] The Puna Plateau in the southern Central Andes
vergence [Jordan et al., 1983; Marrett et al., 1994; Marrett
(Figure 1a) is the second highest continental plateau on Earth and Strecker, 2000; Oncken et al., 2006], crustal thickness
and formed by E-W lithospheric shortening since Eocene-
variation [Allmendinger, 1986; Allmendinger et al., 1989],
Oligocene times [Isacks, 1988; Allmendinger et al., 1997;
delamination of lithospheric mantle [Schoenbohm and
Barnes and Ehlers, 2009]. The structural evolution of this
Strecker, 2009; Montero Lopez et al., 2010] and inherited
plateau in a non-collisional tectonic setting, i.e., eastward
structural anisotropy [Grier et al., 1991; Carrera et al., 2006;
subduction of the Nazca Plate under the South American
Hongn et al., 2007]. It is not resolved to what extent spatial
Plate, is not well understood. Notably, the geometry and
and temporal variations in the kinematics of upper-crustal
kinematic history of first-order dip-slip faults in the eastern
deformation in the Central Andes relate to plate (boundary)
Puna and the adjacent Eastern Cordillera (Figure 1a), located
kinematics.
hundreds of kilometers east of the plate boundary, are still
[3] The Eastern Cordillera is characterized by kilometer-
uncertain. Deformation in both morphotectonic regions scale, basement-involved folds, thrusts and reverse faults
[Grier et al., 1991; Kley, 1996; Kley et al., 1996; Cristallini
1
School of Geography and Earth Sciences, McMaster University, et al., 1997]. Understanding the controlling factors for the
Hamilton, Ontario, Canada. geometry of prominent structures and associated kinematics
2
Now at Helmholtz-Zentrum Potsdam, Deutsches GeoForschungsZentrum is important for unravelling the mechanics of continental
Potsdam, Potsdam, Germany. plateau formation and intracontinental deformational response
Corresponding author: U. Riller, School of Geography and Earth to plate (boundary) kinematics. However, the geometry of
Sciences, McMaster University, 1280 Main St. W., Hamilton, first-order Tertiary and younger fault zones and associated
ON L8S 4K1, Canada. (rilleru@mcmaster.ca) kinematics of deformation in these morphotectonic regions is
2012. American Geophysical Union. All Rights Reserved. still not well known. By contrast, the kinematics of small-scale
0278-7407/12/2011TC003068 brittle shear faults has been the subject of many structural

TC4002 1 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Figure 1. Tectonic setting of the Central Andes. (a) Major morphotectonic units superimposed on digital
elevation model. AP: Altiplano, EC: Eastern Cordillera, P: Puna, PC: Precordillera, PR: Pampean Ranges,
SB: Santa Barbara fold belt. (b) First-order reverse faults and location of study areas, La Poma, Cachi and
Luracatao Valley, superimposed on digital elevation model of the eastern Puna Plateau and the Eastern Cor-
dillera, Province of Salta. Red stippled line delineates the boundary between the Puna Plateau and the East-
ern Cordillera based on the 3000 m elevation contour [Gephart, 1994]. COT: Calama-Olacapato-El Toro
Fault, QT: Quebrada del Toro.

studies in the southern Central Andes [Allmendinger, 1986; regions formed as a consequence of intracontinental short-
Allmendinger et al., 1989; de Urreiztieta et al., 1996; Gapais ening of the western South American Plate margin, imparted
et al., 2000; Cladouhos et al., 1994; Marrett et al., 1994; by convergence between the Nazca Plate and the South
Marrett and Strecker, 2000; Coutand et al., 2001; Maffione American Plate [Isacks, 1988]. Deformation in the regions
et al., 2009]. Nonetheless, only few studies, such as the one commenced in Eocene times and propagated mostly east-
by Marrett and Strecker [2000] for the Quebrada del Toro ward [Mon and Salfity, 1995; DeCelles and Horton, 2003;
(Figure 1b), systematically explored the orientation of short- Deeken et al., 2006; Hongn et al., 2007]. Shortening was
ening directions along individual prominent fault zones. It, accomplished by distributed deformation in the Puna,
thus, remains to be elucidated to what extent shortening whereas localized basement-involved folding and thrusting
directions vary on a given fault zone and whether shortening is more typical for the Eastern Cordillera in NW-Argentina
directions portray plate kinematic changes or local kinematics [Grier et al., 1991; Cristallini et al., 1997]. Horizontal
of lower-order faults. shortening in both regions led to formation of N-S trending
[4] To better understand the factors controlling the compressional basins [Jordan and Alonso, 1987; Kraemer
geometry and kinematics of prominent upper-crustal dip-slip et al., 1999; Hongn et al., 2007] and morphological ranges
faults as well as the variation in shortening directions on the (Figure 1b). Uplift of the ranges bordering the basins was
kilometer-scale in the Eastern Cordillera, we conducted accomplished on N-S striking, mostly sinistral, oblique
comprehensive structural analyses in the La Poma, Lur- reverse faults and NE-striking dextral oblique reverse faults
acatao Valley and Cachi areas (Figure 1b). These areas are [Riller and Oncken, 2003; Maffione et al., 2009]. Shortening
characterized by excellent exposure, known stratigraphy of also affected the basins, the strata of which were folded
Cretaceous to Neogene rocks (Figure 2) and, most impor- during Pliocene times [Marrett et al., 1994].
tantly, are structurally representative of the deformation [6] Orogen-parallel extension of upper crust on the
style in the Eastern Cordillera. The configuration of first- Puna Plateau occurred at an advanced stage of bulk E-W
order dip-slip faults is identified through field mapping, shortening, i.e., at about 10 Ma [Riller et al., 2001]. Promi-
remote sensing and 3D representation. This is com- nent NW-striking fault zones associated with Neogene
plemented by detailed kinematic analyses of small-scale, magmatic centers [Viramonte et al., 1984] are vestiges of
brittle shear faults (767 faults at 67 stations) to elucidate this deformation regime. The fault zones, notably the
principal strain axes and, by inference, the kinematics of Calama-Olacapato-El Toro Fault (Figure 1b), formed by
first-order structures. sinistral transtension [Riller et al., 2001; Petrinovic et al.,
2005, 2006; Ramelow et al., 2006; Acocella et al., 2011].
2. Geological Setting In the southern Puna, late Miocene horizontal extension of
the upper crust has been attributed to local gravitational
[5] The Eastern Cordillera is located between the Puna
instability of the plateau [Schoenbohm and Strecker, 2009;
Plateau to the west and the Santa Barbara foreland fold-and-
Montero Lopez et al., 2010] possibly caused by delamination
thrust belt to the east (Figure 1a). These morphotectonic

2 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Horizontal extension has been documented mainly from the


southern Puna and the Eastern Cordillera [Allmendinger,
1986; Allmendinger et al., 1989; Mercier et al., 1992;
Cladouhos et al., 1994] and attributed to spreading of
topography that lead to a decrease in crustal thickness
[Allmendinger et al., 1989; Marrett et al., 1994].
[8] Basement rocks in the Eastern Cordillera consist of the
low-grade metamorphic, late Neoproterozoic to early Cam-
brian Puncoviscana Formation [Turner, 1960; Mon and
Salfity, 1995], metagranitoid rocks of the Oire Complex
[Blasco et al., 1995] and syntectonic Ordovician granitoid
plutons [Hongn and Riller, 2007]. Collectively, these rocks
are overlain unconformably by Cretaceous to Eocene sedi-
mentary rocks of the Salta Group (Figure 2), which is made
up of the Pirgua, Balbuena and the Santa Barbara Subgroups
[Moreno, 1970; Salfity and Marquillas, 1994]. The Salta
Group formed during a phase of continental rifting prior to
bulk E-W shortening [Marquillas et al., 2005]. The Pirgua
Formation consists of coarse-grained syn-rift deposit at the
base to well-sorted conglomerate and sandstone toward the
top. The Balbuena Group marks the post-rift phase and
includes conglomerate and coarse-grained sandstone of the
Lecho Formation and overlying dolomite of the Yacoraite
Formation [Blasco et al., 1995]. The Santa Barbara Sub-
group is made up of sandstone and conglomerate of the
Mealla, Maiz-Gordo and Lumbrera Formations and is
overlain unconformably by the Quebrada de los Colorados
Formation [Daz et al., 1987] of the Eocene to Miocene
Payogastilla Subgroup [Hongn et al., 2007]. Fluvial and
alluvial deposits of Quaternary age overlie in turn uncon-
formably deformed Cretaceous and Tertiary strata [Hongn
and Seggiaro, 2001]. These strata are mostly preserved in
the valleys, whereas Paleozoic basement rocks that are
emplaced over these strata make up the ranges. As a con-
sequence, Quaternary pediments in the Calchaqu and Lur-
acatao Valleys that are part of the Oran Group [Russo and
Serraiotto, 1979; Carrera et al., 2006] consist mainly of
clasts derived from the Puncoviscana Formation and Ordo-
vician granitoid rocks.
Figure 2. Simplified stratigraphy of the Calchaqu and
Luracatao Valleys modified after Bosio et al. [2009]. 3. Methods
[9] Orientations of faults, foliation surfaces and aplitic dike
of the lithospheric mantle [Kay et al., 1994; Risse et al., margins were measured in Paleozoic granitoid rocks. In
2008; Kay and Coira, 2009]. Quaternary and Tertiary conglomerate, oblate clasts are
[7] Fault-kinematic studies conducted in the eastern Puna aligned concordant to compositional layering. Collectively,
and Eastern Cordillera indicate that bulk E-W shortening these characteristics revealed the orientation of bedding
was accomplished by local WNW-ESE to NW-SE shorten- planes, which were used to construct fold-axial planes and,
ing on orogen-parallel thrust faults in Miocene to Pliocene thus, directions of local shortening. In inaccessible areas, the
times and WSW-ENE shortening on NE-striking Quaternary first-order dip-slip faults were traced on air photos, Google
strike-slip faults [Allmendinger, 1986; Allmendinger et al., Earth images, Advanced Spaceborne Thermal Emission and
1989; Marrett et al., 1994; Cladouhos et al., 1994; Reflection scenes and Shuttle Radar Topography Mission
Coutand et al., 2001]. Based on fault kinematics, a possible data, from which digital elevation models with a resolution of
change in shortening directions during the Pliocene and 90 m were obtained. Fault orientations were also obtained by
Pleistocene has been attributed to a change in plate conver- applying the three-point method using the software GeoTrig
gence rate between the South American Plate and the Nazca (Rockware). The geometry of prominent dip-slip faults was
Plate [Marrett and Strecker, 2000]. However, it has not been modeled and extrapolated to a depth of 3 km using the soft-
systematically investigated to what extent shortening direc- ware Geomodeller v.1.3.1 (Intrepid Geophysics).
tions are controlled by the kinematics of first-order faults [10] Acquisition of fault slip data consisted of recording
and the presence of pre-existing structural anisotropies, the orientations of fault planes, striations (slicken-lines), slip
notably reactivated Cretaceous normal faults [Grier et al., senses and quality of kinematic information for each fault. A
1991; Hongn and Seggiaro, 2001; Carrera et al., 2006]. total of 767 faults at 67 stations were measured (Table 1).

3 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Table 1. Paleo-Stress and Strain Parameters Inferred From Different Techniques of Fault Slip Analyses From the Calchaqu Valley and
the Luracatao Valleya
PBT NDA Inverse
Strain Stress Method
Station n P B T l1 l2 l3 Ratio NEV s1 s2 s3 Ratio NEV Used
Calchaqu Fault
1 8 184/03 272/44 088/50 313/26 214/18 093/58 0.89 1 191/03 037/86 281/02 0.59 2 NDA
2 11 322/25 197/53 076/34 314/33 204/28 083/44 0.77 2 353/31 183/59 085/05 0.59 3 NDA
3 11 261/24 228/39 158/08 259/41 073/48 166/03 0.48 1 134/62 300/27 033/06 0.59 4 NDA
4 12 179/47 024/49 300/01 193/50 028/39 292/08 0.43 2 181/28 031/59 278/13 0.38 4 NDA
5 4 124/81 344/08 258/12 095/79 354/02 264/11 0.69 0 244/29 114/48 350/26 0.55 2 NDA
6 18 323/41 082/07 156/64 295/24 034/19 157/58 0.34 6 241/82 086/07 356/03 0.89 9 NDA
7 12 273/04 176/07 029/83 121/11 031/01 293/79 0.48 4 153/52 003/34 263/15 0.94 4 PBT
8 5 147/05 242/43 042/43 144/07 242/44 047/45 0.49 0 318/00 228/04 052/86 0.20 0 NDA
9 4 296/30 165/35 048/36 306/19 198/43 053/41 0.42 2 262/12 167/21 021/66 0.29 2 NDA

Toro Muerto Fault


10 14 256/34 151/27 065/40 257/37 163/07 064/52 0.40 5 254/22 155/23 023/58 0.05 7 NDA
11 8 067/27 183/38 323/33 066/25 203/57 327/20 0.60 1 091/22 182/04 280/67 0.23 3 NDA
12 10 284/06 180/42 002/57 110/06 205/38 012/51 0.65 7 183/30 032/57 281/13 0.97 1 NDA

Cerro Bajo Dome


13 48 089/11 178/01 273/78 089/12 179/01 273/78 0.56 5 079/17 170/03 268/73 0.57 4 Inverse
14 23 106/09 015/01 276/81 107/08 017/01 280/82 0.52 0 207/00 297/33 117/57 0.76 2 NDA
15 22 111/16 020/00 288/74 112/13 022/02 283/77 0.51 4 028/01 118/00 226/89 0.84 4 NDA
16 17 116/13 026/03 284/77 115/13 025/02 288/77 0.50 0 116/02 206/01 321/88 0.22 0 NDA
17 17 078/49 340/03 249/57 061/37 158/10 260/52 0.66 4 357/11 262/23 111/64 0.50 9 NDA
18 16 085/02 174/11 032/82 105/02 195/04 345/85 0.70 4 069/03 160/04 309/85 0.59 5 NDA
19 14 105/10 015/04 265/80 105/10 015/03 266/79 0.50 0 201/02 294/49 109/41 0.99 5 NDA
20 20 238/01 329/04 063/83 Cannot be calculated 071/01 341/02 181/88 0.94 0 PBT
21 6 328/08 149/18 291/75 059/06 150/09 295/79 0.81 2 171/10 262/07 026/78 0.23 1 NDA
22 5 080/32 183/17 300/42 071/44 199/32 309/28 0.83 0 055/33 256/55 152/10 0.26 1 PBT
23 5 108/01 018/03 213/88 108/01 018/02 215/87 0.50 0 294/03 204/03 074/86 0.23 0 NDA
24 4 266/02 175/04 021/85 267/02 177/06 018/84 0.48 0 079/29 178/15 292/57 0.33 2 NDA

Refugio Fault
25 22 289/39 127/40 249/69 294/26 120/64 025/03 0.18 9 077/28 175/15 290/58 0.94 7 NDA
26 6 085/66 169/03 250/35 144/29 010/51 248/24 0.88 3 356/14 226/69 090/15 0.65 1 NDA
27 9 059/09 151/13 279/78 065/09 155/04 271/80 0.41 1 283/01 192/46 014/44 0.38 3 NDA
28 5 170/43 297/15 043/34 159/21 267/40 048/43 0.39 2 307/11 045/36 203/52 0.85 2 NDA
29 5 016/27 296/28 110/47 358/31 245/33 120/41 0.57 1 153/29 277/46 044/30 0.87 2 NDA
30 13 064/06 298/31 221/03 286/16 019/10 139/71 0.82 3 031/02 300/10 132/80 0.77 4 Inverse
31 5 327/07 220/55 060/32 332/03 238/52 065/38 0.37 2 324/08 216/64 058/24 0.02 2 NDA
32 8 130/10 185/41 280/60 141/10 232/11 014/75 0.08 2 161/63 039/15 302/22 0.81 3 NDA
33 14 256/02 334/37 164/52 261/08 354/21 151/67 0.19 1 084/08 348/37 184/52 0.17 5 PBT
34 8 222/43 146/22 033/39 208/62 319/11 054/26 0.34 3 027/76 166/10 257/09 0.82 3 NDA
35 8 319/34 220/16 103/41 206/59 347/25 086/17 0.93 1 160/04 252/21 058/69 0.80 3 PBT
36 6 186/12 136/41 268/25 019/22 139/51 275/31 0.86 2 067/84 178/02 269/06 0.68 2 NDA
36b 9 289/01 195/20 032/70 292/02 201/18 028/71 0.41 0 098/02 188/16 359/74 0.21 0 NDA
37 7 083/72 314/46 201/19 333/53 163/37 069/05 0.47 2 206/13 301/19 084/67 0.86 3 NDA
38 11 007/23 238/04 151/74 014/19 122/41 265/43 0.35 2 184/15 276/08 033/73 0.54 4 NDA
39 5 313/23 213/34 075/57 312/20 214/20 083/62 0.67 1 347/14 083/23 228/63 0.18 2 NDA
40 5 046/05 125/02 191/83 260/00 350/08 166/82 0.72 1 187/28 296/31 063/46 0.45 2 PBT
41 6 090/16 218/70 003/29 106/19 216/45 359/38 0.42 2 079/10 331/61 174/27 0.24 3 NDA
42 7 060/15 331/31 153/67 054/14 312/18 191/71 0.52 1 208/13 300/09 063/74 0.38 1 NDA
43 14 026/24 267/22 153/35 035/23 290/30 156/50 0.40 6 233/01 142/22 325/68 0.68 4 PBT
44 18 316/02 176/49 032/30 133/07 233/56 039/33 0.27 3 343/05 251/22 086/68 0.56 6 NDA
45 6 186/47 314/37 069/17 213/75 001/13 092/08 0.71 1 184/53 347/36 083/08 0.15 1 NDA
46 5 283/06 179/66 016/24 283/05 181/62 016/26 0.52 0 196/63 335/19 070/14 0.94 1 NDA
46b 7 019/02 340/46 108/10 019/14 115/22 259/63 0.35 0 032/08 294/45 129/44 0.43 2 NDA
47 12 224/22 294/05 351/82 262/05 170/24 003/65 0.52 2 178/12 081/28 288/59 0.12 4 NDA
48 15 319/22 042/28 171/46 108/04 016/28 206/62 0.63 4 334/07 065/10 209/78 0.46 6 NDA
49 6 084/13 342/49 186/39 083/14 337/50 184/37 0.41 0 227/17 325/25 106/59 0.46 2 PBT
50 7 154/09 292/67 197/24 329/04 062/45 235/45 0.47 3 078/35 279/53 175/10 0.80 2 NDA
51 8 057/05 327/29 166/80 054/06 145/05 277/82 0.75 1 105/00 195/71 014/19 0.68 3 NDA
52 6 280/15 187/21 063/61 269/21 177/07 069/68 0.53 0 285/23 038/43 176/38 0.43 0 NDA

Cachi Fault
53 7 200/61 160/06 342/66 200/40 090/22 338/42 0.65 1 053/11 322/03 218/78 0.90 0 Inverse
54 23 242/34 040/37 134/35 214/14 323/52 114/34 0.63 9 193/09 097/33 297/56 0.51 5 NDA
55 16 256/22 157/23 125/75 286/38 020/04 115/52 0.66 7 033/73 267/10 174/14 0.96 8 NDA
56 15 263/31 165/16 052/54 251/32 156/08 054/56 0.68 3 075/09 167/13 312/74 0.75 4 NDA
57 19 258/26 000/19 129/60 261/29 002/19 121/54 0.79 4 009/16 100/03 201/74 0.39 5 NDA

4 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Table 1. (continued)
PBT NDA Inverse
Strain Stress Method
Station n P B T l1 l2 l3 Ratio NEV s1 s2 s3 Ratio NEV Used
58 12 281/02 142/48 011/43 109/08 200/05 324/81 0.53 2 181/03 271/02 039/86 0.25 2 NDA
59 7 330/18 116/71 232/10 335/65 137/24 230/07 0.80 1 342/41 208/39 096/25 0.35 2 NDA
60 14 267/06 357/32 167/60 272/07 005/24 167/65 0.48 2 003/14 270/08 152/73 0.38 2 NDA
61 13 241/43 081/54 340/01 293/50 068/31 173/23 0.33 6 351/08 087/39 252/50 0.94 5 NDA
62 25 113/01 007/52 211/33 304/04 038/43 210/47 0.33 10 095/01 002/64 185/26 0.20 9 PBT
63 4 293/32 209/01 119/55 288/27 019/01 111/63 0.44 1 285/44 167/26 058/34 0.08 2 NDA
64 6 302/09 107/34 234/55 343/12 077/19 222/67 0.63 1 091/06 269/84 001/00 0.73 3 NDA
65 11 250/38 071/65 356/07 256/47 097/41 357/10 0.38 3 190/31 035/56 288/11 0.56 4 NDA

Luracatao Fault
66 13 110/19 360/10 283/52 102/23 199/15 319/62 0.58 3 008/00 277/18 099/72 0.73 5 NDA
67 5 006/08 189/56 081/01 003/05 270/31 102/59 0.26 1 233/02 140/54 325/36 0.79 2 NDA
a
The P-T method (PBT), numerical dynamic analysis (NDA) and direct inversion method (Inverse) were applied for all faults (n) at a given station. The
choice of method used (Method Used) and individual parameters are explained in the text.

Fault slip data was processed using the software package dynamic analysis, whereby the P-T method resulted in simi-
Tectonics FP v. 1.6 [Ortner et al., 2002]. The quality of the lar solutions of principal axes orientations for some fault
fault data was tested using the CHECK algorithm of this populations (Table 1). We, therefore, interpret the axes in
software package. This algorithm ensures that for each fault terms of (infinitesimal) strain.
plane, the respective lineation lies on the fault plane. Fault
planes with lineations deviating more than 10 from the 4. Structure of the La Poma Area
respective fault plane were disregarded for further analysis.
Only corrected fault sets were used to calculate the orienta- [12] The Calchaqu Valley is bound by the N-S striking
tion of principal axes using methods that are based on the Toro Muerto and Calchaqu Faults [Marrett et al., 1994].
Mohr-Coulomb criterion for brittle fracturing. These faults juxtapose rocks of the Puncoviscana Formation
[11] For each fault population per station, principal axis over Cretaceous-Tertiary strata of the Salta Group (Figures 2
orientations were calculated using the P-T method [Turner, and 3). In the La Poma area, Cretaceous and Tertiary strata
1953], the direct inversion method [Angelier and Goguel, form a kilometer-scale, slightly asymmetric structural dome,
1979] and the numerical dynamic analysis [Spang, 1972]. known as the Cerro Bayo Dome [Hongn et al., 2007] that is
The P-T method calculates the P-axis (shortening axis), B-axis cored by Puncoviscana Formation (Figure 3b). The spatial
(intermediate principal strain axis) and T-axis (extension extent of the Cerro Bayo Dome is well defined by the con-
axis) for each fault plane and produces respective mean axes centric strike of inclined strata of the Lecho, Yacoraite, Mealla
for a given fault set. The direct inversion method computes and Quebrada de Los Colorados Formations (Figure 3a). The
the rudimentary paleostress tensor (s1 s2 s3) and stress Dome is asymmetric in map view (Figure 3a) and profile plane
ratio Rstress = (s2 s3)/(s1 s3). The numerical dynamic (Figure 3b). Specifically, it is elongate in the N-S direction and
analysis provides the reciprocal strain tensor (l1 l2 l3) characterized by a curved hinge zone [Hongn et al., 2007].
and strain ratio, Rstrain = (l2 l3)/(l1 l3). For all analyses, Sedimentary strata of the east limb of the Dome dip moder-
the angle (q) between the maximum principal axis and a ately eastward and display decameter-scale, asymmetric folds
given fault plane was assumed to be 30 , which is a reason- with east-dipping fold-axial planes. The strata shallow with
able value for the angle of internal friction of newly formed distance from the hinge zone of the Dome and are locally
faults [Byerlee, 1968] and average angle for fault sets con- overturned in the footwall of the Calchaqu Fault. The western
sisting of pre-existing faults [Sperner and Zweigel, 2010]. and eastern flanks of the Dome, the lateral extents of which
We confirmed the validity of this angle for our data sets by can be estimated by extrapolating the concentric trend of
comparing the pattern and kinematics of representative fault bedding plane trajectories, are obscured by rocks of the Pun-
sets with solutions of principal axes orientations arising from coviscana Formation that were displaced on the Toro Muerto
different angles of internal friction. The direct inversion Fault and Calchaqu Fault, on the west and east respectively
method and numerical dynamic analysis provide the number (Figure 3a).
of fault planes, the measured slip senses of which are oppo- [13] The southwestern portion of the Cerro Bayo Dome is
site to the ones predicted by a particular solution indicated as cut by the Cerro Bayo Fault, a steeply east-dipping reverse
negative expected value NEV (Table 1). The method cho- fault that juxtaposes the Yacoraite and underlying Lecho
sen for a given fault population was based on the NEV and Formations in the hanging wall against locally steeply west-
the best fit of principal axes orientations with respect to fault dipping strata of the Quebrada de los Colorados Formation
orientation and slip sense (see Figure S1 in the auxiliary in the footwall (Figures 2 and 3a). The Cerro Bayo Fault is
material).1 Our results suggest that principal axes of most regarded as a splay of the Calchaqu Fault [Hongn et al.,
brittle fault populations are best represented by the numeric 2007] and is cut by the Toro Muerto Fault (Figure 3a). The
Calchaqu Fault branches into two fault splays, one of which
1 is spatially associated with the Pleistocene basaltic to
Auxiliary materials are available in the HTML. doi:10.1029/
2011TC003068. andesitic Los Gemelos Volcanoes east of the central Cerro

5 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Figure 3. Structure of the La Poma area. (a) Geological units and structure of the La Poma area draped
over an airphoto. Stippled gray semi-circles delineate portion of the Cerro Bayo Dome overridden by the
Puncoviscana Formation on the Toro Muerto Fault, white arrows pointing at each other indicate shorten-
ing directions inferred from brittle fault kinematics. Line A-A indicates trace of cross section in Figure 3b.
(b) Cross-section across the Cerro Bayo Dome along line A-A in Figure 3a.

Bayo Dome [Guzmn et al., 2006]. The spatial affinity of the Cerro Bayo Dome are displaced by a higher-order reverse
splay with the volcanoes may indicate that the Calchaqu fault near the Calchaqu River, herein called the Barbaro
Fault was active in the Quaternary. Similarly, alluvial and Fault (Figures 3a and 4a). Therefore, shortening of the Dome
lacustrine Quaternary deposits covering the east limb of the

6 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Figure 4. Quaternary structures of the La Poma area and the Luracatao Valley. So delineates bedding
planes. (a) Tertiary and Quaternary strata cut by Barbaro Thrust Fault. (b) Cachi Fault juxtaposing Creta-
ceous strata of the Pirgua Subgroup against gently folded Quaternary conglomerate. (c) Angular unconfor-
mity between Quaternary Unit 1 and Unit 2 conglomerates from the Cachi area. (d) Folded and faulted
Unit 1 conglomerate.

and activity of the Calchaqu Fault likely continued both into non-cylindrical deformation in the first and cylindrical
the Quaternary. deformation in the second episode that likely continued into
[14] In order to assess the fault type of, and displacement Quaternary times.
magnitudes on, the Toro Muerto and the Calchaqu Faults,
the orientation of these faults was constrained in 3D 5. Structure of the Southern Luracatao Valley
(Figure 5a). The Toro Muerto Fault dips at 10 to 15 to the
west and, thus, constitutes a bona-fide thrust fault. By con- [16] Granitoid rocks of the Paleozoic Oire Complex are
trast, the Calchaqu Fault dips at 25 to 35 to the east. Thus, displaced over folded strata of the Eocene Quebrada de los
the splay of the Calchaqu Fault west of the Gemelos Vol- Colorados Formation [Bosio et al., 2009] along the Refugio
canoes is shallower than the main segment of this reverse Fault and the Luracatao Fault along the western flank of the
fault and is somewhat curved in map view (Figure 5a). Luracatao Valley (Figure 6). The Refugio Fault cuts the
Judging by the portions of the Cerro Bayo Dome that are Luracatao Fault [Hongn and Seggiaro, 2001]. Both faults
obscured by the Puncoviscana Formation on the Toro dip approximately 40 to the west, with the Luracatao Fault
Muerto Thrust Fault and the Calchaqu Reverse Fault, hori- shallowing toward the intersection with the Refugio Fault
zontal displacements, i.e., shortening magnitudes, are esti- (Figure 5). Interestingly, the Refugio Fault is concordant to
mated at 2.5 km and 3 km, respectively (Figure 3a). These the metamorphic foliation in the metagranite and felsic dikes
estimates are based on the map view distances between in the hanging walls of the faults (see equal-area projection
the fault plane traces and the outermost (extrapolated) tra- in Figure 6). Thus, the orientation of the Refugio Fault and
jectories of inclined bedding planes delineating the circum- numerous higher-order cataclasite zones associated with this
ference of the dome. fault appear to be controlled by the Paleozoic structural
[15] In summary, two deformation episodes are evident anisotropy [Hongn and Seggiaro, 2001].
from the La Poma area. The first episode is characterized by [17] At the eastern valley flank, conglomerate of the
the formation of the Cerro Bayo Dome. This increment must Pirgua Subgroup is in contact with the Quebrada de los
have occurred after the Eocene as it affected the Quebrada de Colorados Formation at the Cachi Fault (Figure 6). Here,
los Colorados Formation. Displacement of the Puncoviscana conglomerate of the Pirgua Subgroup is juxtaposed against
Formation over the western Cerro Bayo Dome and trunca- slightly folded Quaternary strata, which are unconformably
tion of the Cerro Bayo Fault by the Toro Muerto Fault overlying folded strata of the Quebrada de los Colorados
characterizes the second deformation increment. Thrusting Formation (Figure 4b). The Cachi Fault has an average dip
on the Toro Muerto Fault and reverse faulting on the of 60 . The hanging wall of this fault is characterized by
Calchaqu Fault are key elements of this deformation incre- thick conglomerate deposits of the syn-rift facies of the
ment. Deformation of both episodes differs in style, i.e., Pirgua Subgroup, which grade into conglomeratic sandstone

7 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Figure 5. Three-dimensional representations of first-order dip-slip faults of the La Poma and Luracatao
areas. Note that W-dipping faults in both areas are dipping shallower than E-dipping faults. (a) A 3D rep-
resentation of the Toro Muerto Thrust Fault and the Calchaqu Fault in the La Poma area. (b) A 3D rep-
resentation of the Refugio Fault, the Luracatao Fault and the Cachi Fault in the southern Luracatao Valley.

with distance to the fault. Therefore, the steep dip of this plunging anticline cored by Tertiary strata (Figure 7b).
fault can be explained best by its formation as a normal Quaternary strata dip shallower than the Tertiary strata.
growth fault [Hongn and Seggiaro, 2001; Carrera et al., Thus, folding occurred after deposition of the Quaternary
2006] during deposition of the Pirgua Subgroup. Thus, the strata, but may well have started prior to Quaternary times.
orientation of first-order faults in the Luracatao Valley was The anticline is characterized by a moderately SE-dipping
mostly controlled by pre-Andean planar structures. fold-axial plane (see equal-area projection in Figure 7b)
indicating NW-SE shortening. Bedding planes of Unit 2
6. Structure of the Cachi Area steepen toward a SE-dipping reverse fault that dissects the
west limb of the anticline. Here, the geometry of meter-scale
[18] In the Cachi area (Figures 1b and 7a) coarse Quater- folds in Unit 2 along with the overall variation in bedding
nary conglomerate of the Oran Group [Carrera et al., 2006] planes suggests top-to-NW displacement (Figure 4d). Unit 1
cover Tertiary sandstone and can be subdivided into two forms the hanging wall of the reverse fault near overturned
units, Unit 1 and Unit 2 (Figure 7b). Unit 2 covers uncon- strata of Unit 2 (Figure 7b). Collectively, the structural
formably tilted strata of, and is thus younger than, Unit 1 observations point to the presence of a kilometer-scale
(Figure 4c). Unit 1 has a yellowish matrix enveloping poorly cylindrical fault-propagation fold in the Cachi area.
sorted clasts of cordierite schist, biotite gneiss and meta-
granitoid rock. Collectively, these lithologies are typical
for the Cachi Range west of the study area [Turner, 1960].
7. Fault Slip Analysis
Unit 2 consists chiefly of clasts set in a red matrix that are [20] To better understand the kinematics of first-order
derived from the Puncoviscana Formation of the eastern structures and the variability of fault kinematics on the
valley flank. kilometer-scale, higher-order brittle faults genetically related
[19] The synorgenic Quaternary strata [Carrera et al., to the formation of first-order structures were examined in
2006] form a kilometer-scale, shallowly northward the Calchaqu Valley and the Luracatao Valley. Inversion of

8 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

shortening directions and the strike of the Calchaqu Fault


points to a component of sinistral displacement in addition to
reverse sense-of-slip on this fault. Similar to the orientation
of the Toro Muerto Fault, brittle faults associated with this
fault dip shallowly to the WSW and are characterized by
reverse sense-of-slip. Striations plunge almost uniformly
toward the W. This indicates E-W shortening (Figure 3a)
and vertical extension and is consistent with E-directed
thrusting on the Toro Muerto Fault. Brittle faults of the
Cerro Bayo Dome display conjugate sets of NNE-striking
reverse faults that are mostly concordant to bedding planes
of the Yacoraite Formation and the Mealla Formation. This
suggests that the Cerro Bayo Dome formed by flexural slip
on bedding planes. Striations on fault planes plunge mostly
toward the E and W indicating E-W shortening (Figure 3a)
and subvertical extension, which is consistent with doming.
[22] Brittle faults related to the activity of the Cachi Fault
in the Luracatao Valley display variable orientations.
Nonetheless, the majority of faults forms clusters of E-W
striking dextral to dextral-oblique reverse faults and N-S
striking sinistral to sinistral-oblique reverse faults resulting
in mainly E-W to NW-SE shortening (Figure 6). Similarly,
brittle faults associated with the Refugio Fault are highly
variable in orientation and are characterized by reverse
sense-of-shear and strike slip. For the majority of the fault
sets, minimum principal strain axes are remarkably horizon-
tal and trend either WNW-ESE or NNE-SSW (Figure 6),
whereas extension is respectively horizontal and mostly
subvertical (Table 1). It is uncertain, whether the two kine-
matic regimes adhere to temporally distinct kinematic events
or formed as a consequence of locally variable, pre-existing
rock anisotropy during a single deformation event. In sum-
mary, prominent structures in the Calchaqu Valley and the
Luracatao Valley formed under overall E-W shortening.
Figure 6. Geological units and structure of the Luracatao
However, in detail shortening directions vary with position
Valley. Arrows pointing toward each other indicate shorten-
and affinity to a given first-order structure. Except for the
ing directions inferred from brittle fault kinematics. Lower-
Refugio Fault, which may record two distinct kinematic
hemisphere, equal-area projection shows concordance of
regimes, shortening directions are uniform at a given first-
the Luracatao and Refugio Faults with foliation surfaces
order structure (Figure 8).
and aplitic dike margins of the meta-granitoid Oire Com-
plex. Fold axes in the valley center are adopted from Bosio
et al. [2009]. 8. Implication of Structural Data for Causes
of Intracontinental Deformation
higher-order shear faults provided directions of principal [23] Knowledge of the (1) age and kinematics of defor-
strain axes, l1 l2 l3, which respectively correspond to mation, (2) inherited structural anisotropy and mechanical
the directions of extension, intermediate elongation and configuration of upper crust as well as (3) plate kinematic
shortening. In the La Poma area, principal strain axes were parameters is critical to assess causes for intracontinental
inferred from fault slip data acquired in fault rock of the deformation. With regard to (1) and (2), our structural study
Toro Muerto Fault and the Calchaqu Fault as well as in advances our understanding of intracontinental deformation
Cretaceous and Tertiary strata of the Cerro Bayo Dome in the southern Central Andes. Comparison of our data from
(Figure 3a). In the Luracatao Valley, fault slip data were the Eastern Cordillera with equivalent data sets from other
recorded in the Pirgua Subgroup at the Cachi Fault and in parts of the Eastern Cordillera, the Puna and the Pampean
granitoid rocks of the Ordovician Oire Complex that were Ranges (Figure 9 and Table 2) promises to unravel the extent
affected by the Refugio Fault and the Luracatao Fault to which intracontinental deformation in the southern Cen-
(Figure 6). Results of the fault slip analysis are given in tral Andes is controlled by (3) plate (boundary) kinematics.
Table 1 and the geometry of individual fault sets is shown in [24] Deformation in the Eastern Cordillera started in late
Figure S1. Eocene to Early Oligocene times [Deeken et al., 2006] with
[21] At the Calchaqu Fault, brittle faults are highly vari- the local formation of an upper-crustal, basement-cored
able in orientation, but striations plunge dominantly toward dome at La Poma. Interestingly, this area is devoid of syn-
the NW and SE. Reverse faults dominate the fault popula- rift sedimentary rocks of the Pirgua Subgroup, which else-
tions and indicate NW-SE shortening (Figure 3a) and sub- where in the Eastern Cordillera can be many thousands of
vertical extension. The obliquity between the inferred meters thick [Salfity and Marquillas, 1994; Marquillas

9 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Figure 7. Structure of the Cachi area. (a) First-order Quaternary structures east of Cachi. (b) Outcrop
pattern of Quaternary fault-propagation fold in the hanging wall of SE-dipping thrust fault. Lower-
hemisphere, equal-area projection shows shortening direction inferred from the orientation of Quaternary
and Tertiary strata of the fold-propagation fold. Great circles represent the limbs of the fold.

et al., 2005]. Thus, at the onset of E-W shortening, the [26] E- and W-dipping faults differ in terms of shorten-
interface between Paleozoic basement rocks and Cenozoic ing directions inferred from fault slip analysis and inferred
sedimentary cover rocks was at significantly shallower age of activity. Notably, the Cachi Fault displaces Qua-
crustal level at La Poma than in areas replete with Pirgua ternary deposits (Figure 4b) and is characterized by overall
Subgroup strata. It is, therefore, conceivable that the reduced WNW-ESE shortening (Figures 3a, 6, and 8). This short-
structural anisotropy and horizontal layering of the upper ening direction is also evident from the Quaternary fault-
crust preconditioned the La Poma area for the formation of propagation fold in the Cachi area (Figures 7 and 8). By
a kilometer-scale dome. contrast, W-dipping faults do not deform Quaternary
[25] Doming in the La Poma area was followed by reverse deposits and formed under E-W shortening (Toro Muerto
and thrust faulting evident in the Calchaqu Valley and the Thrust) or WNW-ESE and NNE-SSW shortening (Refugio
Luracatao Valley. In both valleys, the E-dipping reverse Fault). Our data shows that overall NW-SE shortening con-
faults, the Calchaqu and Cachi Faults, are steeper than the tinued to be active into the Quaternary, whereas NNE-SSW
W-dipping faults, the Toro Muerto and Refugio Faults. The shortening associated with the Refugio Fault seems to have
spatial association of syn-rift deposits with the Cachi Fault ceased prior to this time. This agrees with fault-kinematic
suggests that the steep dip of this fault is likely a conse- data from the Puna, which is characterized chiefly by NW-
quence of its initiation as a normal fault during rifting in the SE shortening lasting in places from Miocene to Pleistocene
Cretaceous [Hongn and Seggiaro, 2001; Carrera et al., 2006]. times (Figure 9 and Table 2). Evidence for NW-SE short-
This may also apply for the Calchaqu Fault. By contrast, ening in the Pleistocene is at variance with the notion that
W-dipping faults are not associated with syn-rift deposits shortening directions in the Eastern Cordillera and adjacent
and, thus, did most probably not form from normal faults. Puna changed from NW-SE to NE-SW some time during the
The thrust geometry of these faults suggests that they ini- Pliocene [Marrett et al., 1994]. Based on a detailed, but
tiated as new faults during Andean shortening, in agreement spatially rather limited, fault slip analysis in the Eastern
with Mohr-Coulomb theory for the formation of new shear Cordillera (i.e., Quebrada del Toro), this temporal change in
faults at upper crustal level. However, the moderately dip- shortening direction was attributed to a directional change in
ping Refugio Fault is concordant to foliation surfaces and absolute motion of the South American Plate [Marrett and
margins of aplitic dikes in Paleozoic metagranitoid rocks at Strecker, 2000]. Judging by the occurrence of NW-SE
the fault (Figure 6). Thus, orientations of most first-order shortening in Pleistocene time in the Puna and the Eastern
faults in the study areas seem to be effectively controlled by Cordillera (Figure 9 and Table 2), the hypothesis that intra-
pre-Neogene planar structures, recognized also from other continental shortening directions portray plate motion
parts of the Eastern Cordillera [Grier et al., 1991; Hongn directions requires reconsideration.
and Seggiaro, 2001; Carrera et al., 2006].

10 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

southern and northern Puna [e.g., Maffione et al., 2009;


Tibaldi et al., 2009] and Pampean Ranges [Gapais et al.,
2000]. This suggests that Pleistocene shortening directions
in the southern Central Andes (Figure 9 and Table 2) are
unrelated to the post-3.2 Ma N75 E direction of Nazca-South
America plate convergence [DeMets et al., 1994; Norabuena
et al. 1998].
[28] Except for the Refugio Fault, shortening directions
inferred from fault slip data in the La Poma area and the
Luracatao Valley are uniform at, but can vary with orienta-
tion and type of, a given first-order structure (Figure 8). This
has also been observed in other parts of the southern Central
Andes. Specifically, E-W to subvertical shortening and N-S

Figure 8. Synoptic diagrams showing the variation of min-


imum principal strain axes (filled symbols) and maximum
principal strain axes (open symbols) inferred from brittle
fault analysis for each station in the La Poma area, Luracatao
Valley and Cachi area. Strain axes inferred from fault sets
containing more than ten faults are represented by circles
and are more reliable in terms of indicating overall shortening
directions (arrows) than strain axes inferred from fault sets
with less than ten faults, which are represented by squares.
Great circles indicate orientations of prominent faults.

Figure 9. Spatial distribution of shortening directions


[27] In our fault-kinematic data set, sub-vertical shortening inferred from analyses of brittle shear faults compiled from
is apparent in 4 out of 67 stations (Table 1). Therefore, Allmendinger et al. [1989], Marrett et al. [1994], Cladouhos
gravitational collapse does not seem to account for the et al. [1994], de Urreiztieta et al. [1996], Schoenbohm and
observed fault kinematics. Our data are dominated by hori- Strecker [2009], Maffione et al. [2009], Tibaldi et al. [2009],
zontal NW-SE and E-W shortening and subvertical exten- Montero Lopez et al. [2010], and this study. Note NW-SE
sion, particularly at first-order faults that were active in shortening directions persisting to Pleistocene times. See
Quaternary times. NW-SE shortening in Pliocene rocks has Table 2 for details of the analyses. For location of the area
also been documented elsewhere from the Eastern Cordillera, displayed in the figure, see Figure 1a.

11 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Table 2. Compilation of Kinematic Data From the Southern Central Andes


Location Coordinates ( S/ W) Na Shortening Extension Age of Deformationb Sourcec
Loma Blanca 23.0/66.5 20 246/09 155/10 Quaternary* 1
Chorrillos 24.2/66.5 27 079/42 345/04 Quaternary 1
17 103/08 012/06 Tertiary 1
Negra Muerta 24.4/66.2 7 251/48 346/04 Tertiary 1
Pastos Grandes 24.5/66.8 19 115/00 208/87 ? 1
16 039/43 138/09 Quaternary 1
40 015/84 162/05 Post10.6 Ma 1
Cerro Pircas 24.7/67.0 26 232/18 126/40 Quaternary * 1
Sierra de Bequeville 25.0/67.0 23 118/07 256/81 Tertiary 1
14 064/27 334/00 Quaternary * 1
7 065/08 331/28 Quaternary * 1
Salar de Hombre Muerto 25.3/67.1 9 176/69 297/11 Post2 Ma * 1
Salar de Antofalla 25.5/67.5 16 052/50 142/00 Quaternary * 1
Quebrada de Humahuaca 23.2/65.3 11 060/18 161/30 Post1.5 Ma 1
10 279/25 101/65 Pleistocene 1
8 228/62 330/06 MiocenePliocene 1
Quebrada Del Toro 24.5/66.0 128 311/11 091/75 Pre0.98 1
80 061/03 307/82 Post0.98 * 1
14 237/00 329/87 Quaternary 1
La Poma 24.6/66.2 58 299/08 067/77 Tertiary 1
55 060/06 302/78 Quaternary * 1
Pucara 25.7/66.2 12 308/01 207/86 >13.412.11 Ma 1
27 062/09 181/72 Post12.11 Ma 1
Choromoro 26.4/65.4 9 287/02 172/85 Pliocene 1
Tiomayo 22.5/66.5 34 E-W N-S Post6.75 Ma 2
Rinconada 22.3/66.1 18 111 vertical Pre & post 12 Ma 2
47 vertical E-W Quaternary 2
Abra Moreta 22.4/66.0 58 101 vertical Post14.26 Ma 2
132 vertical E-W Quaternary 2
Pan de Azucar 22.6/65.8 22 150 vertical ? 2
Abra Pampa 22.6/65.8 36 93 vertical ? 2
Pumahuasi 22.2/65.8 54 NE-SW NW-SE Quaternary 2
Yavi 22.0/65.4 37 NW-SE vertical Quaternary 2
Quebrada Honda 21.8/65.2 36 vertical N32 E Miocene-younger 2
Abra Tuc Tuca 22.4/65.4 14 137 vertical Post9.57 2
Sierra Aconquija 27.0/66.0 70 WNW vertical Post2.96 Ma 3
Sierra Alto Huasi 27.0/66.6 28 NW-SE vertical Pre2.35 3
65 284/05 193/11 Post2.35 3
Rio Bolson 27.0/66.6 28 WNW vertical Post3.35 Ma 3
La Hoyada 27.0/66.6 45 WNW ENE Post3.35 Ma 3
Pasto Ventura 26.6/67.2 27 E-W N-S Quaternary 3
El Penon 26.6/67.2 19 vertical N-S Quaternary 3
Fiambala BasinNormal Fault 27.0/67.7 20 vertical NW-SE Post 3.7 Ma 4
14 vertical NW-SE Post 3.7 Ma 4
15 vertical NW-SE Post 3.7 Ma 4
Fiambala BasinEastern Fault 27.0/67.7 8 NW-SE ? NeogenePleistocene 4
27.0/67.7 6 NNE-SSW ? NeogenePleistocene 4
26.9/67.7 6 NW-SE ? NeogenePleistocene 4
Fiambala BasinWestern Fault 27.0/67.8 10 WNW-ESE ? MiocenePleistocene 4
Fiambala BasinCentral Fault 27.0/67.7 19 NW-SE ? Pleistocene 4
Honda Fault 26.7/66.8 6 WSW-ENE ? ? 4
Sierra de Cochinoca 22.5/65.8 7 NW-SE N-S OligoceneMiocene 5
Abra Pampa 22.7/65.6 5 WSW-ENE NNE-SSW OligoceneMiocene 5
Tres Cruces 22.8/65.6 6 NNW-SSE WSW-ENE MidLate Miocene 5
Pasto Ventura 26.6/67.2 8 NW-SE NE-SW Quaternary 6
La Hoyada 27.0/66.6 6 E-W NW-SE 72.5 Ma 6
11 NW-SE NE-SW 72.5 Ma 6
11 E-W NNW-SSE 72.5 Ma 6
7 WNW-ESE NE-SW 72.5 Ma 6
12 NNE-SSW NW-SE 72.5 Ma 6
Las Papas Valley 27.0/67.7 9 WNW-ESE ? Post 3.7 Ma 6
Salar de Uyuni 20.0/68.1 6 NW-SE vertical Pliocene 7
20.0/68.1 7 NW-SE NE-SW Pliocene 7
20.0/68.1 13 E-W N-S Miocene 7
Salar de UyuniOllague Area 21.3/68.2 9 vertical E-W ? 7
Salar de UyuniCosca Area 21.0/68.4 9 NW-SE vertical Pliocene 7
21.0/68.4 10 vertical NE-SW PlioceneQuaternary 7
Salar de UyuniIgnimbrites 21.3/68.2 6 NW-SE vertical Pliocene 7
21.3/68.2 6 vertical NW-SE ? 7
Salar de UyuniEast of Ollague 21.3/68.3 9 NW-SE vertical Pliocene 7

12 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Table 2. (continued)
Location Coordinates ( S/ W) Na Shortening Extension Age of Deformationb Sourcec
North of Pipanaco 27.0/66.7 79 NW-SE vertical ? 8
27.0/66.2 47 NW-SE vertical Quaternary * 8
27.0/66.5 58 NE-SW vertical Quaternary * 8
26.7/66.0 10 052/12 156/46 Quaternary * 8
27.0/66.4 13 NW-SE NE-SW Quaternary * 8
Tinogasta 28.0/67.6 44 WNW-ESE vertical Quaternary * 8
West of Pipanaco 28.1/67.6 27 E-W vertical Quaternary * 8
South of Pipanaco 28.7/66.3 41 NE-SW vertical Quaternary * 8
28.7/66.3 12 331/63 069/04 ? 8
East of Pipanaco 28.3/65.9 84 NE-SW vertical Quaternary * 8
28.0/65.5 18 295/13 118/77 Quaternary * 8
28.1/65.8 30 068/11 328/44 Quaternary * 8
27.5/66.0 41 285/13 183/42 Quaternary * 8
Sierra Alto Huasi 27.0/66.6 30 NW-SE Vertical MiocenePliocene 9
Cerro Negro 27.0/66.6 23 NW-SE NE-SW MiocenePliocene 9
Sierra Chango 27.0/66.6 12 NW-SE vertical MiocenePliocene 9
Sierra Hombre Muerto 27.0/66.6 23 NW-SE NE-SW MiocenePliocene 9
Rio El Bolson 27.0/66.6 30 E-W N-S Quaternary 9
Rio Totoral 27.0/66.6 19 E-W N-S Quarernary 9
Sierra de Quilmes 27.0/66.6 32 WNW-ESE NNE-SSW Quaternary 9
Quebrada del ToroSan Bernado Fault 24.5/66.0 29 NW-SE NE-SW Miocene0.98 Ma 10
6 vertical NW-SE Miocene0.98 Ma 10
6 90.4/0.7 180.6/20.9 4.17 Ma- present 10
Quebrada del ToroSol Fault 24.5/66.0 9 NW-SE NNW-SSE Miocene0.98 Ma 10
39 NE-SW vertical 4.17 Mapresent 10
11 NE-SW NW-SE 4.17 Mapresent 10
Quebrada del ToroGlgota fault 24.5/65.0 22 NW-SE NNE-SSW Miocene0.98 Ma 10
30 NE-SW vertical 4.17 Mapresent 10
Quebrada Carachi 24.5/66.0 65 NW-SE NE-SW Miocene0.98 Ma 10
17 NE-SW vertical 4.17 Mapresent 10
Toro Muerto Fault 24.7/66.2 32 E-W ? Tertiary 11
Calchaqui Fault 24.7/66.2 85 NW-SE NE-SW Quaternary 11
Cerro Bayo Dome 24.7/66.2 197 E-W vertical Tertiary-Quaternary 11
Refugio Fault 25.3/66.5 137 NNE-SSW vertical Tertiary 11
Refugio Fault 25.3/66.5 126 WNW-ESE NE-SW Tertiary 11
Luracatao Fault 25.3/66.5 13 WNW-ESE NW-SE Tertiary 11
Cachi Fault 25.3/66.4 85 WNW-ESE NE-SW Quaternary 11
a
Number of faults per population.
b
The asterisk refers to age inferred from similarity of shortening direction with that in adjacent younger rocks.
c
References: (1) Marrett et al. [1994], (2) Cladouhos et al. [1994], (3) Allmendinger et al. [1989], (4) Schoenbohm and Strecker [2009], (5) Maffione
et al. [2009], (6) Montero Lopez et al. [2010], (7) Tibaldi et al. [2009], (8) de Urreiztieta et al. [1996], (9) Allmendinger [1986], (10) Marrett and
Strecker [2000], (11) this study.

extension characterizes NW-SE striking sinistral fault zones, normal stresses, generated by (oblique) plate convergence at
notably the Calama-Olacapato-El Toro Fault Zone [Riller the plate boundary. In our opinion, the fault kinematics
et al., 2001, Petrinovic et al., 2006; Acocella et al., 2011]. observed in the continental interior of the Central Andes is
NW-SE shortening and subvertical extension is mostly more likely controlled by deformation partitioning of upper
associated with orogen-parallel sinistral reverse faults in the crust rather than by directions of plate motion or plate con-
Eastern Cordillera, the Puna [Coutand et al., 2001; Riller and vergence. The importance of deformation partitioning in con-
Oncken, 2003; Maffione et al., 2009] and Pampean Ranges trolling the orientations in principal shortening (compression)
[de Urreiztieta et al., 1996; Gapais et al., 2000]. ENE-WSW axes that deviate significantly from plate convergence vectors
shortening and subvertical extension has been recorded at is also known from other well-studied regions of intraconti-
NE-SW striking dextral reverse fault zones in the Puna nental deformation, such as the Alpine-Himalayan collision
[Marrett et al., 1994] and Pampean Ranges [de Urreiztieta zone in Iran [Lacombe et al., 2006; Navabpour et al., 2007;
et al., 1996; Gapais et al., 2000]. Collectively, this suggests Shabanian et al., 2010; Kargaranbafghi et al., 2011].
that small-scale brittle faults are kinematically related to
prominent (first-order) faults in the southern Central Andes. 9. Conclusions
[29] First-order faults, N-S striking sinistral reverse and
NE-SW striking dextral reverse faults in particular, form a [30] Detailed structural analyses conducted in the Calchaqu
rhombic pattern in the southern Central Andes [Gapais Valley and the Luracatao Valley revealed that late Tertiary to
et al., 2000; Riller and Oncken, 2003]. This pattern has Quaternary intracontinental deformation was chiefly accom-
been explained in terms of deformation partitioning of plished by prominent orogen-parallel reverse faults and folds.
the upper-crust under bulk E-W shortening [Gapais et al., Locally, i.e., at La Poma in the Calchaqu Valley, deformation
2000; Riller and Oncken, 2003; Riller et al., 2012]. E-W started with doming of upper crust, which may be attributed to
shortening in the plateau area and adjacent morphotectonic the thickness variation of Cretaceous to Eocene rift sediments.
regions arises likely from far-field paleostresses, notably The geometry of the reverse faults was significantly influenced

13 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

by Paleozoic dikes and foliation surfaces as well as Cretaceous Coutand, I., P. R. Cobbold, M. de Urreiztieta, P. Gautier, A. Chauvin,
normal faults. D. Gapais, E. A. Rossello, and O. Lopz-Gamund (2001), Style and
history of Andean deformation, Puna Plateau, northwestern Argentina,
[31] Analysis of fault slip data indicates that upper-crustal Tectonics, 20, 210234, doi:10.1029/2000TC900031.
doming in the La Poma area occurred under E-W shortening, Cristallini, E., A. H. Cominguez, and V. A. Ramos (1997), Deep structure
whereas orogen-parallel reverse faults in the three study areas of the Metn-Guachipas Region: Tectonic inversion in Northwestern
formed mostly under NW-SE shortening. Shortening direc- Argentina, J. South Am. Earth Sci., 10, 403421, doi:10.1016/S0895-
9811(97)00026-6.
tions are kinematically related to, and rather uniform with DeCelles, P., and B. K. Horton (2003), Early to middle Tertiary foreland
respect to first-order faults, evident also elsewhere in the basin development and the history of Andean crustal shortening in Boli-
southern Central Andes. Specifically, fault-kinematic regimes via, Geol. Soc. Am. Bull., 115, 5877, doi:10.1130/0016-7606(2003)
115<0058:ETMTFB>2.0.CO;2.
vary with position and structural precondition of upper crust Deeken, A., E. R. Sobel, I. Coutand, M. Haschke, U. Riller, and M. R.
but do not appear to change systematically with time on the Strecker (2006), Development of the southern Eastern Cordillera,
regional scale. Shortening directions inferred from fault slip NWArgentina, constrained by apatite fission track thermochronology: From
early Cretaceous extension to middle Miocene shortening, Tectonics, 25,
analyses compiled from other parts of the Eastern Cordillera, TC6003, doi:10.1029/2005TC001894.
Puna and Pampean Ranges corroborate this notion. Moreover, DeMets, C., R. G. Gordon, D. F. Aargus, and S. Stein (1994), Effect of
age and kinematics of regional deformation are at variance recent revisions to the geomagnetic reversal time scale on estimates of
with hypotheses attributing intracontinental shortening direc- current plate motion, Geophys. Res. Lett., 21, 21912194, doi:10.1029/
94GL02118.
tions in the southern Central Andes to gravitational collapse or de Urreiztieta, M., D. Gapais, C. Le Corre, P. R. Cobbold, and E. Rossello
changes in plate (boundary) kinematics. The kinematics of (1996), Cenozoic dextral transpression and basin development at the
intracontinental deformation in the southern Central Andes southern edge of the Puna plateau, northwestern Argentina, Tectonophy-
sics, 254, 1739, doi:10.1016/0040-1951(95)00071-2.
appears to be rather influenced by pre-existing structural Daz, J., D. Malizzia, and G. Bossi (1987), Anlisis estratigrfico del Grupo
anisotropy and local kinematic regime as a consequence of Payogastilla (Terciario Sup.), in Actas del X Congreso Geolgico Argen-
deformation partitioning of upper-crust under bulk E-W tino, Tucumn, vol. 2, pp. 113117, Asoc. Geol. Argent., Buenos Aires.
Gapais, D., P. Cobbold, O. Bourgeois, D. Roubya, and M. de Urreiztieta
shortening. We, therefore, caution the usage of brittle fault- (2000), Tectonic significance of fault-slip data, J. Struct. Geol., 22,
kinematics in continental interiors as proxy for tracking plate- 881888, doi:10.1016/S0191-8141(00)00015-8.
kinematic changes. Gephart, J. (1994), Topography and subduction geometry in the central
Andes: Clues to the mechanics of a non-collisional orogen, J. Geophys.
[32] Acknowledgments. This study was funded by the National Res., 99, 12,27912,288, doi:10.1029/94JB00129.
Science and Engineering Research Council of Canada. We thank S. Guzmn, Grier, M. E., J. A. Salfity, and R. W. Allmendinger (1991), Andean reacti-
F. Hongn, and I. Petrinovic for many enlightening discussions and logistical vation of the Cretaceous Salta rift, northwestern Argentina, J. South Am.
support and W. A. Morris for the use of Geomodeller. Constructive reviews Earth Sci., 4, 351372, doi:10.1016/0895-9811(91)90007-8.
by L. Schoenbohm, G. Axen, J. Kley, and an anonymous person improved Guzmn, S. R., I. A. Petrinovic, and J. A. Brod (2006), Pleistocene mafic vol-
the content of this work significantly. canoes in the Puna-Cordillera Oriental boundary, NW-Argentina, J. Volcanol.
Geotherm. Res., 158, 5169, doi:10.1016/j.jvolgeores.2006.04.014.
Hongn, F., and U. Riller (2007), Tectonic evolution of the western margin
References of Gondwana inferred from syntectonic emplacement of Paleozoic granit-
Acocella, V., A. Gioncada, R. Omarini, U. Riller, R. Mazzuoli, and L. Vezzoli oid plutons in N.W. Argentina, J. Geol., 115(2), 163180, doi:10.1086/
(2011), Tectono-magmatic characteristics of the Calama-Olacapato-El Toro 510644.
Fault Zone (Puna and Eastern Cordillera), Central Andes, Tectonics, 30, Hongn, F., and R. E. Seggiaro (2001), Hoja Geolgica 2566-III, Cachi, Bol.
TC3005, doi:10.1029/2010TC002854. 248, scale 1:250,000, Serv. Geol. Min. Argent., Buenos Aires.
Allmendinger, R. W. (1986), Tectonic development, southeastern border of Hongn, F., C. del Papa, J. Powell, I. Petrinovic, R. Mon, and V. Deraco
the Puna Plateau, northwestern Argentine Andes, Geol. Soc. Am. Bull., 97, (2007), Middle Eocene deformation and sedimentation in the Puna-Eastern
10701082, doi:10.1130/0016-7606(1986)97<1070:TDSBOT>2.0.CO;2. Cordillera transition (23 26 S): Control by pre-existing heterogeneities on
Allmendinger, R. W., M. Strecker, J. E. Eremchuk, and P. Francis (1989), the pattern of initial Andean shortening, Geology, 35(3), 271274,
Neotectonic deformation of the southern Puna Plateau, northwestern doi:10.1130/G23189A.1.
Argentina, J. South Am. Earth Sci., 2, 111130, doi:10.1016/0895-9811(89) Isacks, B. (1988), Uplift of the central Andean plateau and bending of the
90040-0. Bolivian orocline, J. Geophys. Res., 93, 32113231, doi:10.1029/
Allmendinger, R. W., T. E. Jordan, S. M. Kay, and B. Isacks (1997), The JB093iB04p03211.
evolution of the Altiplano-Puna Plateau of the central Andes, Annu. Rev. Jordan, T. E., and R. N. Alonso (1987), Cenozoic stratigraphy and basin
Earth Planet. Sci., 25, 139174, doi:10.1146/annurev.earth.25.1.139. tectonics of the Andes mountains, 20 28 South Latitude, AAPG Bull.,
Angelier, J., and J. Goguel (1979), Sur une methode simple de determina- 71, 4964.
tion des axes princepaux des contraintes pour une polulation de failles, Jordan, T. E., B. L. Isacks, R. W. Allmendinger, J. A. Brewer, V. A. Ramos,
C. R. Seances Acad. Sci. Ser. D, 288(3), 307310. and C. J. Ando (1983), Andean tectonics related to geometry of sub-
Barnes, J. B., and T. A. Ehlers (2009), End member models for Andean Plateau ducted Nazca plate, Geol. Soc. Am. Bull., 94, 341361, doi:10.1130/
uplift, Earth Sci. Rev., 97, 105132, doi:10.1016/j.earscirev.2009.08.003. 0016-7606(1983)94<341:ATRTGO>2.0.CO;2.
Blasco, G., E. Zapettini, and F. Hongn (1995), Hoja Geolgica San Antonio Kargaranbafghi, F., F. Neubauer, and J. Genser (2011), Cenozoic kinematic
de los Cobres, map, scale 1:250,000, Secr. de Min. de la Nacion, Dir. evolution of southwestern Central Iran: Strain partitioning and accommo-
Nac. del Serv. Geol., Buenos Aires. dation of ArabiaEurasia convergence, Tectonophysics, 502, 221243,
Bosio, P. P., J. Powell, C. Del Papa, and F. Hongn (2009), Middle Eocene doi:10.1016/j.tecto.2010.02.004.
deformation-sedimentation in the Luracatao Valley: Tracking the begin- Kay, S. M., and B. L. Coira (2009), Shallowing and steepening subduction
ning of the foreland basin of northwestern Argentina, J. South Am. Earth zones, continental lithospheric loss, magmatism and crustal flow under
Sci., 28, 142154, doi:10.1016/j.jsames.2009.06.002. the central Andean Altiplano-Puna plateau, in Backbone of the Americas:
Byerlee, J. D. (1968), Brittle-ductile transition in rocks, J. Geophys. Res., Shallow Subduction, Plateau Uplift and Ridge and Terrane Collision,
73(14), 47414750, doi:10.1029/JB073i014p04741. edited by S. M. Kay et al., Geol. Soc. Am. Mem., 204, 229259,
Carrera, N., J. A. Muoz, F. Sbat, R. Mon, and E. Roca (2006), The role doi:10.1130/2009.1204(11).
of inversion tectonics in the structure of the Cordillera Oriental (NW Kay, S. M., B. Coira, and J. G. Viramonte (1994), Young mafic backarc
Argentinean Andes), J. Struct. Geol., 28, 19211932, doi:10.1016/j.jsg. volcanic rocks as indicators of continental lithospheric delamination
2006.07.006. beneath the Argentine Puna Plateau, Central Andes, J. Geophys. Res.,
Cladouhos, T. T., R. W. Allmendinger, B. Coira, and E. Farrar (1994), Late 99, 24,32324,339, doi:10.1029/94JB00896.
Cenozoic deformation in the central Andes: Fault kinematics from the Kley, J. (1996), Transition from basement-involved to thin-skinned thrusting
northern Puna, northwestern Argentina and southwestern Bolivia, in the Cordillera Oriental of southern Bolivia, Tectonics, 15, 763775,
J. South Am. Earth Sci., 7, 209228, doi:10.1016/0895-9811(94)90008-6. doi:10.1029/95TC03868.

14 of 15
TC4002 SANTIMANO AND RILLER: INTRACONTINENTAL DEFORMATION ANDES TC4002

Kley, J., A. H. Gangui, and D. Krger (1996), Basement-involved blind Petrinovic, I. A., U. Riller, and A. Brod (2005), The Negra Muerta Complex,
thrusting in the eastern Cordillera Oriental, southern Bolivia: Evidence southern Central Andes: Geochemical characteristics and magmatic evolu-
from cross-sectional balancing, gravimetric and magnetotelluric data, tion of an episodically active volcanic center, J. Volcanol. Geotherm. Res.,
Tectonophysics, 259, 171184, doi:10.1016/0040-1951(95)00067-4. 140, 295320, doi:10.1016/j.jvolgeores.2004.09.002.
Kraemer, B., D. Adelmann, M. Alten, W. Schnurr, K. Erpenstein, E. Kiefer, Petrinovic, I. A., U. Riller, J. Brod, G. Alvarado, and M. Arnosio (2006),
P. van den Bogaard, and K. Grler (1999), Incorporation of the Paleogene Bimodal volcanism in a tectonic transfer zone: Evidence for tectonically
foreland into the Neogene Puna Plateau: The Salar de Antofalla area, NW controlled magmatism in the southern Central Andes, NW Argentina,
Argentina, J. South Am. Earth Sci., 12, 157182, doi:10.1016/S0895- J. Volcanol. Geotherm. Res., 152, 240252, doi:10.1016/j.jvolgeores.
9811(99)00012-7. 2005.10.008.
Lacombe, O., F. Mouthereau, S. Kargar, and B. Meyer (2006), Late Cenozoic Ramelow, J., U. Riller, R. L. Romer, and O. Oncken (2006), Kinematic link
and modern stress fields in the western Fars (Iran): Implications for the tec- between episodic caldera collapse of the Negra Muerta Collapse Caldera
tonic and kinematic evolution of central Zagros, Tectonics, 25, TC1003, and motion on the OlacapatoEl Toro Fault Zone, NW-Argentina, Int. J.
doi:10.1029/2005TC001831. Earth Sci., 95, 529541, doi:10.1007/s00531-005-0042-x.
Maffione, M., F. Speranza, and C. Faccenna (2009), Bending of the Bolivian Riller, U., and O. Oncken (2003), Growth of the Central Andean plateau by
orocline and growth of the central Andean plateau: Paleomagnetic and tectonic segmentation is controlled by the gradient in crustal shortening,
structural constraints from the Eastern Cordillera (22 24 S, NW Argentina), J. Geol., 111, 367384, doi:10.1086/373974.
Tectonics, 28, TC4006, doi:10.1029/2008TC002402. Riller, U., I. Petrinovic, J. Ramelow, M. Strecker, and O. Oncken (2001),
Marquillas, R. A., C. del Papa, and I. F. Sabino (2005), Sedimentary aspects Late Cenozoic tectonism, collapse caldera and plateau formation in the
and paleoenvironmental evolution of a rift basin: Salta Group (Cretaceous- central Andes, Earth Planet. Sci. Lett., 188, 299311, doi:10.1016/
Paleogene), northwestern Argentina, Int. J. Earth Sci., 94, 94113, S0012-821X(01)00333-8.
doi:10.1007/s00531-004-0443-2. Riller, U., A. R. Cruden, D. Boutelier, and C. Schrank (2012), The causes of
Marrett, R., and M. R. Strecker (2000), Response of intracontinental defor- sinuous crustal-scale deformation patterns in hot orogens: Evidence from
mation in the central Andes to late Cenozoic reorganization of South scaled analogue experiments and the southern Central Andes, J. Struct.
American plate motions, Tectonics, 19, 452467, doi:10.1029/ Geol., 37, 6574, doi:10.1016/j.jsg.2012.02.002.
1999TC001102. Risse, A., R. B. Trumbull, B. Coira, S. M. Kay, and P. van den Boggard
Marrett, R. A., R. W. Allmendinger, R. N. Alonso, and R. E. Drake (1994), (2008), 40Ar/39Ar geochronology of mafic volcanism in the back-arc
Late Cenozoic tectonic evolution of the Puna Plateau and adjacent foreland, region of the southern Puna Plateau, Argentina, J. South Am. Earth Sci.,
northwestern Argentine Andes, J. South Am. Earth Sci., 7, 179207, 26, 115, doi:10.1016/j.jsames.2008.03.002.
doi:10.1016/0895-9811(94)90007-8. Russo, A., and A. Serraiotto (1979), Contribucin al conocimiento de la
Mercier, J. L., M. Sbrier, A. Lavenu, J. Cabrera, O. Bellier, J.-F. Dumont, estratigrafa terciaria en el moroeste Argentino, in Actas del VII Congreso
and J. Machare (1992), Changes in the tectonic regime above a subduction Geolgico Argentino. Neuqun, vol. 1, pp. 715 730, Asoc. Geol. Argent.,
zone of Andean type: The Andes of Peru and Bolivia during the Pliocene- Buenos Aires.
Pleistocene, J. Geophys. Res., 97, 11,94511,982, doi:10.1029/90JB02473. Salfity, J. A., and R. A. Marquillas (1994), Tectonic and sedimentary
Mon, R., and J. A. Salfity (1995), Tectonic evolution of the Andes of northern evolution of the Cretaceous Eocene Salta Group basin, Argentina, in
Argentina, in Petroleum Basins of South America, edited by A. J. Tankard Cretaceous Tectonics of the Andes, edited by J. A. Salfity, pp. 266315,
et al., AAPG Mem., 62, 269283. Vieweg, Wiesbaden, Germany.
Montero Lopez, M. C., F. D. Hongn, M. R. Strecker, R. Marrett, R. Seggiaro, Schoenbohm, L. M., and M. R. Strecker (2009), Normal faulting along the
and M. Sudo (2010), Late Mioceneearly Pliocene onset of N-S extension southern margin of the Puna Plateau, northwest Argentina, Tectonics, 28,
along the southern margin of the Central Andean Puna Plateau: Evidence TC5008, doi:10.1029/2008TC002341.
from magmatic, geochronological and structural observations, Tectono- Shabanian, E., O. Bellier, M. R. Abbassi, L. Siame, and Y. Farbod (2010),
physics, 494, 4863, doi:10.1016/j.tecto.2010.08.010. Plio-Quaternary stress states in NE Iran: Kopeh Dagh and Allah Dagh-
Moreno, J. A. (1970), Estratigrafa y Paleogeografa del Cretcico superior Binalud mountain ranges, Tectonophysics, 480, 280304, doi:10.1016/
en la cuenca del Noroeste Argentine, con especial mencin de los subgru- j.tecto.2009.10.022.
pos Balbuena y Santa Brbara, Asoc. Geol. Argent. Rev., 24, 944. Spang, J. H. (1972), Numerical method for dynamic analysis of calcite twin
Navabpour, P., J. Angelier, and E. Barrier (2007), Cenozoic post-collisional lamellae, Geol. Soc. Am. Bull., 83(2), 467471, doi:10.1130/0016-7606(1972)
brittle tectonic history and stress reorientation in the High Zagros 83[467:NMFDAO]2.0.CO;2.
Belt (Iran, Fars Province), Tectonophysics, 432, 101131, doi:10.1016/ Sperner, B., and P. Zweigel (2010), A plea for more caution in faultslip
j.tecto.2006.12.007. analysis, Tectonophysics, 482, 2941, doi:10.1016/j.tecto.2009.07.019.
Norabuena, E., L. Leffler-Griffin, A. Mao, T. Dixon, S. Stein, I. S. Sacks, Tibaldi, A., C. Corazzato, and A. Rovida (2009), MioceneQuaternary
L. Ocala, and M. Ellis (1998), Space geodetic observations of Nazca structural evolution of the UyuniAtacama region, Andes of Chile and
South America convergence along the central Andes, Science, 279, Bolivia, Tectonophysics, 471, 114135, doi:10.1016/j.tecto.2008.09.011.
358362, doi:10.1126/science.279.5349.358. Turner, F. J. (1953), Nature and dynamic interpretation of deformation
Oncken, O., D. Hindle, J. Kley, K. Elger, P. Victor, and K. Schemmann lamellae in calcite of three marbles, Am. J. Sci., 251, 276298,
(2006), Deformation of the Central Andean upper plate systemfacts, fiction doi:10.2475/ajs.251.4.276.
and constraints for plateau models, in The Andes: Active Subduction Orog- Turner, J. C. M. (1960), Estratigrafa de Nevado de Cachi y sector al oeste
eny, edited by O. Oncken et al., pp. 327, Springer, Berlin, doi:10.1007/ (Salta), Acta Geol. Lilloana, 3, 191226.
978-3-540-48684-8_1. Viramonte, J. G., R. H. Omarini, V. Araa Saavedra, and A. Aparicio,
Ortner, H., F. Reiter, and P. Acs (2002), Easy handling of tectonic data: L. Garcia Cacho, and P. Parica (1984), Edad, genesis y mecanismos de
The programs Tectonics FP for Mac and Tectonics FP for Windows, erupcion de las riolitas granatiferas de San Antonio de los Cobres, Provin-
Comput. Geosci., 28, 11931200, doi:10.1016/S0098-3004(02)00038-9. cia de Salta, in Actas del IX Congreso Geolgico Argentino, Bariloche,
vol. 3, pp. 216233, Asoc. Geol. Argent., Buenos Aires.

15 of 15

You might also like