Hermanns Strecker GSA Bull 1999

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/316149413

Structural and lithological controls on large quaternary bedrock landsliedes in


NW-Argentina

Article · January 1999

CITATIONS READS

3 33

2 authors:

Reginald L. Hermanns Manfred R Strecker


Norwegian University of Science and Technology Universität Potsdam
222 PUBLICATIONS   3,720 CITATIONS    557 PUBLICATIONS   20,014 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Tectonic and climatic control on the evolution of the intermontane Kangra Basin in NW Himalaya, India View project

TerraceM 2 View project

All content following this page was uploaded by Reginald L. Hermanns on 28 November 2019.

The user has requested enhancement of the downloaded file.


Structural and lithological controls on large Quaternary rock avalanches
(sturzstroms) in arid northwestern Argentina

Reginald L. Hermanns*
Manfred R. Strecker† } Institut für Geowissenschaften, Universität Potsdam, Postfach 601553, D-14415
Potsdam, Germany

ABSTRACT tudes of Ms > 6.0 (e.g., Solonenko, 1977; Adams, 1981; Keefer, 1984;
Nikonov, 1988), often resulting in catastrophic effects for communities
Landsat thematic mapper (TM) analysis, aerial photograph inter- (e.g., Hadley, 1964; Plafker and Ericksen, 1978). However, limited re-
pretation, and field studies of the semiarid Puna Plateau, adjacent gional studies in tectonically or volcanically active regions show that the
Eastern Cordillera, and Sierras Pampeanas of Argentina (lat 24°–28°S) occurrence of large landslides is not random in these environments; in-
have revealed the existence of at least 55 rock-avalanche deposits with stead, the distribution of landslides appears to be influenced by a combi-
volumes larger than 106 m3 that formed by the collapse of entire moun- nation of local geologic, structural, and geomorphic factors that predispose
tain fronts. The spatial distribution of landslide deposits is not random, steep mountain fronts to failure during large seismic events (e.g., Eisbacher,
but it clusters along mountain fronts bounded by active faults. Inspec- 1979; Francis and Wells, 1988; Jibson and Keefer, 1993; Crozier, 1995;
tion in the field reveals five principal controls on the distribution of Tibaldi et al., 1995).
these events. The source area of the rock avalanches has two topo- In semiarid to arid northwestern Argentina, large landslide deposits oc-
graphic constraints: (1) vertical relief contrasts between the breakaway cur along steep mountain fronts of the Puna Plateau, the northern Sierras
zone and the mountain front must exceed a threshold of 400 m, and (2) Pampeanas, and the Cordillera Oriental between 24° and 27º 30′S (Fig. 1)
the slope inclinations must be steeper than 20°. Rock avalanches are re- (Wayne, 1994). All of these bedrock landslides have volumes in excess of
stricted to three types of lithology: granites, low-grade metamorphic 106 m3, and the majority of the lobate deposits result from highly mobile
rocks, and coarse clastic sediments. Structural controls are very im- rock avalanches, defined as sturzstroms (e.g., Heim, 1932; Hsü, 1975). The
portant. Rock avalanches are controlled by planar structures such as landslide deposits cluster along steep mountain fronts with documented
bedding planes, exfoliation joints, minor faults, and cleavage that all Quaternary tectonic activity and are readily recognized on thematic mapper
dip toward the valley. In addition, major slide clusters occur along (TM) satellite imagery due to the sparse vegetation cover. The large number
mountain fronts that underwent Quaternary reverse-fault reactivation of landslides occurring along neotectonic mountain fronts with intrinsically
of former transfer faults with strike-slip kinematics. The trigger mech- different lithologic and structural conditions and a relatively homogenous
anism for the majority of these landslides is interpreted to be seismic, climate make this region a perfect site for a comprehensive analysis of the
although the ages of some major slides are about 30 ka, and they may controls on landslide distribution in an active orogen.
correspond to a more humid interval in southern South America.
GEOLOGIC AND TECTONIC SETTING
INTRODUCTION
The southern Puna Plateau, the northern Sierras Pampeanas, and the
The stratigraphic record and geomorphology of high mountains often Cordillera Oriental coincide with a seismic transition zone between a nearly
record mountain-front collapse coupled with catastrophic rock avalanches flat segment of the subducted Nazca plate to the south of 28°S and a 30°
that may exceed volumes of 106 m3 (e.g., Voight and Pariseau, 1978; dipping segment north of 24°S, and are characterized by moderate seismic-
Shaller, 1991). Such large landslides can be triggered by a variety of fac- ity (Bevis and Isacks, 1984; Cahill and Isacks, 1992) (Fig. 1). The border of
tors, including strong and/or long-lasting rainfalls, rapid snow melts, ef- the low-lying Andean foreland is defined by the meridianally oriented west-
fects of deglaciation on pore pressure, and volcanic eruptions (e.g., Heim, ern Sierras Pampeanas and the Cordillera Oriental, which reach altitudes be-
1932; Porter and Orombelli, 1981; Siebert, 1984; Hewitt, 1988), or a com- tween 3000 and 5500 m. These ranges divide the Argentine northwest into
bination of different factors (see Voight, 1978; Eisbacher and Clague, an eastern humid region characterized by high rainfall and lush vegetation,
1984). In addition, in mountain belts characterized by ongoing tectonism, and a western intramontane arid province receiving <200 mm/yr precipita-
rock avalanches are most frequently triggered by strong earthquakes (e.g., tion (Bianchi and Yañez, 1992).
McSaveney, 1978; Evans et al., 1987; Schuster et al., 1992) with magni- The southern part of the Puna Plateau and the northern Sierras Pam-
peanas are defined by reverse-fault–bounded ranges and intervening basins
*E-mail: reginald@geo.uni.potsdam.de. (Allmendinger, 1986; Allmendinger et al., 1989). On average, ranges in the
†E-mail: strecker@rz.uni-potsdam.de. southern Puna Plateau are 20–30 km long and as high as 6000 m; the Sier-

GSA Bulletin; June 1999; v. 111; no. 6; p. 934–948; 14 figures; 1 table.

934







""
##



 ,,








!







,,




QUATERNARY ROCK AVALANCHES, NORTHWESTERN ARGENTINA

Peru
Bolivia
14˚S
gionally consistent and characterized by late Cenozoic northwest-southeast
to west-northwest–east-southeast shortening, followed by northeast-south-
west–oriented shortening (Allmendinger, 1986; Marrett, 1990; Grier et al.,






""
##



 









!







,,




200 km 1991; Marrett et al., 1994). The second, still-active kinematic regime is

Pacific Ocean
Chile
Co much more variable and involves strike-slip and normal faulting at the
rd southern Puna Plateau margin and coeval thrusting in the Sierras Pampeanas
ill
er
aO Argentina and the Cordillera Oriental (Allmendinger et al., 1989; Marrett, 1990; Grier
Altiplano rie et al., 1991). Most range fronts are therefore oblique or perpendicular with
nt
al respect to the current shortening direction (Marrett and Strecker, 1997).
Tr
en






""
##



 




!







,,




Historic and instrumentally recorded accounts of crustal seismicity for
ch

the southern Puna Plateau and its margin are rare. Destructive historic earth-
Axi

quakes occurred in northwestern Argentina farther east in the Santa Barbara


s

inas
ranges at about long 65° W; another event of inferred magnitude Mb 6 oc-

ba nd
curred at the village of La Poma, at about 24°43′S, 66°13′W, in 1930 (Cahill
Nazca as Su et al., 1992). In 1973 a Mb 5.8 earthquake with a thrust mechanism occurred

!



""
## 




Plate at 25°98′S, 67°72′W immediately west of the study area in the southern
Sierr

Puna Plateau (#68 of Chinn and Isacks, 1983), and in 1986 a Mb 4 event
Santa was recorded at Villa Vil at the southern Puna Plateau margin (Assumpção
Barbara and Araujo, 1993). The kinematics of these events are consistent with the
kinematics of the younger deformation regime derived from fault-slip
ipal

Puna Salta analysis.


ra Princ

Cafayate
IDENTIFICATION, TYPES, AND REGIONAL DISTRIBUTION
OF LANDSLIDE DEPOSITS
Cordille

Tucumán
For the purpose of this investigation only deposits of large bedrock fail-
ontal
Precordillera

ures with volumes in excess of 106 m3 were included; these deposits and the
Cordillera Fr

Sierras corresponding breakaway zones are large and well preserved enough to be
Pampeanas identified on satellite images. Landslide deposits were first located by ana-
30˚S
74˚W 62˚W lyzing Landsat TM scenes 231-77, 231-78, 231-79, 232-77, and 231-78, us-
ing spectral bands 5, 4, and 2. Landslide deposits in piedmont regions were
Figure 1. Tectonic provinces of the Central Andes in Argentina, Bo- easily identified due to their lobate morphology with frontal rims and lateral
livia, and Chile, modified after Jordan et al. (1983). Gray frame outlines levees enclosing inner depressions. In narrow intramontane valleys, moun-
area shown in Figure 2. tain fronts were scanned for large breakaway zones, massive, chaotic depos-
its on opposite slopes, and evidence for temporary landslide dams. In addi-
tion, the unconformable nature of landslide debris and landslide-related
ras Pampeanas blocks are 60–100 km long and reach elevations of 5500 m. lacustrine sediments are easily identified because of their spectral contrasts.
Although the Puna Plateau has an average elevation of 3600 m and en- In a second step, the inferred landslide origin of these deposits was verified
dorheic basins, the basins of the northwestern Sierras Pampeanas are at on stereopairs of aerial photographs at a scale of 1: 50 000, which further
1500–2500 m elevation and have outlets (Penck, 1920). The ranges in both helped to constrain landslides of different ages based on surface morphology.
provinces mainly comprise late Precambrian–Early Cambrian low- to high- We identified 55 large landslide deposits (Fig. 2, Table 1), and 85% of the
grade metamorphic rocks and Paleozoic granites (González Bonorino, landslide sites were studied in the field. In Figure 2 and in Table 1, white,
1951; Rapela, 1976; Jezek et al., 1985; González et al., 1991; Mon and light gray, and dark gray circles correspond to increasing levels of confidence
Hongn, 1996). Uplift of the western Sierras Pampeanas basement blocks in the identification of these deposits. Dark gray circles represent landslide
started in late Miocene time, accelerated after 4 Ma, and culminated after deposits visited and verified in the field, and light gray circles correspond to
2.9 Ma, when the deposits of the intramontane basins were folded and partly sites with typical landslide morphologies identified on aerial photographs.
overridden by the metamorphic basement blocks (Strecker et al., 1989). White circles represent deposits of inferred landslide origin that were too re-
In contrast, the Cordillera Oriental is a north-south–trending mountain belt mote to visit.
dominated by west-verging thrusts and reverse faults (Mon, 1976; Grier et al., There are two types of large landslide deposits in the study area; rock
1991). The uplifted blocks consist of Paleozoic granites and the low-grade avalanches (sturzstrom) and composite deposits composed of block slides
metamorphic Puncoviscana Formation, which is covered by Cambrian- and related distal flow deposits. The rock-avalanche deposits, including the
Ordovician quartzites and Cretaceous to Eocene carbonates, and sandstones, distal flow deposits of the block slides, show the typical internal fabric and
siltstones, and conglomerates of the rift-related Salta Group (e.g., Salfity and inverse stratification often described in combination with very mobile and
Marquillas, 1994). The mountain ranges are 3000–4000 m high and incorpo- dry bedrock landslides (e.g., McSaveney, 1978; Eisbacher, 1979; Hewitt,
rate narrow, deeply eroded valleys that drain to the eastern foreland. 1988; Yarnold, 1993). The mobility of such avalanches is commonly ex-
Common to all intramontane basins of these geologic provinces are Neo- pressed by an apparent coefficient of friction expressed by the height/dis-
gene sediments and unconformable pediment-cover gravels that record sus- tance ratio (H/L) (e.g., Heim, 1932; Hsü, 1975), where H is the elevation
tained tectonism since late Pliocene time (Allmendinger, 1986; Strecker difference between the breakaway zone and the lowest point reached by the
et al., 1989; Allmendinger et al., 1989; Grier, 1990). avalanche, and L the horizontal distance between the breakaway and the
The late Cenozoic kinematic evolution of these tectonic provinces is re- distal avalanche limit. In accordance with observations by Nicoletti and

Geological Society of America Bulletin, June 1999 935


  




 ,







 
,





,,, 
,








 

















,







,











,,











,







,









,






 




 

67˚W

 
66˚W






 ,













 











,,,



,






,


































, |








,,




































,

 


















 

,













  





 
   















,










,,,



 




































 



,









,
}}}



‚

} ||
{{
zzz
yyy

,,,


,,






























,







}
|
‚



,






 









~
z
y

,


Salta




,


 



 





 















,,, 



‹
‡
†
‚













 









 










,



, 
€€
‚‚‚
‡‡‡

ƒƒƒ
„„„
……
††
ˆˆˆ
‰‰‰
ŠŠ
‹‹
ŒŒŒ
‘‘‘

ŽŽŽ


}}}
||
{{
zzz














yyy

‘
 ~~~


€
{
,,,
,
,,





































 




‡
†
Œ
‹
‘


,








 
ˆ


„
ƒ
‰
Ž




,


25˚S 25˚S



 
,


,



 





,,,
 
,
 
  





 ,









,,, 






















 ,
,
,,







 ~~~

€€
‚‚‚

ƒƒƒ
„„„
……
††
‡‡‡
ˆˆˆ
‰‰‰
ŠŠ
‹‹
ŒŒŒ

‘‘‘
ŽŽŽ


}}}
||
{{
zzz




















yyy




Ž
Š
‰





,,,
,













































,

,,,






,










 








,,, 












 ,













,,




















~~~

€€
‚‚‚

ƒƒƒ
„„„
……
††
‡‡‡
ˆˆˆ
‰‰‰
ŠŠ
‹‹
ŒŒŒ

ŽŽŽ
‘‘‘


}}}
||
{{
zzz


































,,,


yyy




,,,





























,,,


,






     





 ,





 







,,, ,,,,


















,









 










 ,,













 




,






,

,,,
~~~

€€

‚‚‚
ƒƒƒ
„„„
……
††
‡‡‡
ˆˆˆ
‰‰‰
ŒŒŒ
ŠŠ
‹‹

ŽŽŽ
‘‘‘


}}}


||
{{
zzz








yyy


,,,






























 











,,














,,,















,,,


26˚S 26˚S
Cafayate

S a. de Qui l mes





,,



































,,,,























,
























,

 









 ,
,,


















 



,,,













 
{
|
}





y

,





Figure 2. Generalized geologic map of the study


area and distribution of large rock-avalanche depos-
its, marked by white, light gray, and dark gray dots

rto
ue




,
C . Calchaquies
(after Allmendinger et al., 1983; Strecker et al. 1989; eM
















 








 













,,


 
,















 ,



,,,,































,

















 












,,













 


,








,,,













 




y
{
|
}
‚

€
‡
†
…
Œ
‹
Š


‘











 ƒ
ˆ
~

,




ca

Grier et al., 1991). Sa.—Sierra; C.—Cumbres;


br
Lag. Blan

Lag.—Laguna.
Ho



 ,





Sa.

 ,








 




,










,,














, 

 ,





al
Sa.






















 




















,,





,
 
 ,
,,,,


























 








 




 












, 


‘

 

o Re
ng


ha

Tucumán
.C

,






 
















,

Sa

 {|














 
























 




 




27˚S 27˚S
u ij a








nq
A co
Sa.

N
67˚W 66˚W
0 50 km



,
y
KEY
Neogene to Quaternary volcanic
Reverse fault



ˆ
‡
…
rocks and sediments




 ,

Middle to Upper Tertiary Fault (undefined)
sedimentary rocks



,,
 
Syncline

 
,

Cretaceous siltstones and sandstones



Cretaceous conglomerates Anticline
(La Yesera Formation)
Paleozoic sedimentary rocks Rock-avalanche deposit
Paleozoic granites studied in the field
certain
Precambrian to Paleozoic uncertain
metamorphic rocks

936 Geological Society of America Bulletin, June 1999


TABLE 1. MORPHOMETRIC PARAMETERS OF LANDSLIDE DEPOSITS IN NORTHWESTERN ARGENTINA
Name Location Composition Confidence Volume Height Length Ratio
Lat Long level (106 m3) (H) (L) (H / L)
(°S) (°W) (m) (m)
Sierra Aconquija
Loma d. Aspereza 27°10′ 66°09′ Granite/schist l 62 1000 7000 0.14
Aval. d. Zarzo I 27°11′ 66°11′ Granite l 37 900 6500 0.14
Aval. d. Zarzo II 27°11′ 66°11′ Granite l 15 650 4000 0.16
Aval. d. Zarzo III 27°11′ 66°12′ Granite l N.C.* N.M.† N.M. N.C.
Loma Redonda 27°12′ 66°12′ Granite/schist l 65 1000 7000 0.14
Sierra de Hualfín
Villa Vil I 27°02′ 66°46′ Volcanic breccia l 184 300 2500 0.12
Villa Vil II 27°03′ 66°48′ Volcanic breccia l 247 350 2250 0.16
Villa Vil III 27°04′ 66°48′ Volcanic breccia l 375 400 2500 0.16
Villa Vil IV 27°04′ 66°49′ Volcanic breccia l 243 400 2750 0.15
Villa Vil V 27°07′ 66°48′ Volcanic breccia l 34 400 1500 0.27
Villa Vil VI 27°09′ 66°51′ Volcanic breccia l 15 250 1000 0.25
Cumbres Calchaquies
N.N.§ 26°45′ 65°45′ Granite l N.C. 1000 4750 0.21
N.N. 26°41′ 65°45′ Schist/granite l N.C. 1100 5250 0.21
Sierra Chango Real
Rincón Ruins I 26°38′ 66°24′ Granite l 49 800 5500 0.15
Rincón Ruins II 26°38′ 66°25′ Granite l 3 N.M. N.M. N.C.
N.N. 26°40′ 66°26′ Granite l N.C. N.M. N.M. N.C.
Sierra Laguna Blanca
Boulder A 26°25′ 67°09′ Granite ⊗ N.C. 400 2500 0.16
Boulder B 26°26′ 67°09′ Granite l 152 400 4000 0.10
SLB V 26°30′ 67°10′ Phyllite/granite l N.C. N.M. N.M. N.C.
SLB IV 26°30′ 67°11′ Phyllite/granite l N.C. N.M. N.M. N.C.
SLB VI 26°30′ 67°09′ Phyllite/granite l N.C. N.M. N.M. N.C.
SLB VII 26°30′ 67°10′ Phyllite l N.C. N.M. N.M. N.C.
SLB VIII 26°30′ 67°09′ Phyllite l 264 1600 15000 0.11
SLB III 26°31′ 67°10′ Phyllite l N.C. N.M. N.M. N.C.
SLB II 26°31′ 67°10′ Phyllite l N.C. N.M. N.M. N.C.
SLB I 26°31′ 67°11′ Phyllite/granite l N.C. N.M. N.M. N.C.
Cerro Paranilla
N.N. 26°06′ 65°46′ Granite l 23 500 2200 0.23
N.N. 26°04′ 65°43′ Phyllite l N.C. 600 N.M. N.C.
N.N. 26°05′ 65°44′ Phyllite l 32 700 3500 0.20
Sierra Carahuasi
N.N. 26°04′ 65°43′ Conglomerate l N.C. 850 N.M. N.C.
N.N. 26°05′ 65°44′ Conglomerate l 143 700 3750 0.19
N.N. 26°05′ 65°44′ Conglomerate l 54 700 3100 0.23
Sierra de Vazquez
N.N. 25°59′ 66°36′ Granite m N.C. N.M. N.M. N.C.
Cerro Zorrito
El Paso I 25°59′ 65°45′ Conglomerate l 210 1100 4750 0.23
El Paso II 25°59′ 65°45′ Conglomerate l 225 700 2850 0.25
Casa de los Loros I 25°58′ 65°45′ Conglomerate l N.C. N.M. N.M. N.C.
Casa de los Loros II 25°58′ 65°45′ Conglomerate l 163 1100 3500 0.31
N.N. 25°49′ 65°49′ Conglomerate l N.C. N.M. N.M. N.C.
N.N. 25°54′ 65°49′ Conglomerate ⊗ N.C. N.M. N.M. N.C.
Tonco syncline
I 25°34′ 65°58′ Conglomerate l 70 650 2300 0.28
II 25°34′ 65°58′ Conglomerate l N.C. 650 1500 0.43
III 25°38′ 65°59′ Conglomerate l N.C. N.M. N.M. N.C.
IV 25°38′ 65°59′ Conglomerate l N.C. N.M. N.M. N.C.
Cerro Abra Blanca 25°20′ 65°57′ Conglomerate ⊗ N.C. 500 4000 0.13
Cerros La Laguna
Brealito 25°17′ 66°21′ Conglomerate l 30 700 2250 0.31
Cumbres de Brealito N.C. N.M. N.M. N.C.
N.N. 25°09′ 66°22′ Granite m N.C. N.M. N.M. N.C.
Cordón Abra Blanca
N.N. 25°08′ 67°13′ Siltstone (met.) ⊗ N.C. 600 3250 0.18
N.N. 25°08′ 67°13′ Siltstone (met.) ⊗ N.C. 600 3500 0.17
Quebrada Las Arcas
N.N. 24°50′ 65°48′ Cretaceous m N.C. 400 N.M. N.C.
La Poma
N.N. 24°45′ 66°11′ Phyllite l N.C. N.M. N.M. N.C.
Quebrada del Toro
N.N. 24°35′ 65°52′ Phyllite l N.C. N.M. N.M. N.C.
N.N. 24°36′ 65°50′ Phyllite l N.C. N.M. N.M. N.C.
N.N. 24°36′ 65°50′ Phyllite l N.C. N.M. N.M. N.C.
N.N. 24°40′ 65°49′ Phyllite l N.C. 1300 N.M. N.C.
N.N. 24°35′ 65°48′ Phyllite l N.C. 550 N.M. N.C.
N.N. 24°36′ 65°49′ Phyllite l N.C. N.M. N.M. N.C.
Note: Symbols m, ⊗, and l denote the level of confidence for landslide identification: low (m), medium (⊗), and high (l); met.—
low-grade metamorphic rocks.
*N.C. = not calculable.
†N.M. = not measurable.
§N.N. = no name.

Geological Society of America Bulletin, June 1999 937


HERMANNS AND STRECKER

Alluvial-fan deposits 66˚12′W

Rock-avalanche deposit

Granite A 2700
27˚09′S
Loma de la
Metamorphic rock Aspereza El Cerillo


oE
Ce

l
Reverse fault rillo

Figure 3. Generalized geo- Breakaway zone Buey


logic map of the southwestern Muerto 2900
sector of Sierra Aconquija, Movement direction
northern Sierras Pampeanas, Avalancha
2700 Elevation in meters
showing the localized occur- del Zarzo
rence of granitic sturzstrom I
3960
deposits (after Fauque and

Si erra Aconquija
Strecker, 1988). II A′
Loma III R
ío Z rzo
Redonda a
N
3


c. 6

oP
o tr
Na

e ril
lo s
ta
Ru

0 2 km

Sorriso-Valvo (1991) elsewhere, unconfined rock avalanches in north- REGIONAL EXAMPLES OF LARGE LANDSLIDES
western Argentina have lower H/L values (0.1–0.16) than obstructed rock
avalanches, which have higher H/L ratios (0.2–0.43). The avalanches pre- Southwestern Sierra Aconquija Slides (27°10′S, 66°15′W)
sented in Table 1 are all characterized by a high degree of mobility.
All of the documented landslide deposits are in the arid valleys of north- The 100-km-long Sierra Aconquija is a basement block composed of bi-
western Argentina. Deposits of this type and comparable volume were not otite schist, augen gneiss, and granite. Uplift has occurred along steep re-
found along the steep and heavily vegetated eastern slopes of the Cordillera verse faults on the west side of the range, which has been active throughout
Oriental and the Sierras Pampeanas. Due to the low erosion rates on inac- Quaternary time (Strecker et al., 1989). Despite high relief contrasts of
tive alluvial-fan–covered pediments in the arid valleys and the resulting high 2500–3000 m along the entire range front, rock-avalanche deposits only oc-
preservation potential, we are confident that the majority of the important cur in the areally limited granitic sector in the southwestern part of the range
bedrock landslides were found. Exceptions are the narrow steep-sided val- (Fig. 3), where at least eight Quaternary rock avalanches were generated
leys in the Cordillera Oriental that underwent pronounced erosion due to lat- (Fauque and Strecker, 1988); the five best–preserved examples are shown
eral undercutting; mappable landslide deposits there may only represent the in Figure 3. The lobate sturzstrom deposits are 4–5 km away from the
youngest events in these dynamic environments. mountain front on the fan-covered piedmont and are characterized by lateral
In this study we show that the distribution of the landslides is not random. levees and frontal rims that are several meters higher than the inner parts.
Instead, landslide deposits occur in clusters along mountain fronts with high The internal fabric of the deposits is dominated by large angular blocks
relief contrasts and documented neotectonic activity (Fig. 2). All breakaway ranging from several meters to 20 m in the upper part; the basal layers con-
zones are in the hanging wall of important Neogene thrust faults, which sist of finer grained material ranging from powder-size to entirely fractured
were repeatedly reactivated during Quaternary time. Furthermore, the land- blocks, <1 m in diameter. The lithologic composition of the deposits is es-
slide deposits are spatially restricted to granites, low-grade metamorphic sentially monomictic and consists of granite and a small percentage of other
rocks, and coarse clastic sediments with pronounced planes of weakness. In rock types derived from the contact zone between the granite and the biotite
order to illustrate the complex relations between different conditions that fa- schist. On the basis of landslide and alluvial-fan morphology, the deposits
vor mountain front collapse and large-volume bedrock landslides, repre- are inferred to be late Pleistocene in age. The five avalanche deposits have
sentative examples of well-developed and preserved deposits from typical calculated volumes between 5 and 65 × 106 m3. The breakaway zones in the
structural or lithologic settings are discussed in the following sections. granite are strongly conditioned for failure by exfoliation joints that dip

938 Geological Society of America Bulletin, June 1999


QUATERNARY ROCK AVALANCHES, NORTHWESTERN ARGENTINA

B A

‘‘










}}
||
‚‚

††
‡‡
‹‹
ŒŒ
 

,,












y
{
z
~
€

ƒƒ
„
…
ˆˆ
‰
Š

Ž
 
y
~
Breakaway scarp 4500
A









||
{{
€€

ƒƒ
„„
‡‡
ˆˆ 
,,






y
zz
}
~~

‚‚
……
†† 
y
}
A



} 
||
‚ 

,



{
z
y
~

€





††
‡‡
‹‹
ŒŒ

‘‘









ƒƒ
„
…
ˆˆ
‰
Š

Ž

4000 Figure 4. Profile A–A′ of the
Loma de la Aspereza rock
avalanche in front of Sierra




ƒƒ
„„
‡‡
ˆˆ {
z







‚‚
……
††
Aconquija (see Fig. 3 for loca-
3500 3500 tion). Also shown are orienta-

Ž‰Šƒ„…€~
€





‚
‡ˆ
††
Œ
‘‹‹
tions of exfoliation joints (A) and
schistosity (B) as great circles
(lower hemisphere projection).
3000 Rock-avalanche deposit 3000
Alluvial-fan deposits
Granite (with
exfoliation joints)
Metamorphic basement
1 km with pegmatitic dikes
2500 Schistosity
2500






,,,,,,,
20°–50° toward the piedmont (Fig. 4). In contrast, schistosity in the biotite
schist, which also presents potential planes of weakness, is much more vari-









,,,,,,,
able in its orientation; it is folded and dips primarily toward the mountain
66˚24 W (Fig. 4). As reflected in the lithologic composition, metamorphic rocks in
the contact zone with the granite are involved in mountain-front collapse









,,,,,,,
only where they were intruded by pegmatitic dikes parallel or oblique to
eal

schistosity (Fig. 4).


o R

X1
N
ang









,,,,,,,
El Rincón Slides (26°37′S, 66°25′W)
Ch

26˚37 S
a
err

The two El Rincón sturzstrom deposits occur in front of Sierra Chango


Si

Real at the southeastern edge of the Puna Plateau (Fig. 5). The source rock









,,,,,,,
for these landslides is also granitic basement, which is in reverse-fault con-
tact with folded Neogene conglomerates. The east-vergent reverse fault was
reactivated as a dextrally oblique normal fault during late Quaternary time

,







,,,,,,,
X2
(Allmendinger, 1986). The granite is fractured by minor faults dipping
Sediments 50°–80° west. The rocks are further weakened by northeast-trending ex-
foliation joints that separate banks, 1–3 m thick, and dip 45° southeast, i.e.,









,,,,,,,
Young rock-avalanche deposit
parallel to the topography of the mountain front and the principal sliding di-
Old rock-avalanche deposit
rection. Although the breakaway zone of the older slide cannot be defined
Granite unambiguously due to post-landslide erosion, the collapse is inferred to









,,,,,,,
R ío R

have taken place in the headwaters of the Rincón valley (Fig. 5); the land-
Breakaway zone
slide mass with a volume of 75 × 106 m3 streamed downhill along the course
cóin

of the valley and was confined by the steep valley sides. Closer to the moun-
n









,,,,,,,
Thrust fault
tain front, the mass ran up the northern slope in a sharp curve, incorporated
0 1 2 km
Movement direction Neogene basaltic gravel, and continued streaming for 2.5 km into the pied-
exfoliation minor exfoliation minor mont, attesting to its high mobility. The younger avalanche was deposited









,,,,,,,
joints faults joints faults
X1 X2 within the levees of the older landslide; it has a similar block size and a vol-
ume of 3 × 106 m3. This deposit may represent material remobilized from
the first landslide, which deposited material at a higher position. Erosional
remnants of older granitic rock-avalanche deposits are present 12 km south
a) b) a) b)
n=6 n=5 n=8 n=8 along the mountain front (Fig. 2).

Figure 5. Simplified geologic map of the northeastern portion of Sierra Laguna Blanca Slides (26°35′S, 67°15′W)
Sierra Chango Real at the southeastern edge of the Puna Plateau. Ex-
foliation joints and minor faults are represented as great circles (lower At least nine sturzstroms are in front of the Sierra Laguna Blanca (Fig. 6).
hemisphere projection). This 6000-m-high basement block comprises Precambrian low-grade

Geological Society of America Bulletin, June 1999 939


67˚10′W Alluvium
Rock-avalanche deposits
Pleistocene terrace- and
boulder deposits
Neogene sediments
Granite
Metamorphic rocks
N
4420 Breakaway zone
4030

Reverse fault

Movement direction

4420 Elevation in meters


0 1 2 3 4 km

Figure 6. Geologic map of the


western Sierra Laguna Blanca in

Sierra Laguna Blanca


the eastern Puna Plateau (after
González et al., 1991). Orienta- 87
24
tions of minor faults, exfoliation
joints, and cleavage are repre-
sented as great circles (lower
hemisphere projection); asterisk
denotes locality of measurements. V VI
IV
VII
26˚30′S 20 *
5000
III VIII 4350
3600
I II

5946

n=4 n = 10 n = 12
sf1
sf2
5700
minor faults exfoliation joints cleavage

Figure 7. Planar surfaces in the metamorphic basement rocks of


the western Sierra Laguna Blanca. The exact position of the photo
corresponds to the asterisk in Figure 6. Triangles represent minor
faults, large arrows signify exfoliation joints, the dip symbol corre-
sponds to penetrative cleavage (sf1), and the small arrows show the
minor cleavage (sf2).

1m

940 Geological Society of America Bulletin, June 1999


QUATERNARY ROCK AVALANCHES, NORTHWESTERN ARGENTINA

65˚50′W 65˚45′W

u lt
fa
Rio
ta
go

To
ó

l
ro
r G
za

a
mp
a
oL
Est. Solá Ri

N
o

0 2 km
l Ch o r
Sierr a de

Si
err
a d
Figure 8. Generalized geologic

e P
map of the upper Quebrada del
Toro in the Cordillera Oriental

asch
showing distribution of multiple

a
rock-avalanche deposits related
to mountain-front collapse in
s low-grade metamorphic base-
rca
A

ment (after Schwab and Schäfer,


las

1976; Marrett, 1990; Strecker


de

Alluvium o
Ri and Marrett, 1999). Co.—Cerro;
Rock-avalanche deposit 2900 Est.—Estación.
Inferred landslide deposit 24˚40′S
Lake deposits

Tertiary sedimentary rocks

Cretaceous sedimentary rocks Gólgota


t
a

o
Cambrian to Ordovician G ólg
R io
quartzites 3500
Precambrian
metamorphic rocks
2515
Breakaway surface

Reverse fault
o
Rio Co. B ay
Movement direction
3500 Elevation in meters

metamorphic rocks in the south and Paleozoic granites in the north Allmendinger et al., 1989), tectonism along this mountain front during the
(González et al., 1991). In the contact zone the metamorphic rocks are in- current kinematic regime is documented by normal faulting in conjunction
truded by pegmatitic dikes. On the west the range is bounded by east- with strike-slip faulting with ~12 m offset.
dipping reverse faults, which also offset unconformably overlying Tertiary The northern granitic sector of the range generated two rock avalanches
conglomerates and sandstones, which dip as much as 20° west (González by the collapse of an uplifted unconsolidated boulder deposit containing in-
et al., 1991). dividual blocks that range from 5 to 10 m in size. In front of the southern
The metamorphic basement rocks are characterized by a penetrative part of the range eight landslide deposits can be distinguished due to the su-
schistosity dipping 40°–80° northeast to east, a second minor schistosity perposition of lateral levees and frontal rims; these deposits have been dated
dipping 15° southeast to east, exfoliation joints dipping 45°–70° west to with the 21Ne-cosmogenic-nuclide method, resulting in ages between
west-northwest, and minor thrusts with 55°–70° west-southwest–dipping 139 +16/–22 ka and 389 +17/–24 ka (Hermanns, 1999). The uppermost
fault planes (Figs. 6 and 7). Similar to other areas in the Puna Plateau (e.g., parts of the breakaway zones are at 5700 m altitude, whereas the distal seg-

Geological Society of America Bulletin, June 1999 941





,,,,,







,,,,











yyyy
zzz
{{{{
||||

€€€€
}}}}
~~~

ƒƒƒƒ
„„„„
‚‚‚
HERMANNS AND STRECKER





,,,,,









 ,,,,















yyyy
{{{{
||||

€€€€
zzz
}}}}
~~~
ƒƒƒƒ
„„„„

‚‚‚
…………
‡‡‡‡
ˆˆˆˆ
†††





,,

65˚45 W
Casa de los Loros








,,,










 


,,,,,









 ,,,,













{{{{
||||
yyyy
zzz
}}}}
~~~

€€€€
ƒƒƒƒ
„„„„

‚‚‚
‡‡‡‡
ˆˆˆˆ
…………
†††
,,,,,











,,



N
I
Co. Zorrito
Pica ch o

3224 ss 105 / 34
fau II
lt

,  
,,

C ss 223 / 21











 



 ,,,,














{{{{
||||
yyyy
zzz
}}}}

€€€€
~~~

‚‚‚
ƒƒƒƒ
„„„„
‡‡‡‡
ˆˆˆˆ
…………
†††
,,,,,

















,,


ha
cras

 ,,
, 
1500

fault
lt
El Z orr it o fa u

A B





,,,,,


















,
II

,,



I
as
nc h

,,

,,,

Co El Paso
s
La Rí
o La Y
Río

esera

,,,







26˚S

,,,




, …
ˆ

Alluvium 0 1 2 km A B Movement direction
3224 Elevations in meters
II Young rock-avalanche deposit
Thrust
Lake sediments n = 21

,
n = 20

Old rock-avalanche deposit Cretaceous siltstones and sandstones


I (Las Curtiembres Formation)

 Terrace gravel
Tertiary sediments
Cretaceous conglomerates
(La Yesera Formation)

Figure 9. Geologic map of the southwestern part of Rio de las Conchas valley (after Torres, 1985; Grier et al., 1991) with minor fault-plane
analysis represented as pseudofault-plane solutions (after Grier, 1990); gray pattern corresponds to extension of old kinematic regime, black rep-
resents extension of young kinematic regime. The aerial photograph corresponds to the same area shown in the map. Orientations of landslide
breakaway surfaces and planes of movement are also shown; extension fractures and exfoliation joints are represented in lower hemisphere pro-
jection. Co.—Cerro.

ments of the lowest deposit are at 3700 m and as much as 10.4 km away rounded by 3000–4000-m-high reverse-fault–bounded mountain ranges com-
from the range front. The distance between these locations is 20 km. Due to posed of the metamorphic Precambrian to Cambrian Puncoviscana formation
the superposition it is not possible to quantify volumes for most of these de- (Omarini, 1983). The crystalline basement is overlain by marine Cambrian-
posits. The surfaces of the landslides are mainly composed of low-grade Ordovician strata, and Cretaceous to lower Tertiary carbonates and sandstones
metamorphic rocks and small amounts of granite. The avalanche surfaces in (Reyes and Salfity, 1973). The valley fill is dominated by thick coarse-clastic
the southern half are entirely composed of metamorphic rocks or contain Neogene to Quaternary sediments, which were overthrusted by basement af-
only minute amounts of granite. The relative chronology of these events is ter 0.98 Ma (Schwab and Schäfer, 1976; Marrett et al., 1994).
reflected by the abundance and size of surficial avalanche clasts. Due to in- On the west side of the valley all landslides were generated along the
tense weathering and abrasion the older deposits contain fewer clasts, steep mountain front of Sierra del Choro, which is bounded by a steeply
whereas younger deposits have larger and more abundant clasts. In contrast west dipping reverse fault and subparallel faults that separate metamorphic
to the sturzstroms of purely granitic composition, clast size never exceeds basement and overlying Cretaceous carbonates from the deformed Neogene
1.5 m in diameter in the metamorphic material. Some avalanche rims are and Quaternary units. The massive landslide at Río Gólgota also incorpo-
eroded and expose the upper parts of the internal fabric, characterized by a rates the Neogene coarse clastic rocks from the steep footwall region in the
fractured matrix with blocks to 1 m in diameter and typical of the “jigsaw” basal layer. This inverted stratigraphy is characterized by a basal pulverized
breccias of dry rock avalanches (e.g., Shreve, 1966). reddish material derived from Neogene sediments and overlying metamor-
phic basement fragments with typical jigsaw-puzzle features.
Quebrada del Toro Slides (24°40′S, 65°50′W) On the opposite side of the valley the landslides are limited to the areas
characterized by the foliated basement rocks. Numerous landslide scars in
Numerous landslides exist in the Quebrada del Toro (Fig. 8). This north- the southeastern sector of Sierra de Pascha attest to important landslide ac-
west-trending longitudinal valley is at about 2000 m elevation and is sur- tivity along this sector of the mountain front, which is affected by constant

942 Geological Society of America Bulletin, June 1999


QUATERNARY ROCK AVALANCHES, NORTHWESTERN ARGENTINA

The conglomerates dip 34° east on the east side of the mountain, and 21°
west on the west side. Additional planar structures are 70°–89° southwest-
dipping extension fractures and 40°–75° southeast-dipping exfoliation
joints that are well developed on the southern Cerro Zorrito slopes (Fig. 9).
Two rock avalanches formed on each of the west, south, and east sides of
the mountain. The deposits from the east and south side dammed the Las
Conchas valley, impounding lakes with surface areas >600 km2, which led
to the deposition of laminated buff colored lacustrine sediments. These
landslide deposits have volumes of as much as 2.25 × 108 m3. The avalanche
deposits on the west side occur in the piedmont regions and have no strati-
graphic relation to the lake sediments.
The avalanche debris of the first event (El Paso I) crops out west of the
Río Las Conchas up to 270 m above the present river course; the base of the
deposit is locally exposed by erosion. The contact with the underlying Ter-
tiary and Quaternary strata is marked by a pulverized horizon, which is
overlain by a 2-m-thick layer of heavily fragmented and sharp-edged mate-
rial in coarse- to medium-grained sand fraction. Toward the top, fractured
clasts of varying size occur in a fine-grained matrix of broken material and
become increasingly abundant. This matrix-supported basal section grades
into a clast-supported layer of jigsaw breccia of internally broken clasts. The
top of the deposit is characterized by blocks of up to 20 m in diameter. These
chaotic deposits are unconformably overlain by varved lake sediments in
upstream and downstream directions, attesting to an additional coeval or
younger landslide-generated dam farther downstream; remnants of these
sediments (Casa de los Loros slide I) are still exposed below younger rock-
avalanche debris at Casa de los Loros (Fig. 9).
The El Paso II debris was generated in the second event along the south
face of Cerro Zorrito, where southwest-dipping extension fractures and 45°
southeast-dipping exfoliation joints acted as breakaway surfaces. This de-
posit overlies both the older rock-avalanche (El Paso I) and the lake deposits,
and has the same internal fabric as the older avalanche (Fig. 10). In contrast
to the older deposits, however, typical sturzstrom morphology is still pre-
served. The lake sediments underneath the avalanche contact are folded or
mixed with landslide debris; locally, lake sediments form 13-m-high injected
dikes (Fig. 10) similar to phenomena associated with landslide events de-
scribed by Yarnold (1993) and Friedmann (1997). The dikes and folds docu-
Figure 10. Lake sediments injected into the younger rock-avalanche ment the unconsolidated nature of the lake sediments during this landslide
deposit; direction of landslide movement is top to the left. event. On the basis of stratigraphic relationships, the second rock avalanche
is inferred to have been generated simultaneously with an avalanche on the
east side of Cerro Zorrito, which resulted in a deposit with a volume of
erosion and undercutting by the Río Toro. This region also coincides with a 1.63 × 108 m3 (Casa de los Loros II). At Casa de los Loros (Fig. 9) the debris
strongly fractured sector of basement that was uplifted due to Neogene- of that slide created a natural dam, 220 m high, which caused the river to be
Quaternary reverse-fault reactivation of a former transfer fault that kine- dammed up to the 1700 m contour. During this event, the avalanche over-
matically linked the Gólgota fault to another thrust fault to the south topped a 340-m-high ridge on the opposite slope of the valley and continued
(Strecker and Marrett, 1999). traveling into a small valley east of the ridge. Using this run-up distance and
the calculation principles outlined by Crandell and Fahnestock (1965), a ve-
Cerro Zorrito Slides (25°58′S, 65° 45′W) locity of 294 km/h can be derived for this landslide, which was generated
along a 34° east-dipping bedding plane in the Cretaceous conglomerates. A
Torres (1985) and Gallardo (1988) described large landslide deposits and minimum age of 28 990 ± 150 yr B.P. can be defined for older landslide by
associated lacustrine sediments in front of Cerro Zorrito (3224 m) in the accelerator mass spectrometry 14C dating of fresh-water gastropods from
Quebrada La Yesera, which is located at the eastern outlet of the Calchaquí lake sediments upstream from the landslide (Trauth and Strecker, in press).
Valley (Río de las Conchas) at about 26°S (Fig. 9). Cerro Zorrito consists of Varve counts in the lake sediments (Kleinert et al., 1997) indicate that the du-
Cretaceous conglomerates, which were uplifted during the early kinematic ration of the dams was at least 4 k.y.
phase along the Picacho fault in Neogene time. To the southeast this thrust
was linked to the Chacras thrust fault via a left-lateral transfer fault (El Zor- Tonco (25°30′S, 66°00′W) and Villa Vil Slides (27°05′S, 66°47′W)
rito fault). During the younger kinematic regime the El Zorrito fault was re-
activated as a reverse fault and brought the Mesozoic conglomerates into Four landslides are in the hanging wall of a west-vergent thrust in the
thrust contact with Quaternary piedmont sediments (Grier, 1990). The pre- northern Calchaquí Valley, which places Cretaceous to upper Tertiary rocks
sent relief between the Río de las Conchas and the highest point on Cerro of the Salta Group against Andean foreland-basin strata. Bedrock failure
Zorrito is 1500 m. leading to avalanche formation also occurred in the folded Cretaceous con-

Geological Society of America Bulletin, June 1999 943


,,,,,,,,,
yyyyyyyyy
,,,,,,,,,




yyyyyyyyy
~~~~~~~~~
ƒƒƒƒƒƒƒƒƒ
ˆˆˆˆˆˆˆˆˆ

,,,,,,,,,




yyyyyyyyy
~~~~~~~~~
ƒƒƒƒƒƒƒƒƒ
ˆˆˆˆˆˆˆˆˆ

,,,,,,,,,




yyyyyyyyy
~~~~~~~~~
ƒƒƒƒƒƒƒƒƒ
ˆˆˆˆˆˆˆˆˆ

,,,,,,,,,




yyyyyyyyy
~~~~~~~~~
ƒƒƒƒƒƒƒƒƒ
ˆˆˆˆˆˆˆˆˆ

,,,,,,,,,




yyyyyyyyy
~~~~~~~~~
ƒƒƒƒƒƒƒƒƒ
ˆˆˆˆˆˆˆˆˆ

,,,
























HERMANNS AND STRECKER

,,,























,,,,,,,

yyyyyyy
zzzzzzz




zzzzzzzz
{{{{{{{{
||||||||
}}}}}}}}




















zzzzzzzz
{{{{{{{{
||||||||
}}}}}}}}

€€€€€€€€

‚‚‚‚‚‚‚‚
„„„„„„„„
……………………
††††††††
‡‡‡‡‡‡‡‡
‰‰‰‰‰‰‰‰
ŠŠŠŠŠŠŠŠ
‹‹‹‹‹‹‹‹
ŒŒŒŒŒŒŒŒ
ŽŽŽŽŽŽŽŽ


{|€‘‘‘‘‘‘‘‘
,,,,,,,









yyyyyyy
zzzzzzz

~~~~~~~
ƒƒƒƒƒƒƒ
„„„„„„„
ˆˆˆˆˆˆˆ
‰‰‰‰‰‰‰

ŽŽŽŽŽŽŽ






ƒƒƒƒƒƒƒ
„„„„„„„
ˆˆˆˆˆˆˆ
‰‰‰‰‰‰‰

ŽŽŽŽŽŽŽ




















zzzzzzzz
{{{{{{{{
||||||||
}}}}}}}}

€€€€€€€€

‚‚‚‚‚‚‚‚
„„„„„„„„
……………………
††††††††
‡‡‡‡‡‡‡‡
‰‰‰‰‰‰‰‰
ŠŠŠŠŠŠŠŠ
‹‹‹‹‹‹‹‹
ŒŒŒŒŒŒŒŒ
ŽŽŽŽŽŽŽŽ


‘‘‘‘‘‘‘‘




















zzzzzzzz
{{{{{{{{
||||||||
}}}}}}}}

€€€€€€€€

‚‚‚‚‚‚‚‚
„„„„„„„„
……………………
††††††††
‡‡‡‡‡‡‡‡
‰‰‰‰‰‰‰‰
ŠŠŠŠŠŠŠŠ
‹‹‹‹‹‹‹‹
ŒŒŒŒŒŒŒŒ
ŽŽŽŽŽŽŽŽ


‘‘‘‘‘‘‘‘




















zzzzzzzz
{{{{{{{{
||||||||
}}}}}}}}

€€€€€€€€

‚‚‚‚‚‚‚‚
„„„„„„„„
……………………
††††††††
‡‡‡‡‡‡‡‡
‰‰‰‰‰‰‰‰
ŠŠŠŠŠŠŠŠ
‹‹‹‹‹‹‹‹
ŒŒŒŒŒŒŒŒ
ŽŽŽŽŽŽŽŽ


‘‘‘‘‘‘‘‘




















zzzzzzzz
{{{{{{{{
||||||||
}}}}}}}}

€€€€€€€€

‚‚‚‚‚‚‚‚
„„„„„„„„
……………………
††††††††
‡‡‡‡‡‡‡‡
‰‰‰‰‰‰‰‰
ŠŠŠŠŠŠŠŠ
‹‹‹‹‹‹‹‹
ŒŒŒŒŒŒŒŒ
ŽŽŽŽŽŽŽŽ


‘‘‘‘‘‘‘‘


,,








{{
zz
yy
€€

ƒƒ
„„
……
ˆˆ
‰‰
ŠŠ









ƒƒ
„„
……
ˆˆ
‰‰
ŠŠ

ŽŽ

~
65˚58 W

,,,


























{{{{{{
||||||
}}}}}}















{{{{{{
||||||
}}}}}}
‚‚‚‚‚‚

€€€€€€
………………
††††††
‡‡‡‡‡‡
ŠŠŠŠŠŠ
‹‹‹‹‹‹
ŒŒŒŒŒŒ


‘‘‘‘‘‘









………………
††††††
‡‡‡‡‡‡
ŠŠŠŠŠŠ
‹‹‹‹‹‹
ŒŒŒŒŒŒ


‘‘‘‘‘‘




}}
‚‚
‡‡
ŒŒ




|

†
‹



‡‡
ŒŒ
‘‘



†
‹


~
N

 25˚34 S
0 1 2 km

,,,























Tonco Syncline
II
T onco
31
44
I Río 44

,,,























Figure 11. Geologic map of the
Tonco slides in the folded Meso-
zoic to Tertiary rocks in the

,,,























Cordillera Oriental (modified af-



Œ


‰
Ž
ter Grier, 1990).

‘,,,























24 37 38

,,,























III
o
onc

IV Quaternary to Tertiary
Río T

Breakaway surface
sedimentary rocks
Rock-avalanche deposit






Reverse fault
Cretaceous
siltstones and sandstones Syncline
Cretaceous
conglomerates Slide direction

glomerates (La Yesera Formation) that are part of the western limb of an 66°45′W) in 24° northwest-dipping Tertiary breccias in the Villa Vil area
asymmetric syncline (Fig. 11). (Fig. 2) are genetically similar to the Tonco slides. A total of six large land-
Two of the best-preserved avalanche deposits are located within the slides with volumes between 15 and 375 × 106 m3 are along the western
syncline in front of a remarkable breakaway scarp. The base of the break- limb of an asymmetric, southwest-plunging anticline (L. Fanque, 1996, per-
away is 100 m above the valley floor and reaches the 3185-m-high crest sonal commun.).
of the ridge. A block 50 m tall and 1 km long collapsed along the 35° east-
dipping bedding plane (Fig. 12). A west-vergent thrust fault cuts the west- DISCUSSION AND CONCLUSIONS
ern limb of the syncline and places the conglomeratic units against
younger sandstones and siltstones. The trace of the fault coincides with The distribution of large landslides in the northwestern Argentine Andes
the western limit of the breakaway zone. The debris of the collapsed block clearly shows that they were not generated randomly. Instead, large land-
(I) has the same internal structure as the rock avalanches from Cerro Zor- slides are found in clusters along sectors of mountain fronts representing
rito and constitutes a volume of 70 × 106 m3. A second, minor cliff col- distinct topographic, geomorphic, petrologic, and structural characteristics
lapse (II) occurred along the northern breakaway wall of the first event that either individually or in combination favor mountain-front collapse and
and resulted in a lobe being deposited on top of the older avalanche de- the generation of rock avalanches. In addition, all landslides formed in the
posit. Remnants of two additional slides (III and IV) are found to the hanging walls above tectonically active thrust or reverse faults. In the fol-
southwest (Fig. 11). lowing sections we discuss the importance of these factors and evaluate pos-
A number of landslides at the southeastern Puna Plateau edge (27°S, sible landslide-trigger mechanisms.

944 Geological Society of America Bulletin, June 1999


QUATERNARY ROCK AVALANCHES, NORTHWESTERN ARGENTINA

50 m

Figure 12. The bedding-parallel breakaway surface at Tonco slide I, view toward south. Arrow depicts movement direction. White broken line
marks the breakaway surface, which cuts bedding obliquely in the lower sector.

Conditions Controlling the Location of Large Bedrock Failure tential planes for failure and thus correspond to a type A mountain front.
and Rock Avalanches Sheeting in granitic mountainous regions also appears to favor catastrophic
mountain-front collapse in other tectonically active regions (e.g., Plafker
One of the most important controls on cliff collapse in the Argentine and Ericksen, 1978; Fauque, 1987; Topping, 1993; Weidinger et al., 1996;
northwest is lithologic; although the majority of ranges in this region are González Díaz et al., 1997).
composed of a variety of metamorphic rocks, granites, and fine- to coarse- In contrast, areas in Sierra Aconquija dominated by biotite schist were
grained sedimentary units, volumetrically important landslides are re- only involved in the mountain-front collapse where the schistose fabric was
stricted to low-grade metamorphic rocks, granites, and coarse clastic sedi- intruded and weakened by pegmatitic dikes. The orientation of schistosity
mentary rocks. The outcrops of these rock types at landslide sites are did not play a role in this particular example because the folded fabric dips
characterized by inclined planar surfaces that are ideal zones of weakness toward the mountain. Mountain fronts in metamorphic rocks affected by
that enhance weathering and movement of water, and thus present potential large landslides are all of low metamorphic grade, have a well-developed
planes for failure. Based on our synoptic representation of landslides in this crenulation cleavage, and are further characterized by a high level of disin-
region, three types of mountain fronts with inclined planes of weakness can tegration due to pronounced mechanical weathering.
be discriminated (A, B, C, Fig. 13). Type B mountain fronts define a single breakaway surface coinciding with
Type A contains multiple potential breakaway surfaces intersecting each bedding planes in sedimentary rocks that intersect slopes and may be later-
other and the slope of the mountain through a variety of angles. Typically, ally undercut by rivers (Figs. 13 and 14). This type of setting is reminiscent
breakaway zones of type A are 600–1900 m above the valleys and are semi- of the Saidmarreh slide in Iran (e.g., Watson and Wright, 1963), the Gros
circular and amphitheater-shaped, indicating that the collapse did not affect Ventre slide in Wyoming (e.g., Voight, 1978), or landslides in the Canadian
a single continuous block, but a strongly fractured portion of the mountain Rockies in glacially oversteepened valleys (e.g., Eisbacher, 1979). Examples
front that contained multiple preexisting potential sliding surfaces. This is of this type occur on the east side of Cerro Zorrito. The breakaway surfaces
exemplified by the multiple breakaway surfaces at the southern Cerro Zor- are typically 2–4 km long and dip 24°–34° toward the valley. Lateral cutoffs
rito and Laguna Blanca landslides. of the dislodged landslide mass are sharp and have well-defined surfaces. In
A significant relation between important relief contrasts, lithology, and contrast to Gros Ventre–type slides, failure did not occur along strata of sig-
structural susceptibility to mountain-front collapse is found in the granitic nificant permeability contrasts; rather, the sliding surface developed along
terrains of the study area. For example, the 100-km-long mountain front of bedding planes within relatively uniform Cretaceous conglomerates.
Sierra Aconquija is characterized by high relief contrasts and recurrent Qua- Type C mountain fronts are also characterized by breakaway surfaces
ternary reverse faulting. However, large bedrock landslides occur exclu- parallel to one bedding plane (24°–40°) and may reach several kilometers
sively in front of the small, 10-km-wide southern granitic sector. Like other in extent. In contrast to mountain fronts of type B, the breakaway and/or
faulted granitic mountain fronts in the region (e.g., El Rincón slides), the bedding-plane surfaces do not intersect slopes, but contain gently inclined
granites are fractured by surface-parallel exfoliation joints that present po- planar detachment surfaces in the lower portion that cut obliquely through

Geological Society of America Bulletin, June 1999 945


HERMANNS AND STRECKER


, 
,

 , ,,
B  , ,,,, ,,,,
,
,,,,,,,,,,,,,,,,,,,,


 ,


 ,, ,,,, ,,,,


 , ,, ,,,, , ,, ,
,,,,,,,,,,,,,,,,,,,,,
,,C,,,, ,,,,,, ,,,
,,,,,,,,,,,,,,,,,



, 


, ,, ,,,, ,,,,,,
, ,,,, ,,,, , ,,,
,,,,,,,,,,,,||| 

zzz



, ,zz

,,

yy
{{



, ,,
yy
,
,  ,, , ,|||



, Planar surfaces 

zzz
 || 


,

zz
,,

yy
{{
 {{ 


, ,,

y y Figure 14. Combined bedding and sliding planes at the Casa de los
Loros breakaway zone intersecting river bed. Arrow depicts movement
Sliding surfaces direction of landslide.

planes of weakness is especially important where neotectonic mountain-


Figure 13. Model of three principal types of mountain fronts suscep- bounding faults are characterized by inherent structural complexity result-
tible to major bedrock failure and rock avalanching. (A) Multiple po- ing in numerous minor faults and fracture planes.
tential breakaway surfaces intersecting each other and the slope of the
mountain through a variety of angles. (B) Single breakaway surface co- Triggering Mechanisms
inciding with bedding planes in sedimentary rocks that intersect slopes
and maybe laterally undercut by rivers. (C) Breakaway surfaces par- The spatial distribution of landslides in northwestern Argentina shows
allel to one bedding plane that do not intersect with the slope. that major mountain-front collapse only occurs in mountain-front sectors
that are structurally and lithologically predisposed to failure and where large
amounts of strongly fractured bedrock crop out at high elevations above the
bedding (Tonco and Villa Vil slides). These detachment surfaces occur at valley. The prolonged existence of weakened volumes of rock along steep
positions along the mountain fronts above valley bottoms, where, prior to mountain fronts and occasional removal by landsliding appears to be caused
sliding, overlying finer grained strata were partially eroded, but a buttress of by an arid climate and infrequent strong seismicity. There are no accounts in
finer grained sediments remained intact at the lowest point of the slope pro- the scientific literature of deposits of large rock avalanches along the humid
file (Fig. 13). The position of the obliquely oriented detachment suggests a eastern slopes of the study area; neither did our combined analysis of digi-
reduction in rock strength immediately above the buttress due to removal of tal satellite images and aerial photographs reveal such phenomena. Instead,
layers that maintained confining pressures high enough to keep the moun- mass transfer on the eastern vegetated slopes consists of relatively small ro-
tain front stable. We therefore suggest that due to erosional unloading the tational slips or slump earthflows that continuously affect this environment
area above the buttress became the weakest part of the slope and was prone (e.g., Czajka and Vervoorst, 1956). Pronounced debris flows and slumping
to fail during a seismic event. on the eastern slopes are frequent during phases of higher precipitation (e.g.,
Another factor controlling the occurrence of large landslides is the re- Polanski, 1966; Marcuzzi et al., 1994). In contrast, slumping at mountain
peated activity of mountain-bounding faults. Although lithologic and topo- fronts in the arid parts is neither correlated nor documented in historic
graphic factors are necessary for the generation of landslides, large events records and debris flows occur only during unusually strong rainfall events.
documented in this study only occur along mountain fronts that have un- The difference between the frequency-magnitude distributions of landslide
dergone multiple faulting events that have significantly weakened the co- events in the two contrasting climatic regions suggests that unstable moun-
herence of bedrock in the hanging walls. The role of tectonically induced tain fronts in the arid parts exist much longer than their counterparts on the

946 Geological Society of America Bulletin, June 1999


QUATERNARY ROCK AVALANCHES, NORTHWESTERN ARGENTINA

eastern slopes and/or that greater levels of instability are reached. The re- ground motion. In this regard the situation in the southern Puna Plateau, the
duction of frictional resistance and increase in pore pressures resulting from northern Sierras Pampeanas, and adjacent areas may be similar to the south-
precipitation and infiltration appears to be a negligible factor in landslides eastern Sierras Pampeanas, where no significant deformation has been re-
and rock avalanches in the arid western regions. This suggests that these cat- ported in association with historical earthquakes; however, 14C dating of
astrophic mass movements may have been triggered seismically. faulted soil material indicates events with at least 2.1 m of slip during the past
Although the present climate does not appear to contribute to the precon- 1300 yr along a single fault (Costa and Vita-Finzi, 1996).
ditioning of mountain fronts for landsliding, paleoclimatic variability may
have resulted in wetter periods during which thresholds for mountain-front ACKNOWLEDGMENTS
collapse would have been lowered. Trauth and Strecker (in press) dated sed-
iments from landslide-dammed lakes at the Cerro Zorrito and Quebrada del This work is part of the Collaborative Research Center 276 “Deformation
Toro landslide sites at 30, 29, and 26 ka. This interval corresponds to mod- Processes in the Andes” supported by the German Research Foundation. The
erately cold and relatively wet conditions in subtropical and tropical South success of this project was made possible through the support of many friends
America (e.g., van der Hammen and Absy, 1994; Ledru et al., 1996; Turcq and colleagues, namely R. Alonso, A. Bourdin, I. Capdevila, L. Fauque,
et al., 1997). Although systematic and marked changes occur in the clay R. Gonzales, F. Hongn, J. J. Marcuzzi, J. A. Salfity, A. Villanueva, and
mineral and diatom assemblages of the sediments and suggest significantly J. Viramonte. J. Sosa Gomez provided unpublished geologic maps. The pro-
higher intra-annual and interannual rainfall variations, the sediments do not jections of structural data were produced with the program STEREONET by
support an inference of entirely different climatic conditions in the intra- R. Allmendinger, Cornell University. We gratefully acknowledge thorough re-
montane valleys compared to the present character. This may be related to views by M. Reheis, K. Haselton, F. Nullo, and C. Stark.
the important rain shadow exerted by the meridianally oriented mountain
ranges to the east of the arid valleys that prevents moisture from fundamen- REFERENCES CITED
tally influencing the arid valleys even during times of higher precipitation
Adams, J., 1981, Earthquake-dammed lakes in New Zealand: Geology, v. 9, p. 215–219.
to the east. Higher precipitation on the eastern slopes is documented by the Allmendinger, R. W., 1986, Tectonic development, southern border of the Puna Plateau, north-
asymmetric mountain glaciation during Pleistocene time (Strecker, 1987). western Argentine Andes: Geological Society of America Bulletin, v. 97, p. 1070–1082.
During Pleistocene time the 4–5-km-high eastern ranges underwent Allmendinger, R. W., Ramos, V. A., Jordan, T. E., Palma, M., and Isacks, B. L., 1983, Paleo-
geography and Andean structural geometry, northwest Argentina: Tectonics, v. 2, p. 1–16.
multiple asymmetric glaciations at elevations in excess of 4300 m with Allmendinger, R. W., Strecker, M. R., Eremchuk, J. E., and Francis, P., 1989, Neotectonic
well-developed valley glaciers on the east and limited cirque glaciation on deformation of the southern Puna Plateau, northwestern Argentina: Journal of South Amer-
west-facing slopes (Fox and Strecker, 1991). Mountain glaciation must ican Earth Sciences, v. 2, p. 111–130.
Assumpção, M., and Araujo, M., 1993, Effect of the Altiplano-Puna plateau, South America, on
have caused a much more steady discharge into the ephemeral rivers in the the regional intraplate stresses: Tectonophysics, v. 221, p. 475–496.
intramontane basins. It is therefore conceivable that enhanced scouring at Bevis, M., and Isacks, B. L., 1984, Hypocentral trend surface analysis: Probing the geometry of
Benioff zones: Journal of Geophysical Research, v. 89, no. B7, p. 6153–6170.
mountain fronts in such narrow valleys as the Quebrada del Toro or the Bianchi, A. R., and Yañez, C. E., 1992, Las precipitaciones en el noroeste Argentino (second edi-
Quebrada de Cafayate (Cerro Zorrito) conditioned unstable rock masses tion): Salta, Instituto Nacional de Tecnologia Agropecuaria, 383 p.
that were later dislodged during an earthquake, or were even triggered by Cahill, T., and Isacks, B. L., 1992, Seismicity and shape of the subducted Nazca plate: Journal of
Geophysical Research, v. 97, no. B12, p. 17,503–17,529.
lateral scouring similar to a scenario envisioned for cliff collapse along the Cahill, T., Isacks, B. L., Whitman, D., Chatelain, J.-L., Perez, A., and Ming Chiu, J., 1992, Seis-
Río Grande (e.g., Reneau and Dethier, 1996; Dethier and Reneau, 1996). micity and tectonics in Jujuy Province, northwestern Argentina: Tectonics, v. 11,
This climate-controlled mechanism, however, would have been effective p. 944–959.
Chinn, D. S., and Isacks, B. L., 1983, Accurate source depths and focal mechanisms of shallow
only in narrow valleys with rivers capable of undercutting, a situation that earthquakes in western South America and in the New Hebrides island arc: Tectonics, v. 2,
applies to the locations mentioned here and perhaps to the Villa Vil slides, p. 529–563.
Costa, C. H., and Vita-Finzi, C., 1996, Late Holocene faulting in the southern Sierras Pampeanas
but not to the remainder of the more than 50 landslide localities. of Argentina: Geology, v. 24, p. 1127–1130.
In accordance with other large landslide events in various tectonically ac- Crandell, D. R., and Fahnestock, R. K., 1965, Rockfalls and avalanches from Little Tahoma Peak
tive regions of the world (e.g., Shreve, 1966; Solonenko, 1977; McSaveney, on Mount Rainier, Washington: U.S. Geological Survey Bulletin 1221-A, 30 p.
Crozier, M. J., Deimel, M. S., and Simon, J. S., 1995, Investigations of earthquake triggering for
1978; Plafker and Ericksen, 1978; Adams, 1981; Keefer, 1984; Evans et al., deep-seated landslides, Taranaki, New Zealand: Quaternary International, v. 25, p. 65–73.
1987; Nikonov, 1988) and the fact that all landslide masses in northwestern Czajka, W., and Vervoorst, F., 1956, Die naturräumliche Gliederung Nordwest-Argentiniens:
Argentina cluster along mountain fronts with repeated Quaternary offsets, Petermanns Geographische Mitteilungen, v. 100, p. 89–102.
Dethier, D. P., and Reneau, S. L., 1996, Lacustrine chronology links late Pleistocene climate
we propose that seismic shaking was the principal trigger mechanism that change and mass movements in northern New Mexico: Geology, v. 24, p. 539–542.
caused mountain-front collapse and rock avalanching into piedmont and Eisbacher, G. H., 1979, Cliff collapse and rock avalanches (sturzstroms): Canadian Geotechnical
Journal, v. 16, p. 309–334.
valley regions. Eisbacher, G. H., and Clague, J. J., 1984, Destructive mass movements in high mountains: Haz-
The interpretation of seismic shaking as a trigger mechanism is further ard and management: Geological Survey of Canada Paper 84–16, 230 p.
supported by the coincidence of major rock-avalanche clusters and complex Evans, S. G., Aitken, J. D., Wetmiller, R. J., and Horner, R. B., 1987, A rock avalanche triggered
by the October 1985 North Nahami earthquake, District of Mackenzie, N.W.T.: Canadian
mountain fronts characterized by reverse faulting along former transfer faults Journal of Earth Sciences, v. 24, p. 176–184.
(e.g., Cerro Zorrito, Quebrada del Toro) or normal faulting along former re- Fauque, L., 1987, Avalanchas de roca en la quebrada de Segovia y zonas vecinas de la vertiente
verse faults (e.g., El Rincón, Sierra Laguna Blanca). Due to the kinematic occidental de las sierras del Famatina, Provincia de la Rioja, Argentina: Congreso Ge-
ológico Argentino, 10th, Actas, Volume III, San Miguel de Tucumán, Argentina,
change in northwestern Argentina, the former strike-slip environments are p. 333–336.
now nearly perpendicular or slightly oblique with respect to the neotectonic Fauque, L., and Strecker, M. R., 1988, Large rock avalanche deposits (Sturzströme, sturzstroms)
at Sierra Aconquija, northern Sierras Pampeanas, Argentina: Eclogae Geologicae Helveti-
northeast-southwest shortening directions, causing a concentration of cae, v. 81, p. 579–592.
stresses along these older faults that may ultimately rupture in an earthquake Fox, A. N., and Strecker, M. R., 1991, Pleistocene and modern snowlines in the Central Andes
and cause associated rock avalanching. Although instrumentally recorded (24–28°S): Bamberger Geographische Schriften, v. 11, p. 169–182.
Francis, P. W., and Wells, G. L., 1988, Landsat thematic mapper observations of debris avalanche
large earthquakes are rare in this region (e.g., Chinn and Isacks, 1983), there deposits in the Central Andes: Bulletin of Volcanology, v. 50, p. 258–278.
is evidence for sustained historic and Holocene seismic activity (Strecker Friedmann, J. S., 1997, Rock avalanche elements of the shadow valley basin, eastern Mojave
et al., 1989; Cahill et al., 1992). The limited modern seismic activity thus Desert, California, processes and problems: Journal of Sedimentary Research, v. 67,
p. 792–804.
shows that monitored seismicity does not represent the maximum level of

Geological Society of America Bulletin, June 1999 947


HERMANNS AND STRECKER

Gallardo, E. F., 1988, Geologia del Cuaternario en la confluencia de los Rios Calchaquí y Santa Polanski, J., 1966, Flujos rápidos de escombros rocosos en zonas áridas y volcánicas: Buenos
Maria (Salta): Revista de la Asociación Geológica Argentina, v. 43, p. 435–444. Aires, Editorial Universitaria de Buenos Aires, 67 p.
González Bonorino, F., 1951, Descripción Geológica de la Hoja 12e, Aconquija (Catamarca- Porter, S. C., and Orombelli, G., 1981, Alpine rockfall hazards: American Scientist, v. 69,
Tucumán): Buenos Aires, Argentina, Dirección General de Industria Mineria Bulletin, v. 75, p. 67–75.
p. 1–50. Rapela, C. W., 1976, El basamento metamorfico de la region de Cafayate, Provincia de Salta. As-
Gonzalez Diaz, E. F., Fauque, L., Costa, C., Giaccardi, A., Palomera, P. A., and Pereyra, F., 1997, pectos petrológicos y geoquímicos: Revista de la Asociación Geológica Argentina, v. XXI,
La avalancha de rocas del “Potrero de Leyes,” Sierras Pampeanas australes, Sierra Grande p. 203–222.
de San Luis, Argentina (32° 30′ lat. S): Revista de la Asociación Geológica Argentina, v. 52, Reneau, S. L., and Dethier, D. P., 1996, Late Pleistocene landslide-dammed lakes along the Río
p. 93–107. Grande, White Rock Canyon, New Mexico: Geological Society of America Bulletin, v. 108,
González, O. E., Hongn, F. D., and Mon, R., 1991, Estructura de la Sierra Laguna Blanca y zonas p. 355–385.
aledañas, Provincia de Catamarca: Revista de la Asociación Geológica Argentina, v. 46, Reyes, F. C., and Salfity, J. A., 1973, Consideraciones sobre la estratigrafica del Cretacico (Sub-
p. 299–308. grupo Pirgua) del noroeste Argentino: Congresso Geológico Argentino, 5th, Actas,
Grier, M. E., 1990, The influence of the Cretaceous Salta rift basin on the development of Andean p. 355–385.
structural geometries, NW Argentine Andes [Ph.D. dissert.]: Ithaca, New York, Cornell Uni- Salfity, J. A., and Marquillas, R. A., 1994, Tectonic and sedimentary evolution of the Cretaceous-
versity, 178 p. Eocene Salta Group Basin, Argentina, in Salfity, J. A., ed., Cretaceous tectonics of the An-
Grier, M. E., Salfity, J. A., and Allmendinger, R. W., 1991, Andean reactivation of the Cretaceous des: Earth Evolution Sciences: Braunschweig, Wiesbaden, Friedrich Vieweg und Sohn,
Salta rift, northwestern Argentina: Journal of South American Earth Sciences, v. 4, p. 266–315.
p. 351–372. Schuster, R. L., Logan, R. L., and Pringle, P. T., 1992, Prehistoric rock avalanches in the Olympic
Hadley, J. B., 1964, Landslides and related phenomena accompanying the Hebgen Lake earth- Mountains, Washington: Science, v. 258, p. 1620–1621.
quake of August 17, 1959, in The Hebgen Lake, Montana Earthquake of August 17, 1959: Schwab, K., and Schäfer, A., 1976, Sedimentation und Tektonik im mittleren Abschnitt des Rio
U.S. Geological Survey Professional Paper 435, p. 107–138. Toro in der Ostkordillere NW-Argentiniens: Geologische Rundschau, v. 65, p. 175–194.
Heim, A., 1932, Bergsturz und Menschenleben: Zürich, Beiblatt zur Vierteljahresschrift der Shaller, P. J., 1991, Analysis and implications of large martian and terrestrial landslides
Naturforschenden Gesellschaft in Zürich, no. 30, 217 p. [Ph.D. dissert.]: Pasadena, California Institute of Technology, 586 p.
Hermanns, R. L., 1999, Spatial-temporal distribution of mountain-front collapse and formation Shreve, R. L., 1966, Sherman landslide, Alaska: Science, v. 154, p. 1639–1643.
of giant landslides in the arid Andes of northwest Argentina (24–28° S, 65–68° W) [Ph.D. Siebert, L., 1984, Large volcanic debris avalanches: Characteristics of source areas, deposits, and
dissert.]: Potsdam, Germany, Potsdam University, 123 p. associated eruptions: Journal of Volcanology and Geothermal Research, v. 22, p. 163–197.
Hewitt, K., 1988, Catastrophic landslide deposits in the Karakoram Himalaya: Science, v. 242, Solonenko, V. P., 1977, Landslide and collapses in seismic zones and their prediction: Interna-
p. 64–67. tional Association of Engineering Geology Bulletin, v. 15, p. 4–8.
Hsü, K. J., 1975, Catastrophic debris streams (sturzstroms) generated by rockfalls: Geological So- Strecker, M. R., 1987, Late Cenozoic landscape development, the Santa Maria valley, northwest-
ciety of America Bulletin, v. 86, p. 129–140. ern Argentina [Ph.D. dissert.]: Ithaca, New York, Cornell University, 262 p.
Jezek, P., Aceñolaza, F. G., and Miller, H., 1985, The Puncoviscana trough: A large basin of Late Strecker, M. R., and Marrett, R. A., 1999, Kinematic evolution of fault ramps and the role in de-
Precambrian to Early Cambrian age on the Pacific edge of the Brazilian shield: Geologische velopment of landslides and lakes in the northwestern Argentine Andes: Geology, v. 27, p.
Rundschau, v. 74, p. 573–584. 307–310.
Jibson, R. W., and Keefer, D. K., 1993, Analysis of the seismic origin of landslides: Examples Strecker, M. R., Cerveny, P., Bloom, A. L., and Malizzia, D., 1989, Late Cenozoic tectonism and
from the New Madrid seismic zone: Geological Society of America Bulletin, v. 105, landscape development in the foreland of the Andes: Northern Sierras Pampeanas
p. 521–536. (26°–28°S), Argentina: Tectonics, v. 8, p. 517–534.
Jordan, T. E., Isacks, B. L., Allmendinger, R. W., Brewer, J. A., Ramos, V. A., and Ando, C. J., Tibaldi, A., Ferrari, L., and Pasquarè, G., 1995, Landslide triggered by earthquakes and their re-
1983, Andean tectonics related to geometry of subducted Nazca plate: Geological Society lations with faults and mountain slope geometry: An example from Ecuador: Geomorphol-
of America Bulletin, v. 94, p. 341–361. ogy, v. 11, p. 215–226.
Keefer, D. K., 1984, Landslides caused by earthquakes: Geological Society of America Bulletin, Topping, D. J., 1993, Paleogeographic reconstructions of the Death Valley extended region: Evi-
v. 95, p. 406–421. dence from Miocene large rock-avalanche deposits in the Amargosa Chaos basin, Califor-
Kleinert, K., Trauth, M. H., and Strecker, M. R., 1997, A Pleistocene lacustrine phase in the Que- nia: Geological Society of America Bulletin, v. 105, p. 1190–1213.
brada de Cafayate, NW-Argentina: Evidence for variations in paleohydrology from multi- Torres, M. A., 1985, Estratigrafia de la ladera occidental del Cerro Amarillo y Quebrada de La
tracer analysis: Terra Nova, v. 9, abstract supplement 1, p. 631. Yesera, Departamento de Cafayate, Salta: Revista de la Asociación Geológica Argentina,
Ledru, M.-P., Soares Braga, P. I., Soubiès, F., Fournier, M., Martin, L., Suguio, K., and Turcq, B., v. 40, p. 141–157.
1996, The last 50 000 years in Neotropics (Southern Brazil): Evolution of vegetation and Trauth, M. H., and Strecker, M. R., Formation of landslide-dammed lakes during a wet period be-
climate: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 123, p. 239–257. tween 40, 000–25, 000 yr B. P. in northwestern Argentina: Palaeogeography, Palaeoclima-
Marcuzzi, J. J., Wayne, W. J., and Alonso, R. N., 1994, Geologic hazards of Salta Province, Ar- tology, Palaeoecology, in press.
gentina: International Association of Engineering Geology Proceedings, v. 7, p. 2039–2048. Trauth, M. H., Müller, A. B., and Strecker, M. R., 1998, The role of climate as a preparatory or
Marrett, R. A., 1990, The late Cenozoic tectonic evolution of the Puna plateau and adjacent triggering factor in the generation of catastrophic landslides in NW Argentina: European
foreland, northwestern Argentine Andes [Ph.D. dissert.]: Ithaca, New York, Cornell Uni- Geophysical Society, Annales Geophysicae, Nizza, v. 16, p. C 1202.
versity, 365 p. Turcq, B., Pressinotti, M. M. N., and Martin, L., 1997, Paleohydrology and paleoclimate of the
Marrett, R. A., and Strecker, M. R., 1997, Kinematic reorganization of Andean deformation in the past 33 000 years at the Tamanduá River, central Brazil: Quaternary Research, v. 47,
Quebrada del Toro, NW Argentina: Geological Society of America Abstracts with Pro- p. 284–294.
grams, v. 29, no.6, p. A-48. van der Hammen, T., and Absy, M. L., 1994, Amazonia during the last glacial: Palaeogeography,
Marrett, R. A., Allmendinger, R. W., Alonso, R. N., and Drake, R. E., 1994, Late Cenozoic tec- Palaeoclimatology, Palaeoecology, v. 109, p. 247–261.
tonic evolution of the Puna Plateau and adjacent foreland, northwestern Argentine Andes: Voight, B., 1978, Lower Gros Ventre slide, Wyoming, U.S.A., in Voight, B., ed., Rockslides and
Journal of South American Earth Sciences, v. 7, p. 179–207. avalanches: Amsterdam, Elsevier, p. 277–314.
McSaveney M. J., 1978, Sherman glacier rock avalanche, Alaska, U.S.A., in Voight, B., ed., Voight, B., and Pariseau, W. G., 1978, Rockslides and avalanches: An introduction, in Voight, B.,
Rockslides and avalanches: Amsterdam, Elsevier, p. 197–257. ed., Rockslides and avalanches: Amsterdam, Elsevier, p. 1–67.
Mon, R., 1976, The structure of the eastern border of the Andes in northwestern Argentina: Ge- Watson, R. A., and Wright, H. E., 1963, Landslide on the east flank of the Chuska Mountains,
ologische Rundschau, v. 65, p. 211–222. northwestern New Mexico: American Journal of Science, v. 261, p. 525–548.
Mon, R., and Hongn, F. D., 1996, Estructura del basamento proterozoico y paleozoico inferior del Wayne, W. J., 1994, Large Andean landslides: Indicators of prehistoric seismic epicenters: Geo-
norte argentino: Revista de la Asociación Geológica Argentina, v. 51, no. 1, p. 3–14. logical Society of America Abstracts with Programs, v. 26, no. 7, p. A-216.
Nicoletti, P. G., and Sorriso-Valvo, M., 1991, Geomorphic controls of the shape and mobility of Weidinger, J. T., Schramm, J.-M., and Surenian, R., 1996, On prefatory causal factors, initiating
rock avalanches: Geological Society of America Bulletin, v. 103, p. 1365–1373. the prehistoric Tsergo Ri landslide (Langthang Himal, Nepal): Tectonophysics, v. 260,
Nikonov, A. A., 1988, Reconstruction of the main parameters of old large earthquakes in Soviet p. 95–107.
Central Asia using the paleoseismological method: Tectonophysics, v. 147, p. 297–312. Yarnold, J. C., 1993, Rock-avalanche characteristics in dry climates and the effect of flow into
Omarini, R. H., 1983, Caracterización litológica, diferenciación y génesis de la Formación Pun- lakes: Insights from mid-Tertiary sedimentary breccias near Artillery Peak, Arizona: Geo-
coviscana entre el Valle de Lerma y la Faja Eruptiva de la Puna Plateau. [Ph.D. dissert.]: logical Society of America Bulletin, v. 105, p. 345–360.
Salta, Argentina, Universidad de Salta, 220 p.
Penck, W., 1920, Der Südrand der Puna Plateau de Atacama (NW Argentinien): Akademie der
Wissenschaften Abhandlungen der Mathematisch–Physikalische Klasse, v. 37, 420 p. MANUSCRIPT RECEIVED BY THE SOCIETY MARCH 26, 1998
Plafker, G., and Ericksen, G. E., 1978, Nevados Huascarán avalanches, Peru, in Voight, B., ed., REVISED MANUSCRIPT RECEIVED AUGUST 3, 1998
Rockslides and avalanches: Amsterdam, Elsevier, p. 277–314. MANUSCRIPT ACCEPTED SEPTEMBER 16, 1998

Printed in U.S.A.

948 Geological Society of America Bulletin, June 1999


View publication stats

You might also like