Sébrier, M. & Mercier, J. - Tectonics 1988 - The State of Stress in An Overriding Plate Situated Above A Flat Slab, Andes of Central Peru

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

TECTONICS, VOL. 7, NO.

4, PAGES 895-928, AUGUST 1988

THE STATE OF STRESS IN AN OVERRIDING


PLATE SITUATED ABOVE A FLAT SLAB:
THE ANDES OF CENTRAL PERU

Michel S•brier, Jacques Louis Mercier,


Jos• Machar•, Didier Bonnot, Justo Cabrera,
and Jean Luc Blanc

Laboratoire de G•ologie Dynamique Interne,


Universit• de Paris-Sud, France

Abstract. In the Andes of central Peru Andes, tectonics being compressional, •zz
that are situated above a flat dipping is •3 and •Hmax is •1; in the High Andes,
slab, analysis of both structural data •zz becomes •1, then •Hmax is •2,
collected in the field and available focal allowing oHmin striking N-S to be •3. Along
mechanisms show that the Quaternary and the Pacific Coast, tectonics is not
Recent state of stress is characterized by compressional as expected, and a nearly
N-S trending extension in the Western neutral state of stress may be due to a
Cordillera and along the Pacific Coast, topographical effect related to the nearby
whereas E-W trending compression prevails deep Peru-Chile trench. However, a
in the sub-Andean Lowlands and at the conspicuous difference occurs in the Andes
contact between the Nazca and South of central Peru in respect to the southern
American plates. The Eastern Cordillera Peru Andes: compressional strike-slip
deformations suggest an intermediate state tectonics is observed in the Eastern
of stress (i.e., strike-slip faulting) Cordillera. Indeed, the mean elevation of
characterized by E-W trending compression the Eastern Cordillera is 3.7 km, whereas
and N-S trending extension. Thus major that of the Western Cordillera i• 4.2 km,
fault zones striking NW-SE, i.e., parallel allowing an intermediate state of stress
to the central Peruvian Andes, are normal to occur. Therefore in central Peru a
with a sinistral component in the Western lower mean Andean topography and a
Cordillera (e.g., Cordillera Bianca) and stronger coupling consequence of the flat
reverse sinistral in the Eastern subduction seems to explain that
Cordillera (e.g., Cordillera de compressional tectonics is distributed in
Huaytapallana). This state of stress a more widespread area than in southern
displays similarities with that of Peru. Superficial deformations related to
southern Peru and may be also interpreted the subduction of the aseismic Nazca ridge
as an effect of compensated high are restricted to the coastal domain
topography. In this model the vertical opposite to this ridge. They support The
stress •zz increases with the topography fact that submarine topography of large
and •Hmax is considered as fairly extension may subduct without producing
constant, trending E-W roughly parallel to compressional deformation in the
the convergence. On both edges of the overriding plate which can be reminiscent
of a collisional setting.
Copyright 1988
1. INTRODUCTION
by the American Geophysical Union.
Paper number 8T0202. Along subduction zones, tectonic regime
0278- 7407/88/008T-0202510.00 in the overriding plate is generally
896 S•brier et al.: Stress in the Andes of Central Peru

related to the dip of the oceanic slab; Suarez et al., 1983; Blanc et al., 1983].
this one being a consequence of the slab In this paper we describe all the
buoyancy [Vlaar and Wortel, 1976; Molnar Quaternary deformations we have observed
and Atwater, 1978; Uyeda and Kanamori, in central Peru. We first used Landsat
1979; England and Wortel, 1980]; the less imagery and side-looking airborne radar
buoyant the slab the more steeply it dips. (SLAR) images in the sub-Andes in order to
Thus two end types can be distinguished map the most conspicuous fault zones of
among subduction zones [Uyeda and central Peru. Then we focused our field
Kanamori, 1979]: (1) the Mariana type work essentially on the Quaternary basins
characterized by steep dipping slab and which are associated with major Quaternary
extensional tectonics and (2) the Andean fault zones. Analysis of slip-fault data
type characterized by low-angle dipping has been performed using the CareyVs
slab and dominantly compressional inversion algorithm [Carey and Brunier,
tectonics. In fact, the Andean type 1974; Carey, 1979; S•brier et al., 1985].
displays two different slab geometries Although data sampling is less dense than
[Stauder, 1973, 1975; Barazangi and in southern Peru, it nevertheless permits
Isacks, 1976, 1979]: 30 ø dipping segments study of the interesting occurrence of
(i.e., south Peru-Bolivia-north Chile and extensional and compressional faulting in
central Chile) and flat subducting the High Andes. It is demonstrated that in
segments (central Peru and northcentral the Coastal area and Western Cordillera,
Chile) so that the different situations normal faulting indicates a roughly N-S
can be compared. In southern Peru, S•brier trending extensioB, whereas in the Eastern
et al. [1985] have shown that the Cordillera, in the sub-Andes, and at the
continental overriding plate situated contact between the two plates, focal
above the 30 ø dipping slab is mechanisms [Stauder, 1975; Suarez, 1982;
characterized by compressional deformation Suarez et al., 1983] and field data show
on both edges of the Andes (plate an E-W trending compression. As in
interface and Amazonian foothills), southern Peru, the occurrence of
whereas the High Andes and Pacific extensional tectonics appears to be
Lowlands exhibit extensional deformation. related to the high topography [S•brieret
The latter is interpreted as a topographic al., 1985]. The different tectonic regimes
effect related to the existence of the observed in the High Andes seem to be
Central Andes high plateau and of the essentially explained by different mean
Peru-Chile trench. elevations of the Western and Eastern
In central Peru the oceanic slab Cordilleras, more than 4 km and less than
plunges with a 30 ø dipping angle down to a 4 km, respectively.
depth of approximately 80 km, then it
flattens beneath the High Andes [Stauder, 2. EXTENSIONAL TECTONICS IN THE HIGH
1975; Barazangi and Isacks, 1979• Suarez, ANDES
1982; Bevis and Isacks, 1984]. In this
region recent deformation of the 2.1. Cordillera Blanca Normal Faults:
overriding plate appears to be rather Western Cordillera
complex [S•brier et al., 1982• Suarez et
al., 1983]. First, it has been considered The NNW-SSEstriking Cordillera Blanca
as compressional [M•gard and Philip, (Figure 1, points 33, 34, 35, and 36) is a
1976]; then Soulas [1978] separates a snow-capped range, which is nearly 200 km
Coastal extensional domain from a High long and more than 6 km high. On its
Cordillera compressional one. In fact, western side an active SW facing normal
normal faulting affects the Pacific Coast fault zone separates the High Cordillera
[Soulas, 1978; S•brier and Machar•, 1980; from the 15-km-wide Callejon de Huaylas
Machar• et al., 1986] and part of the High intracordilleran basin (Figure 2). This
Andes [Heim, 1949• Silgado, 1951; basin corresponds to a subsiding Pliocene
Dalmayrac, 1974; M•gard and Philip, 1976; graben located between the Cordilleras
Yonekura et alo, 1979; Mercier and Negra and Bianca to the SW and NE,
S•brier, 1981; Bonnot, 1984] Compressional respectively. This graben has been
faulting affects the Amazonian (sub-Andes) infilled by continental deposits up to
foothills, the contact between the two 2000 m thick which are unconformably
plates [Stauder, 1975; Chinn and Isacks, overlain by Quaternary outwash and moraine
1983; Suarez et al., 1983• Blanc, 1984] deposits [Bonnot, 1984]. The Cordillera
and also some parts of the High Andes Bianca active normal fault zone clearly
[Deza, 1971; Philip and M•gard, 1977; cut the present-day topography [Dalmayrac,
S•brier et al.' Stress in the Andes of Central Peru 897

7eo

SUJO

SU.5

PIURA
ß

•TRI
UICHES

ß X

ST.13(•
ST.31

•2770 •
ST.
5 •CERRO
DE SU.,a

1710
eef.
•AI
ST.
7(•)

RDIk•.
AYACUC
•% HO CUZCO

T/iI
0I 100
I 2$Okm
I ' "':
31)",1
eee• , ,\•-
•1
',/\
I

•i•. 1. •oca•ons o• n•be•ed Qua•e•na• 7 •aul• s•es a•a17•ed •n •is paper.


T•c• so1•d l•nes' ma]o• Qua•e•na• 7 •aul•s obse•ed in •e •eld (ha•c•u•ed on
•e do••ows); •c• dashed l•nes: •n•e•ed Qua•e•a• 7 •aul•s; •c• l•nes
wi• a•acbed so1•d •ian•les' ma]o• sub-A•dea• •e•e•se •aul•s; s•aded a•eas'
01i•o-E•oce•e •olcanics; s•ppled a•eas' •l•o-Qua•e•a• 7 basins (C•, Ca]ababa
bas•n; •, •aAe •u• bas•n). •ocal mechanisms: le•e•s •e•e• •o authors, Ab
[Abe, 1972], S• [S•aude•, 1975], Sp [Dewey a•d Spence, •979], •e [•enn•n•on,
1981] , Su [Sua•e•, 1982; Sua•e• e• al. , 1983]; n•be•s a•e •ose used b 7 •hese
au•Ao•s.

1974; Yonekura et al., 1979] and is is carved in the Upper Miocene grano-
characterized by one of the steepest dioritic stocks of the Cordillera Blanca.
topographical gradients in the Andes, 4.5 Downward, this scarp is generally limited
km of vertical variation along a 15-km by morphological scarplets (Figures 3a,
distance. This fault zone exhibits a 1000- 3b, 3c, 3d) that offset Late Quaternary
m-high faceted scarp (Figures 3a, 3b) that moraines with vertical throws ranging
898 S•brier et al.' Stress in the Andes of Central Peru

Huaytapallana

,,
Acra Ranco

Huallanc

Oueruran

Ssn
Cr, iN) '

Ca•a

Ulta

Yunga
Sen
Cr,$tobal {C) •

nda

San
Cristobal IS)

Llaca

• Ouerococha
Catac'
9
,,

•,•,,
Quenua Ragra

P,nculloc
0 1OKra

Fig. 2. Location of the Cordillera Blanca normal fault system (Figure 1,


points 33, 34, 35, 36). Small stereonet represents the measured slip vector
corresponding to the Quaternary and active kinematics of the major faults.
Stippled areas are higher than 5000 m.

between 6 and 50 m [Yonekura et al., 1979; The southern segment of the fault zone
Bonnot and S•brier, 1985]. However, no is characterized by 15-km-long normal
historical earthquake is known to have faults striking N140øE and displaying a
reactivated the Cordillera Blanca fault left-hand en "echelon" pattern (Figure 2).
zone [Silgado, 1978]; this may be However, the slip vectors corresponding to
explained by a rather long return period the latest fault movement observed on the
(estimated of the order of 1500-3000 major slickensides of the Cordillera
years) for major seismic shocks capable of Blanca fault zone clearly show that NW-SE
producing surface faulting [Bonnot, 1984; to NNW-SSE trending normal faults have a
Bonnot and S•brier, 1985]. sinistral strike-slip component. Thus
S•brier et al.' Stress in the Andes of Central Peru 899

extension does not appear orthogonal to (Figure 1, point 35) we observed 15 normal
the Cordillera Blanca strike but roughly faults having metric downthrows (Figure 4,
N-S. Along the Cordillera Blanca major stereodiagram 35) that postdate
fault zone, three fault sets have been compressional deformation affecting the
separated (Figure 4, sites 33, 34, and 36) Pliocene continental deposits; these data
corresponding to the three northern, yield a N190øE minimumprincipal stress
central, and southern segments. Inversion direction. Thus the Recent kinematics of
of the slip vector data for the southern the Cordillera Blanca normal fault zone
and central segments gives minimum appear to be characterized by a nearly N-S
principal stress directions •3 striking trending extension.
N10øW (site 36, Table 1) and N183øE (site
34, Table 1). In the northern part the 2.2. 1946 Quiches Seismic Normal Faults
fault azimuthal distribution does not
allow direct computation of the stress The Quiches normal active fault is
tensor, so we use a test program to located in the Mara•on fold and thrust
determine the best stress tensor fitting belt of Tertiary age and forms the
the data. The result obtained corresponds northeastern part of the Western
to a minimum principal stress striking Cordillera (Figure 1, point 32). It has
N10øE (site 33, Table 1). In the central been activated during the November 10,
part of the Callejon de Huaylas basin 1946, Ancash earthquake (M = 7.25),
900 S•brier et al.' Stress in the Andes of Central Peru

Fig. 3b

/'

.,,,,..

Fig. 3c
S•brier et al.: Stress in the Andes of Central Peru 901

Fig. 3d

located at 8ø20'S and 77ø50'W, whose focal the geometry of the fault scarp clearly
depth was estimated about 30 km [Heim, indicates a normal movement; and (3)
1949; Ruegg, 1950; Silgado, 1951; Richter, several huge landslides have occurred
1958]. Discontinuous surface ruptures are during the 1946 earthquake and some of the
exposed along a 20-km-long distance antithetic faults could be of such an
(Figure 5). The main fault is outlined by origin. It is even possible that a
two NW-SE trending segments of SW facing landslide component has partly induced the
scarplets which offset Quaternary glacial fault movement at the surface. Although
deposits and Mesozoic limestones and our observations were performed 32 years
sandstones. The southeastern scarplet is after the Ancash earthquake, the magnitude
located at an elevation ranging between of surface offset is such that it is still
4000 and 3600 m and shows a vertical possible to demonstrate a tensional
offset up to 3.5 m (Figure 6). It has a tectonic origin of the Quiches fault.
N138øE mean strike and a 58 ø SW mean dip. Hodgson and Bremner [1953] have
The northwestern segment (Figure 5) is proposed a reverse strike-slip fault plane
located at a mean elevation of 4000 m and solution and concluded that the whole
has a N125øE mean strike and a vertical surface faulting was a secondary effect of
offset of 1 m. Two NE facing scarplets, the 1946 earthquake. On the contrary,
approximately parallel to the main fault, Suarez et al. [1983] have proposed a
are 0.3 and 2 km long and have been purely normal fault plane solution;
considered as antithetic faults [Silgado, although their solution is not strongly
1951]. constrained, it demonstrates a normal
Field observations performed in the component. Therefore both field and
Quiches area [S6brier, 1979] show that (1) seismic data agree with normal faulting
the clear striations seen on the for the Ancash earthquake. Unfortunately,
calcareous exposures of the seismic fault none can specify the direction of
plane [SilEado, 1951] are reverse extension that produced the Quiches fault
stylolithic striations: they do not thus reactivation.

indicate the 1946 seismic slip vector but


are more likely a creep deformation due to 2.3. Ayacucho Basin Normal Faults
an older compressional regime; (2) the
1946 seismic vector cannot be measured The NW-SE striking Ayacucho depression
from the present-day field exposures, yet is an intracordilleran basin with a mean
902 S•brier et al.' Stress in the Andes of Central Peru

"•:::•.o' 33

,2..
53

N

[4!
14

N 534344 :
4532::1l' /
34.• •J

•Ji,o, • 23 2, s2
• J4oso•J,j• •3
o lo 3o 3o o o4•3.o2os•o2,i•'?•sloo
]o 4o

35 350'-5' 36

4, 12

6-

"•,,, o.) 8
3 7

1o,i,,l,,i,,!:,]
. ,s',l% o !,,•!o,!os!e,!,,]
o 1o 20 3o
o io '20

Fig. 4. Normal fault data from Western Cordillera (locations on Figure 1)


used to compute solutions of Table 1. Numbers outside stereonets refer to data
inside the histograms. Arrows attached to fault traces correspond to the
measured slip vector S (Wulff stereonet, lower hemisphere). Thick segmentson
fault traces and histograms showdeviations between measured (S) and predicted
(t) slip vector for each fault plane. Divergent large black arrows give
azimuths of the minimum o3 computed principal stress directions. Solution of
site 33 results from a test.

elevation of nearly 3000 m (Figure 1, equivalent with the Huari latitic lavas
point 42). It has been infilled by and the Rio Vinchos rhyolitic tuffs which
Paleogene and Neogene formations (Figure are dated at 3.8 ñ 0.4 m.y. [M•gard et
7) that have been folded and faulted al., 1984] and 2.45 ñ 0.06 m.y. [Kaneoka
during Tertiary compressional tectonic and Guevara, 1984], respectively. These
pulses [M•gard et al., 1984]. From Latest pediments and fans are mainly cropping out
Miocene, erosional processes have been at the foot of the SW facing, 1000 m high,
dominant with respect to sedimentation. Razuhuilca scarp that bounds this basin to
PliooPleistocene evolution is the NE (Figure 7). This scarp does not
characterized by pediment landforming and exhibit any evidence of Recent activity;
thin alluvial fans that are partly time- it corresponds to a NW-SE striking reverse
S•brier et al.: Stress in the Andes of Central Peru 903

TABLE 1. Parameters of the Deviatoric Stress Tensors Computed From the Quaternary
and Recent Normal Faults of the High Andes and Pacific Lowlands

Principal Stress Directions


Site ND Latitude, Longitude, •1 •2 •3
S W Azimuth Dip Azimuth Dip Azimuth Dip

1 10 15ø22 f 75ø09 f 191 ø 85 ø 092 ø O1 ø 002 ø 05 ø 0.86


9 19 13ø30 • 71ø56 • 324 ø 80 ø 097 ø 07 ø 188 ø 07 ø 0.42
27 17 13ø05 • 76ø20 • 179 ø 80 ø 288 ø 03 ø 019 ø 09 ø 0.65
3O 24 14ø42 ' 75ø17 f 288 ø 85 ø 083 ø 05 ø 173 ø 02 ø 0.92
33 15 9ø05 • 77ø42 ' 100 ø 90 ø 280 ø O0 ø 010 ø O0 ø 0.25
34 54 9ø27 • 77ø30 • 033 ø 80 ø 274 ø 05 ø 183 ø 09 ø 0.55
35 15 9ø38 • 77ø28 • 294 ø 89 ø 100 ø O1 ø 190 ø O1 ø 0.70
36 15 10ø00 • 77ø15 • 131 ø 84 ø 259 ø 04 ø 350 ø 05 ø 0.69
42 44 13ø05 f 74ø10 • 128 ø 70 ø 294 ø 19 ø 026 ø 04 ø 0.54
42a-b-c 15 13ø03 • 74ø08 • 124 ø 74 ø 293 ø 15 ø 024 ø 03 ø 0.35
42d 29 13ø07 • 74ø12 ' 175 ø 71 ø 280 ø 05 ø 012 ø 18 ø 0.59

These parameters are computed from superficial Quaternary and Recent normal
faults of the Pacific Lowlands (sites 1, 27, and 30) and of the High Andes (sites
9, 33, 34, 35, 36 and 42). Site localizations are shown on Figure 1. ND is the
number of striated fault planes used to compute the solutions. Azimuths are
measured clockwise from North; dip is toward the measured azimuth. R = •'2-
•'1/•'3-•'1 is the "stress ratio" or "shape factor" of the stress tensor. Its
value varies between two extremes: 0 as •'2 = •'1 and 1 as •f2 = •'3. Computed
normal faults are on stereonets of Figures 4, 9, and 16.

fault and flexure zone that has been c), and Totorilla (point d). Totorilla
active during and subsequently to the normal fault planes postdate reverse
sedimentation of the Late Miocene Ayacucho faulting that cut the tuffaceous beds of
formation. the Late Miocene Ayacucho formation.
Compressional deformations are observed Quinua (Figure 8) and Jachuacmollo normal
in alluvial fan deposits that rest slickensides were observed affecting Plio-
unconformably on the Late Miocene Ayacucho Pleistocene torrential deposits. Normal
formation. As these fans are of Pliocene sinistral striations observed in the
or Early Quaternary age, it is very likely Quebrada Lluncuna correspond to the last
that these compressional deformations reactivation of the Razuhuilca fault zone.
occur during Early Quaternary as it has All these normal faults have throws
been shown in many localities of the ranging between several decimeters and
Central Andes [Martinez, 1980; Lavenu et several meters. Faults having a roughly E-
al., 1980; S•brier et al., 1980, 1982; W direction are nearly purely normal;
S•brier and Machar•, 1980; Blanc, 1984; faults having a roughly NE-SW to N-S
Bonnot, 1984; Huaman, 1985]. The youngest direction have a dextral strike-slip
fault activity, due to normal faulting, component while faults striking NW-SE have
postdates these compressional deformations a sinistral component (stereonets 42a-b-c
and thus should be of Quaternary age. and 42d, Figure 9). These kinematics agree
During 1981, a shallow seismic crisis with an approximately NNE-SSW lenthening.
occurred in the Ayacucho area. Computation of the principal stress
Unfortunately, no evidence of surface directions confirms this first inference.
faulting, clearly of tectonic origin, has We first processed the Totorilla striated
been observed [Machar•, 1980, S•brier and fault planes and those measured in the
Cabrera, 1981], and no focal mechanisms vicinity of La Quinua (Figure 9,
have been published so that present-day stereonets 42d and 42a-b-c). Numerical
state of stress in the Ayacucho basin is results are similar, showing N12øE and
unknown. N24øE, respectively, trending extensional
In this basin we measured 44 striated principal stress directions (Table 1).
normal fault planes at several localities: Grouping all the data from the Ayacucho
Quebrada Lluncuna (point a on Figure 7), basin (Figure 9, stereonet 42), we
La Quinua (point b), Jachuacmollo (point obtained a minimum principal stress (•3)
904 S•brier et al.' Stress in the Andes of Central Peru

/
/
/
/

0 5Kin

Fig. 5. Location of normal surface ruptures of the 1946 Quiches earthquakes


(Figure 1, point 32)' thick lines, hatchured on the downthrows. Lines with
roman numbers correspond to the 1946 isoseists [Silgado, 1951].

trending N26øE. Computation of the Huaytapallana range to the southwest


uncertainities shows that the minimum [M•gard, ].978]. This range is located in
principal stress axes trend between N6øE the Eastern Cordillera, 18 km to the NE of
and N28øE (Figure 9, stereonet A) with R the intermontane Huancayo basin where
values ranging between 0.35 and 0.7. conspicuous Pleistocene compressional
Therefore normal faults of the Ayacucho deformations have been reported [M•gard,
basin are fairly homogeneous and indicate 1968; Dollfus and M•gard, 1968; Soulas,
a nearly NNE-SSW trending extension. 1975; Cabrera, 1982; S•brier et al., 1982;
Blanc, 1984]. Nevertheless, no clear
3. COMPRESSIONAL STRIKE-SLIP TECTONICS evidence of active deformation is observed
IN THE HIGH ANDES within the Huancayo basin [Blanc, 1984],
and surface faulting related to seismic
activity is restricted to the
3.1. 1969 .Huaytapallana Reverse Fault,
Eastern Cordillera
Huaytapallana fault [Philip and M•gard,
1977; Blanc et al., 1983; Blanc, 1984]
that limits Precambrian from Permo-
The Huaytapallana fault (Figure 1, Triassic rocks. This fault was reactivated
point 41) belongs to the NW-SE trending during two earthquakes: on July 24 (M=5.6)
fault zone which bounds the snow-capped and October 1 (M=6.2) 1969 [Deza, 1971;
S•brier et al.' Stress in the Andes of Central Peru 905

Fig. 6. Field view, taken in 1978, of the 3-m-high normal scarp (arrow heads)
produced by the 1946 Quiches earthquake, altitude 3950 m. In the background,
toward the northeast, the eroded high plateau of the Eastern Cordillera is
separated from the Quiches fault by the 2000-m-deep canyon (black arrow) of
the Mara•on river.

Stauder, 1975; Silgado, 1978]. These propagation of the rupture in favor of an


earthquakes produced a SW facing scarplet asperity-barrier model [Aki, 1984].
[Deza, 1971; Philip and M•gard, 1977] The reverse left-lateral strike-slip
which is actually composed of two major kinematics of the Huaytapallana fault is
segments (Figure 10). The northern one confirmed by superficial seismic
(Figures 11a, lib), 9.5 km long, exhibits deformations [Philip and M•gard, 1977], by
a 2-m-high reverse throw [Blanc et al., slip vector measurements (Blanc et al.
1983; Blanc, 1984]. The southern one, 5.5 [1983] and Figure 12), and by several
km long, shows a 1.8-m-high reverse throw focal mechanisms (Stauder [1975], Suarez
and a 0.7-m-left-lateral offset [Philip et al. [1983], and Figure 10). In order to
and M•gard, 1977]. Field studies show that observe detailedly the seismic rupture,
the seismic faults have NW-SE strikes and trenching has been performed along the
high-angle dips toward the NE. As the southern segment of the Huaytapallana
southern and northern segments are fault [Blanc et al., 1983]. These trenches
separated by a 3.5 km zone where no have exhibited the downthrusted 1969 soil
evidence of seismic surface rupture has and shown that the geometry of the
been observed (Figure 10), we think that superficial fault plane corresponds to an
these two major segments are likely to be imbricated structure due to surperficial
due to the two 1969 earthquakes, rotations and collapses of the upthrown
respectively. Moreover, the International block edge. Measured slip vectors on
Seismological Centre locations are quite metric to decimetric slickensides show an
different for the two shocks, the October inhomogeneous deformation due to these
one being more to the •rW in respect to the superficial effects. However, Blanc et al.
July 24 one; both epicenters being shifted [1983] have shown that this inhomogeneous
toward the NW in respect to the surface deformation is in agreement with a WSW-ENE
ruptures. Thus it is suggested that the shortening (Figure 12) which is in good
southern segment ruptured first and the agreement with four of the six available
northern one later, showing a northwestern focal mechanisms.
906 S•brier et al.' Stress in the Andes of Central Peru

7,•'3o
12 ø 30

13'

AYACU
ß

o lo KM
ß . .

Fig. 7. Structural sketch (modified after MAgard et al. [1984]) of the


Ayacucho basin (Figure 1, point 42). 1, Quaternary alluvial fans and
pediments; 2, Pliocene volcanics; 3, late Miocene Ayacucho formation; 4, upper
Miocene volcanics; 5, early to middle Miocene continental and volcanic rocks;
6, Pre-Miocene formations; 7, anticline; 8, syncline; 9, monoclinal folds; 10,
fault; 11, normal fault; 12, reverse fault; 13, bed dip. Stars with letters
correspond to the Ayacucho sites (Figure 1, point 42): a, Quebrada Lluncuna;
b, La Quinua; c, Jachuacmollo; d, Totorilla. Stars e (Ayacucho airport) and 43
(Mitapasama•an) are sites where Soulas [1978] has reported compressional
Quaternary deformation of early Quaternary age.

No other reverse active fault is known this normal fault could be partly due to a
in the Eastern Cordillera or even in the landslide effect as it is located just on
rest of the High Andes. The only other the edge of the High Andes, bordering upon
Quaternary fault mentioned in the Eastern the 2000-m-deep canyon of the Rio Ulcumayo
Cordillera is the NNW-SSEtrending Cayesh River. Unfortunately, we were not able to
fault (Figure 1, point 38) [MAgard, 1973; carry out field studies on this fault.
MAgard and Philip, 1976]. This one is
3.2. High Andes Focal Mechanisms
described as a 70 ø SW dipping normal fault
which has a 5- to 10-m throw and exhibits In the High Andes, seismicity is
a left-hand "en echelon" pattern. However, generally low, and consequently, there are
S•brier et al.' Stress in the Andes of Central Peru 907

Fig. 8. La Quinua normal fault (Figure 7, point b) cutting late Miocene


Ayacucho Formation (white) and Pliocene to Early Quaternary torrential
deposits (dark) with a 7-m vertical offset. On the right a secondary steep
normal fault is seen with a 2-m vertical offset.

few available focal mechanisms. There are to the microseismic activity of the
only two focal mechanisms showing normal Huaytapallana fault area; the last one is
motion in the whole High Andes, and both the strongest event (June 4, 1980) of this
are located in the Western Cordillera and microseismic activity. Each of these
are of poor quality. One is a solution of solutions shares several similarities
the 1946 Ancash earthquake (see section (Figures 1, 10, and 13): (1) one of the
2.2), the other one is composite and nodal planes has a NW-SE strike
corresponds to microseismic activity corresponding to the central Peru Andean
recorded during 1981 from a local network trend, a strong dip toward the NE, and a
that was located near Chonta in southern reverse left lateral motion; (2) all the P
Peru [Grange et al., 1984b]. The other axes have a roughly E-W trend with a low
High Andes focal mechanisms are located in dip; and (3) the T axes have roughly N-S
the Eastern Cordillera of central Peru and trends with variable inclinations.
show reverse or strike-slip solutions Although the available number of seismic
(Figures 1 and 10). Eight proceed from events is small, their common
World-Wide Standard Seismological Network characteristics suggest that the Eastern
(WWSSN) data and concern three distinct Cordillera deformations correspond
seismic events (July 24, 1969; October 10, preferentially to strike-slip faulting
1969; and October 15, 1971) [Stauder, resulting from an E-W trending compression
1975; Philip and M•gard, 1977; Suarez et and a nearly N-S trending extension.
al., 1983] . The four other ones [Suarez,
1982] were obtained from a local seismic 4. COMPRESSIONAL TECTONICS IN THE
network that was operating during 1980; SUB-ANDES
two mechanisms are from the same shock
(May 15, 1980) and are considered as very In central Peru, similarly to southern
poorly constrained [Suarez, 1982] . One is Peru, altitude decreases markedly from the
composite (May-June 1980) and corresponds edge of the Eastern Cordillera toward the
908 $•brier et al.' Stress in the Andes of Central Peru

ß 26;4' 42 - A

15
18

38 ß

28

0 lO 20 •O 40 50 60 70 80

24ø.3
ø 12".18
o
abc • 42d
1 21 '•::::;:•"••%'-- •/• -•'•-2417

N 14
4• 2932

O lO 20 30 40 50 60 70 60 90 100 O O 20 :10 40 50 60 70 60 90

Fig. 9. Normal fault data from Ayacucho basin (locations on Figures 1 and 7)
used to compute solutions of Table 1. (A) Plotting of o3 axes calculated on
partial data sets from the Ayacucho data. Symbols as on Figure 4.

sub-Andes: 4 km of vertical variation is over wide and weakly deformed synclines.


observed along a 40-km-distance. The sub- Most of these folds have axial
Andean Lowlands of central Peru exhibit terminations on oblique strike-slip faults
clear differences from the southern Pezu (Figure 14). Rocks cropping out in the
ones; they are seismically more active and sub-Andes are of Precambrian to Quaternary
are much wider, forming a NNW-SSE striking ages. Precambrian and Paleozoic exposures
fold and thrust belt, nearly 1000 km long, are generally seen in the central parts of
and about 250 km wide (Figure 1). Their the faulted anticlines, whereas the
altitudes range between 200 and more than Cenozoic deposits crop out within the wide
2000 m, mean elevation being approximately synclines. The sub-Andean belt of central
500 m. Their structure is typically Peru is generally considered to display an
characterized by anticlines upthrusted almost continuous sedimentological record
S&brier et al.' Stress in the Andes of Central Peru 909

75"10'
Due to unfavorable field conditions,
only two areas (Satipo and San Ramon) have
been analyzed. They are located at the
foot of the High Andes and correspond to
piedmont basins that are located on the
fault belt limiting the sub-Andes from the
Eastern Cordillera. These basins are in a
structural location comparable to that of
the Quince Mil and Pillcopata basins of
southern Peru [S6brier et al., 1985]. The
11'50'
Satipo and San Ramon basins are infilled
by undated conglomeratic deposits
Pampa Carnlcero ' attributed to Plio-Pleistocene [S6brier et
al., 1982; Blanc, 1984] because they lie
/! •l::•a
coc
ha 04-06•0
unconformably on folded Mesozoic strata
that are thrusted over Tertiary
continental beds. Structural analyses show
that the oldest, generally tilted,
conglomeratic deposits, considered as
Pliocene to early Pleistocene in age, show
two directions of shortening, N-S and E-W.
In contrast, the youngest conglomeratic
deposits, considered as middle to upper
Quaternary in age, exhibit some few
evidences of a single, roughly E-W
trending, shortening.

4.1. Satipo-Mazamari Basin

ACOPA
The NW-SE trending Satipo-Mazamari
basin (Figure 1, point 47) is an elongated
12'
depression, bounded to the SW by the
0 5 10 Km Eastern Cordillera and to the NE by the
folded ranges of the Ene and Perene rivers
Fig. 10. Location of the 1969 (Figure 14). This depression is partly
Huaytapallana reverse fault (Figure 1, infilled by undated piedmont alluvial
point 41). Numbered star' trenches deposits which exhibit terrace levels
performed across the Huaytapallana fault; cropping out at an elevation from
focal mechanisms' (St) from Stauder [1975] approximately 800 up to 1200 m. In these
and in Philip and M6gard [1977] and (Su) alluvial deposits, which are attributed to
from Suarez et al. [1983] Su-c is a the Plio-Pleistocene [S•brier et al.,
composite solution computed from 1982; Blanc, 1984], two series are
microearthquakes [Suarez, 1982], Su-d is separated: (1) an older one, strongly
the strongest event among them. weathered and composed of yellowish to
reddish torrential conglomerates and (2) a
younger one made of poorly to moderately
from lower Paleozoic to Present and to weathered conglomeratic fan terraces.
have been folded, for the first time, Series 1 is affected by reverse and
during upper Miocene or lower Pliocene strike-slip faults, whereas no clear
[Audebaud et al., 1973]. In fact, for a deformation has been observed in series 2.
long time, angular unconformities and In the Satipo Mazamari area we measured
related strong erosions have been reported striated fault planes at three different
[Koch, 1962; Pardo and Zu•iga, places: Marankiari (Figure 14, point 46),
1973],indicating that compressional and along the road to Puerto Ocopa at km 4
deformations initiated at least during (Figure 14, point 47b) and at km 7.5
Paleogene. The upper Miocene-lower (Figure 14, point 47a). At Marankiari,
Pliocene compressional phase is actually faults were observed in weathered fluvial
the most conspicuous one because it conglomerates possibly equivalent to
corresponds to the major episode of sub- series 1 because they lie unconformably
Andean structuration. over folded Paleogene red beds. Along the
910 S•brier et al.' Stress in the Andesof Central Peru

Fig. 11a

Fig. 11b
Fig. 11. (a) Aerial obliquephotographof the northernsegmentof the 1969
Huaytapallanafault (arrowheads)offseting lagunaQuillacocha(locationon
Figure10). Thenortheastern
sideof this lakehasbeenupliftedby about0.5
m,
area.
displaying a bared light-colored shore. (b) Field view of the samefaulted
S•brier et al.' Stress in the Andes of Central Peru 911

unconformably upon weathered fluvio-


torrential conglomeratic deposits which
2O-

. 3•• 83
41 are tilted southward [Dollfus,
Soulas, 1975, 1978; S•brier
Blanc, 1984]. According to prior
1965•
et al., 1982;

-,• •39
&75•'"3ø
references, the tilted conglomerates are
attributed to Neogene, the overlying
unconformity being related to early
Quaternary compressional deformations
[Blanc, 1984].
Near San Ramon (Figure 14, point 45) we
analyzed, along the San Ramon-Tarma road,
the discontinuous deformation both in the
tilted conglomerates and in the overlying
terrace. Within poorly consolidated
conglomeratic deposits, pressure marks can
be observed between the pebbles;
statistically, these marks indicate the
direction of shortening (i.e., ol).
•o •o &.o •o •o 70 •o •o loo tlo lzo t•o f&.o 1•o 1•o 17o '•oo
Frequently, shearing marks, especially
Fig. 12. Slip vector data measured in striations, are observed on the flat sides
trenches cutting the 1969 Huaytapallana of these pebbles. These striations have
reverse fault (location on Figures 1 and been considered as slips on microfaults
10), used to compute solution 41 on Table affecting a highly discontinuous material
2. Symbols as on Figure 4. and have been analyzed using the same
methods as fault slip data. These methods
have been already used with success to
analyze deformations within conglomeratic
Mazamari-Puerto Ocopa road, faults were formations of the northern Italian
measured within series 1. At the three
Appennine [Fesce, 1987]. In the San Ramon
points, field observations indicate
basin [Dumont, 1985] these analyses yield
clearly that two different kinematics are similar results: two N-S and E-W
present and that they correspond to
roughly N-S and E-W trending shortenings.
In addition, the chronology of striation
crosscuttings shows that here the N-S
shortening postdates the E-W one. In each
site we separated two striation sets
according to field observations. Then we
grouped the Marankiari data with those of
Puente Sonomoro. The older set shows
kinematics that agree with a roughly E-W
to ENE-WSW trending shortening (stereonets
47al and 47bl, Figure 15 and Table 2). The
younger set exhibits kinematics that agree
with an approximately N-S trending
shortening (stereonets 47a2 and 4762,
Figure 15 and Table 2). Therefore the last
faulting observed in the Satipo Mazamari
area is compressional and corresponds to a
N-S trending shortening. However, we think
that this N-S trending shortening is not
the present-day deformation being more
likely of early Quaternary age.

4.2. San Ramon Basin

The San Ramon-La Merced basin (Figure Fig. 13. P and T axes of High Andes focal
1, point 45) is located at a mean altitude mechanisms shown on Figures 1 and 10
of 900 m. It is infilled by a set of three [Stauder, 1975; Suarez, 1982; Suarez et
main torrential fan terraces that lie al., 1983].
912 S•brier et al.' Stress in the Andes of Central Peru
S•brier et al.' Stress in the Andes of Central Peru 913

1A

• SUB.ANDES
7,s•
5• 45
12
7A

6
289.4

•'1

10 13A

?O 50

24 tB

•00 20 30

22
24 X12 33
47al 3sz:•2•" 47a2
15

30 16

19 ,10

282t2•P/
o-/ 29

18 25
23

26 --•-•-<::

•o
1•17
$ o , .23!,0.3,'T"-•..
: : .•.14 : . .
58

0 10 20 ]0 40 50 60 70 80 0 in 20 ]O 40 50 60

•.•••...•47
bl 28 ,'•2_4
47•2

o
•••••• 7/•
3'26
o 1o 20 o 1o 20

Fig. 15. Reverse strike-slip vector data from sub-Andes of central Peru
(locations on Figures 1 and 15) used to compute solutions of Table 2.
Stereonet sub-Andes gives the selected nodal planes of sub-Andean focal
mechanisms shown on Figure 1 and used to compute the sub-Andes tensor of Table
3. Same symbols as on Figure 4.

directions of shortening, similar to those discontinuous deformation has been


affecting the old conglomerates of Satipo- observed, postdating the tilting of the
Mazamari (see section 4.1), are old conglomerates. This deformation is
demonstrated. In addition, a youngest compressional (stereonet 45 on Figure 15)
914 S•brier et al.: Stress in the Andes of Central Peru

TABLE 2. Parameters of the Deviatoric Stress Tensors Computed From Reverse


Strike-Slip Vector Data of the Central Peruvian High Andes and Sub-Andes

Principal Stress Directions


Site ND Latitude, Longitude, al a2 •3 R
S W Azimuth Dip Azimuth Dip Azimuth Dip

41 84 11ø56 f 75ø03 f 075 ø 03 ø 168 ø 32 ø 340 ø 57 ø 0.84


45 25 11ø10 • 75ø25 • 289 ø 04 ø 021 ø 26 ø 190 ø 64 ø 0.65
47a2 15 11ø20 • 74ø31 • 357 ø 12 ø 089 ø 10 ø 220 ø 74 ø 0.71
47al 19 11ø20 f 74ø31 • 282 ø 02 ø 192 ø 08 ø 023 ø 81 ø 0.76
4762 20 11ø21 • 74ø31 • 002 ø 04 ø 272 ø 02 ø 159 ø 85 ø 0.88
47bl 17 11ø21 • 74ø31 f 052 ø 09 ø 320 ø 12 ø 176 ø 74 ø 0.77

These parameters are computed from the reverse strike-slip vector data of
the central Peruvian High Andes (site 41) and Sub-Andes (sites 45, 47al, 47a2,
47bl, and 4762). Same abreviations as on Table 1. Computed data are shown on
stereonets of Figure 12 and 15.

and indicates a N289 ø trending shortening faults provided by these focal mechanisms.
(site 45, Table 2) which is in agreement Computation of the state of stress
with seismic data. requires selection of the seismic plane
from each focal solution. Here we use the
4.3. Sub-Andean Focal Mechanisms Carey-Gailhardis • inverse method. This has
some similarities with the Gephart and
The boundary between the sub-Andean Forsyth•s [1984] technics. It is described
zone and the High Andes is seismically in detail in Carey-Gailhardis and Mercier
very active. In this region 14 focal [1986] and has been used with success for
mechanisms are available (Figure 1): 12 analysis of Tibetan earthquakes [Mercier
proceed from the WWSSNdata and correspond et al., 1987a].
to six distinct events [Stauder, 1975; In order to take into account the scale
Suarez et al., 1983] and two have been variations of the seismic ruptures we have
obtained from a local network that was weighted the data according to their
operating in 1980; one of them is magnitude. Inversion of these data
composite [Suarez, 1982]. All these (stereonet sub-Andes, Figure 15) yields a
seismic events correspond to compressional N86 ø trending compressional axis and an
deformations, i.e. P axes are almost vertical minimum principal stress
suhorizontal and trend statistically E-W axis (Table 3). This calculation
(except event Su 16); T axes are nearly corroborates that the present sub-Andean
vertical (except St 43). We attempt to compression strikes nearly E-W as
search a mean regional state of stress previously proposed [Stauder, 1975; Suarez
which explains kinematics of the seismic et al., 1983]. These results are confirmed

%ABLE 3. Parameters of the Deviatoric Stress Tensors Computed From the Focal
Mechanisms of the High Andes, Pacific Lowlands, and Sub-Andes

Principal Stress Directions


Area NS ND •1 •2 •3 R
Azimuth Dip Azimuth Dip Azimuth Dip

Sub-Andes 7 14 086 ø 01 ø 356 ø 11 ø 177 ø 79 ø 0.71


Forearc 10 12 068 ø 02 ø 158 ø 02 ø 298 ø 87 ø 0.82

These parameters are computed from one of the pair of nodal planes from
focal mechanisms of earthquakes located in the Sub-Andes and Forearc. NS is the
number of earthquakes, and ND the number of available solutions. Other
abreviations as on Table 1. Computeddata on stereonet Sub-Andes and Forearc of
Figures 15 and 16.
S•brier et al.' Stress in the Andes of Central Peru 915

I JI •l 3l 40

I(•.•1

N
930 ^8•N,2.
12

FOREARC
13
^

23

2^

s 173'2'
•2A 10A
1A

• lO 30 0 I0 20
I,•.•)1
Fig. 16. Normal fault data from Pacific coast {locations on Figure 1• used •o
compute solutions of •able 1. Symbols as on Figure •. Stereonet CNF' pole
density of unstriated normal faults observed along the Pacific coast of
central Peru, numbers, pole amount per 1% area. Stereonet forearc' selected
nodal planes from Forearc focal solutions shown on Figure 1 and used to
compute the regional azimuth of the maximum •1 principal stress axis
(convergent large black arrows).

by recent seismological studies [Dorbath Peru [Atherton et al., 1983]. The emerged
et al., 1986] that have been undertaken in coastal domain on which we have performed
the sub-Andes of central Peru with a local field observations is a 30-km narrow
seismic network. strip. This is partly covered with a
discontinuous string of Quaternary
5. EXTENSIONAL AND COMPRESSIONAL alluvial fans, located around the mouths
DEFORMATION IN THE PACIFIC LOWLANDS of the main rivers which flow down from
the High Andes.
In central Peru the prolongation of the
southern Peru Pacific Lowlands is mainly 5.1. Forearc Focal Mechanisms
below the sea level and constitutes an
approximately 100-t•m-wide continental The forearc is a highly active seismic
shelf. Since this depr,•ssed feature has area which corresponds to the shallower
existed for the last 40 Ma, it is part of the Andean Benioff zone. Beneath
interpreted [Machar• et al., 1986] as the the forearc, seismic foci defined a 30 ø
consequence of an anomalously dense dipping Benioff zone [Snoke et al., 1979;
crustal structure related to a lower Suarez, 1982]. Focal mechanisms are
Cretaceous "back arc" basin which extended very roughly grouped into two dipping
along the present coastline of central parallel layers: a deeper one with normal
916 S•brier et al.' Stress in the Andes of Central Peru

Fig. 17. Field view, looking west, of the N125øE trending Montejato normal
fault near Ca•ete (Figure 1, point 27) which cuts Late Eocene Paracas
Formation (white) and Early Quaternary Ca•ete Formation (dark). Vertical
offset is about 20 m.

fault mechanisms located chiefly in the discrepancy could be due to the tendency
oceanic slab and an upper one with of the slab to plunge orthogonally to the
different reverse and thrust type focal plate boundary orientation [Scotese and
mechanisms located in the vicinity or at Rowley, 1985] .
the plate interface [Stauder, 1975; Suarez
et al., 1983]. Here we only consider the 5.2. Surface Faulting Along the
12 thrust type focal mechanisms (Figure Pacific Coast
1), corresponding to 10 earthquakes [Abe,
1972; Stauder, 1975; Dewey and Spence, During Quaternary, the onshore Pacific
1979; Pennington, 1981; Suarez et al., coast of central Peru is a subsiding
1983]. Assuming that all the considered coastline: Pleistocene deposits are mainly
seismic foci are not restricted to the continental. Generally, no stepped marine
plate interface, we have computed a state terraces are observed [S&brier and
of stress using data (stereonet forearc, Machar•, 1980; S•brier et al., 1982;
Figure 16) weighted according to the Machar• et al., 1986]. The only exception
earthquake magnitude. Computation yields a is observed at San Lorenzo Island (point
N68øE trending compression and an almost 25 on Figure 1) which is interpreted as an
vertical minimum stress axis (Table 3). upthrown block separated from the Lima
The gently eastward dipping seismic planes harbor of Callao by an upper Pleistocene
(stereonet Forearc, Figure 16) are fault [S•brier and Machar•, 1980]. The
probably located at the plate interface, significance of emerged Holocene marine
whereas the other ones, markedly deposits is still discussed. Their
different, possibly occur within the constant elevation, between approximately
deeper part of the continental forearc 2 and 4 m, along most of the Peruvian
wedge. The resulting compressional axis coast is not in favor of an Holocene
shows a slight difference from the N80øE tectonic uplift but rather of a post
convergence direction between the Nazca glacial isostatic rebound of the medium-
and South American plates [(Minster and and low-latitudinal coasts caused by
Jordan, 1978]). If significant, this small overburden of the oceanic lithosphere with
S•brier et al.: Stress in the Andes of Central Peru 917

ice-melting water [Clark, 1980]. On the 30, Figure 16 and Table 1). The ratio
contrary, southern and northwestern Peru value, R=0.92, indicates a nearly radial
are characterized by flights of maripe extension in the Rio Grande area. This is
terraces that crop out south of Pisco and in agreement with the R=0.86 value (site
north of Chiclayo, respectively [Bosworth, 1, Table 1) calculated from faults of the
1922; Laharie, 1970; S•brier, 1978; Marcona area (Figure 1, point 1) located
S•brier et al., 1979, 1982; DeVries, 1984: on the southern border of the coastal

Machar• et al., 1986]. uplift associated with the Nazca ridge


Along the Pacific coast of central [S•brier et al., 1985]. These results seem
Peru, Quaternary normal faults have been to correspond to a regional effect related
mentioned with throws ranging between to an oval-shaped updoming in the edge of
several centimeters and several meters the overriding plate predicted by
[Soulas, 1978; S•brier and Machar•, 1980• numerical modeling of the buoyant Nazca
Machar•, 1981; Machar• et al., 1986]. ridge subduction [Moretti, 1982]. It must
These normal faults affect very poorly be emphasized that subduction of the
consolidated deposits and do not generally buoyant Nazca ridge produces a stress
exhibit striated siickensides. Normal field radically different from those
faults we observed along the Pacific coast reported for collisions [Tapponnier and
(Figure 1) have directicns that are Molnar, 1976].
oblique to orthogonal to the coast trend. Although normal faulting is the most
This is confirmed by the plotting of their conspicuous deformation along the Peruvian
poles on a stereonet (CNF, Figure 16). coast, small reverse faul•'s have been also
This fault geometry is in favor of a very reported in deposits that are attributed
roughly N-S trending extension. On the to middle or late Quaternary [S•brier,
coast of NW Peru the WSW-ENE and WNW-ESE 1978; Machar•, 1981; S•brier et al., 1982;
trending Quaternary normal faults that Machar• et al., 1986]. These are few and
have been reported in the Cabo Blanco area isolated (Figure 1, points 21, 23, 24, and
(DeVries [1984]; Figure 1, point 20) are 28). Their throws are of the order of 10
also in agreement with our observations. cm, and their directions appear either
Striated slickensides have been roughly parallel or orthogonal to the
observed in the Montejato graben, 5 km to central Peruvian coast. The most recent

the SE of Ca•ete (Figure 1, point 27); at evidence of reverse faulting is observed


Playa Jahuay, km 180 of the Pan-American along the northern part of the N160øE
highway (point 28); and at Cerro Caucato, trending Illescas fault (Figure 1, point
7 km to the NNE of Pisco (point 29). At 21), where several small reverse faults
Montejato, normal faults offset the cut debris flows related to an upper
conglomeratic deposits of the early Pleistocene marine terrace [S•brier,
Quaternary Ca•ete formation (Figure 17). 1978]. The coastal reverse faults do not
At Cerro Caucato, fault planes cut the exhibit good and/or enough striated
Pliocene Pisco
formation, and normal slickensides; nevertheless, they indicate
striations are
posterior to reverse ones that the tectonic regime may occasionally
corresponding to an E-W trending change on the Peruvian coast, as there is
compression of early Quaternary age. At no field evidence indicating that these
Playa Jahuay, normal faults affect sandy small reverse faults might be an effect of
conglomeratic beds of the Pleistocene sliding in unconsolidated sediments. The
Topara formation. Computation of these small magnitudes of the coastal, either
data yields a N19øE trending minimal compressional or tensional, deformations
principal stress (site 27, Table 1). show that, as in southern Peru, the
In the Rio Grande area (point 30 on coastal state of stress at the surface is

Figure 1) we measured striated fault almost neutral with a tendency to a N-S


planes: at the Rio Santa Cruz confluence, trending extension.
1 km to the West of Cabildo; on the left
riverbank, 1 km to the SE of Cabildo; and 6. STRESS PATTERN, HIGH TOPOGRAPHY, AND
at Malpaso, 5 km to the SSW of Cabildo. FLAT SUBDUCTION IN CENTRAL PERU
All these data have been observed in the
marine Pisco formation of upper Miocene- 6.1. Present-Day Stress Pattern
Pliocene age. They postdate early in Central Peru
Quaternary compressional movements
[Machar& et al., 1986] and are thus On both side of the High Andes (Figure
representative of the Quaternary state of 18), i.e., at the contact between the
stress. Inversion of data gives a N173øE Nazca and South American plates, and in
_trending extensional (•3) axis (streonet the sub-Andean Lowlands, tectonics is
918 S•brier et al.' Stress in the Andes of Central Peru

.•,• ffUANCABAMBA
ß
i!i.i:
I. aPlURA

71::.:,•:..,....
oC
HIC
o

28 4•

o
I

16 ø

Fig. 18. Principal directions of extension and compression deduced from


structural analysis of Quaternary and active faults of central Peru (arrows
attached to filled circles) Numbered sites as in Figure 1. Circles with
principal stress axes correspond to stress regimes computed (forearc and sub-
Andes) from available focal mechanisms shown on Figure 1. Beneath the Forearc,
in the contact zone between the two plates, ENE-WSWcompression occurs. At the
surface (Pacific coast) and in the Western Cordillera, N-S striking, regional
extensional direction appears to be roughly orthogonal to the direction of
convergence (large white arrow) between the two plates [Minster and Jordan,
1978]. In the Eastern Cordillera the few structural and focal mechanism data
suggest strike-slip faulting with E-W trending compression and N-S trending
extension. In the sub-Andes, both structural and seismic data show an E-W
trending compression. The 3000-m line separates the High Andes (Western and
Eastern Cordilleras) from lowlands; stippled areas have an altitude higher
than 5000 m. Double thick dashed line is the vertical projection of the
transition zone between the flat dipping slab and the 30 ø dipping slab [Grange
et al., 1984a].
S•brier et al.' Stress in the Andes of Central Peru 919

•3

Trench
0"2 •.2•r3 •'2 G2

Coastal Lowlands 0'1 •3


•2
• Western Cordillera ffl
' i
,
Eastern
Cordillera
Subandean Lowlands

Brazilian Shield
/

ß ,
ß

0'1

0'2 •'2
•'3 3

O'3
0"3
0'2 0'2

•2

Fig. 19. Tentative model to explain the state of stress in the Andes of
central Peru. Principal lithospheric stress (set of thin solid arrows) in
excess to the reference (sea level) lithostatic stress are oxx (striking E-W),
oyy (striking N-S), and ozz (vertical); oxx and oyy are considered as fairly
constant; ozz amount to the weight of the topography. Convergence between
Nazca and South American plates (thick arrows) is roughly parrallel to oxx.

compressional: oxx striking E-W is ol, oyy topography [Dalmayrac and Molnar, 1981;
striking NoS is o2, and ozz is o3. Thus Froidevaux and Isacks, 1984]. The oxx
the state of stress is such that axis, roughly parallel to the convergence,
oxx>oyy>ozz (Figure 19). In the Western is considered as fairly constant and
Cordillera and along the Pacific coast, N- remains oHmax. Therefore ozz increasing
S trending extensional tectonics occurs with the topography is o3 in the sub-
(Figures 18 and 20): ozz becomes ol, oxx Andes, and ol in the Western Cordillera.
is o2, and oyy is o3. The state of stress Tensional tectonics in the coastal
is such that ozz>oxx>oyy. In the Eastern lowlands might be due to the topographical
Cordillera, data are few but suggest that effect of the nearby deep trench as
compressional strike-slip deformation suggested for the Aegean Arc [Mercier et
occurs. In this case, ozz is o2, oxx is al., 1987b].
ol, and oyy is o3; the state of stress is However, a major difference appears
such that oxx>ozz>oyy (Figure 19). As in between southern and central Peru.
southern Peru [S•brier et al., 1985], the Compressional strike-slip faulting is
major features of the state of stress in known in the High Andes of central Peru.
central Peru may be interpreted as a This has led some authors to consider
result of boundary forces due to compressional tectonics as the typical
convergence and of body forces due to state of stress in the Andean lithosphere
920 S•brier et al.: Stress in the Andes of Central Peru

Indeed, this transition might occur at a


lower altitude if, as suggested by Kono et
al. [1986], the Eastern Cordillera is
© © undercompensated and since its elevation,
according to Airy isostasy, should be
about 3000 m. Normal faults affecting the
Ayacucho basin (see section 2.3), whose
mean elevation is 3700 m, seem to support
the above assumption. However, these
normal faults do not exhibit any
characters of Recent activity. Therefore
it could be also assumed that normal
faulting results from a previous
Quaternary state of stress that should
correspond to a regional setting
characterized by volcanic activity and
higher topography, similar to the present-
day southern Peru one. This is supported
by the fact that the volcanic activity has
stopped after 2.5 Ma in the Ayacucho
basin, while it stopped earlier,
approximately between 5 and 6 Ma, to the
north of it. Therefore an asthenospheric
wedge may have existed beneath the
continental lithosphere of the Ayacucho
basin region up to the Plio-Pleistocene
limit. Finally, the elevation at which the
Fig. 20. Principal stress directions change from a compressional to an
computed from slip vectors of Quaternary extensional tectonic regime occurs is not
and Recent normal faults of the High precisely defined, due to the scarcity of
Andes, mainly Western Cordillera, and the data from the Eatern Cordillera.
Pacific coast of central Peru. Recent
Extensional deformation of the High
Quaternary faulting in these regions Andes postdates early Quaternary
results from a regional extension striking compressional deformations [S•brier and
between N10øW and N26øE.
Machar•, 1980; S•brier et al, 1982; Blanc,
1984; Bonnot, 1984; Machar• et al , 1986'
Cabrera
et al., 1987;Bonnot
et ai.,
situated above flat slab segments [M•gard 1988]. Although these early Quaternary
and Philip, 1976]. Indeed, in the deformations are not precisely dated, this
Cordillera de Huaytapallana (Eastern indicates that the present-day state of
Cordillera), compressional tectonics is stress has probably been active during the
observed 4700 m above the sea level, but last 1 or 2 Ma. The magnitude of the
in the Cordillera Bianca (Western Quaternary and present-day deformations
Cordillera), normal faulting is clearly has not been calculated because we did not
seen at a lower elevation between 3600 and survey the whole of central Peru. Along
4300 m. This seems in contradiction with the western border of the Cordillera
the above mentioned model. In fact, such a Bianca, where normal faulting exhibits the
model concerns large mass distribution at greatest offsets, Quaternary stretching
a wavelength on the order of the could reach up to 8% [Bonnot, 1984; Bonnot
lithosphere thickness [Fleitout and et al., 1988]. However, this is restricted
Froidevaux, 1982]. Thus topography has to to the Callejon de Huaylas graben and thus
be smoothed at a 100-km wavelength and the magnitude of stretching is not
then it appears (Figure 21) that in the representative of all of the High Andes.
Western Cordillera the mean altitude is The N-S trending extension probably does
4200 m, whereas it is only 3700 m in the not produce a more significative amount of
Eastern Cordillera. Roughly at a mean stretching than in southern Peru, where
elevation higher than 4000 m, tectonics is stretching has been estimated to be of the
tensional, whereas below, compressional order of 1% during the last 1 or 2 Ma
strike-slip tectonics occur; azz [S•brier et al., 1985]. In the same
transforms from a a2 axis into a •1 axis manner, shortening due to the
between 3700 and 4000 m of mean elevation. Huaytapallana fault appears to be small,
S•brier et al.' Stress in the Andes of Central Peru 921

sw NE

6OOO

4000

0 ................................... : : ; : . . .

0 ]00 200 300 400

SW NE

$ooo

4000

2000

O 100 200 500 KM

SW NE

60OO

4OO0
¸
2000

0 100 200 :•00 400 KM

Fig. 21. Selected topographic profiles through the Andes of central Peru
constructed taking a point each 5-km distance; data taken from Peruvian IGN
topographic maps at a scale of 1:100,000 except for part of the Eastern
Cordillera and sub-Andes, for which map at a scale of 1:1,000,000 has been
used. In order to determine the mean topographical elevation of the transect a
50-km-long window has been moved along the profile with a step of 5 km. For
each step the mean elevation is that given by the arithmetic mean of the 10
values included within the window. The dashed lines are the mean values of
these mean elevations. Profile locations on Figure 1: (a) Cordillera Bianca
transect, (b) Huancayo basin transect, (c) Ayacucho basin transect.

as this fault has a high dip and a strike- Andes. Pardo considers that the first
slip component. On the coastal area the compressional phase which affected the
magnitude of superficial deformations is sub-Andes is of upper Miocene age. Thus in
very small [S•brier and Machar•, 1980; a rough approximation, for the last 6 Ma,
Machar• et al, 1986], even smaller than in the mean shortening rate should range
the High Andes. Due to the poor geological between 1.6 and 3.3 mm per year.
knowledge of the sub-Andes it is even more Nevertheless, if we assume that the
difficult to specify shortening that was shortening rate must have been higher both
produced by Quaternary compressional during upper Miocene and early Quaternary
deformations. The main difficulty arises compressional tectonic pulses, present-day
from the unprecisely known geometry of the deformation rate of the order of 1-2 mm
sub-Andean faults. However, compared with per year appears reasonable.
the High Andes seismicity, sub-Andean
seismic activity is high and shortening 6.2. High Topography and Benioff Geometry
must be significant. Taking into account
the seismic moments, Suarez et al. [1983] The mean elevation of several profiles
calculated a shortening rate ranging has been calculated using a window moving
between 1 and 2 mm per year. Thus 1-4 km along two-dimensional profiles on a
of shortening would have been produced topographic map (Figure 21). A more
during the Quaternary and the present-day precise calculation has been realized on a
period. According to Pardo [1982], 10-20 test zone in the Cuzco region (J. Cabrera,
km of shortening has affected the manuscript in preparation, 1988) using a
northernmost part of the Peruvian sub- window moving along three-dimensional
922 S•brier et al.: Stress in the Andes of Central Peru

profiles having a width of 30 km in the effect of the overriding continental


horizontal plane; the results are similar lithosphere on the subduction geometry.
to the previous ones within uncertainities Subduction of the aseismic Nazca ridge
of +50 m in elevation. The results show prolongation beneath central Peru appears
that the southern Peru segment situated to be a possible factor to explain
above the 30 ø dipping slab has a mean differences in slab geometry and
elevation of 4300 m, whereas the one in consequently, in the continental
central Peru situated above a flat dipping deformation pattern.
slab has only a 4000 m mean elevation
(Figure 21). Thus there is a difference in 6.3. High Topography and Crustal Thickening
mean elevation of roughly 300 m;
consequently, the •zz value must be lower Andean uplift is essentially of Miocene
in central Peru than in southern Peru. In age [S•brier et al., 1988]. The above
addition, along the flat subduction of mentioned main features of the Andean
central Peru [Barazangi and Isacks, 1976; topography, i.e., mean elevation higher in
Hasegawa and Sacks, 1981; Suarez, 1982; central than in southern Peru and Eastern
Grange et al., 1984a] the oceanic slab Cordillera lower than Western Cordillera,
should be more coupled to the overriding may give some new constraints to this
continental lithosphere thus producing an uplift process. Seismic studies
increase of the •xx value. Therefore both [Cunningham et al., 1986] show that
effects tend to increase compressional crustal thickness could attain 70 km under
deformations in the central Peruvian the Western Cordillera. Several processes
continental lithosphere. This can explain have been put forward to explain such a
why compressional strike-slip tectonics thick crust [Allmendiger, 1986): (1)
may occur in the High Andes of central tectonic consumption of the Pacific margin
Peru whereas it has not been observed in of South America [Rutland, 1971; Mortimer,
the•igh Andes
of southern
Peru. 1972; Hussong and Wipperman, 1981]; (2)
The transition between central and suspect terranes collage [Nur and Ben-
southern Peru occurs in the Abancay Avraham, 1982; Howell, 1986]; (3) magmatic
deflection zone. This transition appears accretion [James, 1971]; (4) crustal
to be gradual, as a conspicuous active shortening by upright isoclinal folding
fault belt does not exist in this zone. [Helwig, 1973] or by continental
These field observations are better in subduction of South America beneath the
agreement with a contortion than with a High Andes [Suarez et al., 1983]. Andean
tearing of the slab. Moreover, rare faults shortening along the central Peru transect
of Quaternary age observed along the has been estimated to be of the order of
Abancay deflection (Figure 1, point 45) 100 km during the whole Cenozoic [M•gard,
are characterized by northern upthrown 1984], this not being sufficient to create
blocks, which are due more probably to a crustal root that could compensate a 4-
differential movements between the Western km-high topography. In the Central Andes,
and Eastern Cordilleras than to an paleomagnetic results [Shackleton et al.,
accommodation of the continental 1979; Knight et al., 1983; May and Butler,
lithosphere above a slab tearing. 1985] and offshore and field studies along
Various factors may control the slab the Peruvian margin [Von Huene et al.,
geometry: (1) anomalously thick slab 1985; Machar• et al., 1986] imply that
corresponding to an aseismic ridge processes 1 and 2 should have none or
[Pilger, 1981]; (2) lateral variation in negligible effects on crustal thickening
the age of the slab [Molnar and Atwater, during Cenozoic. Process 3 has been
1978; Sacks, 1983]; (3) differences in estimated insufficient to account for the
convergence velocity; and (4) differences whole Andean root [Francis and Rundle,
in the composition of the continental 1976; Baker and Francis, 1978]. A lower
lithosphere. It is unlikely that factors 2 elevation of the Eastern Cordillera is not
and 3 could explain the contrast between in agreement with process 4 acting alone,
southern and central Peru because these for if continental subduction had occurred
factors do not appear significantly during Late Neogene-Pleistocene in the
different in these two regions [Mammerickx sub-Andes, it must have thickened the
et al., 1975, 1980; Handschumacher, 1976; Eastern Cordillera crust. The fold and
Minster and Jordan, 1978] . Concerning thrust belt that separates the Western
factor 4, many differences have existed from the Eastern Cordillera (i.e., Mara•on
between central and southern Peru for a imbricated zone of Wilson et al. [1967]
long time, but little is known about the has been also considered as a major thrust
S&brier et al.: Stress in the Andes of Central Peru 923

zone of continental subduction beneath the Molnar, 1981• Froidevaux and Isacks, 1984•
Western Cordillera [Bourgois and Janjou, S•brier et al., 1985]. In this model
1981]. It was active mainly during (Figure 19) the vertical stress •zz
Paleogene and hardly explains the present- increases with the topography, and •Hmax
day greater mean elevation of the Western is considered fairly constant and trends
Cordillera. In fact, the area of major E-W roughly parallel to the convergence.
magmatic activity during the last 30 Ma Therefore in the sub-Andean Lowlands,
(in central Peru, volcanism was active tectonics being compressional, •zz is •3
till 5 Ma b.p.) has the highest mean and •Hmax is •1. In the High Andes, •zz
elevation. In addition, the highest mean becomes •1 and •Hmax is •2; thus the third
elevation of southern Peru suggests that axis oHmin is •3 and trends N-S, allowing
asthenospheric uplift could partly extension to occur along this direction.
compensate the Andean topography. On the coastal lowlands, tectonics is not
Therefore it seems that magmatic accretion compressional as expected• the nearly
and/or asthenospheric uplift, related to neutral state of stress may be due to an
the subducting process, add to crustal effect of topography related to the nearby
shortening as causes of the high Andean deep Peru-Chile trench.
topography and consequently, of However, a conspicuous difference
extensional tectonics in the High Andes. appears between the Andes of central and
southern Peru. Compressional tectonics
7 . CONCLUSIONS seems to affect the High Andes of central
Peru but not those of southern Peru.
In the Andes of central Peru, field Therefore it had been considered that
studies and focal mechanisms make it above a flat slab the overriding plate is
possible to analyze the state of stress in submitted to compressional tectonics
an overriding plate situated above a flat [M&gard and Philip, 1976]. Indeed,
slab. These studies show the following compressional strike-slip faulting in the
(Figure 19): Eastern Cordillera may also be explained
1. In the sub-Andes and at the contact by an effect of topography. The mean
between the Nazca and South American elevation of the Eastern Cordillera is
plates, deformations result from an E-W about 3700 m, whereas it is about 4200 m
trending compression roughly parallel to in the Western Cordillera; because the
the convergence between the two plates, as Eastern Cordillera is undercompensated,
shown by focal mechanisms of earthquakes. this elevation should be lower in an
In the sub-Andes, field data demonstrate isostatic equilibrium. Thus the state of
compressional tectonics, and the most stress in the Eastern Cordillera is
recent Quaternary deformations are in probably intermediate between those of the
agreement with an E-W trending shortening. Lowlands and of the Higher Lands: •zz
2. In the High Andes, two tectonic being •2, •Hmax being •1, and oHmin being
regimes have been shown. In the Western •3 so that we have •xx>•zz>•yy (Figure
Cordillera, Recent and active deformations 19). This state of stress in the High
result from normal faulting. The 200-km- Andes of central Peru probably results
long fault zone of the Cordillera Bianca from two cumulative effects: (1) due to
has a kinematics which agrees with a N-S the flat slab geometry, coupling between
trending extension. In the Eastern the two plates is expected to be stronger,
Cordillera, seismicity and active so that •xx value must be higher, allowing
compressional strike-slip faults result compression at a higher mean elevation and
from a N-S trending extension plus an E-W (2) the mean elevation of the Andes of
trending compression. central Peru is 300 m lower in respect to
3. On the Pacific coast, Quaternary that of southern Peru. Both effects tend
faulting results mainly from a N-S to make tectonics more compressional. The
trending extension, but some rare small altitudinal threshold that controls
reverse faults have also been observed. change, from strike-slip to normal
The state of stress seems to be nearly faulting, lies below 4000 m in central
neutral. Peru. Because of the dense tropical jungle
The main features of the state of of the Amazonian foothills it is
stress in the Andes of central Peru are impossible to observe the tectonic change
similar to those of the Andes of southern from thrust to strike-slip faulting.
Peru [Mercier, 1981• S&brier et al., 1985] According to the evidence of thrusting or
and may be interpreted as an effect of reverse faulting in the sub-Andes of
compensated high topography [Dalmayrac and central Peru, this change must take place
924 S•brier et al.' Stress in the Andes of Central Peru

higher than at 1000 m in elevation. In a motion prediction, J. Geophys. Res., 89,


similar geological setting, i.e., in the 5867-5872, 1984.
Argentine Andes situated above the 28 ø- Allmendiger, R., Tectonic development,
33øS flat subducting segment, the E1 Tigre southeastern border of the Puna plateau,
fault [Bastias et al., 1984] which crops northwestern Argentine Andes, Geol. Soc.
out at a mean elevation of 2000 m, has a Am. Bull., 97(9), 1070-1082, 1986.
strike-slip movement. This may suggest Atherton, M.P., W. S. Pitcher, and V.
that the change from thrust to strike-slip Warden, The Mesozoic marginal basin of
faulting, i.e., the change of •zz axis cental Peru, Nature, 305(5932), 303-305,
from a •3 to a •2 value, should take place 1983.
between 1000 and 2000 m. Audebaud, E., R. Capdevila, B. Dalmayrac,
Analysis of the stress pattern in the J. Debelmas, G. Laubacher, C. Lefevre,
Peruvian Andes shows that deformations R. Marroco, C. Martinez, M. Mattauer, F.
related to the subduction of the buoyant M•gard, and J. Paredes, P. Tomasi, Les
Nazca ridge are restricted to a 200-km- traits geologiques essentiels des Andes
long (NW-SE), 40-km-wide (NE-SW), coastal centrales (Perou - Bolivie), Rev. G•ogr.
updoming, which is located opposite to Phys. G•ol. Dyn., 15(1-2), 73-114, 1973.
this ridge. There deformations are Baker, M. C., and P. W. Francis, Upper
extensional, as they are all along the Cenozoic volcanism in the Central Andes'
Pacific coast of Peru. Therefore Ages and volumes, Earth Planet. Sci.
subduction of the buoyant aseismic Nazca Lett., 41, 175-187, 1978.
ridge is, in any case, comparable to a Barazangi M., and B. Isacks, Spatial
collisional setting where compressional distribution of earthquakes and
tectonics is very active. subduction of the Nazca plate beneath
Finally, the magnitude of deformations South America, Geology, •(11), 686-692,
1976.
appears to be small in the High Andes, and
actually, most of the Andean deformations Barazangi, M., and B. Isacks, Subduction
take place mainly in the two compressional of the Nazca plate beneath Peru'
zones that are located on both edges of Evidence from spatial distribution of
the High Andes: (1) the contact between earthquakes, Geophys. J. R. Astron.
the oceanic Nazca and the South American Soc., 57, 537-555, 1979.
plates and (2) the limit between the hot Bastias, H. E., N. E. Weidmann, and A.M.
lithosphere of the Oligo-Miocene Andean Perez, Dos zonas de fallamiento plio-
cuaternario en la Precordillera de San
magmatic arc, corresponding to the High
Andes and the comparatively colder Juan, in Proceedings Noveno Congreso
lithosphere of cratonic South America, Geo10gico.Argentino, vol. 2, pp.329-341,
i.e., the Lowlands of the Brazilian Bariloche, 1984.
shield. Bevis, M., and B. Isacks, Hypocentral
trend surface analysis ß Probing the
Acknowledgments. Field work has been geometry of Benioff zones. J. Geophys.
supported by Instituto Geofisico del Peru, Res., 89, 6153-6170, 1984.
ORSTOM Lima (Institut Frangais de Blanc, J. L., Etude n•otectonique et
Recherche Scientifique pour le sismotectonique des Andes du P•rou
D•veloppement en Cooperation), and ATP central dans la r•gion de Huancayo,
G•odynamique II and ASP Blocs et thesis, 158 pp., Univ. Paris-Sud, Orsay,
Collisions (Institut National des Sciences 1984.

de l'Uaivers). We are grateful to Servicio Blanc, J. L., M. S•brier, and J. Cabrera,


Estudio microtectonico de la falla
Aerofctografico National del Peru which
authorized publication of aerial sismica de Huaytapallana (Andes del Peru
photographs. central), Rev. Geofis, Instituto
Panamericano de Geografia e Historia,
Mex., 18/19, 5-24, 1983.
REFERENCES Bonnot, D., N•otectonique et tectonique
active de la Cordill&re Blanche et du
Abe, K., Mechanisms and tectonic Callejon de Huaylas (Andes nord-
implications of the 1966 and 1970 Peru p•ruviennes), thesis, 96 pp., Univ.
earthquakes, Phys. Earth. Planet. Paris-Sud, Orsay, 1984.
Inter., •, 365-379, 1972. Bonnot, D., and M. S•brier, Analisis
Aki, K., Asperities, barriers, cinematico y sismogenico de un sistema
characteristic earthquakes and strong de fallas normales activas' ejemplo de
S•brier et al.' Stress in the Andes of Central Peru 925

la Cordillera Blanca (Peru), in Dalmayrac, B., Un exemple de tectonique


Proceedings VI Congreso Geologico vivante ß les failles subactuelles du
Venezolano, vol. 4, pp. 2378-2396, pied de la Cordill&re Blanche (P•rou),
Caracas, 1985. Cah. ORSTOM• S•r. G•ol., •(1), 19-27,
Bonnot D., M. S•brier, and J. L. Mercier, 1974.
Evolution g•odynamique plio- quaternaire Dalmayrac, B., and P. Molnar, Parallel
du bassin intraocordill•rain du Callejon thrust and normal faulting in Peru and
de Huaylas et de la Cordill&re Blanche, constraints on the state of stress,
G•odynamique,3(2), in press, 1988. Earth Planet. Sci. Lett., 55, 473-481,
Bosworth, T. O., Geology of the Tertiary 1981.
and Quaternary periods in the De Vries, T., Neotectonica del area de
northwestern part of Peru, 434 pp., Mc Cabo Blanco, Noroeste del Peru, Bol.
Millan Co., London, 1922. Soc. Geol. Peru, 73, 1-14, 1984.
Bourgois, J., and D. Janjou, Subduction Dewey, J. W., and W. Spence, Seismic gaps
oc•anique, subduction continentale et and source zones of recent larges
surrection andine' lWExemple du P•rou earthquakes in coastal Peru, Pure Appl.
septentrional, C. R. Hebd. S•ances Acad. Geophys., 117(6), 1148-1171, 1979.
Sci., 293, 859-864, 1981.
Deza, E., The Pariahuanca earthquakes,
Cabrera, J., Estratigrafia y neotectonica
Huancayo, Peru' July-October 1969, Bull.
del SW de la cuenca de Huancayo, thesis,
R. Soc. N. Z., •, 77-83, 1971.
140 pp., Univ. Nac. San Agustin,
Dollfus, O., Les Andes centrales du P•rou
Arequipa, 1982.
Cabrera J., M. S•brier, and J. Lo Mercier, et leurs piedmonts (entre Lima et le
P•ren•), Tray. Inst. Fr. Etud. Andines,
Active normal faulting in the High
Plateaus of Central Andes' The Cuzco 10, 404 pp., 1965.
region (Peru), Ann. Tectonicae, •(2), Dollfus, O., and F. M•gard, Les formations
116-138, 1987. quaternaires du bassin de Huancayo et
Carey, E., Recherche des directions leur n•otectonique (Andes centrales
principales de contraintes associ•es au p•ruviennes), Rev. G•ogr. Phys. G•ol.
jeu dWune population de failles, Rev. Dyn., 10(5), 429-440, 1968.
G•ogr. Phys. G•ol. Dyn., 21(1), 57-66, Dorbath, C., T, Dorbath, A. Cisternas, J.
1979. Deverch&re, M. Diament, L. Ocola, and M.
Carey, E., and B. Brunier, Analyse Morales, First results on crustal
th•orique et num•rique d•un mod&le seismicity of the Amazonian foothill of
m•canique •l•mentaire appliqu• • l•tude the central Peruvian Andes, Geophys.
d•une population de failles, C. R. Hebd. Res. Lett., 13(10), 1023-1026, 1986.
S•ances Acad. Sci., 279, 891-894, 1974. Dumont, J. F., First results on the
Carey-Gailhardis, E., and J. L. Mercier, A tectonics of the San Ramon basin (Sub-
numerical method for determining the Andes of central Peru), paper presented
state of stress using focal mechanisms at Workshop on Seismotectonics of
of earthquakes populations' application Central and Southern Andes, Centro
to Tibetan teleseisms and Estudios Cientificos de Santiago, Chile,
1985.
microseismicity of Southern Peru, Earth
Planet. Sci. Lett., 82, 165-179, 1987. England, P., and R. Wortel, Some
Chinn, D. S., and B. L. Isacks, Accurate consequences of the subduction of young
source depths and focal mechanisms of slabs, Earth Planet. Scio Letto, 47,
403-415, 1980.
shallow earthquakes in Western South
America and in the New Hebrides Island Fesce, A., Analisi mesostrutturale sui
Arc, Tectonics, •(6), 529-563, 1983. ciottoli di conglomerato e su superfici
Clark J. A., A numerical model of di faglia della Formazione a Colombacci
worldwide sea level changes on a lungo del valle del Bidente (Forli),
viscoelastic earth, in Earth Rheology, Rend. Soc. Geol. It., 10, 41-43, 1987.
Isostasy and Eustasy, edited by N. A. Fleitout, L., and C. Froidevaux, Tectonic
M6rner, pp.525-534, John Wiley, New and topography for a lithosphere
York, 1980. containing density heterogeneities,
Cunningham, P.S., S. W. Roecker, and D. Tectonics, •(1), 21-56, 1982.
Hatzfeld, Three-dimensional P and S wave Francis P. W., and C. C. Rundle, Rates of
velocity structures of southern Peru and production of the main magma types in
their tectonic implications, J. Geophys. the Central Andes, Geol. Soc. Am. Bull.,
Res., 91, 9517-9532, 1986. 87, 474-480, 1976.
926 S•brier et al.' Stress in the Andes of Central Peru

Froidevaux, C., and B. Isacks, The James, D. E., Plate tectonic model for the
mechanical state of the Altiplano-Puna evolution of Central Andes, Geol. Soc.
segment of the Andes, Earth Planet. Sci. Am. Bull., 82, 3325-3346, 1971.
Lett., 71, 305-314, 1984. Kaneoka, I., and C. Guevara, K-Ar age
Gephart, J. W., and D. W. Forsyth, An determinations of late Tertiary and
improved •ethod for determining the Quaternary Andean volcanic rocks,
regional stress tensor using earthquake Southern Peru, Geochem. J., 18, 233-239,
focal mechanism data' Application to the 1984.
San Fernando earthquake sequence, J. Knight, R. J., N. Mortimer, D. Wilson, A.
Geophys. Res., 89, 9305-9320, 1984. Nur, and M. Villafuerte, Paleomagnetic
Grange, F., P. Cunningham, J. Cagnepain, studies of the Arequipa massif, Peru, in
D. Hatzfeld P. Molnar, L. Ocola, A. Proceedings of the Circumpacific Terrane
Rodriguez, S. W. Stock, and G. Suarez, Conference, Stanford Univ., Calif.,
The configuration of the seismic zone 1983.
and the downgoing slab in southern Peru, Koch, E., Die Tektonik im Subandin des
Geophys. Res. Lett., 11(1), 38-41, Mittel-Ucayali-Gebietes, Ost-Peru,
1984a. Geotekton. Forsch., 15, Stuttgart, 67
Grange, F., D. Hatzfeld, P. Cunningham, P. pp., 1962.
Molnar, S. W. Roecker, G. Suarez, A. Kono, M., K. Heki, and A. Yamamoto,
Rodriguez, and L. Ocola, Tectonic Mountain building in the Central Andes,
implications oF the microearthquake Andes Sciences, •, 125-136.
seismicity and fault plane solutions in Laharie, R., Cronologia del Cuaternario
southern Peru, J. Geophys. Res., 89, Peruano, in Proceedings I Congreso Latin
6139-6152, 1984b. American Geology, vol. 6, pp. 145-157,
Handschumacher, D. W., Post Eocene plate Sociedad Geologica Peru, Lima, 1970.
tectonics of the Eastern Pacific, in The Lavenu, A., M. S•brier, and M. Servant,
Geophysics of the Pacific Ocean Basin N•otectonique des Andes Centrales'
and its Margin, Geophys. Monogr. Ser. P•rou, Bolivie, Bull., •, 56-58, INQUA
vol. 19, edited by G. H. Sutton et al., Neotectonic Comm., Stockholm, 1980.
pp. 177-202, AGU, Washington, D.C., Machar•, J. Informe sobre las zonas
1976.
afectadas por los sismos de Ayacucho,
Hasegawa, A., and I. S. Sacks, Subduction Noviembre 1980, report, 5 pp., Inst.
of the Nazca plate beneath Peru as Geophys. Lima, Peru, 1980.
determined from seismic observations, J. Machar•, J., Geologia del Cuaternario en
Geophys. Res., 86, 4971-4980, 1981. la costa del Peru central, thesis, 197
Heim, A., Observaciones geologicas en la pp., Univ. Nac. Ing., Lima, 1981.
region del terremoto de Ancash de Machar• J., M. S•brier, D. Huaman, and J.
Noviembre de 1946, Soc. Geol. Peru V. Mercier, Tectonica cenozoica de la
Jub., •(6), 28 pp., 1949. margen continental peruana, Bol. Soc.
Helwig, J., Plate tectonic model for the Geol. Peru, 76, 45-77, 1986.
evolution of Central Andes' Discussion, Mammerickx J., R. Anderson, H. Menard, and
Geol. Soc. Am. Bull., 84, 1493-1496, S. Smith, Morphology and tectonic
1973. evolution of the East Central Pacific,
Hodgson, J., and P. Bremner, Direction of Geol. Soc. Am. Bull., 86, 111-118, 1975.
faulting in Ancash, Peru, earthquake of Mammerickx J., E. Herron, and L. Dorman,
Nov. 1946 from teleseismic evidence, Evidence for two fossil spreading ridges
Bull. Seismol. Soc. Am., 43, 121-125, in the southeast Pacific, Geol. Soc. Am.
1953.
Bull., 91, 263-271, 1980.
Howell, D., Des terrains d•plac•s' Les Martinez, C., Structure et •volution de la
terranes, Pour Sci., 99, 18-28, 1986. chaine Hercynienne et de la chaine
Huaman, D., Evolution tectonique Andine dans le nord de la cordill&re des
c•nozoique et n•otectonique du piemont Andes de Bolivie, Tray. Doc. ORSTOM,
pacifique dans la r•gion dWArequipa 119, 352 pp., 1980.
(Andes du sud Perou), thesis, 220 pp., May, S. R., and R. F. Butler,
Univ. Paris-Sud, Orsay, 1985. Paleomagnetism of the Puente Piedra
Hussong, D. M., and L. K. Wipperman, formation, central Peru, Earth Planet.
Vertical movement and tectonic erosion
Sci. Lett., 72(2/3), 205-218, 1985.
of the continental wall of the Peru-
M•gard, F., Geologia del cuadrangulo de
Chile trench near 11ø30wS, Mem. Geol. Huancayo, Bol. Serv. Geol. Miner., 18,
Soc. Am., 154, 509-524, 1981. 123 pp., 1968.
S•brier et al.' Stress in the Andes of Central Peru 927

M•gard, F., Etude g•ologique d•une Pardo, A., Caracteristicas estructurales


transversale des Andes au niveau du de la faja Subandina del Norte del Peru,
P•rou central, thesis, 263 pp., Univ. paper presented at the Symposium on
Sci. Tech. Languedoc, Montpellier, 1973. Exploracion Petrolera en las Cuencas
M•gard, F., Etude g•ologique des Andes du Subandinas de Venezuela, Colombia,
P•rou central, Mem. ORSTOM, 86, 310 pp., Ecuador, y Peru, Bogota, 1982.
Paris, 1978. Pardo, A., and F. Zuniga, Estratigrafia y
M•gard, F., and H. Philip, Plio-Quaternary evolucion tectonica del Mesozoico y
tectono-magmatic zonation and plate Cenozoico de la region de la Selva del
tectonics in the Central Andes, Earth Peru, in Proceedings II Congreso Latin
Planet. Sci. Lett., 33, 231-238, 1976. American Geology-Caracas, pp. 588-608,
M•gard F., D.C. Noble, E. McKee, and H. 1973.
Bellon, Multiple pulses of Neogene Pennington, W. D., Subduction of the
deformation in the Ayacucho intermontane eastern Panama Basin and seismotectonics
basin, Andes of central Peru, Geol. Soc. of the northwestern South America, J.
Am. Bull., 95(9), 1108-1117, 1984. Geophys. Res., 86, 10753-10770, 1981.
Mercier, J. L., Extensional-compressional Philip, H., and F. M•gard, Structural
tectonics associated with the Aegean analysis of the superficial deformation
Arc' Comparison with the Andean of the 1969 Pariahuanca earthquakes
Cordillera of south Peru-north Bolivia, (central Peru), Tectonophysicso, 38,
Philos. Trans. R. Soc. London, Ser. A, 259- 278, 1977.
300, 337-355, 1981. Pilger, R. H., Plate reconstructions,
Mercier, J. L., and M. S•brier, Informe aseismic ridges, and low-angle
sobre el estudio preliminar de la fallas subduction beneath the Andes, Geol. Soc.
recientes de la region de la represa de Am. Bull., 92, 448-456, 1981.
Recreta (Cordillera Bianca, Dep. Richter, C. F., Elementary Seismology, W.
Ancash), report, 8 pp., Electroperu, H. Freeman, New York, 1958
Lima, 1981. Ruegg, W., Le tremblement de terre
Mercier, J. L., R. Armijo, P. Tapponnier, d'Ancash (P•rou) du 10 Novembre 1946 et
E. Carey-Gailhardis, and Han T. L., ses causes g•otectoniques, Tray.
Change from Late Tertiary compression to Institut Fran9ais Etudes Andines, •,
Quaternary extension in southern Tibet 153-166, 1950.
during the India-Asia collision, Rutland, R. W., Andean orogeny and ocean
Tectonics, •(3), 275-304, 1987a. floor spreading, Nature, 233, 252-255,
Mercier, J. L., D. Sorel, and K. Simeakis, 1971.
Changes in state of stress in the Sacks, I. S., The subduction of young
overriding plate of a subduction zone' lithosphere, J. Geophys. Res., 88,
The Agean Arc from the Pliocene to the 3355-3366, 1983.
Present, Ann. Tectonicae, •,(1), 20-39, Scotese, C. R., and D. B. Rowley, The
1987b. orthogonality of subduction' An
Minster, J. B., and T. H. Jordan, empirical rule?, Tectonophysics, 116,
Present-day plate motions, J. Geophys. 173-187, 1985.
Res., 83, 5331-5334, 1978. S•brier, M., La tectonica reciente de la
Molnar, P. and T. Atwater, Interarc zona de Bayovar, Rep. 78-1, Inst.
spreading and cordilleran tectonics as Geofis., Lima, Peru, 1978.
alternates related to the age of the S•brier, M., Informe general de
subducted oceanic lithospere, Earth actividades y resultados cientificos
Planet. Sci. Lett., 41, 330-340, 1978. para los amos 1977-1978, report, 17 pp.,
Moretti, I., Subduction des rides Inst. Geofis., Lima, Peru, 1979.
as•ismiques, thesis, 107 pp., Univ. S•brier, M., and J. Cabrera, Oservaciones
Paris-Sud, Orsay, 1982. neotectonicas en la zona epicentral de
Mortimer, C., The evolution of the los sismos de Ayacucho, report, 21 pp.,
continental margin of northern Chile, in Inst. Geofis., Lima, Peru, 1981
Proceedings 24th International S•brier, M., and J. Machar•, Observaciones
Geological Congress, sect. 8, pp. 48-52, acerca del Cuaternario de la Costa del
Montreal, 1972. Peru central, Bull. Inst. Fr. Etud.
Nur, A., and A. Ben-Avraham, Oceanic Andines, •(1-2), 5-22, 1980.
plateaus, the fragmentation of S•brier, M., R. Marocco, J. J. Gross, S.
continents and mountain building, J__=. Macedo, and M. Montoya, Evolucion
Geophys. Res., 87, 3644-3661, 1982. neogena del Piedemonte Pacifico de los
928 S•brier et al.' Stress in the Andes of Central Peru

Andes del Sur del Peru, in Proceeding II plate, J. Geophys. Res., 78, 5033- 5061,
Congreso Geology Chile, vol.3, pp.71-88, 1973.
Instituto Investigaciones Geologicas, Stauder, W., Subduction the Nazca plate
Santiago, Chile, 1979. under Peru as evidence by focal
S•brier, M., A. Lavenu, and M. Servant, mechanisms and by seismicity, J.
Apuntes recientes sobre la neotectonica Geophys. Res., 80, 1053-1064, 1975.
en los Andes Centrales (Peru-Bolivia), Suarez, G., Seismicity, tectonics, and
Bull. Inst. Fr. Etudo Andines, •(1-2), surface wave propagation in the central
1-3, 1980. Andes, Ph.D. thesis, 259 pp., Mass.
S•brier, M., D. Huaman, J. L. Blanc, J. Inst. of Technol., 1982.
Machar• , D. Bonnot, and J. Cabrera, Suarez, G., P. Molnar., and B.C.
Observaciones acerca de la Neotectonica Burchfiel, Seismicity, fault plane
del Peru, 109 pp., report, Inst. Geofis. solutions, depth of faulting, and active
Peru, Seismicity Seismic Risk Andean tectonics of the Andes of Peru, Ecuador,
Reg., Lima, 1982. and southern Colombia, J. Geophys. Res.,
S•brier, M., J. L. Mercier, F. M•gard, G. 88, 10403-10428, 1983.
Laubacher, and E. Carey-Gailhardis, Tapponnier, P., and P. Molnar, Slip-line
Quaternary normal and reverse faulting field theory and large-scale continental
and the state of stress in the central tectonics, Nature, 264, 319-324, 1976.
Andes of South Peru, Tectonics, 4, Uyeda, S., and H. Kanamori, Back arc
739-780, 1985. opening and the mode of subduction, J.
S•brier M., A. Lavenu, M. Fornari, and J. Geophys. Res., 84, 1049-1061, 1979.
P. Soulas, Tectonic and uplift in Vlaar, N.J., and M. J, Wortel,
Central Andes (Peru, Bolivia, northern Lithospheric aging, instability and
Chili) from Eocene to Present, subduction, Tectonophysics, 32, 331-351,
G•odynamique,3(2), in press, 1988. 1976.
Shackleton, R. M., A. C. Ries, M.P. Von Huene, R., L. D. Kulm, and J. Miller,
Coward, and P. R. Cobbold, Structure, Structure of the frontal part of the
metamorphism and geochronology of the Andean convergent margin, J. Geophys.
Arequipa massif of coastal Peru, J. Res., 90, 5429-5442, 1985.
Geol. Soc. London, 136, 195-214, 1979. Wilson, J. J., L. Reyes, and J. Garrayar,
Silgado, E., The Ancash, Peru earthquake Geologia de los cuadrangulos de
of November 10, 1946, Bull. Seismol. Mollebamba, Tayabamba, Huaylas,
Soc. Am., 41, 83-100, 1951. Pomabamba, Carhuaz y Huari, Bol. Ser.
Silgado, E., Historia de los sismos mas Geol. Miner., 16, 1967.
notables ocurridos en el Peru (1513- Yonekura, N., T. Matsuda, M. Nogami, and
1974), Bol. 3, Ser. C, 130 pp., Inst. S. Kaizuka, An active fault along the
Geol. Miner., Lima, 1978. western foot of the Cordillera Bianca,
Snoke, Jo A., I. S. Sacks, and D. James, Peru, Chigaku Zasshi, 88(1), 1-19, 1979.
Subductions beneath Western South
America' Evidence from converted phases,
Geophys. J. R. Astron. Soc., 59, 219- J. L. Blanc, D. Bonnot, J. Cabrera, J.
225, 1979. Machar•, J. L. Mercier, and M. S•brier,
Soulas, J.P., Las fases tectonicas Laboratoire de G•ologie Dynamique Interne,
jovenes de los Andes centrales del Peru, Unit• Associ•e 730 du Centre National de
Bol. Soc. Geol. Peru, 50, 77-86, 1975. la Recherche Scientifique, BRtiment 509,
Soulas, J.P., Tectonique quaternaire' La Universit• de Paris-Sud, F-91405 ORSAY
c6te pacifique et la chaine andine du Cedex, France.
P•rou central, Rev. G•ogr. Phys. G•ol.
Dyn., 20(5), 399-414, 1978.
Stauder, W., Mechanism and spatial (Received October 2, 1986
distribution of Chilean earthquakes with revised March 1, 1988;
relation to subduction of the oceanic accepted March 2, 1988.)

You might also like