Mapping the continuity of the Nazca Plate through its aseismic part in the Arica Elbow(Central Andes)

Similar documents
2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Subduction geometry in central Peru from a microseismicity survey: first results

Seismic Activity near the Sunda and Andaman Trenches in the Sumatra Subduction Zone

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

Slip distributions of the 1944 Tonankai and 1946 Nankai earthquakes including the horizontal movement effect on tsunami generation

The geometry of the Wadati Benioff zone and lithospheric kinematics in the Hellenic arc

Subduction zones are complex plate boundaries in which variable geometry and structure can be

Earthquake patterns in the Flinders Ranges - Temporary network , preliminary results

MONITORING OF SEISMIC EVENTS IN THE CENTRAL ANDEAN REGION. Lawrence A. Drake, Rodolfo Ayala and S. Cristina Condori. Observatorio San Calixto

Magnitude 7.0 PERU. This region of the Andes is a sparsely populated area, there were no immediate reports of injuries or damage.

DOUBLE-PLANED STRUCTURE OF INTERMEDIATE- DEPTH SEISMIC ZONE AND THERMAL STRESS IN THE DESCENDING PLATE. (Received December 20, 1983)

ANOTHER MEXICAN EARTHQUAKE! Magnitude 7.1, Tuesday Sept. 19, 2017

9th Workshop on Three-Dimensional Modelling of Seismic Waves Generation, Propagation and their Inversion

Tomographic imaging of P wave velocity structure beneath the region around Beijing

Data Repository Item

FOCAL MECHANISMS OF SUBDUCTION ZONE EARTHQUAKES ALONG THE JAVA TRENCH: PRELIMINARY STUDY FOR THE PSHA FOR YOGYAKARTA REGION, INDONESIA

Plate Tectonics 3. Where Does All the Extra Crust Go?

Characteristics of seismic activity before Chile M W 8.8 earthquake in 2010

Estimation of S-wave scattering coefficient in the mantle from envelope characteristics before and after the ScS arrival

Magnitude 7.1 PERU. There are early reports of homes and roads collapsed leaving one dead and several dozen injured.

Seismic structure in southern Peru: Evidence for a smooth contortion between flat and normal subduction of the Nazca plate

Magnitude 7.6 & 7.6 PERU

SURFACE WAVE GROUP VELOCITY MEASUREMENTS ACROSS EURASIA

Late 20 th Century Tests of the Continental Drift Hypothesis

Seismic Velocity Structure in the Crust and Upper Mantle beneath Northern Japan

Time dependence of PKP(BC) PKP(DF) times: could this be an artifact of systematic earthquake mislocations?

4 Deforming the Earth s Crust

crustal thicknesses. earthquakes as in recorded

Improving Global Seismic Event Locations Using Source-Receiver Reciprocity

C3.4.1 Vertical (radial) variations in mantle structure

Seismic ray path variations in a 3D global velocity model

Tracing rays through the Earth

High resolution receiver function imaging of the seismic velocity discontinuities in the crust and the uppermost mantle beneath southwest Japan

CHAPTER 2 THE WAY THE EARTH WORKS: EXAMINING PLATE TECTONICS

USU 1360 TECTONICS / PROCESSES

SUPPLEMENTARY INFORMATION

General Oceanography Geology 105 Expedition 10 - Whole Lotta Shakin' Goin' On

Search for deep slabs in the Northwest Pacific mantle (subduction/lithosphere/residual spheres/lower mantle)

Paul S. Cunningham 1 and Steven W. Roecker 2

Crust and upper mantle P- and S-wave delay times at Eurasian seismic stations

The Earthquake of Padang, Sumatra of 30 September 2009 scientific information and update

Plate Tectonics. Structure of the Earth

Continent-sized anomalous zones with low seismic velocity at the base of Earth s mantle

Inversion of travel times to estimate Moho depth in Shillong Plateau and Kinematic implications through stress analysis of Northeastern India

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

Geophysical Journal International

4-D Geodynamic Modeling With Data Assimilation: Subduction and Continental Evolution

Plate Tectonics. entirely rock both and rock

Coseismic slip distribution of the 1946 Nankai earthquake and aseismic slips caused by the earthquake

Lithospheric structure of the central Andes based on surface wave dispersion

Topic 5: The Dynamic Crust (workbook p ) Evidence that Earth s crust has shifted and changed in both the past and the present is shown by:

EARTHQUAKE LOCATION ACCURACY IN THE ARABIAN-EURASIAN COLLISION ZONE. Sponsored by Air Force Research Laboratory 1

Global Tectonics. Kearey, Philip. Table of Contents ISBN-13: Historical perspective. 2. The interior of the Earth.

Fault Length and Direction of Rupture Propagation for the 1993 Kushiro-Oki Earthquake as Derived from Strong Motion Duration

A) B) C) D) 4. Which diagram below best represents the pattern of magnetic orientation in the seafloor on the west (left) side of the ocean ridge?

Received May 11, 1989; Accepted July 10, * To whom correspondence should be addresseḍ

Geology 101 Study Guide #4

D DAVID PUBLISHING. Deformation of Mild Steel Plate with Linear Cracks due to Horizontal Compression. 1. Introduction

Mantle layering across central South America

of other regional earthquakes (e.g. Zoback and Zoback, 1980). I also want to find out

Earth Science Lesson 3 Dynamic Earth Processes (Grades 8-12 ) Instruction 3-3 Plate Boundaries

Dynamic Subsidence and Uplift of the Colorado Plateau. Supplementary Material

Earthquake Doublet Sequences: Evidence of Static Triggering in the Strong Convergent Zones of Taiwan

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

Magnitude 7.1 NEAR THE EAST COAST OF HONSHU, JAPAN

Supplementary Material

Crustal Activity. Plate Tectonics - Plates - Lithosphere - Asthenosphere - Earth s surface consists of a major plates and some minor ones

12. The diagram below shows the collision of an oceanic plate and a continental plate.

Magnitude 7.1 SOUTH SANDWICH ISLANDS

Magnitude 7.5 PALU, INDONESIA

Lab 1: Plate Tectonics April 2, 2009

28th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Seismology and Structure of the New Hebrides

Tristan volcano complex: oceanic end-point of a major African lineament.

09/06/2017. The orogenic cycle in the Andes: Arc magmatism and delamination as control of the fold and thrust belts

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges

Magnitude 8.2 NORTHWEST OF IQUIQUE, CHILE

Seismic Anisotropy and Mantle Flow in the Izu-Bonin-Mariana Subduction System

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Geo736: Seismicity along mid-ocean ridges

Shear-wave anisotropy beneath the Ryukyu arc

Magnitude 7.9 SE of KODIAK, ALASKA

APPLICATION OF A PASSIVE TOMOGRAPHY METHOD AND CORRELATION WITH ACTIVE SEISMIC OBSERVATIONS IN THE KYPARISSIAKOS GULF, SOUTHWESTERN HELLENIC ARC

3. PLATE TECTONICS LAST NAME (ALL IN CAPS): FIRST NAME: PLATES

Auxiliary Material. Subduction of oceanic asthenosphere: evidence from sub-slab seismic anisotropy. Teh-Ru Alex Song, Hitoshi Kawakatsu

GEOLOGY MEDIA SUITE Chapter 13

Relocation of a global earthquake data set with a 3-D velocity model

Magnitude 8.3 SEA OF OKHOTSK

Supporting Information for An automatically updated S-wave model of the upper mantle and the depth extent of azimuthal anisotropy

A magnitude 7.8 earthquake has occurred km (63.3 mi) ESE of Suva, Fiji at a depth of km (378 miles).

Dynamic Crust Practice

North America subducted under Rubia. Are there modern analogs for Hildebrand s model of North America subducting under Rubia?

GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L19604, doi: /2004gl020366, 2004

MAR110 Lecture #5 Plate Tectonics-Earthquakes

The Earth s Structure from Travel Times

2. Explain why there are these two predominate elevations. (Hint: think about isostasy and the type of materials we used to demonstrate buoyancy).

29th Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

M. Koch and T.H. Münch. Department of Geohydraulics and Engineering Hydrology University of Kassel Kurt-Wolters-Strasse 3 D Kassel

GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L14308, doi: /2008gl034461, 2008

Beneath our Feet: The 4 Layers of the Earty by Kelly Hashway

Transcription:

~r'.!. ]- ELSEVIER Physics of the Earth and Planetary Interiors 101 (1997) 163-173 PHYSICS OFTHE EARTH ANDPLANETARY INTERIORS Mapping the continuity of the Nazca Plate through its aseismic part in the Arica Elbow(Central Andes) Catherine Dorbath ORSTOM and EOPG, URA 1358 CNRS, Paris, France Received 13 March 1996; accepted 14 June 1996 Abstract Along the Peru-Chile trench, the subduction of the Nazca Plate under South America is well underlined by the Wadati Benioff zone. In the Central Andes, the seismicity defines a moderately deepening slab down to about 325 km. The occurrence of a scarce seismicity in the 550-650 km depth range, after a complete seismic quiescence, raises the question of the continuity of the slab. To answer this question we performed a 3-D teleseismic tomography of the mantle at 20 S. On the tomographic image, a higher velocity zone is associated with the subducted plate where it is defined by the Wadati Benioff zone. At greater depth, this higher velocity body can be followed through the model down to the lower mantle. Our results provide clear evidence of the continuity of the slab throughout its seismically quiescent part. Moreover, they allow the imaging of the bending of the Nazca Plate in the Arica Elbow, where the general strike of the Andean chain changes and no deep seismicity has never been recorded. Finally, our results confirm the increase of the dip of the slab between 350 and 550 km in depth. 1. Introduction The western part of South America is one of the major plate boundaries on the Earth. It is the largest region in which ocean-continent convergence is happening today. Moreover, it is the only region where an oceanic plate is subsiding under a major lithospheric plate and not under small continental blocks. The Andes are the result of the subduction of the Nazca Plate beneath the South American Plate along the Peru-Chile trench since the Cretaceous. This continental margin orogenic belt, continuous for more than 7000 km, presents distinct broad-scale tectonic segments; their coincidence with the varia- tions in the geometry of the subducted Nazca Plate is very close (Barazangi and Isacks, 1976; Jordan et al., 1983). The most noticeable of these segments is certainly the Central Andes (15 S to 27 S): the range is at its widest part (Bolivian Orocline) including the Altiplano-Puna high plateau and the general strike of the structures changes from N320 W to nearly NS (Arica Elbow) (Fig. 1). The location of the Wadati Benioff zone provides the most common way used to map subducting lithospheres. Numerous studies concerning the seismicity and shape of the Nazca Plate have been carried out, using a compilation of data from international centres (Sykes and Hayes, 1971; Stauder, 1973, 0031-9201/97/$17.00 1997 Elsevier Science B.V. All rights reserved. PII S0031-9201(96)03206-2

164 C. Dorbath / Physics ~/'the Earth and Planetary Interiors I01 (1997) 163-173 10" S 15" S 20"s 25"s 30" S 75" W 70" W 65" W Fig. 1. Location map. The triangles represent the seismic stations of the 1994 Lithoscope experiment. The thick solid grey line off shore represents the Peru-Chile trench, and the dashed grey line inland the eastern limit of the Andean front of deformation. The pale grey zone represents the Altiplano, and the dark grey one Titicaca lake. The shape of the Nazca Plate is shown via contours to the depth of the centre of the Wadati Benioff zone from Cahill and Isacks (1992). 1975; Barazangi and Isacks, 1976, 1979; Bevis and Isacks, 1984). In the Central Andes, they describe the slab as continuously dipping to the east at a moderate angle of about 30, and neighbouring segments (north of 15 S and south of 27 S) become nearly horizontal as they reach a depth of about 10 km. The updated compilation of Cahill and Isacks (1992) (Fig. 1) suggests that, in fact, this segment of the slab does not have a constant dip along its entire extent, but gradually flattens south of 20 S. Complementary data from local temporary networks operated in the same region support and refine the results obtained from teleseismic studies (Boyd et al., 1984; Grange et al., 1984; Comte and Suarez, 1995; Delouis et al., 1995). Whatever data, teleseismic or local, are used to map the geometry of the upper part of the subducted lithosphere, the downgoing plate is seismically defined in the Central Andes from the trench down to 300-350 km. From this depth down to 500-550 km, a seismic quiescence is observed. Then few very deep earthquakes occur in clusters down to 600-650 km. Similar observations are made for the sub-horizontal subduction beneath central Peru, where the Wadati Benioff zone extends down to 100-125 km, then a seismic quiescence is observed down to 500 km, followed by a deep seismic zone. The lack of any events in the depth range 325-520 km is confirmed by the recent homogeneous and comprehensive catalogue produced by Kirby et al. (1995) who relocated 213 historical and modem deep events. An important question arises: is the slab continuous or not through the aseismic zone? To answer this question we need to use techniques not based on mapping the local seismicity. In this way, Isacks and Barazangi (1973), Snoke et al. (1974) and Barazangi and Isacks (1976) studied the transmission through the region of anomalously high-frequency waves, but their results did not lead to any clear conclusion about slab continuity. Later on, models have been proposed suggesting either the absence of a slab in the aseismic zone (Wortel, 1982, 1984) or the continuity of the subducted plate (Engebretson and Kirby, 1992). Recently, James and Snoke (1990) presented seismic evidence for the continuity of the Nazca Plate beneath central Peru (10-14 S), in a segment where subduction becomes nearly horizontal. This result is based on the analysis of an anomalous P-wave arrival in the P codas of Peru-Brazil deep focus earthquakes recorded on the Carnegie broadband station at Cuzco. The authors concluded that the descending plate exists and is probably continuous, although aseismic, between 150 and 500 km depth, and is very steeply dipping (70 ) below a depth of about 150 or 200 km. Finally, Engdahl et al. (1995) imaged the subducted lithosphere beneath South America by inversion of travel-time residuals of teleseismically well-recorded events that occurred in the region. Their study provides evidence for high-velocity material at depth and therefore for a probable continuity of the slab over regions with seismic quiescence. It also suggests that the slab penetrates the lower mantle. In a recent paper (Dorbath et ai., 1996), we provided corroborative evidence of the existence and continuity of the subducted Nazca Plate through the

C. Dorbath / Physics of the Earth and Planetary Interiors 101 (1997) 163-173 165 aseismic zone beneath the Central Andes along a vertical cross-section at 20 S. This complementary paper will focus on the shape of the aseismic part of the slab below the Bolivian Orocline from 14 to 23 S, in the Arica Elbow region. These results are based on a 3-D tomography of the lithosphere down to a 660 km depth realized by the inversion of travel-time residuals of teleseismic P waves recorded during the 1994 Lithoscope field experiment. Such a method for mapping subducted lithosphere has never been used for the Nazca Plate. However, it has been used for other subduction zones, such as the western Pacific (Hirahara, 1981) or eastern Pacific (Harris et al., 1991). Everywhere, slabs are evidenced by inclined high-velocity zones which well correlate with the Wadati Benioff zones (Hirahara and Hasemi, 1993). For western Pacific regions, which offer a large number of shallow, intermediate and deep focus earthquakes, local earthquake tomography offers a powerful complementary method to teleseismic tomography. Clearly, this assistance is not possible in our study area, as we intend to investigate an aseismic zone and very deep seismicity is scarce. On the other hand, the recent paper by Engdahl et al. 2sl 0s f -2 s -4 s 100 km Azixnuth: 31 ~h: 0.0-10.0 Distance: 2 -.e-: 30.0-40.0 0s s -2 s -4s s s Azitnuth: 315.0-325. Distance: 35.0-50.q Azimuth: 30,0-40.0 Distance: 30.0-40.0-4 s Azinluth: 27.5.0-7.45.0 Distance: 30.0-45.0 Fig. 2. Mean P residuals, relative to the same reference station located in the Altiplano, for six different epicentral regions. Dashed line, observed residuals; grey line, residuals computed using the final 3-D perturbation model obtained by the tomography. The map in the centre shows the distribution, relative to the centre of the Lithoscope array (black cross), of the events selected for the tomography at epicentral distances at less than 100.

166 c. Dorbath / Physics oj the Earth and Planetary Interiors 101 (1997) 163-173 (1995) will provide a interesting basis lor comparison: their large-scale tomographic image of the subducted Nazca Plate is obtained by inverting traveltime residuals of teleseismically recorded South American earthquakes and residuals at South American stations observed from events from a larger study region which includes the South Atlantic and eastern Pacific as well. Our tomographic image is obtained by inverting travel-time residuals of teleseismic events recorded on a local network; its scale, smaller in the upper layers, is of the same order of magnitude as Engdahl's study in the deeper layers. 2. Data From June to November 1994, 41 vertical shortperiod stations from the French Lithoscope network were operated along a 700-km-long profile crossing the Andean chain in its entirety. Therefore, this profile was placed above the subducted plate from the coast, where the upper part of the plate is at a 50 km depth, to its eastern limit as it is defined by the Wadati Benioff zone; the profile ended at the sub- Andean front of deformation, above the very deep seismicity. The profile was designed to be as perpendicular as possible to the structures. The spacing between the stations was about 20 krn (Fig. 1). Among the 250 teleseismic events recorded during the 6-month experiment, we selected 120 events which provided clear P-wave arrivals at more than ten stations. They were evenly distributed in distance and azimuth relative to the seismic network. Absolute travel times were calculated using hypocentral data from the USGS's Preliminary Determination of Epicenter's bulletin and Herrin's tables (Herrin, 1968). Next we calculated residuals at every station lor every selected event, and then we computed the value of these residuals relative to the same reference station situated in the Altiplano. A study of the variation along the profile of these relative residuals was presented in a preliminary paper (Dorbath et al., 1996). However, we show again in Fig. 2 (dotted line) some P-wave residual profiles which are within the scope of this study. To avoid errors due to poor readings or source peculiarity, we plotted profiles of mean relative residuals for some source regions, and not for singular events. Unfortunately, there is a lack of active seismic zones in the eastern and western azimuths, it is aligned with our linear network, which could provide direct and reverse profiles of residuals, and consequently help us to make simple qualitative modelling. Fig. 2 shows the strong azimuthal variations of P relative residuals and particularly the difference between the north-eastern quadrant and the other azimuths. For the azimuths between 0 and 40 E, the more we go west, the more the relative residuals are negative, therefore more rays cross high velocity structures. The decrease from east to west along the profile is more gradual for northern than for north-eastern azimuths, but on both profiles the Table 1 Initial P-velocity model Layer Velocity (km s- 1 ) Thickness (km) Nb of blocks N-S Dimension (km) Nb of blocks E-W Dimension (km) 1 6.00 10 1 Cones 42 Cones 2 6.30 20 6 45 26 30 3 6.80 30 6 45 26 30 4 8.00 40 6 60 20 40 5 8.10 40 6 60 22 40 6 8.20 50 8 75 18 50 7 8.35 50 10 75 20 50 8 8.50 60 10 90 18 60 9 8.80 80 10 120 16 80 10 9.20 80 12 120 16 80 11 9.75 100 14 150 14 100 12 10.30 100 16 150 16 100

C. Dorbath / Physics of the Earth and Planetary Interiors 101 (1997) 163-173 167 Layer 6 : 140-190 km Layer 10 : 380-460 km 2OO 0-200 -4OO -200 0 200 400 Layer 7 : 190-240 km -400-200 0 200 409 ]uu Layer 11 : 460. 560 km -~( -400-200 0 200 400 lun ~ 3 I ~ ~ IOOkm -60Q -400-7,00 0 200 400 600 Inn Layer 12 : 560-660 km Layer 9 : 300-380 km -400-200 0 200 400 Ion O 0.2 0.4 0.6 0.8-6oo -400.zoo o 200 400 mekl Fig. 3. Value of the diagonal element of the resolution matrix corresponding to each block in the seven deeper layers of the tomography. The darkest blocks are the best resolved in the inversion. The numbers indicate how many rays crossing each block were used in the inversion.

168 C. Dorbath / Physics ~[' the Earth and Planetary Interiors I01 (1997) 163-173 (a) Layer 6 : 140-190 km Layer 10 : 380-460 km -200-~. I ' I ' I ' I ' I ' I ' I ' -400 -~0 0 200 400 lan Layer 7 : 190-240 km 20+ I 40o -20o o 200 4o0 km Layer 11 : 460. 560 km....,.,m, 400-200 0 200 400 Ion Layer 8 : 240. 300 km 200 0-200..400-200 0 200 400 Ion,600-400 -200 0 200 400 60014all Layer 12 : 560-660 km Layer 9 : 300-380 km 600 200 o.200 400 200 O -400-200 0 300 40e Ion -200-4O0.-600 4.5-3.5-2.5-1.5-0.5 0.5 1.5 2.5 3.5 4.5 % -600-400 -200 0 200 400 600 k~ Fig. 4. Velocity perturbations (in %) in the layers below 140 km. The regions with higher velocities than the reference model (positive perturbations) are filled with dotted grey patterns. (a) Original block model resulting from the inversion of teleseismic travel-time resiuals, on which iso-perturbation contours have been superimposed. (b) Smoothed perturbation model, together with the seismicity associated with the slab as reported by the NEIS for the past years 20 (black dots). The June 9, 1994 Bolivian earthquake is shown by a larger white dot on Layer 12.

C. Dorbath / Physics of the Earth and Planetary Interiors 101 (1997) 163-173 169 (b) Layer 6 : 140-190 km Layer 10 : 380-460 km -400-7~1 0 200 400 km Layer 7 : 190-240 Ion 400-7~0 0 700 400 km Layer 11 : 460 - [;60 km 400 -:ZOO 0 200 400 Inn Layer 8 : 240. 300 km -400-200 0 200 400 Ion -6~ -400-200 0 200 400 600 km Layer 12 : 560-660 km Layer 9 : 300-380 lun.400-200 0 200 400 inn -4.5-325-2.T-1.5-0.5 0.5 1.5 2,.5 3,.5 4.5 % Fig. 4 (continued). -400-200 0 200 400 600 km variation reaches 4 s, twice that observed for other azimuths. Such large-scale differences emphasize at the same time that the anomalous structures are deep and that the region sampled by the rays is far from the axial symmetry. This assertion has to be sustained by the tomography.

170 C. Dorbath / Physics of the Earth and Planeta O' Interiors 101 (1997) 163-173 3. Inversion The travel-time residuals are inverted using a version modified by Evans and Achauer (1993) of the 'ACH' inversion method (Aki et al., 1977), extensively described in published literature. The parameterization of the initial velocity model is presented in Table 1 and in Fig. 3. In order to obtain a satisfactory explanation of our data, there is a large reduction in the data variance, we were led to build a 3-D thick initial model. In effect, the reduction of the variance is less than 50% for a 2-D 420-km-thick model; it goes up to 73% for a 3-D model of the same thickness, and reaches 82% when we use a 3-D 660-km-thick model. Therefore, our initial model includes the complete upper mantle and the transition region down to the 650-km discontinuity. In this paper we will focus on the final velocity perturbations in the mantle below 140 km. Above this depth, the slab, evidenced by the Wadati Benioff zone, is situated at the western extremity of the model, where the resolution is the lowest. On the other hand, beneath 140 km, the slab is well inside the model and resolved. The value of the diagonal element of the resolution matrix for the seven deeper layers is presented in Fig. 3 together with the number of rays in the blocks. We only kept the blocks crossed by more than ten rays, Whatever the depth, the central part of the layers, crossed by numerous rays with varying paths, is the best resolved. A complementary method with which to check the validity of the final velocity model is to examine how far it explains the observed data. We present in Fig. 2 (grey solid line) the variations along the profile of P-wave residuals computed with the final model for the selected hypocentral regions. We notice a very close correspondence between the observed and computed residuals. The only marked difference is seen at the western extremity of the north-eastern profile, where the divergence reaches 0.4 s. The strong azimuthal variations observed in our data set are satisfactorily explained by the velocity perturbation model obtained by the inversion. We present in Fig. 4 two images of the velocity perturbations obtained for the seven deepest layers of our model: the original velocity perturbations in blocks in Fig. 4(a), and the smoothed image of the perturbations in Fig. 4(b). The origin of the coordi- nates is the centre of our parameterized model, that is 19.5 S, 66.5 W. The seismicity reported by the National Earthquake Information Service for the time period 1974-1994 in the respective depth intervals is superimposed on each layer in Fig. 4(b). The largest deep earthquake (June 9, 1994, Mw = 8.3, depth = 630 kin) instrumentally recorded in Bolivia has been particularized. Down to 325 km, the Wadati Benioff zone defines the position of the slab in the layers, and is more or less clearly associated with a higher velocity zone. From 140 to 240 km (Layers 6 and 7), the relation between the seismicity and the velocity is not straightforward. Between 240 and 300 km (Layer 8), the subducted plate is displayed by a higher velocity band, underlined by the seismicity, which crosses the layer striking about NI60 E. The same velocity structure is observed in Layer 9, associated with the few events which indicate the maximum extent in depth of the Wadati Benioff zone. In Layer 10, the seismic quiescence is total but an elongated higher velocity structure, which clearly prolongates the one observed in the previous layers, is displayed. Finally, we observe again in Layers 11 and 12, down to 660 km, this band of positive anomaly in which the scarce deep seismicity takes place. In the northern part of Layer 10, the high velocity structure seems to change its trend and become more oblique to the west. The lateral extension of both deeper layers allows the confirmation of this change in orientation of the positive anomaly, which is particularly clear at the 560-660 km depth. Finally, whatever the depth, the velocity contrast associated with the higher velocity zone is weak and seldom exceeds 2.5%. 4. Discussion The images obtained by teleseismic tomography provide strong supportive evidence for the continuity of the slab over regions with gaps in seismicity. The close spacing of the temporary seismic stations together with the good distance-azimuth coverage of the seismic sources allow us to image with high resolution the subducting Nazca plate down to the lower mantle discontinuity. From 240 to 660 km depth, the slab is evidenced by a laterally continuous

C. Dorbath / Physics of the Earth and Planetary Interiors 101 (1997) 163-173 171 10" S 15" $ 20"$ 2S" $ 70"W 65"W Fig. 5. Comparative study of the subducted Nazca Plate in the Arica Elbow. The seismicity of the Wadati Benioff zone reported by the NEIS is plotted for the time period 1974-1994. The darker the dots, the deeper the hypocentres. The dashed contours to depth are from the updated catalogue of Kirby et al. (1995), as well as the terminology of the deep earthquake groups. The grey contours have been obtained from the tomography, by assigning the maximum of the positive perturbation related to the slab in each layer to its mean depth (solid line where the resolution is higher than 0.4, dotted line where it is lower than 0.4). The arrow ended lines show the location of the two cross-sections presented in Fig. 6. high-velocity structure, clearly displaying a pronounced bend in the deepest layers of our final model. In the larger scale study by Engdahl et al. (1995), the positive velocity anomaly associated with the slab in the Arica Elbow for the same depth range was of the same order of magnitude but parcelled out. Nevertheless, the interpretation made by these authors also displays a bend in the trace of this higher velocity structure in the transition zone. We present in Fig. 5 a schematization of the slab geometry deduced from Fig. 4. For every layer in which the slab is clearly identified by higher veloci- ties, we have drawn a contour passing through the maximum of the positive anomaly. Doing so, we approximate the position of the slab at the mean depth of the layer. In the regions where the resolution is lower than 0.4, the contour is represented by a dotted line. In the same figure, we have reported the seismicity of the NEIS for the past 20 years and the iso-depth contours obtained by Kirby et al. (1995). As the seismicity of slabs is concentrated in the upper part of the subducting lithosphere, there is not an exact correspondence between the iso-depth contours obtained for the maximum velocity contrast and for the seismicity. Nevertheless, the theoretical difference between both contours is well within the uncertainties of their positions, and we may compare them directly. Our 270-km contour is very short and does not fit well with the 275-km contour, the deepest of the Wadati Benioff zone. On the other hand, our 610-km contour is in good spatial agreement with the 600 km depth events. It presents the same strike as the southern group at 22-23 S as well as the 'jog group' at 13-14 S, and connects with a smooth bend the well-defined linear trend of these two seismic groups. The June 9, 1994 Bolivian earthquake occurred in the 'jog group' which includes mostly big shocks and has a scarce seismicity. This group is characterized by a WNW strike and a dip about 60 to NNE (Kirby et al., 1995). The main shock and 89 aftershocks were recorded by local permanent and portable instruments, and their spatial and temporal aspect was studied by Myers et al. (1995). These authors show that the deep seismicity can be fitted by a smoothly bent slab, which is clearly compatible with our tomographic image. The aftershock sequence was located with high precision; it defines a slab striking N68 W and dipping 45 NE, which is consistent with the seismicity being confined within the downward extension of the subducted plate. Therefore, the precise location and shape of the subducted Nazca plate at 560-660 km depth defined by the tomography is fully consistent with the results inferred by previous studies on the deep seismicity. In addition, tomography provides information on the dip of the slab throughout the seismically quiescent zone. Our network layout allows us to sample only a roof-shaped zone aligned on the seismic profile and consequently the north-south extension of the layers

172 C. Dorbath / Physics of the Earth and Planetary Interiors 101 (1997) 163 173 ~_ 100-2OO-- 3O0--" 5OO-- 600S 7O0 S Trench Distance along strike 100 200 300 4~ ~0 6~ 700 800 900 kill J I, I 1 I I I ; I i I, /, I i I -- -- Cahlllandlsa ~(1992) %% ofthejune9evem ~ f---= Kirby et al. (1995) ~% / I. ~ Thi~ study ~%{ Fig. 6. Profiles through the subducted Nazca Plate. Their location is shown in Fig. 5. The northern profile (dashed line in Fig. 5) is taken from Myers et al. (1995); in its aseismic part, the slab is inferred from Cahill and Isacks (1992) contours. The southern profile (solid line in Fig. 5) is drawn from Kirby et al. (1995) down to the end of the Wadati Benioff zone, and from this study underneath. depends on their depth. It is therefore not possible to draw from our tomography a vertical cross-section of the slab passing through the 1994 earthquake zone (cf. Fig. 5). We present in Fig. 6 two cross-sections of the slab in the Arica Elbow obtained by different methods. Their location is reported in Fig. 5 on the map view. The northern one, striking N30 E (dashed line on Fig. 5, grey line on Fig. 6), is extracted from the Myers et al. (1995) paper. The southern one, striking N65 E (solid line on Fig. 5, black line on Fig. 6), is directly deduced from Fig. 5, and is a compilation of the Kirby et al. (1995) contours in the seismically active part of the slab and of the contours deduced from our tomography underneath. On both profiles we observe that the dip of the slab changes with depth. On the northern profile, the dip of the slab increases with depth; it is interpolated from the intermediate depth seismicity down to the June 1994 shock zone, with a 45 mean dip. On the southern profile, from the trench to a depth of 275 km, the seismicity defines an approximately 30 dipping slab; then the dip, inferred from the tomography, increases to more than 50 in the aseismic part of the slab. Finally, the dip decreases to about 40 where the deep seismicity is present, suggesting a deflection of the slab at the bottom of the upper mantle. When only hypocentral locations are considered, the hypothesis of the continuity of the slab infers in any case that the dip of the slab increases in the aseismic part. Yet Kirby et al. (1995) observe that the deepest events in the 'Southern group' (Fig. 5) are deflected eastward. In this zone, the tomographic images of Engdahl et al. (1995) display a broad region of recumbent slab in the transition zone as well as regions of high-velocity material beneath the 660-km discontinuity. Both studies interpreted these observations by a partial deflection of the slab in the transition zone before continuing into the lower mantle. In contrast, apparent broadening but no deflection of the velocity anomaly occurs on Engdahl's images, at the bottom of the transition zone in the source region of the June 9 event. On the other hand, the constant spacing of the 510- and 610-km contours in Fig. 5 suggests that the deflection of the slab observed on our cross-section should exist in the whole region of the Arica Elbow. After the Grand (1994) large-scale study, the Engdahl et al. (1995) tomography displays clear evidence of the penetration of the slab in the lower mantle. Kirby and Engdahl interpret the differences that they observe of the velocity anomalies between the two regions of deep earthquakes ('Jog Group' and 'Southern Group') by differences in the level of resistance to penetration into the lower mantle. The vertical extent of our tomography does not allow us to image the lower mantle and consequently to be argumentative on this penetration in itself. Nevertheless, the shape of the slab, as it is defined by the high-velocity body on our tomography, does not suggest any difference in the penetration process along the bend in the Arica Elbow. Finally, the geometry of the deep slab in the transition zone of southern Peru is still an open question. If it is relatively easy to connect our 610-km contour with the Northern Group through a reverse bend, the closing up of the 200- and the 510-km contours at about 15 S involves a steepening of the slab. Therefore, the transition, from the 50 dip of the aseismic part of the slab deduced from the tomography, to the 70 dip inferred by the James and Snoke (1990) study beneath Peru, seems to occur in this region of the Arica Elbow, and implies a very complicated geometry of the slab between 13 S and 18 S.

C Dorbath/Physics of the Earth and Planetary Interiors 101 (1997) 163-173 173 Acknowledgements I would like to thank Anne Paul, Georges Poupinet, and every participant of the Lithoscope Andean Experiment (U. Achauer, M. Aldunate, T. Bianchi, J.P. Caminade, V. Farra, R. Fontanilla, M. Fornari, R. Guiguet, J. Guilbert, B. Guillier, G. Herquel, M. Lambert, C. Martinez, F. Masson, T. Monfret, C. Prquegnat, P. Soler and G. Wittlinger) and Bob Engdahl for helpful review and discussion. This project was sponsored by ORSTOM and INSU-CNRS. References Aki, K., Christoffersson, A. and Husebye, E.S., 1977. Determination of the three-dimensional seismic structure of the lithosphere. J. Geophys. Res., 82: 277-296. Barazangi, M. and Isacks, B.L., 1976. Spatial distribution of earthquakes and subduction of the Nazca Plate beneath South America. Geology, 4: 686-692. Barazangi, M. and Isacks, B.L., 1979. Subduction of the Nazca plate beneath Peru: evidence from spatial distribution of earthquakes. Geophys. J.R. Astr. Soc., 57: 537-555. Bevis, M. and Isacks, B.L., 1984. Hypocentral trend surface analysis: probing the geometry of Benioff Zones. J. Geophys. Res., 89: 6153-6170. Boyd, T.M., Snoke, J.A., Sacks, I.S. and Rodriguez, A.B., 1984. High resolution determination of the Benioff zone geometry beneath southern Peru. Bull. Seismol. Soc. Am., 74: 559-568. Cahill, T. and Isacks, B.L., 1992. Seismicity and shape of the subducted Nazca plate. J. Geophys. Res., 97: 17503-17529. Comte, D. and Suarez, G., 1995. Stress distribution and geometry of the subducting Nazca plate in northern Chile using teleseismically recorded earthquakes. Geophys. J. Int., 122: 419-440. Delouis, B., Cisternas, A., Dorbath, L., Rivera, L. and Kausel, E., 1995. The Andean subduction zone between 22 S and 25 S (Northern Chile): precise geometry and state of stress. Tectonophys., in press. Dorbath, C., Paul, A. and the Lithoscope Andean Group, 1996. A tomography of the Andean crust and mantle at 20 S: first results of the Lithoscope experiment. Phys. Earth Planet. Inter., accepted. Engdahl, E.R., Van der Hilst, R.D. and Berrocal, J., 1995. Imaging of subducted lithosphere beneath South America. Geophys. Res. Lett., 22: 2317-2320. Engebretson, D.C. and Kirby, S.H., 1992. Deep Nazca slab seismicity: why is it so anomalous? Eos, Trans. Am. Geophys. Un., 73(43): 379. Evans, J.R. and Achauer, U., 1993. Teleseismic velocity tomography using the ACH method: theory and application to conti- nental scale studies. In: H.M. Iyer and K. Hirahara (Editors), Seismic Tomography. Chapman and Hall, London, pp. 319-360. Grand, S., 1994. Mantle shear structure beneath the Americas and surrounding oceans. J. Geophys. Res., 99:11591-11621. Grange, F., Cunningham, P., Gagnepain, J., Hatzfeld, D., Molnar, P., Ocola, L., Rodrigues, A., Roecker, S.W., Stock, J.M. and Suarez, G., 1984. The configuration of the seismic zone and the downgoing slab in southern Peru. J. Geophys. Res., 11: 38-41. Harris, R.A., Iyer, H.M. and Dawson, P.B., 1991. Imaging the Juan de Fuca Plate beneath Southern Oregon using teleseismic P wave residuals. J. Geophys. Res., 96: 19879-19889. Herrin, E., 1968. Seismological tables for P phases. Bull. Seismol. Soc. Am., 58 1193-1242. Hirahara, K., 1981. Three-dimensional seismic structure beneath southwest Japan: the subducting Philippine Sea plate. Tectonophysics, 79: 1-44. Hirahara, K. and Hasemi, A., 1993. Tomography of subduction zones using local and regional earthquakes and teleseisms. In: H.M. Iyer and K. Hirahara (Editors), Seismic Tomography. Chapman and Hall, London, pp. 519-562. Isacks, B.L. and Barazangi, M., 1973. High-frequency shear waves guided by a continuous lithosphere descending beneath western South America. Geophys. J.R. Astron. Soc., 33: 129-139. James, D.E. and Snoke, J.A., 1990. Seismic evidence for continuity of the deep slab beneath central and eastern Peru. J. Geophys. Res., 95: 4989-5001. Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ramos, V.A. and Ando, C.J., 1983. Andean tectonics related to geometry of subducted Nazca plate. Geol. Soc. Am. Bull., 94: 341-361. Kirby, S.H., Okal, E.A. and Engdahl, E.R., 1995. The June 9, 1994, Bolivian deep earthquake: an exceptional event in an extraordinary subduction zone. Geophys. Res. Lett., 22: 2233-2236. Myers, S.M., Wallace, T.C., Beck, S.L., Silver, P.G., Zandt, G., Vandecar, J. and Minaya, E., 1995. Implications of spatial and temporal development of the aftershock sequence for the Mw 8.3 June 9, 1994 deep Bolivian earthquake. Geophys. Res Lett., 22: 2269-2272. Snoke, J.A., Sack, I.S. and Okada, H., 1974. Empirical models for anomalous high-frequency arrivals from deep-focus earthquakes in South America. Geophys. J. R. Astron. Soc., 37: 133-139. Stander, W., 1973. Mechanisms and spatial distribution of Chilean earthquakes with relation to subduction of the oceanic plate. J. Geophys. Res., 78: 5033-5061. Stander, W., 1975. Subduction of the Nazca plate under Peru as evidenced by focal mechanisms and by seismicity. J. Geophys. Res., 80: 1053-1064. Sykes, L. and Hayes, D., 1971. Seismicity and tectonics of South America and adjacent oceanic areas. Geol. Soc. Am. Programs, Cordilleran section, 3: 206. Wortel, R., 1982. Seismicity and rheology of subducted slabs. Nature, 296: 553-556. Wortel, M.J.R., 1984. Spatial and temporal variations in the Andean subduction zone. J. Geol. Soc. Lond., 141: 783-791.