Vous êtes sur la page 1sur 16
TECTONICS, VOL. 6, NO. 3, PAGES 233-248, JUNE 1987 RELATIVE MOTION OF THE NAZCA (FARALLON) AND SOUTH AMERICAN PLATES SINCE LATE GRETAGEOUS TIME Federico Pardo-Casas’ and Peter Molnar Massachusetts Institute of Technology, Canbridge Abstract. By combining reconstructions of the South American and African plates, the African and Antarctic plates, the Antarctic and Pacific plates, and the Pacific and Nazca plates, we calculated the relative positions and history of convergence of the Nazca and South Aner- ican plates. Despite variations in con- vergence rates along the Andes, periods of rapid convergence (averaging more than 100 mn/a) between the times of anomalies 21 (49.5 Ma) and 18 (42 Ma) and since anomaly 7 (26 Ma) coincide with two phases of relatively intense tectonic activity in the Peruvian Andes, known as the Late Eocene Incaie and Mio-Pliocene Quechua phases. The periods of rela- tively slow convergence (50 to 55 + 30 mn/a at the latitude of Peru and less farther south) between the tines of anomalies 30-31 (68.5 Ma) and 21 and between those of anomalies 13 (36 Ma) and 7 correlate with periods during which tectonic activity vas relatively quiescent. Thus these reconstructions Provide quantitative evidence for a correlation of the intensity of tectonic activity in the overriding plate at Now in Lima, Peru. Copyright 1987 by the American Geophysical Union. Paper number 710132. 0278- 7407/87 /0077-0132$10,00 subduction zones with variations in the convergence rate. INTRODUCTION Since Steinmann’s [1929] treatise on the geology of the Peruvian Andes, his inference that there have been a small number of brief phases of relatively intense tectonic activity, separated by periods of relative (if not complete) quiescence, has pervaded the literature on the tectonics of the Andes. Most Giscussions of the regional tectonics of particularly large areas of the Andes begin with this idea [e.g., Audebaud et al., 1973; Dalmayrac, 1978; Dalnayrac et al., 1980; Iberico, 1986; Laubacher, 1978; Maroceo, 1978; Martinez, 1980; NcKee and Noble, 1982; Mégard, 1978, 1984; Mégard et al., 1984; Noble et al., 1974, 1979], and many detailed investi- gations of small areas have specifically Addressed the timing of tectonic events with the presumption that these events were not local but widespread phenomena [eg. NcKee and Noble, 1982; Mégard, 1984; Mégard et al., 1984; Noble et al., 1974, 1979]. Although the number of "phases" has increased from Steinmann’s original three--the late Cretaceous Peruvian, the Eocene Incaic, and the Pliocene Quechua phases--to as'many as six [Mégard, 1984; Négard et al., 1984], the idea of periods of widespread quiescence punctuated by these brief phases persists. 234 Table 1. and Anomaly Uong™ 5 73-44 ~53.31 6 75.00 ~66.63, 7 60-58 40.38 10 52.08 ~34.68 3 43.69 -30.94 18 51.91 ~32.52 20 54.94 ~33.25, Pat 58.10 -33.95 25 60.53 34.57 30-31 62.96 35.30 31 32 62-44 ~36.56 33 62.81 33.38 34 63.06 -36.63 If such variations in the styles and amounts of deformation were to reflect variations in rates of crustal shortening in the Andes, then one might expect to observe concurrent changes in relative motion between the South American and Nazea (or Farallon) plates. To examine this possibility, we have determined the relative positions of these plates and uncertainties in these positions at the times of several well defined magnetic anomalies. Using these reconstructions, we can determine average rates of convergence for intervals between the ages of the anomalies ‘To reconstruct the Nazea and South American plates, one must reconstruct the Nazca to the Pacific plate, the Pacific to the Antarctic plate, the Antarctic to the African plate, and the African to the South American plate, For reconstruc- tions of the Nazea and Pacific plates we derived parameters and uncertainties for times younger than anomaly 13 (36 Ma), and we combined them with the parameters for reconstructions of the Farallon to the Pacific plate before the time of anomaly 13, which are given by Rosa and Molnar [1987]. For reconstructions of the Pacific to the Antarctic plate we used Stock and Molnar’s [1987] revision of their previous scenario [Stock and Molnar, 1982]. For reconstructions of Antarctica to Africa we used rotation Pardo-Casas and Molnar: South America Plate Motions Best and Partial Uncertainty Rotations End Points of Plate Boundaries wd Azimuths of Transform Faults t*Long™ — Lae® imuth Angle jong® Azimuth Angle 4.04 17.80 =14.90 -50-80 -9.50 078 8.30 19.20 =17.25 -49.85 -14.90 OBO -9.70 -11.09 13.24 29.60 -20.05 -50.85 -18.20 082 16.29 717,90 18.65 -22.35 -52.60 -20.35 085 20.13, 722.31 -25.05 25.60 -17.40 -26.75 44.60 33.80 086 27674 18.20 -28.00 -44.55 -35.45 8B, -31.00 15.20 ~30.28 -47.65 -36.80 089) -33.58 16.28 -31.45 -47.70 ~39.15 090 orth and East are positive. parameters and uncertainties of Molnar et al. [1987]. Finally, we derived para- meters and uncertainties the reconstruc- tions of Africa to South America. We describe the uncertainty in an individual reconstruction of neighboring plates at a specific time in terns of partial uncertainty rotations, small perturbing rotations about three ortho- gonal axes [Stock and Molnar, 1983]. For each reconstruction, one such axis lies in the middle of the reconstructed plate boundary, so that small rotations about it skew the fit of magnetic anomalies and fracture zones and cause an over lap at one end and a gap at the other. The other two axes lie 90° away in directions such that rotations about one axis cause a systematic misfit of fracture zones, and rotations about the other cause a systematic misfit of magnetic anomalies. Thus each reconstruction is described by the position of a pole and a rotation angle, and the uncertainty in the recon- struction is described by the positions of three poles, 90° from one another, and three small angles (Tables 1 and 2). Although, in the future, better methods for evaluating and describing uncertain- ties will surely be developed, this approach allows us avoid underestimating the uncertainties and hence helps prevent us from overinterpreting our results. The calculated reconstructions of the Pardo-Casas and Molnar: for the South Atlantic Ocean (Africa to South South America Plate Motions America) Skeyed Fit___Mismatched Fracture Zones Misnatched Magnetic Anomalies ong" AngleyeAugiey lst") ong’ AnglesoAneley oat? | Lone" Augie; Angle, - ~ eg __ ka deg ___kn__ 0.63 709.89 70.40 0.2225, 53-88 -33-41 0.13 15 1.02 113 B22 68.00 0.2225 54.22 -33.57 0.22 25 1.93 170 6473 67-79 0.3337 54640-31670 0.33 37 219 244 7.50 65.72 0.3337 48.96 -32.98 0.24 27 1.46 1626.10 65.53 0.2225 49-12 ~31.56 016 18 1.38 154 417 66.54 0.3337 54.07 -29.24 0.20 22 0.92 1034.06 65.62 0.2225 54-07 -30-00 0.1315 115 1283.88 64.40 0.28 31 54.08 -30.97 0.16 1B 1:38 1563.71 63.23 0.3337 54.10 31.91 0.20 22 L16 1303447 58.45 062831 5B.71 -37627 0619 21 0.93 104 58.16 0.22 25 58.72 37-49 0.15 17 0-96 107147157677 002225 58.51 39-02 0.09 10 0263 710.85 56.52 0.2225 58.52 -34-87 Os1l 12 0065 720.00 55.37 0.2225 57.95 -34.63 0.09 10 relative positions of the Nazca and South tmerican plates differ somewhat from those of Pilger [1981, 1983, 1984]. We used different paraneters to describe the evolutions of the Indian, South Atlantic and Pacific ocean basins, each of which is based on a reanalysis of the published magnetic anomaly and bathymetric data (Molnar et al., 1987; Rosa and Molnar, 1987; Stock and Molnar, 1987] instead of using parameters listed in published papers as Pilger [1981, 1983, 1984) did, Although the effect of the differences in most of these paraneters is small, the reanalysis of the original data was necessary for evaluating the uncertain ties in the reconstructions. The most important difference between our para- meters and Pilger’s is a major revision in the inferred history of the South Pacific [Stock and Molnar, 1987}. Pre- vious global reconstructions, including those of Pilger [1981, 1983, 1984), either ignored an unrecognized plate boundary there or guessed where it might Lie. Consequently, in general, they misrepresented motion between the Pacific and Antarctic plates. The recognition of this previously suspected plate boundary and a quantification of motion at i have resolved some of the difficulties in reconstructing the Pacific plate to its other neighbors for tines before anomaly 18 (42 Ma) [Stock and Molnar, 1987]. This revision includes no deformation of Antarctica, but the possibility of a few hundred kilometers of displacement between East and West Antarctica cannot be eliminated. It is remotely possible that there has been a relative rotation of these two parts of Antartica about a pole close to them, such that a large systematic error (larger than 300 km) exists in our reconstructions. We con- sider this sufficiently unlikely that we ignore it here. Movement of Fast with respect to West Antarctica of 300 km or less about a pole far from their boundary would alter the reconstructions of the Nazca and South American plates by a comparable anount but would not affect the observed correlation of rates of convergence with the intensity of tee- tonic activity on the South American plate Below, we first describe the recon- structions of the South American and African plates in the South Atlantic and of the Pacific and Nazca (or Farallon) plates in the Central Pacific. We briefly note how these were combined to yield Nazca-South America reconstructions and how uncertainties were calculated Then we examine the history of convergence between them. THE EVOLUTION OF THE SOUTH ATLANTIC BASIN We reexamined profiles of magnetic anomalies from as many published tracks in the South Atlantic as we could find (in 1984), and when we considered the 236 ‘Best Pole and Angle Anomaly Tat™ — Long® Angle 5 56-64 -87.88 -16.30 6 62.38 -93.02 -30.18 7 63.88 94.75 -39.08 10 67.34 -100.08 ~43.77 13 69.85 -106.13 -49.54 ist 73.13 -114.08 -56.86 20% 73-92 -117.66 -60.36 ait 74.76 -122.36 ~69.98 ast 78.94 -136-61 ~71.40 30-317 80.16 151-26 -77.87 {North and East are positive. Pardo-Casas and Molnar South America Plate Motions Table 2. Best and Partial Uncertainty 2.94 -108.64 ~32.27 122.37 NOSE 3.47 -115.09 -31.15 -130.34 N99E 12.72 129.26 -35.08 -128.16 N80E =13.75 -131.67 ~35.01 -130.68 N8OE 713.94 -134.45 -31.36 -136.70 NBOE {Poles and angles from Rosa and Molnar [1986]. Poles, angles, and partial uncertainty axes interpolated from Rosa and Molnar's [1986] values. anomalies to be clear, we measured the Positions of anomalies 5, 6, 13, 20, 31, 32, 33, and 34, according to the standard numbering system of Pitman et al. [1968] For some anomalies we did not measure the coordinates of crossings at the places where their numbering system would assign the position of each, but instead we measured the locations of particularly clear shapes in the magnetic anomaly profiles, such as at the older edges of broad anomalies 5, 31, 32, and 33 and the young edge of anomaly 34, We define the Positions of the parts of the anomalies that we picked by their ages (Table 3) according to the geomagnetic reversal chronologies of Berggren et al. [1985] for the Cenozoic era and of Kent and Gradstein [1985] for the Cretaceous period For readers only interested in the evolution of the South Atlantic basin, the reconstructions and the parameters describing them given here probably are not significantly different from those of Ladd [1974]. Our main goal has not been to improve them but rather to quantify the uncertainties in these reconstruc- tions, and in doing so, revision was easily done. To pick the positions of anomalies we relied heavily on the large maps of Cande and Rabinowitz [1979], of LaBrecque and Rabinowitz [1977], and of Ladd [1974], including some that he used but did not Publish. Although we did not use pre- cisely the same locations of magnetic anomalies that Ladd had measured, in general, ours and his agreed within a few Kilometers. In addition, we used the data of Barker [1979], Bergh and Barrett [1980}, Dickson et al. [1968], LaBrecque and Hayes [1979], and Rabinowitz and LaBreeque [1979]. Finally, from the GEBCO charts [Heezen and Tharp, 1978; Table 3. Ages Assigned to Magnetic ‘Anomalies “Anomaly Ago, Ha 5 10.59 6 19.90 a 25.82 10 30.03 13 35.58 18 42.01 20 45.41, 21 49.55 25, 58.94 30-31. 68.47 31 69.40 32 73.55 33 80.17 3a, 84.00 Pardo-Casas and Molnar: South America Plate Motions 237 Rotations for the Pacific Ocean (Nazca to Pacific) _Skeved Fit Miamatched Fracture Zones Mismatched Magnetic Anomalies Tat “Angle, Angle, Lat® Long” Angle, Angle, Lat’ Uong™ Angie, Angle, _ deg deg "ten 14.77 ~114.93 2.79° Bll -14.49 -21.02 0.90° 100. 69.07 -68.50 0.90° 100 (713.96 -122.13 5.58° 622 -8.73 -29.94 1.80° 200 73.44 -88-84 1.80° 200 ~23.90 -128.76 0.69% 7 9.13 -42.85 0.90" 100 64.21 -152.28 0.13° 15 24.38 -131.22 0.73% al 9.10 -45.38 0.90% 100 63.77 -154.35 0.13% 1s -23.65 -135.63 0.79° 8B 915-4968 0.90" 100 64.44 159.36 0.13% 15 $33.83 133.22 0.72" 80 -54.76 -114.76 0.36" 40 8.73 -37.31 0.36 40 Searle and Johnson, 1982] we digitized 25 km in assigning angles to partial the positions of the Chain and Romanche uncertainty rotations (Table 1). fracture zones, which lie near the equa- tor and north of where magnetic anomalies RECONSTRUCTIONS PACIFIC AND NAZCA (OR are clear, and the positions of the FARALLON PLATES) Falkland and Agulhas fracture zones (Figure 1). Points were digitized at Using Handschumacher et al.'s [1975] each degree of longitude where the frac- catalogue of magnetic anomalies plotted ture zones are clearly defined by the perpendicular to ship’s tracks, we bathymetry measured the positions of the magnetic Using Hellinger’s [1981] program, we anomalies 7, 10, and 13 on the Pacific searched for poles of rotation and angles and Nazca plates. These include the data that brought both magnetic anomalies and used by Handschumacher [1976] and Herron fracture zones of the same age into coin- [1972], From the bathymetric maps of cidence. These reconstructions were then Mammerickx and Smith [1976] we digitized plotted and examined to determine if the positions of the Marquesas, Agassiz, fits could be visually improved, with Mendafia, and Challenger fracture zones on final parameters describing the finite the Pacific and Nazca plates. Then, rotations given in Table 1. The scatter using the procedure described above, we in the positions of some anomalies, such sought parameters that brought these as 20, 33, and 34 on the African side, anomalies into coincidence (Figure 2, made it impossible to match all segments Table 2). The reconstructed positions of of anomalies with misfits of less than 20 the magnetic anomalies lie within 15 km km, but in general the individual posi- of their common lineations, but the tions of magnetic anomalies lie within 15 mapped fracture zones cannot be made to km of the reconstructed segments of plate overlie one another, because the dis- boundaries (Figure 1). Depending upon tances between the fracture zones on the the number of data and the quality of the two plates are different by 100 km. fit, we allowed misfits (overlaps or Accordingly, to quantify the uncertain- gaps) in the reconstructed magnetic anom- ties, we assigned large angles to the aly lineations of 10 to 37 km. Similarly partial uncertainty rotations that the reconstructed positions of the Chain describe the mismatch of fracture zones. and Romanche fracture zones Lie within 25, For the reconstructions of these km of one another. For the times of plates at the time of anomaly 5 (Table 2) anomalies 33 and 34, those of the we took the angular velocity of Minster Falkland and Agulhas fracture zones lie and Jordan [1978] and miltiplied the rate within 10 km of one another (Figure 1). dy 10.59 Ma, the age of the outermost We allowed misfits of fracture zones of edge of anomaly 5, which we have used in 238 I5°5 20°5| 25%5| 36rs| 40rg agg sors| Pardo-Casas and Molnar: South America Plate Motions 25ew _20°w sew tow Sew o sow SW SOW WOW ew IW WO BEE BFE Fig. 1. Map of magnetic anomalies in the South Atlantic, Solid symbols show the measured positions of anomalies on each side of the ridge, and open symbols show positions of anomalies from the African plate rotated, using parameters listed in Table 1, to their corresponding positions on the South American plate. Different symbols show different anomalies, assigned the nunbers written above and below lines correlating them. These numbers correspond to those defined by Pitman et al. [1968]. The seatter of the open and solid symbols about the lineations that they define provides measures both of the scatter in their measured positions (which reflect errors in navigation and identification) and of the quality of fits (which reflect the quality of the reconstructions). In addition the position of the Agulhas fracture zone on the African plate has been rotated to its corresponding position at the time of anomaly 34. The map is an oblique Mercator projection with the pole at 73,44*N, 53.31°W, the pole position for reconstructing positions of anomaly 5. Pardo-Casas and Molnar lors oy ee ~ —— 205 5 : 25°S & 4 § a a yo 2 30°S 218 a a) 8/8 |} i to 74 x ¥ gi 28 3) ——40°s 140°w 135°W I30°W I25°W «I20°W Fig. 2. Map showing reconstructions of the positions of anomalies 7, 10, and 13 from the Nazca plate (open symbols) to those on the Pacific plate (solid symbols), which is held fixed, Plus signs show crossings of the Marquesa (top) and Agassiz (bottom) fracture zones, and X's show rotated positions of the Mendaha (cop) and Challenger (bottom) fracture zones at the times of these three anomalies. Ellipses surrounding X's show the errors in the reconstructions that are allowed by the partial uncertainty rotations other studies. There is little evidence to check this reconstruction. Handschu- macher [1976] showed crossings of anomaly 5 at only 3 locations, two of which may have formed at the same segment of the spread- ing center (profile C-13 and 0c-2-73 in Handschumacher et al, [1975]). We rotated them to one another to check Whether it would be safe to assume a constant rate of spreading and Minster South America Plate Motions 239 and Jordan's [1978] angular velocity Because when rotated, their latitudes differ by 4° and the trend of the ridge crest at that time is undefined, it is difficult to make a definitive compari- son, Qualitatively the relative posi- tions seemed reasonable, but differences of 100 km probably would also be dif- ficult to rule out and we allow for a mismatch of that amount (Table 2) For anomaly 6 we interpolated between the reconstructions for anomaly 5 and the one that we determined for anomaly 7. There are no pairs of crossings of anom- aly 6, one on each plate, generated at the same segment of the ancient ridge crest, and thus no test of the parameters used can be made. Clearly the uncer- tainty at the time of anomaly 6 is large, and we allow for mismatches as large as 200 km (Table 2) For anomalies older than 13, we used parameters that describe the motion of the Farallon plate with respect to the Pacific plate between the times of anom- alies 32, 30-31, 25, 21, 18, and 13 (Rosa and Molnar, 1987] to determine where the Farallon plate lay with respect to the Pacific plate at these earlier times These parameters are based on magnetic anonalies and fracture zones not only from the South Pacific but also from part of the North Pacific; the fit of data from these two areas shows that both the Pacific and Farallon plates were rigid during the interval from about 70 to 35 Ma [Bngebretson et al., 1984; Rosa and Molnar, 1987]. The use of these para- meters carries with it the assumption that the spreading of the Pacific and Nazca plates was symmetric. We quantify this only by doubling the rotation angles that describe both the finite rotations and the partial uncertainty rotations; the angles given by Rosa and Molnar [1987] describe the positions of the Pacific plate with respect to the Paci- fic-Farallon spreading center, not with respect to magnetic anomaly lineations of the same age on the Farallon plate Although the parameters describing the reconstructions are different, the basic evolution of the Nazea plate described here is similar to that given by Hand- schumacher [1976]. ‘The Pacific and Farallon plates separated from one another at a long, roughly northerly trending spreading center from before the time of anomaly 32 (73.5 Ma) until sone time after anomaly 7 but before anonaly 6 240 Pardo-Casas and Molnar: South America Plate Motions Table 4 Latitudes and Longitudes of Poles and Rotation Angles for Reconstructing Points on the Nazca Plate to the South American Plate Anomaly Latitude, deg N Longitude, deg E ‘Angle 5 92.03 -98..24 = 9.92 6 62.62 -107.80 +21.79 7 60.95 -100.00 = 26.66 10 63.43 -103.00 =28.49 13 64.64 +104,93 +30.40 18 68.51 -123.50 -36.52 20 69.87 -133.07 -40,.60 21 70.29 -16d. 69, -46.34, 25 69.76 -174.12 253.11 31. 3.36 4169.48 -58.58 The lack of unambiguous observations of anomaly 6 and the clear jump in the spreading center after anomaly 6 formed [Herron, 1972] make it difficult to define the history of plate motion between the times of anomaly 7 (26 Ma) and anomaly 5 (11 Ma). NAZCA-SOUTH AMERICA RECONSTRUCTIONS AND ‘THEIR UNCERTAINTIES To calculate the positions of the Nazca and South America plates at different times, ve arbitrarily held South America fixed and successively rotated the Nazea plate to the neigh- boring plates (Table 4). For the times of most anomalies, it was necessary to interpolate between reconstructions at different times in at least one of the different oceans. In each case we inter- polated between the reconstructions in the individual ocean, not between the calculated parameters for positions of the Nazca and South American plates. For the South Atlantic we used the spacings between anomalies on the large maps of Cande and Rabinowitz [1979], LaBrecque and Rabinowitz [1977], and Ladd [1974], but for other oceans we simply assumed a constant rate of spreading in the inter- val surrounding the anomaly for which we obtained interpolated parameters. To assign partial uncertainty rotations for such interpolated parameters, we rotated the partial uncertainty pole positions for a neighboring anomaly to their posi- tions appropriate for the interpolated anomaly, and we increased the angles by 258 or 50% depending upon how reliable the interpolation seemed to be. As noted above, the partial uncertainty angles were chosen to be particularly large for anomaly 6 between the Nazca and Pacific plates To determine the uncertainties in the reconstructed positions of the Nazca and South American plates, we successively rotated the appropriate partial uncer- tainty poles to the South American plate and then combined them to determine uncertainties in the positions of the Nazca plate at the corresponding times [see Molnar and Stock, 1985]. To calcu- late the locations of parts on the Nazca plate that formed before the time of anomaly 13 for times since the time of anomaly 13, two separate rotations, and therefore two sets of partial uncertainty rotations, describing the separation of the Nazca (Farallon) and Pacific plates are necessary. One describes the rela- tive movement of the Farallon and Pacific plates before the time of anomaly 13 (taken from Rosa and Molnar [1987] assuming symmetric spreading). The second describes their relative displace- ment since that time (Table 2). Similar- ly, the reconstructions for times older than anomaly 18 include two separate rotations and two sets of partial uncertainty rotations for the South Pacific. One describes the reconstruc- tion of the Pacific plate to antaretica for the time of anomaly 18, and a second describes the relative motion of the Pacific and Antarctic plates before that time [Stock and Molnar, 1987]. THE HISTORY OF CONVERGENCE BETWEEN THE NAZCA AND SOUTH AMERICAN PLATES Since the time of anomaly 25 (59 Ma), the Nazca plate has moved steadily toward Pardo-Casas and Molnar: South America Plate Motions 241 NAZCA TO SOUTH AMERICA l2ow tow Fig. 3 Positions of two points on the Nazca plate, which formed at the tine of anomaly 30-31, plotted with respect to South America at the times of various magnetic anomalies anomaly 21 (49.5 Ma). South America (Figure 3) [Pilger, 1983, 1984]. Between the times of anomalies 30-31 (68,5 Ma) and 21 (49,5 Ma), the Nazca plate seems to have rotated about a pole in southern South America so that it converged with South America in the nor- thern but not the southern Andes, The uncertainties in the reconstructions are so large that no definitive statement can be made about the relative motion in the southern Andes (Figure 4h and 41), except that if there was convergence, it was not fast. A large component of right-lateral strike slip motion may have existed in the central part of the Andes. After approximately the time of anomaly 21 (49.5 Ma), however, changes in the direc- tion of relative motion were small [Pilger, 1983, 1984] and not resolvable given the uncertainties in the reconstructions (Figure 3) The clearest fact presented by these reconstructions is that the rate of convergence between the Nazca and South American plates has not been constant At the latitude of Peru, convergence was most rapid between about 50 and 42 Ma, Note the relatively steady convergence since the time of between the times of anomalies 21 and 18, and perhaps for a few million years before and after this interval. Rates reached 164 + 65 mm/a at the equator and 154 + 58 mm/a at 10°S. In Chile this period of rapid convergence is much less clear than in Peru (Figures 4 and 5), but there is a suggestion of its presence. The average rate of convergence in the preceding 20 my, was relatively low, only 55 # 28 mm/a at 10°S and decreasing southward along the Andes. The conver- gence rate between 36 and 26 Ma also was. relatively low, 50 + 30 mm/a at 10°S in Peru and 35 + 25 mm/a at 40°S in Chile (Figures 4 and 5). Since 26 Ma, the average rate has again been high all along the Andes: 110 + 8 mm/yr at 10°S and 112 + 8 mm/a at 40°S. There could have been other variations in convergence rates, such as a higher rate between 10 and 20 Ma than in the 5-10 Ma before and after it, but such variations are barely resolvable given the uncertainties in the reconstructions. The changes from slow convergence (or even divergence in south- ern latitudes) in late Cretaceous and 242 Pardo-Casas and Molnar: South America Plate Motions NAZCA TO SOUTH AMERCA, ANOMALY 5 TO PRESENT INAZCA TO SOUTH ANERICA, ANOWALY 6 TO ANOMALY 5 vow vow 20w row os. 208 408 Fig. 4a NAZCA TO SOUTH AMERICA, ANOMALY 7 TO ANOWALY & 2ow 7ow {20s —| 208 408 403 Fig. 4¢ Fig. 4d Fig. 4. Sequence of plots showing the displacement of the Nazca plate with respect to South America, between the times of selected magnetic anomalies Note the steady east-northeast to easterly convergence since the time of anomaly 21. Pardo-Casas and Molnar: South America Plate Motions 243 NAZGA TO SOUTH AVERICA, ANOMALY 13 TO ANOMALY 7 NAZEA TO SOUTH AMERICA, ANOMALY 18 TO ANOMALY 13 sow ow tn 208 208 NAZCA TO SOUTH AMERICA, ANOMALY 21 TO ANOMALY 18: NAZCA TO SOUTH AMERICA, ANOMALY 28 TO ANOWALY 21 sow row sow Tow os. 208 244 Pardo-Casas and Molnar NAZCA TO SOUTH AMERICA, ANOMALY 31 TO ANOMALY 25; ‘90 vow 208 South America Plate Motions NAZCA TO. SOUTH AMERICA, ANOMALY 31 TO ANOWALY 21 sow row tp 208 408 early Eocene time to fast convergence in late Eocene to slow again in Oligocene time and finally to fast again in Miocene and Pliocene time, however, are inescapable The periods of rapid convergence correlate remarkably well with two of the periods of high relatively intense tectonic activity in the Peruvian Andes; the late Focene Incaic and the Mio- Pliocene Quechua phases of Steinmann [1929] [Dalmayrac et al., 1980; Mégard, 1984; Mégard et al., 1984; Noble et al., 1979]. The third of Steinmann’s phases occurred at a time before which we can determine reliable reconstructions. The correlations of rapid convergence between the Nazca and South American plates at times when folding and thrust faulting were particularly active and of slower convergence when tectonic activity had been relatively quiescent is probably too clear to be coincidence. In fact, the correlation is better with the revised ages for these phases than it would be if Steinmann’s original more qualitative ages were used. Specifically, whereas Steinmann [1929] and later Mégard [1978] assumed that the Incaic phase spanned the entire Eocene epoch (= 60 to 40 Ma), precise dating in one locality led Noble Fig. 4j et al. [1979] to conclude that the defor- mation was most intense in middle to late Eocene time (= 50-40 Ma), when conver- gence was most rapid. Similarly, precise dating of Neogene rocks led Mégard et al [1984] to subdivide the Quechua phase into three short subphases, the first of which occurred in early Miocene time, and the last in Pliocene time, not just in Pliocene tine as Steinmann (1929) and Mégard [1978] had inferred earlier. One of us (P.M.), in fact, is somewhat enbar- rassed to admit that this correlation is a big surprise; he set off on this study expecting to find the contrary and has always doubted the existence of separate, brief tectonic phases in the Andes [see Molnar and Lyon-Gaen, 1987]. This correlation of rapid subduction with intense deformation in the Andes obviously implies that the tectonic activity of the overriding plate at subduction zones is strongly dependent on the convergence rate of the major plates Such a dependence had been suspected for a long time, but until the work of Engebretson et al. (1984b, 1986] and Jurdy [1984], there was little evidence to support this suspicion. Beginning from different assumptions, these authors reconstructed the positions of the Faral- Pardo-Casas and Molnar Lovecnua Phose— PERU Inaie Prove PERU 1o 2 B60 | To TIME (Ma) BEFORE PRESENT g 9 20 30 40 50 60 70 —Tt — cHiLe 408 200- RELATIVE VELOCITY (mmsa) -I00 Fig. 5. Plot of average rates of convergence as a function of time Average rates were calculated for intervals of time between selected magnetic anomalies, For the periods between the times of anomalies 7 (26 Ma) and the present and between the times of anomalies 30-31 (68.5 Ma) and 21 (49.5 Ma), the average velocities and their uncertainties are also shown. The small uncertainties in these average velocities derive from the long period of time over which they are averaged, Uncertainties in all these rates, however, do not include uncertainties in the ages of magnetic anomalies lon and North American plates at diffe- rent times in the Cenozoic and Mesozoic eras and found a correlation between a period of rapid convergence and the time of Laramide deformation in the western United States. Because the differences in their calculated velocities before and during the Laramide phase are smaller than those that ve report here and the South America Plate Motions 245 uncertainties in their rates are surely at least as large as ours, we renain unconvinced that their proposed correla- tion is resolvable. In any case, it appears that the intense tectonic acti- vity that has built mich of the structure along the western margins of both North and South Anerica is related to rapid subduction of oceanic lithosphere beneath them, In fact, the greater width of the Andes of Peru and Bolivia than that of most of Chile and Argentina might be a consequence of the longer duration of rapid subduction beneath the northern than the southern segment of the belt Rapid convergence alone, however, clearly is not a sufficient condition for Andean margins co be built; convergence in the western Pacific ocean at present is rapid, but Andean margins are absent, or minor if present at all. Molnar and Atwater [1978] suggested that the age of subducting ocean floor might be an impor~ tant factor in controlling the presence or absence of Andean margins or interare spreading. Interare spreading would be associated with subduction of old oceanic Lithosphere, greater than about 100 Ma, and Andean margins with ocean floor younger than 50 Na (see also England and Wortel, 1980]. The results for the andes do not help to clarify this because we cannot know well the age of the ocean floor subducted beneath South America before about 25 Ma, We cannot be sure how the Mesozoic spreading centers respon- sible for the Phoenix Lineations of Meso- zole magnetic anomalies in the western Pacific [Larson and Chase, 1972] evolved during the Cretaceous magnetic quiet interval. Therefore we do not know how the Cretaceous seafloor in the southwest Pacific, or on the Farallon plate, older than anomaly 32 formed. Seafloor of this age was subducted beneath South America only 15 Ma. Tf the evolution of the Paci- fic and Farallon plates in the Cretaceous period were simple, chen the age of ocean- fe lithosphre subducted beneath South america throughout the last 70 Ma would have been younger than 60 Ma, but we can- not prove this because of the uncertain- ties described above. It is thus tenpt- ing to suspect that Andean margins require rapid subduction (2100 mm/a) of young ocean floor (< 60 Ma) conciustons We have determined both the parameters and their uncertainties describing the 246 Pardo-Casas and Molnar: relative positions of the African and South American plates and of the Pacific and Nazea plates since late Cretaceous time. Combining them with those for the Pacific and Antarctic plates and for the Antarctic and African plates, we calou- lated relative positions of the Nazca and South American plates and hence the aver- age rates of convergence between them. Since 70 Ma the Nazca plate has converged with South America, but the rate has varied considerably. The most rapid convergence (>100 mm/a) occurred between about 50 and 42 Ma and since 26 Ma. Between 70 and 50 Ma and between 36 and 26 Ma, the average rates have been only 50-55 + 30 mm/a at the latitudes of Peru and less farther south. These phases of rapid subduction coincide with revised ages [Mégard et al., 1984; Noble et al., 1979] of Steinann’s [1929] two most recent phases of relatively intense tectonic activity in Peru: the late Eocene Incaic phase and the Mio-Pliocene Quechua phase Acknowledgments. We thank J. Ladd, for providing us with large charts of magnetic anomalies plotted perpendicular to ship tracks in the South Atlantic, J. Stock, for helping us with some calcula- tions, C. Stork for help with the Pacific- Nazca reconstructions, G. Suarez for encouragement and guidance in the early stages of this work, and D. C, Engebret- son, R. G. Gordon, and J. Ladd for criti- cal reviews of the manuscript. This research was supported in part by the National Science Foundation through grant O¢E-8400090 and by NASA through grant. NaG5-300 REFERENCES audebaud, E., R. Gapdevilla, B. Dal- mayrac, J. Debelmas, G. Laubacher, ¢ Lefevre, R, Marocco, ¢. Martinez, M. Mattauer, F. Mégard, J. Paredes, and P. Tomasi, Les traits géologiques essentiels des Andes Centrales (Pérou- Bolivie), Rev. Géogr. Phys, Géo! Dyn., 15, 73-114, 1973. Barker, P. F., The history of ridge-crest offset at the Falkland-Agulhas fracture zone from a small circle geophysical profile, Geophys. J. R. Astron, Soc 39, 131-145, 1979 Berggren, W. A., D.V. Kent, J. J. Flynn, and J.A. van Couvering, Cenozoic Geo- chronology, Geol, Soc, Am, Bu 96, 1407-1418, 1985, South America Plate Motions Bergh, H. W., and D. M. Barrett, Agulhas Basin magnetic bight, Nature, 287, 591- 595, 1980 Cande, 5., and P. D, Rabinowitz, Magnetic anomalies bordering the continental mar- gins of Brazil, Offshore Brazil Map Ser., Am, Assoc. Pet. Geol., Tulsa, Okla,, 1979 Dalmayrac, B., Géologie des Andes péruviennes! Géologie de 1a région de Huanuco: sa place dans une transversale des Andes du Pérou central, Trav, Doc ORSTOM, 93, Paris, 1978 Dalnayrac, B., G, Laubacher, and R Marocco, Géologie des Andes péruviennes Caractéres généraux de 1'évolution géologique des Andes péruviennes, Trav Doc, ORSTOM, 122, Paris, 1980. Dickson, G. 0., W. G. Pitman IIT, and J R. Heivtzler, Magnetic anomalies in the South Atlantic and ocean floor spreading, ° L_Geophys, Res., 73, 2087-2100,1968 Engebretson, D. G., A. Cox, and R. G Gordon, Relative motions between Oceanic plates of the Pacific basin, J. Res, . 89, 10291-10310, 1984a. Engebretson, D. C., A. Cox, and G. A Thompson, Correlation of plate motions with continental tectonics: Laramide to basin and range, Tectonics, 3, 115-119, 1984 Engebretson, D. ¢., A. Cox, and R. Gordon, Relative motions between oceanie and continental plates in the Pacific basin, Spec. Pap. 206, Geol, Soc, ém., 1986 England, P., and R, Wortel, Some conse- quences of the subduction of young slabs, Earth Planet, Sei, Lett., 42, 403-415, 1980, Handschumacher, D, W., Post-Eocene plate tectonics of the eastern Pacific, in The Geophysics of the Pacific Ocean Basin and its Margin, Geophys. Monogr: Ser., Vol. 19, edited by G. H. Sutton, MH! Manghnani, R. Moberly, and E. UL Mcafee, pp. 177-202, AGU, Kashington, D. C., 1976. Handschimacher, D. W., $, T. Okamura, and P. K. Wong, Magnetic and Bathy- metric Profiles from the Central and Southeastern Pacific: 10°N-45°S, 70°W- 150°W, Data Rep. 29 Hawaii Inst. Geophys., Honolulu, 1975 Heezen, B.C. and M, Tharp, Generalized Bathymetric Chart of the Oceans GEBGO), 5,12, Gan. Hydrogr. Office, Ottava, Ontario, 1978. Hellinger, §.J., The uncertainties of finite rotations in plate tectonics, J. Geophys. Res.. 86, 9312-9318, 1981. Pardo-Casas and Molnar: South America Plate Herron, E, M., Sea-floor spreading and the Cenozoic history of the south- eastern Pacific, Geol, Soc, Am. Bull 83, 1671-1692, 1972 Iberico, M., Gedlogia del Peru, in Gran Geografica del Perii, edited by J. Mejia Baca and E, Manfer, pp. 225-323, Barcelona, 1986 Kent, D. V., and F. M. Gradstein, A Cretaceous and Jurassic geochronology, Geol, Soc, Am, Bull, 96, 1419-1427, 1985 Jurdy, D. M., The subduction of the Farallon plate beneath North America as derived from relative plate motions, Tectonics, 3, 107-113, 1984 LaBreeque, J. L., and D. E. Hayes, Seafloor spreading history of the Agulhas Basin, Earth Planet. Sci Lett.. 45, 411-428, 1979. LaBrecque, J., and P. D, Rabinowitz, Magnetic anomalies bordering the continental margin of Argentina, Map Ser, Cat, 826, Am. Assoc, Pet, Geol,, Tulsa, Okla., 1977 Ladd, J. W., South Atlantic sea-floor spreading and Caribbean tectonics, Ph.D. thesis, Columbia Univ., New York, 1974, Larson, R. L., and C. G. Chase, Late Mesozoic evolution of the western Pacific, Geol, Soc, Am, Bull,, 83, 3627-3644, 1972 Laubacher, G., Géologie des Andes péruviennes: Géologie de la Gordillére orientale et de 1/Altiplano au nord et nord-ouest du lac Titieaca, Trav. Doo ORSTOM, 95, Paris, 1978 Mammerickx, J, and $. M, Smith, Bathymetry of the Southeast Pacific, Geol, Soc, Am, Map Ser., MG-26, 1976 Marocco, R., Géologie des Andes péruviennes: Un segment E-W de la chaine des Andes péruviennes: la déflection d’Abancay, étude géologique de la cordillére orientale et des hauts plateaux entre Cuzco et San Miguel sud de Pérou (12°30'S & 14°00'S), Trav. Doc. ORSTOM, 94, Paris, 1978, Martinez, C., Structure et évolution de la chaine hercynienne et de 1a chaine andine dans le nord de 1a cordillére des Andes de Bolivie, Trav. Doc ORSTOM, 119, Paris, 1980. McKee, E. H., and D. G, Noble, Miocene volcanism and deformation in the western cordillera and high plateaus of south-central Peru, Geol. Soc, Amer, Bull. 93, 657-662, 1982. Mégard, F., Etude géologique des andes du Pérou central, Mem. ORSTOM, 86, Paris, 1978. Motions 247 Mégard, F., The Andoan orogenic period and its major structures in central and northern Peru, J, Geol, Soc, London, 141, 893-900, 198% Mégard, F., D.C, Noble, B. H. McKee, and H. Bellon, Multiple pulses of Neogene compressive deformation in the Ayacucho intermontane basin, Andes of central Peru, Geol, Soc, am. Bull,. 95, 1108- 1117, 1984 Minster, J.B., and T.H. Jordan, Present- day plate motions, J. Geophys, Res 83, 5331-5354, 1978. Molnar, P. and T. Atwater, Interare spreading and cordilleran tectonics as alternates related co the age of subducted oceanic lithosphere, Earth Planet. Sei, Lett,, 41, 330-340, 1978, Molnar, P., and H, Lyon-Caen, Some simple physical aspects of the structure, support, and evolution of mountain belts, Geol, Soc, Am Men., in press, 1987 Molnar, P., and J.M, Stock, A method for bounding uncertainties in combined plate reconstructions, J, Geophys Res.. 90, 12537-12544, 1985 Molnar, P., F. Pardo, and J, Stock, Uncertainties in the reconstructions of the Indian, African, and Antarctic plates since late Gretaceous tine, Basin Res.. 1, (in press), 1987. Noble, D. C., E. H. McKee, E. Farrar, and U. Petersen, Episodic Cenozoic volcanism and tectonism in the Andes of Peru, Earth Planet, Sei, Lett., 21, 213-220, 1974. Noble, D. C., E. H. McKee, and F. Mégard, Barly Tertiary "Incaic" tectonism, uplift and voleanic activity, Andes of central Peru, Geol, Soc, Am, Bu: 903-907, 1979 Pilger, R. H., Plate reconstructions, aseismic ridges, and low-angle subduction beneath the Andes, Geol Soc. Am, Bull, 92, 448-456, 1981 Pilger, R. H., Kinematics of the South American subduction zone from global plate reconstructions, in Geodynamics of the Eastern Pacific Region, Carib- bean and Scotia Arcs, Geodyn, Ser. vol. 9, edited by R. Cabré, pp. 113- 125, AGU, Washington, D.c., 1983 Pilger, R. H., Cenozoic plate kinematics, subduction and magmatism: South Ameri- can Andes, J, Geol, Soc. London, 141, 793-802, 1984, Pitman, W. G,, E. M. Herron, and J. & Heirtzler, Magnetic anomalies in the Pacific and seafloor spreading, J. Geophys, Res., 73, 2069-2085, 1968 248 Pardo-Casas and Molnar Rabinowitz, P. D., and J. LaBreeque, The Mesozoic South Atlantic Ocean and evolution of its continental margins, JL. Geophys, Res.. 84, 5973-6002, 1979 Rosa, J. W. C., and P. Molnar, Uncertain ties in reconstructions of the Pacific, Farallon, Vancouver, and Kula plates and constraints on the rigidity of the Pacific and Farallon (and Vancouver) plates between 72 and 35 Ma, J Geophys, Res., 92, (in press), 1987. Searle, R. C., and G. L, Johnson, Generalized Bathymetric Chart of the Oceans (GEBCO), 5,08, Can, Hydrogr. Office., Ottawa, Ontario, 1982 Steinman, G., Geologie von Peru, Karl Winter, Heidelberg, 448 p., 1929. Stock, J. M., and P, Molnar, Uncertain- ties in the relative positions of the Australia, Antarctica, Lord Hove, and Pacific plates since the late Creta- South America Plate Motions ceous, J, Geophys, Res. 87, 4697-4714, 1982 Stock, J. M., and P. Molnar, Some geometrical aspects of uncertainties in combined plate reconstructions, Geology, 11, 697-701, 1983 Stock, J., and P. Molnar, A revised history of early Tertiary plate motion in the south-west Pacific, Nature, 325, 495-499, 1987, F, Pardo-Casas, Apartado 11-0262, Lima 11, Peru P. Nolnar, Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, MA 02139 (Received Novenber 3, 1986; revised February 3, 1987; accepted February 4, 1987.)

Vous aimerez peut-être aussi