<<

Earth and Planetary Science Letters 404 (2014) 250–260

Contents lists available at ScienceDirect

Earth and Planetary Science Letters

www.elsevier.com/locate/epsl

Influence of Peruvian flat- dynamics on the evolution of western Amazonia ∗ Caroline M. Eakin a, , Carolina Lithgow-Bertelloni b, Federico M. Dávila b,c a Department of Geology and , Yale University, New Haven, CT 06520, USA b Department of Earth Sciences, University College London, London, UK c CICTERRA-CONICET and Universidad Nacional de Córdoba, Córdoba, Argentina a r t i c l e i n f o a b s t r a c t

Article history: Convection in the Earth’s is mainly driven by cold, dense subducting slabs, but relatively little Received 27 February 2014 is known about how 3D variations in slab morphology and buoyancy affect mantle flow or how the Received in revised form 17 July 2014 surface above deforms in response (i.e. dynamic topography). We investigate this problem by studying Accepted 20 July 2014 the dynamics of an active region of flat-slab subduction located in Peru in South America. Here the Available online xxxx slab geometry is well known, based on the regional seismicity, and we have observations from the Editor: Y. Ricard local geological record to validate our models. Of particular interest is the widespread subsidence and Keywords: deposition of the Solimões Formation across western Amazonia that coincided with the development of dynamic topography the Peruvian flat-slab during the Mid-Late Miocene. This formation covers an extensive area from the flat-slab subduction foredeep to the Purus Arch located ∼2000 km away from the trench. Close to the the preservation Amazonia of several kilometers of sedimentary thicknesses can be easily accounted for by flexure. Based on an Solimões Formation estimate of the Andean loading we predict 2.8 to 3.6 km of accommodation space that spans 100 km. The spatial and temporal history of the Solimões Formation however, particularly the thick distal foreland accumulations up to 1.2 km deep, can only be matched with the addition of a longer-wavelength dynamic source of topography. Following the transition from normal to flat subduction, we predict over 1 km of dynamic subsidence (∼1500 km wide) that propagates over 1000 km away from the trench, tracking the subduction leading edge. This is followed by a pulse of dynamic uplift over the flat segment behind it. We therefore propose that a combination of uplift, flexure and dynamic topography during slab flattening in Peru is responsible for the sedimentation history and landscape evolution of western Amazonia that eventually led to the configuration of the Amazon Drainage Basin we know today. © 2014 Elsevier B.V. All rights reserved.

1. Introduction to account for widespread deformation in continental interiors, such as for the western US during the Laramide (Dickinson and Snyder, 1978; Bird, 1988). Subducting slabs represent the primary buoyancy force in Earth’s mantle and a critical source of surface deformation via their coupling to the overriding plate. Yet the impact of variations 1.1. Tectonic and geological setting in slab morphology and buoyancy (e.g. slab tears or the subduction of oceanic plateaus) on the dynamics of the mantle and how the The Peruvian subduction zone presents itself as an excel- upper plate deforms in response are poorly understood. The impor- lent modern-day natural laboratory to study the interaction be- tance of such along strike variability in subduction models is how- tween this flat-subduction and surface deformation. Extending ever becoming increasingly apparent (e.g. Guillaume et al., 2009; ◦ over 1500 km along strike, the Peruvian flat-slab segment (3 S Dávila and Lithgow-Bertelloni, 2013; Capitanio and Faccenda, ◦ to 15 S) is by far the most extensive region of flat subduction in 2012; Jadamec et al., 2013). Of particular concern are episodes ◦ the world (Fig. 1). The slab subducts at a normal dip (∼40 ) to of “flat” or “shallow” subduction that represent 10% of subduction zones worldwide today (Gutscher et al., 2000b), and often invoked a depth of around 100 km, most likely directly beneath the over- riding continental , and travels horizontally for several hundred kilometers before steepening again to the east (Cahill and * Corresponding author. Tel.: +1 5107178942. Isacks, 1992; Phillips and Clayton, 2014). The region is character- E-mail address: [email protected] (C.M. Eakin). ized by a distinct lack of contemporary arc volcanism and low heat http://dx.doi.org/10.1016/j.epsl.2014.07.027 0012-821X/© 2014 Elsevier B.V. All rights reserved. C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 251

caped towards the north into the Caribbean Sea (e.g. Hoorn et al., 1995), or exited towards the west through the Guayaquil gap into the Pacific Ocean (e.g. Potter, 1997), as the Andes had not yet grown into a full orogenic barrier. During the Late Miocene however, the northern Andes rose rapidly, shutting off drainage pathways to the north and west, and deepening and broaden- ing the foreland basins to the east (Jordan and Gardeweg, 1989; Kroonenberg et al., 1990; Hoorn et al., 1995, 2010a, 2010b; Nobret et al., 1996; Garzione et al., 2006). At the same time, a wide but relatively shallow subsiding system developed in western Amazonia, from eastern Peru to the Purus Arch (Fig. 1), rep- resented by the Solimões/Pebas Formation (Hoorn, 1993, 1994; Campbell et al., 2006; Wesselingh et al., 2006b; Gross et al., 2011; Latrubesse et al., 2010; Rodrigues Nogueira et al., 2013). De- position of this Solimões Formation ceased in the Mio-Pliocene, as the accumulation rate of terrigeneous material on the off- shore Amazon fan increased dramatically (Dobson et al., 2001; Latrubesse et al., 2010). This indicated a switch in the destina- tion of the bulk of Andean derived sediment, from deposition in the western Amazonian basins, to large-scale transportation Fig. 1. Tectonic setting of Peruvian flat-slab subduction and western Amazonia. The across the continent via a fully fledged Amazon River to the inset map highlights the study location and political boundaries for context. The Atlantic margin (Campbell et al., 2006; Figueiredo et al., 2009; total area thought to be covered by the Solimões Formation is outlined by the thick purple line, based on the maps of Hoorn (1994). This includes many differ- Latrubesse et al., 2010). ent individual basins such as the Acre and the Solimões, as well as other foreland At the time of deposition, the Solimões Formation recorded a basins represented by the letters (P: Putumayo, N: Napo, M: Marañon, U: Ucayali, wide area of subsidence that resulted in simultaneous partial in- MD: Madre de Dios). Isopachs of the Solimões Formation thickness in the Acre and filling of many different pre-existing sub-basins in the western Solimões Basins are available and shown in purple after Latrubesse et al. (2010). Amazonian region (labeled in Fig. 1), both foreland (e.g. Madre Volcano locations (red triangles) are from Siebert and Simkin (2002). The gray arrow represents the absolute plate motion of the from HS3-NUVEL1A (Gripp de Dios) and intra-cratonic (e.g. Solimões) (Hoorn, 1994). The and Gordon, 2002). The dotted black lines represent locations of cross-sections in E–W trending series of Acre, Solimões, and Amazonas basins   Fig. 3(a–a ) and location of the flexural calculation in Fig. 4(b–b ). (For interpreta- overlie the Amazonian subdividing the craton’s exposure tion of the references to color in this figure legend, the reader is referred to the at the surface into the northern Guyana and southern Brazilian web version of this article.) Shields (Fig. 1), the result of an extensional event during the Mid- Proterozoic (Wanderley-Filho et al., 2010). Following on during the flow (∼20 mW/m2) (Henry and Pollack, 1988), in contrast to the ◦ Late-Proterozoic some of the original extensional grabens were normally dipping (∼30 ) regions to the north and south. positively inverted to form basement arches which now separate Subduction has been operating along the western coast of Peru the individual sub-basins, for example the Purus Arch which par- since at least the Jurassic (Martinod et al., 2010), with the present titions the Solimões basin from the Amazonas basin and is the plate configuration stable since 23 Ma, following the formation of eastern limit of the Solimões Formation, was an inversion of the the Nazca and Cocos plates (Lonsdale, 2005). The presence of the Cachimbo Graben (Wanderley-Filho and Costa, 1991). These Pre- Peruvian flat-slab is however relatively recent in comparison. Al- cambrian structural highs have therefore played an important role though the precise date is unclear and is still a matter of debate, in sub-division and configuration of sedimentary deposits across slab flattening is likely to have started around the Mid to Late western Amazonia ever since. Miocene (Bissig et al., 2008; Ramos and Folguera, 2009)around the ∼ time when the first reached the trench ( 11–15 Ma: 1.2. Sources of topography and their effect upon the landscape Hampel, 2002; Rosenbaum et al., 2005) and the pattern of arc vol- canism began to change (Aleman, 2006; Bissig et al., 2008). Full The evolving landscape and drainage of lowland Amazonia has flat-slab development is likely a long process that can last from typically been attributed to interaction of climate, sea level and the several to 10 Myr (Arrial and Billen, 2013; Espurt et al., 2008; developing Andean mountain belt. In particular the intense period Martinod et al., 2010; van Hunen et al., 2002a, 2002b). The var- of uplift during the Miocene is thought to have increased erosional ious stages of progression include broadening of the volcanic arc rates, sediment fluxes, and the deformation of inland Amazonia via and crustal deformation, adakitic , thinning and cooling flexural processes (Hoorn et al., 1995, 2010a, 2010b; Figueiredo of the asthenospheric wedge resulting in a decline and eventual et al., 2009; Mora et al., 2010; Latrubesse et al., 2010). Overall cessation of volcanic activity (Gutscher et al., 2000a). Over central most of the hypotheses for how Andean loading and flexural sub- Peru there is evidence that the arc broadened around 11–12 Ma sidence initiated widespread deposition of the Solimões Formation (Aleman, 2006; Bissig et al., 2008; Ramos and Folguera, 2009), are generally qualitative in nature, and, as far as we are aware, that formed 3–6 Ma (Gutscher et al., 2000a) and that no detailed flexural calculations have yet been made to demon- magmatic activity ended by 3–4 Ma (Soler and Bonhomme, 1990; strate such extensive subsidence. Given that the Solimões Forma- Rosenbaum et al., 2005; Pilger, 1981; Espurt et al., 2008; Martinod tion extends considerably far inland (as far east as the Purus Arch, et al., 2010). ∼2000 km from the trench), affecting the pericratonic foreland, Whilst shallow subduction was beginning to take hold beneath such as the Solimões Basin (Fig. 1), additional explanations for such Peru, the surface above was also going through its own trans- a broad episode of subsidence are warranted. The concurrent rela- formation. Eventually this transformation would lead to the de- tionship between basin development and drainage reorganization velopment of the largest drainage basin in the world, the Ama- with the arrival of the Peruvian flat-slab has been noted previ- zon River system, and the most diverse habitat on the planet, ously (e.g. Latrubesse et al., 2010), but only isostatic or flexural the Amazon Rainforest. Initially, during the Early-Mid Miocene, influences associated with plate interactions during flat-subduction it has been suggested that regional drainage of Amazonia es- were considered (i.e. rapid uplift of the Eastern Cordillera and 252 C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 more extensive deformation in the Subandean foreland basins). Overall such broad view techniques tend to have difficulty repli- This isostatic contribution to western Amazonian topography is es- cating the local geology at the surface, which closely depends on timated and discussed further in Appendix A. The dynamic impli- a broad range of factors. Namely the interaction of both tectonic cations of the flat-slab arrival however and the resulting expression and dynamic controls on topography requires greater considera- at the surface (i.e. the dynamic topography) is another potential tion, as well as the along strike variability in subduction style. source of topography that has yet to be considered. For western Amazonia, a direct comparison of , flexural Dynamic topography describes deflections of the Earth’s surface deformation, and dynamics of the Peruvian flat-slab, with the sub- that are supported dynamically by flow of the mantle below. In surface sedimentary record has yet to be made. A representation particular, convection of the mantle generates normal stresses that that takes such factors into account is therefore timely and re- act at the base of the lithosphere, which deflects the lithosphere quired, so that we can obtain a more complete picture of all the generating dynamic topography. For example a buoyant plume will topographic forces that helped mold the unique Amazon Basin. rise, pushing up on the surface, producing dynamic uplift (e.g., Olson and Nam, 1986; Lithgow-Bertelloni and Silver, 1998). Like- 2. Slab geometry and density structure wise a dense slab will sink, pulling down on the surface, creat- ing dynamic subsidence (e.g., Hager, 1984; Mitrovica et al., 1989). The dynamic topography of flat-slab regions is currently a Such dynamic topography is not in isostatic equilibrium and has subject of debate. Flat-slab episodes have often been equated ∼ relatively low amplitude ( 1km) and long wavelength (typically with large shallow mass anomalies, resulting in prominent long- 100’s–1000’s km) (Braun, 2010). wavelength dynamic subsidence and the formation of sedimen- It has long been known that such deflections of the Earth’s sur- tary basins directly above them (e.g. Liu et al., 2011). Dávila et face and its internal boundaries provide vital contributions to the al. (2010) however demonstrated at the modern Chilean/Argentine long wavelength geoid (e.g. Hager et al., 1985). In the last decade flat-slab segment that dynamic subsidence does not occur directly or so the importance of dynamic topography at shorter wave- over the flat slab but is instead offset to the east, where subduction lengths has also taken on increasing prominence, for example in steepens again at the slab leading edge. Likewise for the Alaskan the development of continental basins (Burgess et al., 1997; Heine flat subduction zone, Jadamec et al. (2013) suggest that the po- et al., 2008), in Australia (Gurnis et al., 1998; Heine et al., 2010), in sition of slab steepening is controlling the location of the Cook North America (Spasojevic et al., 2009; Liu and Nummedal, 2004; Inlet Basin. Furthermore, Dávila and Lithgow-Bertelloni (2013) ar- Liu et al., 2011; Jadamec et al., 2013) and in Africa (Pysklywec gue that flat-slab segments should have neutral buoyancy based on and Mitrovica, 1999). For South America there have been recent evidence for abnormally thick alongside a kinetically efforts to study dynamic topography of the continent as a whole hindered basalt to reaction (e.g. Cembrano et al., 2007; (Shephard et al., 2010; Dávila and Lithgow-Bertelloni, 2013)as well Gans et al., 2011). Several modeling studies demonstrate flattening as focused investigation of the Argentine Pampas region (Dávila of a slab when over-thickened oceanic crust is subducted with par- et al., 2010), and subduction of the ridge beneath Patagonia tial to zero (van Hunen et al., 2002a, 2002b; Arrial (Guillaume et al., 2009). Shephard et al.’s (2010) study primar- and Billen, 2013). Under this assumption flat-segments produce lit- ily focused on relating the drainage reorganization of Amazonia tle or no dynamic topography, meanwhile a shift from normal to during the Miocene and the formation of the modern day Ama- flat-slab subduction actually results in relative dynamic uplift over zon River System, to the evolution of subduction driven dynamic the transitioning region. This relationship between flat-slabs and topography. The geometry and density content of the subducted dynamic topography can therefore be constrained further by study- slabs beneath South America were derived from the global S-wave ing present day, active systems such as in Peru and comparing the tomography of Grand (2002). results with geological observations. This type of approach however has two important caveats. Predictions of dynamic topography and their geological expres- Firstly, whilst first order changes in the Amazonian drainage are re- sion have been shown to be highly sensitive to slab morphology produced, the absolute amount of dynamic subsidence is too large. and density structure i.e. the slab model (Dávila et al., 2010; This is because the tomographic models provide a low-resolution Jadamec et al., 2013). Previous predictions over the northern An- image of a spatially large and high-density anomaly in the lower des and western Amazonia region have used simplified models mantle. This expansive starting anomaly drives strong downwarp- of the subducting slab. For example, in global models, subduc- ing on the land above, predicting flooding over most of the South tion zones are replicated by dropping slablets vertically at the American continent. Evidence for flooding to this extent is not trench (e.g. Lithgow-Bertelloni and Gurnis, 1997). In regional mod- seen today nor in the geological record (Harrison et al., 1983; els, Dávila and Lithgow-Bertelloni (2013) took into account a more Hoorn, 2006; Lithgow-Bertelloni and Gurnis, 1997). accurate flat-slab morphology by modeling a horizontal segment ◦ Secondly, variations in the subduction morphology and density under Peru and then a 30 dip further to the east. Other mod- content along strike are not reproduced. The western margin of els based on global S-wave tomography, such as Shephard et al., South America is characterized by an alternating pattern of steep (2010, 2012), typically struggle to resolve subducting slabs above and flat slabs, but these are not easily observed by body-wave to- the transition zone, especially over South America where station mography, due to a lack of crossing ray-paths in the upper mantle. coverage is lacking (Grand, 2002). Only recently have seismic de- This is especially true for northern South America where station ployments been operating in Peru (e.g. PULSE and CAUGHT: http :/ / coverage is particularly poor. This is concerning because sensitiv- www.iris .edu /hq /science _highlights /pulse _and _caught _seismology _ ity kernels indicate that surface dynamic topography is particu- in _the _high _andes, and PeruSE: http :/ /www.gps .caltech .edu /~clay / larly sensitive to upper mantle structure (whereas the geoid has a PeruWeb /PeruSE .html), mostly covering the southern portion of higher sensitivity to the lower mantle) (Hager and Richards, 1989). the flat-slab region. Tomographic resolution of the slab will there- If we want to be able to match and understand the geological fore likely improve in the future as new data and published results history of a region (i.e. western Amazonia), then the details of sub- become available. At the current time however, such a tomographic duction in the upper mantle, such as along strike variations in slab approach is unable to capture the flat-slab shape of the downgoing dip and density structure, are critical, not just to the tectonic com- Nazca plate, both in the most comprehensive global studies (e.g. Li ponent but also to the dynamic signal of topography (Dávila et al., et al., 2008) and even from upcoming preliminary results of the 2010; Dávila and Lithgow-Bertelloni, 2013). regional seismic projects (e.g. Scire et al., 2013). C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 253

Fig. 2. Comparison of dynamic topography predicted by two different slab models for the Peruvian subduction zone, shown in 3D (top row) and 2D (bottom row). The left hand side (a and c) represents the present day 3D flat-slab morphology determined from local seismicity (small black circles give individual locations), whilst ◦ the right hand side (b and d) reflects the uniform 30 dip slab model akin to ‘normal’ subduction before the onset of slab flattening. (a–b) Dynamic topography (blue mesh surfaces) has been vertically exaggerated for illustrative purposes and plotted up to spherical harmonic degree 40. The slab surfaces are color-coded according to the density anomaly assigned. (c–d) Same as figures (a–b) but now in map view, comparing the dynamic topography against the slab contours (white dashed lines). Contours of seafloor age (on which the density anomalies are based) are from Müller et al. (2008). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) Given the importance of the details of the slab model to the to 700 km. An extensive flat-slab segment is clearly reproduced dynamic topography predictions and therefore to connect with the (Figs. 2a and 2c) by an abundance of seismicity in the 70–200 km geology, we aim to map out a realistic and improved geometry depth plane, that extends ∼500 km away from the trench. The of the Nazca slab. Based on the current available resources, this slab model then rapidly drops off towards a cluster of deep seis- is best achieved for the Peruvian flat-slab region by utilizing the micity (500–700 km depth) further inland. Although there is a seismicity (i.e. the Wadati–Benioff Zone) to delineate the slab paucity of intermediate seismicity, the leading edge of subduction shape (Fig. 2a). Over 300 (M > 3.0) from the NEIC cat- in our model is well defined by the sharp transition from shallow alog were used (http://earthquake.usgs.gov/earthquakes/eqarchives/ to deep events. It also matches well with preliminary tomographic epic/). This catalog contains events from 1970 to present and is images of the Nazca slab in the upper mantle (with resolution be- the most complete and up-to-date compilation available for the low 200 km) using teleseismic body-waves (Scire et al., 2013). region. Events with depths shallower than 70 km were removed Density anomalies (relative to background mantle, up to to eliminate events that may have originated in the continen- 50 kg/m3) were also assigned to the slab model based on the tal lithosphere. Blockmean and surface gridding algorithms from age of the subducting oceanic lithosphere at the trench (Müller et GMT (Wessel and Smith, 1991)were then used to interpolate the al., 2008)(Fig. 2). For the segment where the slab is horizontal, points/events and recreate the inferred shape of the slab down the density anomaly was set to zero (δρ = 0 kg/m3) (analogous to 254 C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 the Pampean flat slab of Dávila et al., 2010), to replicate neutral buoyancy. This assumption can be additionally justified by initial surface-wave tomography results from the region (Knezevic An- tonijevic et al., 2013), based on data from the recent seismic deployments. These new images show that the Peruvian flat-slab is predominantly low-velocity, indicating the presence of lower den- sity material relative to ‘normal’ subducted oceanic lithosphere (i.e. dense sinking slabs), which typically display a high-velocity seis- mic signature. This would suggest that the transition from basalt to higher density eclogite within the oceanic crust is indeed un- dergoing a delay in the region of flat subduction. In addition to a more realistic flat slab, we also made a slab ◦ model for normal subduction, i.e. for a slab with uniform 30 dip angle (Figs. 2b and 2d), based on the subduction angle to the north and south of the flat-slab region (Cahill and Isacks, 1992; Phillips and Clayton, 2014). This is representative of how sub- duction may have looked prior to the arrival of the flat slab. By Fig. 3. Profiles illustrating the different contributions on the total topography across comparing the two different slab models we are able to visualize the Amazonian foreland. The thick profiles depict the impact of relative dynamic topography given a transition from ‘normal’ subduction (thick red line) to flat-slab the temporal change in dynamic topography felt by the landscape  since the onset of slab flattening. subduction (thick blue line) (see location of the profile a–a in Fig. 1). This results in relative dynamic uplift beneath the Peruvian Andes (High Cordillera) but relative dynamic subsidence beneath the western Amazon Basin (distal foreland). The Sub- 3. Predicted dynamic topography and flexure Acre/Solimões basin (gray dotted region) is outside the influence of flexure driven by the mountain loading (thin green lines), but does coincide with relative dynamic 3.1. Dynamic topography computation and results subsidence from the arrival of the flat-slab. Depth of the basin (dotted purple line) determined from an isopach map (Latrubesse et al., 2010). Present-day topography (dot–dash black line) taken from ETOPO1 (Amante and Eakins, 2009). (For interpre- We calculate the dynamic topography resulting from the in- tation of the references to color in this figure legend, the reader is referred to the stantaneous viscous flow induced by a prescribed density field web version of this article.) (i.e. the slab), by solving the equations for the conservation of mass and momentum via propagator matrices in a spherical shell eastwards as the flat-slab develops, followed by a pulse of dynamic (Hager, 1984; Hager and O’Connell, 1979, 1981). Density and flow uplift behind it. fields are computed to spherical harmonic degree and order 50 (plotted to degree 40). The mantle is assumed to have a New- 3.2. Suitability of the dynamic topography modeling approach tonian rheology and varies only with radius. The dy- namic topography we calculate at any given time is indepen- The dynamics of subduction systems are controlled by a com- dent of the absolute viscosity but the amplitude does depend petition between the buoyancy forces (i.e. the slab model input), on the relative viscosity contrast between each layer (the loca- which drives the flow, and the rheology (i.e. the viscosity profile), tion of peak dynamic subsidence or uplift is unaffected). The which acts to resist flow. When attempting to model these dynam- rate of change of dynamic topography would however depend on ics there is usually a trade-off between capturing the complexity the absolute viscosity. We present results for a viscosity struc- of the rheology versus the buoyancy. To incorporate both realistic ture of 10 for the lithosphere (which is uniformly 120 km thick) slabs (e.g. 3D geometries derived from seismicity and lateral vari- and 50 for the lower mantle as normalized to the upper mantle, ations in slab buoyancy) as well as viscosity variations that can based on the best correlation with the geoid (Ricard et al., 1993; cover many orders of magnitude (e.g. temperature and strain-rate Lithgow-Bertelloni and Richards, 1998) and consistent with pre- dependence) requires intensive computational power and time, on vious studies (Lithgow-Bertelloni and Gurnis, 1997; Dávila et al., the order of several hundred processors for days at a time for 2010). Using this mantle viscosity structure the dynamic topogra- each model run (Jadamec et al., 2012, 2013), which can be rather phies predicted by each of the slab models are shown in Fig. 2. restrictive. For a comprehensive review of modeling mantle flow ◦ For the uniform 30 dip slab model, representing ‘normal’ sub- in subduction zones see Billen (2008) and the references therein. duction, dynamic subsidence is predicted underneath the Peru– In the case of our flow model, the rheology is relatively simpli- Chile trench (thick black line in Fig. 2b). Taking into account a fied, including only radial variations in viscosity. By contrast, we more realistic flat-slab model (Figs. 2a and 2c), dynamic subsi- can easily include a high degree of detail in our slab morphol- ◦ ◦ dence is considerably offset from trench, by well over 1000 km ogy and buoyancy (on the order of 0.5 × 0.5 ), whilst keeping inland. This depression is located to the east of the flat-slab, di- the computational intensity low and accessible, which allows one rectly over the region where the slab begins to steepen again and to sweep through many different morphological scenarios. Pre- sink back down into the mantle. In contrast, directly over the flat vious similar studies have indicated that a high-resolution im- segment, dynamic topography values are smaller (∼500 m) and age of the slab, with density structure varying over short length positive in value. For both slab models the dynamic subsidence scales, is critical for correlating the dynamics with the regional predicted is long-wavelength (∼1500 km) and comparatively high geology (Dávila et al., 2010; Dávila and Lithgow-Bertelloni, 2013; in amplitude (>1 km). Jadamec et al., 2013). Additionally recent observations of seismic ◦ If we assume that our ‘normal’ slab model (uniform 30 dip) anisotropy below the Peruvian flat-slab strongly suggest that man- is representative of subduction in the region prior to the arrival of tle flow in the region is closely tied to the local slab morphology the flat-slab (see also Dávila and Lithgow-Bertelloni, 2013), then (Eakin and Long, 2013; Eakin et al., 2013). We therefore argue that a transition from normal to flat subduction would predict tran- as the driver of flow, the buoyancy force is the main controlling sient relative dynamic uplift across the Andes but relative dynamic process in subduction driven dynamic topography. subsidence within the distal pericratonic foreland and western If we were able to include rheological complexities in our Amazonia (Fig. 3). Over time therefore we would expect a wave flow models, then lateral variations in viscosity would be the of dynamic subsidence moving out from the trench, propagating next step, most likely in the form of temperature dependent C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 255 viscosity. For example, cold subducting slabs are expected to have accommodation space next to the load but the flexural basin di- a higher viscosity than the ambient mantle (King and Hager, 1994; minishes within about 100 km (gray shaded area in Fig. 4). Moresi and Gurnis, 1996; Zhong and Davies, 1999), and as the vis- We show in Appendix B that even if we consider a simpler cosity of a slab increases, the dynamic subsidence of the surface cross-section across the full orogen (Fig. B.1), and vary the load above also strengthens, deepening the dynamic basin produced distribution (Fig. B.2), the wavelength of the flexural basin remains (Billen et al., 2003; Zhong and Gurnis, 1992, 1994; Moresi and Gur- on the order of 100 km (minimum: 69 km, maximum: 115 km). nis, 1996). If this applied to the Peruvian subduction zone, then Dávila and Lithgow-Bertelloni (2013) also show a similar result of our estimate of dynamic subsidence beneath western Amazonia, <100 km width based on cross-sections from Mégard (1987) and would represent a minimum (i.e. the true dynamic signal could Hermoza et al. (2005) that are less detailed than what we use here. potentially be greater). The actual strength profile of subducting We expect therefore that the flexural subsidence should be an or- slabs is however relatively uncertain and potentially rather com- der of magnitude narrower than the dynamic subsidence. plex (e.g. Hines and Billen, 2012). This uncertainty increases even more for flat-slab regions such as Peru, where the slab geometry 4. Impact on western Amazonia and the material being subducted are atypical (i.e. the flat shape and subduction of the Nazca Ridge), so we might not expect the 4.1. Comparison with the geological record same rheological properties to apply. Even if we had a good idea of what the strength profile should To constrain and interpret dynamic topography, we need to be, the effect of short wavelength viscosity variations on dynamic compare with available geological proxies that are sensitive to to- topography (and therefore the geoid) is poorly understood (Billen pographic changes. Given that sediment dynamics respond to gra- et al., 2003) and difficult to constrain (Moucha et al., 2007). In dient changes, not only in thicknesses but also in facies, the strati- order to significantly impact our results, a large contrast in vis- graphic record is a good proxy with which to compare our dynamic cosity would most likely be required. The most plausible way for topography predictions. We analyze the Solimões Formation that this to occur would be via the presence of melt in the mantle covers a large area of western Amazonia (Fig. 1), extending from wedge to drastically lower the wedge viscosity. The wedge vis- the most proximal foreland depocenters (foredeep), to pericratonic cosity could also be lowered by several orders of magnitude by settings (e.g. the Solimões Basin), located more than 500 km away the dislocation creep deformation mechanism for olivine and the from the easternmost relief. presence of high strain-rates i.e. such as expected in the wedge Given the inaccessibility and vastness of the Amazonia re- (Jadamec and Billen, 2010, 2012; Stadler et al., 2010), though gion, sampling of the formation is difficult and sporadic, with the presence of melt would likely nullify this mechanism. Such most information sourced from riverbank exposures and road a low viscosity wedge (LVW) would act to decouple the slab cuts (Hoorn, 1993, 1994; Räsänen et al., 1995; Latrubesse et from the surface, weaken the dynamic topography and reduce al., 1997, 2007, 2010; Cozzuol, 2006; Wesselingh et al., 2006a; the overall depth of the dynamic basin (Billen and Gurnis, 2001; Gross et al., 2011; Rodrigues Nogueira et al., 2013) or from wells Billen et al., 2003). However this effect is limited, as dynamic drilled by hydrocarbon exploration in the 1970s (Maia et al., 1977; topography is the result of normal stresses, which will still be Del’Arco et al., 1977; Hoorn, 1993; Wesselingh et al., 2006b; able to transmit through the to the surface even Silva-Caminha et al., 2010; Wanderley-Filho et al., 2010). Based on when the asthenosphere is relatively weak ( stresses however these various geological surveys an isopach map of the Solimões would be adversely impacted). Additionally the potential impact Formation is available (see Fig. 1 and Fig. 5c) (Latrubesse et al., of a LVW is restricted above the Peruvian flat-slab where there 2010), that illustrates the thicknesses of the sedimentary beds is very little space available for mantle material between the flat- within the Solimões and Acre basins (>1 km in the Acre Basin). It should be noted that the Solimões Formation is present over a slab and the overlying (Cahill and Isacks, 1992; much wider area (thick purple line in Fig. 1 and Fig. 5c) than that Phillips and Clayton, 2014), but a LVW could exist to the east covered by the availability of the isopachs, and that these isopachs where the slab steepens and a thicker mantle wedge is present. represent only the thickness of the Solimões Formation within the However the entire region is void of contemporary arc-volcanism basins not the overall basin depth (i.e. the Solimões Formation and so the fate of slab volatiles that would be required to initi- overlies older but less spatially extensive Cretaceous and Paleozoic ate melt production and lower the wedge viscosity is very much deposits that are not included). unclear. As well as producing the isopach map the geological field stud- ies indicate that the Solimões unit consists mostly of fine-grained 3.3. Estimation of flexure inter-bedded claystones, siltstones and sandstones (Latrubesse et al., 2010; Rodrigues Nogueira et al., 2013). Interpretations of the Dynamic topography is typically much broader and flatter than paleo-environmental history of this sedimentary record are widely tectonically driven flexure. To demonstrate this, we also calculated varying and extensively debated, ranging from marine influenced flexure profiles for the region (Fig. 3 and Fig. 4), based on a bal- (Hoorn, 1993, 1994; Räsänen et al., 1995; Roddaz et al., 2005), to anced cross-section from Gotberg et al. (2010), which was the best mega-wetland or lacustrine dominated (Wesselingh et al., 2002; and most detailed section available. This cross-section covers only Campbell et al., 2006; Hoorn et al., 2010a, 2010b), to fluvial dom- the eastern part of the cordillera but models show that it is the inated (Latrubesse et al., 2007, 2010; Silva-Caminha et al., 2010; loads which are the closest, up to 200 km from the basin, that have Gross et al., 2011; Rodrigues Nogueira et al., 2013). the most influence on the flexural subsidence (Garcia-Castellanos Given its large extent and widespread coverage the Solimões et al., 2002). We therefore use the cross-section from Gotberg et Formation is therefore a good observational proxy with which to al. (2010) to estimate the extent of crustal shortening and thus ge- understand the subsidence history of the region, and thus to com- ometry of the Andean load that is adjacent to the foredeep. The pare with our dynamic topography and flexural models (Fig. 3). flexure generated by this load depends on the lithosphere rigidity It is important to notice that this sedimentary accumulation ex- i.e. effective elastic thickness (Te) of the plate, in addition to other tends beyond the zone of flexural subsidence expected from the rheological parameters (e.g. Watts, 2001). We used two different cordilleran loading (<300 km) and cannot therefore be easily ex- Te estimates, 25 km and 40 km (cf. Pérez-Gussinyé et al., 2007; plained by purely tectonic processes. The basin geometry, specif- Tassara et al., 2007). These predict 2.8 to 3.6 km of sedimentary ically the location of the depocenter and maximum thickness 256 C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260

Fig. 4. Diagrams illustrating the sequence of steps taken to estimate the flexural component of topography in the region. Firstly, in figures (a) through to (c), the extent of crustal shortening is estimated, and the geometry of the orogenic load reconstructed, based on a balanced cross-section from Gotberg et al. (2010). Then in figure (d) the flexural curves (two green lines) are calculated based on this loading for two different effective elastic thicknesses (Te). The sedimentary accommodation space generated next to the load (i.e. the predicted foredeep), and spanning around 100 km across, is highlighted by the gray shaded triangle. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

(∼1 km), does however correlate remarkably well to the dynamic dynamic subsidence propagating eastward over time tracking the subsidence driven by the flat-slab morphology (Fig. 3 and Fig. 5c). migration of the subduction leading edge. It is also important to emphasize that the development of the Peruvian flat slab and deposition of the Solimões Forma- 4.2. Implications for the evolution of western Amazonia tion are contemporaneous, both having occurred from the Mid to Late Miocene and onwards (Hampel, 2002; Bissig et al., 2008; Based on the above observations, we therefore propose that Ramos and Folguera, 2009; Hoorn, 1993; Wesselingh et al., 2006b; long wavelength dynamic topography associated with the transi- Latrubesse et al., 2010; Gross et al., 2011). In agreement with the tion from normal to flat-slab subduction beneath Peru, as well as flat subduction dynamics in Peru, there are also indications that flexural subsidence in the proximal foreland basins, were both im- the Solimões Formation gets younger towards the east. For ex- portant factors in the widespread subsidence and deposition of the ample the formation is solely Upper Miocene in the Acre Basin Solimões Formation across western Amazonia. Considering the in- to the west of the Iquitos Arch (Gross et al., 2011), but ex- teraction of both tectonic and dynamic controls on topography, and tends into the Pliocene in the Solimões Basin to the east of with observations from the geological record, we hypothesize the this arch (Silva-Caminha et al., 2010; Rodrigues Nogueira et al., following model for the evolution of the Amazonian landscape, as 2013)(Fig. 5). Additionally it has been more precisely dated by illustrated in Fig. 5. Latrubesse et al. (2010) that the Solimões Formation only propa- Initially during the Early Miocene (Fig. 5a–b), before the Nazca gated east of the Iquitos Arch after 9–6 Ma. This eastward young- Ridge arrived, subduction was steeper beneath Peru compared to ing trend is consistent with our prediction that a wave of dynamic the present day. The Andes were beginning to grow and the conti- subsidence propagated out east from the Andes as slab flattening nent was deforming via crustal shortening and flexural processes. advanced in both space and time (Fig. 3). The difference in the At longer wavelengths there was also dynamic subsidence be- age of the deposits across the Iquitos Arch, from older in the west neath the Andes from mantle flow induced by the dipping slab. to younger in the east, can therefore be easily explained by the The flexural and dynamic basin would have slowly filled with transient dynamic signal from the formation of the flat-slab, with sediment, forming the first and oldest deposits of the Solimões C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 257

Fig. 5. Schematic illustrations demonstrating how a change in subduction style could have driven the topographic and sedimentary evolution of western Amazonia since the Miocene, eventually shaping the present day landscape. Figures (a) and (c) represent maps of the predicted topography from our flexural and dynamic calculations both before and with the arrival of the flat slab. The scale bar for both maps is plotted below figure (a). Turquoise arrows represent drainage directions. The thin dotted line on figures (a) and (c) represents the approximate location of cartoon cross-sections figures (b) and (d) respectively. Gray dotted regions indicate sedimentary infill, with region 1 indicating the oldest deposits and 3 the youngest. The letters I and P on the cross-sections mark the estimated locations of the Iquitos and Purus Arches. Figure (e) is a similar cross-section extending from the trench to the Atlantic Ocean, showing the present day configuration of drainage and the location of the sedimentary deposits that preserve the record of landscape evolution hypothesized in figures (b) and (d). Formation (gray shaded region in Fig. 5b, and labeled as 1 in Later as flat-slab subduction took hold during the Late Miocene Fig. 5d). At this stage, the Northern Cordillera was not fully es- (Fig. 5c–d), plate coupling and tectonic deformation intensified tablished and drainage portals allowed access to the Pacific Ocean along the northern Andes. Consequently the northern Cordillera and/or the Caribbean Sea (turquoise arrows in Fig. 5a). underwent a period of rapid growth, increasing the relief, orogenic 258 C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 loading, erosional rates and sediment run-off, as well as trigger- edge, and that the transition from normal to flat subduction in- ing a further deepening and broadening of the flexural basins, duces uplift over the flat-slab region itself. especially across the foredeep (Jordan and Gardeweg, 1989; Kroon- Characterizing patterns of dynamic topography accurately is im- enberg et al., 1990; Hoorn et al., 1995, 2010a, 2010b; Nobret et al., portant if we want to understand how and why basins, as well 1996; Garzione et al., 2006; Roddaz et al., 2010). Additional accom- as the overall landscape, have evolved over time. Our results and modation space for the Solimões Formation was created from both their agreement with the local geology imply that the influence of this flexural subsidence driven by orogenic deformation, as well as the 3D details of subduction on the dynamics, such as along strike the propagation of longer wavelength dynamic subsidence into the variations in morphology, matter greatly to this goal. In the future, distal foreland. Whilst the tectonic signal has the primary control our ability to correctly predict dynamic topography will therefore on topography dominating where thick sedimentary sequences are improve with better knowledge of subduction zone structure. In preserved close to the Andean Cordillera (i.e. several 1000’s of kilo- particular better constraints on slab shape and density, potentially metres thick within the foredeep) (Wanderley-Filho et al., 2010), from seismic and gravitational studies, as well as more information only the dynamic signal can explain the broad spatial distribution on rheological properties, such as lateral variations in viscosity be- of the Solimões Formation (Fig. 5c). tween the slab and the wedge, have the potential to refine the It has been proposed that as uplift progressed, an orogenic bar- expression of flat-slab subduction on the Earth’s surface both in rier was formed, closing off the drainage pathways to the Pacific the present and in the past. Ocean or Caribbean Sea (Hoorn et al., 1995, 2010b; Hungerbühler et al., 2002; Campbell et al., 2006) and driving an eastward sedi- Acknowledgements ment flux. Drainage pathways to the east however were partially blocked by the Purus Arch (e.g. Figueiredo et al., 2009). Sediment The 7th Framework Program, Marie Curie Actions (ERC), sup- eroded and transported from the growing Andes was therefore ports our ANDyN project 251954 (Andes–Nazca Experiment) in carried to the western Amazonian lowlands and trapped in con- the Andes. We also appreciated funds and support by SECyT-UNC, tinental basins that had been formed by flexural and dynamic CONICET, and FONCyT from Argentina, MAPS faculty funds at Uni- subsidence (e.g. the Acre and Solimões Basins) (Fig. 5c–d). These versity College London, and NERC grant NE/J024813/1. We thank basins would have rapidly filled with the large influx of sediment, two anonymous reviewers for suggestions that greatly improved transitioning from a possibly (mega-) wetland or lacustrine low- the manuscript. energy low-circulation environment (similar to water filling up a bathtub), to more fluvial domains as the basin became full, al- Appendix A. Supplementary material lowing for sediment to be transported by rivers across the region rather then being trapped in the depression. Such a paleoenviron- Supplementary material related to this article can be found on- ment evolution is well represented by the depositional sequence line at http://dx.doi.org/10.1016/j.epsl.2014.07.027. of the Solimões Formation and change in sedimentary facies from muddy (low energy conditions interpreted as wetlands) to sandy References (indicative of higher energy conditions, potentially fluvial) over time (Hoorn et al., 2010b; Figueiredo, 2012). Aleman, A.M., 2006. The Peruvian flat-slab. In: Geological Society of America Ab- Once the basin became overfilled, drainage would have under- stracts with Programs, Speciality Meeting, No. 2, p. 21. gone reorganization and it is possible that at this stage the eastern Amante, C., Eakins, B.W., 2009. ETOPO1 1 arc-minute global relief model: proce- rim of the Solimões Basin was breached (e.g. Campbell et al., 2006) dures, data sources and analysis. NOAA Technical Memorandum NESDIS NGDC- connecting western Amazonia and the Andes to the Atlantic Ocean, 24, 19 pp. Arrial, P.A., Billen, M.I., 2013. Influence of geometry and eclogitization on oceanic and thus establishing the transcontinental Amazon River like we plateau subduction. Earth Planet. Sci. Lett. 363, 34–43. see today (Fig. 5e). Large-scale deposition in western Amazonia Billen, M.I., 2008. Modeling the dynamics of subducting slabs. Annu. Rev. Earth would have ceased once sediment was carried in vast quantities to Planet. Sci. 36, 325–356. the Atlantic Ocean. This is corroborated by an order of magnitude Billen, M.I., Gurnis, M., 2001. A low viscosity wedge in subduction zones. Earth increase in the mass accumulation rate of terrigeneous material on Planet. Sci. Lett. 193, 227–236. Billen, M.I., Gurnis, M., Simons, M., 2003. Multiscale dynamic models of the Tonga– the Amazon offshore fan beginning ∼5 Ma (Dobson et al., 2001), at Kermadec subduction zone. Geophys. J. Int. 153, 359–388. the same time that deposition of the Solimões Formation in west- Bird, P., 1988. Formation of the Rocky mountains, western United States – a contin- ern Amazonia ceased (e.g. Latrubesse et al., 2010), indicating that uum computer model. Science 239 (4847), 1501–1507. deposition of the bulk of Andean derived sediment switched from Bissig, T., Ullrich, T.D., Tosdal, R.M., Friedman, R., Ebert, S., 2008. The time–space dis- the Amazonian lowlands to the Atlantic Ocean at this time. tribution of Eocene to Miocene magmatism in the central Peruvian polymetallic province and its metallogenetic implications. J. South Am. Earth Sci. 26, 16–35. Braun, J., 2010. The many surface expressions of mantle dynamics. Nat. Geosci. 3 5. Conclusion (12), 825–833. Burgess, P.M., Gurnis, M., Moresi, L., 1997. Formation of sequences in the cratonic Considering the Peruvian flat-slab morphology in detail pre- interior of North America by interaction between mantle, eustatic, and strati- graphic processes. Geol. Soc. Am. Bull. 109 (12), 1515–1535. dicts broad dynamic subsidence that is significantly offset from the Cahill, T., Isacks, B.L., 1992. Seismicity and shape of the subducted Nazca Plate. trench, by well over 1000 km inland. This is in agreement with J. Geophys. Res. 97 (B12), 17503–17529. the geological record of deep sedimentary accumulations (up to Campbell Jr., K.E., Frailey, C.D., Romero Pitman, L., 2006. The Pan-Amazonian Ucay- 1 km thick) across western Amazonia, which lie too far from the ali peneplain, Late Neogene sedimentation in Amazonia, and the birth of the orogenic front to be explained by flexure alone. This has direct modern Amazon River system. Palaeogeogr. Palaeoclimatol. Palaeoecol. 239, 166–219. relevance for ancient episodes of flat-slab subduction, such as the Capitanio, F.A., Faccenda, M., 2012. Complex mantle flow around heterogeneous sub- Cretaceous-Early Tertiary , when dynamic topog- ducting oceanic plates. Earth Planet. Sci. Lett. 353, 29–37. raphy has been called upon to explain sedimentary basins and ma- Cembrano, J., Lavenu, A., Yañez, G., Riquelme, R., Garcia, M., González, G., Herail, G., rine transgressions within the western US interior (Liu et al., 2008; 2007. Neotectonics. In: Moreno, T., Gibbons, W. (Eds.), The Geology of Chile. The Mitrovica et al., 1989; Liu and Nummedal, 2004). Modeling and ob- Geological Society, London, pp. 147–178. Cozzuol, M., 2006. The Acre vertebrate fauna: diversity, geography and time. J. South servations across the Peruvian foreland reinforce the notion (Dávila Am. Earth Sci. 21, 185–203. and Lithgow-Bertelloni, 2013; Jadamec et al., 2013) that slab flat- Dávila, F.M., Lithgow-Bertelloni, C., 2013. Dynamic topography in South America. tening produces subsidence far inland at the subduction leading J. South Am. Earth Sci. 43, 127–144. C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260 259

Dávila, F.M., Lithgow-Bertelloni, C., Giménez, M., 2010. Tectonic and dynamic con- Hermoza, W., Brusset, S., Baby, P., Gil, W., Roddaz, M., Guerrero, N., Bolaños, R., 2005. trols on the topography and subsidence of the Argentine Pampas: the role of The Huallaga evolution: thrust propagation in a deltaic environ- the flat slab. Earth Planet. Sci. Lett. 295 (1), 187–194. ment, northern Peruvian Andes. J. South Am. Earth Sci. 19 (1), 21–34. Del’Arco, J.O., Santos, R.O.B., Rivetti, M., Olivera Alves, E.D., Fernandes, C.A.C., Silva, Hines, J.M., Billen, M.I., 2012. Sensitivity of the short-to intermediate-wavelength L.L., 1977. Folha SB. 19 Juruá. IeGeologia. Projeto Radambrasil. Levantamento de geoid to rheologic structure in subduction zones. J. Geophys. Res., Solid Recursos Naturais 15, 19–88. Earth 117, B05410. http://dx.doi.org/10.1029/2011JB008978. Dickinson, W., Snyder, W., 1978. of the Laramide orogeny. In: Geol. Hoorn, C., 1993. Marine incursions and the influence of Andean tectonics on the Soc. Am. Mem., vol. 151, pp. 355–366. Miocene depositional history of northwestern Amazonia: results of a palynos- Dobson, D.M., Dickens, G.R., Rea, D.K., 2001. Terrigenous sediment on Ceara Rise: a tratigraphic study. Palaeogeogr. Palaeoclimatol. Palaeoecol. 105, 267–309. Cenozoic record of South American orogeny and erosion. Palaeogeogr. Palaeocli- Hoorn, C., 1994. An environmental reconstruction of the palaeo-Amazon River matol. Palaeoecol. 165, 215–229. system (Middle-Late Miocene, NW Amazonia). Palaeogeogr. Palaeoclimatol. Eakin, C.M., Long, M.D., 2013. Complex anisotropy beneath the Peruvian flat slab Palaeoecol. 112 (3), 187–238. from frequency-dependent, multiple-phase shear wave splitting analysis. J. Geo- Hoorn, C., 2006. The birth of the mighty Amazon. Sci. Am. 294 (5), 52–59. phys. Res., Solid Earth 118. http://dx.doi.org/10.1002/jgrb50349. Hoorn, C., Guerrero, J., Sarmiento, G.A., Lorente, M.A., 1995. Andean tectonics as a Eakin, C.M., Long, M.D., Beck, S.L., Wagner, L.S., Tavera, H., 2013. Characterizing cause for changing drainage patterns in Miocene northern South America. Geol- seismic anisotropy across the Peruvian flat-slab subduction zone: shear wave ogy 23 (3), 237–240. splitting from PULSE. In: AGU Fall Meeting 2013. San Francisco, CA. Hoorn, C., Wesselingh, F.P., Ter Steege, H., Bermudez, M.A., Mora, A., Sevink, J., An- Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna, C., Brus- tonelli, A., 2010a. Amazonia through time: Andean uplift, climate change, land- set, S., 2008. Flat subduction dynamics and deformation of the South American scape evolution, and biodiversity. Science 330 (6006), 927–931. plate: insights from analog modeling. Tectonics 27, TC3011. http://dx.doi.org/ Hoorn, C., Wesselingh, F.P., Hovikoski, J., Guerrero, J., 2010b. The development of the 10.1029/2007TC002175. Amazonian mega-wetland (Miocene; Brazil, Colombia, Peru, Bolivia). In: Hoorn, Figueiredo, J.J.P., 2012. Comment by JP Figueiredo, and Hoorn, C. on ‘Late Miocene C., Wesselingh, F.P. (Eds.), Amazonia, Landscape and Species Evolution: A Look sedimentary environments in south-western Amazonia (Solimões Formation; into the Past. Wiley-Blackwell, Chichester, pp. 123–142. Brazil)’ by Martin Gross, Werner E. Piller, Maria Ines Ramos, Jackson Douglas Hungerbühler, D., Steinmann, J., Winkler, W., Seward, D., Egüez, A., Peterson, D.E., da Silva Paz. J. South Am. Earth Sci. 35, 74–75. Helg, U., Hammer, C., 2002. Neogene stratigraphy and Andean of Figueiredo, J., Hoorn, C., van der Ven, P., Soares, E., 2009. Late Miocene onset of southern . Earth-Sci. Rev. 57, 75–124. the Amazon River and the Amazon deep-sea fan: evidence from the Foz do Jadamec, M.A., Billen, M.I., 2010. Reconciling surface plate motions with rapid three- Amazonas Basin. Geology 37 (7), 619–622. dimensional mantle flow around a slab edge. Nature 465 (7296), 338–341. Gans, C., Beck, S., Zandt, G., Gilbert, H., Alvarado, P., 2011. Continental and oceanic Jadamec, M.A., Billen, M.I., 2012. The role of rheology and slab shape on rapid crustal structure of the Pampean at slab region, western Argentina, using re- mantle flow: three-dimensional numerical models of the Alaska slab edge. J. ceiver function analysis: new high-resolution results. Geophys. J. Int. 186, 45–58. Geophys. Res., Solid Earth 117, B02304. http://dx.doi.org/10.1029/2011JB008563. Garcia-Castellanos, D., Fernandez, M., Torné, M., 2002. Modeling the evolution Jadamec, M.A., Billen, M.I., Kreylos, O., 2012. Three-dimensional simulations of ge- of the Guadalquivir foreland basin (southern Spain). Tectonics 21 (3). http:// ometrically complex subduction with large viscosity variations. In: ACM Pro- dx.doi.org/10.1029/2001TC001339. ceedings of the 2012 Conference on eXtreme Science and Engineering Discovery Garzione, C.M., Molnar, P., Libarkin, J.C., Mc Fadden, B.J., 2006. Rapid late Miocene Environment. rise of the Bolivian Altiplano: evidence for removal of mantle lithosphere. Earth Jadamec, M.A., Billen, M.I., Roeske, S.M., 2013. Three-dimensional numerical models Planet. Sci. Lett. 241, 543–556. of flat slab subduction and the Denali driving deformation in south-central Gotberg, N., McQuarrie, N., Carlotto Caillaux, V., 2010. Comparison of crustal thick- Alaska. Earth Planet. Sci. Lett. 376, 29–42. ◦ ening budget and shortening estimates in southern Peru (12–14 S): implica- Jordan, T.E., Gardeweg, P.M., 1989. Tectonic evolution of the late Cenozoic central ◦ ◦ tions for mass balance and rotations in the “Bolivian orocline”. Geol. Soc. Am. Andes (20 –33 S). In: Ben-Avraham, Z. (Ed.), The Evolution of the Pacific Ocean Bull. 122 (5–6), 727–742. http://dx.doi.org/10.1130/B26477.1. Margins. Oxford University Press, New York, pp. 193–207. Grand, S.P., 2002. Mantle shear wave tomography and the fate of subducted slabs. King, S.D., Hager, B.H., 1994. Subducted slabs and the geoid: 1. Numerical ex- Philos. Trans. R. Soc. Lond. A 360, 2475–2491. periments with temperature-dependent viscosity. J. Geophys. Res. 99 (B10), Gripp, A.E., Gordon, R.G., 2002. Young tracks of hotspots and current plate velocities. 19843–19852. Geophys. J. Int. 150 (2), 321–361. Knezevic Antonijevic, S., Wagner, L.S., Beck, S.L., Zandt, G., Long, M.D., 2013. Ex- Gross, M., Piller, W.E., Ramos, M.I., da Silva Paz, J.D., 2011. Late Miocene sedimentary posing the Peruvian flat slab from shear wave velocity structure. In: AGU Fall environments in south-western Amazonia (Solimões Formation; Brazil). J. South Meeting 2013. San Francisco, CA. Am. Earth Sci. 32, 169–181. Kroonenberg, S.B., Bakker, J.G.M., van der Weil, A.M., 1990. Late Cenozoic uplift Guillaume, B., Martinod, J., Husson, L., Roddaz, M., Riquelme, R., 2009. Neogene up- and paleogeography of the Colombian Andes: constraints in the development lift of central eastern Patagonia: dynamic response to active spreading ridge of high-Andean biota. Geol. Mijnb. 69, 279–290. subduction? Tectonics 28, TC2009. http://dx.doi.org/10.1029/2008TC002324. Latrubesse, E.M., Bocquentin, J., Santos, C.R., Ramonell, C.G., 1997. Paleoenviron- Gurnis, M., Müller, R.D., Moresi, L., 1998. Cretaceous vertical motion of Australia and mental model for the late Cenozoic southwestern Amazonia paleontology and the Australian Antarctic discordance. Science 279, 1499–1504. geology. Acta Amazonica 27, 103–118. Gutscher, M.A., Maury, R., Eissen, J.P., Bourdon, E., 2000a. Can slab melting be caused Latrubesse, E., Silva, S.F., Cozzuol, M., Absy, M.L., 2007. Late Miocene continental by flat subduction? Geology 28, 535–538. sedimentation in the southwestern Amazonia and its regional significance: bi- Gutscher, M.A., Spakman, W., Bijwaard, H., Engdahl, E.R., 2000b. Geodynamics of flat otic and geological evidence. J. South Am. Earth Sci. 23, 61–80. subduction: seismicity and tomographic constraints from the Andean margin. Latrubesse, E.M., Cozzuol, M., da Silva-Caminha, S.A.F., Rigsby, C.A., Absy, M.L., Tectonics 19 (5), 814–833. Jaramillo, C., 2010. The Late Miocene paleogeography of the Amazon Basin and Hager, B.H., 1984. Subducted slabs and the geoid: constraints on mantle rheology the evolution of the Amazon River system. Earth-Sci. Rev. 99 (3), 99–124. and flow. J. Geophys. Res. 89 (B7), 6003–6015. Li, C., van der Hilst, R.D., Engdahl, E.R., Burdick, S., 2008. A new global model for P Hager, B.H., O’Connell, R.J., 1979. Kinematic models of large-scale flow in the Earth’s wave speed variations in Earth’s mantle. Geochem. Geophys. Geosyst. 9 (5). mantle. J. Geophys. Res., Solid Earth 84 (B3), 1031–1048. Lithgow-Bertelloni, C., Gurnis, M., 1997. Cenozoic subsidence and uplift of continents Hager, B.H., O’Connell, R.J., 1981. A simple global model of plate dynamics and man- from time-varying dynamic topography. Geology 25 (8), 735–738. tle convection. J. Geophys. Res., Solid Earth 8 (B6), 4843–4867. Lithgow-Bertelloni, C., Richards, M.A., 1998. The dynamics of Cenozoic and Mesozoic Hager, B.H., Richards, M.A., 1989. Long-wavelength variations in Earth’s geoid: phys- plate motions. Rev. Geophys. 36 (1), 27–78. ical models and dynamical implications. Philos. Trans. R. Soc. Lond. Ser. A, Math. Lithgow-Bertelloni, C., Silver, P.G., 1998. Dynamic topography, plate driving forces Phys. Sci. 328 (1599), 309–327. and the African superswell. Nature 395, 269–272. Hager, B.H., Clayton, R.W., Richards, M.A., Comer, R.P., Dziewonski, A.M., 1985. Lower Liu, S., Nummedal, D., 2004. Late Cretaceous subsidence in Wyoming: quantifying mantle heterogeneity, dynamic topography and the geoid. Nature 313, 541–545. the dynamic component. Geology 32 (5), 397–400. Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian ac- Liu, L., Spasojevic,´ S., Gurnis, M., 2008. Reconstructing Farallon plate subduction be- tive margin: a re-evaluation. Earth Planet. Sci. Lett. 203, 665–679. neath North America back to the Late Cretaceous. Science 322 (5903), 934–938. Harrison, C.G.A., Miskell, K.J., Brass, G.W., Saltzman, E.S., Sloan, J.L., 1983. Continental Liu, S., Nummedal, D., Liu, L., 2011. Migration of dynamic subsidence across the Late hypsography. Tectonics 2 (4), 357–377. Cretaceous United States Western Interior Basin in response to Farallon plate Heine, C., Muller, R., Steinberger, B., Torsvik, T., 2008. Subsidence in intracontinental subduction. Geology 39 (6), 555–558. basins due to dynamic topography. Phys. Earth Planet. Inter. 171, 252–264. Lonsdale, P., 2005. Creation of the Cocos and Nazca plates by fission of the Farallon Heine, C., Müller, R.D., Steinberger, B., DiCaprio, L., 2010. Integrating deep Earth dy- plate. Tectonophysics 404 (3–4), 237–264. namics in paleogeographic reconstructions of Australia. Tectonophysics 483 (1), Maia, R.G., Godoy, H.K., Yamaguti, H.S., Moura, P.D., Costa, F.D., Holanda, M.D., Costa, 135–150. J., 1977. Costa Projeto de Carvao no Alto Solimões. Relatorio Final. Companhia Henry, S.G., Pollack, H.N., 1988. Terrestrial heat flow above the Andean subduction de Pesquisa de Recursos Minerais–Departamento Nacional da Produçáo Mineral, zone in Bolivia and Peru. J. Geophys. Res. 93 (B12), 15153–15162. Manaus, p. 137. 260 C.M. Eakin et al. / Earth and Planetary Science Letters 404 (2014) 250–260

Martinod, J., Husson, L., Roperch, P., Guillaume, B., Espurt, N., 2010. Horizontal sub- tle in the mantle transition zone beneath the Central Andes. In: AGU Fall Meet- duction zones, convergence velocity and the building of the Andes. Earth Planet. ing 2013, San Francisco, CA. Sci. Lett. 299 (3), 299–309. Shephard, G.E., Müller, R.D., Liu, L., Gurnis, M., 2010. Miocene drainage reversal Mégard, F., 1987. Cordilleran Andes and Marginal Andes: a review of Andean geol- of the Amazon River driven by plate–mantle interaction. Nat. Geosci. 3 (12), ◦ ogy north of the Arica Elbow (18 S). In: Monger, J.W.H., Francheteau, J. (Eds.), 870–875. Circum-Pacific Orogenic Belts and Evolution of the Pacific Ocean Basin. In: Shephard, G.E., Liu, L., Müller, R.D., Gurnis, M., 2012. Dynamic topography and Geodynamic Series, vol. 18. American Geophysical Union, Geological Society of anomalously negative residual depth of the Argentine Basin. Gondwana Res. 22 America, pp. 71–95. (2), 658–663. Mitrovica, J.X., Beaumont, C., Jarvis, G.T., 1989. Tilting of the continental interior by Siebert, L., Simkin, T., 2002. Volcanoes of the World: an illustrated catalog of the dynamical effects of subduction. Tectonics 8 (5), 1079–1094. holocene volcanoes and their eruptions: Smithsonian institution, global vol- Mora, A., Baby, P., Roddaz, M., Parra, M., Brusset, S., Hermoza, W., Espurt, N., canism program digital information series. GVP-3, http://www.volcano.si.edu/ 2010. Tectonic history of the Andes and sub-Andean zones: implications for world/. the development of the Amazon Drainage Basin. In: Hoorn, C., Wesselingh, F.P. Silva-Caminha, S.A.F., Jaramillo, C., Absy, M.L., 2010. Neogene palynology of the (Eds.), Amazonia, Landscape and Species Evolution: A Look into the Past. Wiley- Solimões Basin, Brazilian Amazonia. Palaeontographica 283 (1–3), 1–67. Blackwell, Chichester, pp. 38–60. Soler, P., Bonhomme, M., 1990. Relations of magmatic activity to plate dynamics Moresi, L., Gurnis, M., 1996. Constraints on the lateral strength of slabs from three- in central Peru from late cretaceous to present. In: Kay, S., Rapela, C. (Eds.), dimensional dynamic flow models. Earth Planet. Sci. Lett. 138, 15–28. Plutonism from Antarctica to Alaska. In: Geol. Soc. Am. Mem., pp. 173–191. Moucha, R., Forte, A., Mitrovica, J., Daradich, A., 2007. Lateral variations in man- Spasojevic, S., Liu, L., Gurnis, M., 2009. Adjoint models of mantle convection with tle rheology: implications for convection related surface observables and in- seismic, plate motion, and stratigraphic constraints: North America since the ferred viscosity models. Geophys. J. Int. 169, 113–135. http://dx.doi.org/10.1111/ Late Cretaceous. Geochem. Geophys. Geosyst. 10, Q05W02. j.1365-246X.2006.03225.x. Stadler, G., Gurnis, M., Burstedde, C., Wilcox, L.C., Alisic, L., Ghattas, O., 2010. The Müller, R.D., Sdrolias, M., Gaina, C., Roest, W.R., 2008. Age, spreading rates, and dynamics of plate tectonics and mantle flow: from local to global scales. Sci- spreading asymmetry of the world’s ocean crust. Geochem. Geophys. Geosyst. 9 ence 329 (5995), 1033–1038. (4), 18–36. Tassara, A., Swain, C., Hackney, R., Kirby, J., 2007. Elastic thickness structure of Nobret, V., Lavenu, A., Marocco, R., 1996. Concept of continuum as opposed to peri- South America estimated using wavelets and satellite-derived gravity data. Earth odic tectonism in the Andes. Tectonophysics 255, 65–78. Planet. Sci. Lett. 253 (1), 17–36. Olson, P., Nam, S., 1986. Formation of seafloor swells by mantle plumes. J. Geophys. van Hunen, J., van den Berg, A.P., Vlaar, N.J., 2002a. On the role of subduct- Res. 91 (B7), 7181–7191. ing oceanic plateaus in the development of shallow flat subduction. Tectono- Pérez-Gussinyé, M., Lowry, A.R., Watts, A.B., 2007. Effective elastic thickness of South physics 352, 317–333. http://dx.doi.org/10.1016/S00401951(02)00263-9. America and its implications for intracontinental deformation. Geochem. Geo- van Hunen, J., van den Berg, A.P., Vlaar, N.J., 2002b. The impact of the South phys. Geosyst. 8 (5), 1–22. http://dx.doi.org/10.1029/2006GC001511. American plate motion and the Nazca Ridge subduction on the flat subduc- Phillips, K., Clayton, R.W., 2014. Structure of the subduction transition region from tion below South Peru. Geophys. Res. Lett. 29 (14). http://dx.doi.org/10.1029/ seismic array data in southern Peru. Geophys. J. Int. 196, 1889–1905. http://dx. 2001GL014004. doi.org/10.1093/gji/ggt504. Wanderley-Filho, J.R., Costa, J.B.S., 1991. Contribuição à evolução estrutural de bacia Pilger, R.H., 1981. Plate reconstructions, aseismic ridges, and low-angle subduction do Amazonas e sua relação com o embasamento. In: Simpósio de Geologica de beneath the Andes. Geol. Soc. Am. Bull. 92, 338–456. Amazônia, vol. 3. Sociedade Brasileira de Geologia, Anais. Belém, pp. 244–259. Potter, P.E., 1997. The mesozoic and cenozoic drainage of South America: a natural Wanderley-Filho, J.R., Eiras, J.F., Cunha, P.R.C., Van der Vem, P.H., 2010. The Paleozoic history. J. South Am. Earth Sci. 10, 331–344. Pysklywec, R., Mitrovica, J., 1999. The role of subduction-induced subsidence in the Solimões and Amazonas basins and the Acre foreland basin of Brazil. In: Hoorn, evolution of the Karoo Basin. J. Geol. 107, 155–164. C., Wesselingh, F.P. (Eds.), Amazonia, Landscape and Species Evolution: A Look Ramos, V.A., Folguera, A., 2009. Andean flat-slab subduction through time. In: Mur- into the Past. Wiley-Blackwell, Chichester, pp. 9–28. phy, J.B., Keppie, J.D., Hynes, A.J. (Eds.), Ancient Orogens and Modern Analogues. Watts, A.B., 2001. Isostasy and Flexure of the Lithosphere. Cambridge University In: Geological Society London Special Publication, vol. 327, pp. 31–54. Press, Cambridge. Räsänen, M., Linna, A.M., Santos, J.C.R., Negri, F.R., 1995. Late Miocene tidal deposits Wessel, P., Smith, W.H.F., 1991. Free software helps map and display data. Eos 72 in the Amazonian foreland basin. Science 269, 386–389. (41), 441–446. Ricard, Y., Richards, M., Lithgow-Bertelloni, C., LeStunff, Y., 1993. A geodynamical Wesselingh, F.P., Räsänen, M.E., Irion, G., Vonhof, H.B., Kaandorp, R., Renema, W., model of mantle density heterogeneity. J. Geophys. Res. 98 (B12), 21895–21909. Romero Pittman, L., Gingras, M., 2002. Lake Pebas: a palaeoecological recon- Roddaz, M., Viers, J., Brusset, S., Baby, P., Hérail, G., 2005. Sediment provenances and struction of a Miocene, long-lived lake complex in western Amazonia. Cainozoic drainage evolution of the Neogene Amazonian foreland basin. Earth Planet. Sci. Res. 1, 35–81. Lett. 239 (1), 57–78. Wesselingh, F.P., Ranzy, A., Rasanen, M., 2006a. Miocene freshwater Mollusca from Roddaz, M., Hermoza, W., Mora, A., Baby, P., Parra, M., Christophoul, F., Espurt, N., western Brazilian Amazonia. Scr. Geol. 133, 419–437. 2010. Cenozoic sedimentary evolution of the Amazonian foreland basin system. Wesselingh, F.P., Guerrero, J., Räsänen, M., Romero Pitmann, L., Vonhof, H., 2006b. In: Hoorn, C., Wesselingh, F.P. (Eds.), Amazonia, Landscape and Species Evolu- Landscape evolution and depositional processes in the Miocene Amazonian Pe- tion: A Look into the Past. Wiley-Blackwell, Chichester, pp. 61–88. bas lake/wetland system: evidence from exploratory boreholes in northeastern Rodrigues Nogueira, A.C., Silveira, R., Felix Guimarães, J.T., 2013. Neogene- Peru. Scr. Geol. 133, 323–361. Quaternary sedimentary and paleovegetation history of the eastern Solimões Zhong, S., Davies, G.F., 1999. Effects of plate and slab on the geoid. Earth Basin, central Amazon region. J. South Am. Earth Sci. 46, 89–99. Planet. Sci. Lett. 170, 487–496. Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., Duboz, C., 2005. Sub- Zhong, S., Gurnis, M., 1992. Viscous flow model of a subduction zone with a faulted duction of the Nazca Ridge and the Inca Plateau: insights into the formation of lithosphere: long and short wavelength topography, gravity and geoid. Geophys. ore deposits in Peru. Earth Planet. Sci. Lett. 239, 18–32. Res. Lett. 19 (18), 1891–1894. Scire, A.C., Biryol, C.B., Zandt, G., Beck, S.L., Wagner, L.S., Long, M.D., Minaya, E., Zhong, S., Gurnis, M., 1994. Controls on trench topography from dynamic models of Tavera, H., 2013. Tomographic imaging of the Nazca slab and surrounding man- subducted slabs. J. Geophys. Res. 99 (B8), 15683–15695.