You are on page 1of 16

Journal of South American Earth Sciences 94 (2019) 102209

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

Revised chronostratigraphic framework for the Cretaceous Magallanes- T


Austral Basin, Última Esperanza Province, Chile
Benjamin G. Danielsa,∗, Stephen M. Hubbarda, Brian W. Romansb, Matthew A. Malkowskic,
William A. Matthewsa, Anne Bernhardtd, Sebastian A. Kaempfeb, Zane R. Jobee, Julie C. Fosdickf,
Theresa M. Schwartze, Andrea Fildanig, Stephan A. Grahamc
a
Department of Geoscience, University of Calgary, 2500 University Drive NW, Calgary, AB, T2N1N4, Canada
b
Department of Geosciences, Virginia Tech, 926 West Campus Drive, Blacksburg, VA, 24061, USA
c
Department of Geological Sciences, Stanford University, 450 Serra Mall, Building 320, Stanford, CA, 94305, USA
d
Institute for Geological Sciences, Freie Universität Berlin, Malteserstrasse 74-100, Berlin, 12249, Germany
e
Department of Geology and Geological Engineering, Colorado School of Mines, 1516 Illinois Street, Golden, CO, 80401, USA
f
Center for Integrative Geosciences, University of Connecticut, 354 Mansfield Road, Storrs, CT, 06269-1045, USA
g
Equinor Research Center, 6300 Bridge Point Parkway, Building 2, Suite 100, Austin, TX, 78730, USA

ARTICLE INFO ABSTRACT

Keywords: The deposits of ancient sediment-routing systems in basins adjacent to the Andes offer key perspectives into the
Magallanes-Austral Basin geologic evolution of South America, and can provide insight into basin-evolution-controlling mechanisms that
Chronostratigraphy operate on time scales spanning millions of years. The Andes and associated basins of southernmost South
U-Pb geochronology America are important archives of paleogeographic evolution in Patagonia, a region that underwent significant
Strontium isotope stratigraphy
geologic and paleoenvironmental change during the Mesozoic and Cenozoic eras. Here, we present an updated
Basin analysis
chronostratigraphic framework for Upper Cretaceous units (Punta Barrosa, Cerro Toro, Tres Pasos, Dorotea
Patagonia
Formations) in the Magallanes-Austral retroarc foreland basin in Última Esperanza Province, Chile, to assist
workers with deciphering the geologic and paleoenvironmental evolution of southernmost South America. The
framework combines stratigraphic information from over 60 years of research in the basin with a large suite
(N = 51) of U-Pb volcanic ash and detrital zircon depositional ages, which includes new ages (N = 13) from
outcrop exposures near Laguna Sofia, Chile, as well as weighted mean strontium isotope depositional ages
(N = 3) from in-situ inoceramid shells. Results show that each formation contains stratigraphic architecture that
is related to distinct depositional environments, all of which can be temporally constrained using the available
data. The basin fill succession records deposition associated with: (1) largely unconfined submarine fan systems
(Punta Barrosa Formation; 101.0 ± 2.2 Ma to 89.5 ± 1.9 Ma); (2) abyssal plain systems characterized by
coarse-grained sediment starvation (lower Cerro Toro Formation; 89.5 ± 1.9 Ma to 84.7 ± 1.3 Ma); (3) con-
glomerate-rich deep-water channel-levee systems (upper Cerro Toro Formation; 84.7 ± 1.3 Ma to
80.5 ± 0.3 Ma); (4) a slope system characterized by widespread submarine mass-wasting (lower Tres Pasos
Formation; 80.5 ± 0.3 Ma to 77.4 ± 1.1 Ma); and (5) a shelf margin system characterized by shelf-slope
clinoform propagation (upper Tres Pasos and Dorotea Formations; 77.4 ± 1.1 Ma to 67.1 ± 1.8 Ma). Using the
temporal constraints, we provide recommendations for use of specific ages associated with each architectural
interval to facilitate future basin-analysis studies in the region.

1. Introduction 2017a). Stratigraphic information from basins adjacent to the Andes


can inform the mechanisms that control basin evolution and sediment
Andean foreland basins provide key insights into the evolution of routing over numerous time scales, including orogenic uplift (e.g.,
South America over deep geologic time (Jordan et al., 2001; DeCelles Bayona et al., 2008; Hoorn et al., 2010; Siks and Horton, 2011); tec-
and Horton, 2003; Ghiglione et al., 2010; Hoorn et al., 2010; Cuitiño tonically driven changes in basin subsidence (e.g., Romans et al., 2010,
et al., 2012; Di Giulio et al., 2012; Horton, 2017; Malkowski et al., 2011); eustatic sea-level change (e.g., Howell et al., 2005);


Corresponding author.
E-mail address: bgdaniel@ucalgary.ca (B.G. Daniels).

https://doi.org/10.1016/j.jsames.2019.05.025
Received 30 October 2018; Received in revised form 13 May 2019; Accepted 18 May 2019
Available online 23 May 2019
0895-9811/ © 2019 Elsevier Ltd. All rights reserved.
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

paleoclimate fluctuations (e.g., Do Campo et al., 2010; Varela et al., 2. Geologic background
2018); and autogenic controls, such as catchment reconfiguration via
prolonged weathering processes (e.g., Roddaz et al., 2006). Andean 2.1. Geologic history of the Cretaceous Magallanes-Austral Basin
foreland basins also host important Mesozoic and Cenozoic fossils (e.g.,
Pardo-Pérez et al., 2012; Rubilar-Rogers et al., 2012; Otero et al., 2015) Development of the Magallanes-Austral Basin is linked to Andean
that require independent chronostratigraphic control and paleoenvir- orogenesis and uplift in Chilean and Argentine Patagonia during the
onmental context. Furthermore, these basins are critical to the social Mesozoic. During the Jurassic, rifting and subduction at the southern
and economic development of nations on the margins of the Andean end of South America associated with the breakup of Gondwana led to
Cordillera, as they contain considerable hydrocarbon (e.g., Thomas, the development of a volcanic arc and the associated Rocas Verdes
1949a, 1949b; Mathalone and Montoya, 1995) and groundwater backarc basin (Dalziel et al., 1974; Suárez and Pettigrew, 1976; de Wit
(Raychowdhury et al., 2014) resources. and Stern, 1981; Olivero and Martinioni, 2001; Fildani and Hessler,
Stratigraphic successions in Andean foreland basins have been 2005; Calderón et al., 2007). Late Jurassic–earliest Cretaceous mag-
characterized extensively over the past several decades (e.g., the Chaco matism is reflected in mafic volcanic units of the Sarmiento, Capitán
foreland basin; Sempere et al., 1990; Husson and Moretti, 2002; Aracena, Carlos III, and Tortuga ophiolite complexes, as well as silicic
DeCelles and Horton, 2003; Uba et al., 2006; Hulka and Heubeck, volcanic units associated with the Tobífera, El Quemado, and Ibañez
2010). Recent work on Andean foreland basin successions has empha- Formations of western Patagonia (Figs. 1 and 2; Godoy, 1978; de Wit
sized chronostratigraphic characterization, with implications for pa- and Stern, 1981; Allen, 1982; Pankhurst et al., 2000; Calderón et al.,
leoenvironment reconstructions (e.g., Ruskin and Jordan, 2007), eva- 2007; Mpodozis et al., 2011; Calderón et al., 2016). Deposition in the
luation of sedimentation rates (e.g., Stevens Goddard and Carrapa, Rocas Verdes Basin is reflected in volcanic detritus associated with the
2018), and the relative importance of basin-evolution-controlling me- units listed above, as well as marine, siltstone-prone units preserved in
chanisms in various settings along the length of the orogen (e.g., the Río Mayer and Zapata Formations of Argentina and Chile, respec-
Horton, 2017; Sickmann et al., 2018). Though numerous workers have tively (Nullo et al., 1981; Fildani and Hessler, 2005; Richiano, 2014).
gleaned insightful perspectives into Andean foreland systems using site- Rocas Verdes Basin fill thicknesses range from 200 to 500 m in the
specific chronostratigraphic data (e.g., Fildani et al., 2003; DeCelles north to > 1000 m in the south (Biddle et al., 1986); southward in-
et al., 2007; Horton et al., 2015), thorough evaluation of the history of a creasing thickness trends are associated with a prolonged phase of ex-
basin requires chronostratigraphic data sets that are spatially and tension and crustal attenuation in the southern part of the basin relative
temporally exhaustive. Consideration of small data sets of depositional to the northern regions (Stern et al., 1992; Mukasa and Dalziel, 1996;
ages for individual stratigraphic units (e.g., Tunik et al., 2010; Siks and Stern and de Wit, 2003; Calderón et al., 2013; Malkowski et al., 2016).
Horton, 2011), combined with sparse spatial/temporal coverage of Increased spreading rates in the South Atlantic Ocean, coupled with
depositional age constraints for those units (e.g., Romans et al., 2010), elevated convergence and subduction on the western margin of South
may potentially confound interpretations of controls on basin evolution America during the Aptian-Albian, resulted in the transition from an
(e.g., the central Andes; Charrier et al., 2013). Insufficient or in- extensional to compressional setting in southernmost South America,
adequate depositional age constraints may also fundamentally hinder and led to the formation of the Magallanes-Austral retroarc foreland
accurate stratigraphic correlation of units that preserve numerous facies basin (Figs. 1 and 2; Bruhn and Dalziel, 1977; Rabinowitz and La
transitions across space and time (e.g., the foreland basins of Central Brecque, 1979; Dott et al., 1982; Dalziel, 1986; Mpodozis and Ramos,
Argentina; Folguera et al., 2015). In some instances, reliance on limited 1989; Ramos and Aguirre-Urreta, 1994; Fildani and Hessler, 2005). The
chronostratigraphic data to inform correlations has led to the devel- rapid uplift of the Andes during this time, coupled with an inherited
opment of multiple stratigraphic frameworks for a single Andean attenuated lithosphere, led to enhanced subsidence in the basin fore-
foreland basin (e.g., Arbe and Hechem, 1984; Varela et al., 2012, 2013), deep (Fildani and Hessler, 2005; Calderón et al., 2012; Fosdick et al.,
which has resulted in differing interpretations of the basin history 2014), and development of deep-water depositional environments
(Sickmann et al., 2018). Furthermore, depositional ages used in where water depths exceeded 1000 m (Natland et al., 1974; Hubbard
chronostratigraphic analysis require a level of precision that is appro- et al., 2010). Sediment routed to these environments was largely de-
priate for the time intervals being studied. As an example, while paly- rived from fold-thrust belt, arc, and other basin-peripheral sources to
nological data may assist with temporally constraining depositional the north and west (Fig. 1). Sediment sources include: (1) Paleozoic to
phases that take place over multiple geologic epochs (tens of millions of Early Mesozoic (> 200 Ma) metamorphic basement units (e.g., Duque
years), the age range (including its uncertainty) of a palynomorph may de York metamorphic complex; Fig. 1; Calderón et al., 2016); (2) Early
be inadequate for constraint of depositional phases that take place over to Middle Jurassic volcanic units north of Lago Viedma (e.g., Deseado
shorter periods of time (e.g., Bayona et al., 2008). Massif; Fig. 1; Pankhurst et al., 2000); (3) Late Jurassic to Early Cre-
To assist workers focused on deciphering the tectonic evolution and taceous mafic volcanic units of the Sarmiento, Capitán Aracena, Carlos
paleoenvironmental record of Andean foreland basins in southernmost III, and Tortuga ophiolite complexes, and silicic volcanic units of the
South America, we examine chronostratigraphic relationships for Upper Tobífera, El Quemado, and Ibañez Formations (Fig. 1; Calderón et al.,
Cretaceous units (Punta Barrosa, Cerro Toro, Tres Pasos, and Dorotea 2007, 2013); (4) Early Cretaceous backarc basin deposits of the Zapata
Formations) of the Magallanes-Austral foreland basin in Última Formation (Fig. 1; Calderón et al., 2007); and (5) the Early to Late
Esperanza Province, Chile (Fig. 1). Our analysis combines stratigraphic Cretaceous emergent Andean arc, which was episodically active over
data from over 60 years of research with a large database of U-Pb the entire depositional history of the basin (Fig. 1; Hervé et al., 2007).
detrital and volcanic ash zircon-derived depositional ages from in- Increased transfer of coarse clastic detritus to deep-water settings in
trabasinal units (N = 51), which includes newly reported ages (N = 13) the Magallanes-Austral Basin began at ∼115 Ma (Fosdick et al., 2011;
from units exposed near Laguna Sofia, Chile (Figs. 1 and 2). Additional Malkowski et al., 2017a, 2017b), and is reflected in a ∼1000 m thick
geochronologic constraint is provided by strontium isotope ages succession of intercalated sandstones and siltstones associated with the
(N = 3) from in-situ inoceramid shells, which are found within the Aptian-Turonian Punta Barrosa Formation in Chile and Argentina (Katz,
Cerro Toro Formation. The objectives of this study are to: (1) use the 1963; Arbe and Hechem, 1984; Wilson, 1991; Fildani et al., 2009;
depositional ages to construct an updated Upper Cretaceous chronos- Malkowski et al., 2017a, 2017b). The tabular-to-lenticular architecture
tratigraphic framework in Última Esperanza Province; and (2) evaluate observed in Punta Barrosa Formation sandstones led workers to suggest
the utility of the data for future studies on Magallanes-Austral Basin that these deposits accumulated in largely unconfined submarine fan
evolution, as well as related processes (e.g., development of sediment- systems in the basin foredeep (Fig. 2; Fildani et al., 2009; Malkowski
routing systems, paleofaunal histories). et al., 2017a, 2017b). Southward, basin-axial propagation of these fan

2
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Fig. 1. Geologic map of the modern Patagonian fold-thrust belt (modified from Fildani and Hessler, 2005; Romans et al., 2010; Malkowski et al., 2017a). The location
of the study area (shown in Fig. 2A) is demarcated by the red box. CDMC = Cordillera Darwin Metamorphic Complex; DYMC = Duque de York Metamorphic
Complex; EAMC = Eastern Andean Metamorphic Complex; LA = Lago Argentino; LV = Lago Viedma; MAB = Magallanes-Austral Basin; RVB = Rocas Verdes Basin.
(For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

systems continued until ∼90 Ma, when retroforeland convergence and of conglomeratic strata, was established (Fig. 2; Winn and Dott, 1979;
development of duplex structures in the hinterland resulted in eastward Hubbard et al., 2008; Bernhardt et al., 2012).
migration of the foredeep, as well as narrowing and deepening of the By ∼80 Ma, thrust sheets in the hinterland had been intruded by
basin (Fosdick et al., 2011; Romans et al., 2011). The depositional re- epizonal plutons, suggesting an eastward shift in the locus of thrust
sponse to this change is reflected in a ∼2500 m thick stratigraphic fault propagation (Fosdick et al., 2011; Calderón et al., 2012; Ghiglione
succession associated with the Coniacian-Campanian Cerro Toro For- et al., 2014). This is interpreted to have been followed by a ∼10 m.y.
mation in Chile and Argentina (Cecioni, 1957; Arbe and Hechem, 1984; period of intense structural doming, deformation, and extrusion of the
Winn and Dott, 1977, 1979; Hubbard et al., 2008; Romans et al., 2011; accretionary wedge in Chilean and Argentine Patagonia (80-70 Ma;
Ghiglione et al., 2014; Malkowski et al., 2018). The initial phase of Calderón et al., 2012; Calderón et al., 2016). Enhanced denudation of
Cerro Toro Formation sedimentation is recorded by a thick (∼1500 m) the hinterland and fold-thrust belt during deformation, coupled with a
succession of dominantly fine-grained deposits in the basin foredeep decrease in tectonic subsidence rate, together may have led to foredeep
(Fig. 2; Bernhardt et al., 2012). However, by ∼88 Ma (Bernhardt et al., shallowing during this time period (Daniels et al., 2018). Deposition in
2012; Malkowski et al., 2018), sediment-routing systems emanating the basin over the middle to late Campanian is recorded in sandstone
from the hinterland began to transect shelf edge and slope regions, and and siltstone units associated with north-to-south prograding deep-
a southward-verging basin-axial channel system, recorded in ∼1000 m water slope clinoform systems preserved in the ∼2500 m thick Tres

3
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Pasos and Alta Vista Formations in Chile and Argentina, respectively Graham, 2015; Gutiérrez et al., 2017), and the La Anita and Cerro
(Fig. 2; Macellari et al., 1989; Shultz et al., 2005; Hubbard et al., 2010; Fortaleza Formations in Argentina (Sickmann et al., 2018), which can
Ghiglione et al., 2014; Sickmann et al., 2018). These deep-water units reach thicknesses of ∼1000 m (Fig. 2). These linked Campanian-
are genetically related to Campanian-Maastrichtian shelf, marginal Maastrichtian units are interpreted to record the final infilling phase of
marine, and terrestrial sandstone- and conglomerate-prone deposits of the Cretaceous Magallanes-Austral Basin (Hubbard et al., 2010; Romans
the Dorotea Formation in Chile (Covault et al., 2009; Schwartz and et al., 2011). A regional unconformity, which spans the Paleocene and

(caption on next page)

4
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Fig. 2. Overview of outcropping Magallanes-Austral basin-fill units in Santa Cruz Province, Argentina, and Última Esperanza Province, Chile. (A) Satellite image of
the study area (image data: Google, Landsat, Copernicus, 2018, http://www.google.com/earth/index.html, center of image coordinates: 50°42′16.91″S,
72°40′4.06″W). Featured outcrop locations are indicated with red dots; new data is from outcrops indicated with blue dots (see Part C). (B) Stratigraphic column of
the Rocas Verdes and Magallanes-Austral basin-fill units in the study area. E = Eocene; L = Lower; P = Paleocene; Pz = Paleozoic; O = Oligocene. Characteristic
lithology (yellow = sandstone/conglomerate; brown = mass transport deposit; gray = siltstone; light/dark purple = volcanogenic units) and architecture (chan-
nelform, tabular) for each formation is shown. (modified from Daniels et al., 2018; adapted from Wilson, 1991; Fildani and Hessler, 2005; Fosdick et al., 2011;
Schwartz et al., 2017). (C) Geologic map of the study area (modified from Fosdick et al., 2011; Malkowski et al., 2017a; Daniels et al., 2018; Sickmann et al., 2018).
All Upper Cretaceous deep-water units north of Lago Argentino are mapped as part of the Cerro Toro Formation sensu Ghiglione et al. (2014); mapped units
associated with the Punta Barrosa and Alta Vista Formations are not yet well-constrained in this region. Locations of sectors featured in Fig. 5 are indicated.
J = Jurassic; K = Cretaceous; N = Neogene; Pg = Paleogene. (For interpretation of the references to color in this figure legend, the reader is referred to the Web
version of this article.)

much of the Eocene and is interpreted to be related to large-scale fold- et al., 2018; Sickmann et al., 2018). Dates reported in those studies
thrust belt evolution (Fosdick et al., 2013), caps the succession of in- were acquired via conventional sensitive high-resolution ion microp-
terest. robe (reverse geometry) methods (SHRIMP-RG; e.g., DeGraaff-Surpless
et al., 2002), laser ablation inductively coupled plasma mass spectro-
2.2. Study area metry techniques (LA-ICP-MS; e.g., Fedo et al., 2003; Gehrels et al.,
2008; Matthews and Guest, 2017), or thermal ionization mass spec-
We examine deposits from the Punta Barrosa, Cerro Toro, Tres trometry approaches (TIMS, e.g., Parrish and Noble, 2003). Strontium
Pasos, and Dorotea Formations (and age equivalents) along a near- isotope ages (N = 3) were derived from 87Sr/86Sr dates (nsamples = 14)
continuous, north-south oriented outcrop belt that extends for reported by Bernhardt et al. (2012). 87Sr/86Sr dates were acquired using
∼120 km between Laguna Tres de Abril in Argentina (Malkowski et al., the United States Geological Survey's Isotope Laboratory Menlo Park
2018) and Puerto Natales in Chile (Fig. 1). Post-Cretaceous uplift and Multicollector MAT 261, following methods outlined by Howarth and
deformation (Fosdick et al., 2011), along with glaciation (Sagredo McArthur (1997), McArthur et al. (2001), and Bernhardt et al. (2012).
et al., 2011; Solari et al., 2012; García et al., 2014), has produced Specific analytical procedures for each sample set are summarized in
outcrops that preserve largely depositional-dip-oriented and dip-ob- the aforementioned references.
lique perspectives of the basin fill. These exceptionally exposed out- In addition to previously reported ages, we report new U-Pb max-
crops enable comprehensive examination of stratigraphic architecture imum depositional ages from detrital zircons (N = 13) that were ob-
and depositional timing for all Upper Cretaceous units. tained from fine-to medium-grained sandstones from Upper Cretaceous
Magallanes-Austral Basin units that crop out near Laguna Sofia, Chile
3. Stratigraphic and geochronologic data (Figs. 3 and 4). Detrital zircon grains from this area were separated via
standard mineral separation techniques (Fedo et al., 2003; Gehrels
3.1. Stratigraphic data et al., 2008). U-Pb ratios and grain ages were determined at the Uni-
versity of Calgary using LA-ICP-MS, following practices outlined by
Stratigraphic data discussed in this study are derived from previous Matthews and Guest (2017) and Daniels et al. (2018). Isolated zircon
research. Outcropping deposits of the Late Cretaceous Magallanes- grains were first measured using a high-throughput (n > 500), short
Austral Basin in the study area have been mapped and characterized by ablation period (t = 5 s per grain) method to acquire a large amount of
numerous workers over the past six decades (e.g., Cecioni, 1957; Katz, detrital age data for provenance purposes and to increase the likelihood
1963; Scott, 1966; Smith, 1977; Winn and Dott, 1977, 1979; Arbe and of finding large numbers of young, near-depositional-age zircons (i.e.,
Hechem, 1984; Macellari et al., 1989; Wilson, 1991; Coleman, 2000; zircons that approximate true depositional age; Pullen et al., 2014;
Sohn et al., 2002; Beaubouef, 2004; Shultz and Hubbard, 2005; Shultz Daniels et al., 2018). Following identification, the youngest grains in
et al., 2005; Valenzuela, 2006; Crane and Lowe, 2008; Hubbard and the population (n ≈ 20–60 per sample) were repolished and measured
Shultz, 2008; Hubbard et al., 2008; Armitage et al., 2009; Covault et al., again using a longer ablation period (t = 20 s per grain) to minimize
2009; Fildani et al., 2009; Romans et al., 2009; Hubbard et al., 2010; measurement uncertainty associated with each grain age (Daniels et al.,
Jobe et al., 2010; Bernhardt et al., 2011; Campion et al., 2011; 2018). In some cases, grains from the youngest population could ac-
Macauley and Hubbard, 2013; Ghiglione et al., 2014; Schwartz and commodate multiple 22 μm beam spots. These grains were measured
Graham, 2015; Auchter et al., 2016; Pemberton et al., 2016; Gutiérrez multiple times to better constrain individual grain ages and evaluate
et al., 2017; Daniels et al., 2018; Malkowski et al., 2018). Recent potential Pb-loss (Gehrels, 2012; Spencer et al., 2016). Details of the
mapping has incorporated measurement of stratigraphic sections at mineral separation and LA-ICP-MS procedures for these samples are
decimeter-scale, along with analysis of ground- and unmanned-aerial- provided in Appendix A.
vehicle-based photography and satellite imagery, as well as differential
GPS surveying. Mapping techniques have been augmented with pa- 3.2.2. Depositional age calculation
leocurrent data derived from measurements of sole marks, imbricated U-Pb zircon maximum depositional ages from all data sources (see
clasts, cross-stratification, and ripple crests. These data enable robust Appendix B for ages from Laguna Sofia) were determined via weighted
interpretations of sediment paleodispersal patterns in the basin. mean age calculations conducted using the Isoplot add-in (v. 4.15) for
Microsoft Excel (Ludwig, 2012). All maximum depositional ages were
3.2. Geochronologic data computed from 206Pb/238U ratios with a probability of con-
cordance > 1% (Ludwig, 1998). In most cases, maximum depositional
3.2.1. Data sources ages were determined from the weighted mean of the youngest cluster
Geochronologic data includes depositional ages determined from U- of ages (n ≥ 3) in a sample that overlapped at 2σ uncertainty (YC2σ;
Pb detrital and volcanic ash zircon dates, as well as strontium isotope Dickinson and Gehrels, 2009; Coutts et al., 2019); we use this age
dates. Previously reported depositional ages from detrital and volcanic calculation method as it represents a conservative limit of the true
ash zircons (N = 38) were derived from studies that focused on sedi- depositional age (i.e., the YC2σ age is rarely younger than the true
ment provenance and depositional timing in the basin (Fildani et al., depositional age; Coutts et al., 2019). In cases where < 3 ages were
2003; Romans et al., 2010; Bernhardt et al., 2012; Fosdick et al., 2015; present in the youngest age cluster, we computed maximum deposi-
Malkowski et al., 2017a, 2018, 2017b; Schwartz et al., 2017; Daniels tional ages using the weighted mean of the youngest 2 ages that

5
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Fig. 3. Summary of studied outcrops near Laguna Sofia, Chile. (A) Stratigraphic locations of new detrital zircon geochronology samples (indicated with red dots). CT
= Cerro Toro Formation; DO = Dorotea Formation; E = Eocene; L = Lower; LS = Laguna Sofia; O = Oligocene; P = Paleocene; PB = Punta Barrosa Formation;
Pz = Paleozoic; TP = Tres Pasos Formation. Characteristic lithology (yellow = sandstone/conglomerate; brown = mass transport deposit; gray = siltstone; light/
dark purple = volcanogenic units) and architecture (channelform, tabular) for each formation is shown (modified from Daniels et al., 2018; adapted from Wilson,
1991; Fildani and Hessler, 2005; Fosdick et al., 2011; Schwartz et al., 2017). (B) Geographic locations of new detrital zircon geochronology samples (image data:
Google, DigitalGlobe, US Department of State, CNES/Airbus, 2018, http://www.google.com/earth/index.html, center of image coordinates: 51°32′49.69″S,
72°36′46.78″W). Maximum depositional ages for each sample are provided. (C) Normalized probability density plots of detrital zircon ages < 1600 Ma for each
sample from the Laguna Sofia area. Numbers of ages in view are indicated in parentheses. (For interpretation of the references to color in this figure legend, the
reader is referred to the Web version of this article.)

overlapped at 1σ uncertainty (YC1σ; Dickinson and Gehrels, 2009; 4. New U-Pb detrital zircon age results from Laguna Sofia, Chile
Coutts et al., 2019). Each maximum depositional age incorporates all
systematic uncertainties including excess variance related to the long- New U-Pb detrital zircon age results from Upper Cretaceous units
term reproducibility of the 206Pb/238U ratio in a validation reference exposed near Laguna Sofia are summarized in probability density plots
material, uncertainty in the 206Pb/238U ratio of the calibration re- (Fig. 3), which display the original distribution of zircon age popula-
ference material (σy), as well as uncertainty associated with U-Pb decay tions for each sample, as well as weighted mean age plots that show the
constants (λ; Jaffey et al., 1971, following modifications by Mattinson, youngest cluster of ages in each sample (Fig. 4), which provide the basis
1987; Horstwood et al., 2016). The age calculation method uses in- for maximum depositional age determination.
dividual grain ages together with weighted mean ages determined from
multiple measurements on the same grain wherever possible (Fig. 4). 4.1. Punta Barrosa Formation
Volcanic ash layer depositional ages from previous work employed a
weighted mean of dates associated with the youngest population of ages The Punta Barrosa Formation is exposed on Monte Pirámide, im-
in a sample. Ages were computed from 206Pb/238U ratios in all cases. mediately west of Laguna Sofia (Fig. 3; Fildani et al., 2003; Fosdick
Strontium isotope ages from previous work were computed from a et al., 2011). One sandstone sample (LS-PB1) was obtained from an
weighted mean of 87Sr/86Sr dates acquired from inoceramid shells interval of interbedded sandstone and siltstone located approximately
collected at each sample location. Details of depositional age calcula- midway between the base and top of the formation (Fig. 3, Table 1).
tions for these samples are summarized in the studies where ash and Detrital zircon dates from this sample range between
shell ages were first reported (e.g., Bernhardt et al., 2012; Malkowski 3058.3 ± 78.7 Ma and 90.1 ± 5.6 Ma (Fig. 3, Appendix B). Weighted
et al., 2017b; Daniels et al., 2018; Malkowski et al., 2018). mean age calculations that employed the youngest cluster of zircon ages
The reported degree of uncertainty differs between published results resulted in a maximum depositional age of 94.6 ± 1.6 Ma for this
that are integrated into this study. In most cases, random and sys- sample (Figs. 3 and 4, Table 1).
tematic uncertainty is reported; however, it is usually not clear what
systematic uncertainty components were included in the estimate. To
enable effective comparison of ages between data sets, we have in- 4.2. Cerro Toro Formation
cluded all reported random and systematic uncertainty components and
propagated them into individual grain ages and depositional ages The Cerro Toro Formation crops out on Cerro Benítez, located on
where appropriate (Horstwood et al., 2016; Matthews et al., 2017). the south shore of Laguna Sofia (Fig. 3; Winn and Dott, 1979; Hubbard
et al., 2008). The lower siltstone-prone part of the Cerro Toro Forma-
tion crops out on the western flank of Cerro Benítez, whereas

6
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Fig. 4. Weighted mean age plots that show Laguna Sofia maximum depositional ages obtained from the weighted mean of the youngest cluster of dates that overlap
at 2σ uncertainty (YC2σ). The number of grains used in each calculation is provided in parentheses on each plot. Stratigraphic younging direction is up and to the
right; maximum depositional ages that are significantly older than age constraints from underlying strata are indicated in magenta. Individual age results from a
single analysis on a single grain are indicated with black error bars. Age results derived from a weighted mean of multiple analyses on the same grain are indicated
with blue and red error bars. Blue error bars illustrate age uncertainty associated with 95% confidence limits (C.L.) determined from the population standard
deviation (σ) multiplied by 1.96 (student's-t value for an infinite number of points). This uncertainty calculation technique is used by Isoplot if the ages used in the
calculation have a probability of fit > 15% (Ludwig, 2012). Red error bars show age uncertainty associated with 95% confidence limits (C.L.) determined from the
product of the student's-t result for N-1 degrees of freedom (N = number of ages used in the calculation), σ, and the square root of the MSWD for the age cluster. This
is the default uncertainty calculation in Isoplot if dates used to generate a weighted mean age have a probability of fit < 15% (Ludwig, 2012). Both types of colored
error bars are shown where the red error bar calculation method generated a result with a very large range (e.g., LS-PB1); in some cases, red error bars are clipped at
the margins of each plot to improve visibility of the other error bars in the plot (refer to Appendix B for actual age uncertainty ranges). CT = Cerro Toro Formation;
DO = Dorotea Formation; LS = Laguna Sofia; MSWD = mean square of weighted deviates; PB = Punta Barrosa Formation; TP = Tres Pasos Formation. (For in-
terpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

conglomeratic deposits related to the upper part of the formation are 83.9 ± 3.7 Ma (Fig. 3, Appendix B). Analysis of the youngest cluster of
exposed further east (Fig. 3; Hubbard et al., 2008; Bernhardt et al., zircons in that sample yielded a maximum depositional age of
2012). One sandstone sample (LS-CT0) was obtained from a unit to- 86.3 ± 1.5 Ma (Figs. 3 and 4, Table 1). Detrital zircon dates from
wards the top of the lower siltstone-prone part and six sandstone samples in the upper conglomeratic part of the formation (LS-CT1 to LS-
samples (LS-CT1 to LS-CT6) were collected from various intervals CT6) collectively span 3366.3 ± 71.4 Ma to 73.1 ± 6.9 Ma (Fig. 3,
within the upper conglomeratic part (Figs. 3 and 4). Appendix B). Maximum depositional ages from samples LS-CT1 to LS-
Detrital zircon dates from the sample from the lower siltstone-prone CT5 range from 84.7 ± 1.3 Ma to 79.7 ± 2.2 Ma. A maximum de-
part of the formation (LS-CT0) range between 3124.6 ± 86.3 Ma and positional age of 89.5 ± 2.0 Ma was computed from the

7
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Table 1
Sandstone sample information (Laguna Sofia maximum depositional ages).
Sample name Latitude Longitude Elevation Stratigraphic interval Maximum depositional age Description
(°S) (°W) (m) (YC2σ)

LS-DO2 51°36′25.24″ 72°27′01.53″ 838 Dorotea Fm 68.6 ± 1.3 Ma Fine-grained sandstone


LS-DO1 51°36′31.57″ 72°27′36.64″ 701 Dorotea Fm 68.9 ± 1.7 Ma Fine-grained sandstone
LS-TP3 51°35′59.31″ 72°28′59.85″ 265 Tres Pasos Fm 73.4 ± 1.6 Ma Fine-grained sandstone
LS-TP2 51°35′04.19″ 72°30′01.31″ 191 Tres Pasos Fm 76.9 ± 1.1 Ma Fine-grained sandstone
LS-TP1 51°33′44.11″ 72°30′35.89″ 203 Tres Pasos Fm 78.5 ± 1.6 Ma Fine-grained sandstone
LS-CT6 51°32′41.48″ 72°33′21.58″ 155 Cerro Toro Fm (Upper) 89.5 ± 2.0 Maa Medium-grained sandstone
LS-CT5 51°32′54.12″ 72°34′32.30″ 175 Cerro Toro Fm (Upper) 79.7 ± 2.2 Ma Medium-grained sandstone
LS-CT4 51°32′48.43″ 72°35′19.65″ 252 Cerro Toro Fm (Upper) 82.4 ± 1.9 Ma Medium-grained sandstone
LS-CT3 51°32′49.27″ 72°35′27.71″ 355 Cerro Toro Fm (Upper) 80.2 ± 1.4 Ma Medium-grained sandstone
LS-CT2 51°32′45.60″ 72°36′57.42″ 300 Cerro Toro Fm (Upper) 82.5 ± 1.2 Ma Fine-grained sandstone
LS-CT1 51°32′07.48″ 72°37′39.60″ 74 Cerro Toro Fm (Upper) 84.7 ± 1.3 Ma Fine-grained sandstone
LS-CT0 51°32′55.98″ 72°40′49.26″ 156 Cerro Toro Fm (Lower) 86.3 ± 1.5 Ma Fine-grained sandstone
LS-PB1 51°27′25.08″ 72°47′35.98″ 816 Punta Barrosa Fm 94.6 ± 1.6 Ma Fine-grained sandstone

a
Maximum depositional age is significantly older than age constraints from underlying strata.

stratigraphically highest conglomeratic deposits in the formation at the stratigraphic unit in question are used in the framework (Fig. 5; cf.
Cerro Benítez (LS-CT6). Since this age is much older than inferred de- Johnstone et al., 2019).
positional ages from the underlying deposits, it does not accurately We define the maximum age of the base of the Upper Cretaceous
constrain the true depositional age of this unit. In this case, it is plau- Magallanes-Austral Basin succession to be ca. 100 Ma in this part of
sible that: (1) sample sizes were not large enough to obtain a large Última Esperanza Province. This limit is constrained by a U-Pb volcanic
population of near-depositional-age zircons (Pullen et al., 2014); (2) ash age of 101.0 ± 2.2 Ma from the upper part of a siltstone-prone
near-depositional-age zircons were not incorporated into the sediment- package of strata that records the transition from the underlying Zapata
routing system at the time of deposition; or (3) the youngest population Formation into the Punta Barrosa Formation (Fildani and Hessler, 2005;
of zircons was primarily derived from ∼90 Ma Cretaceous arc sources Fosdick et al., 2011), as well as a volcanic ash U-Pb age of
and/or recycling of ∼90 Ma intrabasinal units. 98.1 ± 2.4 Ma from the basal part of the Punta Barrosa Formation at
Laguna Tres de Abril (Figs. 2 and 5; Malkowski et al., 2017a, 2017b).
4.3. Tres Pasos Formation Previous workers have suggested that this basal boundary may be
younger in the southern part of the study area (e.g., Fildani et al., 2003;
Deposits of the Tres Pasos Formation are exposed in ridges east of Bernhardt et al., 2012; Malkowski et al., 2017b); however, absolute age
Laguna Sofia (Fig. 3; Fosdick et al., 2011); three sandstone samples (LS- constraint for equivalent deposits does not yet exist in this region. The
TP1 to LS-TP3) were obtained from exposures associated with these age at the top of the basin fill is variable along depositional dip;
ridges. Detrital zircon dates span 3072.8 ± 107.2 Ma to available absolute ages suggest that this boundary is as old as ca. 71 Ma
71.2 ± 5.6 Ma (Fig. 3, Appendix B). Maximum depositional ages from in the north and as young as ca. 65 Ma in the south (Fig. 5; Fosdick
this interval range from 78.5 ± 1.6 Ma to 73.4 ± 1.6 Ma (Figs. 3 and et al., 2015; Schwartz et al., 2017; Sickmann et al., 2018). Variability in
4, Table 1). this age is related to progressive younging of depositional systems as-
sociated with the Dorotea Formation from north to south, as well as
spatially variable erosion related to the Paleocene-Eocene angular un-
4.4. Dorotea Formation
conformity that truncates the top of the succession (Fosdick et al., 2015;
Schwartz et al., 2017; Sickmann et al., 2018).
Deposits of the Dorotea Formation are exposed in the Sierra Dorotea
range east of the Tres Pasos Formation outcrops (Fig. 3). Two sandstone
samples (LS-DO1 and LS-DO2) were obtained from exposures in the 5.1. Sector 1
range. Detrital zircon dates from Dorotea Formation units range from
3520.9 ± 86.6 Ma to 66.8 ± 1.9 Ma (Fig. 3, Appendix B). Maximum Chronostratigraphic results from Sector 1 incorporate ages from a
depositional ages derived from this interval span 68.9 ± 1.7 Ma to series of exposures located near the northern limit of Última Esperanza
68.6 ± 1.3 Ma (Figs. 3 and 4, Table 1). Province at the border between Chile and Argentina (Figs. 2 and 5). At
the western edge of the sector, depositional timing associated with the
5. Synthesis of Upper Cretaceous chronostratigraphic data in Punta Barrosa Formation is constrained via a volcanic ash zircon age
Última Esperanza Province from near the base of the unit at Laguna Tres de Abril (LTA84;
98.1 ± 2.4 Ma; Fig. 5; Malkowski et al., 2017a), as well as a detrital
New data reported in this study were integrated into a large data- zircon age from higher up in the formation at Lago Dickson (3/5–3;
base of previously reported depositional ages from Upper Cretaceous 90.6 ± 2.5 Ma; Fig. 5; Fildani et al., 2003; Malkowski et al., 2017a).
units to examine spatial/temporal depositional age patterns along the The age of the transition into the basal siltstone-prone part of the Cerro
basin axis (i.e., southward) in Última Esperanza Province (Figs. 2 and Toro Formation is constrained by a 90.4 ± 2.7 Ma volcanic ash bed
5). To facilitate analysis of age relationships along depositional dip, from this interval at Laguna Tres de Abril (LTA81; Fig. 5; Malkowski
depositional ages were binned into five west-to-east-oriented sectors et al., 2018). The age of the top of the siltstone-prone interval is defined
(Figs. 2 and 5); results from these sectors provide the basis for the re- by an ash bed from just below the conglomeratic interval at Laguna Tres
vised chronostratigraphic framework of the basin fill (Fig. 5). Some de Abril, which yields an age of 88.0 ± 3.4 Ma (LTA88; Malkowski
samples from previous studies lacked significant populations (n ≥ 2) of et al., 2018). Accurate, high-precision constraints from within the upper
near-depositional-age zircons, resulting in a maximum depositional age conglomeratic interval are few; we consider the maximum depositional
that was significantly older than other age constraints from the host age from sample LTA78 from Laguna Tres de Abril (88.5 ± 4.2 Ma;
strata (e.g., sample TS12-LCH-1A; Schwartz et al., 2017). As a result, Fig. 5; Malkowski et al., 2018) to adequately characterize this interval.
only ages that approximately coincide with the true depositional age of The age of the basal part of the Tres Pasos Formation, which

8
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

(caption on next page)

9
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Fig. 5. Chronostratigraphic framework for the Magallanes-Austral Basin in Última Esperanza Province, Chile. Labels at top correspond to sectors outlined in Fig. 2.
Architectural information (e.g., tabular, lenticular bodies) is derived from Fildani et al. (2003), Romans et al. (2011), et al. (2018), and Daniels et al. (2018). Detrital
zircon maximum depositional age (DZ MDA) data is color-coded by formation; the MDA calculation technique for each data set (e.g., YC1σ; YC2σ) has also been
included. Age data is derived from Fildani et al. (2003), Romans et al. (2010), Fosdick et al. (2011), Bernhardt et al. (2012), Fosdick et al. (2015), Schwartz et al.
(2017), Malkowski et al. (2017a), Malkowski et al. (2017b), Daniels et al. (2018), Malkowski et al. (2018), Sickmann et al. (2018), as well as data presented in this
study. Only MDAs that approximately temporally coincide with the true depositional ages of the studied stratigraphic intervals are shown. The majority of ash
samples from the Cerro Toro Formation are derived from fine-grained units adjacent to conglomeratic strata; this relationship is shown by a gradational contact
between the two time-equivalent units in the diagram. Depositional ages for stratigraphic units in Sector 5 deposited between ca. 94-79 Ma and ca. 74-72 Ma
(indicated by hatched lines) are poorly constrained; stratigraphic boundaries in these intervals are based on interpretations from Sector 4 and are speculative. CT =
Cerro Toro Formation; DO = Dorotea Formation; LTA = Laguna Tres de Abril; NA = not available; PB = Punta Barrosa Formation; SS = Silla Syncline; TP = Tres
Pasos Formation. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

contains abundant chaotically bedded mass-transport deposits (i.e., the upper part of the Cerro Toro Formation are derived from two facies
Phase 1; Fig. 5; Daniels et al., 2018), is defined by a detrital zircon age associations: (1) conglomeratic strata that record deposition within
from Cerro Divisadero (F04; 82.8 ± 5.2 Ma; Fig. 5; Romans et al., deep-marine slope channels; and (2) siltstone-prone strata that reflect
2010), as well as a detrital zircon age from Cerro Cagual from higher up deposition in out-of-channel zones (i.e., levees and overbank environ-
in section (15-RZ-DZ1; 78.5 ± 1.3 Ma; Fig. 5; Daniels et al., 2018). ments; Beaubouef, 2004; Crane and Lowe, 2008; Hubbard et al., 2008).
Temporal information linked to the boundary between the basal part of Depositional ages from conglomeratic facies are derived from outcrops
the formation and the overlying strata, which feature abundant silt- at Sierra del Toro and those exposed in the Silla Syncline (Figs. 2 and 5;
stone-prone slope deposits (i.e., Phase 2–4; Fig. 5; Daniels et al., 2018), Crane and Lowe, 2008; Jobe et al., 2010; Bernhardt et al., 2011, 2012).
is captured via a detrital zircon age from Cerro Cagual of A depositional age of 84.0 ± 1.6 Ma was computed for channel de-
78.0 ± 1.3 Ma (sample 15-RZ-DZ2; Fig. 5). No other depositional ages posits near the base of the upper part of the Cerro Toro Formation at
are available for accurate temporal characterization of the upper Tres Sierra del Toro (SdT-Co; Bernhardt et al., 2012); this age is interpreted
Pasos Formation in this sector. to record the onset of deposition within the main channel belt asso-
Depositional age information associated with the contact between ciated with the Cerro Toro Formation depositional system (Hubbard
the Tres Pasos Formation and Dorotea Formation is not available in this et al., 2008; Jobe et al., 2010; Malkowski et al., 2018). Outcrops at the
sector; we employ a depositional age from the upper part of the Cerro Silla Syncline yielded a depositional age of 81.8 ± 3.7 Ma (SS-Ndskld;
Fortaleza Formation in the Zona Centro region in Argentina to place Fig. 5; Bernhardt et al., 2012); these deposits record deposition within a
limits on the maximum age of the base of the Dorotea Formation in this tributary channel system that routed sediment to the main channel belt
sector (EP13; 76.2 ± 2.2 Ma; Fig. 5; Sickmann et al., 2018). Two det- further south (Bernhardt et al., 2012). Ages from out-of-channel facies
rital zircon depositional ages from the Río de las Chinas outcrops pro- range from 84.8 ± 2.0 Ma to 81.7 ± 1.7 Ma; notably, these ages are
vide robust temporal data for the upper part of the Dorotea Formation largely contemporaneous with ages from conglomeratic units (Fig. 5;
in this sector (sample JCF 09–208 = 69.7 ± 2.1 Ma; sample TS11-RT- Bernhardt et al., 2012; Daniels et al., 2018). A volcanic ash sample from
2A = 68.9 ± 2.3 Ma; Fosdick et al., 2015; Schwartz et al., 2017). The fine-grained strata exposed on western Cerro Cazador (15-ECB-01;
upper limit of the Dorotea Formation is marked by an unconformity; we 82.8 ± 0.3 Ma; Daniels et al., 2018) constrains large-scale lateral facies
interpret that deposits just below this unconformity are Maastrichtian relationships in the upper part of the Cerro Toro Formation in a key
in age at this location based on the aforementioned U-Pb ages, as well outcrop showing stratigraphic continuity with the overlying Tres Pasos
as nearby vertebrate fossil assemblages (Riccardi and Rolleri, 1980; Formation.
Macellari et al., 1989). Deposits from the basal part of the Tres Pasos Formation in this
sector (i.e., Phase 1; Daniels et al., 2018) yield an age of 80.7 ± 1.5 Ma
5.2. Sector 2 (15-CAZ-04; Fig. 5); we consider this age to represent the maximum age
limit of Tres Pasos Formation deposition in this sector (cf. Daniels et al.,
Chronostratigraphic results from Sector 2 include information from 2018). Strata from the upper part of the Tres Pasos Formation (i.e.,
outcrops spanning Cerro Ferrier in the west to Cerro Cazador in the east Phases 2–4) return depositional ages of 75.7 ± 1.7 Ma (15-CAZ-02)
(Figs. 2 and 5). In this sector, the Punta Barrosa Formation is temporally and 75.2 ± 1.7 Ma (15-CAZ-01), respectively (Fig. 5; Daniels et al.,
constrained through two detrital zircon samples obtained by Fildani 2018). Due to a lack of depositional age constraints for the contact
et al. (2003); one of these samples was re-analyzed by Malkowski et al. between the Tres Pasos and Dorotea Formations, we interpret the age of
(2017a) to acquire a larger number of near-depositional-age zircons. the base of the Dorotea Formation to be younger than 75.2 ± 1.7 Ma in
Sample Pb-01-04, which was collected from siltstone-dominated strata this area (Fig. 5). Although maximum depositional ages could not be
that record the transition between the underlying Zapata Formation calculated from Dorotea Formation samples, the top of the formation is
and the Punta Barrosa Formation, yielded an age of 97.4 ± 6.1 Ma considered to be Maastrichtian in age on the basis of U-Pb ages from
(Fig. 5; Fildani et al., 2003). Sample 2/21–3, which was collected from Sector 1 and nearby vertebrate fossil assemblages (Riccardi and Rolleri,
the lower Punta Barrosa Formation, yielded an age of 96.4 ± 2.7 Ma 1980; Macellari et al., 1989).
(Fig. 5; Fildani et al., 2003; Malkowski et al., 2017a).
The basal contact of the Cerro Toro Formation in this sector is 5.3. Sector 3
temporally constrained via a depositional age from a volcanic ash bed
in the vicinity of the contact, which was dated to be 89.5 ± 1.9 Ma in Chronostratigraphic results from Sector 3 feature depositional ages
age (SS-09-ash5; Fig. 5; Bernhardt et al., 2012). No detrital zircon from outcrops near Lago Porteño in the west to those near Laguna
samples from the lower siltstone-prone interval of the Cerro Toro For- Figueroa in the east. The maximum age of the base of the Punta Barrosa
mation were available in this sector (Bernhardt et al., 2012). However, Formation is characterized by a volcanic ash bed associated with
the age of this interval is well documented through results from two sample JCF 09–217, which yields an age of 101.0 ± 2.2 Ma (Fosdick
volcanic ash beds (SdT-07-ash2 and SdT-06-ash1), which yield re- et al., 2011). Depositional age constraints for strata within the Punta
spective ages of 86.0 ± 1.3 Ma and 86.0 ± 0.8 Ma. Notably, these Barrosa Formation are limited in this sector; a sole 91.5 ± 3.7 Ma-aged
ages overlap at 2σ uncertainty with weighted mean strontium isotope volcanic ash bed (CC-09-ash1; Bernhardt et al., 2012) provides in-
ages derived from in-situ inoceramid shells from the same interval formation on the transition zone between the upper Punta Barrosa
(87.5 ± 0.3 Ma and 87.2 ± 1.1 Ma; Bernhardt et al., 2012). Ages from Formation and the lower Cerro Toro Formation. Depositional age data

10
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

from volcanic ash beds in the lower Cerro Toro Formation at this lo- glaciation in the area (Sagredo et al., 2011), outcrop exposures asso-
cation range from 84.2 ± 2.6 Ma to 82.7 ± 2.5 Ma (CC-09-ash5 and ciated with Upper Cretaceous units are sparse. A detrital zircon sample
CC-09-ash4; Bernhardt et al., 2012); strontium isotope ages from in-situ from the Punta Barrosa Formation yielded an age of 96.4 ± 1.9 Ma in
inoceramid shells from the same interval produce a weighted mean age this area (3/11–3; Fig. 5; Fildani et al., 2003; Malkowski et al., 2017a).
of 87.1 ± 0.3 Ma. Bernhardt et al. (2012) suggested that the younger No exposures of the lower Cerro Toro Formation have been documented
volcanic ash depositional age may incorporate a higher degree of in detail at this location. Outcrops associated with the upper Cerro Toro
analytical error due to the spatial position of the reference material on Formation are exposed further south at Cerro Rotundo (Hubbard et al.,
the zircon mount; as such, we consider the depositional age range for 2008); however, no depositional age constraints exist for those outcrops
the lower Cerro Toro Formation to plausibly span 91.5 ± 3.7 Ma to at this time. A depositional age of 76.6 ± 2.0 Ma was reported from a
84.2 ± 2.6 Ma at this location. Depositional age information from the conglomeratic unit (sample CB-C) on Cerro Ballena by Bernhardt et al.
upper Cerro Toro Formation is limited at this location, as all of the (2012); this unit, which is positioned west of the trend of the main
relatively small samples from the Cordillera Manuel Señoret Cerro Toro Formation channel belt, has been interpreted to represent a
(nages ≈ 60) lack a significant population of near-depositional-age zir- tributary channel to the main channel belt, which is potentially exposed
cons (samples CC and VC; Romans et al., 2010). at Cerro Rotundo ∼35 km to the south (Bernhardt et al., 2012).
The age of the basal part of the Tres Pasos Formation (i.e., Phase 1; No deposits from the Tres Pasos Formation were available to be
Daniels et al., 2018) is characterized by a detrital zircon sample (15-CS- sampled in this sector because this interval of the stratigraphy is in the
01; 80.6 ± 1.4 Ma), as well as a volcanic ash sample (15-CS-02; subsurface as a result of southward plunging structure (Fosdick et al.,
80.5 ± 0.3 Ma) from the same interval. Two detrital zircon samples 2011). Sample JCF 09–226, which was derived from the upper Dorotea
(15-PIC-04 and 15-PIC-02) from the upper Tres Pasos Formation (i.e., Formation, yielded an age of 67.1 ± 1.8 Ma. This age overlaps at 2σ
Phases 2–4; Daniels et al., 2018) yield ages of 77.4 ± 1.1 Ma and uncertainty with ages from units in the Sierra Dorotea located to the
74.3 ± 1.5 Ma respectively. Above the Tres Pasos Formation, a single north (Fig. 5), as well as with an age reported by Hervé et al. (2004)
sample from near the base of Dorotea Formation (15-CC-01) yields an from the same approximate location (67.4 ± 1.5 Ma). As with Sector 4,
age of 70.6 ± 1.5 Ma (Daniels et al., 2018). No depositional ages from the uppermost deposits of the Dorotea Formation are considered to be
the upper part of the Dorotea are available in this sector; as a result, we Danian in age at this location based on foraminiferal assemblages
interpret that the upper part of the Dorotea Formation is Maastrichtian (Malumián and Caramés, 1997); this interpreted age is also supported
in age (< 70.6 ± 1.5 Ma) at this location using the aforementioned age by the maximum depositional age obtained from sample JCF 09–226,
from the base of the formation, as well as information from vertebrate which overlaps at 2σ uncertainty with the earliest part of the Danian.
fossil assemblages exposed further north (Riccardi and Rolleri, 1980;
Macellari et al., 1989). 5.6. Recommended depositional ages for Cretaceous Magallanes-Austral
Basin units
5.4. Sector 4
To facilitate future chronostratigraphic analyses for Upper
Temporal information from Sector 4 incorporates the detrital zircon Cretaceous units of the Magallanes-Austral Basin in Última Esperanza
data from Laguna Sofia discussed above (Fig. 3), as well as a previously Province, we provide recommended ages for each key stratigraphic
reported age from the Dorotea Formation on Sierra Dorotea (sample SD- boundary (e.g., formation boundary) in the succession from 50°49′S to
06; Romans et al., 2010). In this sector, the depositional age of the 51°32′S (Sectors 1–4; Table 2). The ages are commonly derived from the
Punta Barrosa Formation is characterized by the sole detrital zircon closest deposits to each boundary (Fig. 5). Age precision, defined as the
sample from Monte Pirámide (LS-PB-1; 94.6 ± 1.6 Ma; Figs. 3 and 4). percentage related to the quotient of the 2σ uncertainty and the
The lower Cerro Toro Formation, which yields an 86.3 ± 1.5 Ma det- weighted mean age, is commonly < 3% (Table 2). This level of preci-
rital zircon age in this sector, is temporally coincident with ages from sion provides high-resolution temporal constraints for each formation,
the inoceramid-bearing siltstone-prone interval further north (Sector 3; as well as for intraformational stratigraphic intervals (e.g., upper Cerro
Fig. 5). Detrital ages from the upper Cerro Toro Formation span Toro Formation) where bounding ages do not overlap at 2σ uncertainty.
84.7 ± 1.3 Ma to 79.7 ± 2.2 Ma (Fig. 5); these ages are temporally As previously mentioned, we suggest that a 101.0 ± 2.2 Ma age
coincident with conglomeratic units associated with the same interval should be used as the maximum bounding age for the base of the Punta
further north (Fig. 5). Barrosa Formation, as well as the base of the Upper Cretaceous
The basal part of the Tres Pasos Formation (i.e., Phase 1) is poorly Magallanes-Austral Basin succession, from 50°49′S to 51°32′S (Table 2).
exposed in this sector; as a result, only the upper part of the formation The age of the Punta Barrosa Formation – lower Cerro Toro Formation
(i.e., Phases 2–4) is accurately constrained via depositional ages. As boundary is variable from north to south (90.6 ± 2.5 Ma to
previously mentioned, ages in the formation at this location span 89.5 ± 1.9 Ma; Table 2). Notably, these ages overlap at 2σ uncertainty;
78.5 ± 1.6 Ma to 73.4 ± 1.6 Ma (Figs. 3 and 4); the 78.5 ± 1.6 Ma though the 90.6 ± 2.5 Ma age can be considered a maximum age for
age can be considered a bounding age for the base of Phase 2 of the Tres the base of the interval, we recommend that the 89.5 ± 1.9 Ma age be
Pasos Formation in this area. The base of the Dorotea Formation has used as a maximum age for this boundary from 51°07′S to 51°32′S
been temporally constrained via sample SD-06, which yields an age of (Sectors 2–4), as it is more precise. The age of the boundary between
71.9 ± 2.1 Ma (Romans et al., 2010). This age overlaps within un- the lower and upper Cerro Toro Formation also varies from north to
certainty with the age from LS-DO1 (68.9 ± 1.7 Ma; Figs. 3 and 4), south, ranging from 88.0 ± 3.4 Ma to 84.0 ± 1.6 Ma (Table 2). While
which is from the same stratigraphic package. Sample LS-DO2, which is it may appear that the base of the upper conglomeratic unit becomes
from the upper Dorotea Formation, yields an age of 68.6 ± 1.3 Ma younger southward, it is important to note that all of these ages overlap
(Figs. 3 and 4). The top of the Dorotea Formation in this sector is Da- at 2σ uncertainty. Though sector-specific ages can be used for future
nian in age, based on nearby foraminiferal assemblages (Malumián and analyses on this interval as appropriate (Table 2), we recommend use of
Caramés, 1997). the most precise age (84.7 ± 1.3 Ma) for basin-scale evaluations of this
unit in the region at this time.
5.5. Sector 5 The age for the boundary between the upper Cerro Toro Formation
and the base of the MTD-prone part of the Tres Pasos Formation ranges
Sector 5 incorporates chronostratigraphic data from outcrops ex- from 80.7 ± 1.5 Ma to 80.2 ± 1.4 Ma from 50°49′S to 51°32′S
posed on Cerro Ballena (west of Puerto Natales) and at the southern end (Table 2); we recommend that the volcanic ash age of 80.5 ± 0.3 Ma
of Sierra Dorotea (Figs. 2 and 5). Due to the elevated erosion via be used to characterize this boundary as it is derived from a sample that

11
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

lies closest to the boundary (15-CS-02; Daniels et al., 2018). The age

Volcanic ash age from near the top of the Zapata-Punta Barrosa transition
range of the boundary between the MTD-dominated basal part of Tres
Maastrichtian-Danian paleofaunal assemblages (Macellari et al., 1989; Pasos Formation and overlying sandstone and siltstone turbidite-
dominated part of Tres Pasos Formation spans 78.5 ± 1.6 Ma to
Not enough information to constrain boundary in Sector 2. 77.4 ± 1.1 Ma from 50°49′S to 51°32′S (Table 2); since this in-
traformational stratigraphic boundary is somewhat diachronous in this

Highest-precision age constraints used south of Sector 2.


part of the basin, and the depositional ages associated with this
boundary are similar in terms of precision, we recommend that sector-
specific ages be used where appropriate (Table 2). The age of the li-
Bounding age from Sector 1 used in Sector 2.

Bounding age from Sector 2 used in Sector 1.

Bounding age from Sector 2 used in Sector 3.


thostratigraphic contact between the Tres Pasos and Dorotea Forma-
tions ranges from 76.2 ± 2.2 Ma in the north to 70.6 ± 1.5 Ma further
south. Since previous workers have shown that deposits of the Dorotea
Formation become younger southward, we recommend that sector-
Malumián and Caramés, 1997)

specific ages be used (Table 2). As previously mentioned, the top of the
zone (Fosdick et al., 2011).

Dorotea Formation is poorly constrained via U-Pb ages. However, we


recommend that workers employ the youngest site-specific maximum
depositional ages from the Dorotea Formation for chronostratigraphic
studies at this time (Table 2), as these ages are of high precision and
generally overlap at 2σ with the ages of fossil assemblages (Maas-
Comments

trichtian-Danian) from near the top of the formation (e.g., Macellari


et al., 1989; Malumián and Caramés, 1997). Samples from the top of the
formation could be targeted in the future to more precisely constrain
Sector 4 (51°32′05″S)

the duration of the preserved Dorotea Formation.


101.0 ± 2.2 Mad

We are unable to provide recommended ages for coeval units ex-


67.1 ± 1.8 Mae

89.5 ± 1.9 Mac


84.7 ± 1.3 Ma
71.9 ± 2.1 Ma
78.5 ± 1.6 Ma

80.2 ± 1.4 Ma

posed in areas south of 51°32′S (i.e., Sector 5), as information on the


depositional timing of the Punta Barrosa, Cerro Toro, and Tres Pasos
Formations is limited in this region. Initial results suggest that deposits
associated with the aforementioned stratigraphic intervals could be
Recommended absolute ages for key stratigraphic boundaries (Magallanes-Austral Basin, Última Esperanza Province, Chile).

younger in the south (Fig. 5; McAtamney et al., 2011; Bernhardt et al.,


Sector 3 (51°17′37″S)

2012); however, additional samples at key stratigraphic boundaries are


101.0 ± 2.2 Ma
68.9 ± 2.3 Mab

84.0 ± 1.6 Mac


89.5 ± 1.9 Mac
70.6 ± 1.5 Ma
77.4 ± 1.1 Ma

80.5 ± 0.3 Ma

needed to assess the spatial/temporal development of stratigraphic


architecture in this region.

6. Utility of the results for future studies on Magallanes-Austral


Note 2: Bolded ages represent the highest precision age constraints for each boundary across the sectors.

Basin evolution
Sector 2 (51°06′44″S)

The chronostratigraphic framework provides a template for strati-


101.0 ± 2.2 Mad
68.9 ± 2.3 Mab

78.0 ± 1.3 Mab

89.5 ± 1.9 Ma
80.7 ± 1.5 Ma

84.0 ± 1.6 Ma

graphic correlation of basin fill units to the north and south of Última
Esperanza Province, and helps to place robust temporal constraints on
the interpreted paleogeographic evolution of the basin over the Late
a

Cretaceous (Macellari et al., 1989; Romans et al., 2011; Ghiglione et al.,


NA

2014; Sickmann et al., 2018). These results can be integrated with


Note 1: All absolute ages are U-Pb ages. Uncertainty is reported at the 2σ level.
Sector 1 (50°49′14″S)

paleobiological and/or paleontological findings in the basin in Última


101.0 ± 2.2 Mad

Esperanza Province to provide a thorough record of evolutionary pat-


80.7 ± 1.5 Mac
68.9 ± 2.3 Ma

76.2 ± 2.2 Ma
78.0 ± 1.3 Ma

88.0 ± 3.4 Ma
90.6 ± 2.5 Ma

terns through time (e.g., Pardo-Pérez et al., 2012; Rubilar-Rogers et al.,


2012; Manríquez et al., 2019). Future studies of these patterns will be
critical to comprehensive characterization of the paleoenvironmental
Bounding age from time-equivalent strata in Sector 3 used.
Bounding age from time-equivalent strata in Sector 1 used.

Bounding age from time-equivalent strata in Sector 5 used.


Bounding age from time-equivalent strata in Sector 2 used.
NA = not available (no absolute age constraints available).

evolution of the region, with implications for better understanding the


development of paleoflora and paleofauna at the high southern lati-
Top of Upper Tres Pasos Fm (Phases 2–4)/Base of Dorotea Fm
Top of Lower Tres Pasos Fm (Phase 1)/Base of Upper Tres Pasos

Base of Punta Barrosa Fm (base of Upper Cretaceous basin fill)

tudes across deep geologic time (e.g., Barreda et al., 2012).


Top of Lower Cerro Toro Fm/Base of Upper Cerro Toro Fm
Top of Upper Cerro Toro Fm/Base of Lower Tres Pasos Fm
Top of Dorotea Formation (top of Upper Cretaceous basin fill)

Numerous studies on ancient Andean foreland systems have at-


Top of Punta Barrosa Fm/Base of Lower Cerro Toro Fm

tempted to decipher the relative importance of basin-evolution-con-


trolling mechanisms in Andean settings and their relationship to sedi-
ment-routing patterns across deep time (e.g., Ramos et al., 2002; Gómez
et al., 2005; Uba et al., 2007; Garzione et al., 2008; DeCelles et al.,
2011; Schwartz et al., 2017). The approximate depositional duration
associated with each of the four Upper Cretaceous formations (ca. 5–10
m.y.) is sufficiently long enough to elucidate the effects of several basin
Key Stratigraphic Boundary

evolution controlling mechanisms when the recommended bounding


ages and their 2σ uncertainty values are considered (cf. Daniels et al.,
Fm (Phases 2–4)

2018). Basin-evolution-controlling mechanisms in Andean settings that


operate over 5–10 m.y. timescales include: (1) Andean uplift and re-
(Phase 1)

lated structural deformation, which can impact basin configuration and


sediment supply (e.g., Giambiagi et al., 2003; Calderón et al., 2012;
Table 2

Eude et al., 2015); (2) tectonically controlled variations in basin sub-


d
b
a

sidence (e.g., Fosdick et al., 2014; Gianni et al., 2018); (3) eustatic sea-
c

12
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

level change (e.g., Aguirre-Urreta et al., 2011; Gianni et al., 2018); and of Andean controls on basin evolution. The chronostratigraphic in-
(4) long-term changes in paleoclimate regime, including periods of formation reported here may also assist with constraint of the long-term
regional warming or cooling (e.g., Do Campo et al., 2010; Varela et al., effects of sea-level change and paleoclimate variation on the basin fill;
2018). however, we acknowledge that additional data related to the timing of
Of the mechanisms summarized above, it is plausible that the regional sea-level and chemostratigraphic proxy studies with which to
Magallanes-Austral Basin chronostratigraphic framework is best suited infer paleoclimate changes are required before this will be possible.
to evaluate basin evolution in response to Andean uplift and deforma- Our compilation of the chronostratigraphy of Upper Cretaceous
tion. Depositional age data, when integrated with structural observa- units in the Última Esperanza Province of Chile and surrounding areas
tions from the Andes (e.g., Calderón et al., 2012), as well as a growing in Argentina also provides value to those interested in the regional
geochronologic and thermochronologic record of plutonism and de- geology of the Magallanes-Austral Basin. In particular, the framework
formation occurring within and adjacent to the basin (e.g., Hervé et al., provides a template for stratigraphic correlation of time-equivalent
2007; Calderón et al., 2012; Fosdick et al., 2015; Süssenberger et al., basin fill units exposed in Argentina further north, as well as those
2018), offers an opportunity to examine changes to sediment routing exposed south of Puerto Natales, Chile.
and basin evolution in response to specific changes in Andean uplift and
deformation, such as the development of duplex structures during the Acknowledgements
Coniacian (e.g., Fosdick et al., 2011). The depositional age data may
also assist with understanding the role of relative sea-level change, as This work represents the sum of numerous contributions from two
well as paleoclimate variation in Chilean and Argentine Patagonia decades of research based in Chilean and Argentine Patagonia, largely
during the Late Cretaceous. However, as relatively few regional studies conducted via Stanford University (Hubbard, Romans, Malkowski,
on sea-level change and paleoclimate variations in the Magallanes- Bernhardt, Jobe, Fosdick, Schwartz, Fildani, Graham), the University of
Austral Basin currently exist (e.g., del Río and Martínez, 2015; Varela Calgary (Daniels, Hubbard, Matthews), and Virginia Tech (Romans,
et al., 2018), a thorough evaluation of how these phenomena affected Kaempfe). Funding for work at Stanford University was provided by
the basin over the Late Cretaceous may not yet be possible. This sponsoring corporations of the Stanford Project on Deep-water
chronostratigraphic framework can be considered alongside new tem- Depositional Systems (SPODDS), which have included Aera Energy,
porally constrained results on sea-level changes, paleoclimate varia- Anadarko, BHP Billiton, BP, Chevron, ConocoPhillips, Emerald Trail,
tions, and chemostratigraphy in the region over the Late Cretaceous as Eni, ExxonMobil, Hess, Husky, Karoon Gas, Maersk, Marathon, NEOS,
they become available. Nexen, Occidental Petroleum, Petrobras, Reliance Industries, Repsol
Furthermore, the framework offers the opportunity to evaluate YPF, Rohöl-Aufsuchungs A.G. (RAG), Saudi Aramco, Schlumberger,
other geological elements that can assist with further comprehension of Shell, Unocal, and Venoco. Funding for work at the University of
the evolutionary history of foreland basins as a whole. For example, the Calgary and Virginia Tech was generously provided by the industrial
framework offers a means to interrogate the long-term connection be- sponsors of the Chile Slope Systems Joint Industry Project (Anadarko,
tween hinterland sources of sediment and depositional sinks in a ret- BG Group, BHP Billiton, BP, Chevron, ConocoPhillips, Equinor, Hess,
roarc foreland basin, as well as compute sedimentation rates for strata Maersk, Marathon, Nexen, Repsol, and Shell). Additional funding was
deposited along a ∼120 km segment of the Andean Cordillera to ex- provided through a Natural Sciences and Engineering Research Council
amine along-orogen variations in sediment input over time. Though of Canada (NSERC) Discovery Grant (RGPIN/341715-2013) to
results generated from these future analyses will indeed be specific to Hubbard, and by the University of Calgary Silver Anniversary
the Magallanes-Austral Basin, they could be applied to help better un- Fellowship and Queen Elizabeth II graduate scholarships to Daniels. We
derstand the evolution of retroarc foreland basins in other analogous are grateful to various landowners for granting access to the outcrops,
settings worldwide (e.g., the Western foreland basin of Taiwan; Nagel including Hotel Explora, the Alarcón family, Mauricio Álvarez
et al., 2013). Kusanovic and Hella Roehrs Jeppesen, Armando Álvarez Saldivia, Jose
Antonio Kusanovic and Tamara Mac Leod, Christian Cárdenas
7. Conclusions Alvarado, Raúl Cárdenas Rodríguez, Jorge Vidal Nuñez, and Felipe
Vidal Nuñez, as well as Estancia Cerro Guido, Estancia El Chingüe,
Compilation of chronostratigraphic data (i.e., U-Pb zircon, stron- Estancia Las Chinas, and Estancia 3R. We are grateful to Tomislav Goic
tium isotope, and fossil ages) from the Cretaceous Magallanes-Austral and Alejandrina Mac Leod for friendship and support during our field
Basin, along with thirteen new U-Pb maximum depositional ages from work. This work benefitted from inspiring discussions with Neal
detrital zircon results from near Laguna Sofia, Chile, provides a revised Auchter, Zach Sickmann, Glenn Sharman, and many other academic
chronostratigraphic framework for the basin in the Última Esperanza and industrial colleagues, for which we are extremely grateful. Finally,
Province of Chile. The new framework offers high-precision (com- we wish to thank two anonymous reviewers for their comments and
monly < 3% for each age) temporal constraints for numerous suggestions, which helped improve the manuscript, as well as Dr. James
bounding surfaces associated with each formation in the basin (re- Kellogg for editorial assistance.
commended absolute age range: 101.0 ± 2.2 Ma to 67.1 ± 1.8 Ma),
which can be used to temporally constrain discrete phases of basin Appendix A and Appendix B. Supplementary data
evolution related to deposition of these units. These data will assist
future workers with the evaluation of paleontological and/or paleo- Supplementary data to this article can be found online at https://
biological trends in the high southern latitudes, as well as elucidation of doi.org/10.1016/j.jsames.2019.05.025.
basin-evolution-controlling mechanisms that operate on 5–10 m.y.
timescales. For the purposes of deciphering the relative importance of References
basin-evolution-controlling mechanisms across these timescales, we
conclude that the depositional age constraints reported here are most Aguirre-Urreta, B., Tunik, M., Naipauer, M., Pazos, P., Ottone, E., Fanning, M., Ramos,
V.A., 2011. Malargüe Group (Maastrichtian-Danian) deposits in the Neuquén Andes,
appropriate for examinations on the effects of Andean uplift and de-
Argentina: implications for the onset of the first Atlantic transgression related to
formation. Depositional ages can be used to compute sedimentation western Gondwana break-up. Gondwana Res. 19, 482–494. https://doi.org/10.1016/
rates that are plausibly precise enough to be compared directly against j.gr.2010.06.008.
Allen, R.B., 1982. Geología de la Cordillera Sarmiento, Andes Patagónicos entre los 51°00′
a large database of geochronologic and thermochronologic data from y 52°50′ Lat. Sur, Magallanes, Chile. Serv. Nac. Geol. Miner. Sernageomin, Boletín 38,
the arc, fold-thrust belt, and basin. This may enable constraint of the 1–46.
timing of deformation in the area, facilitating a thorough investigation Arbe, H.A., Hechem, J.J., 1984. Estratigrafía y facies de depósitos marinos profundos del

13
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Cretácico Superior, Lago Argentino, Provincia de Santa Cruz. IX Congreso Geológico Daniels, B.G., Auchter, N.C., Hubbard, S.M., Romans, B.W., Matthews, W.A., Stright, L.,
Argentino Actas 5, 7–41. 2018. Timing of deep-water slope evolution constrained by large-n detrital and
Armitage, D.A., Romans, B.W., Covault, J.A., Graham, S.A., 2009. The influence of mass- volcanic ash zircon geochronology, Cretaceous Magallanes Basin, Chile. Geol. Soc.
transport-deposit surface topography on the evolution of turbidite architecture: the Am. Bull. 130, 438–454. https://doi.org/10.1130/B31757.1.
Sierra Contreras, Tres Pasos Formation (Cretaceous), southern Chile. J. Sediment. de Wit, M.J., Stern, C.R., 1981. Variations in the degree of crustal extension during for-
Res. 79, 287–301. https://doi.org/10.2110/jsr.2009.035. mation of a back-arc basin. Tectonophysics 72, 229–260. https://doi.org/10.1016/
Auchter, N.C., Romans, B.W., Hubbard, S.M., 2016. Influence of deposit architecture on 0040-1951(81)90240-7.
intrastratal deformation, slope deposits of Tres Pasos Formation, Chile. Sediment. del Río, C.J., Martínez, S.A., 2015. Paleobiogeography of the Danian molluscan assem-
Geol. 341, 13–26. https://doi.org/10.1016/j.sedgeo.2016.05.005. blages of Patagonia (Argentina): Palaeogeography, Palaeoclimatology. Palaeoecology
Barreda, V.D., Cúneo, N.R., Wilf, P., Currano, E.D., Scasso, R.A., Brinkhuis, H., 2012. 417, 274–292. https://doi.org/10.1016/j.palaeo.2014.10.006.
Cretaceous/Paleogene floral turnover in Patagonia: drop in diversity, low extinction, DeCelles, P.G., Horton, B.K., 2003. Early to middle Tertiary foreland basin developments
and a Classopollis spike. PLoS One 7, e52455. https://doi.org/10.1371/journal.pone. and the history of Andean crustal shortening in Bolivia. Geol. Soc. Am. Bull. 115,
0052455. 58–77. https://doi.org/10.1130/0016-7606(2003)115<0058:ETMTFB>2.0.CO;2.
Bayona, G., Cortés, M., Jaramillo, C., Ojeda, G., Jairo Aristizabal, J., Reyes-Harker, A., DeCelles, P.G., Carrapa, B., Gehrels, G.E., 2007. Detrital zircon U-Pb ages provide pro-
2008. An integrated analysis of an orogen-sedimentary basin pair: latest Cretaceous- venance and chronostratigraphic information from Eocene synorogenic deposits in
Cenozoic evolution of the linked Eastern Cordillera orogeny and the Llanos foreland northwestern Argentina. Geology 35, 323–326. https://doi.org/10.1130/G23322A.1.
basin of Colombia. Geol. Soc. Am. Bull. 120, 1171–1197. https://doi.org/10.1130/ DeCelles, P.G., Carrapa, B., Horton, B.K., Gehrels, G.E., 2011. Cenozoic foreland basin
B26187.1. system in the central Andes of northwestern Argentina: implications for Andean
Beaubouef, R.T., 2004. Deep-water leveed-channel complexes of the Cerro Toro geodynamics and modes of deformation. Tectonics 30, TC6013. https://doi.org/10.
Formation, upper Cretaceous, southern Chile. AAPG (Am. Assoc. Pet. Geol.) Bull. 88, 1029/2011TC002948.
1471–1500. https://doi.org/10.1306/06210403130. DeGraaff-Surpless, K., Graham, S.A., Wooden, J.L., McWilliams, M.O., 2002. Detrital
Bernhardt, A., Jobe, Z.R., Lowe, D.R., 2011. Stratigraphic evolution of a submarine zircon provenance analysis of the Great Valley Group, California: evolution of an arc-
channel-lobe complex system in a narrow fairway within the Magallanes foreland forearc system. Geol. Soc. Am. Bull. 114, 1564–1580. https://doi.org/10.1130/0016-
basin, Cerro Toro Formation, southern Chile. Mar. Pet. Geol. 28, 785–806. https:// 7606(2002)114<1564:DZPAOT>2.0.CO;2.
doi.org/10.1016/j.marpetgeo.2010.05.013. Di Giulio, A., Ronchi, A., Sanfilippo, A., Tiepolo, M., Pimentel, M., Ramos, V.A., 2012.
Bernhardt, A., Jobe, Z.R., Grove, M., Lowe, D.R., 2012. Palaeogeography and diachronous Detrital zircon provenance from the Neuquén Basin (south-central Andes):
infill of an ancient deep-marine foreland basin, Upper Cretaceous Cerro Toro Cretaceous geodynamic evolution and sedimentary response in a retroarc-forearc
Formation. Magallanes Basin Basin Res. 24, 269–294. https://doi.org/10.1111/j. basin. Geology 40, 559–562. https://doi.org/10.1130/G33052.1.
1365-2117.2011.00528.x. Dickinson, W.R., Gehrels, G.E., 2009. Use of U-Pb ages of detrital zircons to infer max-
Biddle, K.T., Uliana, M.A., Mitchum Jr., R.M., Fitzgerald, M.G., Wright, R.C., 1986. The imum depositional ages of strata: a test against a Colorado Plateau Mesozoic data-
stratigraphic and structural evolution of the central and eastern Magallanes Basin, base. Earth Planet. Sci. Lett. 288, 115–125. https://doi.org/10.1016/j.epsl.2009.09.
southern South America. In: Allen, P.A., Homewood, P. (Eds.), Foreland Basins. 013.
Wiley-Blackwell Publishing, Oxford, UK, pp. 41–61. https://doi.org/10.1002/ Do Campo, M., del Papa, C., Nieto, F., Hongn, F., Petrinovic, I., 2010. Integrated analysis
9781444303810.ch2. for constraining palaeoclimatic and volcanic influences on clay-mineral assemblages
Bruhn, R.L., Dalziel, I.W.D., 1977. Destruction of the Early Cretaceous marginal basin in in orogenic basins (Palaeogene Andean foreland, Northwestern Argentina). Sediment.
the Andes of Tierra del Fuego. In: In: Talwani, M., Pitman, W.C. (Eds.), Island Arcs, Geol. 228, 98–112. https://doi.org/10.1016/j.sedgeo.2010.04.002.
Deep Sea Trenches and Back-Arc Basins: Washington, vol. 1. American Geophysical Dott, R.H., Winn Jr., R.D., Smith, C.H.L., 1982. Relationship of late Mesozoic and early
Union, Maurice Ewing Series, pp. 395–405. https://doi.org/10.1029/ME001p0395. Cenozoic sedimentation to the tectonic evolution of the southernmost Andes and the
Calderón, M., Fildani, A., Hervé, F., Fanning, C.M., Weislogel, A., Cordani, U., 2007. Late Scotia arc. In: Craddock, C. (Ed.), Antarctic Geoscience: Madison, Wisconsin.
Jurassic bimodal magmatism in the northern sea-floor remnant of the Rocas Verdes University of Wisconsin Press, pp. 193–202.
Basin, southern Patagonian Andes. J. Geol. Soc. Lond. 164, 1011–1022. https://doi. Eude, A., Roddaz, M., Brichau, S., Brusset, S., Calderon, Y., Baby, P., Soula, J.-C., 2015.
org/10.1144/0016-76492006-102. Controls on timing of exhumation and deformation in the northern Peruvian eastern
Calderón, M., Fosdick, J.C., Warren, C., Massonne, H.-J., Fanning, C.M., Fadel Cury, L., Andean wedge as inferred from low-temperature thermochronology and balanced
Schwanethal, J., Fonseca, P.E., Galaz, G., Gaytán, D., Hervé, F., 2012. The low-grade cross section. Tectonics 34, 715–730. https://doi.org/10.1002/2014TC003641.
Canal de las Montañas Shear Zone and its role on the tectonic emplacement of the Fedo, C.M., Sircombe, K.N., Rainbird, R.H., 2003. Detrital zircon analysis of the sedi-
Sarmiento Ophiolitic Complex and Late Cretaceous Patagonian Andes orogeny, Chile. mentary record. In: In: Hanchar, H.M., Hoskin, P.W.O. (Eds.), Zircon: Chantilly,
Tectonophysics 524–525, 165–185. https://doi.org/10.1016/j.tecto.2011.12.034. Virginia, Mineralogical Society of America, Reviews in Mineralogy and
Calderón, M., Prades, C.F., Hervé, F., Avendaño, V., Fanning, C.M., Massonne, H.J., Geochemistry, vol. 53. pp. 277–303.
Theye, T., Simonetti, A., 2013. Petrological vestiges of the Late Jurassic–Early Fildani, A., Hessler, A.M., 2005. Stratigraphic record across a retroarc basin inversion:
Cretaceous transition from rift to back-arc basin in southernmost Chile: new age and Rocas Verdes–Magallanes basin, Patagonian Andes. Geol. Soc. Am. Bull. 117,
geochemical data from the Capitán Aracena, Carlos III, and Tortuga ophiolitic com- 1596–1614. https://doi.org/10.1130/B25708.1.
plexes. Geochem. J. 47, 201–217. https://doi.org/10.2343/geochemj.2.0235. Fildani, A., Cope, T.D., Graham, S.A., Wooden, J.L., 2003. Initiation of the Magallanes
Calderón, M., Hervé, F., Fuentes, F., Fosdick, J.C., Sepúlveda, F., Galaz, G., 2016. Tectonic foreland basin: timing of the southernmost Patagonian Andes orogeny revised by
evolution of the Mesozoic Andean metamorphic complexes and the Rocas Verdes detrital zircon provenance analysis. Geology 31, 1081–1084. https://doi.org/10.
ophiolites in southern Patagonia. In: Ghiglione, M.C. (Ed.), Geodynamic Evolution of 1130/G20016.1.
the Southernmost Andes. Springer International Publishing, Cham, CH, pp. 7–36. Fildani, A., Hubbard, S.M., Romans, B.W., 2009. Stratigraphic evolution of deep-water
https://doi.org/10.1007/978-3-319-39727-6_2. architecture: examples of controls and depositional styles from the Magallanes Basin,
Campion, K.M., Dixon, B.T., Scott, E.D., 2011. Sediment waves and depositional im- Chile. SEPM Field Trip Guide 10, 73.
plications for fine-grained rocks in the Cerro Toro Formation (upper Cretaceous), Folguera, A., Zárate, M., Tedesco, A., Dávila, F., Ramos, V.A., 2015. Evolution of the
Silla Syncline, Chile. Mar. Pet. Geol. 28, 761–784. https://doi.org/10.1016/j. Neogene Andean foreland basins of the southern Pampas and northern Patagonia
marpetgeo.2010.07.002. (34°-41°S), Argentina. J. South Am. Earth Sci. 64, 452–466. https://doi.org/10.1016/
Cecioni, G.O., 1957. Cretaceous flysch and molasses in departamento Ultima Esperanza, j.jsames.2015.05.010.
Magallanes Province, Chile. AAPG (Am. Assoc. Pet. Geol.) Bull. 41, 538–564. Fosdick, J.C., Romans, B.W., Fildani, A., Bernhardt, A., Calderón, M., Graham, S.A., 2011.
Charrier, R., Hérail, G., Pinto, L., García, M., Riquelme, R., Farías, M., Muñoz, N., 2013. Kinematic evolution of the Patagonian retro-arc fold-and-thrust belt and Magallanes
Cenozoic tectonic evolution in the Central Andes in northern Chile and west central foreland basin, Chile and Argentina, 51°30′S. Geol. Soc. Am. Bull. 123, 1679–1698.
Bolivia: implication for paleogeographic, magmatic and mountain building evolution. https://doi.org/10.1130/B30242.1.
Int. J. Earth Sci. 102, 235–264. https://doi.org/10.1007/s00531-012-0801-4. Fosdick, J.C., Grove, M., Hourigan, J.K., Calderón, M., 2013. Retroarc deformation and
Coleman, J.L., 2000. Reassessment of the Cerro Toro (Chile) sandstones in view of exhumation near the end of the Andes, southern Patagonia. Earth Planet. Sci. Lett.
channel-levee-overbank reservoir continuity issues. In: In: Weimer, P., Slatt, R.M., 361, 504–517. https://doi.org/10.1016/j.epsl.2012.12.007.
Coleman, J., Rosen, N.C., Nelson, H., Bouma, A.H., Styzen, M.J., Lawrence, D.T. Fosdick, J.C., Graham, S.A., Hilley, G.E., 2014. Influence of attenuated lithosphere and
(Eds.), Deep-water Reservoirs of the World, GCSSEPM Research Conference, vol. 20. sediment loading on flexure of the deep-water Magallanes retroarc foreland basin,
pp. 252–258. Southern Andes. Tectonics 33, 2505–2525. https://doi.org/10.1002/2014TC003684.
Coutts, D.S., Matthews, W.A., Hubbard, S.M., 2019. Assessment of widely used methods Fosdick, J.C., Grove, M., Graham, S.A., Hourigan, J.K., Lovera, O., Romans, B.W., 2015.
to derive depositional ages from detrital zircon populations. Geosci. Front. https:// Detrital thermochronologic record of burial heating and sediment recycling in the
doi.org/10.1016/j.gsf.2018.11.002. (in press). Magallanes foreland basin, Patagonian Andes. Basin Res. 27, 546–572. https://doi.
Covault, J.A., Romans, B.W., Graham, S.A., 2009. Outcrop expression of a continental- org/10.1111/bre.12088.
margin-scale shelf-edge delta from the Cretaceous Magallanes Basin, Chile. J. García, J.-L., Hall, B.L., Kaplan, M.R., Vega, R.M., Strelin, J.A., 2014. Glacial geomor-
Sediment. Res. 79, 523–539. https://doi.org/10.2110/jsr.2009.053. phology of the Torres del Paine region (southern Patagonia): implications for gla-
Crane, W.H., Lowe, D.R., 2008. Architecture and evolution of the paine channel complex, ciation, deglaciation and paleolake history. Geomorphology 204, 599–616. https://
Cerro Toro Formation (upper Cretaceous), Silla Syncline, Magallanes basin, Chile. doi.org/10.1016/j.geomorph.2013.08.036.
Sedimentology 55, 979–1009. https://doi.org/10.1111/j.1365-3091.2007.00933.x. Garzione, C.N., Hoke, C.D., Libarkin, J.C., Withers, S., MacFadden, B., Eiler, J., Ghosh, P.,
Cuitiño, J.I., Pimentel, M.M., Ventura Santos, R., Scasso, R.A., 2012. High resolution Mulch, A., 2008. Rise of the Andes. Science 320, 1304–1307. https://doi.org/10.
isotopic ages for the early Miocene “Patagoniense” transgression in Southwest 1126/science.1148615.
Patagonia: stratigraphic implications. J. South Am. Earth Sci. 38, 110–122. https:// Gehrels, G.E., 2012. Detrital zircon U-Pb geochronology: current methods and new op-
doi.org/10.1016/j.jsames.2012.06.008. portunities. In: Busby, C., Perez, A.A. (Eds.), Tectonics of Sedimentary Basins. Wiley-
Dalziel, I.W.D., 1986. Collision and cordilleran orogenesis: an Andean perspective. In: In: Blackwell Publishing, Chichester, UK, pp. 47–62.
Coward, M.P., Ries, A.C. (Eds.), Collision Tectonics: Geological Society, London, Gehrels, G.E., Valencia, V.A., Ruiz, J., 2008. Enhanced precision, accuracy, efficiency,
Special Publication, vol. 19. pp. 389–404. and spatial resolution of U-Pb ages by laser ablation-multicollector-inductively cou-
Dalziel, I.W.D., De Wit, M.J., Palmer, K.F., 1974. Fossil marginal basin in the southern pled plasma-mass spectrometry. Geochem. Geophys. Geosyst. 9, Q03017. https://doi.
Andes. Nature 250, 291–294. https://doi.org/10.1038/250291a0. org/10.1029/2007GC001805.

14
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

Ghiglione, M.C., Quinteros, J., Yagupsky, D., Bonillo-Martínez, P., Hlebszevtich, J., Sci. 14, 775–798. https://doi.org/10.1016/S0895-9811(01)00072-4.
Ramos, V.A., Vergani, G., Figueroa, D., Quesada, S., Zapata, T., 2010. Structure and Katz, H.R., 1963. Revision of Cretaceous stratigraphy in Patagonian cordillera of Ultima
tectonic history of the foreland basins of southernmost South America. J. South Am. Esperanza, Magallanes Province, Chile. AAPG (Am. Assoc. Pet. Geol.) Bull. 47,
Earth Sci. 29, 262–277. https://doi.org/10.1016/j.jsames.2009.07.006. 506–524.
Ghiglione, M.C., Likerman, J., Barberón, V., Beatriz Giambiagi, L., Aguirre-Urreta, B., Ludwig, K.R., 1998. On the treatment of concordant uranium-lead ages. Geochem.
Suarez, F., 2014. Geodynamic context for the deposition of coarse-grained deep- Cosmochim. Acta 62, 665–676. https://doi.org/10.1016/S0016-7037(98)00059-3.
water axial channel systems in the Patagonian Andes. Basin Res. 26, 726–745 10 Ludwig, K.R., 2012. Manual for Isoplot 3.75: Berkeley, vol. 30. Berkeley Geochronology
.1111/bre.12061. Center, Special Publication No. 5. Rev. January, pp. 75 2012.
Giambiagi, L.B., Ramos, V.A., Godoy, E., Alvarez, P.P., Orts, S., 2003. Cenozoic de- Macauley, R.V., Hubbard, S.M., 2013. Slope channel sedimentary processes and strati-
formation and tectonic style of the Andes, between 33° and 34° south latitude. graphic stacking, Cretaceous Tres Pasos Formation slope system, Chilean Patagonia.
Tectonics 22, 1041. https://doi.org/10.1029/2001TC001354. Mar. Pet. Geol. 41, 146–162. https://doi.org/10.1016/j.marpetgeo.2012.02.004.
Gianni, G.M., Dávila, F.M., Echaurren, A., Fennell, L., Tobal, J., Navarrete, C., Quezada, Macellari, C.E., Barrio, C.A., Manassero, M.J., 1989. Upper Cretaceous to Paleocene de-
P., Folguera, A., Giménez, M., 2018. A geodynamic model linking Cretaceous or- positional sequences and sandstone petrography of southwestern Patagonia
ogeny, arc migration, foreland dynamic subsidence and marine ingression in (Argentina and Chile). J. South Am. Earth Sci. 2, 223–239. https://doi.org/10.1016/
southern South America. Earth Sci. Rev. 185, 437–462. https://doi.org/10.1016/j. 0895-9811(89)90031-X.
earscirev.2018.06.016. Manríquez, L.M.E., Lavina, E.L., Fernández, R.A., Trevisan, C., Leppe, M.A., 2019.
Godoy, E., 1978. Observaciones en el complejo ofiolítico de isla Milne Edwards – Cerro Campanian-Maastrichtian and Eocene stratigraphic architecture, facies analysis, and
Tortuga (Isla Navarino), Magallanes, Chile. In: 7th Congreso Geológico Argentino, paleoenvironmental evolution of the northern Magallanes Basin (Chilean Patagonia).
Actas, vol. 2. pp. 625–636. J. South Am. Earth Sci. https://doi.org/10.1016/j.jsames.2019.04.010.
Gómez, E., Jordan, T.E., Allmendinger, R.W., Cardozo, N., 2005. Development of the Malkowski, M.A., Grove, M., Graham, S.A., 2016. Unzipping the Patagonian Andes – long-
Colombian foreland-basin system as a consequence of diachronous exhumation of the lived influence of rifting history on foreland basin evolution. Lithosphere 8, 23–28.
northern Andes. Geol. Soc. Am. Bull. 117, 1272–1292. https://doi.org/10.1130/ https://doi.org/10.1130/L489.1.
B25456.1. Malkowski, M.A., Schwartz, T.M., Sharman, G.R., Sickmann, Z.T., Graham, S.A., 2017a.
Gutiérrez, N.M., Le Roux, J.P., Vásquez, A., Carreño, C., Pedroza, V., Araos, J., Oyarzún, Stratigraphic and provenance variations in the earth evolution of the Magallanes-
J.L., Pablo Pino, J., Rivera, H.A., Hinojosa, L.F., 2017. Tectonic events reflected by Austral foreland basin: implications for the role of longitudinal versus transverse
palaeocurrents, zircon geochronology, and palaeobotany in the Sierra Baguales of sediment dispersal during arc-continent collision. Geol. Soc. Am. Bull. 129, 349–371.
Chilean Patagonia. Tectonophysics 695, 76–99. https://doi.org/10.1016/j.tecto. https://doi.org/10.1130/B31549.1.
2016.12.014. Malkowski, M.A., Sharman, G.R., Graham, S.A., Fildani, A., 2017b. Characterisation and
Hervé, F., Godoy, E., Mpodozis, C., Fanning, M., 2004. Monitoring magmatism of the diachronous initiation of coarse clastic deposition in the Magallanes-Austral foreland
Patagonian batholith through the U-Pb SHRIMP dating of detrital zircons in sedi- basin, Patagonian Andes. Basin Res. 29, 298–326. https://doi.org/10.1111/bre.
mentary units of the Magallanes Basin. In: In: Carcione, J., Donda, F., Lodolo, E. 12150.
(Eds.), GEOSUR, International Symposium on the Geology and Geophysics of the Malkowski, M.A., Jobe, Z.R., Sharman, G.R., Graham, S.A., 2018. Down-slope facies
Southernmost Andes, the Scotia Arc and the Antarctic Peninsula, Actas 4-06: variability within deep-water channel systems: insights from the Upper Cretaceous
Bolletino di Geofisica Teorica ed Applicata, vol. 45. pp. 113–117. Cerro Toro Formation, southern Patagonian. Sedimentology 65, 1918–1946. https://
Hervé, F., Pankhurst, R.J., Fanning, C.M., Calderón, M., Yaxley, G.M., 2007. The South doi.org/10.1111/sed.12452.
Patagonian batholith: 150 my of granite magmatism on a plate margin. Lithos 97, Malumián, N., Caramés, A., 1997. Upper Campanian-Paleogene from the Río Turbio coal
373–394. https://doi.org/10.1016/j.lithos.2007.01.007. measures in southern Argentina: micropaleontology and the Paleocene/Eocene
Hoorn, C., Wesselingh, F.P., ter Steege, H., Bermudez, M.A., Mora, A., Sevink, J., boundary. J. South Am. Earth Sci. 10, 189–201. https://doi.org/10.1016/S0895-
Sanmartín, I., Sanchez-Meseguer, A., Anderson, C.L., Figueiredo, J.P., Jaramillo, C., 9811(97)00015-1.
Riff, D., Negri, F.R., Hooghiemstra, H., Lundberg, J., Stadler, T., Särkinen, T., Mathalone, J.M.P., Montoya, M., 1995. Petroleum geology of the sub-Andean basins of
Antonelli, A., 2010. Amazonia through time: Andean uplift, climate change, land- Peru. In: In: Tankard, A.J., Suarez Soruco, R., Welsink, H.J. (Eds.), Petroleum Basins
scape evolution, and biodiversity. Science 330, 927–931. https://doi.org/10.1126/ of South America: Tulsa, OK, vol. 62. American Association of Petroleum Geologists
science.1194585. Memoir, pp. 423–444.
Horstwood, M.S.A., Košler, J., Gehrels, G.E., Jackson, S.E., McLean, N.M., Paton, C., Matthews, W.A., Guest, B., 2017. A practical approach for collecting large-n detrital
Pearson, N.J., Sircombe, K., Sylvester, P., Vermeesch, P., Bowring, J.F., Condon, D.J., zircon U-Pb data sets by quadrupole LA-ICP-MS. Geostand. Geoanal. Res. 41,
Schoene, B., 2016. Community-derived standards for LA-ICP-MS U-(Th-)Pb geo- 161–180. https://doi.org/10.1111/ggr.12146.
chronology – uncertainty propagation, age interpretation and data reporting. Matthews, W.A., Guest, B., Madronich, L., 2017. Latest Neoproterozoic to Cambrian
Geostand. Geoanal. Res. 40, 311–332. https://doi.org/10.1111/j.1751-908X.2016. detrital zircon facies of western Laurentia. Geosphere 14, 243–264. https://doi.org/
00379.x. 10.1130/GES01544.1.
Horton, B.K., Anderson, V.J., Caballero, V., Saylor, J.E., Nie, J., Parra, M., Mora, A., 2015. Mattinson, J.M., 1987. U-Pb ages of zircons: a basic examination of error propagation.
Application of detrital zircon U-Pb geochronology to surface and subsurface corre- Chem. Geol. 66, 151–162. https://doi.org/10.1016/0168-9622(87)90037-6.
lations of provenance, paleodrainage, and tectonics of the Middle Magdalena Valley McArthur, J.M., Howarth, R.J., Bailey, T.R., 2001. Strontium isotope stratigraphy:
Basin of Colombia. Geosphere 11, 1790–1811. https://doi.org/10.1130/GES01251.1. LOWESS version 3: best fit to the marine Sr-isotope curve for 0-509 Ma and ac-
Horton, B.K., 2017. Sedimentary record of Andean mountain building. Earth Sci. Rev. companying look-up table for deriving numerical age. J. Geol. 109, 155–170. https://
178, 279–309. https://doi.org/10.1016/j.earscirev.2017.11.025. doi.org/10.1086/319243.
Howarth, R.J., McArthur, J., 1997. Statistics for strontium isotope stratigraphy: a robust McAtamney, J., Klepeis, K., Mehrtens, C., Thomson, S., Betka, P., Rojas, L., Snyder, S.,
LOWESS fit to the marine Sr-isotope curve for 0 to 206 Ma, with look-up table for 2011. Along-strike variability variability of back-arc basin collapse and the initiation
derivation of numeric age. J. Geol. 105, 441–456. https://doi.org/10.1086/515938. of sedimentation in the Magallanes foreland basin, southernmost Andes (53-54.5°S).
Howell, J.A., Schwarz, E., Spalletti, L.A., Veiga, G.D., 2005. The Neuquén Basin: an Tectonics 30, TC5001. https://doi.org/10.1029/2010TC002826.
Overview, vol. 252. Geological Society of London Special Publications, pp. 1–14. Mpodozis, C., Ramos, V., 1989. The Andes of Chile and Argentina. In: In: Ericksen, G.E.,
https://doi.org/10.1144/GSL.SP.2005.252.01.01. Cañas Pinochet, M.T., Reinemund, J.A. (Eds.), Geology of the Andes and its Relation
Hubbard, S.M., Shultz, M.R., 2008. Deep burrows in submarine-fan channel deposits of to Hydrocarbon and Mineral Resources: Circum-Pacific Council for Energy and
the Cerro Toro Formation (Cretaceous), Chilean Patagonia: implications for firm- Mineral Resources Earth Science Series, vol. 11. pp. 59–90.
ground development and colonization in the deep sea. Palaios 23, 223–232. https:// Mpodozis, C., Mella, P., Padva, D., 2011. Estratigrafía y megasecuencias sedimentarias en
doi.org/10.2110/palo.2006.p06-127r. la Cuenca Austral-Magallanes, Argentina y Chile: VIII Congreso de exploración y
Hubbard, S.M., Romans, B.W., Graham, S.A., 2008. Deep-water foreland basin deposits of desarrollo de hidrocarburos. (Mar del Plata, Argentina).
the Cerro Toro Formation, Magallanes Basin, Chile: architectural elements of a sin- Mukasa, S.B., Dalziel, I.W.D., 1996. Southernmost Andes and south Georgia Island, north
uous basin axial channel belt. Sedimentology 55, 1333–1359. https://doi.org/10. Scotia ridge, zircon U-Pb and muscovite 40Ar/39Ar age constraints on tectonic
1111/j.1365-3091.2007.00948.x. evolution of southwestern Gondwanaland. J. South Am. Earth Sci. 9, 349–365.
Hubbard, S.M., Fildani, A., Romans, B.W., Covault, J.A., McHargue, T.R., 2010. High- https://doi.org/10.1016/S0895-9811(96)00019-3.
relief slope clinoform development: insights from outcrop, Magallanes Basin, Chile. J. Nagel, S., Castelltort, S., Wetzel, A., Willett, S.D., Mouthereau, F., Lin, A.T., 2013.
Sediment. Res. 80, 357–375. https://doi.org/10.2110/jsr.2010.042. Sedimentology and foreland basin paleogeography during Taiwan arc continent
Hulka, C., Heubeck, C., 2010. Composition and provenance history of Late Cenozoic se- collision. J. Asian Earth Sci. 62, 180–204. https://doi.org/10.1016/j.jseaes.2012.09.
diments of southeastern Bolivia: implications for Chaco foreland basin evolution and 001.
Andean uplift. J. Sediment. Res. 80, 288–299. https://doi.org/10.2110/jsr.2010.029. Natland, M.L., González, E., Cañón, A., Ernst, M., 1974. A system of stages for correlation
Husson, L., Moretti, I., 2002. Thermal regime of fold and thrust belts – an application to of Magallanes basin sediments: Boulder, Colorado. Geol. Soc. Am. Mem. 139, 126.
the Bolivian sub Andean zone. Tectonophysics 345, 253–280. https://doi.org/10. https://doi.org/10.1130/MEM139-p1.
1016/S0040-1951(01)00216-5. Nullo, F.E., Proserpie, C.A., Blasco, G., 1981. El Cretácico de la Cuenca Austral entre el
Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., Essling, A.M., 1971. Precision lago San Martín y Río Turbio. In: In: Wolkheimer, V., Musacchio, E. (Eds.), Cuencas
measurement of half-lives and specific activities of 235U and 238U. Phys. Rev. 4, Sedimentarias de América del Sur, vol. 1. pp. 181–220.
1889–1906. https://doi.org/10.1103/PhysRevC.4.1889. Olivero, E.B., Martinioni, D.R., 2001. A review of the geology of the Argentinian Fueguian
Jobe, Z.R., Bernhardt, A., Lowe, D.R., 2010. Facies and architectural asymmetry in a Andes. J. South Am. Earth Sci. 14, 175–188. https://doi.org/10.1016/S0895-
conglomerate-rich submarine channel fill, Cerro Toro Formation, Sierra del Toro, 9811(01)00016-5.
Magallanes, Chile. J. Sediment. Res. 80, 1085–1108. https://doi.org/10.2110/jsr. Otero, R.A., Soto-Acuña, S., Salazar, C., Oyarzún, J.L., 2015. New elasmosaurids
2010.092. (Sauropterygia, Plesiosauria) from the late Cretaceous of the Magallanes basin,
Johnstone, S.A., Schwartz, T.M., Holm-Denoma, C.S., 2019. A stratigraphic approach to Chilean Patagonia: evidence of a faunal turnover during the Maastrichtian along the
inferring depositional ages from detrital geochronology data. Front. Earth Sci. Weddellian Biogeographic Province. Andean Geol. 42, 237–267. https://doi.org/10.
https://doi.org/10.3389/feart.2019.00057. (in press). 5027/andgeoV42n1-a05.
Jordan, T.E., Schlunegger, F., Cardozo, N., 2001. Unsteady and spatially variable evo- Pankhurst, R.J., Riley, T.R., Fanning, C.M., Kelley, S.P., 2000. Episodic silicic volcanism
lution of the Neogene Andean Bermejo foreland basin, Argentina. J. South Am. Earth in Patagonia and Antarctic Peninsula: chronology of magmatism associated with the

15
B.G. Daniels, et al. Journal of South American Earth Sciences 94 (2019) 102209

break-up of Gondwana. J. Petrol. 41, 605–625. https://doi.org/10.1093/petrology/ Slope Systems: Processes and Products: London, vol. 244. Geological Society
41.5.605. [London] Special Publication, pp. 27–50. https://doi.org/10.1144/GSL.SP.2005.244.
Pardo-Pérez, J., Frey, E., Stinnesbeck, W., Fernández, M.S., Rivas, L., Salazar, C., Leppe, 01.03.
M., 2012. An ichthyosaurian forefin from the Lower Cretaceous Zapata Formation of Sickmann, Z.T., Schwartz, T.M., Graham, S.A., 2018. Refining stratigraphy and tectonic
southern Chile: implications for morphological variability within Platypterygius. history using detrital zircon maximum depositional age: an example from the Cerro
Palaeobiodivers. Palaeoenviron. 92, 287–294. https://doi.org/10.1007/s12549-012- Fortaleza Formation, Austral Basin, southern Patagonia. Basin Res. 30, 708–729.
0074-8. https://doi.org/10.1111/bre.12272.
Parrish, R.R., Noble, S.R., 2003. Zircon U-Th-Pb geochronology by isotope dilution- Siks, B.C., Horton, B.K., 2011. Growth and fragmentation of the Andean foreland basin
thermal ionization mass spectrometry (ID-TIMS). Rev. Mineral. Geochem. 53, during eastward advance of fold-thrust deformation, Puna plateau and Eastern
183–213. https://doi.org/10.2113/0530183. Cordillera, northern Argentina. Tectonics 30, TC6017. https://doi.org/10.1029/
Pemberton, E.A.L., Hubbard, S.M., Fildani, A., Romans, B.W., Stright, L., 2016. The 2011TC002944.
stratigraphic expression of decreasing confinement along a deep-water sediment Smith, C.H.L., 1977. Sedimentology of the Late Cretaceous (Santonian–Maastrichtian)
routing system: outcrop example from southern Chile. Geosphere 12, 114–134. Tres Pasos Formation, Ultima Esperanza District, Southern Chile. Master’s thesis.
https://doi.org/10.1130/GES01233.1. University of Wisconsin, Madison, WI, pp. 129.
Pullen, A., Ibáñez-Mejía, M., Gehrels, G.E., Ibáñez-Mejía, J.C., Pecha, M., 2014. What Sohn, Y.K., Choe, M.Y., Jo, H.R., 2002. Transition from debris flow to hyperconcentrated
happens when n=1000? Creating large-n geochronological datasets with LA-ICP-MS flow in a submarine channel (the Cretaceous Cerro Toro Formation, southern Chile).
for geologic investigations. J. Anal. At. Spectrom. 29, 971–980. https://doi.org/10. Terra. Nova 14, 405–415. https://doi.org/10.1046/j.1365-3121.2002.00440.x.
1039/C4JA00024B. Solari, M.A., Le Roux, J.P., Hervé, F., Airo, A., 2012. Evolution of the Great Tehuelche
Rabinowitz, P.D., La Brecque, J., 1979. The Mesozoic South Atlantic ocean and evolution Paleolake in the Torres del Paine National Park of Chilean Patagonia during the Last
of its continental margin. J. Geophys. Res. 84, 5973–6002. https://doi.org/10.1029/ Glacial Maximum and Holocene. Andean Geol. 39, 1–21. https://doi.org/10.5027/
JB084iB11p05973. andgeoV39N1-a01.
Ramos, V.A., Aguirre-Urreta, M.B., 1994. Cretaceous evolution of the Magallanes basin. Spencer, C.J., Kirkland, C.L., Taylor, R.J.M., 2016. Strategies towards statistically robust
In: Salfity, J.A. (Ed.), Cretaceous Tectonics of the Andes, Earth Evolution Sciences. interpretations of in situ U-Pb zircon geochronology. Geosci. Front. 7, 581–589.
Vieweg+Teubner Verlag, Wiesbaden, pp. 316–345. https://doi.org/10.1007/978-3- https://doi.org/10.1016/j.gsf.2015.11.006.
322-85472-8_7. Stern, C.R., de Wit, M.J., 2003. Rocas Verdes ophiolites, southernmost South America:
Ramos, V.A., Cristallini, E.O., Pérez, D.J., 2002. The Pampean flat-slab of the central remnants of progressive stages of development of oceanic-type crust in a continental
Andes. J. South Am. Earth Sci. 15, 59–78. https://doi.org/10.1016/S0895-9811(02) margin back-arc basin. Geol. Soc. Lond. Spec. Publ. 218, 665–683. https://doi.org/
00006-8. 10.1144/GSL.SP.2003.218.01.32.
Raychowdhury, N., Mukherjee, A., Bhattacharya, P., Johannesson, K., Bundschuh, J., Stern, C.R., Mukasa, S.B., Fuenzalida, R., 1992. Age and petrogenesis of the Sarmiento
Bejarano Sifuentes, G., Nordberg, E., Martin, R.A., del Rosario Stormiolo, A., 2014. ophiolite complex of southern Chile. J. South Am. Earth Sci. 6, 97–104. https://doi.
Provenance and fate of arsenic and other solutes in the Chaco-Pampean Plain of the org/10.1016/0895-9811(92)90020-Y.
Andean foreland, Argentina: from perspectives of hydrogeochemical modeling and Stevens Goddard, A., Carrapa, B., 2018. Effects of Miocene-Pliocene global climate
regional tectonic setting. J. Hydrol. 518, 300–316. https://doi.org/10.1016/j. changes on continental sedimentation; a case study from the southern Central Andes.
jhydrol.2013.07.003. Geology 46, 647–650. https://doi.org/10.1130/G40280.1.
Riccardi, A.C., Rolleri, E.O., 1980. Cordillera Patagónica Austral. Actas Segundo Simposio Suárez, M., Pettigrew, T.H., 1976. An upper Mesozoic Island arc back-arc system in the
de Geología Regional Argentina 2, 1173–1306. southern Andes and south Georgia. Geol. Mag. 113, 305–400. https://doi.org/10.
Richiano, S., 2014. Lower Cretaceous anoxic conditions in the Austral Basin, south- 1017/S0016756800047592.
western Gondwana, Patagonia Argentina. J. South Am. Earth Sci. 54, 37–46. https:// Süssenberger, A., Schmidt, S.T., Wemmer, K., Baumgartner, L.P., Grobéty, B., 2018.
doi.org/10.1016/j.jsames.2014.04.006. Timing and thermal evolution of fold-and-thrust belt formation in the Ultima
Roddaz, M., Viers, J., Brusset, S., Baby, P., Boucayrand, C., Hérail, G., 2006. Controls on Esperanza District, 51°S Chile: constraints from K-Ar dating and illite characteriza-
weathering and provenance in the Amazonian foreland basin: insights from major tion. Geol. Soc. Am. Bull. 130, 975–998. https://doi.org/10.1130/B31766.1.
and trace element geochemistry of Neogene Amazonian sediments. Chem. Geol. 226, Thomas, C.R., 1949a. Geology and petroleum exploration in Magallanes Province, Chile.
31–65. https://doi.org/10.1016/j.chemgeo.2005.08.010. AAPG (Am. Assoc. Pet. Geol.) Bull. 33, 1553–1578.
Romans, B.W., Hubbard, S.M., Graham, S.A., 2009. Stratigraphic evolution of an out- Thomas, C.R., 1949b. Manantiales field, Magallanes Province, Chile. AAPG (Am. Assoc.
cropping continental slope system, Tres Pasos Formation at Cerro Divisadero, Chile. Pet. Geol.) Bull. 33, 1579–1589.
Sedimentology 56, 737–764. https://doi.org/10.1111/j.1365-3091.2008.00995.x. Tunik, M., Folguera, A., Naipauer, M., Pimentel, M., Ramos, V.A., 2010. Early uplift and
Romans, B.W., Fildani, A., Graham, S.A., Hubbard, S.M., Covault, J.A., 2010. Importance orogenic deformation in the Neuquén Basin: constraints on the Andean uplift from U-
of predecessor basin history on the sedimentary fill of a retroarc foreland basin: Pb and Hf isotopic data of detrital zircons. Tectonophysics 489, 258–273. https://doi.
provenance analysis of the Cretaceous Magallanes basin, Chile (50-52°S). Basin Res. org/10.1016/j.tecto.2010.04.017.
22, 640–658. https://doi.org/10.1111/j.1365-2117.2009.00443.x. Uba, C.E., Heubeck, C., Hulka, C., 2006. Evolution of the late Cenozoic Chaco foreland
Romans, B.W., Fildani, A., Hubbard, S.M., Covault, J.A., Fosdick, J.C., Graham, S.A., basin, Southern Bolivia. Basin Res. 18, 145–170. https://doi.org/10.1111/j.1365-
2011. Evolution of deep-water stratigraphic architecture, Magallanes Basin, Chile. 2117.2006.00291.x.
Mar. Pet. Geol. 28, 612–628. https://doi.org/10.1016/j.marpetgeo.2010.05.002. Uba, C.E., Strecker, M.R., Schmitt, A.K., 2007. Increased sediment accumulation rates and
Rubilar-Rogers, D., Otero, R.A., Yury-Yáñez, R.E., Vargas, A.O., Gutstein, C.S., 2012. An climatic forcing in the central Andes during the late Miocene. Geology 35, 979–982.
overview of the dinosaur fossil record from Chile. Earth Sci. Rev. 37, 242–255. https://doi.org/10.1130/G224025A.1.
https://doi.org/10.1016/j.jsames.2012.03.003. Valenzuela, A.A.S., 2006. Proveniencia sedimentaria de Estratos de Cabos Nariz y
Ruskin, B.G., Jordan, T.E., 2007. Climate change across continental sequence boundaries: Formacion Cerro Toro, Cretacico Tardio-Paleoceno, Magallanes, Chile. PhD thesis.
Paleopedology and lithofacies of Iglesia Basin, Northwestern Argentina. J. Sediment. Universidad de Chile, Santiago, Chile, pp. 153.
Res. 77, 661–679. https://doi.org/10.2110/jsr.2007.069. Varela, A.N., Poiré, D.G., Martin, T., Gerdes, A., Goin, F.J., Gelfo, J.N., Hoffmann, S.,
Sagredo, E.A., Moreno, P.I., Villa-Martínez, R., Kaplan, M.R., Kubik, P.W., Stern, C.R., 2012. U-Pb zircon constraints on the age of the Cretaceous Mata Amarilla formation,
2011. Fluctuations of the Última Esperanza ice lobe (52°S), Chilean Patagonia, during southern Patagonia, Argentina: its relationship with the evolution of the Austral
the last glacial maximum and termination 1. Geomorphology 125, 92–108. https:// basin. Andean Geol. 39, 359–379. https://doi.org/10.5027/andgeoV39n3-a01.
doi.org/10.1016/j.geomorph.2010.09.007. Varela, A.N., Gómez-Peral, L.E., Richano, S., Poiré, D.G., 2013. Distinguishing similar
Schwartz, T.M., Graham, S.A., 2015. Stratigraphic architecture of a tide-influenced shelf- volcanic source areas from an integrated provenance analysis: implications for
edge delta, upper Cretaceous Dorotea Formation, Magallanes-Austral basin, foreland Andean basins. J. Sediment. Res. 83, 258–276. https://doi.org/10.2110/jsr.
Patagonia. Sedimentology 62, 1039–1077. https://doi.org/10.1111/sed.12176. 2013.22.
Schwartz, T.M., Fosdick, J.C., Graham, S.A., 2017. Using detrital zircon U-Pb ages to Varela, A.N., Sol Raigenborn, M., Richiano, S., White, T., Poiré, D.G., Lizzoli, S., 2018.
calculate Late Cretaceous sedimentation rates in the Magallanes-Austral basin, Late Cretaceous paleosols as paleoclimate proxies of high-latitude southern hemi-
Patagonia. Basin Res. 29, 725–746. https://doi.org/10.1111/bre.12198. sphere: Mata Amarilla formation, Patagonia, Argentina. Sediment. Geol. 363, 83–95.
Scott, K.M., 1966. Sedimentology and dispersal patterns of a Cretaceous flysch sequence, https://doi.org/10.1016/j.sedgeo.2017.11.001.
Patagonian Andes, southern Chile. AAPG (Am. Assoc. Pet. Geol.) Bull. 50, 72–107. Wilson, T.J., 1991. Transition from back-arc to foreland basin development in the
Sempere, T., Hérail, G., Oller, J., Bonhomme, M.G., 1990. Late Oligocene-Early Miocene southernmost Andes: stratigraphic record from the Ultima Esperanza District, Chile.
major tectonic crisis and related basins in Bolivia. Geology 18, 946–949. https://doi. Geol. Soc. Am. Bull. 103, 98–111. https://doi.org/10.1130/0016-7606(1991)
org/10.1130/0091-7613(1990)018<0946:LOEMMT>2.3.CO;2. 103<0098:TFBATF>2.3.CO;2.
Shultz, M.R., Hubbard, S.M., 2005. Sedimentology, stratigraphic architecture, and ich- Winn, R.D., Dott Jr., R.H., 1977. Large-scale traction produced structures in deep-water
nology of gravity-flow deposits partially ponded in a growth-fault-controlled slope fan-channel conglomerates in southern Chile. Geology 5, 41–44. https://doi.org/10.
minibasin, Tres Pasos Formation (Cretaceous), southern Chile. J. Sediment. Res. 75, 1130/0091-7613(1977)5<41:LTSIDF>2.0.CO;2.
440–453. https://doi.org/10.2110/jsr.2005.034. Winn Jr., R.D., Dott Jr., R.H., 1979. Deep-water fan-channel conglomerates of Late
Shultz, M.R., Fildani, A., Cope, T.D., Graham, S.A., 2005. Deposition and stratigraphic Cretaceous age, southern Chile. Sedimentology 26, 203–228. https://doi.org/10.
architecture of an outcropping ancient slope system: Tres Pasos Formation, 1111/j.1365-3091.1979.tb00351.x.
Magallanes Basin, southern Chile. In: In: Hodgson, D.M., Flint, S.S. (Eds.), Submarine

16

You might also like