<<

Sedimentology (2020) doi: 10.1111/sed.12715

Controls on Early desert sediment provenance in south-west Gondwana, Botucatu Formation (Brazil and Uruguay)

GABRIEL BERTOLINI*† , JULIANA C. MARQUES† , ADRIAN J. HARTLEY‡ , ATILAA.S.DA-ROSA§ , CLAITON M. S.SCHERER† ,MIGUELA.S.BASEI¶ and JOSE C. FRANTZ† *School of Geosciences, King’s College, University of Aberdeen, Aberdeen, AB24 3FX, UK (E-mail: [email protected]) †Instituto de Geociencias,^ Universidade Federal do Rio Grande do Sul, Porto Alegre 90040-060, Brazil ‡Geology & Petroleum Geology, King’s College, University of Aberdeen, Aberdeen, AB24 3FX, UK §Departamento de Geociencias,^ Universidade Federal de Santa Maria, Av. Roraima, 1000 Predio 17, Santa Maria, RS 97105-900, Brazil ¶Instituto de Geociencias,^ Universidade de Sao~ Paulo, Sao~ Paulo 05508-080, Sao~ Paulo, Brazil

Associate Editor – Subhasish Dey

ABSTRACT The Lower Cretaceous Botucatu Formation records the development of wide- spread dry–aeolian desert sedimentation throughout the Parana Basin in south- west Gondwana. To reconstruct the provenance of the aeolian sediment, petro- graphy, granulometric analysis, U-Pb detrital zircon ages have been determined from along the southern basin margin in Rio Grande do Sul state (southern Bra- zil) and Uruguay (Tacuarembo region). The dataset reveals a mean composition

Qt89F8L3, comprising very fine to medium-grained quartozose and feldspatho- quartzose framework. Heavy mineral analysis reveals an overall dominance of

zircon, tourmaline and rutile grains (mean ZTR0.84) with sporadic garnet, epi- dote and pyrolusite occurrences. The detrital zircon U-Pb ages are dominated by to (515 to 650 Ma), to (900 to 1250 Ma) and to (1.8 to 2.2 Ga) material. The detrital zircon dataset demonstrates a significant lateral variation in sediment provenance: Cambrian to Neoproterozoic detrital zircons dominate in the east, while Tonian to Stenian and Orosirian to Rhyacian ages predominate in the west of the study area. Sandstones are quartz-rich with dominantly durable zircon, tourmaline and rutile heavy mineral suite, with subtle but statistically significant along- strike differences in heavy mineral populations and detrital mineralogy which are thought to record local sediment input points into the aeolian . The similar age spectra of Botucatu desert with proximal Parana Basin units, the predominance of quartzose, and zircon, tourmaline and rutile components, sug- gests that recycling is the mechanism responsible for the erg feeding. Keywords Botucatu desert, Cretaceous Parana Basin, desert accumulation, detrital zircon, heavy minerals, provenance analysis.

INTRODUCTION understanding how and why large accumula- tions of aeolian sediment are developed and Determining the origin and transport paths of preserved (Dickinson & Gehrels, 2003; Rahl sediment in aeolian systems is important for et al., 2003). Provenance studies of dune field

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists 1 2 G. Bertolini et al. sediment supply and erg sand-delivery systems input from the south. In addition, the close reveal a range of desert sand supply mecha- proximity of the Botucatu to basement highs (for nisms. For example, sands in the Namib Desert example, the Rio-Grandense and Uruguayan are sourced by longshore drift from the Orange shields), may allow establishment of a direct River delta (Garzanti et al., 2014), sands in the ‘source to sink’ connection. Sediment prove- Taklamakan Desert are supplied by fluvial sys- nance studies from the southern part of the tems draining the surrounding mountain ranges basin will therefore be important in reconstruct- (Rittner et al., 2016), and the Rub‘ Al Khali erg ing sediment routing systems, understanding (Arabian Desert) is derived from a mix of sand controls on sediment supply and assessing the sources including aeolian transport and defla- potential for first cycle input direct from tion as well as fluvial, alluvial and coastal basement terranes. inputs (Garzanti et al., 2017; Farrant et al., The primary objective of this study is to con- 2019). In the rock record, a study of the strain and clarify the nature of the sediment ergs of the Colorado Plateau was able to recon- source for the Botucatu dunes along a transect struct a transcontinental south-east/north-west of the southern Parana Basin margin (south-east palaeoriver and palaeowind system responsible Brazil and Uruguay). Due to the continental for transportation and dispersal of sediment in scale of the aeolian system, the Botucatu Forma- Laurentia (Dickinson & Gehrels, 2009). When tion covers a number of distinct geological ter- studying the sedimentary record, however, it is ranes providing a wide range of potential often difficult to properly reconstruct aeolian sediment source areas and pathways. Prove- sediment supply systems due to the incomplete- nance analysis utilizing U-Pb detrital zircon ness of the succession and post-depositional dia- , heavy mineral analysis, petrogra- genetic modification of the original mineralogy phy and granulometric analysis has been under- (Mcbride, 1985; Cardenas et al., 2019). As a taken to ascertain whether significant variations consequence, with the exception of the Per- in sediment provenance signatures across the mian–Jurassic Colorado Plateau ergs (Dickinson southern basin margin can be identified. & Gehrels, 2003, 2009, 2010; Rahl et al., 2003), there is a lack of studies that address the recon- struction of sediment source terranes, pathways, GEOLOGICAL CONTEXT mixing, recycling and transport distance in ancient large-scale ergs. To address this gap, this The Botucatu Formation forms part of the study has undertaken a provenance analysis of intracratonic Parana Basin fill succession – an the Botucatu Formation, a Cretaceous aeolian 8 km thick (Zalan et al., 1990) sedimentary- unit deposited across south-west Gondwana. magmatic package of Late to Early The Cretaceous Botucatu dune field extended Cretaceous age (Fig. 2). In the south (Brazil, continuously across 1.5 million km2 of the inte- Argentina, Uruguay and Paraguay), the basin-fill rior of the south-west Gondwana is divided into six supersequences (Milani et al., and represents part of what was probably the 1998): Rio Ivaı (Ordovician–), Parana most voluminous continuous dune field in Earth (), Gondwana I ( to Lower history (Scherer & Goldberg, 2007; Fig. 1). The ), Gondwana II (Triassic), Gondwana III Botucatu sand sea forms part of the intracratonic ( to Lower Cretaceous) and Bauru Parana Basin-fill succession and represents the (). The Rio Ivaı, Parana and last phase of sedimentation in the Parana Basin Gondwana I supersequences represent marine prior to break-up of the South American and and shoreline deposits. The Gondwana II, Gond- African continents. The Parana Basin developed wana III and Bauru supersequences comprise ter- across a wide range of different pre- restrial sediments and associated volcanic strata. basement terranes (Palaeoproteorozoic to Cam- The Botucatu Formation and the Serra Geral For- brian) with differing ages and compositions. mation form part of the Gondwana III superse- Consequently, when combined with previous quence, also known as the Sao~ Bento Group work on palaeowind directions, it should be (Schneider et al., 1973). The Botucatu palaeoerg possible to distinguish different provenance sig- is overlain by 700 to 1700 m of lava-flows from natures in the Botucatu Formation. Previous the Parana–Etendeka Large Igneous Province studies of the formation (e.g. Bigarella & Sala- (Serra Geral Formation). Interbedded basaltic muni, 1961) show that coarser sediment prevails lava flows and aeolian sediments indicate co- in the southern part of the basin suggesting existence between, and a transition from, aeolian

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 3

Fig. 1. Botucatu desert extent in South America within the Parana and Chaco-Parana Basins with sample locations. The desert had a bimodal wind pattern, with southerly directions in low latitudes and north-east/east directions in the study (Scherer, 2000; Scherer & Goldberg, 2007). sedimentation to lava emplacement (Scherer, Formation (Amarante et al., 2019). The Botucatu 2002). Formation also has equivalent units in Argen- The age of the top of the Botucatu Formation tina (San Cristobal Formation), Paraguay (Mis- is constrained by overlying lavas of the Serra siones Formation), Bolivia (Ichoa Formation) Geral Formation, which have yielded ages of ca and Namibia (Twyfelfontain Formation) (Scherer 134 Ma (Valagininan) by Ar/Ar (whole-rock & Goldberg, 2007). Outcrops in Rio Grande do grains/plagioclase) and U-Pb methods (Zircon Sul State and Uruguay likely represent the and Badeleeite) (Ernesto et al., 1999; Thiede & upwind margin of the erg as illustrated by: (i) Vasconcelos, 2010; Janasi et al., 2011; Pinto variations in thickness from 0 to 100 m along et al., 2011; Baksi, 2018; Rossetti et al., 2018). the southern margin to up to 400 m thickness in The determination of the onset of Botucatu the basin centre (Milani et al.,1998); and (ii) deposition is imprecise due to a lack of dated palaeocurrent data demonstrating a consistent interbedded volcanics and/or palaeontological north-east/east palaeowind direction along the constraints. The Botucatu Formation overlies the southern basin margin (Scherer & Goldberg, Upper Jurassic Guara and Batovi Member (Soto 2007; Fig 1). & Perea, 2008; Perea et al., 2009; Francischini Scherer (2000) described four facies across the et al., 2015), suggesting an earliest Cretaceous southern Botucatu margin: thick, large-scale age ( to ) for palaeoerg trough cross-stratification; large-scale planar accumulation. cross-stratification; medium-scale trough cross- The study area is located along the southern stratification; and low-angle cross-stratification. margin of the Parana Basin, in Rio Grande do The relatively monotonous and simple sedimen- Sul State (Brazil) and Central Uruguay. In Uru- tary facies are controlled by dune morphology, guay, the Botucatu Formation is equivalent to illustrating a consistent predominance of the Rivera Member of the Tacuarembo crescentic dunes and linear draa bedforms.

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 4 G. Bertolini et al.

Fig. 2. Parana Basin stratigraphic chart with the six supersequences: Rio Ivaı; Parana; Gondwana I; Gondwana II; Gondwana III; Bauru (adapted from Milani et al., 2007).

The Botucatu is considered to represent a Piramboia Formation is a wet aeolian dry-aeolian system (sensu Havholm & Kocurek, system, composed of red fine to coarse sand- 1994), indicated by the absence of deposits or stones (Dias & Scherer, 2008). sedimentary features such as wet interdune The Parana Basin basement framework com- deposits or adhesion structures generated by prises a of complex Cambrian to Palaeo- water-table or flooded/wet interdune areas. rocks forming the following terranes In the study area, the Botucatu Formation (from east to west): Dom Feliciano Belt, overlies three distinct units of the Gondwana I Tacuarambo Terrane (Sul-Rio-Grandense and II supersequences of the Parana Basin-fill: Shield), Nico Perez and Piedra Alta Terranes (i) the Jurassic Guara Formation and Batovi (Uruguayan Shield). For simplification, the over- Member of Tacuarembo Formation (Uruguay) in all crustal framework displays a distribution of the west; (ii) the Triassic Caturrita Formation in Neoproterozoic ages in the east (Dom Feliciano the central region; and (iii) the Permian to Juras- Belt) and Palaeoproterozoic age terranes in the sic Piramboia Formation in the east. The Guara west (Tacuarembo, Nico Perez and Piedra Alta Formation (Batovi Member) is an Upper Jurassic Terranes) (Fig. 11). The Dom Feliciano Belt fluvial–aeolian unit, comprising a south/south- forms part of the central to east Rio-Grandense west directed distributive fluvial system and Shield and includes three main orogenic events: small to medium size dunes (Scherer & Lavina, Passinho (0.9 to 0.86 Ga), Sao~ Gabriel (0.77 to 2006; Amarante et al., 2019). The Triassic Catur- 0.68 Ga) and Dom Feliciano (0.65 to 0.54 Ga) rita Formation comprises amalgamated sand- (Philipp et al., 2016). Additionally, the Dom stones intercalated with siltstones that represent Feliciano terrane includes a Neoproterozoic– high energy, low sinuosity river deposits (Zer- Cambrian volcano-sedimentary succession fass et al., 2003; Jenisch et al., 2017). The (Camaqua~ Basin) and a Neoproterozoic

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 5 metamorphic belt (Porongos Belt) (Janikian Formation lavas (intratrap sandstones) to et al., 2012; Hofig€ et al., 2018). The granitic– observe possible lateral and chronological gneissic Tacuarembo, Piedra Alta and Nico changes in provenance signal. In the studied Perez Terranes are located on the Western Sul- area, the Botucatu Formation outcrop ranges Rio-Grandense and Uruguayan Shields. These from 0.5 to 30.0 m in thickness and mostly com- terranes yield Orosirian to ages, with prises fine to medium sand. The predominant peaks from 1.8 to 2.2 Ga (Oyhantcßabal et al., sedimentary facies are large-scale cross-beds, 2018a,b). In general, there is a westward with occasional low-angle beds, trough cross- increase in the age of terranes; however, excep- bed sets and locally massive sandstones. Scherer tions occur like the Neoproterozoic magmatism (2000) considered these facies to represent parts in both the Tacuarembo and Nico Perez blocks of aeolian dune morphologies (simple to com- (Oyhantcßabal et al., 2018a,b), and Palaeoprotero- pound crescentic dunes and complex linear zoic successions are present as the Rhyacian Rio draas) with very local ephemeral fluvial stream dos Ratos Complex and Encantadas Block in the deposits. Table 1 documents the sedimentary east (Gregory et al., 2015). facies as well as the sample number, location, stratigraphic level (base, top and interbedded) and the provenance techniques performed. The MATERIALS AND METHODS samples were collected mostly from large scale cross-stratified and trough-stratified sandstones To constrain the provenance of the Botucatu (Ste), with some from low-angle stratified (Sle) sandstones in the south-western part of the and massive sandstones (Sm). Figure 3 shows Parana Basin, sedimentary petrography, heavy the 11 studied profiles (based on Scherer, 2000) mineral analysis, grain-size analysis and detrital and the position of each sample within the zircon geochronology were undertaken. Twenty- stratigraphy. one samples (Fig. 1) were collected along the Sample preparation was undertaken separately southern margin of the basin from sandstones at for each analysis. For granulometric and heavy the base and top of the formation, together with mineral analysis, disaggregation was carefully sandstones from within the Serra Geral undertaken with a rubber point pestle and

Table 1. Sample locations by stratigraphic position and location: Heavy Mineral (HM); Bulk Petrography (BP); Detrital Zircon (DZ); Granulometric Analysis (GR).

Provenance techniques Sample Facies Stratigraphic Coordinates Region code code level Latitude/Longitude sampled BP DZ HM GR

PB-01 Sle Base À50.482903/À29.886139 East X PB-02A Ste Top À50.395787/À29.877412 East X X X X PB-02C Sm Intratrap À50.395787/À29.877412 East X X PB-03 Ste Top À50.529168/À29.830188 East X PB-09 Ste Top À50.577807/À29.743651 East X X X PB-18 Ste Top À52.414231/À29.726391 East X X X X PB-31B Ste Intratrap À51.131414/À29.632367 East X X X PB-19 Sm Intratrap À52.634681/À29.55371 Central X PB-20 Sm Intratrap À52.638663/À29.261931 Central X X X PB-21 Ste Top À52.631227/À29.566123 Central X X X PB-22 Ste Intratrap À53.0209/À29.406328 Central X X X X PB-23 Ste Base À54.68608/À29.481004 Central X X X X PB-24 Ste Top À54.749004/À29.365739 Central X X X PB-29 Ste Intratrap À54.764559/À29.36278 Centra X PB-04 Sm Base À55.412019/À30.717301 West X X X PB-10 Ste Base À55.72246/À31.564436 West X X X X PB-11 Ste Top À55.897349/À31.147843 West X X X PB-12 Ste Intratrap À55.913833/À31.142028 West X X X PB-14 Sle Top À56.216685/À30.486345 West X X PB-16 Ste Top À55.517067/À30.852035 West X PB-17 Ste Base À55.349101/À30.82337 West X

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 6 G. Bertolini et al.

Fig. 3. Sedimentary profiles (adapted from Scherer, 2000) with the stratigraphic position of the sampling where the letters a to k refer to the profile locations: a – Tacuarembo (Uy); b – Tranqueras (Uy); c – Santana do Livra- mento (Br); d – Jaguari (Br); e – Sao~ Pedro do Sul (Br); f – Santa Maria (Br); g – Sobradinho (Br); h – Herveiras (Br); i – Ivoti (Br); j – Santo Antonio^ da Patrulha (Br); k – Osorio (Br). mortar to avoid grain breakage. For geochrono- A complementary elemental analysis was under- logical studies, sample disaggregation was car- taken with an energy-dispersive X-ray spectrome- ried out using a mechanical jaw crusher (when ter (EDS) on a scanning electron microscope necessary) and a plate mill. (SEM) to aid with heavy mineral identification. Nineteen petrographic thin sections were The analysis was carried out at the Universidade described and subject to point counting Federal do Rio Grande do Sul using a Jeol 6610- (300 points per sample) using the Gazzi and LV (15 kV beam voltage and 12 mm working dis- Dickinson method (Ingersoll et al., 1984). The tance; Jeol Limited, Tokyo, Japan). results are displayed following the scheme of Detrital zircon grains were analyzed from 10 Garzanti et al. (2018) (Fig. 4). samples for U-Pb at the University of Sao~ Paulo, Grain-size analyses were undertaken on 14 sam- using a Finnigan Neptune mass spectrometer and ples, with 100 g of each sandstone. The process an excimer ArF laser (k =193 nm) ablation system consisted of sample cleaning (removal of external (Thermo Fisher Scientific, Waltham, MA, USA). exposed surfaces), disaggregation, drying and The laser ablation – multi-collector – inductively sieving. The petrographic data reveal no silt sized coupled plasma – mass spectrometry (LA-MC-ICP- grains, justifying the sand-class interval measur- MS) uses 5 mJ, 6 Hz and 29 lm diameter beam as ing (63 to 125 lm, 125 to 250 lm, 250 to 500 lm analytical conditions. The207Pb/206Pb ages were and 500 to 1000 lm). Samples were weighed after used for zircons older than 1.3 Ga and the each step and the difference recalculated and 206Pb/238U ages for younger grains. The discor- assigned to each respective grain-size proportion. dance tolerance is 10%; values above this limit Between 150 g and 400 g of 14 samples were were dismissed. The parameters used are selected for heavy mineral analysis. Samples discussed in Sato et al. (2011). were dried and sieved (63 to 125 lm) following The GRADISTAT software (Blott & Pye, 2001) the grain-size window interval of Morton & was used for statistical analysis and to deter- Hallsworth (1994) via dense liquid separation mine sediment distribution. The Provenance R À (Bromoform À2.80 g ml 1). The mainly fine sand Package (Vermeesch et al., 2016) was used to sediment framework composition of the Botu- run the cumulative age densities (CAD), kernel catu Formation is compatible with hydraulic densities estimators and Multidimensional Scal- equivalence of heavy mineral assemblage, ruling ing (MDS) plots. out significant hydraulic sorting bias effects (e.g. Morton & Hallsworth, 1994; Hurst et al., 2017). Heavy mineral concentrates were mounted on RESULTS petrographic slides, with identification of 150 translucent minerals using the ribbon technique Petrography reveals a mean composition of following the method of Mange & Maurer (2012). Qt89F8L3, compositionally classified (Garzanti

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 7

Fig. 4. Sandstone classification scheme (Garzanti et al., 2018) and lithics plot. Samples point to quartzose and fel- sic plutonic compositions. (Q – Total quartz; F – K-feldspar and plagioclase; L – total lithics; Lv – volcanic rock fragments; Lp – Plutonic lithics; Lm and Ls – metamorphic and sedimentary lithics; qQ – pure-quartzose; Q – Quartzose; qFQ – quartz-rich feldspatho-Quartzose).

et al., 2018) as pure-quartzose (PB-04 and PB- ZTR and very local occurrences of garnet, epi- 23), quartzose (PB-01, PB-03, PB-10, PB-16, PB- dote and pyrolusite. The heavy mineral analysis 17, PB-18, PB-19, PB-21, PB-22, PB-24 and presents a mean ZTR0.84, ranging from 33.5 to PB-29A) and quartz-rich feldspatho-quartzose 99.4% and is characterized by the predominance (PB-2A, PB-2C, PB-09 and PB-20) (Fig. 4) sand- of zircon, reaching almost 98% in sample PB-24. stones. Monocrystalline quartz is predominant The area contains zircons that display different (93 to 100%) over polycrystalline quartz. K-feld- levels of roundness, varying from rounded oval spar is most frequent over plagioclase (mean grains to elongate bi-pyramidal prismatic grains. 70%), however centre region samples show an Samples PB-2A, PB-31B and PB-20 show peaks increase in plagioclase (60 to 70%). The lithic in Fe-garnet and epidote, and sample PB-09 fragment distribution (Fig. 4B) shows both contains a high abundance of pyrolusite. igneous (felsic plutonic) and volcanic (basaltic Minerals such as corundum and sulphides, lathwork to glassy trachyte) fragments. In the were identified with the aid of the SEM and the eastern and central regions moderately sorted latter in particular may represent authigenic fine sand predominates with poorly sorted med- phases. ium sand in the west. Detrital zircon age spectra display a complex The heavy mineral population of the 21 stud- signature, with ages ranging from 155 Ma to ied samples comprises 14 mineral species, with 3290 Ma (Fig. 6). The wide age distribution a dominant signal from resistant species – zir- reflects the complex geological history of the con, rutile and tourmaline. Local occurrences of basement rocks and potential recycling from anomalously high amounts of garnet, epidote, older sedimentary units. In order to track pyrolusite and apatite, and a few grains of other possible sediment pathways using the detrital species such as alumino-silicates (kyanite, anda- zircon (DZ) record, the ages were separated lusite or sillimanite), spinel, corundum, mon- into seven groups that correspond to signifi- azite, fluorite, xenotime, sulphides (cassiterite cant crustal construction periods in South and pyrite) are also recorded. Figure 5 shows America (Bahlburg et al., 2009): Late Palaeo- the heavy mineral assemblage distribution zoic to Mesozoic (Gondwanides Cycle), Devo- across the studied region, illustrating the high nian to Cambrian (Famatian Cycle), Cambrian

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 8 G. Bertolini et al.

Fig. 5. Heavy mineral suite distribution across the regions, with prevalent zircon, tourmaline and rutile (ZTR) assemblage.

to Neoproterozoic (Brasiliano/Pampean/Pan-African Geral lavas are also shown in Fig. 3. The prove- cycles), Tonian to Stenian (Grenvilian Cycle), nance dataset reveals a lateral variation Mid- to , Orosirian to described in further detail below. Rhyacian and to Palaeoarchean (Fig. 5). Eastern region The detrital zircon data indicate that Early Neoproterozoic to Cambrian ages (515 to Three sections were sampled in the eastern 650 Ma) are the main contributor to the DZ pop- region: Santo Antonio^ da Patrulha (PB-03 and ulation with Late Palaeozoic to Mid Mesozoic PB-09), Osorio (PB-01, PB-02A and PB-02C) and (150 to 359 Ma), Tonian–Stenian (0.9 to 1.25 Ga) Ivoti (PB-31B) (see Fig. 3). Bulk petrography and Orosirian to Rhyiacian ages (1.8 to 2.25 Ga) shows a mean composition of Qt85F11L4, with a forming subsidiary contributions. The detrital relative change in feldspar content (5 to 15%). zircon signal appears to be relatively homoge- The granulometric analysis reveals a unimodal neous but grouping the samples by region distribution, fine-sand, moderately (PB-02A and reveals a transitional signature across the south- PB-31B) to moderately well-sorted (PB-09). The ern margin of the Botucatu desert, with increas- heavy mineral analyses (Fig. 5) revealed that ing Cambrian to Neoproterozoic ages (31.6 to samples PB-02A and PB-31B contain 54.3% and 52.5%) and a decrease in ages older than 12.5% of almandine, respectively, PB-02A con- 900 Ma towards the east. tains 11.5% epidote and sample PB-20 com- The geographic location of each sample and a prises 55.4% pyrolusite. Detrital zircon dating compilation of the different dataset types are reveals an age signature ranging from 155 Ma to illustrated in Fig. 7. Uranium–lead (U-Pb) detri- 3231 Ma, with main peaks in the Late Palaeo- tal zircon ages are plotted using a probability zoic to Mid Mesozoic (11%), Neoproterozoic to density plot, the bulk petrography, heavy min- Cambrian (52%) and Tonian to Stenian (10%). eral and granulometric analyses are illustrated Eastern region samples show a significant using pie charts. The sampling level in the upward increase in the proportion of C-NP zir- stratigraphy and location relative to the Serra cons from 49% at the base (PB-02A) to 72% in

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 9

Fig. 6. Normalized probability density plots located on the schematic stratigraphic profile. The coloured areas demonstrate the most prominent age intervals. interbedded sandstones (PB-31B) at the top of feldspars from 0 to 10% and total lithics from 0 the Botucatu Formation. to 6%. Grain-size analysis reveals a unimodal distribution, moderately well (PB-18, PB-20 and PB-24) to moderately sorted (PB-21, PB-22 and Central region PB-23), fine-sand to medium-sand (PB-23). The central portion samples were collected from Heavy mineral analysis (Fig. 5) shows a variable five sections (Fig. 3): Jaguari (PB-23, PB-24 and ZTR value, ranging from 51.5 to 99.5%, with a PB-23), Sao~ Pedro do Sul (PB-25), Santa Maria garnet peak (30.3%) occurring in sample PB-20. (PB-26 and PB-27), Sobradinho (PB-22) and Her- Samples 18, 21, 22, 23 and 24 have high zircon, veiras (PB-18, PB-21, PB-19 and PB-20). Petro- tourmaline and rutile (ZTR) values, reaching graphy reveals a mean composition of Qt89F7L3, >95% zircon in sample 24. Iron-garnet and epi- classified as quartz-enriched feldspatho-quart- dote are observed in sample 20 (>35%). Detrital zose and quartzose sandstones. Bulk petrogra- zircon ages range from 174 to 3290 Ma, includ- phy shows total quartz varying from 83 to 99%, ing Early Neoproterozoic to Cambrian (41%),

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 10 G. Bertolini et al.

Fig. 7. (A) Geological map (adapted from Wildner et al., 2006) with the sampling spots. (B) Provenance data compilation.

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 11

Tonian–Stenian (12%) and Orosirian–Rhyacian zircons form the main component of the popula- (10%) peaks. tion with the Tonian–Stenian (0.9 to 1.25 Ga) and Orosirian–Rhyacian (1.80 to 2.25 Ga) pro- viding subsidiary contributions in all regions. Western region More detailed analysis allows lateral variations Seven samples were taken from the western part in age groups to be distinguished. Variations in of the study area from three sections (Fig. 3): detrital zircon age distributions across the Tacuarembo (PB-13), Tranqueras (PB-10, PB-11 regions are compiled in Table 2 and demon- and PB-12) and Santana do Livramento (PB-4, strate a relative increase in C-NP ages in the PB-5, PB-14, PB-15, PB-16 and PB-17) in Uruguay eastern region, with a predominance of >900 Ma and Brazil. Petrographically the samples have a grains in the western region. The central region predominantly quartzose composition (mean is transitional between both regions containing Qt94F5L0). The granulometric analysis shows uni- intermediate values in all age periods. modal sorting, moderate–well (PB-11), moder- The lateral variations described above are likely ately (PB-04 and PB-10) to poorly sorted (PB-12 to be related to changes in either basement units and PB-14A), medium (PB-04, PB-11 and PB- immediately adjacent to the outcrop or to recy- 14A) to fine sand (PB-10 and PB-12). Heavy min- cling of underlying sedimentary successions, or a eral analysis (Fig. 5) reveals high ZTR concentra- combination of both. These variations are also tion (83 to 96%) with zircons dominant. The expressed in Fig. 8 as cumulative density plots region has a higher than average rutile content (A) and multidimensional scaling plots (B). The (13%) contrasting with a mean of 5% from other CDP demonstrates that the C-NP is the most vari- regions. Aluminium–silicon (Al-Si) polymorphs able detrital zircon age group from east to west, (mean 5%) appear consistently in the west but with almost 20% variation, whilst the MDS plots only occur occasionally in the east and central illustrate the transitional characteristics of the lat- regions. Garnet (Fe-Almandine) appears only in eral changes. The average compositional dataset sample PB-04. Detrital zircon grains from sam- (granulometry, petrography and heavy minerals) ples PB-10, PB-11 and PB-12 display ages ranging also displays a lateral modification from west to from 155 to 2717 Ma, with mainly Early Neopro- east. The petrography shows an increase of lithics terozoic to Cambrian (32%), Tonian to Stenian and feldspar over quartz components (Q94-5F5- (18%) and Orosirian to Rhyacian (14%) peaks. 10L0-5), and the granulometry demonstrates a medium to very fine sand transition. The heavy mineral data present an increase of the mean ZTR DISCUSSION from 80 to 89% (east to west). The western region also shows an increase of Al-Si minerals, suggest- ing greater influence of metamorphic sources. Lateral variations In the study area, the Parana Basin is devel- The detrital zircon geochronology dataset illus- oped above a basement composed of Neoarchean trates that Cambrian–Neoproterozoic (C-NP) aged to Cambrian rocks. The contact between

Table 2. Detrital zircon age proportions from the southern margin of the Botucatu Formation organized by regions.

%

Event Time interval (Ga) East Centre West

Late Palaeozoic to Mid Mesozoic 0.150–0.359 11.027 9.695 9.294 Early Famatian (Ordovician–Silurian) 0.359–0.515 8.365 6.648 5.948 Early Neoproterozoic to Cambrian Ages (Late Brasiliano) 0.515–0.650 52.471 40.720 31.599 Early Brasiliano 0.650–0.900 9.886 9.418 7.807 Tonian–Stenian Ages (Grenvillian Cycle) 0.900–1.250 9.886 11.634 18.216 Mid-Mesoproterozoic to Statherian ages 1.250–1.800 3.422 6.648 9.665 Orosirian to Rhyacian ages 1.800–2.200 2.662 9.695 12.639 Neoarchean to Palaeoarchean 2.200–3.400 2.281 5.540 4.833

The data show an overall increase in >900 Ma ages to the west and Brasiliano peak to the east.

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 12 G. Bertolini et al.

Fig. 8. Cumulative probability plot (A) and multidimensional scaling (MDS) plots (B) illustrating the lateral vari- ability of detrital zircon U-Pb ages. Western region is enriched in older ages (>1.25 Ga) whilst the eastern region is dominated by Early Brasiliano ages (515 to 620 Ma). The central portion of the basin is transitional between the two. The bold central lines in plot (A) correspond to the sum of values for each region.

basement and basin-fill occurs beneath Permian C-NP ages in the eastern region (48.8 to 71.6%) units (Itarare and Rio Bonito formations), as could indicate increased direct input from these such there is no direct contact between base- source terrains at this location. ment and the Botucatu Formation. In addition, A lack of direct first cycle input into the the basal contact of the Botucatu changes later- majority of the Botacatu Formation is indicated ally, from the Upper Jurassic Guara Formation/ by the limited presence of detrital zircon grains Rivera Member (west), Triassic Caturrita Forma- of Jurassic age (180 to 155 Ma). These grains tion (central) and Permian Piramboia Formation occur only in the extreme west of the Botucatu (east). No detrital zircon age data are available exposure (Central Argentina) where they are from these underlying units, so it is not possible considered to be derived directly from the to ascertain whether zircons were derived Andean magmatic arc (Peri et al., 2016). The directly from these strata. Zircon ages from the Tonian–Stenian ages (Grenvilian Cycle) are not underlying basement also show distinct age present in basement terranes close to the study domains across the margin. The Rio-Grandense area and could indicate an original source in and Uruguayan Shields contain zircon popula- areas such as the Sierra Pampeanas (Argentina) tions of 500 to 900 Ma (Cambrian–Neoprotero- and Natal–Namaqua Belt (south-west Africa) zoic contribution) and 1250 to 3400 Ma (Casquet et al., 2008; Rapela et al., 2010; Bial (including Orosirian–Rhiacian subsidiary contri- et al., 2015). The relative increase in zircons bution) (Philipp et al., 2016). Tectonic domains derived from strata that yielded Tonian–Stenian such as the Piedra Alta, Nico Perez and ages in the western region together with the pre- Tacuarembo terranes, yielded multiple Neopro- vailing north-east wind pattern could indicate terozoic to Neoarchean ages, but overall ages are that first-cycle grains were derived from the composed predominantly of >1.25 Ga rocks south-west Gondwana margin terranes (Sierra (Santos et al., 2003; Oyhantcßabal et al., 2012; Pampeanas). Canile et al. (2016) show that juve- Oriolo et al., 2016). The Dom Feliciano Belt in nile zircons [Positive ƐHf(t)] from the Natal– Rio Grande do Sul State and the coastal region Namaqua belt (in present day Namibia and of Uruguay comprises a Cambrian to Neoprotero- South Africa) are present in older Parana zoic granite belt (Basei et al., 2008; Philipp Basin units, demonstrating the existence of pre- et al., 2016). The progressive upward increase in vious recycling. The Tonian–Stenian age

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 13 signature probably records the presence of recy- Congo River sediment described by Garzanti cled grains from a mixture of Natal–Namaqua et al. (2019), suggesting that Botucatu sediment Belt and Western Sierra Pampeanas sources. has a predominant poly-cyclic history. To understand the poly-cyclic sediment history of the Botucatu sediment and its relationship with Sedimentary recycling Parana Basin Carboniferous–Jurassic units, the Uranium–lead (U-Pb) detrital zircon dating is a detrital zircon data presented here is compared useful tool that allows the resulting age spectra with other datasets from older strata in the from a sedimentary rock to be linked potentially southern Parana Basin. to an original sediment source (Gehrels, 2014). Cumulative age density and multidimensional However, due to the resistance of zircon to scaling biplots (Fig. 9A and B) are utilized to transport, burial and even metamorphic events, assess and compare the DZ populations present detrital zircon age spectra commonly include in the Carboniferous Itarare Group (Rio do Sul grains recycled from older strata such that the Formation), Carboniferous–Permian Guata final detrital zircon age spectra within a sedi- (Palermo, Sideropolis, Paraguacu and Triunfo mentary rock represent a composite of first-cycle formations), Permian Passa Dois Group (Irati, and recycled grains (Dickinson & Gehrels, 2009; Serra Alta, Teresina, Morro Pelado and Serrinha Hadlari et al., 2015). As a consequence, if solely formations), Triassic Santa Maria formation utilizing the detrital zircon dataset it is virtually (Santa Maria and Santa Catarina region – based impossible to establish whether the age peaks in on Canile et al., 2016; Philipp et al., 2018). The the Botucatu are derived directly from basement MDS biplot (Fig. 9B) shows a distribution of (first-cycle grains) or by multiple phases of units enriched in Palaeoproterozoic sources (Rio deposition and reworking (poly-cycling grains). Bonito and Palermo formations), Permian–Trias- By combining the detrital zircon data with pet- sic ages (Irati, Serra Alta and Teresina forma- rographic and heavy mineral data it is possible tions), Cambrian–Neoproterozoic ages (Triassic to more precisely determine the sediment origin. and Eastern Botucatu) and a zone of mixed ages The petrographic and heavy mineral dataset (centre, west, north Botucatu, Itare, Serrinha and shows a highly quartzose and ZTR-rich compo- Triunfo formations) can be defined. The MDS sition. A quartzose sedimentary factory requires shows that the Upper Carboniferous–Lower Per- multiple steps of diagenesis, weathering and mian units (Rio Bonito and Palermo formations) recycling in order to obtain a pure-quartzose correspond to a western Botucatu signature, sug- and high ZTR end-member (for example, the gesting that these units were the primary source.

Fig. 9. Parana Basin polycyclic nature revealed by cumulative age densities (CAD) (A) and multidimensional scal- ing (MDS) plots (B), demonstrating an overlapping of detrital zircon ages from Carboniferous to Cretaceous strata (data from Canile et al., 2016; Philipp et al., 2018).

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 14 G. Bertolini et al.

The conglomeratic end-member of the Triassic of Fe-Garnet, epidote and Mn oxides (samples unit (Santa Cruz Sequence – Philipp et al., PB-31B, PB-20, PB-2A and PB-09). Iron–Garnet 2018) contains an assemblage enriched in Cam- and epidote are common heavy mineral consti- brian–Neoproterozoic ages, whilst finer grained tuents. Manganese oxides, identified by SEM components display a complex Permian to analysis, are common in Serra Geral lavas, and Palaeoproterozoic age distribution. The similar- their presence in the Botucatu is indicative of ity between the conglomerate end-members (Tri- an authigenic phase. The heavy mineral and assic) and eastern Botucatu samples (MDS and petrographic composition of the Botucatu For- CAD) suggests that eastern Botucatu samples mation is considered to be representative of the contain a relatively smaller amount of recycled original sediment source composition with little material in comparison to central/west samples, subsequent modification by burial diagenesis especially in intratrap (PB-31B) samples. and weathering, the reasons for which are dis- cussed below. Morton & Hallsworth (2007) described the Controls on heavy mineral assemblages effects of burial related dissolution on heavy The MDS biplot of the heavy mineral data mineral assemblages. These authors demon- (Fig. 10) defines two distinct clusters: (i) a well- strated that epidote and garnet remain stable defined group, referred to here as the average until respectively 1.0 km and 3.5 km in the Botucatu composition; and (ii) isolated samples North Sea Basin. Ando et al. (2012), in a study of referred to as local detrital component. The cores from Bengal Bay, showed that the progres- average Botucatu composition includes the sive dissolution of amphibole occurred at burial majority of studied samples and has a mean depths of up to 1.5 km and for epidote at depths composition of ZTR0.9, with samples distributed of up to 2.5 km, transitioning to a garnet and throughout the basin, irrespective of strati- ZTR assemblage below 2.5 km depth. The Botu- graphic level (basal, top or interbedded sand- catu burial history suggests relatively shallow stone) or the sampling region (west, centre or burial of up to 1.5 km in the basin depocentre in east). The local detrital component has a mean the north (Milani et al., 2007). The thinner lava- composition of ZTR0.6 related primarily to peaks pile and absence of upper-basalt sedimentary deposits along the southern margin, points to a relatively shallower burial depth of 700 to 800 m (lava-pile thickness in the south). Limited burial suggests that the presence of garnet and epidote is not controlled by burial related dissolution but likely due to provenance controls. The common presence of micro-porosity and mouldic porosity (on lithics and feldspars) and feldspar dissolution may suggest a degree of modification through surface weathering of out- crops. However, Francßa et al. (2003) describes feldspar dissolution in well cores, such that the process also occurs at depth. The sampling occurs mostly on roadside outcrops in similar climatic conditions (humid and hot) across the studied area. If differential dissolution was the driving mechanism that generated local compo- sitional variations, intermediate heavy mineral compositions would be expected with highly weathered/dissolved garnet and epidotes in ZTR enriched samples, which does not occur. Thus, the local variations apparently are not controlled by differential surface weathering, indicating Fig. 10. Multidimensional scaling (MDS) biplot on that the heavy mineral and petrographic changes heavy mineral dataset demonstrates an average compo- are depositional. sition [zircon, tourmaline and rutile (ZTR) rich] and Reconstructions of desert depositional systems local detrital input (garnet, epidote and pyrolusite). (e.g. Pye & Tsoar, 2008) illustrate the coexistence

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 15

Fig. 11. Source to sink map demonstrating the impact of surrounding basement and external sources on the Botu- catu detrital zircon (DZ) composition. Pie-charts display the mean petrography and heavy mineral compositions, mirroring a lateral variation with DZ and granulometric data. At local level, fluvial inputs generate heavy mineral anomalous compositions. Data for the basement are from Saalmann et al. (2011) and the desert extent is from Scherer & Goldberg (2007). of different depositional environments such as reworked laterally by aeolian processes. Aeolian rivers, alluvial fans and deltas that could pro- systems promote an increase in grain roundness, vide sediment to a dune field. An erg margin, as but at the same time maintain stable heavy min- in the study area, is prone to different styles of eral proportions (Resentini et al., 2018). interaction between depositional systems, partic- ularly at the fluvial–aeolian interface (Al- Multiscale desert filling Masrahy & Mountney, 2015). Whilst there are very few preserved records of waterlain deposits Despite the similarities in the petrographic, in the dry–aeolian Botucatu system, fluvial input heavy mineral and detrital zircon compositions, is likely to have occurred around the erg mar- subtle differences are present in the dataset, gins with subsequent aeolian reworking remov- indicating the presence of additional controls ing that sedimentary signature. Provenance on sandstone composition. The data set reveals studies from present day deserts such as Takla- two important characteristics: (i) lateral varia- makann (Rittner et al., 2016), Arabia (Garzanti tions in detrital zircon age spectra; and (ii) dif- et al., 2013) and Namibia (Garzanti et al., 2012), ferences in heavy mineral and petrographic data demonstrate that shifts in provenance can be that suggest local sediment input points. The correlated with fluvial, alluvial fan or deltaic lateral provenance variations (see Lateral varia- inputs. The localized heavy mineral anomalies tions section) are illustrated on Fig. 11, demon- (Fig. 5) concomitant with increase in quartz strating that the detrital zircon, granulometry, content suggest a point-sourced alluvial system petrography and heavy mineral assemblages

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 16 G. Bertolini et al. change from east to west. Additionally, the fig- statistical tools (multidimensional scaling, ure shows the underlying units (Parana Basin) cumulative density plots and density plots) and basement terranes that compose the Sul- indicate that: Rio-Grandense and Uruguayan shields. As dis- cussed in the Sedimentary recycling section, the 1 Cambrian to Neoproterozoic ages (515 to detrital zircon composition of Botucatu and 650 Ma) are the main detrital zircon component, Parana Basin units displays a statistical similar- with subsidiary contributions from Orosirian to ity, suggesting that underlying units represent Rhyacian (1.8 to 2.2 Ga) and Tonian to Stenian the final source for the Botucatu Desert. Fig- ages (0.9 to 1.25 Ga). ure 11 displays the overlap of the lateral varia- 2 A progressive lateral change in detrital zir- tions and underlying units of Botucatu (Guara con ages, together with changes in the underly- Formation, Santa Maria Formation and Pir- ing basin stratigraphy and basement domains, amboia Formation) suggesting these units as the suggests a polycyclic provenance. most feasible sediment source. Furthermore, 3 The eastern region records an upward Fig. 11 illustrates how ultimately the basement increase of Cambrian to Neoproterozoic detrital age framework contributes to the lateral detrital zircons; the central and western regions show zircon variation. The prevalent >900 Ma ter- no significant changes. ranes in the west (Tacuarembo, Nico Perez and 4 There is a progressive downwind decrease Piedra Alta) contrasting with the C-Np ages in in grain size from medium to very fine sand over the east (Dom Feliciano Belt) probably con- a distance of 600 km from downwind. tributed to the Ordovician to Jurassic Parana 5 An average quartz–zircon, tourmaline and Basin filling, later being reworked for Botucatu. rutile (ZTR) enriched composition predomi- The grain size, heavy mineral and petrography nates along the southern Botucatu margin; vari- partly corroborate the lateral changes in the ations allow recognition of local sediment detrital zircon population, supporting the notion input points within an otherwise homogenous that reworking of rocks underlying the Parana aeolian system which likely relate to local flu- Basin was the main cause of the lateral varia- vial systems developed along the erg margin. tion. The large-scale control (over more than 600 km along the basin margin) shows that aeo- lian processes were not able to completely ACKNOWLEDGEMENTS homogenize the overall composition of the sand supplied to the Botucatu erg. The authors gratefully acknowledge support The anomalous heavy mineral values charac- from Shell Brasil through the ‘BG05: UoA- terize local sedimentary input; however, it is not UFRGS-SWB Sedimentary Basins’ project at clear whether this input was sourced directly UFRGS and the strategic importance given by from basement material, reworked from older, ANP through the R&D levy regulation. The underlying sedimentary successions or resulted authors would like to thank Renan Guilherme from a combination of both. Figure 11 illustrates de Souza and Adriano Reis for help in the fluvial systems draining Parana Basin units Tacuarembo/Uy Field Trip. This work is part of and possible basement terranes. The evidence the current dual-degree PhD from Gabriel Ber- for local sediment input supports the idea of a tolini at Universidade Federal do Rio Grande do number of distinct sediment pathways into the Sul and University of Aberdeen. We thank the Botucatu erg, resulting in distinct, identifiable editors, and Dr Eduardo Garzanti, Dr Luca compositional signatures. Caracciolo and Dr Sebastian Oriolo for their kind and constructive reviews. We thank to Conselho Nacional de Desenvolvimento Cientı´- CONCLUSIONS fico e Tecnologico (CNPq) by the financial sup- port (Grant: 203786/2017-3). Provenance analysis of the aeolian sediments of the Botucatu Formation along the southern margin of the Parana Basin has been under- DATA AVAILABILITY STATEMENT taken to provide insights into depositional con- trols. The provenance dataset of detrital zircon The data that supports the findings of this study U-Pb dating, heavy mineral analysis, petrogra- are available in the supplementary material of phy and granulometric analysis using multi- this article (Tables S1 to S4).

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 17

REFERENCES (Parana Basin), southern Brazil. J. S. Am. Earth Sci., 25, 405–415. Dickinson, W.R. and Gehrels, G.E. (2003) U-Pb ages of Al-Masrahy, M.A. and Mountney, N.P. (2015) A detrital zircons from Permian and Jurassic eolian classification scheme for fluvial-aeolian system interaction sandstones of the Colorado Plateau, USA: paleogeographic in desert-margin settings. Aeol. Res., 17,67–88. implications. Sed. Geol., 163,29–66. Amarante, F.B., Scherer, C.M.S., Aguilar, C.A.G., Reis, A.D., Dickinson, W.R. and Gehrels, G.E. (2009) U-Pb ages of Mesa, V. and Souto, M. (2019) Fluvial-eolian deposits of detrital zircons in Jurassic eolian and associated the Tacuarembo Formation (Norte Basin - Uruguay): sandstones of the Colorado Plateau: evidence for depositional models and stratigraphic succession. J. S. transcontinental dispersal and intraregional recycling of Am. Earth Sci., 90, 355–376. sediment. Geol. Soc. Am. Bull., 121, 408–433. Ando, S., Garzanti, E., Padoan, M. and Limonta, M. (2012) Dickinson, W.R. and Gehrels, G.E. (2010) Insights into North Corrosion of heavy minerals during weathering and American paleogeography and paleotectonics from U-Pb diagenesis: a catalog for optical analysis. Sed. Geol., 280, ages of detrital zircons in Mesozoic strata of the Colorado 165–178. Plateau, USA. Int. J. Earth Sci., 99, 1247–1265. Bahlburg, H., Vervoort, J.D., Du Frane, S.A., Bock, B., Ernesto, M., Raposo, M.I.B., Marques, L.S., Renne, P.R., Augustsson, C. and Reimann, C. (2009) Timing of crust Diogo, L.A. and De Min, A. (1999) Paleomagnetism, formation and recycling in accretionary orogens: insights geochemistry and 40Ar/39Ar dating of the North-eastern learned from the western margin of South America. Earth- Parana Magmatic Province: tectonic implications. J. Sci. Rev., 97, 215–241. Geodynamics, 28, 321–340. Baksi, A.K. (2018) Parana flood basalt volcanism primarily Farrant, A.R., Mounteney, I., Burton, A., Thomas, R.J., limited to~ 1 Myr beginning at 135 Ma: new 40Ar/39Ar Roberts, N.M., Knox, R.W. and Bide, T. (2019) Gone with ages for rocks from Rio Grande do Sul, and critical the wind: dune provenance and sediment recycling in the evaluation of published radiometric data. J. Volcanol. northern Rub’al-Khali, United Arab Emirates. J. Geol. Soc., Geoth. Res., 355,66–77. 176, 269–283. Basei, M.A.S., Frimmel, H.E., Nutman, A.P. and Preciozzi, Francßa, A.B., Araujo, L.M., Maynard, J.B. and Potter, P.E. F. (2008) West Gondwana amalgamation based on detrital (2003) Secondary porosity formed by deep meteoric zircon ages from Neoproterozoic Ribeira and Dom leaching: botucatu eolianite, southern South America. Feliciano belts of South America and comparison with AAPG Bull., 87, 1073–1082. coeval sequences from SW Africa. Geol. Soc. London. Francischini, H., Dentzien-Dias, P.C., Fernandes, M.A. and Spec. Publ., 294, 239–256. Schultz, C.L. (2015) Dinosaur ichnofauna of the Upper Bial, J., Buettner, S.H., Schenk, V. and Appel, P. (2015) The Jurassic/Lower Cretaceous of the Parana Basin (Brazil and long-term high-temperature history of the central Namaqua Uruguay). J. S. Am. Earth Sci., 63, 180–190. Metamorphic Complex: evidence for a Mesoproterozoic Garzanti, E., Ando, S., Vezzoli, G., Lustrino, M., Boni, M. continental back-arc in southern Africa. Res., and Vermeesch, P. (2012) Petrology of the Namib Sand 268, 243–278. Sea: long-distance transport and compositional variability Bigarella, J.J. and Salamuni, R. (1961) Early Mesozoic wind in the wind-displaced Orange Delta. Earth-Sci. Rev., 112, patterns as suggested by dune bedding in the Botucatu 173–189. Sandstone of Brazil and Uruguay. Geol. Soc. Am. Bull., 72, Garzanti, E., Vermeesch, P., Ando, S., Vezzoli, G., 1089–1105. Valagussa, M., Allen, K., Kadi, K.A. and Al-Juboury, A.I. Blott, S.J. and Pye, K. (2001) GRADISTAT: a grain size (2013) Provenance and recycling of Arabian desert sand. distribution and statistics package for the analysis of Earth-Sci. Rev., 120,1–19. unconsolidated sediments. Earth Surf. Proc. Land., 26, Garzanti, E., Vermeesch, P., Padoan, M., Resentini, A., 1237–1248. Vezzoli, G. and Ando, S. (2014) Provenance of passive- Canile, F.M., Babinski, M. and Rocha-Campos, A.C. (2016) margin sand (Southern Africa). J. Geol., 122,17–42. Evolution of the Carboniferous-Early Cretaceous units of Garzanti, E., Vermeesch, P., Al-Ramadan, K.A., Ando, S., Parana Basin from provenance studies based on U-Pb, Hf Limonta, M., Rittner, M. and Vezzoli, G. (2017) Tracing and O isotopes from detrital zircons. Gondwana Res., 40, transcontinental sand transport: from Anatolia-Zagros to 142–169. the Rub’ al Khali Sand Sea. J. Sed. Res., 87, 1196–1213. Cardenas, B.T., Kocurek, G., Mohrig, D., Swanson, T., Garzanti, E., Dinis, P., Vermeesch, P., Ando, S., Hahn, A., Hughes, C.M. and Brothers, S.C. (2019) Preservation of Huvi, J., Limonta, M., Padoan, M., Resentini, A., Rittner, autogenic processes and allogenic forcings in set-scale M. and Vezzoli, G. (2018) Dynamic uplift, recycling, and aeolian architecture II: the scour-and-fill dominated climate control on the petrology of passive-margin sand Jurassic page sandstone, Arizona, USA. J. Sed. Res., 89, (Angola). Sed. Geol., 375,86–104. 741–760. Garzanti, E., Vermeesch, P., Vezzoli, G., Ando, S., Botti, E., Casquet, C., Pankhurst, R.J., Rapela, C.W., Galindo, C., Limonta, M., Dinis, P., Hahn, A., Baudet, D., De Grave, J. Fanning, C.M., Chiaradia, M., Baldo, E., Gonzalez- and Yaya, N.K. (2019) Congo River sand and the Casado, J.M. and Dahlquist, J.A. (2008) The equatorial quartz factory. Earth-Sci. Rev., 102918. Mesoproterozoic Maz terrane in the Western Sierras Gehrels, G. (2014) Detrital zircon U-Pb geochronology Pampeanas, Argentina, equivalent to the Arequipa- applied to tectonics. Annu. Rev. Earth Planet. Sci., 42, Antofalla block of southern Peru: implications for West 127–149. – Gondwana margin evolution. Gondwana Res., 13, 163 175. Gregory, T.R., de Fatima Bitencourt, M., Nardi, L.V.S., Dias, K.D.N. and Scherer, C.M. (2008) Cross-bedding set Florisbal, L.M. and Chemale Jr, F. (2015) thickness and stratigraphic architecture of aeolian systems: Geochronological data from TTG-type rock associations of an example from the Upper Permian Piramboia Formation the Arroio dos Ratos Complex and implications for crustal

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology 18 G. Bertolini et al.

evolution of southernmost Brazil in Oyhantcßabal, P., Wagner-Eimer, M., Wemmer, K., Schulz, times. J. S. Am. Earth Sci., 57,49–60. B., Frei, R. and Siegesmund, S. (2012) Paleo-and Hadlari, T., Swindles, G.T., Galloway, J.M., Bell, K.M., Neoproterozoic magmatic and tectonometamorphic Sulphur, K.C., Heaman, L.M., Beranek, L.P. and Fallas, evolution of the Isla Cristalina de Rivera (Nico Perez K.M. (2015) 1.8 billion of detrital zircon recycling Terrane, Uruguay). Int. J. Earth Sci., 101, 1745–1762. calibrates a refractory part of Earth’s sedimentary cycle. Oyhantcßabal, P., Oriolo, S., Philipp, R.P., Wemmer, K. and PloS one, 10, e0144727. Siegesmund, S. (2018a) In: The Nico Perez Terrane of Havholm, K.G. and Kocurek, G. (1994) Factors controlling Uruguay and Southeastern Brazil. In: Geology of aeolian sequence stratigraphy: clues from super bounding Southwest Gondwana (Eds Siegesmund, S., Basei, M., surface features in the Page Sandstone. Oyhantcßabal, P. and Oriolo, S.), pp. 161–188. Springer, Sedimentology, 41, 913–934. Cham. Hofig,€ D.F., Marques, J.C., Basei, M.A.S., Giusti, R.O., Oyhantcßabal, P., Cingolani, C.A., Wemmer, K. and Kohlrausch, C. and Frantz, J.C. (2018) Detrital zircon Siegesmund, S. (2018b) The Rıo de la Plata craton of geochronology (U-Pb LA-ICP-MS) of syn-orogenic basins Argentina and Uruguay. In: Geology of Southwest in SW Gondwana: new insights into the - Gondwana (Eds Siegesmund, S., Basei, M., Oyhantcßabal, of Porongos Complex, Dom Feliciano Belt, P. and Oriolo, S.), pp. 89–105. Springer, Cham. southern Brazil. Precambrian Res., 306, 189–208. Perea, D., Soto, M., Veroslavsky, G., Martınez, S. and Hurst, A., Morton, A., Scott, A., Vigorito, M. and Frei, D. Ubilla, M. (2009) A fossil assemblage in (2017) Heavy-mineral assemblages in sandstone intrusions: Gondwana: biostratigraphy and correlations of the panoche giant injection complex, California. J. Sed. Res., Tacuarembo Formation, Parana Basin, Uruguay. J. S. Am. 87, 388–405. Earth Sci., 28, 168–179. Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Peri, V.G., Naipauer, M., Pimentel, M. and Barcelona, H. Pickle, J.D. and Sares, S.W. (1984) The effect of grain size (2016) Eolian deposits of the shouthwestern margin of the on detrital modes: a test of the Gazzi-Dickinson point- Botucatu paleoerg: reconstruction of the Gondwana counting method. J. Sed. Res., 54, 103–116. landscape in central northern Argentina. Sed. Geol., 339, Janasi, V.A., Freitas, V.A. and Heaman, L.H. (2011) The 234–257. onset of flood basalt volcanism, Northern Parana Basin, Philipp, R.P., Pimentel, M.M. and Chemale, F., Jr (2016) Brazil: a precise U-Pb baddeleyite/zircon age for a Tectonic evolution of the Dom Feliciano Belt in Southern Chapeco-type dacite. Earth Planet. Sci. Lett., 302, 147–153. Brazil: geological relationships and U-Pb geochronology. Janikian, L., de Almeida, R.P., Fragoso-Cesar, A.R.S., de Souza Braz. J. Geol., 46,83–104. Martins, V.T., Dantas, E.L., Tohver, E. and D’Agrella-Filho, Philipp, R.P., Schultz, C.L., Kloss, H.P., Horn, B.L., Soares, M.S. (2012) Ages (U–Pb SHRIMP and LA ICPMS) and M.B. and Basei, M.A. (2018) SW stratigraphic evolution of the Neoproterozoic volcano- Gondwana paleogeography and sedimentary dispersal sedimentary successions from the extensional Camaqua~ revealed by integration of stratigraphy and U-Pb zircon Basin, Southern Brazil. Gondwana Res., 21,466–482. analysis: the Santa Cruz Sequence, Parana Basin, Brazil. J. Jenisch, A.G., Lehn, I., Gallego, O.F., Monferran, M.D., S. Am. Earth Sci., 88, 216–237. Horodyski, R.S. and Faccini, U.F. (2017) Stratigraphic Pinto, V.M., Hartmann, L.A., Santos, J.O.S., McNaughton, distribution, and paleoenvironments of N.J. and Wildner, W. (2011) Zircon U-Pb geochronology Spinicaudata in the Triassic and Jurassic of the Parana from the Parana bimodal volcanic province support a brief Basin. J. S. Am. Earth Sci., 80, 569–588. eruptive cycle at~ 135 Ma. Chem. Geol., 281,93–102. Mange, M.A. and Maurer, H. (2012) Heavy Minerals in Pye, K. and Tsoar, H. (2008) Aeolian Sand and Sand Dunes. Colour. Springer Science & Business Media, Berlin. Springer Science & Business Media, Berlin. Mcbride, E.F. (1985) Diagenetic processes that affect Rahl, J.M., Reiners, P.W., Campbell, I.H., Nicolescu, S. and provenance determinations in sandstone. In: Provenance Allen, C.M. (2003) Combined single-grain (U-Th)/He and of Arenites (Ed. Zuffa, G. G.), pp. 95–113. Springer, U/Pb dating of detrital zircons from the , Dordrecht. Utah. Geology, 31, 761–764. Milani, E.J., Faccini, U.F., Scherer, C.M.S., Araujo, L.M. and Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E., Cupertino, L.M. (1998) Sequences and stratigraphic Galindo, C., Fanning, C.M. and Dahlquist, J.M. (2010) The hierarchy of the Parana Basin (Ordovician to Cretaceous), Western Sierras Pampeanas: protracted Grenville-age Southern Brazil. Bol. IG USP Serie Cientıfica, 29, 125–173. history (1330-1030 Ma) of intra-oceanic arcs, subduction– Milani, E.J., Melo, J.H.G., Souza, P.A., Fernandes, L.A. and accretion at continental-edge and AMCG intraplate Francßa, A.B. (2007) Bacia do Parana. Bol. Geocienc.^ magmatism. J. S. Am. Earth Sci., 29, 105–127. Petrob., 15, 265–287. Resentini, A., Ando, S. and Garzanti, E. (2018) Quantifying Morton, A.C. and Hallsworth, C. (1994) Identifying roundness of detrital minerals by image analysis: sediment provenance-specific features of detrital heavy mineral transport, shape effects, and provenance implications. J. assemblages in sandstones. Sed. Geol., 90, 241–256. Sed. Res., 88, 276–289. Morton, A.C. and Hallsworth, C. (2007) Stability of detrital Rittner, M., Vermeesch, P., Carter, A., Bird, A., Stevens, T., heavy minerals during burial diagenesis. Dev. Sed., 58, Garzanti, E., Ando, S., Vezzoli, G., Dutt, R., Xu, Z. and 215–245. Lu, H. (2016) The provenance of Taklamakan desert sand. Oriolo, S., Oyhantcßabal, P., Basei, M.A., Wemmer, K. and Earth Planet. Sci. Lett., 437, 127–137. Siegesmund, S. (2016) The Nico Perez Terrane (Uruguay): Rossetti, L., Lima, E.F., Waichel, B.L., Hole, M.J., Simoes,~ from Archean crustal growth and connections with the M.S. and Scherer, C.M.S. (2018) Lithostratigraphy and Congo Craton to late Neoproterozoic accretion to the Rıo volcanology of the Serra Geral Group, Parana-Etendeka de la Plata Craton. Precambrian Res., 280, 147–160. Igneous Province in southern Brazil: towards a formal

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology Controls on desert sediment provenance 19

stratigraphical framework. J. Volcanol. Geoth. Res., 355, Soto, M. and Perea, D. (2008) A ceratosaurid (Dinosauria, 98–114. theropoda) from the late Jurassic-Early Cretaceous of Saalmann, K., Gerdes, A., Lahaye, Y., Hartmann, L.A., Uruguay. J. Verteb. Paleontol., 28, 439–444. Remus, M.V.D. and Laufer,€ A. (2011) Multiple accretion at Thiede, D.S. and Vasconcelos, P.M. (2010) Parana flood the eastern margin of the Rio de la Plata craton: the basalts: rapid extrusion hypothesis confirmed by new prolonged Brasiliano orogeny in southernmost Brazil. Int. 40Ar/39Ar results. Geology, 38, 747–750. J. Earth Sci., 100, 355–378. Vermeesch, P., Resentini, A. and Garzanti, E. (2016) An R Santos, J.O.S., Hartmann, L.A., Bossi, J., Campal, N., Schipilov, package for statistical provenance analysis. Sed. Geol., A., Pineyro,~ D. and McNaughton, N.J. (2003) Duration of the 336,14–25. Trans-Amazonian Cycle and its correlation within South Wildner, W., Ramgrab, G.E., Lopes, R.C. and Iglesias, America based on U-Pb SHRIMP geochronology of the La C.M.F. (2006). Mapa Geologico do Estado do Rio Grande Plata Craton, Uruguay. Int. Geol. Rev., 45,27–48. do Sul, Escala 1:750.000. CPRM - Companhia de Pesquisa Sato, K., Basei, M.A.S., Ferreira, C.M., Sproesser, W.M., de Recursos Minerais, Porto Alegre. Vlach, S.R.F., Ivanuch, W. and Onoi, A.T. (2011). U-Th- Zalan, P.V., Wolff, S.J.C.J., Conceicßao,~ J.D.J., Marques, A., Pb analyses by excimer laser ablation/ICP-MS on MG Astolfi, M.A.M., Vieira, I.S., Appi, V.T. and Zanotto, O.A. Brazilian xenotime. Goldschmidt Conference Abstracts. (1990). Bacia do Parana. Origem e evolucßao~ das bacias Miner Mag 1801. sedimentares, pp.135–168. Scherer, C.M.S. (2000) Eolian dunes of the Botucatu Zerfass, H., Lavina, E.L., Schultz, C.L., Garcia, A.J.V., Faccini, Formation (Cretaceous) in southernmost Brazil: U.F. and Chemale, F., Jr (2003) Sequence stratigraphy of morphology and origin. Sed. Geol., 137,63–84. continental Triassic strata of Southernmost Brazil: a Scherer, C.M.S. (2002) Preservation of aeolian genetic units contribution to Southwestern Gondwana palaeogeography by lava flows in the Lower Cretaceous of the Parana Basin, and palaeoclimate. Sed. Geol., 161,85–105. southern Brazil. Sedimentology, 49,97–116. Scherer, C.M. and Goldberg, K. (2007) Palaeowind patterns Manuscript received 30 October 2019; revision during the latest Jurassic–earliest Cretaceous in accepted 28 January 2020 Gondwana: evidence from aeolian cross-strata of the Botucatu Formation, Brazil. Palaeogeogr. Palaeoclimatol. – Palaeoecol., 250,89 100. Supporting Information Scherer, C.M. and Lavina, E.L. (2006) Stratigraphic evolution of a fluvial–eolian succession: the example of the Upper Jurassic-Lower Cretaceous Guara and Botucatu Additional information may be found in the online formations, Parana Basin, Southernmost Brazil. Gondwana version of this article: Res., 9, 475–484. Schneider, R.L., Muhlmann, H., Tommasi, E., Medeiros, Table S1. Heavy mineral data. R.A., Daemon, R.F. and Nogueira, A.A. (1973) Revisao~ Table S2. Petrographic data. estratigrafica de Bacia do Parana An. XXVIII Congr. Bras. Table S3. Granulometric data. Geol. Soc. Bras. Geol., 1,41–65. Table S4. Detrital zircon U-Pb age data.

© 2020 The Authors. Sedimentology © 2020 International Association of Sedimentologists, Sedimentology