<<

, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology

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

The oldest known profiles on Earth: 3.46 Ga Panorama Formation, Western

Gregory J. Retallack

Department of Geological Sciences, University of Oregon, Eugene, Oregon 97403-1272, USA

ARTICLE INFO ABSTRACT

Keywords: Subaerial volcanic flank and floodplain facies of the 3.46 Ga Panorama Formation have been recognized on the Archean basis of trough cross-bedded sandstones, lapilli tuffs, and abundant nodularized barite crystals, like those Pilbara noted in other Archean non-marine facies. These barite-nodular layers are interpreted as alluvial for Barite the following reasons. They show different degree of bedding disruption scaled to nodule size beneath a sharp Acid sulfate upper boundary, like desert profiles. They show multiple generations of cracking, skins, and up-profile Atmospheric destruction of feldspar and fragments, compatible with weathering of labile constituents of the parent material. Loss of alkali and alkaline earth elements and up profile are also features of chemical weathering. Geochemical mass balance calculations (tau analysis) shows that the profiles lost mass and labile elements, unlike unaltered and tuffs. Rare earth element analysis also shows light rare earth retention as in weathering, as opposed to sedimentation or hydrothermal alteration. Barite of nodules in the paleosols is mobilized to concentration under very acidic conditions (pH < 3) and is very stable under less acidic condi- tions. These Archean alluvial paleosols may have formed by acid sulfate weathering by sulfuric acid, rather than the currently more common hydrolysis by carbonic acid. This archaic system of widespread acid sulfate weathering has since been marginalized to a few playa lakes, deep water tables, and sulfur springs.

1. Introduction (Tarr, 1933) or Wüstenrosen (Dietrich, 1975). Such features are turning out to be abundant in Archean alluvial paleosols in both Western Archean paleosols at major geological unconformities have been Australia (Retallack, 2014a; Retallack et al., 2016), and South studied for some time (Buick et al., 1995; Rye and Holland, 1998), but (Nabhan et al., 2016). This paper documents additional examples from Archean alluvial paleosols have eluded detection until recently the Panorama Formation of the Pilbara Craton of Western Australia, as (Retallack, 2014a; Nabhan et al., 2016; Retallack et al., 2016). Without evidence for the idea that acid-sulfate weathering (sulfuricization of clear indications like root traces, alluvial paleosols can be Staff, 2014) may have been more widespread on land difficult to distinguish from ordinary sedimentary beds (Retallack, during the Archean than it has been subsequently. The paleosols de- 2011; Maslov and Grazhdankin, 2011; Kolesnikov et al., 2012, 2015). scribed here are currently the oldest known paleosols on Earth (Ohmoto Field indications of Precambrian paleosols include vertical threads of et al., 2014; Retallack, 2014a; Nabhan et al., 2016; Retallack et al., likely organisms (Retallack, 2011, 2013a), calcareous nodules 2016), although paleosols as old as 3.7 Ga have been recognized on (Retallack, 2011), and gypsum sand crystals (Retallack, 2012, 2013a), Mars, and those also have sulfate nodules (Retallack, 2014c). Two of but confident identification of Precambrian alluvial paleosols requires a the Martian paleosols may appear controversial because described as variety of petrographic and geochemical analyses (Retallack, 2011, playa lake deposits with early diagenetic sulfates (Schieber et al., 2012, 2013a). A variety of lines of evidence for Precambrian alluvial 2017), otherwise known as (Food and Agriculture paleosols was reviewed by Retallack (2016): oxidation rinds, Organization, 1974), or acid sulfate soils (Fitzpatrick et al., 2008). texture, tsunamites, playa minerals, boron content, boron isotopes, carbon/sulfur ratios, total/reactive iron ratios, lapilli and crystal tuffs, 1.1. Geological setting ice deformation, birefringence fabrics, evaporitic sand crystals, geo- chemical tau analysis, and carbon‑oxygen isotopic covariance. Of these, The principal locality studied for this work is the Panorama Tarhan et al. (2015) found that the most convincing were evaporite Formation in the area here informally called “Panorama Portal”, where sand crystals and nodules, of the kind commonly called desert roses the road to Dresser Mine leaves the floodplain of the North Shaw River,

E-mail address: [email protected]. http://dx.doi.org/10.1016/j.palaeo.2017.10.013 Received 4 April 2017; Received in revised form 4 October 2017; Accepted 9 October 2017 0031-0182/ © 2017 Elsevier B.V. All rights reserved.

Please cite this article as: Retallack, G.J., Palaeogeography, Palaeoclimatology, Palaeoecology (2017), http://dx.doi.org/10.1016/j.palaeo.2017.10.013 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 1. Geological map of a paleovolcano in the 3.46 Ga Panorama Formation, North Pole Dome, Pilbara Craton, Western Australia (after van Kranendonk, 2000; Bolhar et al., 2005), showing location of measured sections of paleosols. and enters hills 30 km south of the Port Hedland-Marble Bar highway, elements and Th values, similar to local tholeiites and komatiites. In and 10 km northwest of Panorama homestead (abandoned), in east contrast, most silicifi ed rhyolitic tuffs of the Panorama Formation have Pilbara Shire, Western Australia (Figs. 1–2). Coordinates of the central light rare earth enrichment and lack negative Eu anomalies, consistent ridge of this locality (Fig. 2A) are 21.01580°S 119.38178°E. with melting of mantle eclogite, unlike other local Archaean rhyolites of The geological age of the Panorama Formation is uncertain in the the Duffer Formation (Cullers et al., 1993). Unsilicified calcareous tuffs study area, but UePb ages from rhyolite zircons have been determined of the Panorama Formation are similar in rare earth element compo- in both Glen Herring Creek as 3430 ± 4 Ma (Nelson, 2002), and on sition to rhyolitic tuffs, but have greater light rare earth enrichment, Panorama Ridge as 3458 ± 1.9 Ma (Thorpe et al., 1992). This latter and high Ba and Sr (Fig. 5). These calcareous tuffs are not as enriched in date is equivalent to the UePb age of 3459 ± 18 Ma of the intrusive light rare earth elements as carbonatite tuffs(Woolley and Kempe, North Pole Monzogranite (Thorpe et al., 1992), which may have been a 1989), including Archean (3.01 Ga) carbonatite from Greenland subvolcanic intrusion to the active Panorama paleovolcano (van (Larsen and Rex, 1992; Bizzarro et al., 2002). Kranendonk, 2000; Bolhar et al., 2005). The study area exposes an unusually thick sequence of the 2. Materials and methods Panorama Formation (Fig. 2A), regarded as flank facies of a paleo- volcanic stratocone (van Kranendonk, 2000; Bolhar et al., 2005). This The primary activity of this fieldwork was measuring sections of the volcanic edifice remained a paleotopographic high because it is over- Panorama Formation (Fig. 3), regarded as a subaerial volcanic edifice lain by Euro Basalt in the study area, whereas 28 m of Strelley Pool by Konhauser et al. (2002). Also documented were beds considered Formation unconformably overlies the Panorama Formation elsewhere likely paleosols (Figs. 2C–I, 4), because of field characters of sharp top (Fig. 1). The Panorama paleovolcano erupted onto pillow basalts of the and gradational change below, dilational cracking patterns, and eva- Mt. Ada Basalt (van Kranendonk, 2000; Bolhar et al., 2005). An area poritic sand crystals. Potential paleosol beds were all named in the identified as an ancient vent of the paleovolcano by van Kranendonk field, within four distinct categories of Jurl, Ngumpu, Nganga and (2000) and Bolhar et al. (2005) includes cross-cutting breccia pipes, Wanta, using descriptive words from Nyamal local aboriginal languages with large blocks of banded iron formation of a caldera lake (Fig. 2B), (Burgman, 2007). Each potential pedotype is a bed with a distinctive and of carbonate-altered Mt. Ada Basalt. The ancient stratovolcano sequence of horizons, and the diagnostic features of each are given in flank facies is at least 1 km of agglomerates, rhyolite flows, and tuffs Table 1. These are field categories to be tested as paleosol profiles by (Fig. 3–4). These tuffs include silicate accretionary lapilli (van laboratory data. Kranendonk, 2000), which disperse in water, and thus are evidence that Oriented thin sections prepared vertical to bedding were point the volcano flanks were subaerial (Retallack, 2014b). Central subaerial counted using a Swift automated stage and Hacker electronic counting volcanoes have also been inferred from a pattern of paleocurrents in box (Supplementary material Tables S1–S2). Point counting was ad- fluvial facies of the Panorama Formation radiating from regional cen- justed for pervasive silicification and recrystallization to crystallites ters (DiMarco and Lowe, 1989). minimally 5 μm across (Lepot et al., 2013), so that 5 μm grain size and The original nature of the Panorama paleovolcano is best inferred smaller were counted as clay, rather than the standard soil-clay cutoff at from trace elements, because of widespread silicification and meta- 2 μm of . Rock fragments were mainly sedimentary, with morphic alteration (Brown et al., 2006). Some Panorama Formation rare metamorphic and basaltic clasts. Evaporite minerals are now tuffaceous siltstones have high Sc and Cr values, and low rare earth quartz pseudomorphs, and identified from past research on polished

2 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 2. Field photographs of Panorama paleovolcano, banded iron formation, and paleosols and polished slabs of paleosol surface horizons covered with laminites, and of barite sand crystals in Panorama Formation east Pilbara Craton, Western Australia. (A) ridge west of Panorama Portal (289–294 m in Fig. 3:21.02865°S 119.38087°E); (B) jasper-siderite banded iron west of Panorama Portal (10-20 m in Fig. 3:21.03722°S 119.37359°E); (C) Nganga paleosol west of Panorama Portal (648 m in Fig. 3:21.03365°S 119.37915°E); (D) Jurl paleosol west of Panorama Portal (293 m in Fig. 3:21.03229°S 119.97799°E); (E), Jurl paleosol southeast of Panorama Portal (311 m in Fig. 3:21.02554°S 119.38371°E), (F) Jurl paleosol north of Panorama Portal (610 m in Fig. 3:21.01605°S 119.37734°E); (G) A horizon Wanta paleosol (R3774) west of Panorama Portal (63 m in Fig. 3: 21.03182°S 119.37310°E), (H) A horizon of Jurl paleosol (R3777) west of Panorama Portal (62 m in Fig. 3: 21.03182°S 119.37310°E); (I) By horizon of Jurl paleosol (R3755) west of Panorama Portal (294 m in Fig. 3: 21.03229°S 119.97799°E). Sand crystals are at arrows. slabs, large thin sections and electron microprobe analyses for com- arefined blank correction in the CAMCOR instrument facility of the parable local Archean cherts (Lambert et al., 1978; Brasier et al., 2002; University of Oregon (Donovan et al., 2011). Rare earth elements were Sugitani et al., 2003; Pinti et al., 2009). Original evaporite mineralogy normalized to chondrite of Anders and Grevesse (1989). Bulk densities could not be determined in thin section while point counting, but the were determined from 20 to 40 g samples using paraffin at the Uni- peudomorphs have a distinctive chalcedonic fabric. Opaque material versity of Oregon (Retallack, 1997a), with errors from 10 replicate counted includes both heavy minerals and organic matter. Two separate density determinations of one of the Panorama Formation chert speci- 500-point counts were made for grain size and for mineral composition mens (R4309). Thin sections and their source rocks are archived in the of each sample, with errors (2σ) of ± 2% for common (> 8%) com- Condon Collection of the Museum of Natural and Cultural History of the ponents (Murphy, 1983). Chemical analyses of selected samples were University of Oregon, in Eugene, Oregon. performed by ALS Chemex of Vancouver, British Columbia, using XRF on glass discs for major elements, ICP on glass discs for trace elements, and FeO by Pratt titration (Supplementary material Tables S4–S6). The 3. Burial alteration standard for the analyses was Canadian Certified Reference Materials Project standard SY-4, diorite gneiss from near Bancroft, Ontario. Errors Soil formation is a form of early diagenesis, but reconstructing pa- (2σ) are from 89 replicate analyses of the standard in the same la- leosols requires a thorough understanding of subsequent alterations, boratory. Titania values were below detection (< 0.01 wt%) for the particularly metamorphism and silicification of individual beds. Jurl paleosol (R3775-R3779), but these were re-analyzed by ion mi- croprobe using an aggregation of multiple spectrometer intensities and

3 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 4. Detailed sections of Panorama Formation in the paleovolcano flank facies, east Pilbara Craton, Western Australia. Blue boxes to right of column show thickness (of A-Bk) and degree of development of paleosols (after Retallack, 1997a), white grains of barite are shown as volume percent, and hue is from a Munsell chart. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

fl Fig. 3. Geological section of Panorama Formation in the paleovolcano flank facies, east hydrothermal uids, because they lack large positive europium Pilbara Craton, Western Australia (with paleoenvironmental interpretations from van anomalies due to high temperature plagioclase and Cl-complex Kranendonk, 2000). formation (Fig. 5), found in known hydrothermal chert of the Pilbara region (Bolhar et al., 2005; Sugahara et al., 2010). 3.1. Metamorphism Greenschist facies regional metamorphism and pyrophyllitic hy- drothermal alteration are compatible with likely burial of the Panorama – Panorama Formation paleosols, metasediments, rhyolites and tuffs Formation to depths of 14 17 km, and deep burial alterations of the suffered chlorite-calcite-epidote lower greenschist facies burial meta- paleosols must be disentangled from ancient soil formation. Around morphism (ca. 350 °C) like the overlying Strelley Pool Formation and North Pole Dome, the overlying Strelley Pool Formation is 28 m thick, Euro Basalt (Cullers et al., 1993; van Kranendonk and Pirajno, 2004; Euro Basalt 9400 m, Sulfur Springs Group 4941 m (van Kranendonk, Allwood et al., 2006). Clayey rocks of the Panorama Formation were 2000), Gorge Creek Group 2610 m (van Kranendonk et al., 2006), for a also influenced by contact hydrothermal alteration to pyrophyllite (also total of 16,980 m of overburden. Near Strelley Pool however, the ca. 350 °C) by intrusion of the North Pole Monzogranite (Brown et al., thickness of Euro Basalt is only 1000 m (Buick et al., 1995), so over- 2006). The cherts studied here however were not altered by burden beneath the Lalla Rookh and Fortescue Groups would have been

4 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 5. Rare earth element analyses of putative paleosols of the Panorama Formation of Western Australia normalized to chondritic values of Anders and Grevesse (1989):(A) Wanta and Nganga pedotypes; (B) Jurl and Ngumpu pedotypes.

8580 m. The Lalla Rookh Formation of the overlying De Gray Group 3.3. Early diagenetic silicification was confined to fault basins (Krapež, 1984), so its 3000 m did not overlie the studied localities. Fortescue Group basalts and sediments Most tuffs in the Panorama Formation have been extensively sili- northwest of Marble Bar are only 5240 m thick (Thorne and Trendall, cified like stromatolitic dolostones, sandstones and conglomerate of the 2001; van Kranendonk, 2010), so toward the east total burial depth overlying Strelley Pool Formation (Allwood et al., 2007). Clues to sili- may have been only 13,826 m. cification of Panorama Formation paleosols come from organic micro- in comparable cherts of the Strelley Pool Formation (Sugitani et al., 2010; Wacey et al., 2010, 2011), which Hickman (2008) con- 3.2. Late diagenesis siders equivalent to the microfossiliferous “Kitty's Gap Chert” (Westall et al., 2006). Microfossils are evidence of silicification before the cells Illitization is a common deep burial enrichment of potassium could decay: which takes about 160 days for cyanobacteria under an- (Nesbitt and Young, 1989; Fedo et al., 1995; Novoselov and de Souza oxic conditions (Harvey et al., 1995). Microfossils of the Strelley Pool Filho, 2015). Nevertheless, geochemical data on the putative paleosols Formation are mainly robust large spindles, which are widely accepted chosen for detailed analysis shows weathering rather than illitization as genuine for reasons of carbon isotopic composition, Raman spec- trends (Fig. 6A). Illitization has affected red laminites (part of Wanta troscopy, and three-dimensional imaging (Sugitani et al., 2010; Wacey profile) and weathered upper portions of calcareous tuffs (Nganga et al., 2010, 2011; Schopf et al., 2007; Brasier et al., 2015). These ob- profile) of the Panorama Formation with marked potash enrichments servations contradict the idea that cherts of the Strelley Pool and Pa- (Fig. 6B). In contrast the cherty paleosols (Jurl profiles) show minor norama Formations are silcretes formed in Archean outcrops during depletion, not enrichment in potash, probably because of protection by Cenozoic weathering (van Kranendonk, 2000). Australian Cenozoic early silicification, as documented for the overlying Strelley Pool For- silcretes are chemically distinct, with much more titania (0.7–48.0 wt mation (Buick and Barnes, 1984; Hickman, 2008; Allwood et al., 2010). %: Wopfner, 1978; Taylor and Eggleton, 2001) than cherts of the Pa- Silicification from dissolved silica released by illitization of clays norama Formation (0.006–0.14 wt%: Supplementary material Table and pressure solution of sand grains is a common late diagenetic al- S3). teration of sedimentary rocks (Siever, 1962), but unlikely for extensive The following paragraphs consider three plausible mechanisms of silicification of the Panorama Formation. Not only is there little evi- early diagenetic silicification of paleosols of the Panorama Formation: dence of a trend toward illite (Fig. 6A) or a potash enrichment in these (1) hydrothermal, (2) biogenic, and (3) evaporitic. Silicification in hot rocks (Fig. 6B), but they were relatively poor in clay (Fig. 7). Parent springs would be compatible with chert dikes and pyrophyllitic al- material in these plots is below the feldspar line (50% Al2O3)inFig. 6A teration associated with intrusion of the North Pole Monzogranite, and and at origin in Fig. 6B. Cherts of the Panorama Formation do show nearby vent facies of the paleovolcano in the Panorama Formation pervasive crystal domains up to 5 μm in size, larger than usual for (Kato and Nakamura, 2003; van Kranendonk, 2006). Furthermore, red geologically younger cherts with “pinpoint birefringence”, and perhaps cherty banded iron formation of the basal Panorama Formation shows a due to diagenetic (neomorphic) recystallization, as for cherts of the strong positive europium anomaly and heavy rare earth enrichment overlying Strelley Pool Formation (Lepot et al., 2013). characteristic of hydrothermal alternation, probably within waters of a

Table 1 Summary of Panorama Formation paleosol definition and classification.

Pedotype Nyamal meaning Diagnosis US taxonomy (Soil FAO map unit (Food and Australian classification (Burgman, 2007) Survey Staff, 2014) Agriculture Organization, 1974) (Isbell, 1996)

Jurl Salt Black massive cherty surface (A horizon) over Haplogypsid Orthic Solonchak Gray Sodosol gypsum crystal pseudomorphs in gray chert (By horizon) Nganga Sand Brown sandstone surface (A) horizon over white Calcaric Arenic Rudosol calcareous sandstone (C) horizon Ngumpu Narrow Black massive cherty surface (A horizon) over Fluvent Eutric Stratic Rudosol laminated gray chert (C horizon). Wanta Crazy Red cracked and laminated surface (A horizon) Halaquept Gleyic Solonchak Gray-orthic Tenosol over barite mottle pseudomorphs (By horizon)

5 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 6. Illitization of paleosols in the Panorama Formation in the paleovolcano flank facies, east Pilbara Craton, Western Australia, showing (A) weathering trends toward alumina pole of Nesbitt and Young (1989) on a truncated triangular diagram, and (B) potash enrichment as tau values following Novoselov and de Souza Filho (2015). volcanic summit caldera (Bolhar et al., 2005). However, cherts of the 4.1. Field observations paleovolcano flank facies have a different and flatter rare earth element pattern (Fig. 6), comparable with patterns in the Strelley Pool Forma- Many cherty beds of the Panorama Formation have common early tion considered evidence for low temperature silicification (van diagenetic pseudomorphs of evaporite minerals which range in form Kranendonk, 2000; van den Boorn et al., 2010; Allwood et al., 2010). from elongate to short crystals (Fig. 2D–F at arrows), variously cor- An alternative source of silica in low temperature sedimentary en- roded and or extended as rounded nodules (Fig. 2I at arrows). These vironments of many Precambrian microfossiliferous cherts is waters of vary in abundance and size from bed to bed, from small masses with higher silica concentrations than today, because predating evolution of little disrupting effect (Fig. 2D) on primary bedding to abundant no- siliceous microbes such as diatoms and radiolarians (Eriksson and dules up to 2 cm in diameter which totally obscure bedding (Fig. 2F). Warren, 1983; Maliva et al., 2005). This view is undermined by de- Furthermore, these nodular and crystalline masses are neither at the monstration that silica spicules are secreted by sulfate-reducing bac- very top nor base of their respective beds: they are sparse in the upper teria (Birnbaum and Wireman, 1985; Birnbaum et al., 1989), which 10 cm or so of a bed then increase in abundance below that and taper have independently been inferred from microfossils and sulfur isotopic off in abundance some distance above the plane or cross-bedded base of studies of the overlying 3.42 Ga Strelley Pool Formation (Wacey et al., the bed (Fig. 2E–F). In all these ways, paleosols of the Panorama For- 2010, 2011; Bontognali et al., 2012), and underlying 3.48 Ga Dresser mation are superficially similar to salt-rich (By) horizons of soils Formation (Shen et al., 2001; Ueno et al., 2008). Recrystallization of (Retallack and Huang, 2010; Amundson et al., 2012), and paleosols of biogenic spicules may have contributed to silicification of the Panorama the 3.0 Ga Farrel Quartzite of Western Australia (Retallack et al., 2016), Formation. and 3.2 Ga Moodies Group of South Africa (Nabhan et al., 2016). The most likely mechanism of early silicification is formation as The bedding disruption fades downward from truncated upper playa cherts, comparable with magadiite (Hay, 1968; Eugster, 1969). surfaces of the paleosols, including those without evaporites

Minerals such as gypsum (CaSO4.2H2O) are often precipitated at high (Fig. 2C–F), and this may represent the upper surface of the original pH (> 9), which mobilizes silica, and such alkaline solutions dissolve soil. These upper surfaces have organic carbon crusts with sharp upper biogenic opaline or volcanic silica in preference to quartz, creating si- surface and wispy downward extensions (Fig. 2H), perhaps from mi- lica gels in playas and intertidal flats (Eugster, 1969; Smith, 1979; crobial soil crusts (Retallack, 2014a). Some upper surfaces to beds with Larsen, 2008). This model of caustic rather than neutral solutions, numerous evaporate pseudomorphs have a few laminations of hematite, perhaps aided by sulfate-reducing bacteria is appealing for silica per- including contorted layers (Fig. 2G), comparable with the cumulic mineralization of many fossiliferous Precambrian intertidal to supra- surface horizons of lowland soils (Retallack, 1997a). Most soils have tidal cherts (Maliva et al., 2005). sharp upper surfaces, but lowland soils have increments of additional before burial by a terminating tuff or river sand.

4. Testing likely paleosols 4.2. Petrographic observations

A variety of beds were identified in the field as potential paleosols Bedding disruption is especially obvious in thin sections of the largely on the basis of destruction of primary sedimentary features, and upper surface of putative paleosols which show lamination above a these beds were sampled for a battery of geochemical and petrographic sharp contact to near-vertical, chaotic fabric in thin sections cut vertical tests to determine whether they were paleosols, or unaltered tuffsor to regional bedding (Fig. 7A–E). This near-vertical fabric reveals mul- sedimentary beds. Most of these formed the top 30–40 cm of trough tiple generations of bedding disruption by cracking which have cross bedded sandstone (angular discordance to bed top in Fig. 2D–E). homogenized original bedding, but there also are cases where lamina- tion is only mildly disrupted by vertical structures (Fig. 7F). Other indications of soil formation in thin section are clay skins

6 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 7. Petrographic thin sections of paleosols in the Panorama Formation of Western Australia, under crossed nicols (A–C, F–H) and plane light (D–E): showing filamentous vertical disruptions of surface horizons of type Wanta (A), of type Jurl clay (B,D), of type Nganga loam (C) and of type Ngumpu clay loam (E–F); G, clay skins in surface horizon of type Jurl clay; H, rock fragments weathered to clay in A horizon of type Nganga clay loam. All thin sections were cut vertical to regional bedding and are curated in the Condon Collection University of Oregon: A, R3774; B, D, G, R3756; C, R3765; E–F, R3761; H, R3766. rimming weathered grains, which are embayed and dilated into frag- growth or simple precipitation in marine evaporites (Warren, 2006; ments (Figs. 7G–H). Such dilation indicates alteration within a soil Ziegenbalg et al., 2010) and playa lakes (Hay, 1968; Renaut and under little confining pressure, unlike burial diagenetic alterations Tiercelin, 1994). Also different are void-filling, inclusion-free, barites of which result in overall reduction in volume. Furthermore, the clay skins hydrothermal vugs and veins, such as those of the 3.48 Ga Dresser have high birefringence (Fig. 7G), a feature of soils and paleosols called Formation (Shen et al., 2001; Ueno et al., 2008). Different again are sepic plasmic fabric (Retallack, 1997a). spring tufas with cavity-filling pendant and erect barite fans, and la- Evaporite pseudomorphs of the Panorama Formation in thin section minated to stromatolite-like isopachous barite coatings, also inclusion- show abundant included grains characteristic of desert roses (Tarr, free (Donovan et al., 1988; Bonny and Jones, 2008a, 2008b). 1933; Dietrich, 1975), as revealed by common sedimentary inclusions, The original mineral of Archean chert pseudomorphs of evaporite and corroded and dusty rims (Fig. 8C–E). Some of these pseudomorphs minerals in the Panorama Formation may have been barite (BaSO4), are radially arranged like desert roses (Fig. 8C), and others are heavily despite low Ba (8.2–3670 ppm) remaining after thorough early diage- nodularized (Fig. 8B). Graded beds overlying the paleosols have re- netic silicification (Supplementary materials Table 3). Barite inclusions deposited crystal pseudomorphs (Fig. 8A), demonstrating that they had within other Archean pseudomorphs are near stoichiometric, with 59 to formed and were eroded during breaks in deposition. Comparable ob- 65 wt% BaO (Sugitani et al., 2003; Nabhan et al., 2016). The crystals servations of sand crystals and redeposition reveal paleosols in the when best preserved have the squat form of barite (Fig. 8E), unlike

3.2 Ga Moodies Group of South Africa (Nabhan et al., 2016). hexagonal needles of nahcolite (NaHCO3), fishtail twins of selenite- Sand crystals are distinct from crystals formed in open water or gypsum (CaSO4.2H2O), or cubes of anhydrite (CaSO4). Barite is the saturated sediments, which are clean of inclusions because of displacive most common of pseudomorphed evaporite minerals reported from

7 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 8. Petrographic thin sections of evaporate pseudomorphs in the Panorama Formation of Western Australia, under plane light (A–C, E) and crossed nicols (D), showing redeposited barite at base of graded bed (A), nodularized barite (B), crystal rosette of barite (C), corroded (left) and sand crystal (right) of barite (D), and sand crystal of barite (E). Paleosols are type Jurl clay (A–B, E), Jurl clay loam (C) and type Wanta silt loam (D). Sand crystals are within dashed outlines. Specimens in Condon Collection, University of Oregon are R3756A (A), R3756B (B), R4309 (C), R3775 (D) and R3757 (E). other Archean cherts: 3.0 Ga Farrel Quartzite (Sugitani et al., 2003), weatherable minerals until the very top, and so is grouped with the 3.4 Ga Kromberg Formation (Lowe and Worrell, 1999), 3.4 Ga Strelley weakly developed profiles (Ngumpu). Pool Formation (Lowe, 1983; Allwood et al., 2007), 3.5 Ga Apex Chert (Pinti et al., 2009) and 3.5 Ga Dresser Formation (Lambert et al., 1978; 4.4. Major element geochemical trends Buick and Dunlop, 1990). In the Phanerozoic, gypsum-bearing paleo- sols are common (Retallack and Huang, 2010), but barite-bearing pa- Individual beds sampled as likely paleosols also show plausible leosols are rare and restricted to a few stratigraphic horizons (Tarr, trends in major element geochemical composition, expressed as mole- 1933; Beattie and Haldane, 1958; Olson, 1967; Skinner and Taylor, cular (molar) weathering ratios (Fig. 9C). Parent material in these thin 1967; Keyser, 1968; Laznicka, 1976; Bown and Kraus, 1988; Khalaf and alluvial paleosols is the material directly below (labelled C for inter- El Sayed, 1989; de Brodtkorb, 1989; Speczik and El Sayed, 1991; Garces preted sediment, or R for interpreted rock). Furthermore these thin and Aguilar, 1994; Kreuser, 1995; Dietrich, 1975; Schrank and profiles are unusually deeply weathered with high base loss (alumina Mahmoud, 2000; Retallack and Kirby, 2007; Retallack, 2009; Neuhaus, over bases) for soils with evaporites (Amundson et al., 2012). Alumina 2010; Buck et al., 2010; Brock-Hon et al., 2012; McCarthy and Plint, over bases shows little weathering for the one profile (Nganga), but 2013; Jennings et al., 2015). expected down-profile decline for others, especially marked in two superposed profiles (Wanta on Jurl profile). Another pronounced fea- 4.3. Petrographic trends ture is the high ratio of ferrous over ferric iron, like that of a water- logged (gleyed) soils in the modern world (Vepraskas and Sprecher, Point counting within individual beds also reveals changing pro- 1997). Unlike the little-weathered calcareous tuff (Nganga profile), the portions of minerals compatible with increased intensity of weathering cherty profiles have local enrichments of ferric iron, very low alkalies, from bottom to top: increased proportion of clay, quartz, and opaque alkaline earths and alumina, and modest salinization. These are features minerals toward the surface, and declining abundance of feldspar and of acidic modern soil-forming processes, such as podzolization rock fragments (Fig. 9B). Degree of weathering revealed by this metric (Retallack, 1997a). The combination of podzolization-gleization with varies from minimal (Ngumpu, Nganga) to moderate (Wanta, Jurl) in evaporite minerals is a distinctive feature of these and other Archean proportion with bedding disruption by crystal pseudomorphs. Some paleosols (Nabhan et al., 2016; Retallack et al., 2016), and some minerals such as quartz and opaque oxides are resistant to weathering, modern soils (Benison and Bowen, 2015; Jennings et al., 2015). but feldspar and rock fragments weather to clay in soils (Retallack, 1997a). Furthermore, the extent of these weathering trends vary with 4.5. Tau analysis the degree to which original sedimentary bedding has been homo- genized by growth of nodules and cracking (Fig. 9A), revealing deeper Geochemical mass balance or “tau analysis” is a method for calcu- chemical weathering associated with physical homogenization. lating both the mobilization of particular elements (mass transfer) and Barite pseudomorphs were found only in the cherty beds, not in the the change in volume (strain) of soil formation (Brimhall et al., 1992). highly calcareous bed considered an alkaline tuff from its unusual Mass transfer (τw,j in mass fraction) is derived from chemical con- abundance of light rare earth elements (Fig. 5). That thick calcareous centration of the element in soils (Cj,w in weight %) compared with bed (Nganga profile) shows muted changes in proportions of parent material (Cp,w in weight %), correcting for bulk density of the

8 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 9. Mineral composition and molecular weathering ratios of paleosols in the Panorama Formation at Panorama Portal, Western Australia.

− 3 − 3 soil (ρw in g cm ) and parent material (ρp in g cm ). Strain (εi,w as a Soils and paleosols have a distinctive location in this style of plot- mass fraction) is estimated from an immobile element in soil (such as Ti ting geochemical data: in lower left quadrant of volume collapse and used here). The relevant Eqs. (1) and (2) (below) are the basis for elemental loss when plotted as mass transfer with strain (Fig. 10A), and calculating divergence from parent material composition. with depletion up-profile when plotted as mass transfer with depth (Fig. 10B). This is where the Panorama beds plot with a few exceptions, ⎡ ρCw⋅ jw, ⎤ primarily of the Nganga profile, which is a little weathered tuff τjw, = [1]εiw, +−1 ⎢ ρC⋅ ⎥ (Fig. 9C), and so plots in the dilate-and-gain quadrant of mass transfer ⎣ p jp, ⎦ (1) with strain like a sedimentary or tuffaceous bed (Fig. 10A). The Ngumpu profile also shows little chemical differentiation, and may also ρCp⋅ jp, ε = ⎡ ⎤ − 1 have been a very weakly developed soil. Tight clustering of silica, iw, ⎢ ⎥ ⎣ ρCw⋅ jw, ⎦ (2) alumina, alkalies and alkaline earth elements in this kind of analysis

9 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 10. Tau analysis of paleosols in the Panorama Formation of Western Australia, showing (A) mass transfer with strain and (B) mass transfer with depth, and calculated by method of Brimhall et al. (1992).

(Fig. 10A) over distances of only a few tens of centimeters (Fig. 10B) is 5. Interpretation of paleosols evidence of chemical weathering. The named beds of the Panorama Formation appear to be paleosols, which are of interest because they yield information not only about the 4.6. Trace element trends diversity of landforms of the past, but about a variety of paleoenvir- onmental conditions ranging from time for formation to climate and Chemical differences over short depths within the profiles are also atmospheric composition (Retallack, 1997a). Each of the profiles seen in the abundance of rare earth elements, which can vary by an named on the basis of their field appearance (Table 1) represents a order of magnitude within a single profile of the Panorama Formation different combination of formative influences, which are interpreted in (Fig. 5). The light rare earth enrichment of the Nganga bed is attributed the following paragraphs (Table 2). to alkaline tuff parent material different from that of other beds, but depleted in rare earth elements in the subsurface (C) horizons (Fig. 5A). Profiles with evaporite pseudomorphs (Wanta and Jurl) also have de- 5.1. Classification of panorama formation paleosols pletion in lower levels (Fig. 5A). A profile lacking evaporites and with clear bedding (Ngumpu profile) shows little change in rare earth ele- Field observations, together with geochemical and petrographic ment composition at different levels, and a flatter pattern than the other information (Fig. 9) on the paleosols allow interpretation within clas- profiles (Fig. 5B). These depletions within profiles over only a few sifications of modern soils of the various field pedotypes recognized in centimeters and heavy-rare-earth depletion (Jurl and Wanta versus the Panorama Formation (Table 1). Point count data (Fig. 9) also give Ngumpu and Nganga) are evidence of acidic humid weathering at the the option of naming particular paleosols from soil textural categories level of the evaporate pseudomorphs (Nesbitt, 1979). These weathering of their surface horizons, such as Jurl silt loam versus Jurl clay loam. trends can be contrasted with sedimentary redistribution and marine Persistence of bedding, thin profiles, and muted chemical variation diagenesis of sediments. Sedimentary processes separate rare earth mark the Nganga and Ngumpu profiles as in the U.S Taxonomy elements into clay rather than sand (Morey and Setterholm, 1997), but (Soil Survey Staff, 2014). Surface lamination like banded iron forma- the Panorama beds are poor in clay and do not show large changes in tion, an aquatic facies of the Panorama Formation (Bolhar et al., 2005), grain size (Fig. 9). Marine diagenesis of sediments (halmyrolysis) does may be evidence that Wanta paleosols were intermittently waterlogged, not change rare earth concentrations significantly (Clauer et al., 1990; as well as developing small barite nodules and crystals below when Setti et al., 2004), unlike the relationships seen in the Panorama For- better drained, so gleyed . Although banded iron formations mation (Fig. 5). have been considered deep water, this does not seem to be the case for

10 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Table 2 Summary of Panorama Formation paleosol interpretations.

Pedotype Climate Organisms Topography Parent material Soil duration

Jurl Humid (1154–1343 ± 182 mm MAP) temperate (8.7–10.3 ± 0.5 °C MAT) Microbial earth Floodplain Quartz-lithic silt 2408–70,151 years Nganga Not diagnostic for climate Uncertain Coastal dunes Calcareous sand 10–1000 years and gravel Ngumpu Not diagnostic for climate Microbial earth Streamside bar Quartz-lithic sand 10–1000 years Wanta Not diagnostic for climate Microbial earth dominated Floodplain Quartz-lithic sand 1571–69,505 years by iron-oxidizing bacteria swale

Note: MAP is mean annual precipitation and MAT is mean annual temperature.

Table 3 Occurrences of pedogenic barite.

Ma Geological age Formation Location Reference

0.015 ± 0.0011 Lufkin loam soil College Station, Texas Jennings et al., 2015 0.015 ± 0.0011 Pleistocene Falba fine sandy loam soil Huntsville, Texas Carson et al., 1982 0.014 ± 0.0010 Pleistocene Bluford silt loam soil Bluford, Illinois Darmody et al., 1989 0.100 ± 0.05 Pleistocene High terrace Bellavista, Peru Stoops and Zavaleta, 1978 2 ± 2 Pleistocene High terrace calcrete Mormon Mesa, Nevada Brock-Hon et al., 2012 5 ± 2 Pliocene Brucedale Parna Wagga Wagga, New South Wales Beattie and Haldane, 1958 10 ± 6 Mio-Pliocene Kuwait Group Kuwait Khalaf and El Sayed, 1989 10 ± 1 Edelény variegated clay Rudabánya, Hungary de Brodtkorb, 1989 16 ± 1 Cucaracha Formation Panama Retallack and Kirby, 2007 16 ± 1 Barstovian Fort Randall Formation Bijou Hills, South Dakota Skinner and Taylor, 1967 26 ± 0.2 Chattian Rockenberg Formation Rockenberg, Germany Neuhaus, 2010 27 ± 2 Arikareean Monroe Creek Formation Ovid, Colorado de Brodtkorb, 1989 35 ± 5 Chadronian Carroza Formation La Papa, Mexico Buck et al., 2010 36 ± 1 Priabonian Qasr el Sagha Formation Fayum Depression, Egypt Bown and Kraus, 1988 50 ± 5 Ypresian El Naab Formation Bahariya Oasis, Egypt Speczik and El Sayed, 1991 90 ± 10 Cenomanian Dunvegan Formation Clayhurst, British Columbia McCarthy and Plint, 2013 90 ± 10 Cenomanian Sabaya Formation Kharga Oasis, Egypt Schrank and Mahmoud, 2000 90 ± 10 Cenomanian Bahariya Formation Bahariya Oasis, Egypt Speczik and El Sayed, 1991 135 ± 5 Berriasian Morrison Formation Como Bluff, Wyoming Retallack, 2009 151 ± 2 Tithonian Morrison Formation Thermopolis, Wyoming Jennings et al., 2015 200 ± 5 Hettangian Elliott Formation Bethleham, South Africa Keyser, 1968 235 ± 2 Carnian Cacheuta Formation Cacheuta, Argentina de Brodtkorb, 1989 250 ± 3 Griesbachian Upper Beaufort Group Thaba Ncha, South Africa Keyser, 1968 260 ± 2 Guadalupian Unidad Roja Superior Jaca, Spain Garces and Aguilar, 1994 270 ± 2 Kungurian Mbuyura Formation Ruhuhu, Tanzania Kreuser, 1995 270 ± 2 Kungurian Garber Sandstone Norman, Oklahoma Tarr, 1933, Olson, 1967 270 ± 2 Kungurian Rotliegend Rothenberg, Gembeck, Germany Dietrich, 1975 359 ± 10 Late Horn River Formation Twitya River, NWT, Canada Laznicka, 1976 2765 ± 10 Archean Mt Roe Basalt Whim Creek, Western Australia Teitler et al., 2015 3016 ± 13 Archean Farrel Quartzite Mt Grant Western Australia Retallack, 2014a 3220 ± 10 Archean Moodies Group Barberton, South Africa Nabhan et al., 2016 3458 ± 1.9 Archean Panorama Formation North Pole Dome, Western Australia Herein banded iron formation in the Panorama Formation (Bolhar et al., 2005), wooded “wheat belt”, for a range of mean annual precipitation of and likely photoferrotrophic origin is not possible in deep water (Crowe 260–340 mm (Benison and Bowen, 2013, 2015). This region has mean et al., 2008). There are no baritic soils in the US taxonomy, although annual temperature of 18–20 °C (Bureau of Meteorology, 2016). These some have abundant barite at depth, and baritic paleosols are parts of Western Australia have a mix of desert soils developed on locally abundant (Table 3). Nevertheless, Gypsids (with gypsum rather deeply weathered lateritic residuum from thick Ultisol and pa- than barite in Soil Survey Staff, 2014) are a match for Jurl paleosols in leosols of the Miocene (McKenzie et al., 2004). These Solonchak soils profile form and evaporite density. Comparable considerations can be are also called “acid saline lakes” (Benison et al., 2007; Benison and used to classify the paleosols in classifications of Australia (Isbell, 1996) Bowen, 2013), even though dry more often than flooded. A distinctive and of the Food and Agriculture Organization (1974). Classification of feature of these lakes despite abundant salts is very acidic waters the Jurl paleosols in the FAO system as a Solonchak (among others in (pH 1.4–3.9), due to oxidation of sulfides unbuffered by local carbonate Table 1) opens the possibility of finding modern units com- in soils or (Benison and Bowen, 2015). Halite and gypsum are parable with the Panorama paleosol assemblage, which can be de- the principal evaporite minerals of the acidic lakes (Benison and scribed by the FAO code as Zo with inclusions of Zg, Jd, and Rc (Orthic Bowen, 2013), but barite is also recorded in Western Australian acid Solonchak, with Gleyic Solonchak, Dystric Fluvisol and Calcaric Re- sulfate soils with pH < 3 (Fitzpatrick et al., 2008). Another analog is gosol). The closest match to this in all mapped soil units around the cold sulfur springs where barite is precipitated at pH 7.3–8.4 from world is map unit Zo36-2a, with associated Zg, and inclusions of Rc, Rd, oxidation and neutralization of anoxic and acidic brines from the sub- Xk and Yh (Calcaric Regosol, Dystric Regosol, Haplic Xerosol, and surface (Donovan et al., 1988; Bonny and Jones, 2008a, 2008b). Haplic Yermosol) on alluvial sediments widespread in Western Aus- tralia, but most extensive near Lakes Disappointment and Carnegie (Food and Agriculture, Organization, 1978). Such soils are found in 5.2. Parent material regions of arid grasslands and shrublands, but extend into the formerly Two distinct parent materials for paleosols of the Panorama

11 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Formation are revealed by rare earth element analyses (Fig. 5) and formulae: point counting (Fig. 9), and these are now calcareous tuff and cherty AS= 3920· 0.34 (3) tuff. The calcareous tuff is an unsilicified lithic tuff with only thin and 2 weakly developed paleosols (Nganga pedotype). The cherty tuffs were with r = 0.57, s.e. = ± 1800 yrs also largely a lithic tuff, although acicular shapes are suggestive of a AG=+3987· 5774 4 vitric component, which may have aided silicification. Interbedded with r2 = 0.95, s.e. = ± 15,000 yrs flows and chemical composition of the Panorama Formation has been If the sand crystals and nodules of the Panorama Formation were taken as evidence that these silicified tuffs were originally rhyodacitic calcite, their size in 13 Jurl paleosols would be evidence of in composition (Cullers et al., 1993). These were parent materials to 4208 ± 1800 yr, and in 7 Wanta paleosols, of 3371 ± 1800 yr most of the paleosols recognized: Ngumpu, Jurl and Wanta pedotypes. (Supplementary information Table S6). If on the other hand, they were gypsum crystals and had densities measured in the field, the duration of 5.3. Paleotopographic position 13 Jurl paleosols average 55,151 ± 15,000 years, and 7 Wanta pa- leosols average 44,505 ± 15,000 years. These estimates vary by an Evidence that the parent tuffs formed a well-drained volcanic apron order of magnitude, but provide an envelope of possibilities. A 130Xe comes from facies analysis using observations such as accretionary la- isotopic estimate of exposure time before burial of < 40,000 years was pilli (van Kranendonk, 2000), and measurements such as paleocurrents obtained for bedded (non-pedogenic) barite of the 3.48 Ga Dresser and stratigraphic thickness variations (Cullers et al., 1993). Calcareous Formation (Pujol et al., 2009). capped by Nganga paleosols have cross bedding 1.5 m thick (Fig. 2C), which may have been coastal eolian dunes with as much 5.5. Paleoclimate relief. Then there is the evidence of abundant barite sand crystals in Jurl and Wanta paleosols, which requires desiccation and precipitation Chemical index of alteration (I) is a general ratio of alumina over in a porous medium (Tarr, 1933; Dietrich, 1975). The distribution of bases, calculated as follows barite, scarce at the surface and peaking in abundance a short distance below (Figs. 2D–F, 9), is comparable with the distribution of gypsum in 100(mAl O ) I = 23 excessively drained modern desert soils (Retallack and Huang, 2010; [(mAl23 O )(++ mCaO )( mNa 2 O )( + mK 2 O )] (5) Amundson et al., 2012). Barite in modern soils forms near the water This ratio is < 65 in glacial-frigid climates and > 80 in tropical table, because it requires neutralization, evaporation and oxidation climates, but between these limits in temperate climates (Nesbitt and from highly acidic and chemically reducing (Jennings and Young, 1989). Burial illitization would reduce this value toward cooler Driese, 2014; Jennings et al., 2015). Barite nodules in Jurl and Wanta climates, but was not a discernible problem for most of the early-sili- paleosols do not extend much below 30 cm from the top of the profile, cified paleosols (Fig. 6). The 19 available analyses of the cherty pa- and this may have been the level of water table in dry playas, com- leosols (Ngumpu, Jurl, Wanta), and of their parent materials as a guide parable with Western Australian saline lakes (Benison and Bowen, to sedimentary source terranes, have a chemical index of alteration of 2015). 67 ± 12, which suggests cool temperate to periglacial climate. Another indication of high water table may be laminae of hematite More specific pedogenic paleothermometers are based on soils of in the surface horizons of Wanta paleosols (Fig. 2G). These are most like modern woody vegetation (Sheldon et al., 2002; Gallagher and banded iron formations elsewhere in the Panorama Formation (Bolhar Sheldon, 2013), not applicable to likely microbial earths of Archean et al., 2005), thought to have formed in an aqueous medium by mi- paleosols (Retallack, 2014a). A paleothermometer based on modern crobial photoferrolysis (Crowe et al., 2008). Although widely con- soils under lichen-shrub tundra vegetation of Iceland (Óskarsson et al., sidered marine, banded iron formations also have been recorded from 2012) may be the best available option, predicting temperature (T in freshwater crater lakes (Bolhar et al., 2005), and fluviodeltaic facies °C) from chemical index of weathering (W), another base depletion (Pufahl et al., 2013). This situation is the reverse of the situation in metric, as in the following formula modern soils, in which waterlogged soils are gray-green and well drained soils red-brown (Vepraskas and Sprecher, 1997). Such upside- 100(mAl O ) W = 23 down relationships have been predicted as a consequence of [(mAl23 O )(++ mCaO )( mNa 2 O )] (6) anoxic Archean atmosphere (Walker, 1987; Domagal-Goldman, 2014). TW= 0.21 –8.93 (7) 5.4. Time for formation with r2 = 0.81, s.e. = ± 0.5 °C. These calculations give the following temperate mean annual pa- Paleosols within the Panorama Formation show different degrees of leotemperatures for the lower A horizons of moderately developed destruction of bedding and ripple marks as if they were moderately paleosols: 10.3 ± 0.5 °C for type Jurl clay, 9.1 ± 0.5 °C for Jurl silt (Jurl, Wanta) to very weakly developed (Ngumpu, Nganga) in the field loam, and 8.7 ± 0.5 °C for Jurl clay loam. These results are consistent scale of Retallack (1997a). The application of this scale to these pa- with “temperate paleoclimate” (< 40 °C) interpreted for the 3.4 Ga leosols is problematic because their nodules are barite, rather than Buck Reef Chert of South Africa, from combined oxygen and hydrogen gypsum (Retallack and Huang, 2010), or low‑magnesium calcite of isotopes (Hren et al., 2009), revising earlier estimates of 55-85 °C from modern soils (Retallack, 2005).There are no modern chronosequences oxygen isotopes alone (Knauth and Lowe, 2003). Temperatures < 75 °C for barite in soils (Jennings and Driese, 2014; Jennings et al., 2015). are indicated by 6–24 vol% original quartz in the Panorama Formation Barite is rare under acidic (< pH 3) conditions of saline lakes with paleosols (Fig. 9): most quartz would have dissolved at higher tem- gypsum and halite (Fitzpatrick et al., 2008), but barite also precipitates peratures (Sleep and Hessler, 2006). One counterargument is that ac- from waters of pH 7.3–8.4 in cold sulfur springs (Donovan et al., 1988; tive volcanism may have buried quartz before it dissolved (Lowe, Bonny and Jones, 2008a, 2008b). 2007), but quartz was exposed for millennia given the degree of che- Despite these caveats, nodules in soils take millennia to coalesce mical weathering of the paleosols (Figs. 5 and 10). Temperate paleo- (Retallack, 1997a). Two transfer function for estimating paleosol ages temperatures are reasonable considering likely mid-paleolatitude posi- (A in years) come from diameter of low magnesium calcite nodules (S in tion of the Panorama Formation: between 20.5 ± 5o for the ca. cm) in radiocarbon-dated soils of New Mexico (Retallack, 2005), and 3.46 Ga upper Apex Chert (Suganuma et al., 2006), and 59.0 ± 8.8o abundance of gypsum crystals (G as % surface area) of the Negev and for the 2.86 ± 0.2 Ga Millindinna Complex of the West Pilbara Atacama Deserts (Retallack, 2013a), according to the following (Schmidt and Embleton, 1985).

12 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

A widely used indicator of mean annual precipitation (mm) for sun. Other greenhouse gases are needed, including water vapor, CH4, paleosols (Sheldon et al., 2002) uses chemical index of alteration C2H6,SO2, and COS (carbonyl sulfide: Ohmoto et al., 2014; Haqq-Misra without potash (A of Eq. (5) without mK2O), in the following re- et al., 2008; Ueno et al., 2009; Rosing et al., 2010; Kasting and lationship. Kirschvink, 2012). is allowed by low (160 ppm) H2 values likely for the Archean (Kasting and Kirschvink, 2012). An atmosphere P = 221e0.0197W (8) with three times the current mass of N2 and a H2 mixing ratio of 0.1, 2 with r = 0.72, s.e. = ± 182 mm. would also have created an adequate greenhouse (Wordsworth and

This paleohyetometer gives humid mean annual precipitation: Pierrehumbert, 2013), but N2 in the atmosphere was limited to 1.1 to 1343 ± 182 mm for the type Jurl clay, 1154 ± 182 mm for Jurl clay 0.5 bars judging from nitrogen and argon isotopic ratios in fluid in- loam, 1200 ± 182 mm for Jurl silt loam. These estimates are based on clusions of the 3.5 Ga Dresser Formation of Western Australia (Marty soils of vascular unlike Archean vegetation, if any, so must be et al., 2013). The gas SO2 proposed theoretically for early Mars (Halevy treated with suspicion. However, other indications of weathering such et al., 2007) and early Earth (Claire et al., 2014) is compatible with the as light rare earth enrichment of the Panorama Formation (Fig. 5) are paleovolcanic setting and abundance of sulfate in paleosols of the Pa- comparable with modern weathering under forests (Nesbitt, 1979; norama Formation. Morey and Setterholm, 1997). Abundant evaporite pseudomorphs in paleosols within the 5.7. Paleoatmospheric oxygen Panorama Formation give the superficial appearance of soils of arid paleoclimates (Retallack and Huang, 2010). Unlike gypsum reported Very low levels of atmospheric oxygen are indicated by high ferrous from some Archean paleosols (Nabhan et al., 2016), crystals and no- to ferric iron ratios of paleosols in the Panorama Formation (Fig. 9), dules pseudomorphed by chert in the Panorama Formation may have comparable with swamp soils today exhausted of oxygen by microbial – been barite (BaSO4), as indicated by elevated Ba (8.2 3670 ppm) and respiration (Vepraskas and Sprecher, 1997). However, paleosols of the – crystal form (Fig. 8D E). Modern pedogenic barite crystals and spher- Panorama Formation were not waterlogged, as indicated by sand ulites are mostly microscopic (Kohut and Durdas, 1993; Shahid and crystals and nodules of barite (Fig. 8), like those of other Archean well Jenkins, 1994; Howari et al., 2002), but some soils have barite nodules drained paleosols (Nabhan et al., 2016). Laminations of hematite and up to 3.8 cm in diameter (Jennings and Driese, 2014; Jennings et al., goethite are a feature of the uppermost surface of Wanta paleosols 2015) like those of some paleosols (Retallack and Kirby, 2007; Jennings (Fig. 2G), comparable with aquatic banded iron formation in the Pa- et al., 2015). Pedogenic barite is best known in humid forested soils norama Formation (Bolhar et al., 2005). The 3.5 Ga Marble Bar Chert is – with 1000 1200 mm mean annual precipitation, and is controlled more also an aquatic banded‑iron formation with original hematite (Hoashi by microbially induced precipitation, and pH or redox changes at the et al., 2009), best explained by chemolithotrophy of iron-oxidizing water table, than by paleoclimate (Retallack and Kirby, 2007; Jennings bacteria under reducing conditions (Konhauser et al., 2002; Crowe et al., 2015). In cold sulfur springs barite is precipitated from oxidation et al., 2008). In most cases Archean and Cenozoic hematite can be and neutralization of saline anoxic brines from the subsurface (Donovan distinguished by Munsell hue, because oxidation of the modern outcrop et al., 1988; Bonny and Jones, 2008a, 2008b). Barite is favored by acid is brownish red (10YR-7.5YR), whereas Archean hematite is fire engine sulfate weathering at low pH (< 3) and is insoluble at higher pH, but red (5R to 10R). Exceptions are Archean rocks locally stained by – gypsum forms and also is soluble at higher pH (4 9: Carson et al., overlying and Cenozoic paleosols (McLoughlin, 1996; Morris 1982). There is no clear relationship between depth of barite and mean and Ramanaidou, 2007), or local oxidation along deep fissures (Kato annual precipitation as there is for gypsum (Retallack and Huang, et al., 2009; Li et al., 2012). 2010). Deeply weathered Panorama paleosols are a guide to soil exposure

to O2, and thus partial pressure of atmospheric O2 (pO2 in atmo- 5.6. Paleoatmospheric spheres). This can be done by modifying the method of Sheldon (2006), using integrated whole profile oxidation of iron and manganese instead Paleosols of the Panorama Formation are thin but deeply weathered of base loss, by Eq. 9 below. (Fig. 9) and could be used to calculate soil CO2 using the paleoba- F rometer of Sheldon (2006), if barite did not introduce the probability of pO = 2 KP Dα acid sulfate weathering found in modern baritic soils (Jennings and Aκ⎢ OO22+ ⎥ ⎣ 1000 L ⎦ (9) Driese, 2014). The model of Sheldon (2006) is predicated on alkali and alkaline earth depletion by carbonic acid from soil CO2, but these Cjp, ZD= jw, M = 2 ∑ ρp ∫ τδZjw,() z baritic paleosols may have been weathered by sulfuric acid as well. The 100 Z=0 (10) strong chemical weathering observed is compatible with high Archean CO levels of 30,000 to 300,000 ppmv (100–1000 PAL) inferred by −0.51 2 B = Lowe and Tice (2004) from nahcolite formation at assumed Archean 0.49 − 1 e0.27k (11) temperatures of 70 °C, but such high temperatures are not supported by chemical composition of Panorama paleosols (section 6.5), oxygen and Variables needed for these calculations are F (mol O2/cm2) hydrogen isotopic studies (Hren et al., 2009), or preservation of quartz = summed molar mass transfer gains of Fe2O3 and MnO through the −3 (Sleep and Hessler, 2006). A minimum level of 2500 ppmv (8.9 PAL) profile using Eq. (10); ρp (g.cm ) = bulk density of parent material; CO2 has been inferred from 3.2 Ga ferrous‑carbonate weathering rinds Cj,p (weight %) = chemical concentration of an element (j) in parent (Hessler et al., 2004), but such fluvial pebbles may have been isolated material (p); τ,j,w (mass fraction) = mass transfer of a specified (j) from the atmosphere by groundwater. Estimates of 1500–9000 ppmv element in a (w); Z (cm) = depth in soil represented by

(5–32 PAL) CO2 for 3.0 Ga come from chemical weathering of the analysis; A (years) = duration of soil formation calculated here using ff Jerico Dam paleosol of South Africa (Grandsta et al., 1986). This much Eq. (3); KO2(mol./kg.bar) = Henry's Law constant for O2 (=0.00125, CO2 would give acid rain (pH 4.0–4.5 according to Ohmoto et al., range 0.0012–0.0013 from Aachib et al., 2004); P (cm) = mean annual 3 2014), and soil-microbial CO2 and H2SO4 could drop pH to 3 precipitation calculated here using Eq. (8); κ (s cm /mol year) = sec- favoring acid sulfate weathering (Carson et al., 1982). onds per year divided by volume per mole of gas at standard tem- 2 ff Even the most extreme of these estimates is short of the amount perature and pressure (=1409); DO2(cm /s) = di usion constant for O2 needed for a greenhouse capable of maintaining likely temperate in air (=0.203 at 20 °C, range from 0 to 40 °C is 0.179–0.227 from

Archean paleotemperatures (Hren et al., 2009) given the faint young Denny, 1993); α (fraction) = ratio of diffusion constant for O2 in soil

13 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

divided by diffusion constant for O2 in air (=0.2, range 0.09–0.32 from Modern acid-sulfate soils are not inimical to life but have a distinctive Aachib et al., 2004); L (cm) = original depth to water table (after de- array of acidophilic microbes, including Actinobacteria, Proteobacteria compaction using fractional compaction B); B (fraction) = compaction and Bacteroidetes in acidic Gypsids () of Western Australia of due to burial (using Eq. (11) after Sheldon and Retallack, (Benison et al., 2008; Mormile et al., 2009). 2001); k (km) = depth of burial (=13.826 km following van A line of evidence commonly used to infer life in Precambrian pa- Kranendonk, 2000; van Kranendonk et al., 2006). Each one of these leosols does not work for the Panorama Formation. Significant molar variables has its own error, and Gaussian error propagation has been depletion of phosphorus (Fig. 10) requires organic ligands at pH > 5.5 calculated for O2 following Hughes and Hase (2010) and equations and stable aluminum (Neaman et al., 2005). Aluminum may appear given in Supplementary information Table S7. erratic or retained compared with other elements (Figs. 6,9), but tau

The results for Archean O2 are 711 ± 131 ppm analysis shows that aluminum was mildly leached from the paleosols (0.0034 ± 0.00063 times preindustrial atmospheric level or PAL) for except for ferruginous top of Wanta (Fig. 10). This and abundant barite the Jurl clay, 75 ± 191 ppm (0.00026 ± 0.00091 PAL) for the Jurl are evidence of pH < 3 (Carson et al., 1982; Fitzpatrick et al., 2008). clay loam, 88 ± 222 ppm (0.00042 ± 0.0011 PAL) for the Jurl silt Thus apatite may have been dissolved abiotically by sulfuric acid in loam, and 57 ± 208 ppm (0.00027 ± 0.00099 PAL) for the Wanta these paleosols. silt loam. These are maximal levels of O2 because based on the short soil There is evidence of life in intertidal stromatolites before and after formation times inferred from Eq. (3), rather than Eq. (4) which would deposition of the Panorama Formation from permineralized micro- give minimal levels of O2 an order of magnitude lower. These estimates fossils in the overlying Strelley Pool Formation (Sugitani et al., 2010, of atmospheric O2 based on Panorama paleosols are within the range of 2013; Wacey et al., 2010, 2011; Brasier et al., 2015; Duda et al., 2016) 20–1000 ppm (0.00095–0.0047 PAL) O2 estimated by a different and underlying Mt. Ada Basalt (Awramik et al., 1983, 1988; van method of calculating oxygen demand of the 3.0 Ga Jerico Dam pa- Kranendonk, 2006; Hickman, 2013) and Dresser Formation (van leosol of South Africa and associated uraniferous conglomerates Kranendonk et al., 2008). Even older claims of aquatic life in stroma- (Grandstaff et al., 1986). Other comparable estimates of 12–63 ppm tolites and banded iron formations come from 3.7-Ga rocks west

(0.0006–0.003 PAL) O2 come from lack of oxidative Cr and U recycling Greenland (Nutman et al., 2016) and 4.2-Ga rocks of northern Quebec in 2.9 Ga Nsuze paleosols of South Africa (Crowe et al., 2013), and of (Dodd et al., 2017).

20–200 ppm (0.00095–0.0095 PAL) O2 from cerium anomalies in the 3.0–3.3 Ga Keonjhar paleosol deep weathering remnant of India 6. Archean acid sulfate weathering

(Mukhopadhyay et al., 2014). Low atmospheric O2 is also indicated by redox-sensitive minerals, including pyrite (FeS), uraninite (UO2), side- Baritic paleosols are common and widespread in silicified tuffsof rite (FeCO3) and gersdorfffite (NiAsS), common (57–85%) in heavy the 3.46 Ga Panorama Formation. Although exposure of the Panorama mineral separations from fluvial siliciclastic sediments some 3.2–2.7 Ga Formation is imperfect (Fig. 2A), detailed sections show an average of in age in the Pilbara region (Rasmussen and Buick, 1999). Consequent 1.4 baritic paleosols per meter, and with 90 m of suitable facies in the lack of an ozone layer in the stratosphere is also indicated by mass- formation (Fig. 3), the formation may contain at least 126 successive independent fractionation of sulfur isotopes in Archean pyrites and baritic paleosols. Furthermore, baritic paleosol intervals can be traced barites of the Pilbara region (Ueno et al., 2008; Shen et al., 2009). In over 4 km laterally from the fault separated volcanic vent facies toward contrast, Ohmoto et al. (2014) has argued for 100,000 ppm (0.48 PAL) the northern margin of outcrop (Fig. 1). Individual paleosols can be atmospheric O2 based on the 3.42 Ga pre-Strelley paleosols of Western walked 2 km laterally along the most resistant ridges (Fig. 2A). These Australia (Ohmoto et al., 2014), and considers the chemically reduced paleosols formed on a small dacitic stratovolcano (Fig. 11), but baritic condition of the Jerico Dam paleosol due to hydrothermal fluids or paleosols in the 3.2 Ga Moodies Group of South Africa formed in quartz burial gleization by organic acids after formation in an oxidizing at- sandstones of braided streams in a stable cratonic back-arc basin with mosphere of at least 3000 ppm (0.014 PAL) O2 (Ohmoto, 1996). My reworked rhyodacitic tuffs(Nabhan et al., 2016). In South Africa, own field inspection of the pre-Strelley “” of Ohmoto et al. baritic paleosols are common within 220 m of section traceable re- (2014) found that they were instead banded iron formations unrelated gionally for 80 km laterally. Baritic nodular paleosols were thus wide- to the paleosol. There is no clear hydrothermal enrichment of heavy spread globally during the Archean, and this contrasts with the rarity of rare earth elements in the Jericho Dam profile (Kimberly and barite in modern and soils (Table 3). Grandstaff, 1986) and burial gleization to 3.3 m depth in that paleosol Baritic paleosols represent a distinctive style of acid-sulfate weath- is unlikely considering an order of magnitude less gleization in paleo- ering rare in three distinct settings of modern soils. First are thin pro- sols of productive forest ecosystems (Retallack, 1997b). files of Gypsids (Solonchaks, Sodosols) in and around acidic playa lakes in subhumid to semiarid regions (McKenzie et al., 2004; Fitzpatrick 5.8. Organisms et al., 2008; Benison and Bowen, 2013, 2015). Second are at depths of a meter or so near the water table in acidic Ultisols (Podzoluvisols) of Direct evidence of life in paleosols of the Panorama Formation re- humid uplands (Stoops and Zavaleta, 1978; Darmody et al., 1989; mains elusive, although Jurl paleosols in the 3.0 Ga Farrel Quartzite Jennings and Driese, 2014; Jennings et al., 2015). Third are thick ter- included microfossils interpreted as photosynthetic sulfur bacteria, races of Entisols (Solonchaks) around cold sulfur springs of humid cli- methanogens and actinobacteria (Retallack et al., 2016). Microbes may mates (Donovan et al., 1988; Bonny and Jones, 2008a, 2008b). The explain how paleosols in the Panorama Formation have such oxidized playa option may be the best analog to Archean paleosols considering minerals as barite (Wanta and Jurl) and hematite (Wanta) in the nearly their profile form, but barite is rare in such settings (Fitzpatrick et al., anoxic atmosphere inferred for these paleosols (section 5.7). Photo- 2008), and not included in the list of minerals by Benison and Bowen synthetic sulfur bacteria (Chromatiaceae and Chlorobiaceae; Youssef (2013). The deep water table option sometimes has common and large et al., 2010) play a major role in precipitation of barite in modern cold nodules (Jennings et al., 2015), and is like barite nodules in the 2.65-Ga sulfur springs within 12 m of the source, as demonstrated by diel Mt. Roe paleosol (Teitler et al., 2015), but is not like the shallow baritic fluctuations in aqueous sulfate and incubations, and stable isotopic paleosols of Panorama Formation (Fig. 9), Farrel Quartzite (Retallack comparisons at Zodletone Spring, Oklahoma (Senko et al., 2004). He- et al., 2016), or Moodies Group (Nabhan et al., 2016). The barite de- matite in banded iron formations is thought to have formed in an anoxic posits of cold sulfur springs match the abundance of barite in Archean aqueous medium by microbial photoferrolysis (Konhauser et al., 2002; paleosols, but are limited in lateral distribution and have very different Crowe et al., 2008), and a comparable origin is likely for banded iron textures, including cavity-filling barite fans and stromatolite-like iso- formation within the Panorama Formation (Bolhar et al., 2005). pachous barite coatings (Donovan et al., 1988; Bonny and Jones,

14 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Fig. 11. Graphical abstract of reconstructed soils in the 3.46 Ga Panorama Formation of Western Australia.

2008a, 2008b). 7. Conclusions Baritic paleosols are rare through the Phanerozoic (Table 3), when they are found mainly at times of CO2 greenhouse crises (Retallack and Baritic paleosols are common and widespread, not only in volcanic Kirby, 2007). Gypsum-bearing paleosols are much more common at apron facies of the 3.46-Ga Panorama Formation, but in thick paleosols other times in the fossil record (Retallack and Huang, 2010). At several between flows of 2.76-Ga Mt. Roe Basalt of Western Australia, on points during -Triassic transition, for example, CO2 estimated floodplain facies of the 3.2-Ga Moodies Group of South Africa, and from stomatal index reached an astounding 7832 ppm (Retallack, coastal plain facies of the 3.0-Ga Farrel Quartzite of Western Australia 2009), with consequent decline in both pH and Eh of paleosols (Table. 3). At 3.46 Ga, baritic paleosols of the Panorama Formation are (Retallack et al., 2011; Retallack, 2013b). Such transient atmospheric currently the oldest known on Earth. Baritic nodular paleosols of acid crises can be regarded as a reversion of life history toward Precambrian sulfate weathering may have been widespread during the Archean, in biotic and atmospheric conditions, enabling a resurgence of baritic contrast with the rarity of barite in modern soils and Phanerozoic pa- paleosols. leosols. Barite in soils requires extreme fluctuation of pH and redox, and A common thread in all occurrences of barite in soils is extreme can only be transported and concentrated in solution at pH < 3. In acid fluctuation of pH and redox (Carson et al., 1982; Jennings et al., 2015). sulfate modern soils, such low pH is created by sulfuric acid from the Barium and barite is remarkably insoluble at pH > 3, which promotes oxidation of pyrite in an oxidizing atmosphere. Archean paleosols such its persistence and precipitation, but can only be transported and con- as those of the Panorama Formation have utilization of iron and man- centrated in solution at pH < 3 (Tarr, 1933; Carson et al., 1982). In ganese indicative of anoxic atmosphere, so that oxidation may have acid sulfate soils such low pH is created by sulfuric acid from the oxi- been provided by photosynthetic sulfur-oxidizing bacteria. Unlike post dation of pyrite, and the pyrite in turn requires an anoxic environment Archean weathering dominated by fungi and weak acid (carbonic acid), (Carson et al., 1982). Thus unlike post Archean weathering dominated Archean weathering may have been largely produced by sulfur bacteria by the weak acid, carbonic acid, Archean weathering may have been and strong acid (sulfuric acid). Archaic and widespread surficial acid- largely produced by the strong acid, sulfuric acid. This difference may sulfate weathering has been driven underground and into marginal have contributed to the surprising degree of weathering of Archean environments by Palaeoproterozoic atmospheric oxidation, into playa shales and paleosols weathered to bauxites (Serdyuchenko, 1968; Dash lakes, cold sulfur springs, and deep water tables within soils. et al., 1987) considering likely low productivity microbial earth com- munities (Retallack, 2014a), or perhaps abiotic weathering sometimes assumed (Rye and Holland, 1998). This whole archaic system of Acknowledgments widespread acid sulfate weathering has been driven underground and into marginal environments. Data on baritic paleosols of the Proter- Jeremy Retallack helped during fieldwork. Also helpful were dis- ozoic and early are lacking (Table 3), so it remains a mystery cussions with Abigail Allwood, Bruce Runnegar, Debra Jennings, Steven when and how that marginalization occurred. Driese, Martin Van Kranendonk, Malcolm Walter, James Farquhar, and Roger Buick.

15 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Appendix A. Supplementary data 233–254. Brown, A.J., Cudahy, T.J., Walter, M.R., 2006. Hydrothermal alteration at the Panorama Formation, North Pole Dome, Pilbara Craton, Western Australia. Precambrian Res. Supplementary data to this article can be found online at https:// 151, 211–233. doi.org/10.1016/j.palaeo.2017.10.013. Buck, B.J., Lawton, T.F., Brock, A.L., 2010. Evaporitic paleosols in continental strata of the Carroza Formation, La Popa Basin, Mexico: record of Paleogene climate and salt tectonics. Geol. Soc. Am. Bull. 122, 1011–1026. References Buick, R., Barnes, K.R., 1984. Cherts in the Warrawoona Group: Early Archean silicified sediments deposited in shallow-water environments. In: Muhling, J., Groves, D.I., Blake, T.S. (Eds.), Archean and Proterozoic Basins of the Pilbara, Western Australia — Aachib, M., Mbonimpa, M., Aubertin, M., 2004. Measurement and prediction of the Evolution and Mineralization Potential. Univ. Western Australia Geol. Dept. Univ. oxygen diffusion coefficient in unsaturated media with applications to soil covers. Extens. Publ. 9, pp. 37–53. Water Air Soil Pollut. 156, 163–193. Buick, R., Dunlop, J.S.R., 1990. Evaporitic sediments of early Archean age from the Allwood, A.C., Walter, M.R., Marshall, C.P., 2006. Raman spectroscopy reveals thermal Warrawoona, North Pole, Western Australia. 37, 247–277. palaeoenvironments of c.3.5 billion-year-old organic matter. Vib. Spectrosc. 41, Buick, R., Thronetee, J.R., McHaughton, N.J., Smith, J.B., Barley, M.E., Savage, M., 1995. 190–197. Record of emergent continental crust ~3.5 billion years ago in the Pilbara craton of Allwood, A.C., Walter, M.R., Burch, I.W., Kamber, B.S., 2007. 3.42 billion-year-old Australia. Nature 375, 574–576. stromatolite reef from the Pilbara Craton of Western Australia: ecosystem-scale in- Bureau of Meteorology, 2016. Australian Government Website. http://www.bom.gov.au/ sights to early life on Earth. Precambrian Res. 158, 198–227. climate/data/ (accessed April 7, 2016). Allwood, A.C., Kamber, B.S., Walter, M.R., Burch, I.W., Kanik, I., 2010. Trace elements Burgman, A., 2007. Nyamal Dictionary, English-Nyamal Finderlist and Topical Wordlist. record depositional history of an Early Archean stromatolitic carbonate platform. Wangka Maya, Port Hedland (27 p). Chem. Geol. 270, 148–163. Carson, C.D., Fanning, D.S., Dixon, J.B., 1982. Alfisols and ultisols with acid sulfate Amundson, R., Dietrich, W., Bellugi, D., Ewing, S., Nishiizumi, K., Chong, G., Owen, J., weathering in Texas. In: Kittrick, J.A., Fanning, D.S., Hossner, L.R. (Eds.), Acid Finkel, R., Heimsath, A., Stewart, B., Caffee, M., 2012. Geomorphological evidence Sulfate Weathering. Soil Sci. Soc. Amer. Spec. Publ. 10pp. 127–146. for the late Pliocene onset of hyperaridity in the Atacama Desert. Geol. Soc. Am. Bull. Claire, M.W., Kasting, J.F., Domagal-Goldman, S.D., Stuëken, E., Buick, R., Meadows, 124, 1048–1070. V.S., 2014. Modeling the signature of sulfur mass-independent fractionation pro- Anders, E., Grevesse, N., 1989. Abundances of the elements: meteoritic and solar. duced in the Archean atmosphere. Geochim. Cosmochim. Acta 141, 365–380. Geochim. Cosmochim. Acta 53, 197–214. Clauer, N., O'Neil, J.R., Bonnnot-Courtois, C., Holtzapfel, T., 1990. Morphological, che- Awramik, S.M., Schopf, J.W., Walter, M.R., 1983. Filamentous fossil bacteria from the mical, and isotopic evidence for an early diagenetic evolution of detrital smectite in Archean of Western Australia. Precambrian Res. 20, 357–374. marine sediments. Clay Clay Miner. 38, 33–46. Awramik, S.M., Schopf, J.W., Walter, M.R., 1988. Carbonaceous filaments from North Crowe, S.A., Jones, C., Katsev, S., Magen, C., O'Neill, A.H., Sturm, A., Canfield, D.E., Pole, Western Australia: are they bacteria in Archean stromatolites? Precambrian Res. Haffner, G.D., Mucci, A., Sundby, B., Fowle, D.A., 2008. Photoferrotrophs thrive in an 39, 303–309. Archean Ocean analogue. U.S. Nat. Acad. Sci. Proc. 105, 15938–15943. Beattie, J.A., Haldane, A.D., 1958. The occurrence of palygorskite and barytes in certain Crowe, S.A., Døssing, L.N., Beukes, N.J., Bau, M., Kruger, S.J., Frei, R., Canfield, D.E., parna soils of the Murrumbidgee region, New South Wales. Austral. J. Sci. 20, 2013. Atmospheric oxygenation three billion years ago. Nature 501, 535–538. 274–275. Cullers, R.L., DiMarco, M.J., Lowe, D.R., Stone, J., 1993. Geochemistry of silicified, felsic, Benison, K.C., Bowen, B.B., 2013. Extreme sulfur-cycling in acid brine lake environments volcaniclastic sediments from the early Archean panorama formation, Pilbara block, of Western Australia. Chem. Geol. 351, 154–167. Western Australia: an evaluation of depositional and post-depositional processes with Benison, K.C., Bowen, B.B., 2015. The evolution of end-member continental waters: the special emphasis on rare earth elements. Precambrian Res. 60, 99–116. origin of acidity in southern Western Australia. GSA Today 25 (6), 1–10. Darmody, R.G., Harding, S.D., Hassett, J.J., 1989. Barite authigenesis in surficial soils of Benison, K.C., Bowen, B.B., Oboh-Ikuenobe, F.E., Jagniecki, E.A., LaClair, D.A., Story, mid-continental United States. In: Miles, D.L. (Ed.), Water-Rock Interaction. A.A. S.L., Mormile, M.R., Hong, B.-Y., 2007. Sedimentology of acid saline lakes in southern Balkema, Rotterdam, pp. 183–186. Western Australia: newly described processes and products of an extreme environ- Dash, B., Sahni, K.N., Bowes, D.A., 1987. Geochemistry and original nature of ment. J. Sediment. Res. 77, 366–388. Precambrian knondalites in the Eastern Ghats, Orissa, India. Trans. Roy. Soc. Benison, K.C., Jagniecki, E.A., Edwards, T.B., Mormile, M.R., Storrie-Lombardi, M.C., Edinburgh. Earth Sci. 78, 115–127. 2008. “Hairy blobs:” microbial suspects preserved in modern and ancient extremely Denny, M.W., 1993. Air and Water: The Biology and Physics of Life's Media. Princeton acid lake evaporites. Astrobiology 8, 1–15. Univ. Press, Princeton (343 p). Birnbaum, S.J., Wireman, J.W., 1985. Sulfate-reducing bacteria and silica solubility: a Dietrich, M., 1975. Wüstenrose. Verein. Freunde Mineral. Geol. 26. pp. 217–219. possible mechanism for evaporite diagenesis and silica precipitation in banded iron DiMarco, M.J., Lowe, D.R., 1989. Stratigraphy and sedimentology of an early Archean formations. Can. J. Earth Sci. 22, 1905–1909. felsic volcanic sequence, Eastern Pilbara block, Western Australia, with special re- Birnbaum, S.J., Wireman, J.W., Borowski, R., 1989. Silica precipitation induced by the ference to the duffer formation and implications for crustal evolution. Precambrian anaerobic sulfate reducing bacterium Desulfovibrio desulfuricans:effects upon cell Res. 44, 147–169. morphology and implications for preservation. In: Crick, R.E. (Ed.), Origin, Dodd, M.S., Papineau, D., Grenne, T., Slack, J.F., Rittner, M., Pirajno, F., O'Neil, J., Little, Evolution, and Modern Aspects of Biomineralization in Plants and Animals. Springer, C.T., 2017. Evidence for early life in Earth's oldest hydrothermal vent precipitates. New York, pp. 507–516. Nature 543, 60–64. Bizzarro, M., Simonetti, A., Stevenson, R.K., David, J., 2002. Hf isotope evidence for a Domagal-Goldman, S., 2014. The upside-down biosphere: evidence for the partially hidden mantle reservoir. 30, 771–774. oxygenated oceans during the Archean Eon. In: Shaw, G.H. (Ed.), Earth's Early Bolhar, R., Van Kranendonk, M.J., Kamber, B.S., 2005. A trace element study of siderite- Atmosphere and Surface Environment. Geol. Soc. Amer. Spec. Pap. 504. pp. 97–99. jasper banded iron formation in the 3.45 Ga Warrawoona Group, Pilbara Craton: Donovan, R.N., Younger, P., Ditzell, C., 1988. Some aspects of the geology of Zodletone Formation from hydrothermal fluids and shallow seawater. Precambrian Res. 137, Mountain, Southwestern Oklahoma. In: Hayward, O.T. (Ed.), Geological Society of 93–114. America Centennial Field Guide Southeast Section. Geological Society of America, Bonny, S.M., Jones, B., 2008a. Petrography and textural development of inorganic and Boulder, pp. 99–102. biogenic lithotypes in a relict barite tufa deposit at Flybye Springs, NT, Canada. Donovan, J.J., Lowers, H.A., Rusk, B.G., 2011. Improved electron probe microanalysis of Sedimentology 55, 275–303. trace elements in quartz. Am. Mineral. 96, 274–281. Bonny, S.M., Jones, B., 2008b. Experimental precipitation of barite (BaSO ) among 4 Duda, J.P., Van Kranendonk, M.J., Thiel, V., Ionescu, D., Strauss, H., Schäfer, N., Reitner, streamers of sulfur-oxidizing bacteria. J. Sediment. Res. 78, 357–365. J., 2016. A rare glimpse of paleoarchean life: of an exceptionally pre- Bontognali, T.R.R., Sessions, A.L., Allwood, A.C., Fischer, W.W., Grotzinger, J.P., served microbial mat facies from the 3.4 Ga Strelley Pool Formation, Western Summons, R.E., Eiler, J.M., 2012. Sulfur isotopes of organic matter preserved in 3.45- Australia. PLoS One 11, e0147629. billion-year-old stromatolites reveal microbial metabolism. U.S. Nat. Acad. Sci. Proc. Eriksson, K.A., Warren, J.K., 1983. A paleohydrologic model for Early Proterozoic dolo- 109, 15146–15151. mitization and silicification. Precambrian Res. 21, 299–321. Bown, T.M., Kraus, M.J., 1988. Geology and paleoenvironment of the Oligocene Jebel Eugster, H.P., 1969. Inorganic bedded cherts from the Magadi area, Kenya. Contrib. Qatrani Formation and adjacent rocks, Fayum depression, Egypt. US Geol. Surv. Prof. Mineral. Petrol. 22, 1–31. Pap. 1452, 1–60. Fedo, C.M., Nesbitt, H.W., Young, G.M., 1995. Unraveling the effects of potassium me- Brasier, M.D., Green, O.R., Jephcoat, A.P., Kleppe, A.K., van Kranendonk, M.J., Lindsay, tasomatism in sedimentary rocks and paleosols, with implications for paleo- J.V., Stewart, A., Grassineau, N.V., 2002. Questioning the evidence for Earth's oldest weathering conditions and provenance. Geology 23, 921–924. fossils. Nature 416, 76–81. Fitzpatrick, R., Degens, B., Baker, A., Raven, M., Shand, P., Smith, M., Rogers, S., George, Brasier, M.D., Antcliffe, J., Saunders, M., Wacey, D., 2015. Changing the picture of Earth's R., 2008. Avon Basin, WA Wheatbelt: Acid sulfate soils and salt efflorescences in open earliest fossils (3.5–1.9 Ga) with new approaches and new discoveries. U.S. Nat. drains and receiving environments. In: Fitzpatrick, R., Shand, P. (Eds.), Inland Acid Acad. Sci. Proc. 112, 4859–4864. Sulfate Soil Systems across Australia. CRC LEME Open File Report 249, pp. 189–204. Brimhall, G.H., Chadwick, O.A., Lewis, C.J., Compston, W., Williams, I.S., Danti, K.J., Food and Agriculture Organization, 1974. Soil Map of the World, 1:500,000. I. Legend. Dietrich, W.E., Power, M.E., Hendricks, D., Bratt, J., 1992. Deformational mass UNESCO, Paris (59 p). transport and invasive processes in soil evolution. Science 255, 695–702. Food and Agriculture Organization, 1978. Soil Map of the World, 1:500,000. X. Brock-Hon, A.L., Robins, C.R., Buck, B.J., 2012. Micromorphological investigation of Australasia. UNESCO, Paris (221 p). pedogenic barite in Mormon Mesa petrocalcic horizons, Nevada USA: implication for Gallagher, T.M., Sheldon, N.D., 2013. A new paleothermometer for forest paleosols and genesis. Geoderma 179, 1–8. its implications for Cenozoic climate. Geology 41, 647–651. de Brodtkorb, M.K., 1989. Barite nodules and concretions. In: de Brodtkorb, M.K. (Ed.), Garces, B.L.V., Aguilar, J.G., 1994. Permian saline lakes in the Aragón-Béarn Basin, Non-Metalliferous Stratabound Ore Fields. Van Nostrand Reinhold, New York, pp.

16 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Western Pyrenees. In: Renaut, R.W., Last, W.M. (Eds.), Sedimentology and support selective preservation in cell-like structures. Geochim. Cosmochim. Acta 112, Geochemistry of Modern and Ancient Saline Lakes. Soc.Sediment. Geol. Spec. Publ. 66–86. 50pp. 267–290. Li, W., Johnson, C.M., Beard, B.L., 2012. U–Th–Pb isotope data indicate Phanerozoic age Grandstaff, D.E., Edelman, M.J., Foster, R.W., Zbinden, E.M., Kimberley, M.M., 1986. for oxidation of the 3.4 Ga Apex Basalt. Earth Planet. Sci. Lett. 319-320, 197–206. Chemistry and mineralogy of Precambrian paleosols at the base of the Dominion and Lowe, D.R., 2007. A comment on weathering of quartz as an Archean climatic indicator Pongola Groups (Transvaal, South Africa). Precambrian Res. 32, 97–131. by N.H. Sleep and A.M. Hessler. Earth Planet. Sci. Lett. 253, 530–533. Halevy, I., Zuber, M.T., Schrag, D.P., 2007. A sulfur dioxide climate feedback on early Lowe, D.R., 1983. Restricted shallow-water sedimentation of early Archean stromatolitic Mars. Science 318, 1903–1907. and evaporitic strata of the Strelley Pool Chert, Pilbara Block, Western Australia. Haqq-Misra, J.D., Domagal-Goldman, S.D., Kasting, P.J., Kasting, J.F., 2008. A revised Precambrian Res. 19, 239–283. hazy methane greenhouse for the Archean Earth. Astrobiology 8, 1127–1137. Lowe, D.R., Tice, M.M., 2004. Geologic evidence for Archean atmospheric and climatic

Harvey, H.R., Tuttle, J.H., Bell, J.T., 1995. Kinetics of phytoplankton decay during si- evolution: fluctuating levels of CO2,CH4, and O2 with an overriding tectonic control. mulated sedimentation: changes in biochemical composition and microbial activity Geology 32, 493–496. under oxic and anoxic conditions. Geochim. Cosmochim. Acta 59, 3367–3377. Lowe, D.R., Worrell, G.F., 1999. Sedimentology, mineralogy, and implications of silicified Hay, R.L., 1968. Chert and its sodium-silicate precursors in sodium-carbonate lakes of evaporites in the Kromberg Formation, Barberton Greenstone Belt, South Africa. In: East Africa. Contrib. Mineral. Petrol. 17, 255–274. Lowe, D.R., Byerly, G.R. (Eds.), Geologic Evolution of the Barberton Greenstone Belt, Hessler, A.M., Lowe, D.R., Jones, R.L., Bird, D.K., 2004. A lower limit for atmospheric South Africa. Geol. Soc. Amer. Spec. Publ. 329pp. 167–188. carbon dioxide levels 3.2 billion years ago. Nature 428, 736–738. Maliva, R.G., Knoll, A.H., Simonson, B.M., 2005. Secular change in the Precambrian silica Hickman, A.H., 2008. Regional review of the 3424–3350 Ma Strelley Pool Formation, cycle: insights from chert . Geol. Soc. Am. Bull. 117, 835–845. Pilbara Craton, Western Australia. West. Austral. Geol. Surv. Rec. 2008/15, 1–61. Marty, B., Zimmermann, L., Pujol, M., Burgess, R., Philippot, P., 2013. Nitrogen isotope Hickman, A.H., 2013. Geology of the North Shaw 1:100 000 Sheet. Western Australia composition and density of the Archean atmosphere. Science 342, 101–104. Geological Survey, Perth (1p). Maslov, A.V., Grazhdankin, D.V., 2011. Petrologiya, geokhimiya, litokhimicheske Hoashi, M., Bevacqua, D.C., Otake, T., Watanabe, Y., Hickman, A.H., Utsunomiya, S., oso'ennosti porod verkhnego benda s prizhnakami vliyaniya proshessov pedogeneza Ohmoto, H., 2009. Primary hematite formation in an oxygenated sea 3.46 billion Okrestnosti r Ust'-Katav (petrology, geochemistry and lithochemical features of upper years ago. Nat. Geosci. 2, 301–306. Vendian rocks with signs of pedogenic processes) near the city of Ust-katau. Trudy Howari, F.M., Goodell, P.C., Miyamoto, S., 2002. Spectral properties of salt crusts formed Inst. Geol. Geophys. Ural Branch Russian Acad. Sci. 158, 113–118. on saline soils. J. Environ. Qual. 31, 1453–1461. McCarthy, P.J., Plint, A.G., 2013. A pedostratigraphic approach to nonmarine sequence Hren, M.T., Tice, M.M., Chamberlain, C.P., 2009. Oxygen and hydrogen isotope evidence stratigraphy: a three-dimensional paleosol-landscape model from the for a temperate climate 3.42 billion years ago. Nature 462, 205–208. (Cenomanian) Dunvegan Formation, Alberta and British Columbia, Canada. In: Hughes, I., Hase, I.P.A., 2010. Measurements and their Uncertainties: A Practical Guide to Driese, S.G. (Ed.), New Frontiers in and Terrestrial . Modern Error Analysis. Oxford Univ. Press, Oxford (136 p). Soc. Econ. Paleont. Mineral. Spec. Pap. 44pp. 159–177. Isbell, R.F., 1996. The Australian Soil Classification. C.S.I.R.O, Collingwood (152 p). McKenzie, N., Jacquier, D., Isbell, R., Brown, K., 2004. Australian Soils and Landscapes. Jennings, D.S., Driese, S.G., 2014. Understanding barite and gypsum precipitation in CSIRO, Melbourne (416 p). upland acid-sulfate soils: an example from a Lufkin series toposequence, South-cen- McLoughlin, S., 1996. Early Cretaceous floras of Western Australia. West. Austral. Mus. tral Texas, U.S.A. Sediment. Geol. 299, 106–118. Rec. 18, 19–65. Jennings, D.S., Driese, S.G., Dworkin, S.I., 2015. Comparison of modern and ancient Morey, G.B., Setterholm, D.R., 1997. Rare earth elements in weathering profiles and se- barite-bearing acid-sulfate soils using micromorphology, geochemistry and field re- diments of Minnesota: implications for provenance studies. J. Sediment. Res. 67, lationships. Sedimentology 62, 1078–1099. 105–115. Kasting, J., Kirschvink, J., 2012. Evolution of a habitable planet. In: Impey, C., Lunine, J., Mormile, M.R., Hong, B.-Y., Benison, K.C., 2009. Molecular analysis of the microbial Funes, J. (Eds.), Frontiers of Astrobiology. Cambridge Univ. Press, New York, pp. communities of Mars analog lakes in Western Australia. Astrobiology 9, 920–930. 115–131. Morris, R.C., Ramanaidou, E.R., 2007. Genesis of the channel iron deposits (CID) of the Kato, Y., Nakamura, K., 2003. Origin and global tectonic significance of Early Archean Pilbara region Western Australia. Austral. J. Earth Sci. 54, 733–756. cherts from the Marble Bar Greenstone Belt, Pilbara Craton, Western Australia. Mukhopadhyay, J., Crowley, Q.G., Ghosh, S., Ghosh, G., Charkrabarti, K., Misra, B., Precambrian Res. 125, 191–243. Heron, K., Bose, S., 2014. Oxygenation of the Archean atmosphere: new paleosol Kato, Y., Suzuki, K., Nakamura, K., Hickman, A.H., Nedachie, M., Kusakabe, M., constraints from eastern India. Geology 42, 923–926. Bevacqua, D.C., Ohmoto, H., 2009. Hematite formation by oxygenated groundwater Murphy, C.P., 1983. Counting pores and illuvial clay in thin section. Geoderma 31, more than 2.76 billion years ago. Earth Planet. Sci. Lett. 278, 40–44. 133–150. Keyser, A.W., 1968. Some indications of arid climate during the deposition of the Beaufort Nabhan, S., Luber, T., Scheffler, T., Heubeck, C., 2016. Climatic and geochemical im- Series. Geol. Surv. South Africa Annals 5, 77–78. plications of Archean pedogenic gypsum in the Moodies Group (∼3.2 Ga), Barberton Khalaf, F.I., El Sayed, M.I., 1989. Occurrence of barite-bearing balls within the Greenstone Belt, South Africa. Precambrian Res. 275, 119–134. Mio–Pliocene terrestrial sand (Kuwait group) in southern Kuwait, Arabian gulf. Neaman, A., Chorover, J., Brantley, S.L., 2005. Element mobility patterns record organic Sediment. Geol. 64, 197–202. ligands in soils on early Earth. Geology 33, 117–120. Kimberly, M.M., Grandstaff, D.E., 1986. Profiles of elemental concentrations in Nelson, D.R., 2002. Tuffaceous rhyolite, shark well (sample 160221). Western Australia Precambrian paleosols in basaltic and granitic parent materials. Precambrian Res. 32, Geological Survey Geochronology Dataset 238, 106–108. 133–154. Nesbitt, H.W., 1979. Mobility and fractionation of rare earth elements during weathering Knauth, L.P., Lowe, D.R., 2003. High Archean climatic temperature inferred from oxygen of a granodiorite. Nature 279, 206–210. isotope geochemistry of cherts in the 3.5 Ga Swaziland Supergroup, South Africa. Nesbitt, H.W., Young, G.M., 1989. Formation and diagenesis of weathering profiles. J. Geol. Soc. Am. Bull. 115, 566–580. Geol. 97, 129–147. Kohut, C.K., Durdas, M.J., 1993. Evaporite mineralogy and trace-element content of salt- Neuhaus, S., 2010. New age for Paleogene/ clastics at the northern termination affected soils in Alberta. Can. J. Soil Sci. 73, 399–409. of the Upper Rhine Graben (Hesse, Germany). Zeit. Deutsche Gesell. Geowiss. 161, Kolesnikov, A.V., Grazhdankin, D.V., Maslov, A.V., 2012. Arumberia-type structures in the 303–322. upper Vendian of the Urals. Dokl. Earth Sci. 447, 1233–1239. Novoselov, A.A., de Souza Filho, C.R., 2015. Potassium metasomatism of Precambrian Kolesnikov, A.V., Marusina, V.V., Nagovitsin, K.E., Maslov, A.V., Grazhdankin, D.V., paleosols. Precambrian Res. 262, 67–83. 2015. in the aftermath of the Kotlinian crisis: Asha group of the South Nutman, A.P., Bennett, V.C., Friend, C.R., Van Kranendonk, M.J., Chivas, A.R., 2016. Urals. Precambrian Res. 263, 59–78. Rapid emergence of life shown by discovery of 3,700-million-year-old microbial Konhauser, K.O., Hamade, T., Raiswell, R., Morris, R.C., Ferris, F.G., Southam, G., structures. Nature 537, 535–538. Canfield, D.E., 2002. Could bacteria have formed the Precambrian banded iron for- Ohmoto, H., 1996. Evidence in pre-2.2 Ga paleosols for the early evolution of atmospheric mations? Geology 30, 1079–1082. oxygen and terrestrial biota. Geology 24, 1135–1138. Krapež, B., 1984. Sedimentation in a small, fault-bounded basin — Lalla Rookh Ohmoto, H., Watanabe, Y., Lasaga, A.C., Naraoka, H., Johnson, I., Brainard, J., Chorney, Sandstone, East Pilbara Block. In: Muhling, J., Groves, D.I., Blake, T.S. (Eds.), A., 2014. Oxygen, iron and sulfur geochemical cycles on the early Earth: problems Archean and Proterozoic Basins of the Pilbara, Western Australia — Evolution and and contradictions. In: Shaw, G.H. (Ed.), Earth's Early Atmosphere and Surface Mineralization Potential. Univ. Western Austral. Geol. Dept. Univ. Extens. Publ. 9pp. Environment. Geol. Soc. Amer. Spec. Pap. 504pp. 55–95. 89–110. Olson, E.C., 1967. Early Permian of Oklahoma. Oklahoma Geol. Surv. Circ. Kreuser, T., 1995. Tectonic and climatic controls of lacustrine sedimentation in prerift 74, 1–111. and rift settings in the Permian–Triassic of East Africa. J. Paleolimnol. 13, 3–19. Óskarsson, B.V., Riishuus, M.S., Arnalds, O., 2012. Climate-dependent chemical weath- Lambert, I.B., Donnelly, T.H., Dunlop, J.S.R., Groves, D.I., 1978. Stable isotopic compo- ering of volcanic soils in Iceland. Geoderma 189–190, 635–651. sitions of early Archean sulfate deposits of probably evaporitic and volcanogenic Pinti, D.L., Mineau, R., Clement, V., 2009. Hydrothermal alteration and microfossil ar- origins. Nature 276, 808–811. tefacts of the 3,465-million-year-old apex chert. Nat. Geosci. 2, 640–643. Larsen, D., 2008. Revisiting silicate authigenesis in the Pliocene-Pleistocene Lake Tecopa Pufahl, P.K., Pirajno, F., Hiatt, E.E., 2013. Riverine mixing and fluvial iron formation: a beds, southeastern California: depositional and hydrological controls. Geosphere 4, new type of Precambrian biochemical sediment. Geology 41, 1235–1238. 612–639. Pujol, M., Marty, B., Bumard, P., Philippot, P., 2009. Xenon in Archean barite: weak decay Larsen, L.M., Rex, D.C., 1992. A review of the 2500 Ma span of alkaline-ultramafic, po- of 130Ba, mass dependent isotopic fractionation and implication for barite formation. tassic and carbonatitic magmatism in West Greenland. Lithos 28, 367–402. Geochim. Cosmochim. Acta 73, 6834–6846. Laznicka, P., 1976. Barite nodules of possibly late diagenetic origin from Twitya River Rasmussen, B., Buick, R., 1999. Redox state of the Archean atmosphere: evidence from area, Mackenzie Mountains, Northwest Territories. Can. J. Earth Sci. 13, 1446–1455. detrital heavy minerals in ca. 3250–2750 Ma sandstones from the Pilbara Craton, Lepot, K., Williford, K.H., Ushikubo, T., Sugitani, K., Mimura, K., Spicuzza, M.J., Valley, Australia. Geology 27, 115–118. J.W., 2013. Texture-specific isotopic compositions in 3.4 Gyr old organic matter Renaut, R.W., Tiercelin, J.-J., 1994. Lake Bogoria, Kenya Rift Valley — a sedimentological

17 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

overview. In: Renaut, R.W., Last, W.M. (Eds.), Sedimentology and Geochemistry of Stoops, G.J., Zavaleta, A., 1978. Micromorphological evidence of barite neoformation in Modern and Ancient Saline Lakes. Soc. Sedim. Geol. Spec. Publ. 50pp. 101–124. soils. Geoderma 20, 63–70. Retallack, G.J., 1997a. A Colour Guide to Paleosols. Wiley, Chichester (346 p). Sugahara, H., Sugitani, K., Mimura, K., Yamashita, F., Yamamoto, K., 2010. A systematic Retallack, G.J., 1997b. Palaeosols in the upper Narrabeen Group of New South Wales as rare-earth elements and yttrium study of Archean cherts at the Mount Goldsworthy evidence of Early Triassic palaeoenvironments without exact modern analogues. greenstone belt in the Pilbara Craton: implications for the origin of microfossil- Austral. J. Earth Sci. 44, 185–201. bearing black cherts. Precambrian Res. 177, 73–87. Retallack, G.J., 2005. Pedogenic carbonate proxies for amount and seasonality of pre- Suganuma, Y., Hamano, Y., Niitsuma, S., Hoashi, M., Hisamitsu, T., Niitsuma, N., cipitation in paleosols. Geology 33, 333–336. Kodama, K., Nedachi, M., 2006. Paleomagnetism of the marble bar Chert member, Retallack, G.J., 2009. Greenhouse crises of the past 300 million years. Geol. Soc. Am. Bull. Western Australia: implications for apparent polar wander path for Pilbara Craton 121, 1441–1455. during Archaean time. Earth Planet. Sci. Lett. 252, 360–371. Retallack, G.J., 2011. Neoproterozoic glacial loess and limits to snowball earth. Geol. Soc. Sugitani, K., Mimura, K., Suzuki, K., Nagamine, K., Sugisaki, R., 2003. Stratigraphy and London J. 168, 1–19. sedimentary petrology of an Archean volcanic/sedimentary succession at Mt. Retallack, G.J., 2012. Were Ediacaran siliciclastics of South Australia coastal or deep Goldsworthy in the Pilbara block, Western Australia: implications of evaporite marine? Sedimentology 59, 1208–1236. (nahcolite) and barite deposition. Precambrian Res. 120, 55–79. Retallack, G.J., 2013a. Ediacaran life on land. Nature 493, 89–92. Sugitani, K., Lepot, K., Nagaoka, T., Mimura, K., van Kranendonk, M.J., Oehler, D.Z., Retallack, G.J., 2013b. Permian and Triassic greenhouse crises. Gondwana Res. 24, Walter, M.R., 2010. Biogenicity of morphologically diverse carbonaceous micro- 90–103. structures from the ca 3400 Ma Strelley Pool Formation in the Pilbara Craton, Retallack, G.J., 2014a. Precambrian life on land. Palaeobotanist 63, 1–15. Western Australia. Astrobiology 10, 899–920. Retallack, G.J., 2014b. Volcanosedimentary paleoenvironments of Ediacaran fossils in Sugitani, K., Mimura, K., Nagaoka, T., Lepot, K., Takeuchi, M., 2013. Microfossil as- Newfound land. Geol. Soc. Am. Bull. 126, 619–638. semblage from the 3400 Ma Strelley pool formation in the Pilbara Craton, Western Retallack, G.J., 2014c. Paleosols and paleoenvironments of early Mars. Geology 42, Australia: results from a new locality. Precambrian Res. 226, 59–74. 755–758. Tarhan, L.G., Droser, M.L., Gehling, J.G., 2015. Depositional and preservational en- Retallack, G.J., 2016. Field and laboratory tests for recognition of Ediacaran paleosols. vironments of the Ediacara Member, Rawnsley Quartzite (South Australia): assess- Gondwana Res. 36, 94–110. ment of paleoenvironmental proxies and the timing of ‘ferruginization. Precambrian Retallack, G.J., Huang, C.-M., 2010. Depth to gypsic horizon as a proxy for paleopreci- Res. 257, 124–136. pitation in paleosols of sedimentary environments. Geology 38, 403–406. Tarr, W.A., 1933. The origin of the sand barites of the lower Permian of Oklahoma. Am. Retallack, G.J., Kirby, M.X., 2007. Middle Miocene global change and paleogeography of Mineral. 18, 260–272. Panama. Palaios 22, 667–679. Taylor, G., Eggleton, R.A., 2001. Regolith Geology and . Wiley, Retallack, G.K., Sheldon, N.D., Carr, P.F., Fanning, M., Thompson, C.A., Williams, M.L., Chichester (375 p). Jones, B.G., Hutton, A., 2011. Multiple Early Triassic greenhouse crises impeded Teitler, A., Philippot, P., Gérard, M., Le Hira, G., Fluteau, F., Ader, M., 2015. Ubiquitous recovery from Late Permian mass extinction. Palaeogeogr. Palaeoclimatol. occurrence of basaltic-derived paleosols in the Late Archean Fortescue group, Palaeoecol. 308, 233–251. Western Australia. Precambrian Res. 267, 1–27. Retallack, G.J., Krinsley, D.H., Fischer, R., Razink, J.J., Langworthy, K., 2016. Archean Thorne, A.M., Trendall, A.F., 2001. Geology of the Fortescue Group, Pilbara Basin, coastal-plain paleosols and life on land. Gondwana Res. 40, 1–20. Western Australia. Geol. Surv. West. Austral. Bull. 144, 1–249. Rosing, M.T., Bird, D.K., Sleep, N.H., Bjerrum, C.J., 2010. No climate paradox under the Thorpe, R.I., Hickman, A.H., Davis, D.W., Mortensen, J.K., Trendall, A.F., 1992. faint early sun. Nature 464, 744–747. Constraints to models for Archean lead evolution from precise zircon U-Pb geo- Rye, R., Holland, H.D., 1998. Paleosols and the evolution of atmospheric oxygen: a cri- chronology for the Marble Bar region, Pilbara Craton, Western Australia. In: Glover, tical review. Amer. J. Sci. 298, 621–672. J.E., Ho, S.E. (Eds.), The Archean: Terrains, Processes and Metallogeny. Univ. West. Schieber, J., Bish, D., Coleman, M., Reed, M., Hausrath, E.M., Cosgrove, J., Gupta, S., Austral. Geol. Extens. Publ. 22, pp. 395–407. Minitti, M.E., Edgett, K.S., Malin, M., 2017. Encounters with an unearthly mudstone: Ueno, Y., Ono, S., Rumble, D., Maruyama, S., 2008. Quadruple sulfur isotope analysis of understanding the first mudstone found on Mars. Sedimentology 64, 311–358. ca. 3.5 Ga dresser formation: new evidence for microbial sulfate reduction in the Schmidt, P.W., Embleton, B.J.J., 1985. Prefolding and overprint magnetic signatures in early Archean. Geochim. Cosmochim. Acta 72, 5675–5691. Precambrian (2.9–2.7 Ga) igneous rocks from the Pilbara Craton and Hamersley Ueno, Y., Johnson, M.S., Danielache, S.O., Eskeberg, C., Pandey, A., Yoshida, N., 2009. Basin, NW Australia. J. Geophys. Res. 90 (B4), 2967–2984. Geological sulfur isotopes indicate elevated OCS in the Archean atmosphere, solving Schopf, J.W., Kudryavtsev, A.B., Czaja, A.D., Tripathi, A.B., 2007. Evidence of Archean faint young sun paradox. U.S. Nat. Acad. Sci. Proc. 106, 14784–14789. life: stromatolites and microfossils. Precambrian Res. 158, 141–155. van den Boorn, S.H.-J.M., van Bergen, M.J., Vroon, P.Z., de Vries, S.T., Nijman, W., 2010. Schrank, E., Mahmoud, M.S., 2000. New taxa of angiosperm pollen and associated pa- Silicon isotope and trace element constraints on the origin of ~3.5 Ga cherts: im- lynomorphs from the early late Cretaceous of Egypt (Mahgrabi Formation, Kharga plications for early Archean marine environments. Geochim. Cosmochim. Acta 74, Oasis). Rev. Palaeobot. Palynol. 112, 167–188. 1077–1103. Senko, J.M., Campbell, B.S., Henriksen, J.R., Elshahed, M.S., Dewers, T.A., Krumholz, van Kranendonk, M.J., 2000. Geology of the North Shaw 1:100,000 sheet. Explanatory L.R., 2004. Barite deposition resulting from phototrophic sulphide-oxidizing bacterial Notes Geological Survey of Western Australia 1–86. activity. Geochim. Cosmochim. Acta 68, 773–780. van Kranendonk, M.J., 2006. Volcanic degassing, hydrothermal circulation and the Serdyuchenko, D.P., 1968. Metamorphosed weathering rusts of the Precambrian: their flourishing of early life on Earth: a review of the evidence from c. 3490–3420 Ma metallogenis and petrogaphic fabric. In: Hejtman, B. (Ed.), Precambrian Geology. rocks of the Pilbara Supergroup, Pilbara Craton, Western Australia. Earth Sci. Rev. Proc. 13th Int. Geol. Congr. Prague 4pp. 37–42. 74, 197–240. Setti, M., Marinoni, L., López-Galindo, L., 2004. Mineralogical and geochemical char- van Kranendonk, M.J., 2010. Geology of the 1:100,000 Coongan Sheet. Geol. Surv. West. acteristics (major, minor, trace elements and REE) of detrital and authigenic clay Australia Explan. Notespp. 1–67. minerals in a Cenozoic sequence from Ross Sea, Antarctica. Clay Miner. 39, 405–421. van Kranendonk, M.J., Pirajno, F., 2004. Geochemistry of metabasalts and hydrothermal Shahid, S.A., Jenkins, D.A., 1994. Mineralogy and micromorphology of salt crusts from alteration zones associated with c. 3450 Ga chert and barite deposits: implications for the Punjab, Pakistan. In: Ringrose-Voase, A.J., Humphreys, G.S. (Eds.), Soil the geological setting of the Warrawoona group, Pilbara Craton, Australia. Geochem. Micromorphology: Studies in Management and Genesis. Elsevier, Amsterdam, pp. Explor. Environmental Anal. 4, 253–278. 799–800. van Kranendonk, M.J., Hickman, A.H., Smithies, R.H., Williams, I.R., Bagas, L., Farrell,

Sheldon, N.D., 2006. Precambrian paleosols and atmospheric CO2 levels. Precambrian T.R., 2006. Revised lithostratigraphy of Archaean supracrustal and intrusive rocks in Res. 147, 148–155. the northern Pilbara Craton, Western Australia. Geol. Surv. West. Austral. Rec. 2006 Sheldon, N.D., Retallack, G.J., 2001. Equation for compaction of paleosols due to burial. (/15), 1–55. Geology 29, 247–250. van Kranendonk, M.J., Philippot, P., Lepot, K., Bodorkos, S., Pirajno, F., 2008. Geological Sheldon, N.D., Retallack, G.J., Tanaka, S., 2002. Geochemical climofunctions from North setting of the Earth's oldest fossils in the ca. 3.5 Ga dresser formation, Pilbara Craton, American soils and application to paleosols across the Eocene-Oligocene boundary in Western Australia. Precambrian Res. 167, 93–124. Oregon. J. Geol. 110, 687–696. Vepraskas, M.J., Sprecher, S.W., 1997. Summary. In: Vepraskas, M.J., Sprecher, S.W. Shen, Y., Buick, R., Canfield, D.E., 2001. Isotopic evidence for microbial sulphate re- (Eds.), Aquic Conditions and Hydric Soils: The Problem Soils. Soil Sci. Soc. Amer. duction in the early Archaean era. Nature 410, 77–81. Spec. Publ. 150pp. 153–156. Shen, Y., Farquhar, J., Masterson, A., Kaufman, A.J., Buick, R., 2009. Evaluating the role Wacey, D., McLoughlin, N., Whitehouse, M.J., Kilburn, M.R., 2010. Two coexisting sulfur of microbial sulfate reduction in the early Archean using quadruple isotope sys- metabolisms in a ca. 3400 Ma sandstone. Geology 38, 1115–1118. tematics. Earth Planet. Sci. Lett. 279, 383–391. Wacey, D., Kilburn, M.R., Saunders, M., Cliff, J., Brasier, M.D., 2011. Microfossils of Siever, R., 1962. Silica solubility, 0°–200 °C, and the diagenesis of siliceous sediments. J. sulphur-metabolizing cells in 3.4-billion-year-old rock of Western Australia. Nat. Geol. 70, 127–150. Geosci. 4, 698–702. Skinner, M.F., Taylor, B.E., 1967. A revision of the geology and of the Bijou Walker, J.C.G., 1987. Was the Archaean biosphere upside down? Nature 329, 710–712. Hills, South Dakota. Am. Mus. Novit. 300, 1–53. Warren, J.K., 2006. Evaporites; Sediments, Resources and Hydrocarbons. Springer, Berlin Sleep, N.H., Hessler, A.M., 2006. Weathering of quartz as an Archean climatic indicator. (1035 p). Earth Planet. Sci. Lett. 241, 594–602. Westall, F., de Vries, S.T., Nijman, J.N., Marchesini, D., Severine, A., 2006. The 3.446 Ga Smith, G.I., 1979. Subsurface stratigraphy and geochemistry of late evapor- “Kittys Gap Chert”, an Early Archean Microbial Ecosystem. In: Reimold, W.U., ites, Searles Lake, California. U.S. Geol. Surv. Prof. Pap. 1043, 1–130. Gibson, R.L. (Eds.), Processes on the Early. Earth. Geol. Soc. Amer. Spec. Pap. 405pp. Soil Survey Staff, 2014. Keys to Soil Taxonomy. Natural Resources Conservation Service, 105–131. Washington DC (600 p). Woolley, A.R., Kempe, D.R.C., 1989. Carbonatites: nomenclature, average chemical Speczik, S., El Sayed, A.A.Y., 1991. Fluid inclusion studies of Bahariya barites (Western compositions and element distribution. In: Bell, K. (Ed.), Carbonatites. Unwin Desert, Egypt). Acta Geol. Pol. 41, 109–116. Hyman, London, pp. 1–14.

18 G.J. Retallack Palaeogeography, Palaeoclimatology, Palaeoecology xxx (xxxx) xxx–xxx

Wopfner, H., 1978. Silcretes of Northern South Australia and adjacent regions. In: within a complex ecosystem (Zodletone Spring, OK, USA): the role of the rare bio- Langford-Smith, T. (Ed.), Silcrete in Australia. Dept. Geogr. Univ., New England, sphere. Plos One 5, e12414-1241. Armidale, pp. 93–141. Ziegenbalg, S.B., Brunner, B., Rouchy, J.M., Birgel, D., Pierred, C., Böttcher, M.E., Caruso, Wordsworth, R., Pierrehumbert, R., 2013. Hydrogen-nitrogen greenhouse warming in A., Immenhauser, A., Peckmann, J., 2010. Formation of secondary carbonates and Earth's early atmosphere. Science 339, 64–67. native sulfur in sulfate-rich Messinian strata, Sicily. Sediment. Geol. 227, 37–50. Youssef, N.H., Couger, M.B., Elshahed, M.S., 2010. Fine-scale bacterial beta diversity

19