<<

Anatomy of a Sub-Cambrian in Wisconsin: Mass Fluxes of Chemical and Climatic Conditions in North America during Formation of the Cambrian Great Unconformity

L. Gordon Medaris Jr.,1,* Steven G. Driese,2 Gary E. Stinchcomb,3 John H. Fournelle,1 Seungyeol Lee,1,4 Huifang Xu,1,4 Lyndsay DiPietro,2 Phillip Gopon,5 and Esther K. Stewart6

1. Department of Geoscience, University of Wisconsin, Madison, Wisconsin 53706, USA; 2. Department of Geosciences, Terrestrial Research Group, Baylor University, Waco, Texas 76798, USA; 3. Department of Geosciences and Watershed Studies Institute, Murray State University, Murray, Kentucky 42071, USA; 4. NASA Astrobiology Institute, University of Wisconsin, Madison, Wisconsin 53706, USA; 5. Department of Earth Sciences, University of Oxford, South Parks Road, Oxford OX1 3AN, United Kingdom; 6. Wisconsin Geological and Natural History Survey, Madison, Wisconsin 53705, USA

ABSTRACT A paleosol beneath the Upper Cambrian Mount Simon Sandstone in Wisconsin provides an opportunity to evaluate the characteristics of Cambrian weathering in a subtropical climate, having been located at 207S paleolatitude 500 My ago. The 285-cm-thick paleosol resulted from advanced chemical weathering of a gabbroic protolith, recording a total mass loss of 50%. Weathering of hornblende and plagioclase produced a pedogenic assemblage of quartz, chlorite, kaolinite, goethite, and, in the lowest part of the profile, siderite. Despite the paucity of quartz in the protolith and 40% removal of

SiO2 from the profile, quartz constitutes 11%–23% of the pedogenic mineral assemblage. Like many other and Cambrian in the Lake Superior region, the paleosol experienced potassium metasomatism, now con- taining 10%–25% mixed-layer illite-vermiculite and 5%–44% potassium feldspar. Estimates of mean annual precipi- tation and mean annual temperature are 1777 mm y21 and 20.17C, respectively, which are consistent with a paleo- latitude of 207S. For an atmospheric CO2 concentration of 4000–6000 ppm at 550–500 Ma, the duration of weathering is constrained to have been between 20,000 and 100,000 y. When the effects of erosion and influence of protolith com- position are considered, the degree, or maturity, of weathering for the Wisconsin paleosol and four other sub-Cambrian paleosols is comparable to that for two modern in subtropical and temperate climates, despite the lack of land plants in Cambrian time. Such correspondent degrees of weathering likely result from the effects of elevated levels of atmospheric CO2 and microbial activity on weathering in Cambrian time.

Online enhancements: appendix.

Introduction The first Phanerozoic continental-scale marine trans- transgressive shoreface system (Peters and Gaines gression is marked by the Cambrian Great Uncon- 2012). formity, the formation of which was preceded by ex- Sub-Cambrian paleosols provide significant insights tensive chemical weathering of continental crust, into the characteristics of chemical weathering and followed by widespread physical reworking of , the contribution of labile elements to the hydro- regolith, and basement during the advance of a sphere and, ultimately, to seawater at 550–500 Ma, but many such paleosols were stripped away by the erosive effect of the transgressive Cambrian ocean. Manuscript received April 19, 2017; accepted January 13, 2018; electronically published March 20, 2018. During exploration by Flambeau Mining in 1996 * Author for correspondence; email: medaris@.wisc for volcanogenic massive sulfide deposits in Trempea- .edu. leau County, Wisconsin, a weathered profile beneath

[The Journal of Geology, 2018, volume 126, p. 261–283] q 2018 by The University of Chicago. All rights reserved. 0022-1376/2018/12603-0001$15.00. DOI: 10.1086/697037

261

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 262 L . G . M E D A R I S E T A L .

Locality and Overview of Drill Core TR119 Core TR119 was taken from a drill hole located in Trempealeau County, Wisconsin (44.44387N, 91.31627W), which in Cambrian time at 500 Ma was located ∼207 south of the equator, as were three other North American sub-Cambrian paleosols (fig. 1). All would have been situated in a wind belt of tropical easterlies (trade winds). Drilling through the Upper Cambrian siliciclastic Mount Simon Formation reached a nonconformity with gabbro basement at a depth of 93 m, passed through the paleosol for 2.85 m, and continued an- other 10.4 m in unweathered gabbro. The basal Mount Simon is bimodal in grain size and compo- sition, consisting of subangular to subrounded quartz pebbles and granules in an arkosic sandy matrix. The paleosol progresses downward through Bt (∼52 cm), Bts (∼95 cm), and Bk/Cr (∼112 cm) horizons, below which incipient weathering continues for another ∼26 cm, where amphibole is partly replaced by sid- erite but plagioclase remains unweathered. The degree (maturity) of weathering and possible effects of K-metasomatism in paleosols may be Figure 1. Locality map and paleolatitudes at 500 Ma for evaluated by application of the familiar A‐C N‐K p sub-Cambrian paleosols in North America. BH Butler plot (fig. 2; Nesbitt and Young 1984), in which A, C , p Hill outcrop, Missouri; SC Squaw Creek outcrop, Texas; N, and K are whole-rock molar quantities of Al O , SQ p drill core SQ8, Minnesota; TR p drill core TR119, 2 3 CaO (in silicate), Na O, and K O, respectively. The Wisconsin. Map used with permission from Key Time 2 2 Slices of North America q 2013 Colorado Plateau Geo- TR119 gabbro protolith, being devoid of K-feldspar, ‐ systems. plots along the A C N join, as does the incipiently weathered sample at a depth of 271 cm, in which

the Mount Simon Formation (Upper Cambrian) was recovered in drill core TR119, thus providing an exceptional opportunity to investigate the proper- ties of a sub-Cambrian paleosol. We have examined this paleosol in detail, establishing its depth varia- tion in pedogenic features, mineralogy, and whole- rock chemical compositions; calculating mass fluxes of weathering and K-metasomatism; determining past climate using several paleoclimate functions; and evaluating the degree of weathering. We have applied a similar analysis to three other sub-Cambrian paleo- sols in North America (Driese et al. 2007) and one in Israel (Sandler et al. 2012) to provide a continental- scale comparison of sub-Cambrian weathering during this important episode of Earth history. Finally, mass fluxes of weathering have been calculated for modern subtropical weathering of a granite (Liu et al. 2016) and for modern temperate weathering of dia- Figure 2. A‐C N‐K plot for gabbro protolith and com- base and granite (Bazilevskaya et al. 2012) to provide a bined saprolite and regolith in drill core TR119. Italicized comparison of weathering characteristics before and numbers are depths in centimeters below the Upper Cam- after the advent of land plants. brian nonconformity. reg p regolith; sap p saprolite.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 263 plagioclase remains unweathered. In the remaining weathered profile, plagioclase was completely re- moved by hydrolysis, and the weathering trend ef- fectively reached the A-apex, with weathered sam- ples consisting of kaolinite 1 quartz 1 chlorite 1 goethite 5 siderite. However, the bulk compositions of all weathered samples, except that at a depth of 271 cm, are dis- placed significantly from the predicted weathering trend toward the K-apex (fig. 2), signifying appre- ciable postweathering addition of potassium, which is a common phenomenon in sub-Cambrian paleo- sols and Cambro-Ordovician siliciclastic in the midcontinent region (Bethke and Marshak 1990; Driese et al. 2007). X-ray diffraction (XRD) analysis (see below) reveals that mixed-layer illite and vermiculite is abundant in samples at depths of 3 and 9 cm and that illite-vermiculite and K-feldspar are abundant in deeper samples in the weathered pro-

file, accounting for the relative enrichment in K2O.

Gabbro Protolith Texture and Mineralogy. The TR119 protolith is a medium-grained subhedral granular gabbro, con- sisting largely of subequal amounts of plagioclase and amphibole with a subordinate amount of diop- side that is partly replaced by amphibole (fig. 3A). Sample 315.7 is slightly silica saturated, containing a small amount of interstitial quartz (fig. 3B), and sample 314.8 is slightly undersaturated, being de- void of quartz. Although igneous texture is gener- ally preserved, the gabbro has locally been recrys- tallized under low-grade metamorphic conditions; some grains of igneous amphibole have been replaced by a fine-grained assemblage of actinolite and chlo- rite (fig. 3A), and plagioclase has been partly replaced by saussurite (a fine-grained mixture of clinozoisite and albite; fig. 3A,3C). Such low-grade metamor- phism is widespread in Paleoproterozoic rhyolite, gran- ite, and diorite in southern Wisconsin, where these rocks contain lower greenschist facies mineral as- semblages, but igneous textures are preserved (Me- daris et al. 2003). Selected protolith minerals were analyzed by elec- tron microprobe techniques (table A1; analytical meth- – odsaregivenintheappendix[appendixandtablesA1 Figure 3. Images of gabbro protolith. A, Photomicrograph A5 are available online]). Igneous plagioclase is (plane polarized light). a1c p metamorphic actinolite and slightly zoned labradorite, ranging in composition chlorite; d p igneous diopside; m p igneous magnesio- from 66 mol% An in cores to 58% An in rims of hornblende; p p igneous plagioclase; s p metamorphic grains, with a mean value of 60 mol% An. In contrast, saussurite. B, Photomicrograph (crossed polarizers). q p recrystallized plagioclase in saussurite is near end- quartz. C, Backscattered electron image of saussurite. e p fl member albite. Igneous clinopyroxene is diopside, epidote (shades of gray re ect different Fe2O3 contents); with a Ca∶Mg∶Fe ratio of 44.9∶41.1∶13.9 and a mag- p p plagioclase (shades of gray reflect different proportions nesium number (100 # Mg=½Mg 1 FeTotal) of 74.7. of albite).

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 264 L . G . M E D A R I S E T A L .

Igneous amphibole ranges in composition from mag- nesiohornblende to magnesiohastingsite and is com- monly zoned with respect to TiO2,havingTiO2-rich cores (∼2.5 wt%) and TiO2-poor rims (!0.5 wt%). Recrystallized amphibole is actinolite, which con- tains ∼2.2 wt% Al2O3 and !0.1 wt% TiO2 and has a magnesium number of 72.4. Metamorphic chlorite associated with actinolite is the variety ripidolite (Si p 5:4 apfu; FeTotal p 3:6 apfu), with a magne- sium number of 61.7. Epidote (clinozoisite) in saussurite originated by recrystallization of labra- dorite to albite and epidote and ranges in Fe2O3 content from ∼0.74 to ∼0.27 wt%, corresponding to Fe31 numbers (100 # Fe31=½Fe31 1 Al) of 24.5 and 8.9. Bulk-Rock Geochemistry. Whole-rock chemical analyses of two protolith samples were provided by the GeoAnalytical Laboratory at Washington State University, where major and selected trace element concentrations were measured by X-ray fluores- cence spectrometry (table A2; analytical details are given in the appendix). The TR119 protolith is a Figure 4. Modal proportions of minerals in the TR119 subalkaline, metaluminous gabbro that contains paleosol above the zone of incipient weathering (deter- p p ∼50 wt% SiO2 and ∼15 wt% Al2O3. The two ana- mined by X-ray diffraction). c chlorite; f potassium lyzed samples are closely comparable in major ele- feldspar; g p goethite; i p illite-vermiculite; k p kaolinite; p p p ment concentrations, displaying !10% relative proto protolith; q quartz; s siderite. standard deviation among the major elements, but are highly variable with respect to minor and trace elements (table A2). Both samples are close to silica which are characterized by domains of particles saturation; sample 315.7 is slightly saturated (2.95 wt% aligned in parallel extinction patterns in response normative quartz) and contains a small amount of to periodic wetting and drying (Bullock et al. 1985). interstitial quartz (fig. 3B), whereas sample 314.8 is Such fabric is now largely defined by postweathering slightly undersaturated (2.5 wt% normative olivine) illite-vermiculite but is thought to be inherited from and is devoid of quartz. Samples 315.7 and 314.8 are the weathering stage. An additional micromorpho- both relatively magnesian and undifferentiated, hav- logical feature is a very fine-grained association of ing magnesium numbers of 76.6 and 77.7 and differ- goethite, chlorite, quartz, and postweathering illite- entiation indices (normative orthoclase 1 albite 1 vermiculite, termed “sesquioxides,” in some cases quartz) of 21.8 and 19.0, respectively. occurring as a pseudomorph after original amphibole. Cryptic features include vermiform tubular shapes that are 10–15 mm in diameter and 40–50 mmlongas Paleosol Mineralogy and Micromorphology well as goethite preserved both as spherules with fine The paleosol consists of pedogenic kaolinite, quartz, fibrous bristles and filaments and as larger colloform chlorite, goethite, and siderite and postweathering, aggregates. Pedogenic siderite occurs toward the base metasomatic illite-vermiculite and K-feldspar. Pro- the weathered profile, below a depth of ∼185 cm, as portions of these minerals were determined by XRD, both spherulitic and nonspherulitic crystal masses. and all phases except siderite occur throughout the The “Cr horizon” refers to saprolite formed from weathered profile (table A3; fig. 4; see the appendix decomposed protolith that retained primary igneous for details on the XRD method). textures, despite extensive mineral alteration. Pedogenic horizons for the TR119 paleosol pro- Two different processes in the weathering of pla- file have been identified on the basis of micro- gioclase and amphibole were responsible for gener- morphological features observed in thin sections of ation of the bulk of the pedogenic minerals (fig. 5A). selected samples (table 1). B horizons show evidence Kaolinite was produced by the hydrolysis of plagio- for destruction of primary igneous rock texture and clase and occurs as aggregates of 10–50-mmgrainsat development of pedogenic fabrics in clay-rich ma- the site of former plagioclase (fig. 5B). Amphibole terials known as “birefringence fabrics” (b-fabrics), was weathered by a combination of hydrolysis and

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 265

Table 1. Micromorphological Features of Mount Simon Sandstone and Regolith and Saprolite Horizons, Drill Core TR119, Trempealeau County, Wisconsin Depth below Depth from top Horizon surface (ft) of paleosol (cm) Brief thin-section descriptions of micromorphological features NA 304.9 13 Basal Mount Simon Sandstone; bimodal texture with coarse mono- and polycrystalline quarts and partially weathered K-feldspars; organic matter and pedolith (reworked paleosol) grains in finer fraction Bt 305.1 to 306.2 23to237 Pedogenic clay accumulation at top of paleosol; birefringence (b-fabric) in clays; Fe-oxides and oxyhydroxides; fine peds; fine tubular organic structures (29cm);5–10-mm diameter “bacterial” Fe grains (29to237 cm) Bts 307.2 to 308.7 267 to 2113 Pedogenic clay and Fe-Al-organic “sesquioxide” accumulation; b-fabric; fine to weak peds; 5–20-mm-diameter “bacterial” Fe grains; 50-mm-long by 10–20-mm-diameter vermiform siliceous tubular pseudofossils Bs 308.7 to 309.5 2113 to 2137 Pedogenic clay and Fe-Al-organic “sesquioxide” accumulations on fractures and redoximorphic fabric with zones of Fe depletions and enrichments; weak peds; 5–20-mm bacterial Fe grains with nuclei; 50-mm-long by 10–20-mm-diameter vermiform siliceous tubular pseudofossils (137 cm) B/Cr 309.5 to 311.2 2137 to 2180 Partial retention of gabbro rock structure; zones of 20–50-mm- diameter bright UV phosphate aggregates (2137 cm); no peds; 5–20-mm-diameter “bacterial” Fe grains with nuclei; zones of 20–50-mm chloritized aggregates and siderite or ankerite grains (2180 cm) Bk/Cr 311.2 to 313.1 2180 to 2247 Partial retention of gabbro rock structure; multiple zones of siderite/ankerite, some as 5–20-mm-diameter lenticular pseudomorphs after gypsum and some as 100–200-mm-diameter spherules; no peds; 5–20-mm-diameter “bacterial” Fe grains with filamentous bristles (2213 cm); prominent crack with Fe and Mn oxides (2247 cm) Cr/R 313.1 to 313.9 2247 to 2271 Weakly weathered gabbro parent material with patches of Fe oxide and oxyhydroxide as well as clay (sericite-saussurite) alteration; minor siderite/ankerite fracture fills R/Cr 313.9 to 315.7 2271 to 2326 Very weakly weathered gabbro parent material with minor patches of Fe oxide and oxyhydroxide as well as clay (sericite- saussurite) alteration; minor siderite/ankerite fracture fills (2271 cm) Note. NA p not applicable. oxidation, yielding 10–50-µm grains of closely asso- 300-mm rounded aggregates with small, euhedral ciated quartz and chlorite at the site of former am- grains at the rims of the aggregates (fig. 6C). In both phibole (fig. 5C,5D;see“Discussion” for further samples, siderite displays discontinuous, sharp com- explanation). In contrast to metamorphic chlorite, positional zoning, with a core-to-rim increase of which is ripidolite in protolith, pedogenic chlorite is ∼5 mol% in siderite and rhodochrosite components picnochlorite, in which Si increases from 5.8 apfu in and an ∼5 mol% decrease in magnesite and calcite most of the weathered profile to 6.2 apfu in the up- components (fig. 6C,6D; table A4). permost weathered sample and FeTotal decreases from All weathered samples except one contain 3–5wt% 3.2 to 2.4 apfu (table A4). very fine-grained goethite, which is widely dispersed Siderite first appears in small amounts in incip- in general but is locally concentrated in thin diffuse iently weathered sample 313.9, where it (along layers or zones in some samples. An exception to this with chlorite) partly replaces amphibole (fig. 6A). pattern is sample 309.5, which contains 10 wt% In the weathered profile, siderite is confinedtothe goethite concentrated in thin, discrete, subhorizontal lower ∼100 cm of the section, with abundances of layers and veins; goethite in this sample displays a 13% and 20% in samples 312.0 and 313.1, respec- colloform texture that may be microbial in origin tively. In sample 313.1, siderite occurs in veins, (fig. 7; see “Discussion”). where it is associated with goethite (fig. 6B), and Some samples in the B horizon contain a dis- as dispersed euhedral grains ≤50 mmindiameter tinctive vermiform feature that resembles the mi- (fig. 6D). In sample 312.0, siderite occurs as 100– crofossil Siphonophycus (fig. 8), based on compar-

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 266 L . G . M E D A R I S E T A L .

Figure 5. Images of pedogenic minerals. A, Sample 305.1: photomicrograph (plane polarized light). a p fine-grained mixture of quartz and chlorite after amphibole; k p fine-grained kaolinite after plagioclase. B, Sample 305.1: backscattered electron (BSE) image of kaolinite (k) after plagioclase. Thin, bright layers in kaolinite are illite-vermiculite. C, Sample 309.5: BSE image of quartz (q) and chlorite (c) after amphibole. D, Sample 309.5: BSE image. Detail of quartz and chlorite after amphibole. ison with examples provided in Anderson et al. rounded grains, ranging in diameter from 20 to (2017, their fig. 8). The vermiform features have 50 mm, that display adularia habit (fig. 9A,9B). The been largely replaced by illite, but they retain well- feldspar is effectively an end-member potassic phase, defined walls composed of organic matter, as re- containing 0.1 wt% of Na2OandnoCaO(tableA4). vealed by UV fluorescence (not shown here). This Transmission electron microscopy and selected area feature is interpreted to be microbial in origin (see electron diffraction demonstrate that the K-feldspar “Discussion”). has triclinic symmetry (microcline). The cluster tex- Metasomatic, mixed-layer illite-vermiculite is abun- ture of microcline, which is characteristic in the dant throughout the weathered profile (10–25 wt%), TR119weatheredprofile (fig. 9A,9B), has been pro- where it forms mats of extremely small grains dis- duced by potassium metasomatism elsewhere (Pi- persed through the samples (fig. 9A,9B,9D)andlo- rajno 2009). cally replaces kaolinite (fig. 9C). This potassic mixed- layer silicate contains 7.3 apfu of Si (based on a total of 22 apfu of ) and significant quantities of Chemical Changes in TR119 Paleosol FeO (5.0 wt%) and MgO (2.9 wt%; table A4). Relative to Protolith Metasomatic K-feldspar is abundant (13–44 wt%) in the midsection of the weathered profile, but it is Evaluation of Immobile Elements. Quantitative esti- less so (5–9 wt%) in the uppermost and lower parts mates of compositional changes during weathering (table A3; fig. 4). K-feldspar occurs as clusters of can be obtained by comparing the compositions of

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Figure 6. Backscattered electron images of pedogenic siderite. A, Sample 313.9: siderite (s) replacing igneous am- phibole (a). B, Sample 313.1: cross-cutting siderite vein (s) associated with locally abundant goethite (g). C, Sample 312.0: colloform siderite with compositional zoning. cores p 80 mol% siderite; rims p 86 mol% siderite. D, Sample 313.1: euhedral siderite. Core and rim compositions are the same as in C.

Figure 7. Goethite in sample 309.5 displaying colloform texture. A, Photomicrograph, plane polarized light. B, Back- scattered electron image.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Figure 8. Vermiform features in sample 308.7 that are interpreted to be microfossils, possibly Siphonophycus spe- cies. A, Photomicrograph, plane polarized light. B, Backscattered electron image. f p feldspar; i p illite-vermiculite; q p quartz.

Figure 9. Backscattered electron images of metasomatic illite-vermiculite (i), metasomatic potassium feldspar (f), and pedogenic chlorite (c). A, Sample 307.2: clusters of K-feldspar associated with mats of illite-vermiculite. B, Sample 308.7: coarser-grained K-feldspar associated with mats of illite-vermiculite and pedogenic chlorite. C, Sample 305.1: kaolinite (k) partly replaced by illite-vermiculite. D, Sample 305.1: extremely fine-grained felted mat of illite-vermiculite in pedogenic quartz.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 269

weathered and unweathered materials while nor- Depth Variation in Composition. CaO, Na2O, and malizing to an immobile element (or oxide) by means Sr, whose percent changes approach 2100, were ef- of the equation fectively completely removed by weathering, and

MgO, Ba, SiO2, and Al2O3 were substantially re- % change p 100 # ½(cj,w =cj,p)=(ci,w =ci,p) 2 1, moved (fig. 12). Although not shown, removal of

TiO2 was comparable to that for Al2O3, the inte- where cj,w is the concentration of element j in weath- grated changes for the two being 224% and 227%, ered material, cj,p is the concentration of element j in respectively. protolith, ci,w is the concentration of immobile ele- Fe2O3 and MnO are strongly partitioned in the ment in weathered material, and ci,p is the concen- profile, being enriched at depths of 213 and 247 cm tration of immobile element in protolith. (due to the precipitation of siderite) and depleted at Likely candidates as immobile elements during levels above this. Integrated percent changes for weathering are Al, Ti, Zr, and Nb, although Nb is Fe2O3 and MnO are relatively small, as are mass excluded from consideration here because of its fluxes, indicating that these two constituents were extremely low concentration (0.8 ppm) and high partitioned but not removed by weathering. Percent relative standard deviation (141%) in the gabbro changeswithdepthforP2O5 scatter about a value of protolith. Accurate calculation of compositional zero with a negligible mass flux, indicating that fi changes in a weathered pro le depends on the pro- P2O5, like Fe2O3 and MnO, was partitioned within tolith being compositionally homogeneous, and but not removed from the profile. this is not the case for TR119 gabbro, where Al2O3, Potassium (K2OMEAS) was added throughout the TiO2, and Zr have relative standard deviations of weathered profile, as predicted by the A‐C N‐K plot 5%, 46%, and 38%, respectively. In addition, al- (fig. 2), and Rb was added as well. The extremely though the concentrations of Al2O3,TiO2,andZr large values of change, 1000%–2000% for K2OMEAS generally increase upward in the TR119 profile, as and 500% for Rb, originate from their low concen- they should with removal of labile constituents, trations in protolith, 0.32 wt% and 7 ppm, respec- there are numerous excursions from a smooth trend tively, compared with their concentrations in the that are consistent with the existence of composi- paleosol (table A5). The premetasomatic percent fi tional heterogeneities in the protolith ( g. 10). change in potassium (K2OCALC) due to weathering can To correct for the variability of immobile ele- be calculated by judicious application of the A‐C N‐K ments due to compositional heterogeneity, linear plot, the method for which is described by Medaris fi least squares ts have been determined for the depth et al. (2015). Following that method, K2O is seen variations of Al2O3,TiO2, and Zr in regolith, saprolite, to have been substantially removed by weathering, and protolith, as illustrated by the dashed lines in with an integrated change of 267% (fig. 12). How- figure 10. Samples at depths of 213 and 247 cm were ever, this procedure provides only a minimum value fi excluded from the least squares ts due to the pres- for the integrated percent removal of K2O due to ence of siderite and high Fe2O3 content (∼19 wt%), weathering, the value of which is controlled by the which effectively dilutes the concentration of other position of protolith and weathered sample in the oxides, especially Al2O3. These least squares fits for A‐C N‐K plot—that is, the more potassic the pro-

Al2O3,TiO2, and Zr were applied to each sample, tolith, the smaller the calculated percent removal of whose concentration sums were then normalized to K2O. Alternatively, an estimate of the maximum in-

100% to generate transformed compositions for use tegrated percent removal of K2O due to weathering in subsequent calculations (table A5). (K2OEST) may be obtained by assigning the percent

After accounting for heterogeneity in the immo- change of K2O through the weathered profile as equal bile element candidates, the question remains, Which to that for Na2O. The minimum (K2OCALC)andmax- elements are truly immobile? This question may be imum (K2OEST) percent removals for K2O integrated answered by plotting ratios of Al2O3, TiO2, and Zr over the TR119 weathered profile are 267.1% and with depth, as illustrated in figure 11. The ratio 290.2%, respectively.

Al2O3/TiO2 is effectively constant with depth, but the ratios Zr/Al2O3 and Zr/TiO2 increase upward in the weathered profile, demonstrating that Al O and 2 3 Mass Fluxes TiO2 were relatively mobile compared with Zr and were removed to some extent by weathering. Con- The mass flux (removal or addition) of material in a sequently, in this instance Zr is the best choice as an weathered profile can be calculated from the equation immobile element for calculating percent changes in the p r ∫Zt paleosol. mj pcj,p0 j,w dZ,

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 270 L . G . M E D A R I S E T A L .

Figure 10. Depth variation in Al2O3,TiO2, and Zr concentrations in regolith and saprolite (circles) and in protolith (squares). Dashed lines are linear least squares fits to all data except for samples at depths of 213 and 247 cm (see text for explanation). Relative standard deviations for the three constituents in protolith are given in the insets.

where mj is the mass flux of constituent j, rp is the compositions of regolith, saprolite, and average pro- density of the protolith, cj,p is the concentration of tolith. Total mass flux for the weathered profile is 22 constituent j in the protolith, and tj,w is the transfer 26.0 mol cm (table 2), and mass fluxes for indi- (percent change) of constituent j in the weathered vidual oxides are included in figure 12. Mass fluxes profile integrated over the depth of weathering, Z for individual oxides depend to a large extent on

(Chadwick et al. 1990). their protolith concentrations; for example, SiO2 is Mass fluxes for the TR119 weathered profile have the most abundant oxide in the protolith (50 wt%) and been calculated accordingly, using the transformed accounts for the greatest mass flux (22.7 mol cm22)

Figure 11. Depth variation in ratios of Al2O3,TiO2, and Zr in regolith and saprolite (circles) and in protolith (squares). Dashed lines are linear least squares fits to all data except for samples at depths of 213 and 247 cm (see text for explanation).

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 271

Figure 12. Depth variation in percent changes of oxides and elements in the weathered profile relative to Zr in the average protolith. Numbers in parentheses are mass fluxes (mol cm22) for individual oxides integrated over the depth of weathering. in the total, while CaO, which exhibits a greater Values of mass flux would be more intuitively percent change in the profile than does SiO2 (fig. 12), meaningful if expressed in terms other than moles contributes less mass flux (21.5 mol cm22) due to its per square centimeter. For example, mass flux may smaller concentration (11.6 wt%) in the protolith. be expressed as percentage of mass removed (%MR), Oxides that exhibit negative percent changes which is defined as through the profile and have small concentrations in the protolith necessarily yield small negative mass p # mass removed fl p %MR 100 , uxes. Such is the case for potassium (K2Oprotolith complete mass removal

0:32 wt%), with mass fluxes for K2OCALC and K2OEST being 20.018 and 20.024 mol cm22, respectively. where “mass removed” for an oxide is its calcu-

Although Fe2O3 is relatively abundant in the pro- lated mass removed from a weathered profile and tolith (8.5 wt%), its mass flux is only 20.07 mol cm22, “complete mass removal” for that oxide is its mass which indicates that iron was retained overall in the removed for a transfer, tj,w, of 100% throughout the paleosol during weathering despite partitioning in the profile. Resulting values of %MR for the TR119 profile. profile are as follows: Na2O, 90.2; CaO, 89.7; MgO,

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).

This content downloadedfrom 144.092.207.008 onJuly30,2018 13:50:27PM Table 2. Comparison of Weathering Parameters among Sub-Cambrian Paleosols

Weathered Minimum age Depth of weathering, TMF RMF K2O added profile Location Protolith of weathering mean5SE (cm) CIAa FIWb CDF (mol cm22)b %MRb (103 # mol cm23)b (mol cm22)

TR119 Wisconsin Gabbro Late Cambrian 285 5 14 97 (CIACALC ) 89.5, 89.8 .61 26.015, 26.021 50.0, 50.1 221.11, 221.13 .30 c SQ8 Minnesota Monzogranite Late Cambrian 269 5 127 90 (CIACALC ) 73.0, 75.8 .38 22.54, 22.57 23.4, 23.7 29.44, 29.55 .30 c Butler Hill Missouri Alkali feldspar Late Cambrian 472 5 78 70 (CIACALC ) 47.8, 83.1 .38 22.37, 22.86 12.2, 14.7 25.01, 26.05 .33 granite c Squaw Creek Texas Alkali feldspar Middle Cambrian 230 5 70 98 (CIAMEAS ) 79.0 .33 22.66 28.2 211.60 0 granite d Timna Israel Quartz monzodiorite Early Cambrian 760 5 40 87 (CIACALC ) 62.0, 72.3 .32 27.27, 27.61 24.1, 25.2 29.57, 210.01 .68 e Longnan China Monzogranite Modern 1063 93 (CIAMEAS ) 59.9 .30 210.11 23.2 29.52 0 f Piedmont Virginia Syenogranite Modern 2050 86 (CIAMEAS ) 70.0 .72 226.75 31.7 213.00 0 f Piedmont Virginia Diabase Modern 95 92 (CIAMEAS ) 80.8 .68 22.42 54.0 225.50 0 Note. %MR p percentage of mass removed; CDF p chemical depletion fraction (cf. Zr, except Ti for Squaw Creek and Piedmont); FIW p feldspar index of weathering; RMF p reduced mass flux; TMF p total mass flux. a Chemical index of alteration (CIA), either calculated (CIACALC ) or measured (CIAMEAS ). b Maximum and minimum values for FIW, TMF, %MR, and RMF in K-metasomatized profiles. c Driese et al. (2007). d Sandler et al. (2012). e Liu et al. (2016). f Pavich et al. (1989). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 273

67.2; K2OCALC, 67.1; SiO2, 40.6; and Al2O3, 27.1. Total original PPM (PPM1.0), demonstrating that the ex- %MR is simply the sum of calculated mass fluxes clusion of Zr does little to affect the model results for all of the oxides divided by the sum of their (G. E. Stinchcomb, L. C. Nordt, S. G. Driese, W. E. mass fluxes, taking tj,w to be 100% for each oxide Lukens, and D. J. Peppe, unpub. manuscript). The throughout the profile. The total %MR for the residual errors between models are similar, and the

TR119 profile is 50.0, that is, one-half of the proto- root mean squared errors for PPM1.1,MAPp 223 mm lith mass was removed by weathering. Note, how- and MAT p 2:527C, are comparable to those for ever, that %MR is a relative quantity and that total PPM1.0, MAP p 228 mm and MAT p 2:467C. mass flux is an absolute quantity (see “Discussion”). Mean Annual Precipitation. The two uppermost TR119 paleosol samples were selected for applica- tion of MAP estimates because the upper part of Paleoclimate Functions and Atmospheric PCO 2 the profile has characteristics most appropriate for Empirical relations among bulk geochemistry, weath- comparison with B horizons of the modern soils ering, and climate of modern soils have been used to used to develop the PPM model (Stinchcomb et al. develop paleoclimate proxies (a.k.a. climofunctions) 2016). Estimates of MAPCIA-K (Sheldon et al. 2002) and soil characterization proxies (e.g., pH) that allow and MAPPPM (Stinchcomb et al. 2016), based on the for interpretations of past climate and soil-forming average of the two uppermost paleosol samples, are conditions (Sheldon and Tabor 2009; Lukens et al., 1509 and 1777 mm y21, respectively, which are forthcoming). However, accurate application of such within error of each other (table 3). functions requires that properties of the paleosol in Mean Annual Temperature. MAT was estimated question lie within the range of properties of modern by the PPM model of Stinchcomb et al. (2016) and soils on which the climofunctions and character- the MATC calibration of Sheldon (2006a), which is ization proxies are based. The most obvious failure defined as of this requirement is that sub-Cambrian paleosols 7 p : 1 : lacked a vegetative cover, although organic ligands MAT(C) C 46 94C 3 99, may have been provided by microbial activity. With this caveat in mind and in lieu of acceptable alter- where C p mAl2O3=mSiO2. natives, proxies developed by Sheldon et al. (2002), MATPPM for the TR119 paleosol is 20.17C, which

Sheldon (2006a), and Stinchcomb et al. (2016) were is somewhat higher than the MATC value of 16.77C used to estimate mean annual precipitation (MAP) (table 3). It is worth noting that both MAT esti- and mean annual temperature (MAT), and a calibra- mates are within error of present-day subtropical tion by Lukens et al. (forthcoming) was used to esti- temperatures. mate pH, a master soil variable where water avail- pH. A regression-based bulk geochemical trans- ability is a first-order control (Slessarev et al. 2016). The fer function, AlCa, was developed by Lukens et al. model developed by Stinchcomb et al. (2016), PPM1.0, (forthcoming) for estimating the paleo-pH of B ho- was modified to exclude the element Zr from the rizons of paleosols and is defined as analysis, as not all paleosols studied had reported Zr. p # ½ = 1 : The PPM without Zr (PPM1.1) was compared with the AlCa 100 Al2O3 (Al2O3 CaO)

Table 3. Comparison of Climofunctions for the TRl19 and Other Cambrian Paleosols MAP (mm y21)b MAT (7C)b Weathered PPM1:1 PPM1:1 21 a c d e profile MAPCIA-K (mm y ) Low Best HighMATC (7C) Low Best High pH T (y) TRI19 1509 1456 1777 2097 16.7 17.7 20.1 23.0 5.2 38,000 SQ8f 1366 825 1182 1539 15.0 12.7 15.9 19.0 6.0 16,000 Butler Hillf 1442 204 605 1007 8.2 8.7 12.2 16.0 5.3 33,000 Squaw Creekf 1566 719 1104 1489 10.9 10.0 13.2 16.0 4.8 17,000 Timnag 1529 122 1096 1469 13.0 12.4 15.7 19.0 5.3 128,000 Note. MAP p mean annual precipitation; MAT p mean annual temperature. a After Sheldon et al. (2002). SE p 181 mm y21. b After Stinchcomb et al. (2016). c After Sheldon (2006a). SE p 0.67C. d After Lukens et al. (forthcoming). Root mean squared error p 0.83. e Formation time for atmospheric CO2 p 5000 ppm. f Driese et al. (2007). g Sandler et al. (2012).

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 274 L . G . M E D A R I S E T A L .

Regression of second- and third-order polynomial that siderite formation requires saturated condi- fits of modern soil pH (n p 619) versus AlCa pH tions, along with reducing conditions and available yields R2 values of 0.66 and 0.67, respectively, and organic matter (Driese et al. 2010; Ludvigson et al. root mean square errors of 0.86 and 0.84, respec- 2013). This can be achieved by either endosatura- tively. tion (saturation achieved by rising water table from For the upper two paleosol samples, the average below) or episaturation (saturation achieved by wa- pH value is 5.2 (table 3), which is consistent with ter table perched on top of paleosol). In this in- advanced acid attack on aluminosilicate minerals stance, there is no evidence for gley (low-) and the absence of plagioclase and amphibole from conditions in the upper part of the paleosol, so an the weathered profile. This pH value is maintained endosaturation scenario makes more sense. During within error through the profile to lower depths, siderite precipitation, the water table in the TR119 where it increases to more basic values of 6.1 and profile must have been relatively high, certainly 6.5 at depths of 213 and 247 cm, respectively, which higher than the saprolite/protolith interface at 285 cm are compatible with the occurrence of siderite at and, conservatively, at a depth as shallow as the top these depths. Such inferred acidic conditions in the of the siderite zone at 185 cm. Raising the depth of upper part of the profile also suggest a net surplus in the water table from 285 to 185 cm decreases the available moisture, which is consistent with our value of PCO2 by a factor of 1.5 (fig. 13B). MAP reconstructions. Among variables in the equation, the parameter T

Atmospheric PCO2. Sheldon (2006b) proposed a (formation time) is the most uncertain and has a thermodynamic model for estimating paleoatmos- large influence on the magnitude of calculated PCO2; pheric CO2 levels from the mass balance of paleosol a factor of 10 difference in time yields a factor of weathering. The essential equation is 10 difference in calculated PCO2.However,this

strong dependence of calculated PCO2 on time can = ≈ ½ = 3 1 k a = M t PCO2 (KCO2 r) 10 ( DCO2 ) L , be turned into an advantage because if the level of atmospheric CO2 during weathering is known, then where M/t is the time-averaged flux of CO2 required the time of weathering can be estimated. According 22 21 for the observed weathering (mol cm y ), PCO2 is to the GEOCARBIII model of Berner and Kothavala the partial pressure of atmospheric CO2 (atm), KCO2 (2001), the concentration of CO2 in the atmosphere is the Henry’s law constant for CO2, r is precipita- at 550–500 Ma was between 4000 and 6000 ppm, tion (cm y21), k is a constant (1:43 # 103 scm3 mol21 corresponding to values of 14.4 and 21:5 # PIAL, 21 y ), DCO2 is the diffusion constant for CO2 in air respectively, where PIAL refers to a preindustrial 2 21 (0.162 cm s ), a is the ratio of the diffusion con- atmospheric level of 280 ppm. These levels of CO2 stant for CO2 in soil divided by that in air, and L is constrain weathering times to between 32,000 and the depth to the water table (cm). 21,000 y for a water table depth of 185 cm (fig. 13B).

During silicate weathering, 2 mol of CO2 are re- These values are minima, however, because an un- quired for the liberation of 1 mol of the base cations, known portion of the weathered profile may have Mg, Ca, Na, and K. Although Si is not considered in been removed by erosion. For example, if 50% of the Sheldon’s model, it has been included here because paleosol were eroded, the time of weathering would

41% SiO2 has been removed from the TR119 pal- have been between ∼200,000 and ∼140,000 y (fig. 13B). eosol. In contrast to the MO oxides, 4 mol of CO2 is required for the removal of 1 mol of SiO .Inthe 2 Comparison of Weathering Parameters TR119 profile, the mass flux removal of MgO, CaO, 22 in Sub-Cambrian Paleosols Na2O, and K2OCALC is a total of 2.9 mol cm ,which 22 requires a CO2 flux of 5.8 mol cm to account for To evaluate weathering and climatic conditions in the observed magnitude of weathering. When SiO2 Cambrian time during formation of the Great Un- is included, these values increase to 5.6 and 16.6 mol cm22, conformity, properties of the TR119 paleosol are respectively. The effect of including the mass flux of compared with those of three other sub-Cambrian

SiO2 is to increase the calculated PCO2 value by a paleosols in North America (Driese et al. 2007) and factor of 2.8, as illustrated in figure 13A. one in Israel (Sandler et al. 2012), including SQ8

In calculating PCO2, depth to the water table (L)is (drill core beneath Upper Cambrian Mount Simon asignificant parameter, but its former position in Sandstone, southeastern Minnesota), Butler Hill (drill the TR119 profile is uncertain. Siderite formation is core beneath Upper Cambrian Lamotte Sandstone, intricately linked to depth to the water table; most southeastern Missouri), Squaw Creek (an outcrop models of pedogenic siderite formation and obser- beneath Middle Cambrian Hickory Sandstone, cen- vations of pedogenic siderite in paleosols indicate tral Texas), and Timna (outcrop beneath Lower Cam-

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 275

Figure 13. Cross plots of PCO2 # PIAL and formation time of weathering, calculated following Sheldon (2006b). Note the differences in scale for the two panels. A, Comparison of results between including the mass flux of SiO2 in the calculation and excluding it. Both are for a water table depth (L) of 185 cm. B, Comparison of results illustrating the effect of table depth (L) at 185 and 285 cm and the effect of 30% and 50% erosion. All calculations include the mass flux of SiO2 and levels of atmospheric PCO2 (4000 and 6000 ppm) at 550–500 Ma from the GEOCARBIII model of Berner and Kothavala (2001). The stippled area constrains the estimated range in formation time of weathering for the TR119 paleosol. brian Amudei Shelomo Sandstone, southern Israel). be evaluated as a function of a single factor by hold- Transformed whole-rock compositions and values ing other factors constant, such as of K O were generated for these four paleosols 2 CALC s p f(cl, u) , following the procedure described for TR119, and o,r,p,t,::: their resultant salient properties are summarized in where s varies as a function of climate with other table 2. Note that the profile at the Squaw Creek factors being held constant. The term u is an un- locality escaped K-metasomatism, whereas the other certainty factor, recognizing that factors assumed to three were metasomatized. In addition, Zr analyses be constant actually vary. fi are unavailable for the Squaw Creek pro le, so TiO2 The paleosols compared here are similar in being was used as the immobile normalizing oxide in this overlain by flat-lying, fluvial, or shallow-marine instance. A sub-Cambrian paleosol in South Dakota sandstones and thus formed under subdued paleo- that developed from a metagabbro protolith (Horo- topography. They are also similar in having formed dyskyj et al. 2012) was not included for comparison prior to the advent of land plants in Ordovician time, because it experienced Mg- and Ca-metasomatism and organic ligands involved in chemical weathering in addition to K-metasomatism, which precludes an were derived from microbial activity and simple accurate assessment of weathering parameters. plant forms, such as algae, rather than more complex Comparison of the paleosols is based on the fun- plant forms. The major difference among the paleo- damental state-factor theory of soils (Jenny 1941, sols is protolith lithology, which varies from gabbro 1980): to alkali feldspar granite. They also differ in the min- imum age of weathering, which ranges from Late S or s p f(cl, o, r, p, t, :::), Cambrian to Early Cambrian. Numerous methods have been proposed to eval- where S is soil, s is some soil property, cl is climate, o uate the degree, or maturity, of chemical weather- is organisms, r is relief, p is parent material, and t is ing, including the chemical index of alteration (CIA; time. The ellipsis represents other possible factors Nesbitt and Young 1982), the chemical depletion unaccounted for that influence S. A property s can fraction (CDF; Riebe et al. 2003), the feldspar index

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 276 L . G . M E D A R I S E T A L . of weathering (FIW; Medaris et al. 2015), and re- resulting from vigorous physical reworking coupled duced mass flux (RMF; Medaris et al. 2017). In with advanced sub-Cambrian chemical weathering weathered profiles, one can distinguish between enhanced by subaerial microbial mats and absolute and relative degrees of weathering, where (Dott 2003). However, K-feldspar is an abundant con- absolute values are given by an element ratio for stituent in silty fractions of the Cambrian sedimen- an individual weathered sample and relative values tary sequences, revealing its persistence in source are given by mass fluxes averaged over an entire regions for Cambrian detritus (Odom 1975; Odom weathered profile; among the four functions con- et al. 1976). sidered here, CIA and CDF are absolute measures of Chemical Depletion Fraction. For a soil undergo- weathering, and FIW and RMF are relative mea- ing steady-state formation, erosion, and weathering, sures of weathering. total denudation flux is equal to the sum of chemical Chemical Index of Alteration. The chemical ma- weathering flux and physical erosion flux. The func- turity of weathering may be evaluated quantitatively tion CDF was introduced by Riebe et al. (2003) to by means of the CIA (Nesbitt and Young 1982), express the relation between chemical weathering which is defined as flux, total denudation flux, and the ratio of an im- mobile element in soil and protolith, as follows: Al O CIA p molar 2 3 # 100:   Al O 1 CaO 1 Na O 1 K O (X) W 2 3 2 2 CDF p 1 2 protolith p , (X)soil D A CIA value of 100 corresponds to the end stage of chemical weathering, in which CaO ,Na2O, and where X is the concentration of an immobile ele- K2O are completely removed from the paleosol. Di- ment, W is the chemical weathering flux, and D is rect calculation of CIA in K-metasomatized sam- the total denudation flux. Although this equation ples, however, yields spurious values due to dis- was derived to evaluate the relative rates of chemi- placement of whole-rock compositions from the cal removal and total denudation during weather- predicted weathering trend by the addition of potas- ing, it can also be used as an estimate of the degree of sium. The effect of K-metasomatism may be cor- chemical weathering because CDF values increase rected by projecting a line from the K-apex through a as labile elements are removed by weathering and metasomatized sample to an intersection with the concentrations of immobile elements correspond- predicted weathering trend to yield a premetaso- ingly increase. matic, calculated CIA value (CIACALC; Fedo et al. CDF values for the four granitic paleosols are 1995). However, the ultimate value for CIACALC in a comparable, 0.32–0.38, but that for TR119 is much weathered profile is controlled by protolith compo- larger, 0.61 (table 2), reflecting a greater relative sition and intersection of its predicted weathering chemical weathering flux and degree of weathering trend with the A-K join in the A‐C N‐K diagram— of TR119 gabbro compared with weathering of gra- that is, the higher the protolith K2O content, the nitic rocks in the other paleosols. lower the limiting value of CIACALC. Feldspar Index of Weathering. Because plagioclase The Squaw Creek paleosol, which is not metaso- and potassium feldspar are abundant phases in most matized, yields a measured CIA value of 98 (table 2), common igneous rocks (∼50–80 vol%), the amount which reflects complete removal of plagioclase and of feldspar removed from paleosols is a good mea- K-feldspar in the upper part of the profile. TR119, sure of the overall degree of weathering. Accord- whose gabbroic protolith is devoid of K-feldspar and ingly, to evaluate the relative removal of feldspar in plots very close to the A‐C N join (fig. 2), yields a a weathered profile, Medaris et al. (2015) introduced CIACALC value of 97, consistent with complete re- the parameter FIW, which is defined as follows: moval of plagioclase throughout the profile. Values of CIA for the other three paleosols are 70, 87, MF CALC FIW p 100 # CNK , and 90 (table 2), which are directly correlated MF100 2 (R p 0:97) with their protolith K2O concentra- tions. Although K-feldspar remains at depth in these where MFCNK is the mass flux of CaO, Na2O, and three profiles, it was removed from the uppermost K2O removed from a paleosol, either measured di- levels, and the premetasomatic CIA values were rectly in unmetasomatized paleosol or calculated in most likely ≥95. K-metasomatized paleosol, and MF100 is the mass

Supermature quartz arenites are a conspicuous flux calculated for complete removal of CaO, Na2O, feature of Cambrian sedimentary rocks in the North and K2O over an entire weathered profile—that is, American interior and the Arabian-Nubian Shield, complete removal of both plagioclase and K-feldspar.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 277

Although the extent of feldspar removal from in- 25.0, 29.4, and 29.6 for the other three paleosols dividual samples is effectively illustrated by a CIA (table 2). plot, the FIW has the advantage of quantitatively K-metasomatism. As was done for TR119 (fig. 12), evaluating the removal of feldspar over an entire the depth variations of K2OMEAS and K2OCALC were paleoweathered profile. FIW values range from 0 in determined for the three metasomatized granitic unweathered protolith to 190 in mature paleosols. paleosols, and the total mass flux of K2O added to FIW values never reach 100 because even though each of the four metasomatized profiles has been feldspar may be completely removed from the upper calculated (table 2). Remarkably, the total mass levels of mature paleosols, some feldspar com- fluxes of K2O for the three metasomatized North monly remains at deeper levels. In addition, even in American paleosols are essentially the same (0.30– paleosol domains from which feldspar has been 0.33 mol cm22) regardless of protolith composition. completely removed, small amounts of CaO, Na2O, In contrast, the mass flux of K2O for the Timna 22 and K2O remain (on the order of 1 wt% total). paleosol in Israel is twice as large, at 0.68 mol cm , The minimum and maximum FIW values for but note that the preserved depth of weathering for TR119, whose protolith is devoid of K-feldspar, are the Timna paleosol is also approximately twice that 89.5 and 89.8, respectively (table 2), indicating that of the others. plagioclase was removed by weathering through most of the paleosol, as illustrated by the depth dis- tribution of CaO and Na2O in the weathered profile Discussion (fig. 12). Similar plots (not shown) for the other four Pedogenic Minerals. A remarkable feature of the granitic paleosols reveal that plagioclase was largely TR119 paleosol is the abundance of quartz (11%– removed through the profiles in each. In contrast to 23%; table 3) when little quartz occurs in the gabbro plagioclase, significant amounts of K-feldspar re- protolith and 40% SiO was removed from the main at depth. The FIW value for the unmetaso- 2 weathered profile (22.7 mol cm22; fig. 12). This ap- matized Squaw Creek paleosol is 79, and minimum parently anomalous situation arises from weather- and maximum FIW values for the other three meta- ing of amphibole, the reaction for which is expressed somatized granitic paleosols are 48 and 83, 62 and qualitatively as 72, and 73 and 76. The different ranges and averages of FIW values for the three metasomatized paleosols amphibole 1 H2O 1 CO2 → quartz 1 chlorite reflect the different K2O contents of their granitic 21 21 1 protoliths and represent not different magnitudes of 1 goethite 1 Mg 1 Ca 1 Na weathering but rather different protolith composi- 1 2 1 K 1 HCO3 , tions.

Reduced Mass Flux. In a given catchment basin, where SiO2 is precipitated as quartz rather than be- depths of weathering and total mass fluxes vary with ing removed in solution, as in the case for hydrolysis topographic position, that is, being different on ridge of feldspar. The labile elements in amphibole, Ca, lines, hill slopes, and valley floors (Bazilevskaya Na, and K, are removed in solution, as is some Mg, et al. 2012). Thus, direct comparison of total mass with the remaining Mg being transferred to chlorite. fluxes is useless for distinguishing different degrees Through the entire profile, the mass flux of MgO is of chemical weathering among paleosols that may 21.2 mol cm22 (fig. 12), which corresponds to 67% have formed at different topographic positions from bulk removal. Iron is both precipitated as goethite different protoliths. However, mass flux can be used and incorporated in chlorite rather than being re- as a basis for comparison by dividing total mass flux moved from the profile, which is consistent with its (mol cm22) by depth of weathering (cm) to provide negligible total mass flux, 20.07 cm22 (fig. 12). an average mass flux (mol cm23) over the depth of The stability fields for quartz, Fe-chlorite, and weathering, which is designated as the RMF (Me- goethite in terms of Eh and pH in the system, Fe-Al- daris et al. 2017): Si-O-H, were calculated by means of the Geochem- ist’s Workbench (GWB 11) software package with total mass flux the thermos database. Figure 14 was generated us- RMF p 1000 # , depth of weathering ing the following input parameters: T p 207C, 24 25 PTotal p 1 atm, a(Fe) p 10 M, a(Al) p 10 M, and 2 where the factor 1000 is introduced for convenience a(Si) p 2 # 10 4 M (note that quartz is stable in the 2 of scale. system at a(Si) ≥ 1:58 # 10 4 M). These parameters The RMF values are 221.1 for TR119, 211.6 for were chosen when considering the mineral assem- the unmetasomatized Squaw Creek paleosol, and blage of the paleosol and potential surface water

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 278 L . G . M E D A R I S E T A L .

Figure 14. Cross plots of Eh and pH for the systems Fe-Al-Si-O-H (left) and Fe-Al-Si-O-C-H (right)atT p 207C and p fi PTotal 1 atm, illustrating the stability elds for quartz, goethite, Fe-chlorite, and siderite. See text for input parameters. conditions, following Driscoll (1985) and Lee and form of soluble silicic acid, according to the reac- Wilkin (2010). Quartz is in equilibrium with goe- tion thite and Fe-chlorite at Eh ≊ 0:1 V and pH ≊ 5:6. 21 This latter value is in good agreement with the es- plagioclase 1 H2O 1 CO2 → kaolinite 1 Ca timated pH (5.2) determined by the method of Lu- 1 2 1 Na 1 HCO3 1 H4SiO4, kens et al. (forthcoming).

The stability range of siderite was evaluated by which resulted in a total loss of 40% SiO2 from the adding C to the system, using the input parameters profile. The paradox originating from the bulk re- 2 22 above, and setting a(HCO3 ) p 10 M(fig. 14). moval of SiO2 and concurrent precipitation of ped- Siderite is in equilibrium with quartz, goethite, and ogenic quartz is resolved by recognizing that pedo- Fe-chlorite at Eh ≊ 0:1 V and pH ≊ 6:0. As expected, genic quartz was produced by the weathering of this pH value is greater than that for the siderite-free hornblende while SiO2 was removed in solution in assemblage, and it is in good agreement with the pH the form of silicic acid due to hydrolysis of plagio- range (6.1–6.5) estimated for the siderite-bearing clase, resulting in a net negative mass flux of SiO2 in samples by the method of Lukens et al. (forthcom- the profile. ing). Note that allowing for a variation in a(Al) be- Mass Flux of Weathering. To provide perspective tween 1024 and 1026 M yields an even closer corre- on the degree of sub-Cambrian weathering and the spondence between the pH values calculated here influence of protolith composition, we compared mass and those determined by the method of Lukens et al. flux parameters of the investigated sub-Cambrian (forthcoming). paleosols, which formed at ∼207S paleolatitude, and

Triangular plots of Si-FeTotal-Al and Si-Mg-Al (fig. 15) three modern weathered profiles (table 2), one of illustrate the topological relations among amphi- which formed from monzogranite in a subtropical bole, quartz, chlorite, and goethite and allow propor- setting in China (latitude, 257N; Liu et al. 2016) and tions of pedogenic phases to be estimated through two of which formed from diabase and syenogranite application of the lever rule. In the Fe-bearing sys- in a temperate climate in the Piedmont of Virginia tem the proportion of quartz to chlorite to goethite (latitude, 397N; Bazilevskaya et al. 2012). Because is 48∶47∶5, and in the Mg-bearing system the pro- plagioclase is more susceptible to weathering than portion of quartz to chlorite to Mg is 40∶52∶8, with K-feldspar and quartz, one would expect to find 8% Mg being removed in solution. greater depths and total mass fluxes of weathering in The weathering of plagioclase was responsible for mafic rocks than felsic rocks. However, the reverse is precipitation of kaolinite and loss of SiO2 in the found in modern weathered profiles worldwide, and

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 279

Figure 15. Molar plots of mineral compositions in the TR119 paleosol. A, Molar plot of Si-FeTotal-Al. In this com- positional space, igneous amphibole weathers to a pedogenic assemblage of quartz (48%), chlorite (47%), and goethite (5%). B, Molar plot of Si-Mg-Al. In this compositional space, amphibole weathers to an assemblage of quartz (40%) and chlorite (52%), and 8% Mg is removed in solution. Am p amphibole; Chl p chlorite; Qtz p quartz. in the Piedmont syenogranite exhibits a greater depth Although protolith composition is not correlated of weathering and total mass flux than diabase (2050 with the absolute parameters, depth of weathering cm and 226.8 mol cm22 vs. 95 cm and 22.4 mol cm22, and total mass flux, it is correlated with the relative respectively; table 2; Bazilevskaya et al. 2012). This parameters, RMF and %MR. As illustrated in fig- phenomenon is attributed to differences in the dis- ure 16, there is a perfect correlation between RMF tribution and densities of micron-sized pores and and %MF, which is not surprising because both are microfractures that control fluid ingress in the two functions of total mass flux. The utility of this cross types of protolith. plot is the demonstration that both RMF and %MR Thus, it appears that physical properties of pro- increase from felsic to mafic protolith compositions tolith may override the influence of mineral com- (compare the results for the most felsic protolith, position with respect to depth of weathering and Butler Hill granite, with those for the two most total mass flux. Indeed, depths of weathering and mafic protoliths, TR119 gabbro and Piedmont dia- total mass fluxes in the five sub-Cambrian paleo- base). The effects of possible erosion on values of sols and three modern weathered profiles are not RMF and %MR are shown for the two extreme sub- correlated with composition, regardless of the choice Cambrian protolith compositions, Butler Hill and of compositional parameter, such as mineral modes, TR119, which are calculated for 50% erosion. weight percent oxides, or normative differentiation The correlation of protolith compositions with rel- index. Direct comparison of depths of weathering and ative weathering parameters is further illustrated in total mass fluxes among sub-Cambrian paleosols and figure 17, where RMF and %MR are plotted against those in modern profiles (table 2) is problematic be- the differentiation index, which is the sum of nor- cause of the likely removal of some portions of the mative quartz, orthoclase, and albite in protolith. For sub-Cambrian profiles by erosion. Although depths the three modern profiles, values of R2 for RMF and of weathering and total mass fluxes of sub-Cambrian %MF are 0.99 and 0.97, respectively, although these profiles derived from felsic protoliths are less than high values are dictated by the large gap in data those in the modern profiles, the depth of weathering points. For the sub-Cambrian paleosols, values of R2 and total mass flux for the TR119 gabbro are signifi- for RMF and %MR are 0.78 and 0.79, respectively, cantly greater than those for the Piedmont diabase but some of the scatter in these data undoubtedly (285 cm and 26.0 mol cm22 vs. 95 cm and 22.4 mol arises from some unknown degree of erosion of the cm22) despite the possibility that some portion of the profiles. Assuming 50% erosion of each of the sub- TR119 paleosol may have been removed by erosion. Cambrian profiles, values of R2 for both RMF and

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 280 L . G . M E D A R I S E T A L .

Phosphorus net flux in TR119 is likewise very small, and P is partitioned with depth (fig. 12), al- though not coinciding with Fe partitioning. Phos- phorus distribution in the Timna profile is the most unusual among the studied profiles, with extensive net gains in the upper 150 cm that resemble the ef-

Figure 16. Cross plot of averages of reduced mass flux (RMF) and percentage of mass removed (%MR) in sub- Cambrian paleosols and the Longnan and Piedmont modern weathered profiles. White circles show observed values for sub-Cambrian paleosols, and gray circles show values prior to 50% erosion for Butler Hill and TR119. RMF and %MR values for the modern Longnan and Pied- mont weathered profiles are indicated by square and dia- mond symbols, respectively.

%MR increase to 0.88, and the resulting least squares fits are in close agreement with those for the modern profiles. Thus, in terms of RMF and %MR, the rela- tive degree of weathering in the sub-Cambrian paleo- sols is comparable to that in modern soils in sub- tropical to temperate settings. Such correspondence may originate from a fortuitous combination of an elevated level of atmospheric CO2 (15–20 # PIAL) and a reduced influence of organic ligands (absence of vegetative cover) in the sub-Cambrian paleosols compared with that in modern soils. Although Fe is partitioned with depth in the TR119 (fig. 12), Butler Hill, Squaw Creek, and Timna profiles, its total mass flux in each is negligible. The negligible mass flux for Fe in the TR119 profile is similar to that calculated for the 1700 Ma Baraboo profile in south-central Wisconsin, for which varia- tions in percent changes of Fe with depth are attrib- fl uted to activities of various microbes utilizing either Figure 17. Cross plots of averages of reduced mass ux reduction or oxidation reactions as part of their me- (RMF) and percentage of mass removed (%MR) in sub- Cambrian paleosols versus differentiation index in pro- tabolism (Medaris et al. 2017). In contrast, Fe was toliths. White circles show observed values for sub- substantially removed (71%) from the existing SQ8 Cambrian paleosols, and gray circles show values prior to fi fi pro le, but the preserved thickness of this pro le is 50% erosion for sub-Cambrian paleosols. RMF and %MR relatively small for a felsic protolith (230 cm), and it is values for the modern Longnan and Piedmont weathered possible that erosion destroyed what may have been profiles are indicated by square and diamond symbols, an original partitioned distribution of Fe. respectively.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 281 fects of P bioconcentration in modern weathered dependent of potassium, and we attribute these dif- profiles (not shown). Horodyskyj et al. (2012) found ferences to the high K2OCALC value throughout the geochemical evidence for P biocycling by mycor- saprolite, which is similar to K2OMEAS values in the rhizal fungi in the upper 30 cm of a 3-m-thick red protolith. The obvious limitation of using the PPM paleosol profile, the Elk Point paleosol (overlain by for reconstructing climate in this case is the inclu- the Mount Simon Sandstone of Middle Cambrian sion of K2O, where this oxide has been modified by age), formed on 1.73 Ga metagabbro in South Da- metasomatism. Even when K2O values are adjusted kota. However, because minerals containing P are (K2OCALC), the MAPPPM estimates remain lower than commonly variably distributed in protoliths, bulk those from MAPCIA-K, except for TR119. geochemical analyses for P can be sample biased, The best estimates of MATPPM are all consistently and this may explain the noise observed for P in higher than those predicted by MATC, with only many of the profiles examined. SQ8, Squaw Creek, and Timna MATC predictions

Metasomatism. Potassium metasomatism is a being within error of MATPPM ranges. Overall, the common phenomenon in sub-Cambrian paleosols range of MATPPM estimates approach values more in and Cambro-Ordovician sedimentary rocks in the line with modern-day subtropical temperatures than midcontinent region of North America, and such do the MATC estimates. metasomatism has been attributed to continental- As described above, the level of atmospheric CO2 scale brine migration related to the Pennsylvanian- at 550–500 Ma is estimated to have been between Permian Allegheny and Ouachita orogenies (Bethke 4000 and 6000 ppm (Berner and Kothavala 2001). and Marshak 1990; Fedo et al. 1995). In the TR119 For a concentration of 5000-ppm CO2 in the atmo- drill core, evidence for metasomatism is provided by sphere, the formation times of weathering for the K-feldspar overgrowths on -size detrital K-feldspar TR119, SQ8, Butler Hill, Squaw Creek, and Timna in the Mount Simon Sandstone and an abundance paleosols are thus 38,000, 16,000, 33,000, 17,000, of fine-grained illite-vermiculite (10%–25%) and K- and 128,000 y, respectively (table 3). These results feldspar (5%–44%) in the weathered profile (table A3; were obtained by including the mass flux of SiO2 in figs. 4, 9). The age of K-metasomatism in drill core each paleosol and, because of the uncertainty in TR119 is yet to be determined, due to the analytical water table depth, setting the water table depth at difficulties in dating the extremely fine-grained as- the saprolite/protolith contact in each case. Con- sociated potassic phases, illite and K-feldspar. How- sequently, the differences among the results for the ever, Liu et al. (2003) obtained 40Ar/39Ar ages for dif- times of weathering arise largely from different depths ferent textural varieties of metasomatic K-feldspar in of weathering. For example, the depth of weathering weathered granite located in west-central Wisconsin in the Timna paleosol is 760 cm, but if the water in the vicinity of TR119. Coarse-grained K-feldspar table is assigned a depth of 250 cm, similar to that in yielded plateau ages of 446–427 Ma (Silurian), and the North American paleosols, the time of weath- fine-grained K-feldspar, which in some respects is ering is reduced from 128,000 to 46,000 y. Regard- texturally similar to that in TR119, gave integrated less, the time required for a steady-state weathered ages of 365–359 Ma (Late Devonian). However, profile to be established in the paleosols was be- these latter ages should be viewed with caution, due tween 104 and 105 y. These estimates of formation to the possible effects of recoil in such fine-grained times for the sub-Cambrian paleosol are in the same material. range as those for modern soils in subtropical envi- Paleoclimate Functions. Consistent with the ap- ronments, which range widely (from 103 to 106 y) and proach taken with TR119, we excluded all but the which are highly dependent on other soil-forming upper two samples in each of the four sub-Cambrian factors, such as climate, organisms, relief, and parent paleosols from North America and Israel for calcu- materials (Birkeland 1999). The most extreme case lating paleoclimate functions, the logic being the for long duration of soil formation in the tropics same as with TR119. For example, such samples and subtropics is the deeply weathered soil most likely represent the uppermost hori- order, which today forms on very old landscapes zons and thus more closely reflect those subsoil with formation times of 106 to 107 y and from par- samples used to develop the PPM (Stinchcomb et al. ent materials that have undergone multiple cycles of 21 2016). Values of MAPPPM (1777 mm y ) and MATPPM weathering, erosion, redeposition, and renewed weath- (20.17C) for TR119 are the highest paleoclimate ering ( Staff 2014). estimates compared with those for SQ8, Butler Hill, Microfossils. Previous authors have reported mi-

Squaw Creek, and Timna (table 3). The MAPPPM es- crobial features in Precambrian marine rocks, now timates for Butler Hill are significantly lower than going back to as old as 3.7 Ga, preserved in stro- the corresponding MAPCIA-K estimates, which are in- matolites in the Isua supracrustal belt in Greenland

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). 282 L . G . M E D A R I S E T A L .

(Nutman et al. 2016). Vermiform features occur in protolith at 207S paleolatitude prior to transgression Precambrian sedimentary successions, many of which of the Cambrian sea onto the North American cra- are of a larger size than what we observe in TR119 ton. The paleosol is the product of a high degree (e.g., Donaldson 1967; Seilacher et al. 2000). The of weathering, having experienced a 50% loss in vermiform features in the TR119 profile are com- mass, with a steady-state weathered profile having posed of vermiculite-illite (fig. 8) but are not the same formed in an interval of ∼40,000 y. When the effects size and shape as the Archaeoscilla species and of erosion and influence of protolith composition Primaevifilum species recently reexamined in the are considered, the TR119 paleosol and four other 1.88 Ga Gunflint Chert by Brasier et al. (2015), who sub-Cambrian paleosols exhibit an advanced degree interpreted them as pseudofossils even though some of weathering, in both magnitude and rate of weath- of the features contain organic residues. Instead, ering, comparable to that of modern soils in sub- these 10–15-mm-diameter tubular structures in the tropical and temperate climates, despite the lack of TR119 paleosol more closely resemble the Siphono- land plants in Cambrian time. phycus species microfossils illustrated in Knoll et al. (1991) and in Anderson et al. (2017), and we therefore interpret them to be true microfossils (albeit re- ACKNOWLEDGMENTS placed by illite) rather than pseudofossils. An or- ganic origin for the vermiform features in TR119 Samples of the TR119 drill core were provided by supports the possibility of paleosol microbial activity the Wisconsin Geological and Natural History Sur- that enhanced weathering through the production vey and are stored at its Mount Horeb Research and of organic ligands. The colloform goethite in TR119 Education Center. B. Hess performed a heroic task (fig. 7) is of uncertain origin, although we posit that it in preparing high-quality polished thin sections from might also be a microbial feature. the friable portion of the TR119 drill core. Support was provided by the NASA Astrobiology Institute to H. Xu and S. Lee (award NNA13AA94A; X-ray dif- Conclusions fraction) and by the National Science Foundation to The TR119 paleosol provides exceptional insight J. H. Fournelle (grant EAR-1337156; electron micro- into the weathering characteristics of a gabbroic probe SXFive FE).

REFERENCES CITED

Anderson, R. P.; McMahon, S.; Bold, U.; Macdonald, F. A.; Chadwick, O. A.; Brimhall, G. H.; and Hendricks, D. M. and Briggs, D. E. G. 2017. Paleobiology of the early 1990. From a black box to a grey box—a mass balance Ediacaran Shuurgat Formation, Zavkhan Terrane, south- interpretation of . 3:369– western Mongolia. J. Syst. Palaeontol. 15:947–968. 390. Bazilevskaya, E.; Lebedeva, M.; Pavich, M.; Rother, G.; Donaldson, J. A. 1967. Precambrian vermiform struc- Parkinson, D. Y.; Cole, D.; and Brantley, S. L. 2012. tures: a new interpretation. Can. J. Earth Sci. 4:1273– Where fast weathering creates thin regolith and slow 1276. weathering creates thick regolith. Earth Surface Pro- Dott, R. H., Jr. 2003. The importance of eolian abrasion in cess. Landf. 38:847–858. supermature quartz sandstones and the paradox of Berner, R. A., and Kothavala, Z. 2001. GEOCARBIII: a weathering on vegetation-free landscapes. J. Geol.

revised model of atmospheric CO2 over Phanerozoic 111:387–405. time. Am. J. Sci. 301:182–204. Driese, S. G.; Ludvigson, G. A.; Roberts, J. A.; Fowle, D. A.; Bethke, C. M., and Marshak, S. 1990. Brine migrations González, L. A.; Smith, J. J.; Vulava, V. M.; and McKay, across North America—the plate tectonics of ground- L. D. 2010. Micromorphology and stable-isotope geo- water. Annu. Rev. Earth Planet. Sci. 18:287–315. chemistry of historical pedogenic siderite formed in Birkeland, P. W. 1999. Soils and geomorphology. Oxford, PAH-contaminated alluvial clay soils, Tennessee, USA. Oxford University Press, 430 p. J. . Res. 80:943–954. Brasier, M. D.; Antcliffe, J.; Saunders, M.; and Wacey, D. Driese, S. G.; Medaris, L. G.; Ren, M.; Runkel, A. C.; and 2015. Changing the picture of Earth’s earliest fossils Langford, R. P. 2007. Differentiating pedogenesis from (3.5–1.9 Ga) with new approaches and new discover- diagenesis in early terrestrial paleoweathering sur- ies. Proc. Natl. Acad. Sci. U.S.A. 112:4859–4864. faces formed on granitic composition parent mate- Bullock, P.; Fédoroff, N.; Jungerius, A.; Stoops, G.; Tur- rials. J. Geol. 115:387–406. sina, T.; and Babel, U. 1985. Handbook for soil thin Driscoll, C. T. 1985. Aluminum in acidic surface waters: section description. Wolverhampton, UK, Waine Re- chemistry, transport, and effects. Environ. Health Pers- search, 152 p. pect. 63:93–104.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c). Journal of Geology SUB-CAMBRIAN PALEOSOL IN WISCONSIN 283

Fedo, C. M.; Nesbitt, H. W.; and Young, G. M. 1995. and kinetic considerations. Geochim. Cosmochim. Unraveling the effects of potassium metasomatism Acta 48:1523–1534. in sedimentary rocks and paleosols, with implications Nutman, A. P.; Bennett, V. C.; Friend, C. R. L.; Van for paleoweathering conditions and provenance. Geol- Kranendonk, M. J.; and Chivas, A. R. 2016. Rapid ogy 23:921–924. emergence of life shown by discovery of 3,700-million- Horodyskyj, L. B.; White, T. S.; and Kump, L. R. 2012. year-old microbial structures. Nature 535:535–537. Substantial biologically mediated deple- Odom, I. E. 1975. Feldspar grain-size relations in Cam- tion from the surface of a Middle Cambrian paleosol. brian arenites, Upper Mississippi Valley. J. Sediment. Geology 40:503–506. Petrol. 45:639–650. Jenny, H. 1941. Factors of soil formation: a system of Odom, I. E.; Doe, T. W.; and Dott, R. H., Jr. 1976. Nature quantitative . New York, Dover, 281 p. of feldspar-grain size relations in some quartz-rich ———. 1980. The soil resource: origin and behavior. sandstones. J. Sediment. Petrol. 46:862–870. Berlin, Springer, 377 p. Pavich, M. J.; Leo, G. W.; Obermeier, S. F.; and Estabrook, Knoll, A. H.; Swett, K.; and Mark, J. 1991. Paleobiology J. R. 1989. Investigations of the characteristics, origin, of a Neoproterozoic tidal flat/lagoonal complex: the and residence time of the upland residual mantle of Draken Conglomerate Formation, Spitsbergen. J. Pal- the Piedmont of Fairfax County, Virginia. U.S. Geol. eontol. 65:531–570. Surv. Prof. Pap. 1352, 58 p. Lee, T. R., and Wilkin, R. T. 2010. Iron hydroxy carbonate Peters, S. E., and Gaines, R. R. 2012. Formation of the formation in zerovalent iron permeable reactive bar- “Great Unconformity” as a trigger for the Cambrian riers: characterization and evaluation of phase stability. explosion. Nature 484:363–366. J. Contam. Hydrol. 116:47–57. Pirajno, F. 2009. Hydrothermal processes associated Liu, J.; Hay, R. L.; Deino, A.; and Kyser, T. K. 2003. with meteorite impacts. In Pirajno, F., ed. Hydrother- Age and origin of authigenic K-feldspar in upper- mal processes and mineral systems. Berlin, Springer, most Precambrian rocks in the North American p. 1097–1130. midcontinent. Geol. Soc. Am. Bull. 115:422–433. Riebe, C. S.; Kirchner, J. W.; and Finkel, R. C. 2003. Liu, W.; Liu, C.; Brantley, S. L.; Xu, Z.; Zhao, T.; Liu, T.; Long-term rates of chemical weathering and phys- Yu, C.; et al. 2016. Deep weathering along a granite ridge- ical erosion from cosmogenic nuclides and geochem- line in a subtropical climate. Chem. Geol. 427: 17–34. ical mass balance. Geochim. Cosmochim. Acta 67: Ludvigson, G. A.; González, L. A.; Fowle, D. A.; Roberts, 4411–4427. J. A.; Driese, S. G.; Villarreal, M. A.; Smith, J. J.; Sandler, A.; Teutsch, N.; and Avigad, D. 2012. Sub- and Suarez, M. B. 2013. Paleoclimatic implications Cambrian pedogenesis recorded in weathering profiles and modern process studies of pedogenic siderite. In of the Arabian-Nubian Shield. 59:1305– Driese, S. G., and Nordt, L. C., eds. New frontiers in 1320. and terrestrial paleoclimatology. SEPM Seilacher, A.; Meschede, M.; Bolton, E. W.; and Lu- Spec. Publ. Vol. 104, p. 79–87. ginsland, H. 2000. Precambrian “fossil” Vermiforma Lukens, W. E.; Nordt, L. C.; Stinchcomb, G. E.; Driese, as a tectograph. Geology 28:235–238. S. G., and Barnard, B. Forthcoming. Reconstructing Sheldon, N. D. 2006a. glacial-interglacial pH of paleosols from geochemical proxies. J. Geol. climate cycles in Hawaii. J. Geol. 114:367–376. Medaris, L. G., Jr.; Boerboom, T. J.; Jicha, B. R.; and Singer, ———.2006b. Precambrian paleosols and atmospheric

B. S. 2015. Metasaprolite in the McGrath Gneiss, CO2 levels. Precambrian Res. 147:148–155. Minnesota, USA: viewing Paleoproterozoic weather- Sheldon, N. D.; Retallack, G. J.; and Tanaka, S. 2002. ing through a veil of metamorphism and metasoma- Geochemical climofunctions from North American tism. Precambrian Res. 257:83–93. soils and application to paleosols across the Eocene- Medaris, L. G., Jr.; Driese, S. G.; and Stinchcomb, G. E. Oligocene boundary in Oregon. J. Geol. 110:687–696. 2017. The Paleoproterozoic Baraboo paleosol revisi- Sheldon, N. D., and Tabor, N. J. 2009. Quantitative ted: quantifying mass fluxes of weathering and meta- paleoenvironmental and paleoclimatic reconstruc- somatism, chemical climofunctions, and atmospheric tion using paleosols. Earth-Sci. Rev. 95:1–52.

pCO2 in a chemically heterogeneous protolith. Pre- Slessarev, E. W.; Lin, Y.; Bingham, N. L.; Johnson, J. E.; cambrian Res. 301:179–194. Dai, Y.; Schimel, J. P.; and Chadwick, O. A. 2016. Medaris, L. G., Jr.; Singer, B. S.; Dott, R. H., Jr.; Naymark Water balance creates a threshold in soil pH at the A.; Johnson, C. M.; and Schott, R. C. 2003. Proterozoic global scale. Nature 540:567–569. climate, tectonics, and metamorphism in the southern Soil Survey Staff. 2014. Illustrated guide to soil taxon- Lake Superior region and Proto-North America: evi- omy. Lincoln, NE, National Soil Survey Center, Nat- dence from Baraboo Interval quartzites. J. Geol. 111: ural Resources Conservation Service, US Department 243–257. of Agriculture, 552 p. Nesbitt, H. W., and Young, G. M. 1982. Early Proterozoic Stinchcomb, G. E.; Nordt, L. C.; Driese, S. G.; Lukens, climates and plate motions inferred from major ele- W. E.; Williamson, F. C.; and Tubbs, J. D. 2016. A ment chemistry of lutites. Nature 299:715–717. data-driven spline model designed to predict paleo- ———. 1984. Prediction of some weathering trends of climate using paleosol geochemistry. Am. J. Sci. 316: plutonic and volcanic rocks based on thermodynamic 746–777.

This content downloaded from 144.092.207.008 on July 30, 2018 13:50:27 PM All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).