<<

Sedimentology (2004) 51, 851–884 doi: 10.1111/j.1365-3091.2004.00655.x

Morphology and distribution of soils in the Permo-Pennsylvanian Wichita and Bowie Groups, north-central , USA: implications for western equatorial Pangean palaeoclimate during icehouse–greenhouse transition

NEIL JOHN TABOR* and ISABEL P. MONTAN˜ EZ *Department of Geological Sciences, Southern Methodist University, Dallas, TX 75275, USA (E-mail: [email protected]) Department of Geology, University of California, Davis, CA 95616, USA

ABSTRACT Analysis of stacked Permo-Pennsylvanian palaeosols from north-central Texas documents the influence of palaeolandscape position on pedogenesis in aggradational depositional settings. Palaeosols of the Eastern shelf of the Midland basin exhibit stratigraphic trends in the distribution of soil horizons, structure, rooting density, clay mineralogy and colour that record long-term changes in soil-forming conditions driven by both local processes and regional climate. Palaeosols similar to modern histosols, ultisols, vertisols, inceptisols and entisols, all bearing morphological, mineralogical and chemical characteristics consistent with a tropical, humid climate, represent the Late Pennsylvanian suite of palaeosol orders. Palaeosols similar to modern alfisols, vertisols, inceptisols, aridisols and entisols preserve characteristics indicative of a drier and seasonal tropical climate throughout the Lower strata. The changes in palaeosol morphology are interpreted as being a result of an overall climatic trend from relatively humid and tropical, moist conditions characterized by high rainfall in the Late Pennsylvanian to progressively drier, semi-arid to arid tropical climate characterized by seasonal rainfall in Early Permian time. Based on known Late Palaeozoic palaeogeography and current hypotheses for atmospheric circulation over western equatorial Pangea, the Pennsylvanian palaeosols in this study may be recording a climate that is the result of an orographic control over regional- scale atmospheric circulation. The trend towards a drier climate interpreted from the Permian palaeosols may be recording the breakdown of this pre- existing orographic effect and the onset of a monsoonal atmospheric circulation system over this region. Keywords Late Palaeozoic palaeoclimate, palaeosols, Permo-.

INTRODUCTION ition from icehouse to greenhouse states and a probable global-scale reorganization of atmo- Outcrop and numerically based modelling stud- spheric circulation systems (Rowley et al., 1985; ies of the Late Palaeozoic geological record indi- Cecil, 1990; Parrish, 1993; West et al., 1997; Gibbs cate that Earth’s climate system underwent et al., 2002; and references therein). This plethora significant evolutionary change during this per- of studies indicates that the mid-latitude and iod. It is generally believed that climate change equatorial regions of Pangea became progres- occurred in response to the formation of the sively more arid, with more seasonal preci- supercontinent Pangea, an accompanying trans- pitation, throughout the Late Palaeozoic,

2004 International Association of Sedimentologists 851 852 N. J. Tabor & I. P. Montan˜ez strongly supporting the development of northern results of this study help to refine the nature and hemisphere monsoonal circulation during this timing of climate change in western equatorial interval. Furthermore, it has been hypothesized Pangea during the early stages of supercontinent that monsoonal circulation began in the Late formation, major change in global atmospheric Carboniferous and reached its peak in the Middle circulation and the onset of deglaciation. based primarily upon lithological and palaeontological studies of North American and European terrestrial strata (Parrish, 1993). GEOLOGICAL SETTING Geological proxy records (e.g. coals, evaporites, aeolianites, tillites) and numerical models can The study area in north-central Texas (Fig. 1) was provide an excellent framework for the long-term located in the western coastal zone of equatorial development of climate systems over Pangea. Few Pangea during the Late Palaeozoic (Ziegler et al., detailed records exist, however, that offer insight 11997) and remained between 0 and 5N of the into higher resolution spatial and temporal varia- equator throughout this time period (Golonka tions in climate regimes that accompanied the et al., 1994; Scotese, 1999; Loope et al., 2004). evolution of Pangean monsoonal circulation. Major tectonic elements such as the Muenster Significantly, recent studies of terrestrial deposits Arch and the Wichita, Arbuckle and Ouachita and their associated palaeosols in the south-west Mountains (Fig. 1) had developed by latest USA suggest that monsoonal circulation was well- Pennsylvanian time (Oriel et al., 1967), and tec- established over western equatorial Pangea tonic quiescence has characterized the region by earliest Permian time (Kessler et al., 2001; 2since (Donovan et al., 2001). Late Pennsylvanian Soreghan G.S. et al., 2002; Soreghan M.J. et al., (Virgilian) and Early Permian (Wolfcampian and 2002; Tabor & Montan˜ ez, 2002). These studies, Leonardian) terrestrial strata of north-central coupled with continuing discrepancies in Texas, along with subordinate marine sediments, model-data comparisons for low-latitude Pangea were deposited upon the broad, gently sloping (Patzkowsky et al., 1991; Rees et al., 2001; Gibbs Eastern shelf of the Midland basin as it subsided et al., 2002), highlight the necessity of understand- throughout the Late Palaeozoic (Fig. 1; Brown ing the evolution of specific Pangean climate et al., 1987; Hentz, 1988). A minor component of zones. detrital sediments representing the Upper Penn- The palaeosol-rich, Permo-Pennsylvanian ter- sylvanian and Lower Permian strata of the Eastern restrial strata of north-central Texas hold the Midland basin may have been derived from Early potential to provide insight into climate change Palaeozoic sediments and metamorphic rocks of over low-latitude Pangea, given that climate the distal Ouachita and Arbuckle Mountain variation would have been most pronounced in chains (Hentz, 1988). However, the principal western equatorial Pangea because of the confi- source of detrital sediments during this time guration of the supercontinent and its attendant was probably derived from second-cycle terrigen- effects on atmospheric circulation and precipita- ous-clastic sediments from Middle Pennsylva- tion patterns (Dubiel et al., 1991; Parrish, 1993). nian fan-delta and fluvial facies that were Furthermore, rich faunal and floral collections eroded from the proximal Ouachita foldbelt and interstratified regionally correlatable fluvial and Muenster Highlands (Fig. 1; Brown, 1973). sandstone bodies and fusulinid-bearing lime- Although there is no evidence for a major change stones provide a chronostratigraphic framework in the lithology of source lands during Permo- in which to define spatial and temporal trends in Pennsylvanian deposition, a progressive decrease palaeosol morphology, mineralogy and geochem- upsection in grain size and thickness of fluvial istry and the degree of palaeopedogenic devel- sandstones, coupled with a decrease in strati- opment. This study evaluates the influence of graphic frequency of thick, multistorey sand- local- to regional-scale autogenic and larger scale stones, is interpreted as recording significant allogenic processes on the stratigraphic distribu- denudation of the source areas by the close of tion of palaeosol morphologies and compositions the Early Permian (Hentz, 1988). through detailed field, petrographic, mineralogi- The primarily terrestrial deposits of the Bowie cal and geochemical analyses of 200 Permo- and Wichita Groups on the north-eastern portion Pennsylvanian palaeosols. These palaeosols of the Eastern shelf reach a thickness of up to provide a detailed data set for comparison with 530 m in the east and thin westward towards previously published palaeoclimate trends based marine-dominated strata (Fig. 2). Three terrestrial on geological data and numerical models. The facies are recognized in the Late Pennsylvanian

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 853

Fig. 1. Regional distribution of source terranes and physiographic provinces in the latest Pennsylvanian to Early Permian, Eastern Shelf of the Midland Basin. Province boundaries are defined for the period of maximum regression in the earliest Permian. Structural and tectonic elements shown are for the pre-Virgilian palaeogeography of north- central Texas; the Muenster and Ouachita highlands remained significant topographic features in the Early Permian. Encircled numbers delineate locations of palaeosol-bearing stratigraphic sections used in this study. Diagram modified after Brown et al. (1987) and Hentz (1988). and Early Permian terrestrial strata of the Alluvial floodplain deposits formed primarily in Eastern Midland basin: (1) sand- and gravel-rich association with meandering streams given their channel-bar facies; (2) point-bar facies; and (3) stratigraphic predominance in the western and floodplain facies (Hentz, 1988). Sand and gravel- southern portions of the study area. Based on rich, channel-bar facies were deposited by brai- regional facies distributions, Hentz (1988) defined ded stream systems fringing the north-eastern three physiographic provinces for the latest highlands, whereas the point-bar facies record Pennsylvanian to Early Permian landscape of penecontemporaneous meandering stream sys- the Eastern shelf: (1) the piedmont province tems that developed on the Eastern shelf to the dominated by sand- and gravel-rich channel-bar south and west of the braided stream system. facies and, to a lesser degree, point-bar facies; (2)

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 854 N. J. Tabor & I. P. Montan˜ez

Fig. 2. Regional correlation of Late Pennsylvanian and Early Permian strata in north-central Texas. Light grey beds delineate regionally extensive fluvial sandstone sets; black horizons are coals; white regions represent alluvial mudstones and intercalated siltstones and thin-bedded, fine-grained sandstones; dark grey beds are marine lime- stones. Encircled numbers show the stratigraphic distribution of studied palaeosols – numbers correspond to sites in Fig. 1. Diagram modified after Hentz (1988). the upper coastal plain province characterized by through correlation of multiple fusulinid-bear- point-bar facies; and (3) the lower coastal plain ing limestones and intercalated multistorey flu- province dominated by floodplain facies and, to a vial sandstone deposits, which were described lesser degree, point-bar facies. (ss units in Fig. 2) and regionally correlated by A high-resolution chronostratigraphic frame- Hentz (1988). The Pennsylvanian–Permian work for Permo-Pennsylvanian terrestrial strata boundary (301 ± 2 Ma; Rasbury et al., 1998) of north-central Texas has been developed occurs within a few metres of the contact

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 855

between the Markley and Archer City Forma- ment colours were identified from dry samples tions of the Bowie Group (Dunbar & Committee using Munsell colour charts (Munsell Color, 3on Stratigraphy, 1960). The Wolfcampian–Leo- 1975). Palaeosol classification primarily follows nardian boundary (283 ± 2 Ma; Ross et al., the USDA Soil Taxonomy system (Soil Survey 1994) is defined by the fusulinid-bearing Elk Staff, 1975, 1998). City Limestone at the base of the Petrolia Palaeosol matrix was sampled at 0Æ1–0Æ2m Formation of the Wichita Group (Fig. 2). vertical spacings; rhizoliths and nodules were sampled where present in the profile. Palaeosol matrix samples were disaggregated overnight by BURIAL HISTORY ultrasonic agitation in a dilute Na2CO3 solution and analysed for grain-size distribution. Percent- The entire stratigraphic thickness of the Late age sand, silt and clay was determined by Pennsylvanian and Early Permian rocks in the centrifugation and filtration. Thin sections study area is 1250 m, with the lower 1000 m (n ¼ 78) of representative samples were exam- dominated by sedimentary rocks characteristic ined for their micromorphology and mineralogy of terrestrial depositional facies (Hentz, 1988; according to the approaches and terminology of 4Nelson et al., 2001; Tabor & Montan˜ ez, 2002). Brewer (1976) and Wright (1990). Deposition in this area during the Triassic and X-ray diffraction of the < 2 lm size fraction was Jurassic was insignificant (Oriel et al., 1967; carried out for identification of clay mineralogy. McGowen et al., 1979). However, Samples were exchange saturated with K or Mg strata were probably deposited in the area, as on filter membranes and transferred to glass slides shown by outliers of Cretaceous marine rocks in as oriented aggregates. A split of the Mg-saturated the extreme western portion of the study area, clays was also treated with glycerol. Oriented but these probably did not exceed 330 m in aggregates of all Mg-treated samples were ana- thickness (Barnes et al., 1987). Tertiary and early lysed at 25 C. K-treated samples were analysed at Quaternary burial was insignificant in this area 25 C, 300 C and 500 C after 2 h of initial (Barnes et al., 1987). Therefore, the maximum heating at their respective temperatures. Step burial depth of Virgilian strata in the Eastern scan analyses were performed on a Diano 8500 Midland basin probably did not exceed 1600 m X-ray diffractometer with CuKa radiation be- in thickness. This shallow burial history is tween 2 and 302h with a step size of 0Æ042h reflected by relatively mild burial temperatures and a 1 s count time. that never exceeded 40–45 C in the Early Permian strata along the flanks of the Midland basin (Bein & Land, 1983). PALAEOSOL TYPES AND PALAEOENVIRONMENTAL IMPLICATIONS METHODS Field-scale indicators of pedogenesis in the stud- Thirty-one stratigraphic sections were dug or ied Permo-Pennsylvanian terrestrial strata include picked back 30–60 cm to provide a fresh surface colour, mottling, structure, slickensides, accumu- for description and sampling. The geographic and lation of oxides, clays and carbonate and fossil stratigraphic positions of these 31 sites are shown root traces. Palaeosol profiles were subdivided in Figs 1 and 2. Palaeosols (n ¼ 190) from the 31 into horizons on the basis of downprofile changes stratigraphic sections of the Bowie and Wichita in macro- and micromorphological features, clay Groups were logged and described in detail mineralogy and abundance and geochemical within the chronostratigraphic and sedimentolog- characteristics. Eight major pedotypes (sensu ical framework defined by Hentz (1988). Palaeo- Retallack, 1994; Table 1; Figs 3 and 4) are recog- sol profiles were recognized and described using nized based on distinctive characteristics that established criteria (Retallack, 1988; Kraus & include the aforementioned macromorphological Aslan, 1993; Kraus, 1999). Profile tops were features as well as profile thickness, degree of identified on the basis of a marked change in development (maturity), relationship to strati- grain size and/or colour, as well as preservation of graphically adjacent palaeosols (i.e. composite, primary sedimentary structures. Profile bases compound or cumulate palaeosols; sensu Marriott were delineated at the lowest occurrence of & Wright, 1993; Kraus, 1999) and clay mineral- unaltered parent material. Palaeosol and sedi- ogy. Palaeosols assigned to each pedotype are

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Table 1. Morphology, mineralogy and chemical indices of representative profiles of pedotypes. 856

Pedotype Nodule and stratigraphic Horizon Type profile and rhizolith £ 2 lm Montan P. I. & Tabor J. N. occurrence depth (m)* Macromorphology variation mineralogy Clay mineralogy Province A: Histosol Oi (0–0Æ20) Dark grey to black, massive Organic layers variable – – Lower coastal Markley Fm. and organic-rich claystone. thickness (3–70 cm); Whole Neuropteris leaves. thin clastic partings Non-calcareous. common. Rare jarosite Aj (0Æ20–0Æ27) Light grey (5Y 7/1) claystone replacement of rhizo- Jarosite (40%) K, I (62%) with fine granular structure. liths. Common medium to coarse yellow (5Y 8/8) mottles.

Dense network of thin ˜ez (< 1 cm), tabular yellow (5Y 8/8) rhizoliths. Non- calcareous. 04ItrainlAscaino Sedimentologists, of Association International 2004 C(0Æ27–0Æ35) Grey (5Y 6/1), massive silty – K, I (59%) claystone. Non-calcareous. B: Ultisol ABt1 (0–0Æ15) Red (10R 4/4) claystone with Clay accumulation in – K, I (67%) Lower coastal Markley Fm. granular to subangular upper profiles varies blocky structure. Common from moderate to sig- discrete, fine-scale yellow nificant. (5Y 8/6) vermicular mottles. Abundant Fe- (less so Mn-) oxide coatings and oriented clay skins on matrix aggre- gates. Non-calcareous. Bt2 (0Æ15–0Æ60) Red (10R 4/3) claystone with Haematite, K, I (78%) subangular blocky structure. goethite (10%) Few discrete, fine-scale yellow (10YR 8/8) subspher- ical mottles. Abundant Fe- (less so Mn-) oxide coatings and oriented clay skins. Sedimentology Non-calcareous. BCg(0Æ60–1Æ02) Mottled grey (5Y 6/1) and Haematite, K, I (42%) red (10R 4/4) silty claystone goethite (15%) with subangular blocky structure. Discrete, brown- ,

51, yellow (10YR 6/8) to red (5R 4/8) vertically oriented 851–884 tubular mottles and nodules. Non-calcareous.

04ItrainlAscaino Sedimentologists, of Association International 2004 Table 1. (Continued).

Pedotype Nodule and stratigraphic Horizon Type profile and rhizolith £ 2 lm occurrence depth (m)* Macromorphology variation mineralogy Clay mineralogy Province Cgv (1Æ02–1Æ20) Grey (5/N) massive silty Haematite (60%) K, I (43%) claystone. Large ( 10 cm); horizontal, red (5R 4/8) Fe- and Mn-oxide nodules that coalesce to form a parti- ally indurated polygonal network in plan view (Fig. 5B). Non-calcareous. C: Inceptisol Ag (0–0Æ28) Olive grey (5Y 5/2) Variable horizon Haematite K, I (22%) Lower and upper Markley Fm. mudstone with fine granular development from (10%) coastal; piedmont structure. Abundant fine- to poor to moderate. Ped medium-scale, subspherical structure varies from white (5Y 2Æ5/1) mottles; moderately to well- Sedimentology fine-scale Fe-oxide nodules. developed, and med- palaeosols from palaeoclimate Permo-Carboniferous Non-calcareous. ium to coarse angular and wedge-shaped. Bv (0Æ28–0Æ49) Olive (5Y 4/4) mudstone Haematite K, I (25%) with medium angular blocky (25%)

, structure. Fine- to medium- 51, scale, dusky red (10R 4/4) 851–884 haematite nodules coalesced to form a semi-indurated layer. Non-calcareous. BCssg(0Æ49–1Æ31) Dark olive grey (5Y 3/2) – K, I (32%) mudstone with wedge- to lenticular-shaped structural aggregates defined by slickensides. Common, medium- to coarse-scale yellow (10YR 7/6) to dusky red (10R 4/4) mottles. Non-calcareous. Cg (131-cover) Dark olive grey massive to – K, I (29%) poorly laminated mudstone. Abundant medium-scale, spherical to laminar light reddish brown (5YR 6/4) mottles. Non-calcareous. 857 858

Table 1. (Continued). Montan P. I. & Tabor J. N.

Pedotype Nodule and stratigraphic Horizon Type profile and rhizolith £ 2 lm occurrence depth (m)* Macromorphology variation mineralogy Clay mineralogy Province D: Vertisol ABw (0–0Æ15) Pale red (5R 4/2) mudstone Haematite nodules – I, I/S, K (28%) Upper Markley Fm. with medium rhombohedral may be lacking. coastal; piedmont structure; superimposed fine granular structure. Abundant fine- to medium-scale, light green-grey (8/5 BG) mottles.

Clastic-filled dykes extend ˜ez through horizon. Non-cal- careous.

04ItrainlAscaino Sedimentologists, of Association International 2004 Bgc (0Æ15–0Æ35) Grey (6/N) mudstone with Haematite I/S, I, K (33%) medium rhombohedral (10%) structure. Common medium- scale dusky red (10R 3/4) mottles and haematite nod- ules. Clastic dykes extend to bottom of horizon. Non-cal- careous. Bss (0Æ35–0Æ75) Reddish-brown (5YR 5/4) – I, K (39%) mudstone with wedge- shaped structural aggregates defined by slickensides. Common medium- to coarse- scale, light green-grey mot- tles (8/5 BG). Non-calcar- eous. 2C (0Æ75–1Æ07) Light green-grey (8/10Y) – I, K (36%) poorly laminated mudstone.

Sedimentology Non-calcareous. E: Entisol AC (0–0Æ91) Dusky red (7Æ5R 3/2) massive Variable development – K, I (31%) Lower and upper Markley Fm. to weakly bedded fine- of redoximorphic fea- coastal; piedmont Archer City Fm. grained muddy sandstone. tures including gleyed Nocona Fm. Branching networks of fine- matrix, mottling and/

, Petrolia Fm. scale mottling and thick, or Fe-oxide nodules. 51, Waggoner Ranch Fm. vertical, tubular light grey (7/ Weak development of 851–884 N) mottles. Non-calcareous. slickensides Table 1. (Continued). 04ItrainlAscaino Sedimentologists, of Association International 2004 Pedotype Nodule and stratigraphic Horizon Type profile and rhizolith £ 2 lm occurrence depth (m)* Macromorphology variation mineralogy Clay mineralogy Province Cc (0Æ91–1Æ40) Weak red (5R 3/3) fine- Haematite K, I (29%) grained muddy sandstone. (5%) Common, fine to coarse dus- ky red (7Æ5R 4/3) mottles and haematite nodules. Non-cal- careous. F: Alfisol AB (0–0Æ08) Dark yellow-brown (10YR 6/ Variable thickness – S, K (30%) Lower and upper Archer City Fm. 4) mudstone with subangular profiles (0Æ7 m to 2 m). coastal; pied- Nocona Fm. blocky structure. Few fine- to In some type F palaeo- mont Petrolia Fm. medium-scale, pale red sols, carbonate is not Waggoner (10YR 4/4) vermicular mot- present or occurs over- Ranch Fm. tles. Non-calcareous. lying or superimposed Bt1 (0Æ08–0Æ23) Pale red (10R 6/4) silty clay- on clay-rich horizons – S, K (56%) stone with medium angular (i.e. polygenetic soils). Sedimentology

blocky structure. Common, palaeosols from palaeoclimate Permo-Carboniferous medium- to coarse-scale very pale red (10R4/3) vertical tubular mottles. Abundant clay skins on ped surfaces. ,

51, Non-calcareous. Bt2 (0Æ23–0Æ53) Very pale red (10YR 4/3) – S, K (61%) 851–884 silty claystone with medium angular blocky structure. Common, medium- to coarse-scale reddish-yellow (7Æ5YR 6/6) vertical tubular mottles. Abundant clay skins on ped surfaces. Non-calcar- eous. Btk (0Æ53–0Æ79) Brown-yellow (10YR 5/3) Calcite (10%) S, K, I/S (57%) silty claystone with medium angular blocky structure. Abundant clay skins on ped surfaces. Abundant stage II carbonate nodules. BCss (0Æ79–1Æ06) Red (7Æ5R 4/6) mudstone – S, I/S, K (33%) with wedge-shaped struc- tural aggregates defined by

slickensides. Very weakly 859 calcareous. 860

Table 1. (Continued). Montan P. I. & Tabor J. N.

Pedotype Nodule and stratigraphic Horizon Type profile and rhizolith £ 2 lm occurrence depth (m)* Macromorphology variation mineralogy Clay mineralogy Province C(1Æ06–1Æ76) Dark yellowish-brown (10YR 4/4), – I/S, S, K (33%) massive to weakly laminated mudstone. Non-calcareous. G: Vertisol ABk (0–0Æ27) Reddish-brown (2Æ5YR 4/4) mud- Carbonate devel- Calcite (10%) S, K (33%) Lower and upper Archer City Fm. stone with medium prismatic opment lacking coastal; piedmont Nocona Fm. structure; superimposed medium in some type G

Petrolia Fm. angular blocky structure. Abun- palaeosols. ˜ez Waggoner dant stage II carbonate nodules Ranch Fm. (Fig. 8G). Bkss1 (0Æ27–0Æ57) Reddish-brown (2Æ5YR 4/4) Calcite (15%) S, K (30%) 04ItrainlAscaino Sedimentologists, of Association International 2004 mudstone with wedge-shaped structural aggregates defined by slickensides. Trace reddish-brown (5YR 5/3) spherical mottles. Clastic dykes extend through horizon. Abundant stage II carbonate nodules. Bkss2 (0Æ57–0Æ98) Reddish-brown (2Æ5YR 4/3) Calcite (10%) S, K (30%) mudstone with wedge-shaped structural aggregates defined by slickensides. Clastic dykes extend through horizon. Abundant stage II carbonate nodules. BCss (0Æ98–1Æ40) Reddish-brown (2Æ5YR 4/3) – S, I/S, K (27%) mudstone with wedge-shaped structural aggregates defined by slickensides. Few to abundant medium-scale, reddish-brown Sedimentology (2Æ5YR 5/3) mottles. Clastic dykes extend to base. Very weakly calcareous. C(1Æ40–2Æ11) Dusky red (2Æ5YR 3/2) poorly – S, I/S (23%) laminated mudstone. Few medium

, grey (6/5BG) spherical mottles. 51, Very weakly calcareous. 851–884 Table 1. (Continued). 04ItrainlAscaino Sedimentologists, of Association International 2004 Pedotype Nodule and stratigraphic Horizon Type profile and rhizolith £ 2 lm occurrence depth (m)* Macromorphology variation mineralogy Clay mineralogy Province H: Aridisol Bkm (0–0Æ22) Reddish-grey (2Æ5YR 6/1) Degree of carbonate Calcite (95%) I, K Lower and upper Archer City Fm. mudstone; stage III carbon- development varies; coastal; piedmont Nocona Fm. ates. morphologies can be Petrolia Fm. Bk1 (0Æ22–0Æ39) Yellow (10YR 7/6) mudstone nodular to tubiform Calcite (35%) I, K (22%) Waggoner with medium subangular (e.g. Fig. 8E). Matrix Ranch Fm. blocky structure. Abundant structure varies be- stage II carbonate nodules. tween prismatic, angu- Bk2 (0Æ39–0Æ58) Pale red (5R 4/3) mudstone lar blocky and Calcite (10%) I, I/S, K (25%) with medium to coarse subangular blocky. angular blocky structure. Some type H palaeo- Abundant stage II carbonate sols in Waggoner nodules. Ranch Fm. have slick- 2Bss (0Æ58–0Æ76) Pale red (10R 5/3) sandy enside planes filled by – I, I/S, K (22%) mudstone with wedge- calcium carbonate at Sedimentology

shaped aggregate structure depth. palaeosols from palaeoclimate Permo-Carboniferous defined by weakly developed slickensides. Calcareous. 3Cs (0Æ76–1Æ30) Reddish-grey (10R 5/1) san- – I, I/S, K (25%) dy mudstone with wedge- ,

51, shaped aggregate structure defined by weakly developed 851–884 slickensides. Few fine- to medium-scale grey-brown (10YR 5/2) subspherical mottles. Very weakly calcareous. 4C (1Æ30–1Æ60) Grey-brown (10YR 5/2) –– massive to weakly bedded fine-grained sandstone. Moderate development of fine- to medium-scale yellow-brown (10YR 5/6) vermicular mottles. Very weakly calcareous.

* Down-profile depth from the interpreted palaeosurface. Percentage of the matrix represented by Fe-oxide, jarosite or carbonate nodules and rhizoliths. Clay mineralogy presented in relative order of abundance in < 2 lm size fraction; numbers in parentheses are percentage of matrix represented by this size

fraction: K ¼ kaolinite, I ¼ illite, I/S ¼ interlayered illite–smectite, S ¼ smectite. Clays referred to in this study as illite are formally classified as mica-like 861 minerals or sericite. 862 N. J. Tabor & I. P. Montan˜ez

Fig. 3. Key of symbols and horizon nomenclature used in Figs 4, 7 and 9–13.

interpreted as being the product of similar soil- Type A palaeosols forming environments in which local and regional processes combined to produce a distinctive Description profile. Pedotypes are thus analogous to a soil The type profile of pedotype A is composed of series in a modern landscape (Retallack, 1990; three non-calcareous layers that define a thin Bestland et al., 1997). profile (0Æ4 m; Figs 3 and 4A, Table 1). The In the following section, the pedotypes are uppermost dark grey to black, massive layer is described based, in part, on individual measured organic rich and dominates the profile (Fig. 5A). profiles judged to be most representative of all The underlying light grey claystone layer exhibits palaeosols assigned to a given group. Major angular structure, is composed of quartz silt variations in morphology or composition from and low cation-exchangeable clays (kaolinite the type profile are discussed and summarized in and subordinate amounts of illite; Table 1) and Table 1. The pedotypes are related to the USDA lacks labile minerals (biotite, muscovite, feld- soil taxonomy classification (Soil Survey Staff, spars) present in the unaltered parent material. 1975, 1998), the spatial and stratigraphic distri- The basal grey claystone layer contains a bution of the associated palaeosols delineated small amount of labile minerals and lacks clear and the palaeoenvironmental implications pedogenic features. Redoximorphic features assessed. (e.g. Vepraskas, 1994) are abundant in type A

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 863

Fig. 4. Measured sections of palaeosol profiles representative of each of the eight pedotypes. (A) Type A palaeosol from the Late Pennsylvanian Markley Fm., site 1 (Fig. 1). (B) Type B palaeosol, Markley Fm., site 1. (C) Type C palaeosol, Markley Fm., site 4. (D) Type D palaeosol, Markley Fm., site 5. (E) Type E palaeosol, Markley Fm., site 6. (F) Type F palaeosol from the Archer City Fm., site 11. (G) Type G palaeosol from the Waggoner Ranch Fm., site 26. (H) Type H palaeosol, Waggoner Ranch Fm., site 28.

palaeosols and include overall low chroma matrix Interpretation and classification colours (gleying), tabular networks of downward- Laterally discontinuous, organic matter-rich bifurcating rhizoliths and medium to coarse layers in type A palaeosols are interpreted to be yellow mottling. Rhizoliths in the type palaeosol surficial accumulations (O horizons) that formed are replaced by jarosite. by in situ accumulation of plant material (Figs 4A Type A palaeosols are developed within silty and 5A). Angular blocky structure in the under- mudstones of the floodplain facies; all type A lying claystone layer is interpreted as being ped palaeosols are laterally discontinuous over tens to structure that formed by differential shearing of hundreds of metres (e.g. see sections 1 and 2 in plastic materials (clays) in the profile. The gleyed Fig. 6). These palaeosols occur as thin (averaging matrix, lack of weatherable labile minerals, low <0Æ5 m) composite profiles (sensu Marriott & organic content and abundance of kaolinite in the Wright, 1993; Kraus, 1999) with common thin £ 2 lm size fraction suggest that the claystones clastic partings. They are stratigraphically restric- directly underlying O horizons formed as eluvial ted to sites 1–3 in the lower half of the Upper A horizons in response to intense hydrolysis or Pennsylvanian Markley Formation (below ss11 in acidolysis in poorly drained profiles (Fig. 4A; cf. Fig. 2). 5van Breeman & Harmsen, 1975).

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 864 N. J. Tabor & I. P. Montan˜ez

Fig. 5. Macromorphologies of Permo-Pennsylvanian palaeosols. (A) Couplets of dark–light banding are type A palaeosols; dark layers are O horizons, light layers are A horizons. Approximately 1Æ7 m thick dark horizon at base of outcrop is a type B palaeosol. Palaeosols are from the Upper Pennsylvanian Markley Fm., site 2. (B) Haematite redox concentrations, oriented into a polygonal pattern (1Æ5 m across), in the Cgv horizon of a type B palaeosol, Markley Fm., site 1. (C) Slickensides in a type G palaeosol, Lower Permian Nocona Fm., locality 15. Knife (0Æ25 m long) for scale; blade points in the direction of mineral orientation along the slickensurface. (D) Type G palaeosol showing pedogenic slickensides in cross-section, Lower Permian Waggoner Ranch Fm., site 26. White ‘blebs’ are pedogenic carbonate nodules. Field of view is 1Æ5 m. (E) Calcareous rhizoliths from a type H palaeosol, Waggoner Ranch Fm., site 31. 35 mm lens cap for scale. (F) Type G palaeosol, upper Waggoner Ranch Fm., site 31. Overlying light-coloured beds are marine calcareous silts and limestones of the Maybelle Limestone.

Type A palaeosols are interpreted as being conditions and anoxic pore waters that promote histosols (Soil Survey Staff, 1998) given the in situ accumulation of organic matter. The relative thickness and concentration of organic predominance of gley matrix colours and low matter in the O horizons of the palaeosol profiles chroma mottles and root traces in type A palae- (Fig. 4A). Modern histosols form in low-lying, osols, as well as the paucity of discrete Fe-oxide waterlogged regions characterized by always wet nodules, supports prolonged saturation of these

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 865

Fig. 6. Palaeosol-bearing sections of Upper Pennsylvanian Markley For- mation. Palaeosol tops are delinea- ted by wavy ‘lines’ to the left of each measured section. Capital letters refer to palaeosols of a given pedo- type; circles delineate pedotype profiles discussed in the text. Sec- tions 1 and 2 are from two outcrops separated by 8Æ5 km; correlation be- tween outcrops based on sandstone set 10 (ss10) in the middle of the Markley Fm. Measured section 4 underlies sandstone set 12 (ss12) in the upper portion of the Markley Fm. Measured section 5 underlies ss14 from the uppermost Markley Fm., proximal to the Pennsylva- nian–Permian boundary. See Fig. 2 for stratigraphic position of sand- stone sets. profiles (Simonson & Boersma, 1972; Wilding Late Pennsylvanian type A palaeosols are et al., 1983; Schwertmann, 1985). The shallow interpreted as having formed in poorly drained, and tabular distribution of jarosite-replaced roots shallow depressions on the regionally extensive in fossil A horizons further reflects the poorly coastal plain given their morphological charac- drained palaeosoil conditions (cf. Retallack, 1988, teristics and laterally discontinuous nature. The 1990). Jarosite, however, may represent post- predominance of stacked, thin, composite profiles pedogenic oxidation of Fe-sulphides despite its (Fig. 6) reflects slow, episodic sediment accumu- neomorphic replacement of roots (cf. Kraus, lation in geomorphic lows on the latest Pennsyl- 1998). vanian coastal plain of western equatorial Pangea.

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 866 N. J. Tabor & I. P. Montan˜ez

massive grey silty claystone (Fig. 4B; Table 1). Redoximorphic features are present throughout the profile as redox depletions (i.e. low chroma mottles and matrix colours) and as redox concen- trations (i.e. Fe- and, less commonly, Mn-oxide coatings on matrix aggregates and Fe-oxide nod- ules). Oblate Fe-oxide nodules, ranging from 0Æ05 to 0Æ2 m in length, are laterally oriented and form a partially indurated horizon at the base of the type profile (Fig. 5B). Type B palaeosols with less well-differentiated horizons commonly lack thick accumulations of Fe-oxides. Kaolinite and trace amounts of illite dominate the £ 2 lm size fraction of all claystones; labile miner- als are generally lacking (Fig. 7A). The abundance of the clay-dominated, £ 2 lm size fraction decrea- ses downprofile (e.g. from 78% to 43%; Table 1). Type B palaeosols occur primarily as thin to moderately thick (< 1Æ5 m), strongly developed profiles within mudstones of the floodplain facies (Figs 5A and 6). Type B palaeosols are strati- graphically associated with type A palaeosols. Type B palaeosols are primarily developed upon mudstones and are stratigraphically limited to the lower half of the Upper Pennsylvanian Markley Formation (below ss11 in Fig. 2).

Interpretation and classification Redoximorphic colouration and the abundance of moderate to large Fe-oxide nodules in the lower portion of profiles (BCg and Cgv horizons in Fig. 4B) indicate that type B palaeosols were waterlogged for prolonged periods (i.e. hydro- morphic soils), given that the degree of chemical weathering of parent material and the size and Fig. 7. XRD patterns of clay minerals from the £ 2 lm abundance of Fe-oxide nodules commonly in- size fraction in select type palaeosols. Clay mineralogy crease with increased duration of waterlogging of: (A) a Bt horizon; type B palaeosol, Upper Pennsyl- (Simonson & Boersma, 1972; Sobecki & Wilding, vanian Markley Fm.; (B) Bt horizon, type G palaeosol, 1983; Retallack, 1990). However, the predomin- Lower Permian (lower Leonardian) Petrolia Fm.; (C) Bk horizon, type H palaeosol, Lower Permian (mid- ance of kaolinite in the clay fraction indicates that Leonardian) Waggoner Ranch Fm. K ¼ kaolinite; these hydromorphic soils were characterized by I ¼ illite; S ¼ smectite; C ¼ hydroxy-interlayered episodic or seasonal changes in soil moisture minerals; Q ¼ quartz. conditions (Wilding et al., 1983; Schwertmann, 1985; Dixon & Skinner, 1992). Clay-rich horizons in type B palaeosols are Type B palaeosols interpreted as being zones of clay accumulation Description or argillic horizons (ABt1 and Bt2 in Fig. 4B) The type profile of pedotype B (Figs 4B and 6) is that formed by downward translocation of clay composed of four distinct units: (1 and 2) two minerals (i.e. illuviation). Concentration of clay upper, red claystones that exhibit fine angular to minerals, in particular kaolinite, in the inferred subangular blocky structure and oriented clay Bt horizon and the presence of continuous, orien- and Fe-oxide coatings along matrix aggregates ted clay coatings (argillans of Brewer, 1976) on and lining channel voids; (3) an underlying, peds and lining channel voids suggest extensive mottled red/grey silty claystone that exhibits pedogenic kaolinite formation. The presence of subangular blocky structure; and (4) a basal, Fe-oxide (and less commonly Mn-oxide;

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 867 ferri-mangans of Brewer, 1976) lining peds and to reddish-brown mottles and Fe-oxide nodules) voids, in addition to the downprofile increase in and reductions (gleyed matrix colours) occur Fe-oxide nodules, records downward transloca- throughout type C palaeosols (Fig. 8A). A semi- tion of Fe and Mn in addition to clay minerals. indurated layer of coalesced submillimetre-sized Translocation and mineral accumulation also indi- haematite nodules (Fig. 8D) occurs near (28–49 cm cate that type B hydromorphic palaeosols under- depth) the interpreted palaeosurface of the type went periods of free drainage (Birkeland, 1999). profile (Fig. 4C); analogous Fe-oxide layers are The haematite- and goethite-cemented layers at lacking in some type C palaeosols. The < 2 lm size the base of type B palaeosols are analogous to fraction contains a moderate amount of clay modern plinthites that form in soil horizons minerals (22–32%) that are dominated by illite characterized by a seasonally fluctuating water and subordinate amounts of kaolinite and hydro- table (Duchaufour, 1982; Soil Survey Staff, 1998). xy-interlayered mineral (HIM) clays. The thickness of the palaeoplinthite horizon Type C palaeosols occur as thick compound ( 1 m) and its depth below the palaeosol surface and composite profiles (Fig. 6) developed in suggest that the local palaeowater table probably intercalated thinly bedded mudstones, siltstones fluctuated within 1–2 m of the land surface and fine-grained sandstones. They are strati- (cf. Duchaufour, 1982). graphically restricted to overbank mudstone Type B palaeosols are interpreted as ultisols deposits associated with channel and crevasse- based on morphological evidence for argillic hori- splay sandstone deposits in the Upper Pennsyl- zons, prolonged saturation and extensive hydro- vanian Markley Formation above fluviogenic lysis interrupted by episodes of free drainage. cycle ss11 (sites 4 and 5; Figs 2 and 6). Modern ultisols form in warm, humid environ- ments (e.g. modern tropical rainforests) character- Interpretation and classification ized by high rates of chemical weathering. These The laterally continuous, semi-indurated Fe-oxide soils minimally require periodic drainage in layer in the upper part of the type profile is order to form argillic horizons through illuviation interpreted as being a palaeoplinthite (Bv horizon (Retallack & German-Heins, 1994; Gill & Yemane, in Fig. 4C). The development of a plinthite proxi- 1996). Although Late Pennsylvanian ultisols mal to the palaeosol surface records a shallow, formed penecontemporaneously with type A pala- fluctuating palaeogroundwater table (Table 1; cf. eosols, ultisols are interpreted as having formed on Duchaufour, 1982). Further evidence for fluctu- stable landscapes such as interchannel highs that ating soil moisture conditions is the occurrence may have developed during minor episodes of of large-scale wedge- to lenticular-shaped peds fluvial downcutting. This interpretation is based defined by slickensides in basal mudstones. on evidence for their strong pedogenic develop- Pedogenic slickensides form in clay-rich modern ment and for their having been at least episodically soils as a result of soil mass movement and struc- well-drained in the upper argillic horizons of tural modification driven by wetting and drying the profile (cf. Markewich & Pavich, 1991; cycles (Dudal & Eswaran, 1988; Wilding & Tessier, Bestland et al., 1997), and on the inferred low 1988). Limited development of large-scale peds sedimentation rates recorded by their thin, and slickensides in type C palaeosols, however, composite profiles (cf. Marriott & Wright, 1993). suggests modest fluctuations in soil moisture content in the lower portions of the profile and/ or profile immaturity (Brewer, 1976; Jim, 1990), Type C palaeosols given that these features form rapidly in soils Description characterized by repeated wetting and drying The type profile of pedotype C consists of four (Yaalon & Kalmar, 1978; Birkeland, 1999). poorly to moderately developed horizons (Table 1) Type C palaeosols are classified as inceptisols that exhibit fine to medium angular blocky struc- given their moderate development of ped structure ture. Larger scale wedge- to lenticular-shaped and horizonation. These probably immature pro- structural units are defined by randomly oriented files formed in floodplain deposits proximal to slickensides that overprint the finer scale angular fluvial channels, as evidenced by their close lateral blocky structure in the lower mudstone (BCssg in and stratigraphic proximity to channel-filling and Fig. 4C). Although all type C palaeosols exhibit crevasse-splay sandstones; their compound and some structure that is indicative of pedogenic devel- cumulate profiles record steady, and sometimes opment, most profiles only have weak horizonation. rapid, sedimentation in these environments (cf. Abundant redoximorphic concentrations (yellow Marriott & Wright, 1993; Kraus, 1999).

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 868 N. J. Tabor & I. P. Montan˜ez

Fig. 8. Photomicrographs of redoximorphic features in Upper Pennsylvanian palaeosols. (A) Haematite nodules (e.g. dark spots in area marked by white ovoids) and ferrans (dark area) in a claystone (Bv horizon) with angular blocky structure; type C palaeosol. Scale bar ¼ 1 mm. (B) Redoximorphic depletions (light grey) and concentrations (dark grey) in an AC horizon; type E palaeosol. Scale bar ¼ 1 mm. (C) Redoximorphic depletions (light grey) and con- centrations (black) of both diffuse and nodular haematite in a Cc horizon; type E palaeosol. Scale bar ¼ 1 mm. (D) A plinthite horizon altered to an ironstone; the ironstone is composed predominantly of haematite nodules (black) and a minor amount of clay minerals; type C palaeosol. Scale bar ¼ 1 mm.

fraction in the type profile is dominated by Type D palaeosols hydroxy-interlayered minerals and smectite and Description lesser amounts of kaolinite (Table 1). The type profile of pedotype D is composed of Type D palaeosols are developed within mud- four red to grey mudstone units that exhibit stones of the sand- and gravel-rich channel-bar abundant redoximorphic features including small facies (Fig. 6) of the uppermost Upper Pennsyl- ( 10 mm) haematite nodules, manganese oxide vanian Markley Formation (above ss13 in Fig. 2). coatings and low chroma soil matrix (or gley) Their distribution is limited to the eastern portion colours (Fig. 4D; Table 1). The top two units of the study area (sites 5 and 6 in Fig. 1). exhibit medium-scale (50–200 mm long) rhombo- hedral structural aggregates; the uppermost clay- Interpretation and classification stone exhibits superimposed angular blocky Well-developed, medium-scale rhombohedral- structure defined by clay-lined microslicken- and wedge-shaped structure and associated ped- sides. Clay-rich mudstones in the lower half of ogenic slickensides in type D palaeosols are the profile exhibit wedge-shaped structural aggre- analogous to ‘parallelepipeds’ in Holocene verti- gates defined by larger scale slickensides. Clay sols (Dudal & Eswaran, 1988). In modern soils, and Mn-oxides typically coat slickensides. Sand- vertic features develop during mass movement filled V-shaped dykes (Figs 3D and 6) cross-cut and shearing in smectite-rich soils that become the upper half of the type profile. The < 2 lm size saturated during seasonal rainfall (Yaalon &

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 869

Fig. 9. Palaeosol-bearing sections of Lower Permian Archer City Formation. Palaeosol tops are delineated by the wavy lines to the left of each measured section. Capital letters refer to palaeosols of a given pedotype; circle delineates pedotype profile discussed in the text. Correlation of three sites (sites 9, 10 and 11 in Fig. 2) based on sandstone set 8 (ss8). Sites 9 and 10 are sep- arated by 0Æ3 km, and sites 10 and 11 are separated by 0Æ2 km. See Fig. 3 for key to symbols.

stress cutans (Jim, 1990) due to differential wetting, and shrinking and swelling forces driven by repeated wetting and drying of the soil matrix. Decreasing soil moisture and shrinking of soil clays resulted in cracking of the soil matrix during drier intervals. The V-shaped dykes that cross-cut the upper portions of the profiles resulted from back-filling of these surficial cracks from sediment-laden water during subsequent flood events. Type D palaeosols are classified as vertisols based on abundant and well-developed vertic features, their clay content (28–39%) with a significant percentage of smectite and the pres- ence of deep V-shaped cracks developed within the upper horizons (Soil Survey Staff, 1998). Modern vertisols form in seasonally moist and typically tropical to warm-temperate climates, with four to eight dry months yearly (Dudal & Eswaran, 1988; Buol et al., 1997). The relatively low kaolinite content and preservation of expand- able 2:1 layer-lattice clays (smectite) in type D palaeosols coupled with the occurrence of clastic dykes indicate periods of drying that were suffi- ciently long to preserve expandable clay minerals and to allow for significant shrinkage of the clay- rich matrix (Duchaufour, 1982; Retallack, 1990). However, the presence of gley matrix colours and common to abundant redox depletions requires that periods of palaeosoil moisture deficiency in type D palaeosols were temporally limited 6(Daniels et al., 1971; Duchaufour, 1982; Buol et al., 1997). Based on their stratigraphic associ- ation with overbank mudstones as well as the morphological and chemical characteristics of these palaeovertisols, they are interpreted as having formed in interfluve muds on the Late Pennsylvanian upper coastal plain and piedmont.

Type E palaeosols Description Kalmar, 1978; Dudal & Eswaran, 1988; Wilding & The representative profile of pedotype E con- Tessier, 1988). The thin clay coatings on super- sists of two units that are very weakly devel- imposed angular blocky peds probably formed as oped in muddy fine-grained sandstone and are

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 870 N. J. Tabor & I. P. Montan˜ez

Fig. 10. Palaeosol-bearing sections of Lower Permian Nocona Formation. Palaeosol tops are delineated by the wavy lines to the left of each measured section. Capital letters refer to palaeosols of a given pedotype. Corre- lation of the siltstone- and claystone-dominated section from site 16 with the sandstone-dominated section at site 15 (80 km apart) based on contemporaneity of Elk City Limestone and ss11 defined by Hentz (1988). Key to symbols as in Fig. 3.

Fig. 11. Palaeosol-bearing sections of Lower Permian Petrolia Formation at sites 17 and 18. Palaeosol tops are delineated by the wavy lines to the left of each meas- ured section. Capital letters refer to palaeosols of a given pedotype. Key to symbols as in Fig. 3.

mottles (Figs 6 and 8B); superposition of different generations of mottles is observed. The lower pale red unit contains fine to coarse dark red mottles and haematite nodules (Fig. 8C). The < 2 lm size fraction in the profile is dominated by kaolinite and illite (Table 1). In addition, some type E palaeosols also exhibit weakly developed slic- kensides. Type E palaeosols occur as thick (1–2 m) cumulate and compound and composite palaeo- distinguished by their colour and nature of the sols throughout the Permo-Pennsylvanian terrest- redoximorphic features (Fig. 4E; Table 1). The rial strata in the study area (sites 1–3, 5, 6, 9–11, upper red unit exhibits a vertically oriented, 12–15, 19–22, 24, 25 and 29 in Fig. 1). They are branching network of light-grey, fine- (< 10 mm developed within or stratigraphically adjacent to wide) to coarse-scale (10–25 mm wide) tubular finely laminated to slightly bioturbated silty

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 871

Fig. 12. Correlated transect (67 km long) of four palaeosol-bearing sec- tions of the Lower Permian Petrolia Formation beneath the regionally traceable Beaverburk Limestone. The two sections (site 22) on the left are siltstone and claystone domin- ated, whereas sections from sites 20 and 21 contain abundant sandstones of the point-bar facies. Palaeosol tops are delineated by the wavy lines to the left of each measured section. Capital letters refer to pal- aeosols of a given pedotype. Key to symbols as in Fig. 3. mudstones and siltstones and fine- to medium- depletions and concentrations in many type E grained sandstones of the channel bar and point palaeosols indicates a certain degree of hydro- bar facies (Figs 6, 9, 10, 12 and 13). morphy (waterlogging; cf. Simonson & Boersma, 1972; Fanning & Fanning, 1989). Interpretation and classification In contrast, type E palaeosols that formed in Type E palaeosols are interpreted as entisols sandy siltstones typically lack redoximorphic based on the very weak development of pedogen- features. Given that many type E palaeosols ic structure and horizonation (Soil Survey Staff, commonly occur as weakly developed, com- 1998). Modern entisols exhibit little to no devel- pound or cumulate profiles in proximal positions opment of horizonation, and typically form upon to fluvial siltstones and sandstones and estuarine unstable, wet landscapes (Buol et al., 1997). siltstones and mudstones (south-western portion Entisols record stunted horizon development of study area), they are interpreted as having resulting from prolonged saturation and/or lim- formed on unstable landscapes proximal to ited periods of pedogenesis (tens to hundreds of fluvial channels (e.g. levees and crevasse splays) years) caused by rapid sedimentation rates (Buol on the coastal plain and piedmont and wetlands et al., 1997). The occurrence of redoximorphic of the lower coastal plain.

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 872 N. J. Tabor & I. P. Montan˜ez

Fig. 13. Correlated transect (31 km long) of three palaeosol-bearing sections (from sites 27, 29 and 30) of the Lower Permian Waggoner Ranch Formation beneath the Maybelle Limestone. Palaeosol tops are delineated by the wavy lines to the left of each measured section. Capital letters refer to palaeosols of a given pedotype; circle delineates pedotype profile discussed in text. Key to symbols as in Fig. 3.

subangular blocky structure. Three significantly Type F palaeosols more clay-rich layers (56–61% clay vs. 30–33% in Description other units) underlie the uppermost mudstone. The type profile of pedotype F is composed of six These are red to brown, silty claystones that well-developed horizons (Fig. 4F; Table 1) that exhibit fine to medium subangular blocky struc- are distinguished by their clay content, structure, ture. Thick, continuous and oriented clay coat- mottling and mineral accumulations. The upper- ings (argillans) occur around detrital grains and most horizon is defined by a dark yellow-brown, line aggregates in all three claystones. The upper non-calcareous mudstone with fine to medium two claystone layers are non-calcareous, whereas

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 873 the lowermost claystone contains discrete, cm- mann, 1985, e.g. Kraus & Aslan, 1993). The dark scale stage II carbonate nodules (sensu Machette, yellowish-brown mudstones with fine-scale 1985) and weakly developed calcareous rhizo- branching root structures that cap most type F liths (Figs 9–12) that exhibit pedogenic fabrics profiles are interpreted to be organic-rich AB (cf. Deutz et al., 2002). Faint tubiform red and horizons. reddish to yellow mottles in the mudstone and The presence of well-differentiated profiles claystone layers define a pattern similar to with argillic horizons suggests that type F palaeo- branching root structures (cf. Retallack, 1988, sols were alfisols (Soil Survey Staff, 1998). 1990) and probably record deeply penetrating Modern alfisols form in humid to subhumid, root systems. Smectite dominates the < 2 lm size mid-latitude to subtropical regions that have fraction of the claystone layers, with trace > 75–100 cm of annual precipitation (Strahler & amounts of kaolinite and illite (Table 1). The Strahler, 1983; Buol et al., 1997). Given that lower two horizons in the type profiles are pedogenic carbonate typically accumulates in red and yellowish-brown mudstones (Table 1); soils receiving < 70 cm mean annual precipitation the upper horizon exhibits wedge-shaped struc- in low-latitude regions (McFadden, 1988; tural aggregates defined by randomly oriented Birkeland, 1999; Royer, 1999), carbonate-bearing slickensides, whereas the lower horizon pre- type F palaeosols may record formation at the serves disturbed primary sedimentary lamina- subhumid to semi-arid climatic boundary. The tions (Fig. 4). predominance of smectite and trace amounts of Some variants of the type profile lack the kaolinite in type F palaeosols is consistent with a horizon of stage II carbonate nodules. Moreover, subhumid to semi-arid palaeoclimate. In addition, some type F palaeosols contain carbonate-rich some type F palaeosols are characterized by the horizons that overlie or are superimposed on the development of pedogenic carbonates above or clay-rich horizons with clay skins (cutans). These within argillic horizons (polygenetic palaeosols). dominantly strongly developed palaeosols may This superposition of calcic-over-argillic horizons occur as composite profiles and are found in is climatically ‘out of phase’ and suggests that Lower Permian mudstones of the floodplain these type F palaeosols formed under two dis- facies (Figs 9–12) throughout the lower and upper tinctly different climatic regimes: an earlier, wet- coastal plain and piedmont physiographic prov- ter climate that leached carbonate and formed inces (sites 9, 16 and 19 in Fig. 1). Given that argillic horizons in the profile and a subsequent these palaeosols are developed in fine-grained drier climate that accumulated carbonate and mudstones and require good drainage, these formed calcic horizons in the profile (cf. Soil palaeosols probably formed upon areas of the Survey Staff, 1975; Buol et al., 1997). These floodplains that were removed from frequent polygenetic alfisols are stratigraphically limited flooding and depositional events associated with to the upper Archer City and Nocona Formations major stream systems (e.g. Buol et al., 1997). as well as the uppermost Waggoner Ranch For- mation between the Maybelle and Lake Kemp Interpretation and classification Limestones (Fig. 2). The clay enrichment, abundant continuous argil- The development of multiple, well-defined lans on detrital grains and peds and the non- horizons, which commonly include superim- calcareous matrix in the upper portions of type F posed argillic horizons, and the composite nature profiles are evidence of clay translocation. Trans- of these palaeosols record prolonged periods (tens location is facilitated through leaching of Ca2+ to hundreds of ky) of pedogenesis (cf. Bestland from clay exchange sites leading to the develop- et al., 1997). These characteristics coupled with ment of argillic (Bt) horizons (Franzmeier et al., evidence for well-drained profiles suggest that 1985; Birkeland, 1999). Pedogenic carbonate can type F palaeosols developed upon stable land- precipitate in lower horizons of the profile if scapes distributed throughout the Early Permian soluble minerals are incompletely leached from Eastern shelf for which the local groundwater the soil horizon (Buol et al., 1997). The occur- table was relatively deep. Evidence of large, deep rence of pedogenic carbonates in lower portions fossil roots in type F palaeosols in conjunction of type F palaeosols, coupled with limited with the presence of well-developed argillic development of redoximorphic features and illu- horizons and blocky ped structure suggest that viated clays deep within the profile, indicate that woodlands may have developed on such stable these palaeosols formed well above the palaeo- areas of the Early Permian alluvial landscape (cf. water table and were well-drained (cf. Schwert- Mack, 1992).

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 874 N. J. Tabor & I. P. Montan˜ez

defined by weakly developed slickensides in the Type G palaeosols lower half of the profile. Mottling is limited to the Description lowermost horizons and typically exhibits high The type G profile is composed of five well- chroma. A thick, indurated layer (‡ 150 mm) of developed mudstone horizons (Fig. 4G; Table 1) stage III carbonate accumulation (¼ petrocalcic that exhibit wedge-shaped structure defined by horizon; Soil Survey Staff, 1975, 1998; Machette, clay-coated slickensides (middle three horizons; 1985) near the top of the profile and extensive Fig. 5C and D) and contain stage II carbonate development of stage II carbonate nodules and nodules and rhizoliths (upper three horizons; calcareous rhizoliths in lower horizons of the Fig. 5F), although a few other type G palaeosols profile (Fig. 5H) help to distinguish the type H in the upper Waggoner Ranch Formation may palaeosol and its pedotypes from all other car- exhibit 50–100 mm thick stage III carbonate bonate-bearing palaeosols. Hydroxy-interlayered accumulations near the bases of the profiles. minerals and mica-like minerals with trace Redoximorphic features are generally lacking amounts of kaolinite and chlorite make up the with red hues dominating matrix colours. Sand- <2 lm size fraction in the type profile (Fig. 7C). filled clastic dykes extend throughout most of The type profile has lower clay contents (< 2 lm the profile (to a depth of 1Æ4 m). The < 2 lm size size fraction in matrix of 22–25%) than all other fraction in the type profile is dominated by pedotypes. smectite and lesser amounts of kaolinite and Type H palaeosols may occur as compound, mica-like minerals (sericite; Table 1; Fig. 7B). cumulate and composite profiles developed in Type G palaeosols developed in mudstones of Lower Permian mudstones of the floodplain the floodplain facies throughout the studied facies (Figs 9–13). Their occurrence at many Lower Permian strata (Figs 9, 10 and 12). Their sites (sites 7–9, 16–22 and 24–31 in Fig. 1) development at numerous sites (sites 7–9, 11–14, reflects their widespread distribution in the 16, 18, 23, 24 and 26–31 in Fig. 1) reflects their study area. widespread distribution in the study area. Interpretation and classification Interpretation and classification Type H palaeosols are interpreted as having been Analogous to type D palaeosols, type G palaeosols similar to modern aridisols that formed under are interpreted as vertisols based on their semi-arid to arid climate conditions based on well-developed vertic features, their smectite- their extensive carbonate accumulation, in par- dominated clay composition and the presence of ticular as petrocalcic horizons, and relatively low V-shaped cracks developed to significant profile clay contents in comparison with all other ped- depths. Type G palaeosols, however, exhibit two otypes (Soil Survey Staff, 1998). In reality, these major differences from type D palaeosols: (1) palaeosols may only be classified as inceptisols accumulation of pedogenic carbonate; and (2) a in the US Soil Taxonomy, as climate conditions paucity of redoximorphic features. The lack of are not clearly understood during Early Permian redoximorphic features in type G palaeosols time (Soil Survey Staff, 1998). Nonetheless, the indicates that these well-drained profiles formed presence of chlorite in the fine clay fraction well above the palaeowater table. Furthermore, suggests that type H palaeosols formed under the significant carbonate accumulation, including conditions of low soil moisture given that chlor- stage II and III carbonate horizons in the profiles, ite is not stable in humid environments (cf. and the presence of deeply penetrating V-shaped Yemane et al., 1996). clastic-filled dykes require prolonged periods of Type H palaeosols are interpreted as having drying of the profiles and prolonged duration of formed on infrequently flooded, stable regions of pedogenesis. Early Permian floodplains given that they record well-drained and oxidizing soil conditions and that the development of thick petrocalcic hori- Type H palaeosols zons requires extended periods of soil formation Description and soil moisture deficiency (Machette, 1985; The type profile of pedotype H consists of six Wright & Marriott, 1996). The occurrence of moderately to well-developed horizons (Fig. 4H; wedge-shaped aggregates defined by weakly Table 1) that exhibit medium subangular to developed slickensides in horizons underlying coarse angular blocky structure in the upper the calcic horizons, however, records fluctuating horizons and wedge-shaped structural aggregates soil moisture conditions.

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 875

DISCUSSION channel-margin deposits and shallow floodplain 7scours (cf. Bown & Kraus, 1987; Kraus, 1986). For Permo-Pennsylvanian palaeosols of north-central the latest Pennsylvanian period, the metre-scale Texas exhibit clear stratigraphic trends in pedo- intercalation of type A and type B palaeosols in type distribution and their mineralogical and floodplain strata deposited on the lower palaeo- geochemical compositions. These superimposed coastal plain (Figs 6 and 15A) necessitates the trends are interpreted here as recording the presence of topography on the palaeolandscape interplay of autogenic factors inherent in the during pedogenesis. Although both palaeosol depositional system (e.g. variability in sediment types were hydromorphic, the presence of well- accumulation rate and parent material, local developed argillic horizons in type B palaeosols topography, variations in floodplain hydrology indicates episodic free drainage of these profiles or soil drainage conditions) and basin- to regio- in comparison with the penecontemporaneous nal-scale allogenic processes such as climate, water-logged type A palaeosols. However, the tectonic subsidence and eustasy. The overall shallow depth to the palaeowater table inferred stratigraphic distribution of the studied palaeo- from palaeoplinthites in the ultisols, as well as sols delineates the time of appearance and limited field relationships of channel-filling sand- disappearance of pedotypes over the Permo- stones and floodbasin strata, indicates that the Pennsylvanian interval, as well as defining palaeotopography was probably not greater than a long-term variations in the inferred degrees of few metres. hydromorphy, free drainage and chemical weath- Larger scale lateral variations (kilometres to ering of the soil profiles and estimated soil tens of km) defined by differences in the strati- moisture conditions. These long-term variations graphic distribution of pedotypes observed with- are interpreted here as recording the predominant in time-equivalent mudstone-dominated intervals role that climate played in the development of bracketed by laterally correlatable, major sand Permo-Pennsylvanian palaeosols associated with bodies (ss units in Figs 2, 6, 10, 12 and 13) further the Eastern shelf of the Midland basin. Before indicate that the geomorphic position in the discussing the application of these palaeosols to Permo-Pennsylvanian alluvial basin influenced constraining climate over western equatorial Pan- soil development, probably through the influence gea, the role of autogenic and basin-scale allo- of river channel avulsion and broader patterns of genic processes other than climate in determining floodplain drainage (cf. Kraus & Aslan, 1993; the development of the pedotypes and their Kraus, 2002). Overall, these inferred meso- and observed spatial distribution is first assessed. macroscale palaeosol–landscape associations Superimposed upon the longer term strati- indicate that differences in sediment accumula- graphic trend are smaller scale patterns in pedo- tion rates and topographically (or parent material) type distribution defined by palaeosols within controlled drainage clearly influenced soil devel- thin stratigraphic intervals (metres to a few tens opment on the Eastern shelf alluvial basin (e.g. metres; e.g. Fig. 14A and B). Lateral variations in Fig. 14B; Sobecki & Wilding, 1983; Wilding et al., pedotype distribution, degree of maturity of pal- 1991; Arndorff, 1993; Bestland et al., 1997; Kraus, aeosols and stratigraphic relationships between 1999, 2002; Kraus & Aslan, 1999). In turn, the palaeosols are also observed between age-equiv- imprint of landscape variability is recorded in alent deposits located hundreds of metres to tens the vertical stacking patterns of pedotypes at the of kilometres apart (Figs 6, 9, 10, 12 and 13). metre to tens of metre scale and accounts for the Reconstruction of palaeocatenas (e.g. fluvial stratigraphic ‘noise’ observed in the longer term channels and their associated floodplains) in this trend. study was limited to a few laterally continuous Larger scale (tens to hundreds of metres) strati- outcrops (hundreds of metres to 1Æ2 km long) graphic variations in the Permo-Pennsylvanian exposed along Holocene stream valleys or in palaeosols define a long-term trend upon which badland exposures (e.g. Figs 9 and 13; Tabor, the aforementioned smaller scale variations are 1999). The observed topography coupled with superimposed. The long-term trend records an palaeocatenary relationships inferred from metre- overall decrease in the degree of hydromorphy and scale vertical stacking patterns of pedotypes intensity of chemical weathering of the soil profiles indicates that geomorphic position on the aggrad- concomitant with an increase in evidence for free ing floodplain influenced soil development drainage of profiles and lower soil moisture through differences in topographically controlled throughout latest Pennsylvanian (Virgilian) to drainage and sediment accumulation provided by Early Permian time. This trend is defined by

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 876 N. J. Tabor & I. P. Montan˜ez

AB

Fig. 14. (A) Schematic stratigraphic column of fluvio-alluvial sedimentary rocks from the Early Permian succession of the Eastern Midland Basin, north-central Texas. The distribution of pedotypes between major fluvial channel deposits suggests that depositional rate, position of the palaeogroundwater table and duration of pedogenesis across the Permo-Pennsylvanian landscape played an important role in the stratigraphic stacking pattern of palaeosol morphologies. (B) Interpreted distribution of soil types across the Early Pennsylvanian landscape of the Eastern Midland Basin, north-central Texas. The stacking pattern of pedotypes between major fluvial channel sandstone deposits is interpreted as representing migration of fluvial systems from proximal (positions 0 and 1) to more distal (positions 4 and 6) and back to proximal locations (positions 8 and 9). See text.

floodplain palaeosols with morphologies that equivalent trend is defined by the distribution of reflect conditions of chemical weathering and soil type C and E profiles (Fig. 15). Type C and E pro- moisture regime (types A, B, D, F, G and H); no files are immature, poorly developed palaeosols

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 877 found throughout the three physiographic prov- The longer term trend defined by the studied inces. Their cosmopolitan nature probably reflects palaeosols could reflect a change in regional base the abundance of young, episodically ‘disturbed’ level governed by tectonic activity or eustasy that and poorly drained environments proximal to may not have been directly linked to climate fluvial channels and ponds upon the floodplain change (cf. Marriott & Wright, 1993; Wright & (Buol et al., 1997). Marriott, 1993). The palaeoshoreline migrated In particular, the longer term variation in north-north-eastward throughout the latest Penn- pedotype development records a relatively rapid sylvanian to Early Permian in response to a shift over tens of metres from Late Pennsylvanian relative sea-level rise, as recorded by the incur- hydromorphic floodplain palaeosols (histosol- sion of marine deposits on to the Eastern Shelf like type A; ultisol-like type B) to Early Permian (Figs 2 and 14; Hentz, 1988). This shift in base well-drained and oxidized profiles (alfisol-like level is expressed in the basinward shift in the type F; calcic vertisol-like type G; and aridisol- geomorphic position of palaeosol-bearing inter- like type H). Vertisols (type D) with abundant vals in the studied Permo-Pennsylvanian strata redoximorphic features but morphological evi- through time (Table 2; Fig. 15). This shift should dence for periods of drying (e.g. clastic dykes in be recorded by a progressive increase in hydro- the upper portions of profiles) make up an morphy in response to shallower regional ground- important component of latest Pennsylvanian water tables or overall increased maturity of mature palaeosols. These palaeosols are inter- floodplain palaeosols with increasing distance preted as recording intermediate soil moisture from sediment source areas (Wright & Marriott, conditions and the initiation of a drying trend 1993; Birkeland, 1999; Kraus, 1999). The ob- during the transition at the Pennsylvanian–Per- served long-term trend of decreasing degree of mian boundary interval. hydromorphy and intensity of chemical weather- The Early Permian palaeosols (types F, G and ing of the palaeosol profiles concomitant with H) record the onset of significant pedogenic increasing evidence for free drainage of profiles carbonate accumulation throughout the region, and overall falling soil moisture levels, however, which could reflect a decrease in average soil is contrary to the trends anticipated in response moisture levels and/or an increase in Ca2+ dust to a rise in base level. This indicates the predom- flux to the region. Moreover, a continuum of soil inant influence of an allogenic process(es) other moisture conditions is defined by these three than base level on the observed long-term trend in Early Permian pedotypes (alfisols ¼ wettest to pedotype distribution. However, the greater abun- aridisols ¼ driest) by the previously discussed dance of type G (calcic vertisols) and type F differences in their morphologies, abundance (alfisols) on landscapes proximal to the palaeo- and distribution of redoximorphic features and shoreline (0 to 30 km; Table 2) and a domin- clay composition. Indicators of seasonal or epi- ance of aridisols (type H) in regions of the alluvial sodic fluctuations in soil moisture conditions basin proximal to sediment sources (> 60 km occur in all mature palaeosols in the study from base level; Table 2) does indicate that a area except histosols. However, the degree to decline in topographic relief and sediment accu- which the morphological indicators of soil mois- mulation rates down palaeoslope on the Eastern ture fluctuations and, increasingly, better soil Shelf may have had a subordinate influence on drainage are developed in the studied palaeo- the observed larger scale stratigraphic variations sols increases significantly from weakly devel- in palaeosol morphology and maturity. The weak oped features in ultisols of the lower half of manifestation of these basin-wide trends, how- the Virgilian strata to strongly developed features ever, attests to the very low palaeoslopes on the in earliest Early Permian alfisols and calcic Late Palaeozoic Eastern shelf. Temporal variation vertisols. in Permo-Pennsylvanian parent material in the The stratigraphic distribution of Early Permian study area as an alternative allogenic control on pedotypes (F, G and H) does not define a strong the large-scale stratigraphic variations in pedo- temporal trend given that all three pedotypes type is not considered likely given the lack of any co-exist in contemporaneous intervals. Type F observed systematic long-term variation in parent palaeosols (alfisols), however, become somewhat material derived from the Ouachita or Arbuckle less abundant upwards through the Early uplands (Hentz, 1988). Permian strata (Waggoner Ranch Fm.) with a It is suggested that climate was the predomin- concomitant increase in the abundance of type H ant control on the origin of large-scale strati- palaeosols (aridisols; Fig. 15). graphic variations in pedotype distribution in

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Table 2. Relationship of Permo-Pennsylvanian palaeosol sites relative to estimated base-level on the Eastern shelf of the Midland Basin. 878

Distance from relative sea level and palaeosol morphology* .J ao .P Montan P. I. & Tabor J. N.

0–10 km 10–20 km 20–30 km 30–40 km 40–50 km 50–60 km > 60 km

Formation Site # Type Site # Type Site # Type Site # Type Site # Type Site # Type Site # Type Waggoner Ranch 31 H,F,G 26 H,G,I 29 H,G 30 H¢,E¢ –– –––– 27 H,G,I – – – – – – – – – – 28 H,G,E – – – – – – – – – – –––– 25H,E–––– –––– – – – – 24 H,E,G – – – – – – – – 23 G¢ ––––––––––– Petrolia – – 22 H,F,G,E – – – – – – – – 19 ˜ez H,F,E

––– ––––– ––20H,E

04ItrainlAscaino Sedimentologists, of Association International 2004 ––– ––––– ––21H,E – – 18 H – – – – – – – – – – –17H,F–––––– –––– Nocona – – 16 H,G,F,E – – – – – – – – 14 G,E –––– –––––– ––15E 13 G,E 12 G,E Archer City – – – – – – – – 9 G,H,F,E – – – – –––– ––––10H,G–––– –––– ––––11G,H–––– –––– –––––– 8G,H–– –––– –––––– 7G,E–– Markley – – – – – – – – – – – – 1 A,B,C – – – – – – – – – – – – 2 A,B,C –––– –––––– ––3A –––– –––––– ––4C –––– –––––– ––5C,E – – – – – – – – – – – – 6 C,D,E

Sedimentology * The distance from base level was determined by measuring the distance down depositional dip from the study sites to the vertically dashed lines in Fig. 2 that represent the stratigraphic cut-off between marine-dominated and terrestrial-dominated sedimentary strata (Hentz, 1988). These are likely to be minimum estimates, as soil formation occurred on the floodplains during times of base level fall or lowstand (e.g. mid-continent cyclothems). The position of each of the study sites and their associated palaeosol morphologies are given in relation to (1) distance from base level (inferred from dashed lines in Fig. 2); and (2) stratigraphic occurrence. The purpose of this table is to assess the distribution of palaeosols relative to base level. Although more data should be gathered to ,

51, test rigorously the effects of base level (and base level change) on the morphology of palaeosols in the study area, it is apparent that there is no robust change in palaeosol morphology as a function of distance from base level for any one time period. 851–884 Permo-Carboniferous palaeoclimate from palaeosols 879

Fig. 15. Schematic diagram of the Permo-Pennsylvanian landscape through time based on palaeosol morphological features. Roughly contemporaneous study sites from Figs 1 and 2 are plotted on each time period, as well as the palaeosol types associated with each study site. See text for discussion.

the study area. The Late Palaeozoic is hypothes- Pangea out of the Intertropical Convergence Zone ized to have been characterized by major changes into the more arid low-latitude belt; (2) rain in climate in response to the assembly of the shadow effects created by uplift of the Alleghe- supercontinent Pangea, a transition from ice- nian and Ouachita mountains; and (3) northward house to green-house states and major changes in diversion of moisture-laden easterlies by the atmospheric circulation (Veevers & Powell, 1987; development of northern hemisphere monsoonal Parrish, 1993; Crowley, 1994). Climatically sen- circulation. sitive depositional and palaeontological records Significantly, recent studies have suggested for low-latitude Pangea (Rowley et al., 1985; that monsoonal circulation and attendant west- Cecil, 1990; Dubiel et al., 1991; DiMichele & erly winds were well-established over western Aronson, 1992; Parrish, 1993; West et al., 1997; equatorial Pangea by earliest Permian time Kessler et al., 2001; Rees et al., 2001) coupled (Kessler et al., 2001; Soreghan, G.S. et al., 2002; with numerical model results (Parrish et al., Soreghan, M.J. et al., 2002; Tabor & Montan˜ ez, 1986; Kutzbach & Gallimore, 1989; Patzkowsky 2002; Loope et al., 2004), despite model results 8et al., 1991; Gibbs et al., 2002) suggest that the indicating that continental-scale monsoonal cir- mid-latitude and equatorial regions of Pangea culation would have strengthened progressively became progressively more arid and seasonal to its peak in the Middle Triassic (Parrish, 1993). throughout the Late Palaeozoic. The long-term The results of these lithological and geochemical aridification trend has been attributed to several studies, along with other discrepancies in model- processes including: (1) northward drift of data comparisons for low-latitude western Pangea

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 880 N. J. Tabor & I. P. Montan˜ez

(Patzkowsky et al., 1991; Rees et al., 2001; Gibbs western equatorial Pangea, this rapid change in et al., 2002), highlight the need for a better regional climatic conditions may record a tec- understanding of specific climate zones over Late tonic moderation of monsoonal conditions. Up- Palaeozoic Pangea. Pennsylvanian to Permian lift of the equatorial mountain chain in the Early palaeosols from north-central Texas provide addi- Pennsylvanian (Chesterian) would have acted to tional constraints on the nature and origin of intensify the normal low pressure area of the climate change over low-latitude Pangea given equatorial belt, thus inhibiting the development that climate change would have been most pro- of fully monsoonal conditions until their erosion nounced in western equatorial Pangea (Dubiel in the latest Pennsylvanian and Early Permian et al., 1991; Parrish, 1993). dampened their climatic influence (Rowley The long-term variation in soil moisture condi- et al., 1985; Otto-Bliesner, 1993). Notably, the tions and, in turn, precipitation inferred from the delay in the development of fully monsoonal large-scale stratigraphic variations in Permo- conditions may have promoted relatively rapid Pennsylvanian pedotypes confirms the previ- intensification of monsoonal conditions over the ously defined long-term aridification trend for study area in the Early Permian as has been western equatorial Pangea. Notably, however, the invoked to explain the origin of rapid variation palaeosol data in this study indicate that the in lithofacies of mid-continent cyclothems in the long-term drop in inferred soil moisture levels latest Pennsylvanian to earliest Permian (West accelerated at end-Pennsylvanian time. Incipient et al., 1997). morphological indicators of seasonality preserved Lastly, pedogenic evidence for intermediate- in type D palaeosols of the study area suggest an scale (103)105 years) climate fluctuations in abrupt onset of seasonality in conjunction with a western equatorial Pangea occurs in the poly- rapid decline in net soil moisture budget in the genetic alfisols characterized by well-developed latest Pennsylvanian. calcic horizons overlying or superimposed on The long-term aridification trend inferred from argillic horizons. These polygenetic palaeosols the palaeosol data cannot be interpreted as are restricted stratigraphically to the lower half of reflecting northward drift of Pangea given that Lower Permian (Wolfcampian) deposits of north- palaeomagnetic studies indicate that the area central Texas, and reappear for a few tens of remained within a few degrees of the equator metres in the uppermost Leonardian-age throughout the Early Permian (Scotese, 1999; Waggoner Ranch and lowermost Clear Fork For- Loope et al., 2004). The observed temporal trend mations. These polygenetic profiles are analogous may record the development of a rain shadow in to those that occur within slightly younger, western Pangea in response to uplift of the Lower Permian mid-continent cyclothems that equatorial Alleghenian and Ouachita mountains have been interpreted to record orbitally forced (Rowley et al., 1985; Parrish, 1993; Ziegler et al., 20 ky climate oscillations associated with wax- 1997). However, the degree to which long-term ing and waning of Early Permian continental aridification through the latest Pennsylvanian glaciers (Miller et al., 1996). Such precessional and Early Permian can be explained by a rain forcing of climate would probably have been shadow effect is strongly dependent on the dampened following deglaciation and a trans- elevation of the equatorial highlands, for which ition to greenhouse states. If polygenetic palaeo- a significant degree of uncertainty exists (Otto- sols from the eastern Midland basin record 9Bliesner, 1993, 1998; Fluteau et al., 2001; Gibbs precessional-scale climate fluctuations, then the et al., 2002). Moreover, the orographic effect of abrupt shift to more stable climatic conditions these prominent equatorial highlands on tropical marked by the loss of polygenetic palaeosols in precipitation (i.e. onset of aridity) should have the Early Permian (mid-Leonardian) succession been well-established by Late Carboniferous time may be a sensitive record of the onset of degla- rather than in the Early Permian as the palaeosols ciation. The brief reappearance of polygenetic of the Eastern shelf suggest (cf. Kessler et al., alfisols in the later part of the Early Permian 2001; Soreghan G.S. et al., 2002). might record a brief renewed episode of glaci- The overall long-term palaeopedogenic trend ation or a climatic cooling analogous to the late and its inferred rapid decline in precipitation Quaternary Younger Dryas. This interpretation, if levels and intensification of seasonality in west- correct, places additional constraints on most ern equatorial Pangea in the latest Pennsylvanian recent estimates of a late Early Permian age for and earliest Permian are strong evidence for the onset of deglaciation (Veevers, 1994; Gibbs northern hemisphere monsoonal circulation. For et al., 2002).

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 881

CONCLUSIONS providing an introduction to the study area. John Nelson and Dan Chaney provided excellent field Late Pennsylvanian and Early Permian sedimen- guidance. This project was funded by NSF grant tary strata of the Eastern Midland basin exhibit a EAR-9814640 to I. P. Montanez and an NSF-IGERT change in palaeosol morphological, mineralogical graduate fellowship to N. J. Tabor. Funding was and chemical characters across the Pennsylva- also provided by grants to N.J.T. from the Geolo- nian–Permian boundary. This change is revealed gical Society of America, American Association of by palaeosols with well-developed redoximor- Petroleum Geologists, Society for Sedimentary phic depletions and concentrations in the Late Geology and Sigma Xi. We thank Peter Haughton, Pennsylvanian that are replaced by palaeosols Paul McCarthy and Susan Marriott for their that generally lack redoximorphic indicators in thoughts and helpful comments on this manu- the Early Permian. Furthermore, many of the script. All these reviewers (but particularly S.M.) Early Permian palaeosols are characterized by have contributed significantly towards improving pedogenic carbonate accumulation that is consis- the quality and clarity of this contribution. tent with a net soil moisture deficiency, whereas Late Pennsylvanian palaeosols exhibit no pedo- genic carbonate. The clay mineralogical compo- REFERENCES sition of these palaeosols indicates a change from more highly weathered and probably humid Arndorff, L. (1993) Calcretes from Hollviken-II drill core, conditions in the Late Pennsylvanian to less Skane, southern Sweden. In: Lundadagarna I; Historical Geology and Paleontology (Ed. M. Siverson), Lund Publ. extreme weathering and potentially semi-arid to Geol., 109, 3. arid conditions in the Early Permian. These Barnes, V.E., Hentz, T.F. and Brown, L.F., Jr (1987) Geologic changes may be a response to changing soil Atlas of Texas; Wichita Falls – Lawton Sheet. University of moisture regimes from humid, poorly drained to Texas at Austin, Bureau of Economic Geology, Austin, TX. dry, well-drained conditions in the Late Pennsyl- Bein, A. and Land, L. (1983) Carbonate sedimentation and vanian to Early Permian. diagenesis associated with Mg-Ca-chloride brines; the Per- mian San Andres Formation in the Texas panhandle. J. Sed. The inferred change in Late Pennsylvanian– Petrol., 53, 243–260. Early Permian soil moisture regimes of western Bestland, E.A., Retallack, G.J. and Swisher, C.C., III (1997) equatorial Pangea is likely to be related to climate Stepwise climate change recorded in Eocene-Oligocene change. It is suggested, based on the preserved paleosol sequences from central Oregon. J. Geol., 105, 153–172. lithological, mineralogical and chemical charac- Birkeland, P.W. (1999) Soils and Geomorphology, 3rd edn. teristics of these palaeosols, that a change Oxford University Press, New York, 430 pp. occurred from a humid tropical climate charac- Bown, T.M. and Kraus, M.J. (1987) Integration of channel and terized by high rainfall in the Pennsylvanian to a floodplain suites. I. Developmental sequence and lateral progressively drier tropical climate characterized relations of alluvial paleosols. J. Sed. Petrol., 57, 587–601. van Breeman, N. and Harmsen, K. (1975) Translocation of iron by seasonal precipitation throughout the Early in acid sulfate soils; I, soil morphology and the chemistry Permian. Furthermore, these data indicate the and mineralogy of iron in a chronosequence of acid sulfate onset of monsoonal-type precipitation patterns soils. Soil Sci. Soc. Am., 39, 1140–1147. nearly coincident with the Pennsylvanian–Per- Brewer, R. (1976) Fabric and Mineral Analysis of Soils. mian boundary. Although these strata preserve a Krieger, New York, 482 pp. general trend of progressively drier and seasonal Brown, L.F., Jr (1973) Pennsylvanian rocks of North-Central Texas: an introduction. In: Pennsylvanian Depositional precipitation through the Permian, polygenetic Systems in North-Central Texas (Ed. L.F. Brown, Jr), Bur. soils preserved within the lower half of the Early Econ. Geol. Guidebook, 14, 1–9. Permian (Wolfcampian) and uppermost strata of Brown, L.F., Jr Solis Irarte, R.F. and Johns, D.A. (1987) Re- the lower Leonardian Waggoner Ranch Formation gional Stratigraphic Cross Sections, Upper Pennsylvanian and Lower Permian Strata (Virgilian and Wolfcampian may indicate an oscillatory climate pattern Series), North-central-Central Texas. Bureau of Economic between more humid and drier climate during Geology, University of Texas at Austin, Austin, TX, 27 pp. deposition of these stratigraphic packages Buol, S.W., Hole, F.D., McCracken, R.J. and Southard, R.J. (1997) Soil Genesis and Classification. Iowa State Univer- sity Press, Ames, IA, 527 pp. Cecil, C.B. (1990) Paleoclimate controls on stratigraphic repeti- ACKNOWLEDGEMENTS tion of chemical and siliciclastic rocks. Geology, 18, 533–536. Crowley, T.J. (1994) Pangean climates. In: Pangea: Paleocli- We are deeply indebted to Bill DiMichele and Dan mate, Tectonics, and Sedimentation During Accretion, Chaney, Department of Palaeobiology, Museum Zenith, and Breakup of a Supercontinent (Ed. G.D. Klein), of Natural History, Smithsonian Institution, for Geol. Soc. Am. Spec. Paper, 288, 25–40.

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 882 N. J. Tabor & I. P. Montan˜ez

Daniels, R.B., Gamble, E.E. and Nelson, L.A. (1971) Relations Kraus, M.J. (1986) Integration of channel and floodplain suites between soil morphology and water table levels on a dis- II. Vertical relations of alluvial paleosols. J. Sed. Petrol., 57, sected North-central Carolina coastal plain surface. Soil Sci. 602–612. Soc. Am. Proc., 35, 157–175. Kraus, M.J. (1998) Development of potential acid sulfate pale- Deutz, P., Montan˜ ez, I.P. and Monger, H.C. (2002) Morphol- osols in Paleocene floodplains, Bighorn Basin, Wyoming ogies and stable and radiogenic isotope compositions of USA Palaeogeogr. Palaeoclimatol. Palaeoecol., 144, 203–224. pedogenic carbonates in Late Quaternary relict and buried Kraus, M.J. (1999) Paleosols in clastic sedimentary rocks: their soils, New Mexico: an integrated record of pedogenic geologic significance. Earth-Sci. Rev., 47, 41–70. overprinting. J. Sed. Res., 72, 809–822. Kraus, M.J. (2002) Basin-scale changes in floodplain paleosols: DiMichele, W.A. and Aronson, R.B. (1992) The Pennsylva- implications for interpreting alluvial architecture. J. Sed. nian-Permian vegetational transition: a terrestrial analogue Petrol., 72, 500–509. to the onshore-offshore hypothesis. Evolution, 46, 807–824. Kraus, M.J. and Aslan, A. (1993) Eocene hydromorphic Dixon, J.B. and Skinner, H.C.W. (1992) Manganese minerals in paleosols: significance for interpreting ancient floodplain surface environments. Catena Suppl., 21, 31–50. processes. J. Sed. Petrol., 63, 453–463. Donovan, R.N., Collins, K. and Bridges, S. (2001) Permian Kraus, M.J. and Aslan, A. (1999) Paleosol sequences in sedimentation and diagenesis on the Northern Margin of the floodplain environments: a hierarchical approach. In: Pal- Wichita uplift. Okl. Geol. Surv. Circ., 104, 171–184. aeoweathering, Palaeosurfaces and Related Continental Dubiel, R.F., Parrish, J.T., Parrish, J.M. and Good, S.C. (1991) Deposits. (Ed. M. Thiry), Int. Assoc. Sedimentol. Spec. The Pangean megamonsoon – evidence from the Upper Publ., 27, 303–321. Triassic Chinle Formation, Colorado Plateau. Palaios, 6, Kutzbach, J.E. and Gallimore, R.G. (1989) Pangean climates: 347–370. megamonsoons of the megacontinent. J. Geophys. Res., 94, Duchaufour, P. (1982) Pedology: Pedogenesis and Classifica- 3341–3357. tion. George Allen & Unwin, London, 448 pp. Loope, D.B., Steiner, M.B. and Rowe, C.M. (2004) Tropical Dudal, R. and Eswaran, H. (1988) Chapter 1: Distribution, westerlies over Pangaean sand seas. Sedimentology, 51, properties and classification of Vertisols. In: Vertisols: Their 315–322. Distribution, Properties, Classification and Management McFadden, L.D. (1988) Climatic influences on rates and pro- (Eds L.P. Wilding and R. Puentes), pp. 1–22. Texas A&M cesses of soil development in Quaternary deposits of University Publishing Center, College Station, TX. Southern California. In: Paleosols and Weathering Through Dunbar, C.O. and Committee on Stratigraphy (1960) Correla- Time; Principles and Applications (Eds J. Reihardt and tion of the Permian formations of North-central America. W.R. Sigleo), Geol. Soc. Am. Spec. Publ., 216, 153–177. Geol. Soc. Am. Bull., 71, 1763–1806. McGowen, J.H., Granata, G.E. and Seni, S.J. (1979) Deposi- Fanning, D.S. and Fanning, M.C.B. (1989) Soil: Morphology, tional framework of the Lower Dockum Group (Triassic) Genesis, and Classification. Wiley and Sons, New York, 395 pp. Texas Panhandle. Bur. Econ. Geol. Report Invest., 97, 60 pp. Fluteau, F., Besse, J., Broutin, J. and Ramstein, G. (2001) The Machette, M.N. (1985) Calcic soils of southwestern United Late Permian climate. What can be inferred from climate States. Geol. Soc. Am. Spec. Paper, 203, 1–21. modeling concerning Pangea scenarios and Hercynian range Mack, G.H. (1992) Paleosols as an indicator of climatic change altitude? Palaeogeogr. Palaeoclimatol. Palaeoecol., 167, at the Early–Late Cretaceous boundary, southwestern New 39–71. Mexico. J. Sed. Petrol., 62, 483–494. Franzmeier, F.P., Bryant, R.B. and Steinhardt, G.C. (1985) Markewich, H.W. and Pavich, M.J. (1991) Soil chrono- Characteristics of Wisconsinan glacial tills in Indiana and sequence studies in temperate to subtropical, low-latitude, their influence on Argillic horizon development. Soil Sci. low relief terrain with data from the eastern United States. Soc. Am. J., 49, 1481–1486. Geoderma, 51, 213–239. Gibbs, M.T., Rees, P.M., Kutzbach, J.E., Ziegler, A.M., Marriott, S.B. and Wright, V.P. (1993) Paleosols as indicators Behling, P.J. and Rowley, D.B. (2002) Simulations of Per- of geomorphic stability in two Old Red Sandstone alluvial mian climate and comparisons with climate-sensitive sedi- suites, South Wales. J. Geol. Soc. London, 150, 1109–1120. ments. J. Geol., 110, 33–55. Miller, K.B., McCahon, T.J. and West, R.R. (1996) Lower Gill, S. and Yemane, K. (1996) Implications of a Lower Permian (Wolfcampian) paleosol-bearing cycles of the US Pennsylvanian Ultisol for equatorial Pangean climates and midcontinent; evidence of climatic cyclicity. J. Sed. Res., early oligotrophic forest ecosystems. Geology, 24, 905–908. 66, 71–84. Golonka, J., Ross, M.I. and Scotese, C.R. (1994) Phanerozoic Munsell Color (1975) Munsell Soil Color Charts. Munsell paleogeographic and paleoclimatic modeling maps. In: Color, Baltimore, MD, 24 pp. Pangea: Global Environments and Resources (Eds A.F. Nelson, W.J., Hook, R.W. and Tabor, N.J. (2001) Clear Fork Embry, B. Beauchamp and D.J. Glass), Can. Soc. Petrol. Group (Leonardian, Lower Permian) of North-Central Texas. Geol. Mem., 17, 1–47. Okl. Geol. Surv. Circ., 104, 167–169. Hentz, T.F. (1988) Lithostratigraphy and paleoenvironments of Oriel, S., Myers, D.A. and Crosby, E.J. (1967) Paleotectonic Upper Paleozoic continental red beds, north-central-central investigations of the Permian system in the United States. Texas: Bowie (new) and Wichita (revised) Groups. Bur. US Geol. Surv. Prof. Paper, 515, Chapter C, 80 pp. Econ. Geol. Report Invest., 170, 49 pp. Otto-Bliesner, B.L. (1993) Tropical mountains and coal for- Jim, C.Y. (1990) Stress, shear deformation and micromorpho- mation: a climate model study of the Westphalian (306 MA). logical clay orientation: a synthesis of various concepts. Geophys. Res. Lett., 20, 1947–1950. Catena, 17, 431–447. Otto-Bliesner, B.L. (1998) Effects of tropical mountain eleva- Kessler, J.L.P. and Soreghan, G.S. and Wacker, H.J. (2001) tions on the climate of the Late Carboniferous: Climate Equatorial aridity in western Pangea: Lower Permian loes- model simulations. In: Tectonic Boundary Conditions for site and dolomitic paleosols in northeastern New Mexico, Climate Reconstructions (Eds T.J. Crowley and K.C. Burke), USA. J. Sed. Res., 71, 817–832. Oxford Monogr. Geol. Geophys., 39, 100–115.

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 Permo-Carboniferous palaeoclimate from palaeosols 883

Parrish, J.T. (1993) Climate of the supercontinent Pangea. climate in upper Paleozoic loessite (lower Cutler beds, J. Geol., 101, 215–233. Utah). Geol. Soc. Am. Bull., 114, 1019–1035. Parrish, J.M., Parrish, J.T. and Ziegler, A.M. (1986) Permian- Soreghan, M.J., Soreghan, G.S. and Hamilton, M.A. (2002) Triassic paleogeography and paleoclimatology and impli- Paleowinds inferred from detrital-zircon geochronology of cations for therapsid distribution. In: The Ecology and upper Paleozoic loessite, western equatorial Pangea. Geol- Biology of Mammal-Reptiles (Eds N. Hotton, P.D. MacLean, ogy, 30, 695–698. J.J. Roth and C. Roth), pp. 109–131. Smithsonian Institution Strahler, A.N. and Strahler, A.H. (1983) Modern Physical Press, Washington, DC. Geography, 2nd edn. John Wiley and Sons, New York, 532 Patzkowsky, M.E., Smith, L.H., Markwick, P.J., Engberts, C.J. pp. and Gyllenhaal, E.D. (1991) Application of the Fujita-Zie- Tabor, N.J. (1999) Permo-Pennsylvanian alluvial paleosols gler paleoclimate model: Early Permian and Late Cretaceous (north-central Texas): high-resolution proxy records of the examples. Palaeogeogr. Palaeoclimatol. Palaeoecol., 86, evolution of early Pangean Paleoclimate. Unpubl. Master’s 67–85. Thesis, University of California, 121 pp. Rasbury, E.T., Hanson, G.N., Meyers, W.J., Holt, W.E., Gold- Tabor, N.J. and Montan˜ez, I.P. (2002) Shifts in late Paleozoic stein, R.H. and Saller, A.H. (1998) U-Pb dates of paleosols; atmospheric circulation over western equatorial Pangea: constraints on late Paleozoic cycle durations and boundary Insights from pedogenic mineral d18O compositions. Geol- ages. Geology, 26, 403–406. ogy, 30, 1127–1130. Rees, P.M., Ziegler, A.M., Gibbs, M.T., Kutzbach, J.E., Veevers, J.J. (1994) Pangea. Evolution of a supercontinent and Behling, P.J. and Rowley, D.B. (2001) Permian phytogeo- its consequences for earth’s paleoclimate and sedimentary graphic patterns and climate data/model comparisons. environments. In: Pangea: Paleoclimate, Tectonics, and J. Geol., 110, 1–31. Sedimentation During Accretion, Zenith, and Breakup of a Retallack, G.J. (1988) Field recognition of paleosols. Supercontinent (Ed. G.V. Klein), Geol. Soc. Am. Spec. Pa- In: Paleosols and Weathering Through Geologic Time: per, 288, 13–23. Principles and Applications (Eds J. Reinhardt and Veevers, J.J. and Powell, M. (1987) Late Paleozoic glacial W.R. Sigleo), Geol. Soc. Am. Spec. Paper, 216, 1–20. episodes in Gondwanaland reflected in transgressive- Retallack, G.J. (1990) Soils of the Past: An Introduction to regressive depositional sequences in Euramerica. Geol Soc. Paleopedology. Unwin-Hyman, Boston, 520 pp. Am. Bull., 98, 574–487. Retallack, G.J. (1994) A pedotype approach to latest Creta- Vepraskas, M.J. (1994) Redoximorphic features for identi- ceous and earliest Tertiary paleosols in eastern Montana. fying aquic conditions. NC Agric. Res. Service Tech. Bull., Geol. Soc. Am. Bull., 106, 1377–1397. 301 pp. Retallack, G.J. and German-Heins, J. (1994) Evidence from West, R.R., Archer, A.W. and Miller, K.B. (1997) The role of paleosols for the geological antiquity of rain forest. Science, climate in stratigraphic patterns exhibited by Late Paleozoic 265, 499–502. rocks exposed in Kansas. Palaeogeogr. Palaeoclimatol. Pal- Ross, C.A., Baud, A. and Menning, M. (1994) A timescale for aeoecol., 128, 1–16. project Pangea. In: Pangea; Global Environments and Re- Wilding, L.P. and Tessier, D. (1988) Genesis of vertisols: shrink- sources (Eds B. Beauchamp and D.J. Glass), Can. Soc. Petrol. swell phenomena. In: Vertisols: Their Distribution, Proper- Geol. Mem., 17, 81–83. ties, Classification and Management (Eds L.P. Wilding and Rowley, D.B., Raymond, A., Parrish, J.T., Lottes, A.L., R. Puentes), pp. 55–81. Texas A&M University Publishing Scotese, C.R. and Ziegler, A.M. (1985) Carboniferous Center, College Station, TX. paleogeographic, phytogeographic, and paleoclimatic Wilding, L.P., Milford, M.H. and Vepraskas, M.J. (1983) reconstructions. Int. J. Coal Geol., 5, 7–42. Micromorphology of deeply weathered soils in the Texas Royer, D.L. (1999) Depth to pedogenic carbonate horizon as a Gulf Coast Plains. In: Soil Micromorphology, Vol. 2 (Eds paleoprecipitation indicator? Geology, 27, 1123–1126. P. Bullock and C.P. Murphy), pp. 567–574. AB Academic Schwertmann, U. (1985) Occurrence and formation of iron Publishers, Hertfordshire. oxides in various pedoenvironments. In: Iron in Soils and Wilding, L.P., Williams, D., Cook, T., Miller, W. and Clay Minerals (Eds J.W. Stucki, B.A. Goodman and 11 Eswaran, H. (1991) Close interval spatial variability of U. Schwertmann), NATO Adv. Study Inst. Series C, 217, vertisols: a case study in Texas. Proceedings of the Sixth 267–308. International Soil Correlation Meeting, pp. 232–247. US Scotese, C.R. (1999) PALEOMAP Animations ‘Paleogeogra- Department of Agriculture – Soil Conservation Service, phy’. PALEOMAP Project. Department of Geology, Univer- Washington, DC. sity of Texas, Arlington, TX. Wright, V.P. (1990) A micromorphological classification of Simonson, G.H. and Boersma, L. (1972) Soil morphology and fossil and recent calcic and petrocalcic microstructures. water table relations. II. Correlation between annual water Proc. Int. Working Meeting Soil Micromorphol., 8, 401– table fluctuations and profile features. Soil Sci. Soc. Am. 407. Proc., 36, 649–653. Wright, V.P. and Marriott, S.B. (1993) The sequence strati- Sobecki, T.M. and Wilding, L.P. (1983) Calcic horizon distri- graphy of fluvial depositional systems: the role of floodplain bution and soil classification in selected soils of the Texas sediment storage. Sed. Geol., 86, 203–210. Coast Prairie. Soil Sci. Soc. Am. Proc., 46, 1222–1227. Wright, V.P. and Marriott, S.B. (1996) A quantitative approach Soil Survey Staff (1975) Soil Taxonomy. US Department of to soil occurrence in alluvial deposits and its application to Agriculture Handbook, 436, 754 pp. the Old Red Sandstone of Britain. J. Geol Soc. London, 153, Soil Survey Staff (1998) Keys to Soil Taxonomy. US Depart- 907–913. ment of Agriculture Natural Resources Conservation Yaalon, D.H. and Kalmar, D. (1978) Dynamics of cracking Service, Washington, DC, 325 pp. and swelling clay soils: displacement of skeletal grains, Soreghan, G.S., Elmore, R.D. and Lewchuk, M.T. (2002) optimum depth of slickensides, and rate of intra-pedonic Sedimentologic-magnetic record of western Pangean turbation. Earth Surf. Proc., 3, 31–42.

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884 884 N. J. Tabor & I. P. Montan˜ez

Yemane, K., Kahr, G. and Kelts, K. (1996) Imprints of post- Postglacial Environmental Changes – Pleistocene, Car- glacial climates and paleogeography in the detrital clay boniferous-Permian and Proterozoic (Ed. I.P. Martini), mineral assemblages of an Upper Permian fluviolacustrine pp. 111–146. Oxford University Press, Oxford. Gondwana deposit from north-central Malawi. Palaeogeogr. Palaeoclimatol. Palaeoecol., 125, 27–49. Manuscript received 2 July 2003; Ziegler, A.M., Hulver, M.L. and Rowley, D.B. (1997) Permian revision accepted 22 March 2004. World Topography and Climate. In: Late Glacial and

2004 International Association of Sedimentologists, Sedimentology, 51, 851–884