<<

International Journal of Coal Geology 154–155 (2016) 136–154

Contents lists available at ScienceDirect

International Journal of Coal Geology

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

Sequence of fluvially-dominated strata of the Mid-Late Conemaugh Formation, Central Appalachian Basin

Ronald L. Martino

Department of Geology, Marshall University, Huntington, WV 25755, United States article info abstract

Article history: Sequence stratigraphic analysis of nonmarine, fluvially-dominated strata is particularly challenging compared to Received 14 August 2015 paralic and marine sequences due to rapid facies changes and limited lateral continuity of key beds. The Received in revised form 29 December 2015 Conemaugh Formation of central is a prime example of this type of stratigraphic interval. This Accepted 30 December 2015 study describes the stratigraphy and sedimentary facies from 68 outcrops in the lower 100 m of the Mid-Late Available online 31 December 2015 Pennsylvanian, nonmarine Conemaugh Formation in this area. Facies identified include 1) fluvial–upper estua- Keywords: rine channel sandstones, 2) crevasse splay channel and sheet sandstones, 3) lacustrine shale, mudstone, and Paleosols claystone, 4) lacustrine/palustrine limestone, 5) paludal coal and carbonaceous shale, and 6) hackly mudstone Sequence stratigraphy and claystone paleosols. Conemaugh Fm. Mature, polygenetic, high-chroma calcic vertisols and calcisols are the regionally developed interfluvial sequence Nonmarine cyclothems boundaries (IFSBs) of seven fourth-order sequences (major cyclothems) between the Upper Freeport and Har- lem coal horizons. Initial paleosol development occurred under well-drained conditions and strongly seasonal, semiarid–arid climate approaching and during glacial maxima at a time of minimal accommodation space on the interfluves. Subsequent rising water table associated with rising base level occurred as interglacial sea level rose and the climate became less seasonal and more humid. This initially led to deposition of thin, carbonaceous mud and peat followed by lacustrine and palustrine limestones. The presence of spirorbid microconchids indi- cates that the lakes were at least intermittently connected to the sea, which reached to within 50–80 km of all outcrops in the study during five transgressions. The tops of these limestones represent maximum flooding sur- faces. They are overlain by coarsening–upward lake-fill sequences formed during high accommodation on the in- terfluves during the HST. A nonmarine sequence stratigraphic model and a polygenetic paleosol model are proposed for the Conemaugh in central West Virginia. This study underscores the importance of recognizing re- gionally developed IFSB paleosols and microconchid limestones in correlation and sequence stratigraphic analysis. © 2015 Elsevier B.V. All rights reserved.

1. Introduction perception that deposition was dominated by autocyclic processes such as delta switching and river avulsion that precluded the develop- The study of Pennsylvanian cyclothems in North America has had a ment of widespread marker beds (e.g. Donaldson, 1979). long history dating back to the early 1900s (e.g. Weller, 1930; Sequence stratigraphic studies of fluvial successions are hindered by Wanless and Weller, 1932; Wanless and Shepard, 1936). Cylcothems rapid lateral facies changes and inability to distinguish time-significant have more recently been viewed within the context of sequence stratig- surfaces. The recognition of widespread mature paleosols as interfluvial raphy (sensu Vail et al., 1977; e.g. Miall, 2010). The bulk of sequence sequence boundaries between incised valley fills (IVFs) has led to signif- stratigraphic studies to date have addressed stratigraphic intervals icant progress in this area (Wright and Marriott, 1993; Shanley and with marine components. Marine–coastal cyclothems have been de- McCabe, 1994; Gibling and Bird, 1994; McCarthy et al., 1999; McCarthy, scribed from the lower Conemaugh of the north- 2002). The purpose of this study is to develop a high-resolution ern and central Appalachian Basin (e.g. Stout, 1947; Sturgeon and stratigraphic framework by correlating paleosol-bounded terres- Hoare, 1968; Busch and Rollins, 1984; Busch and West, 1987; Martino, trial cyclothems in central West Virginia with their downdip, 2004). Henry et al. (1979) and Windolph (1987) were unable to distin- marine-influenced equivalents. Sedimentary facies architecture will guish or correlate Conemaugh cyclothems in central West Virginia. be analyzed and compared with nonmarine sequence stratigraphic Their efforts were hindered by 1) the absence of marine shales/lime- models. The sequence stratigraphic significance of “nonmarine” (brack- stones and the paucity of laterally persistent coal beds, and 2) the ish-freshwater) limestones will be evaluated.

http://dx.doi.org/10.1016/j.coal.2015.12.016 0166-5162/© 2015 Elsevier B.V. All rights reserved. R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 137

Fig. 1. Map showing outcrop locations used in this study. The regional paleoslope was toward the northwest. The transgressive maximum line is based on projected maximum extent of Brush Creek and Ames marine units (modified from Busch and West, 1987). Conemaugh outcrops southeast of this line (i.e. updip) are comprised of nonmarine cyclothems whereas those northwest of it (i.e. downdip) contain marine-cored cyclothems (Martino, 2004).

2. Methods described at 68 outcrops (Fig. 1, Table 1). Paleosols were identified in the field using soil structure, horizonization, and root traces (Retallack, 1988). A total of 2251 m of strata was measured through the lower 80–100 m of Elevations were determined with an American Paulin System Micro altime- the Conemaugh Formation and component sedimentary facies were ter, or from satellite imagery using Google Earth software. Lithostratigraphic

Table 1 Outcrop location coordinates in latitude and longitude and UTM coordinates (WGS 1984 Zone 17N datum).

Loc. Latitude Longitude Northing Easting Loc. Latitude Longitude Northing Easting

BX2 38.68838 −80.68688 4282243 527231.8 K51 38.3855 −81.83572 4248919 427010.8 BX3 38.66811 −80.77741 4279971 519364 K52 38.38717 −81.7918 4249071 430848.4 BX5 38.68706 −80.68956 4282096 526999.5 K53 38.39175 −81.81247 4249594 429048.2 BX6 38.70611 −80.65954 4284219 529602.4 K54 38.45545 −81.49139 4256465 457125.7 BX7 38.69973 −80.66369 4283510 529244 K55 38.45793 −81.49527 4256741 456788 BX8 38.62732 −80.72181 4275458 524214.8 K56 38.45793 −81.49527 4256741 456788 BX9 38.62833 −80.71881 4275571 524475.4 K57 38.52115 −81.35114 4263699 469390.3 BX14 38.65836 −80.72679 4278901 523771.1 K58 38.52115 −81.35114 4263699 469390.3 BX15 38.65836 −80.72679 4278901 523771.1 K59 38.52939 −81.33951 4264609 470407.8 BX20 38.64676 −80.7194 4277616 524418.4 K60 38.42554 −81.52571 4253163 454112.2 BX22 38.61624 −80.84808 4274203 513225.5 K61 38.37817 −81.60347 4247948 447289.7 C1 38.24939 −82.28829 4234271 387273.9 K62 38.36272 −81.84536 4246399 426146 C2 38.26804 −82.25398 4236300 390303.4 K63 38.3676 −81.82152 4246922 428233.8 C3 38.26871 −82.20683 4236319 394429.8 K64 38.41507 −81.67188 4252083 441344.3 C4 38.27642 −82.23976 4237212 391559.9 K65 38.42002 −81.64309 4252615 443861.9 C5 38.25644 −82.28838 4235054 387276.5 L1 38.29213 −82.18998 4238898 395937.2 C6 38.26382 −82.29592 4235882 386628.4 L2 38.27939 −82.12436 4237413 401658.7 C7 38.32634 −82.21858 4242727 393486 L3 38.26064 −81.83011 4235060 427376.6 C8 38.32783 −82.20703 4242879 394497.2 L4 38.26696 −81.82177 4235754 428112.4 K36 38.39691 −81.59226 4250020 448282.5 L5 38.23305 −81.83867 4232005 426600.2 K37 38.4191 −81.55082 4252461 451915.5 L7 38.24726 −81.8335 4233578 427066.5 K38 38.44799 −81.51548 4255649 455018.8 L11 38.2826 −82.17377 4237823 397340.9 K39 38.44022 −81.51693 4254787 454887.2 L12 38.28259 −82.17381 4237821 397337.5 K40 38.30428 −81.7007 4239809 438734.7 R1 38.57037 −81.26346 4269134 477049.4 K41 38.36352 −81.72219 4246397 436907.8 R2 38.58355 −81.20568 4270585 482085.6 K42 38.33656 −81.70144 4243391 438697.4 R3 38.551 −81.29602 4266994 474206.1 K43 38.40991 −81.65176 4251498 443096.8 R4 38.54639 −81.30682 4266485 473263.3 K44 38.41429 −81.6242 4251967 445506.1 R5 38.58218 −81.21691 4270435 481107.2 K45 38.37554 −81.79396 4247781 430649.3 R6 38.5632 −81.13186 4268314 488512.1 K47 38.38151 −81.60131 4248317 447480.9 R7 38.56283 −81.14188 4268275 487639.4 K48 38.38522 −81.61397 4248736 446378.2 R8 38.57196 −81.2605 4269311 477307.3 K49 38.40384 −81.62 4250806 445864.9 R9 38.56941 −81.15995 4269008 486066.1 K50 38.40111 −81.66085 4250527 442295.9 R10 38.57085 −81.26767 4269189 476682.5 138 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 correlations were based on similarities in well-developed paleosols taking Sandy Grove Sandstone (Windolph, 1987) and the Pittsburgh coal (West into account their thickness, type, and elevation, and were guided in part by Virginia Geological and Economic Survey Coal Bed Mapping Project; Krebs their relations to the elevation of structural contours on the top of the and Teets, 1914).

Fig. 2. Stratigraphic framework showing (from left to right) North American subsystem, North American series, eastern European stages, central and western European series and stages, global eastern European stages, and lithostratigraphic units in West Virginia. Lithostratigraphic framework for Conemaugh is based on Fonner (1987),andMartino (2004) and relies on coal beds and marine units which are absent or limited in the study area. Twomile Limestone and Sandy Grove Sandstone are units that have been mapped in the Charleston area (Windolph, 1987). Ranges of selected palynomorphs are from Peppers (1996) and Eble et al. (2009). R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 139

3. Stratigraphic framework and coal geology cycles and climate variations associated with the eccentricity of the earth's orbit (e.g. Busch and Rollins, 1984; Heckel, 2008). The late Middle to Late Pennsylvanian Conemaugh Group/Forma- “ ” tion, known as the Lower Barren Measures due to the paucity of min- 4.3. Paleogeography able coals, extends from the top of the Upper Freeport coal to the base of the Pittsburgh coal (Wanless, 1939; Fig. 2). Where the Ames Limestone During the Pennsylvanian, the Central Appalachian Basin drifted is present, the Conemaugh Group is divided into the Glenshaw and northward with Larussia and was positioned within a few degrees of Casselman Formations (Fig. 2). The Ames Limestone and other marine the equator by the Late Pennsylvanian (Blakely, 2007; Rosenau et al., markers in Fig. 2 are well-developed in Wayne County, West Virginia, 2013a, Fig. 3). During the Middle and Late Pennsylvanian, rivers flowed but are absent throughout most of the study area where the Conemaugh west and north across West Virginia, draining the Allegheny Orogen. is treated as a formation (e.g. Henry et al., 1979, Fig. 72). In the Channel belts in the tropical coastal plain were flanked by flood basin Charleston, West Virginia area, the Conemaugh is 180 m thick (Krebs lakes and swamps (Arkle, 1974; Donaldson, 1979; Martino, 2004). and Teets, 1914). Where the Upper Freeport coal is not present, the Eight transgressions, marked by marine limestone and shale, are record- base of the Conemaugh is placed at the lowest occurrence of red ed in the Glenshaw Formation of Ohio and Pennsylvania (Busch and mudrocks (Rice, 1986). The Brush Creek, Bakerstown, and Harlem West, 1987). Six of these transgressions extended into northern and coals, locally mined in the northern Appalachian Basin, tend to be higher westernmost West Virginia, down dip from the current study area. in sulfur and ash compared to coals from the Lower and Middle Pennsyl- vanian (Cecil et al., 1985). 5. Sedimentary facies 4. Geologic setting 5.1. Facies 1 — large scale channel-fills 4.1. Tectonics 5.1.1. Description During the Middle–Late Pennsylvanian, thrust-loading in the Facies 1 consists of single and multistory bodies of micaceous sand- Appalachian Orogen from crustal collisions associated with the closure stones with erosional basal contacts (Fig. 5F). The thickness of single- fi of the Rheic Ocean induced basin subsidence and provided sediment story lls average 7.75 m and ranges from 4.1 to 11 m, with multistory fi fi fi accommodation space (Quinlan and Beaumont, 1984). Subsidence lls reaching a maximum of 26 m. The lls ne upward within a story was greatest in the foredeep of eastern West Virginia and decreased and are commonly interbedded with, or capped by siltstone, shale, northwestward toward the cratonic platform in Ohio and Kentucky. and occasionally coal. Conglomeratic channel lag deposits are some- Thrust loading alternated with relaxation producing transgressive– times present with siderite, limestone, coal spar, and vein quartz peb- fi fi regressive tectophases lasting several million years. Higher frequency bles. The lower portions of the lls are typically trough cross-strati ed fi glacioeustatic transgressive–regressive cycles are embedded within with sets from 20 to 50 cm thick. Compound cross-strati cation is com- – the tectophases (Busch and Rollins, 1984; Heckel, 1994; Ettensohn, mon with large scale sets from 1.5 to 7 m thick, and foreset dips of 8 11 fi 2008). At the beginning of the Late Pennsylvanian, deformational load- degrees. The upper portions of the lls are ripple cross-laminated or ing forced foreland basin subsidence to shift from the central to the parallel laminated. Cross-bed dip azimuths and trough axes are northern Appalachian Basin. The Brush Creek Limestone through the unimodal within a story and range from WSW to NNE. The vector Ames Limestone interval may represent an overall long-term transgres- mean is 315° and the magnitude, R, is 58% (Fig. 4). Plant are fi sion from the west in response to the early stages of this tectophase rare and include fossilized, mud- lled logs, and ferns. Mudstone lenses (Ettensohn, 2008). During the Late Pennsylvanian, accommodation up to 6 m thick and 70 m wide are locally present. The mudstones vary space was generally lower than during the Early–Middle Pennsylvanian from dark gray to greenish gray silty shales to red hackly, calcareous due to a decreased rate of tectonic subsidence (Greb et al., 2008). mudstones. Although most channel-fill sandstones contain unimodal, down-dip 4.2. Paleoclimate and eustasy oriented paleocurrent indicators, four locations in the study area have possible tidal structures. At K57, four bundles of thicker shale-draped During the Early to Middle Pennsylvanian, the climate in the central siltstone laminations alternate with thinner couplets. At BX22, a fi Appalachian Basin varied from tropical everwet to long wet season/ 9.75 m thick channel sandstone nes upward into rhythmically bedded, fi short dry season. The climate during the late Middle and Late Pennsyl- very ne sandstone-silty shale with horizontal laminations. Thicker vanian had significantly less rainfall, and fluctuated between humid bundles of sand laminae alternate with thinner bundles of silt laminae subtropical to semiarid (Donaldson et al., 1985; Cecil, 1990). An abrupt and carbonaceous shale. Four thick-thin cycles are locally developed climate change occurred near the Westphalian–Stephanian boundary over a 50 cm interval. At K36, the top 10 cm of a compound cross- fi fi and was characterized by global warming, stronger seasonality, and strati ed channel- ll is burrow mottled. At K63, 7 m of coarse, trough fi shorter wet phases. It coincided with the regional extinction of tree cross-strati ed sandstone are overlain by 3.1 m of inclined heterolithic lycopsids and the appearance of widespread, red vertisols and aridosols. strata. Aridity was greatest during deposition of the Glenshaw Formation (Cecil, 1990; Pfefferkorn et al., 2008). 5.1.2. Interpretation Widespread ice sheets developed in Gondwana at the beginning of Facies 1 is interpreted as the product of alluvial, upper estuarine, and the Stephanian and extended to within 25 degrees of the equator possibly upper delta plain distributary channels and is similar to those (Fielding et al., 2008; Isbell et al., 2003). Busch and Rollins (1984) recog- previously described from downdip areas (Martino, 2004). This inter- nized 11 paleosol-bounded cyclothems in the Glenshaw Formation in pretation is consistent with the unimodal paleocurrents and vector Ohio and Pennsylvanian. These have been correlated between the Ap- mean northwest flow which is down the regional paleoslope. The R palachian and Illinois Basins using conodonts and palynomorphs. Re- value of 58% may reflect moderate channel sinuosity and radiating del- gional analysis indicates that only the greater highstands invaded the taic distributaries. The local occurrence of burrowing, shale-draped Appalachian Basin due to its high shelf position (e.g. Busch and West, foresets, and rhythmic laminations may indicate fluvial channels at 1987; Heckel, 1995, 2008); however, even these marine units are diffi- the fluvial–tidal transition of the upper estuary (Greb and Martino, cult to trace updip into the study area. Major cyclothems had an average 2005). Upper delta plain distributaries differ from alluvial facies in duration of 400 kyr and have been attributed to glacioeustatic sea level having smaller depth and width, and more widespread lakes and 140 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 swamps in overbank areas (Reading and Collinson, 1996; Olariu and 5.4.2. Interpretation Bhattacharya, 2006). Facies 4 is interpreted as the deposits of flood-basin lakes (or possi- bly bays) where low energy deposition of mud from suspension oc- curred. Red clays may have been derived by erosion of well-drained 5.2. Facies 2 — small scale channel-fills upland soils that were deposited in the lake waters. Rather than indicat- ing subaerial exposure during deposition, ferric iron-stained clays were 5.2.1. Description preserved in the lakebed where there was insufficient organic matter to Facies 2 consists of very fine to fine, greenish gray, micaceous, calcar- reduce the iron. Fissility in shales is prevalent where flocculation of clays eous sandstones 0.5–4 m thick. Scour-fill trough cross-stratification was inhibited (Moon and Hurst, 1984). Slickensides may have formed predominates with 10–50 cm thick sets. The channel-fills fine upward from periods of subaerial exposure and shrink and swell during incipi- into siltstone and shale. Lateral dimensions are 3–15 m. Ripple cross- ent pedogenesis, or by dewatering of subaqueous, clay-rich sediment lamination and parallel lamination are common toward the top of the and clay flocculation and shrinkage during syneresis (White, 1961; fills. Plant fossils are rare. The channel-fills typically overlie thin- Potter et al., 1980). bedded mudstones and shales of facies 4. Lake levels in coastal plains are usually less than 5 m deep (Elliott, 1974; Reading and Collinson, 1996) and lake levels are vulnerable to 5.2.2. Interpretation seasonal and longer term variations in climate. Upward-coarsening of Facies 2 is interpreted as the product of crevasse splay channels. Basal the lacustrine shale into siltstone reflects shallowing and infilling of scours, fining-upward, and less-than-meter-scale thicknesses are consis- the lake by the distal margins of encroaching levees and crevasse splays tent with a flood basin origin (Elliott, 1974; Reading and Collinson, 1996). (Elliott, 1974; Horne et al., 1978; Reading and Collinson, 1996). Crevasse channel and splay lobe deposits that fine upward and have erosional bases reflect waning flow from flooding that breached the le- 5.5. Facies 5 — dark gray–black, molluscan shale/claystone vees of fluvial and deltaic channels (Collinson, 1996; Leeder, 2011). Splays deposits contribute to the infilling of flood basin lakes (Collinson, 1996). 5.5.1. Description Facies 5 is found at only three localities in Braxton County and over- lies the Brush Creek coal. The dark gray to black shale/claystone facies is 5.3. Facies 3 — sheet sandstone and shale 30–50 cm thick but fossils are restricted to a thin zone of 2–3 cm. The fossils include the gastropod Bellerophon, several bivalve genera includ- 5.3.1. Description ing Phestia and a nuculid, an Orbiculoid brachiopod, and plant fragments Facies 3 is typically 5–6 m thick and consists of interbedded fine (Cordaites). The facies contains siderite nodules and thin beds and sandstone to shale. Sandstones are sharp based with scour-fill, trough coarsens upward into siltstone and very fine sandstone. cross-strata that grade up into ripple cross-lamination and parallel lam- ination. There are typically four or five fining-upward sandstone-shale 5.5.2. Interpretation bedsets ranging from 1 to 1.5 m thick. The couplets are laterally persis- Facies 5 was formed in a low-energy, nearshore marine setting as tent across the outcrops for at least 100 m up to a km or more. Small rising sea level drowned a coastal swamp. The types of invertebrates in- channel-fills 1–2 m thick and 5–10 m wide locally cut into the top of dicate normal to restricted marine salinity (Boardman et al., 1984). The the sandstones. Root traces and plant fossils are occasionally present. high organic content of the shale may be due to proximity to coastal Trace fossils are rare and include sand-filled vertical burrows peat swamps and limited clastic influx as sediments were initially 2–10 mm wide occur in parallel laminated, very fine sandstone (K38), trapped in estuaries until they filled. Siderite nodules and beds form and a large Diplichnites trackway (K44, Fig. 3A). during early diagenesis under reducing conditions and are common in organic-rich mudrocks deposited in fresh- and brackish-water settings (Tucker, 2001). Siderite beds may be formed by the sudden influx of 5.3.2. Interpretation freshwater into more saline waters (Woodland and Stenstrom, 1979). Crevasse-splay lobes with erosional bases and textures and struc- tures reflect waning flow from flooding that breaches the levees of flu- 5.6. Facies 6 — coal/carbonaceous shale vial and deltaic channels (Collinson, 1996; Leeder, 2011). They have a sheet-like to lobate geometry that can extend up to several kilometers 5.6.1. Description away from the main channel, and are typically finer grained than the Facies 6 consists of seams of coal, bone coal, and carbonaceous shales main channel-fill. Successive flood events are capable of producing up to 1.75 m thick and usually overlies root-traced, hackly mudstone or meter-scale, stacked bedsets of sand-shale like those described above. light gray to white underclays (Facies 7). Roof rock is most often a dark The large Diplichnites trackway was made by Arthropleurid myriopods gray shale with plant fragments. In some cases, it consists of hackly in sand and mud flats that bordered ephemeral flood-basin lakes mudstone with crude, thin bedding and slickensides. Plant fossils in- (Martino and Greb, 2009). clude locally abundant Cordaites and occasionally Neuropteris and Annularia. Smooth-shelled ostracods are present at R2, and abundant 5.4. Facies 4 — thin-bedded mudstone/claystone/shale Conchostracans occur at C7. Coals rarely occur near the top of channel-fills interbedded with dark gray shales. Coal seams typically 5.4.1. Description contain thin interbeds of carbonaceous shale and hackly mudstone. Lat- Facies 4 is widely developed and 1–8 m thick. Mudstones display eral extent is often limited by channel washouts or coal pinchouts. crude, thin bedding or are hackly. Slickensides are locally developed. Olive green and dark greenish gray shades predominate, but dark 5.6.2. Interpretation gray, red, and variegation (red/green/gray mottling) also occur. The facies The coals that overlie underclays or root-traced, hackly mudstones is commonly darker at the base and coarsens upward in to silty shale and accumulated in coastal peat swamps. The shale roof facies with plant siltstone; it typically overlies facies 6, 7, or 8 and is capped by facies 3 or 6. fossils but lacking root traces indicates that swamps became lakes as Thin beds and nodules of argillaceous, micritic limestone and siderite the water table rose and became too deep for standing vegetation. are locally present. Plant fossils commonly include Cordaites, and rarely Channel-fill coals accumulated in deactivated channels following avul- Neuropteris, Alethopteris, Pecopteris, and Sphenopteris (K59). At K65, sion. Mudstone partings resulted from clastic influx from alluvial chan- Sigillaria is preserved in situ (Fig. 3B). nels, most likely due to catastrophic flooding, perhaps at 500- to 1000- R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 141

Fig. 3. A. Diplichnites trackway 23 cm wide preserved as convex hyporelief. Sample in float from very fine grained, flaggy sandstone from crevasse splay facies at K44. B. Arborescent lycopod Sigillaria preserved as mudstone cast in situ above protosol at K65. C. Flint clay facies (white) at Clendenin Exit of I-79 (K57) pedogenically overprinted by calcic vertisol. Staff is 1.5 m. D. Flint clay lens (white with red margin) of lacustrine origin from I-79 roadcut northeast of Amma (R8). The lens grades laterally into thin bedded mudstone. 1.1 m of staff showing for scale. Bedding is chaotic between and above flint clay bodies at this location. E. Two benches of lacustrine/palustrine limestone and intervening mudstone overlying variegated, mottled, calcic vertisol along Rt. 119 near Priestly (L3). F. Trough cross-stratified, ostracodal grainstone from upper limestone bench in E. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) year intervals (using a tropical peat accumulation rate from modern an- and weather light gray to white, or yellow-orange if pyritic or sideritic. alogs of 0.25 mm/year; Cecil et al., 1993). Cordaites were most common The facies is up to 1.9 m thick and consist of individual beds 5–50 cm along the margins of swamps and expanded into their interior during thick interbedded with olive to dark gray claystone, mudstone, and oc- ecologic stress (Eble, 1998). Ostracods with thin smooth shells are usu- casionally coal. Allochems include pelloids, limeclasts, ostracodes, ally found in freshwater facies; Conchostracans are found mainly in spirorbid microconchids, fish bones and teeth, coprolites, and blue- freshwater to brackish facies which may interfinger with marine facies green algae (Fig. 6). Algal laminations are common, and the tops of or pinchout over short distances (Tasch, 1980). Carbonaceous, fissile some beds are brecciated. Birdseye, laminar, and tubular fenestrae are shales lacking root traces formed in anoxic lakes or bays bordering the common. The carbonates nearly always occur immediately above a swamps where clays mixed with macerated organics. hackly mudstone (paleosol) and are overlain by Facies 4 (Figs. 3E, 5A, C). At K44 and K45, intraclastic rudstone with rounded ripups and 5.7. Facies 7 — limestone/claystone bone fragments fill shallow channels up to 50 cm deep and 10–20 m wide which are cut into the underlying paleosol (Fig. 5D). 5.7.1. Description Facies 7 encompasses a range of carbonate lithologies; subfacies in- 5.7.2. Interpretation clude mudstones, peloidal packstones and wackestones and locally, Facies 7 accumulated in lacustrine, palustrine, and, possibly, near- rudstones and grainstones. The limestones are medium to dark gray shore bay environments. The deposition of the lime mudstones, 142 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154

5.8. Facies 8 — flint clay

5.8.1. Description Flint clay occurs in northeastern Kanawha, Roane, and Braxton Counties. The kaolinite-rich claystone is brittle with conchoidal fracture and is predominantly red or white with subordinate mottles or lenses of white, red, gray, and green. The thickness is usually 30–100 cm and varies from 5 to 290 cm. Two facies associations occur. (1) In most cases, the flint clay occurs as a separate interval within the variegated hackly mudstone facies (Facies 9, pedotype D; Fig. 3C). Internal struc- ture is massive, mottled, and occasionally convoluted. Peds and cutans are commonly present. (2) In some outcrops, the flint clay is thin- bedded or laminated and occurs at the top of variegated hackly mud- stone or carbonaceous shale (Fig. 3D). The flint clay fills small, shallow channels cut into the underlying hackly mudstones. At some locations, it is brecciated with angular, granule- to pebble-sized clasts occurring in a claystone matrix and is overlain or interbedded with laminated shale or micritic limestone.

5.8.2. Interpretation Flint clays can form in soils during pedogenesis when acidic waters from swamps remove silica, iron, and alkaline/alkali cations through di- Fig. 4. Paleocurrent rose based on 117 measurements of foreset dip azimuths and trough alysis hydrolysis, converting illite to a colloidal gel that crystallizes to axes from 3-D exposures of cross-stratification in fluvial-upper estuarine channel kaolinite. Flint clays may also be deposited by selective flocculation of sandstones (Facies 1). colloidal kaolin in paludal–lacustrine transitions where the pH changes from acidic to basic (e.g. Williams et al., 1968; Staub and Cohen, 1979; Keller, 1981; Bragonier, 1989). A depositional origin is likely where wackestones, and packstones occurred in a low-energy aquatic setting, the flint clay facies lacks peds and cutans, or where it overlies a paleosol whereas the intraclastic rudstones and cross-laminated ostracod or carbonaceous shale and is overlain by laminated silty shale or micritic grainstones indicate traction transport by short-lived flows. Lacustrine limestone. As the paleosol or swamp was drowned, the lake margin carbonates are formed mainly by biochemical or biologically mediated would have provided the ideal zone of mixing for alkaline and acidic precipitation. Uptake of carbon dioxide by algae and phytoplankton in- waters, allowing for flocculation of kaolin colloids (e.g. Keller, 1981; creased pH which facilitated inorganic precipitation of micrite. Fenestral Bragonier, 1989). These beds are often brecciated due to water loss dur- cavities are characteristically developed in the littoral and supralittoral ing exposure and dessication. In western Pennsylvania, laminated flint zone from trapped gas bubbles and dessication, decaying algal mats, clay occurs laterally between the Upper Freeport coal and nonmarine and burrows or, possibly, decaying roots (Tucker, 2001; Pratt, 2010). limestone and shales, reflecting lateral facies changes from paludal to la- Palustrine carbonates form by pedogenic overprinting of nearshore custrine environments (Bragonier, 1989). The flint clay facies that occur deposits of extremely shallow lakes with fluctuating lake levels and sandwiched between paleosols have sharp contacts and a lens-like ge- densely vegetated shorelines (Freylet and Plaziat, 1982; Platt and ometry; these are also likely to be lacustrine in origin and represent rel- Wright, 1991). ict bedding in the parent material preserved within the paleosol profile. The facies relations of most carbonates directly above paleosols sug- gests they formed by drowning of stable landscapes by lakes or brackish 5.9. Facies 9 — pedotypes bays. The same low clastic influx that favored paleosol development would also have facilitated the development of carbonates. The low tur- Paleosols were described following the criteria of Retallack (1988). bidity benefitted suspension-feeding Micronconchids and blue-green Mack et al. (1993) developed a predominantly descriptive classification algae, and prevented dilution of carbonates by siliciclastic sediment. of paleosols that includes 9 orders called pedotypes. This system was During seasonally low lake levels, exposure can cause brecciation of needed because of the limitations in matching modern soils where the lacustrine mud substrates, and storms can cause runoff in shallow climatic, textural, and mineralogic criteria for soils classes are known channels along lake margins producing intraclastic conglomerates to their ancient counterparts where this information is often obscured (Platt and Wright, 1991). or impossible to interpret with a high degree of confidence. In this Spirorbiform microconchids are semi-infaunal, suspension-feeding study, six pedotypes (A–F) were distinguished. In addition, polygenetic lophophorates that have been reported from Pennsylvanian marine, and compound paleosols were also identified (Fig. 7). The stratigraphic brackish and freshwater settings (Taylor and Vinn, 2006; Zatoń and context of the pedotypes is illustrated in Figs. 8–9). Peck, 2013). Their presence indicates a marine connection and probable brackish influence (Schultz, 2009; Gierlowski-Kordesch and Cassle, 5.9.1.1. Pedotype A — description 2015). Two occurrences of microconchid-bearing ‘nonmarine’ lime- Pedotype A consists of coal or bone coal up to 1.0 m thick. The Upper stones in the Glenshaw Formation have been re-interpreted as brackish, Freeport coal at K40 consists of 82 cm of bright laminated coal overlain clear water, nearshore facies (Morris, 1967; Busch and West, 1987)In by 18 cm of bone coal and 35 cm of carbonaceous mudstone with slick- West Virginia, the Cambridge Limestone was found to have an open- ensides, abundant Cordaites,andAnnularia and Neuropteris (Figs. 7A, 8). marine fauna that graded landward into an assemblage containing The contact with overlying dark gray shale is sharp. The seatrock is only Spriorbid micronconchids and a few ostracods (Morris, 1967). 1.1 m thick and begins with dark gray mudstone with crude thin bed- The Noble Limestone in southeastern Ohio contains an open-marine ding and root traces that becomes light gray in the upper 30 cm. crinoidal biofacies that grades landward into a fenestral limestone biofacies with abundant Spirorbid microconchids and ostracods. The 5.9.1.2. Pedotype A — interpretation limestone has been interpreted to have formed in intertidal to Coal represents a histic epipedron, and is a Histosol (Mack et al., supratidal ponds adjacent to the Noble Seaway (Busch and West, 1987). 1993). Histosols require the accumulation of peat in poorly drained, R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 143

Fig. 5. A. Thick vertisol with conjugate shears (arrows) capped by lacustrine limestone and shale from K56 near Elkview Exit of I-79. Staff = 1.5 m. B. Microconchid packstone with shark’s tooth from Twomile Limestone at K45 east of St. Albans. C. Twomile Limestone interbedded with olive green flint clay above thick, polygenetic, calcic vertisol in its type area near Guthrie (K43). D. Shallow channel-fill limestone incised into vertisol near St. Albans (K45). E. Rudstone channel lag, I-77 north, Edens Fork Exit (K44). F. Multistory channel sandstone from amalgamation of Upper and Lower Mahoning Sandstone IVFs along I-79 at K47. Channel-fills (arrows) are 8–9mthick. anoxic environments where microbial decomposition is inhibited and 10–35 cm, green; 35–85 cm, red/green variegated; 85–2.65 cm, red; where influx of clastics is limited (Retallack, 2001). The vertical succes- and 2.65–2.9 cm, green. sion of lithologies at K40 indicates the development of a planar mire with highly acidic, anoxic conditions which was drown by rising water 5.9.1.4. Pedotype B — interpretation table. Portions of the mire that were increasingly more proximal to Pedotype B is classified as a Vertisol. Vertisols are distinguished by the lake received higher rates of clastic influx forming bone coal and thick accumulations of smectitic clays that expand and contract with carbonaceous shale. wetting and drying. Large scale, downward tapering cracks may form. These form in sparsely wooded areas of low relief with subhumid to 5.9.1.3. Pedotype B — description semiarid climates with a prolonged dry season (Retallack, 2001). Pedotype B paleosols range in thickness from 0.5 to 3.0 m. The up- Thick argillic horizons (Bt) in excess of 1 m are characteristic of strongly permost horizon is a thin carbonaceous mudstone which overlies dark to very strongly developed soils suggest prolonged periods of landscape greenish gray mudstone grading down into red mudstone/claystone. stability on the order of at least 105 years (Retallack, 2001). Soil structure includes angular blocky peds, slickensides, and conjugate shears. An example of Pedotype B at K55 consists of a 2.9 m thick, hackly 5.9.1.5. Pedotype C — description mudstone containing angular blocky peds, slickensides, and steeply in- Pedotype C is similar to Pedotype B, but contains abundant micritic clined fractures (Fig. 7B). The base of the paleosol is transitional with limestone nodules in red, or variegated mudstone up to 6 m thick thin-bedded greenish gray mudstone. The top is sharply overlain by (Figs. 3E, 5A, D). At K58, an example occurs that is 4.67 m thick dark gray shale. Colors from top down are: 0–10 cm, medium gray; (Fig. 7C). The upper 2.5 m consists of red hackly mudstone with angular 144 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154

Fig. 6. Photomicrographs. All at 40×, PPL. Scale bar = 500 microns. A. Spar-filled microconchid (above arrows) in peloidal, micrite matrix, Twomile Limestone near West Hamlin (L12). B. Disarticulated ostracods (arrows) in micrite matrix (L12). C. Pelodal dismicrite (P-pelloids), clotted fabric, with a few ostracods (L12). D. Ostracod shells (O), blocky phosphatic bone fragments (P), and spar-filled birdseye structure (BE) from Twomile Limestone near West Hamlin (L12). E. Blue-green algae (BG). Longitudinal section through bundles with long, narrow (2.5 micron) filaments with a thin micritic wall. Spar-filled tubular fenestrae (TF) in upper left. Twomile Limestone near West Hamlin (L1). F. Blue green-algae (arrows) with scalloped outline and micritic texture (L1). blocky peds, slickensides, and steeply inclined fractures filled with at the top and overlies dark gray claystone which becomes light olive hematite/limonite. The lower 2.17 m has similar structures and is gray toward the base. The profile contains abundant root traces and sev- variegated red/greenish gray with micritic limestone nodules. Its base eral micritic and sideritic nodules, as well as slickensides, angular blocky is transitional with olive, thin-bedded siltstone. peds, and large-scale inclined fractures.

5.9.1.6. Pedotype C — interpretation 5.9.1.8. Pedotype D (Gleysols) — interpretation Pedotype C is classified as a calcic vertisol (Mack et al., 1993). Calcic Pedotype D is classified as a Gleysol. Gleysols form where the water vertisols that have a concentration of carbonate nodules within the B table is consistently high, or predominantly high throughout most of the horizon (Bk) are indicative of at least semiarid conditions lasting year, producing low redox conditions. Low chroma colors such as gray 104 years (Retallack, 2001). or green are characteristic, and carbonaceous horizons are common (Mack et al., 1993). 5.9.1.7. Pedotype D — description Pedotype D consists of dark greenish gray, hackly mudstones that 5.9.1.9. Pedoype E — description grade up into dark gray mudstones and is restricted to the Upper Free- Pedotype E has a shallow Bk or K horizon within 1 m of the surface of port to Brush Creek coal interval. An example of Pedotype D from K37 the profile. A 2.9 m thick example occurs at K39 which has a sharp upper consists of an 80 cm thick claystone sharply overlain by greenish gray contact with red shale and a base that is transitional with underling shale (Fig. 7D). A black carbonaceous claystone band 1 cm thick occurs greenish gray silty shale (Fig. 7E). The upper 45 cm consists of red R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 145

Fig. 7. Pedotypes. A: histosol, Upper Freeport coal (K40); B: vertisol (K55); C: calcic vertisol (K58); D: gleysol (K37); E: calcisol (K39); F: protosol (K37); G: polygenetic paleosol, Brush Creek coal horizon (BX5); H: compound paleosol, 3 profiles each capped by carbonaceous mudstone (K55). Munsell color designations (hue, chroma) are: 1) red: 5R 4/6 to 10R 3/3; 2) greenish gray: 5G 5/1 to 5GY 6/1; 3) dark greenish gray: 5GY 4/1; 4) olive: 5Y 4/3; 5) dark gray: N3; 6) medium gray: N5; 7) light gray: N7. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) hackly mudstone with angular blocky peds and slickensides. From 45 to development, indicate these are in the strongly developed stage 85 cm, light gray micritic limestone nodules coalesce to form an irregu- (Retallack, 1988, 2001). lar bed (K horizon). The lower 2.05 m consists of red hackly mudstone as at the top, but with steeply inclined fractures filled with micrite. Frac- 5.9.1.11. Pedotype F — description ture margins are reduced to green. Pedotype F consists of hackly mudstones with crude, thin bedding. Soil profiles are usually thin and lack well-differentiated horizons. 5.9.1.10. Pedoype E — interpretation Root traces are preserved in greenish gray to dark gray varieties but Pedotype E is classified as a calcisol based on the prominence of its are absent in red and variegated types. Two examples occur at K37 calcic horizon (Mack et al., 1993). The proximity of the calcic horizon (Fig. 7F). The lower one is 20 cm thick with crudely bedded greenish to the surface suggests this pedotype is an Aridosol (Retallack, 1988, gray siltstone that grades up into hackly mudstone with root traces. 2001). Aridosols are formed in semiarid to arid regions where limited The upper one is similar, but is 60 cm thick. rainfall is unable to completely leach soluble minerals very far down- ward, resulting in densely packed or coalescing carbonate nodules 5.9.1.12. Pedotype F — interpretation close to the surface at the depth of the average wetting (Retallack, Pedotype F is classified as a Protosol (Mack et al., 1993). Protosols in- 2001). This stage of carbonate accumulation requires pedogenesis clude weakly developed paleosols ranging from Entisols (incipient soil) over 105 years. The thickness and distinctness of the horizons in the Inceptisols (young soils) Mack et al., 1993; Retallack, 2001). Entisols pedotypes B, C, and D, especially the thick (N1.5 m, after compaction), reflect minimal soil development. Stratification is usually well- red, argillic horizons, and calcareous horizons with Stage 2–3 preserved. Root traces are often present and organic matter may 146 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154

Fig. 8. Cross-section of composite sections through southern portion of study area (Fig. 1) from northern Wayne County eastward to Priestly along Rt. 119, then northward to the Charleston area. Location numbers at top of stratigraphic columns show central location for composite section which also incorporates nearby locations. Paleosols representing interfluvial sequence boundaries are numbered from 1 (base of Upper Freeport Coal) to 9 (base of Harlem Coal) after Martino (2004, Fig. 12). Horizons 2 and 3 are poorly developed in the study area due to extensive development of the Mahoning Sandstone between the Upper Freeport and Brush Creek coals. The Wayne County section is based on outcrops in northern Wayne County (Martino, 2004) and is included to show correlation of downdip, marine-cored cyclothems with nonmarine cyclothems updip to the east. Pedotypes (A–G) are labeled in parentheses. Asterisks indicate pedotype used in Fig. 7; not all pedotype examples in Fig. 7 are included in composite sections. IFSB paleosols are split into a and b divisions where avulsion locally introduced a lens of clastics that interrupted pedogenesis. accumulate at the surface. Inceptisols represent slightly longer The upper 30 cm consists of carbonaceous shale and carbonaceous pedogenesis than Entisols, which obscures or obliterates primary depo- claystone, which overlies 3.1 m of greenish gray, hackly mudstone with sitional features. Such soils are typical of short periods of time for forma- shrink-swell features including angular blocky peds and inclined frac- tion due to frequent clastic influx as in floodplains that are proximal to tures. Large, steeply inclined, elongated siderite nodules are present in active drainage lines (Retallack, 2001). the upper 1 m of the mudstone; this interval weathers yellow-orange. A lateral transition of the 3.1 m interval occurs from greenish gray to 5.9.1.13. Polygenetic paleosols — description patchy areas of variegated (dull red, gray, and green), hackly mudstones. A polygenetic paleosol has attributes that point to contrasting pedo- genic conditions (e.g. Driese and Ober, 2005; Rosenau et al., 2013a,b). 5.9.1.14. Polygenetic paleosols — interpretation An example at BX5 is 3.4 m thick; it is transitional with underlying gray The paleosol in Fig. 7G records changing paleohydrologic conditions shale and sharply overlain by crudely bedded olive mudstone (Fig. 7G). that may be linked to climatic and sea level changes. The paleosol R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 147

Fig. 9. Cross-section of composite sections through northern portion of study area from Guthrie (just north of Charleston) to Flatwoods.

contains angular blocky peds, pedogenic slickensides, conjugate shears dark gray to black carbonaceous shale or carbonaceous hackly mud- that indicate a well-drained phase of pedogenesis and the initial stone. The upper paleosol is 2.2 m thick; 1 cm thick carbonaceous development of a vertisol which likely had high chroma coloration. A shale overlies 30 cm of green hackly mudstone which occurs above a subsequent rise in water table produced poorly drained, low Eh, low 1.88 m of variegated (red/green) mudstone with micritic limestone pH conditions that led to gley overprinting, producing low chroma nodules. This upper profile, as well as the other two, contain slicken- colors in the upper portion of the profile due to Fe reduction. These con- sides and angular blocky peds. ditions also promoted siderite precipitation which may have formed The middle profile is 1.45 m thick and consists of 15 cm of dark gray around decaying roots (rhizoconcretions), and anaerobic conditions hackly mudstone above variegated (red–green–yellow) mudstone. The which facilitated the preservation of organic matter. Similar polygenetic lower profile is a 75 cm thick hackly mudstone that is dark gray in the paleosols in Pennsylvanian have been reported elsewhere in the Appa- upper 20 cm and passes downward to red and gray variegation, then be- lachian and Illinois Basins (Gardner et al., 1988; Driese and Ober, 2005; comes green at the base. Rosenau et al., 2013a). IFSB paleosols in the Conemaugh are polygenet- ic; they formed when well-drained, high-chroma vertisols, calcic 5.9.1.16. Composite paleosols — interpretation vertisols and calcisols were subsequently gleyed in their upper part dur- A composite paleosol profile consists of stacked paleosols that over- ing high water table conditions (see Section 7.3). lap, whereas compound paleosol profiles are separated by thin beds of nonpedogenically modified sediment (Wright, 1992). Both composite 5.9.1.15. Composite paleosols — description (Fig. 7H) and compound paleosols occur in the study area and indicate A 4.5 m thick composite profile occurs at K55 recording three that pedogenesis was locally interrupted by influx of sediment (Wright, stacked paleosols (Fig. 7). The top of each paleosol is demarcated by 1992). 148 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154

Fig. 10. Sequences, systems tracts, and stratigraphic surfaces defined in relation to the base-level and the transgressive–regressive curves. SU—subaerial unconformity; c.c.—correlative conformity; BSFR—basal surface of forced regression; MRS—maximum regressive surface; MFS—maximum flooding surface; (A)—positive accommodation (base-level rise); NR—normal regression; FR—forced regression; LST—lowstand systems tract; TST—transgressive systems tract; HST—highstand systems tract; FSST—falling-stage systems tract; RST—regressive systems tract; TR—transgressive–regressive sequence; IFSB—paleosol formed during negative accommodation on interfluves. Late Pennsylvanian major cyclothems have average duration of 400 kyr. Modified from Catuneanu (2006).

Fig. 11. Nonmarine sequence stratigraphic model for fourth order cyclothems of the Conemaugh Formation in central West Virginia. R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 149

Fig. 12. Three stage model for development of IFSB paleosols in the Conemaugh Formation. Modified from Driese and Ober (2005, Fig. 9).

6. Stratigraphy Lycospora is in the Upper Freeport coal; it is greatly reduced in the Mahoning and is absent above it (Peppers, 1996; Eble et al., 2009). 6.1. Results Where the Upper Freeport is absent, the lowest occurrence of wide- spread red mudstone is used to mark the base of the Conemaugh Eleven composite sections were developed from individual outcrops (Rice, 1986). In the Huntington area (Fig. 1), the lowest red mudstone for the following areas: northwestern Lincoln County (West Hamlin), is typically 15 m above the Upper Freeport coal. The Brush Creek coal oc- northeastern Lincoln County (Priestly), westernmost Kanawha County curs 28 m above the base of the Upper Freeport coal. It is distinguished (Ruth, St. Albans), north Charleston (Eden Fork, Mink Shoals), north- from the Mahoning coal below by the absence of Lycospora, and from eastern Kanawha County (Big Chimney, Clendenin), and southern the Bakerstown above by the presence of Laevigatosporites globosus Roane and Braxton Counties (Amma, Wallback, Flatwoods; Figs. 1, 8, and Punctatosporites granifer (Peppers, 1996; Eble et al., 2009). The 9). The tops of thick, regionally persistent paleosols were used in con- Brush Creek coal to Bakerstown coal interval is 24.5 m; the base of the junction with the Upper Freeport and Brush Creek coals to correlate Ames Shale/Limestone marine unit occurs 48.5 m above the Brush the sections. An additional composite section was developed for north- Creek coal (Martino, 2004; Fig. 8); the interval between the Pittsburgh ern Wayne County where marine-cored cyclothems are present coal and the Ames Limestone is 90 m in Cabell and Putnam Counties (Martino, 2004). (Fonner and Chappell, 1987; Fonner, 1987). The Pittsburgh coal is dis- tinguished from coals above and below it by the Thymospora theissenii 6.2. Stratigraphic correlation epibole (Kosanke, 1988; Eble et al., 2009). Early workers (e.g. Krebs and Teets, 1914) mistook what is likely the The Upper Freeport coal has been mined in the Hamlin, McCorkle, Brush Creek coal for the Bakerstown coal in the study area. At K50 near and Sutton areas, but is otherwise thin or absent. In the Appalachian Guthrie (Fig. 1), this coal occurs 136.6 m below the Pittsburgh coal (as Basin, palynologic analyses indicate the last abundant occurrence of mapped by the WVGES; Fig. 9), 20.7 m lower than its expected position 150 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154

Fig. 13. Paleosol/limestone cycles at K44, Eden Fork exit of I-79 N. This outcrop is located in an interfluvial portion of the coastal plain where IVFs are absent. The Twomile Limestone is exposed at base of the outcrop. Inverted triangles indicate coarsening upward trends truncated by fluvial and crevasse channels of the HST. Composite stratigraphic column for this location is in Figs. 8 and 9.SeeFigs. 10 and 11 for sequence stratigraphy.

if it were the Bakerstown. In Braxton County (BX2, 7, 19; Fig. 1), the 7. Sequence stratigraphy “Bakerstown” coal of Hennen and Gawthrop (1917) occurs 30.5 m above the Upper Freeport coal (Fig. 9), 22 m lower than expected if it 7.1. Climatic, tectonic, and eustatic controls were the Bakerstown, and it is overlain by a marine shale that likely correlates to the Lower Brush Creek marine shale (Figs. 2, 9). Sequence stratigraphic models have been developed for The Twomile Limestone in its type area (K43, K44, K50, Fig. 1)was nonmarine fluvial successions (e.g. Shanley and McCabe, 1993, apparently mistaken for the equivalent of the Ames Limestone (Krebs 1994; Wright and Marriott, 1993; Catuneanu, 2006, Fig. 2.19; Miall, and Teets, 1914); its average position in this area is 119.2 m below the 2010, Fig. 6.10), and models have been developed for Late Pennsyl- Pittsburgh coal, 29 m lower than expected (Figs. 1, 8). Keroher et al. vanian cyclothems in equatorial settings (e.g. Falcon-Lang, 2004; (1966) and Rice et al. (1994) placed the Twomile Limestone above the Martino, 2004; Dolby et al., 2011). Sequence boundaries within the Bakerstown coal below the Morgantown Sandstone. In contrast, Henry Conemaugh formed during falling base level and are marked by et al. (1979) and Outerbridge (1989) tentatively correlated the high relief unconformities at the base of incised valleys and strongly Twomile with the Brush Creek Limestone. Kosanke (1988) reported developed paleosols on the interfluves. They developed during Laevigatosporites globosus from three assemblages north of Charleston glacioeustatic lowstands associated with the expansion of Gondwa- in the Twomile type area. These include sample 502-C from 23 m na ice sheets (Martino, 2004). below the Twomile Limestone (1.6 km SE of K47), sample 852 I-M at ap- Aggradation and degradation of fluvial systems depend on the inter- proximately the same stratigraphic position as the Twomile Limestone play between the rate at which accommodation is produced vs. sedi- (1.1 km NNW of K44), and sample 462 from 14 m above the Twomile ment supply and energy flux of the depositional system (e.g. Miall, Limestone (0.75 km SW of K47). These results, along with the absence 2010). In the proximal (upstream) portions of rivers, episodes of of Lycospora in the samples, indicate that the Twomile Limestone occurs downcutting and aggradation are controlled by climate and tectonics between the Mahoning and Bakerstown coal horizons. The ‘Bakerstown’ which cause variations in discharge and sediment load (Miall, 2010). coal of Krebs and Teets (1914) that occurs 6 m below the Twomile at Climatic and tectonic cycles may operate independently from base K50 and up to 13 m below it elsewhere is in the Brush Creek–Wilgus level/sea level cycles (Blum and Tornqvist, 2000; Miall, 2010). Sea coal interval. Since this coal occurs 15 m above the lowest red bed mud- level cycles influence aggradational/degradational cycles of river sys- stones, it is probably the Brush Creek coal, a placement also suggested tems in their downstream reaches. In the Mississippi River, this influ- by Windolph (1987). Based on this analysis, the Twomile Limestone is ence is seen 220 km from the mouth (Miall, 2010) and could reach as correlated with the Upper Brush Creek Limestone reported by Martino far inland as 300–400 km for low gradient streams with a high sediment (2004) in the Huntington area (Figs.1,2,8). supply (Blum and Tornqvist, 2000). R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 151

The outcrops in this study are 28 km or less from the projected transgressive systems tracts (McCarthy and Plint, 1998; McCarthy paleoshorelines during the maximum transgression of the Lower et al., 1999; Fig. 10). Lowstand paleosols are relatively thick with well- Brush Creek and Ames Seas (Fig. 1). The Upper Brush Creek, Cambridge defined horizons. Exposure time for Milankovitch driven eccentricity and Portersville transgressions reached to within 50–80 km or less of all cycles is usually 104 to 105 years, and the soils are usually well- outcrops (Busch and West, 1987). A total of eight Glenshaw transgres- drained and of regional extent due to lowered base level (Miall, 2010). sions are recognized in West Virginia and/or eastern Ohio and south- Climate and base level cycles may be interdependent at the time scale western Pennsylvania (Busch and Rollins, 1984; Busch and West, of Milankovitch periodicities (Catuneanu, 2006). Major cyclothems 1987). It is therefore likely that sea level/base level changes played a have a duration of about 400 kyr (Heckel, 2008), and IFSB paleosols in dominant role in the development of sequences and systems tracts of nonmarine cyclothems of this study may represent much of this time the Conemaugh updip from these marine locations in central West (Fig. 10). Virginia. In Figs. 8 and 9, well-developed paleosols interpreted as IFSBs are capped by limestones, coals or carbonaceous shales. Good examples 7.2. Incised valley-fills (IVFs) are illustrated in Figs. 5A, C, and D and 8 (Eden Fork and St. Albans sec- tions, horizons 4–9). The close association of well-drained, soils Valley fills are elongate bodies larger than a single channel that (vertisols, calcisols) with overlying paludal and/or lacustrine facies indi- range from 8 to 100 m thick and 0.5 to 40 km in maximum width cates contrasting pedogenic conditions and little or no accommodation (Schumm and Ethridge, 1994; Dalrymple et al., 1994). Within an IVF, space on the interfluves during falling stage, lowstand, and early trans- the lowest story is usually the coarsest and often conglomeratic; gressive systems tracts (Gibling and Bird, 1994; Tandon and Gibling, upper stories are finer grained with a higher frequency of heterolithic 1997; Falcon-Lang, 2004; Hanneman and Wideman, 2010). Rising strata and mudstone plugs. The amalgamation of channel deposits and water table can produce gley soil overprinting of high chroma soils the abrupt increase in grain size within IVFs is expected above sequence and a polygenetic soil with features that reflect contrasting hydromor- boundaries (Emery and Myers, 1996). Coastal plain rivers incised their phic conditions (e.g. Driese and Ober, 2005). These polygenetic valleys during falling base level associated with falling glacioeustatic paleosols are typically present along horizons 4–9inFigs. 8 and 9.A sea level. Crowley and Baum (1991) estimated the magnitude of sea wetter climate accompanied interglacial highstands (e.g. Heckel, 1994; level changes due to Gondwana glaciation to be between 45–75 m Falcon-Lang, 2004; Catuneanu, 2006). The higher sediment flux associ- (minimum ice cover) and 150–190 (maximum ice cover) and those ated with increased runoff would have impeded the development of reaching the Appalachian Basin in its high shelf position required sea limestones unless 1) clastics were filtered by vegetation that fringed level changes of approximately 100 m (Heckel, 2008). the lakes or coastal bays, or 2) clastics were initially trapped in incised IVFs occur in association with all cycles bounded by mature valleys before base level rose above the interfluves. paleosols (Figs. 8, 9) but three are especially prominent. The Upper IFSB paleosols developed in three stages as portrayed in Fig. 12 and Lower Mahoning Sandstone IVFs (Fig. 8, Ruth and Eden Fork sec- (modified from Driese and Ober, 2005, Fig. 9). In Stage I, highly seasonal tions between IFSBs 1 and 4) are 15–19 m thick and separated by the rainfall with short wet/long dry seasons caused oxidizing conditions, Mahoning coal of Windolph (1987). In many places near Charleston, shrink/swell features, and precipitation of micritic limestone nodules, the Mahoning and Upper Freeport coals have been removed by producing a calcic vertisol. Base level was lowest during glacial paleovalley incision (e.g. K42, K47, K37, Figs. 1, 8, 9). A second promi- lowstand. In Stage 2, rising sea level and wetter climate caused water nent IVF, ranging from 13 to 25 m in thickness, occurs above the table to rise and led to impure peat accumulation, reducing conditions Twomile Limestone and locally truncates it (Fig. 9, Priestly, St. Albans, in the upper part of the soil, and gleying. During Stage 3, sea level Eden Fork sections between IFSBs 4 and 7). This IVF correlates with reached highstand, and with wetter climate, water table continued to the Saltsburg-Buffalo Sandstone compound IVF in the Huntington area rise, causing drowning of swamp by lakes and bays and continued gley- (Martino, 2004). A third large IVF, correlated with the Grafton Sand- ing of soil. Calcic vertisols and aridosols formed in the Conemaugh dur- stone IVF of Martino (2004) is represented by the Sandy Grove Sand- ing the Missourian and Early Virgilian lowstands due to higher rates of stone of Windolph (1987; Fig. 9, Flatwoods section above IFSB 8). The evaporation associated with global warming. Coals were thin, impure, top of the Sandy Grove IVF occurs from 70 to 85 m below the Pittsburgh and localized to topographic lows. Lacustrine and palustrine or coastal coal. It is typically 15–20 m thick; greater thicknesses (up to 29 m) likely marine carbonates formed in place of coals or directly above carbona- indicate a compound IVF due to incision into the underlying Saltsburg ceous horizons prior to highstand. IVF. Bundles of thicker and thinner laminae and limited burrowing are 7.4. Paleosol/limestone cycles associated with upper estuarine channels near the fluvial-tidal transi- tion (Greb and Martino, 2005). Inclined heterolithic strata form in A cyclic development of calcic vertisols overlain by microchonchid meandering channels of mixed load streams in fluvial and tidal settings limestones is associated with horizons 5–8 at and above the Twomile (Thomas et al., 1987). Fluvial-tidal transition facies are important in Limestone; these cycles particularly well-developed in the Guthrie- fluvially dominated successions as they may represent the only updip Eden Fork area (IFSBs 5, 6, and 8 in Figs. 8, 9). At K44 (Figs. 8, 9, 13), expression of marine flooding surfaces (Greb and Martino, 2005). the Twomile Limestone is 2 m thick and occurs 120.4 m below the Pitts- Mud-filled channels (Fig. 9, Eden Fork section between IFSBs 5 and burgh coal as mapped by Windolph (1987). Two other microconchid 7) and heterolithic channel fills occur in the upper portion of the IVFs limestones occur at 107 m and 88 m below the Pittsburgh coal (Eden and result from decreased fluvial gradient and stream power associated Fork section, Fig. 9). Each limestone overlies a mature paleosol 1.9– with rising relative base level. 4.7 m thick, and is overlain by shale or mudstone that coarsens upward. These cycles are similar in scale and architecture to the paleosol- 7.3. Interfluvial sequence boundaries bounded cycles of the Huntington area that contain marine limestones and shales directly above a mature IFSB paleosol (Martino, 2004). Interfluvial sequence boundaries (IFSBs) in coastal plain deposits are The paleosols and limestones formed at a time when clastic influx represented by well-developed paleosols formed between incised river was minimal. In both cases, a rising water table drowned the initially valleys during prolonged subaerial exposure (Van Wagoner et al., 1990; well-drained soils. Given the proximity of the study area to the Shanley and McCabe, 1993, 1994; Emery and Myers, 1996; Plint et al., paleoshorelines during maximum transgressions, it is quite likely that 2001; Catuneanu, 2006; Miall, 2010). Sediment bypassing and negative the limestones were deposited in brackish bays or coastal lakes that accommodation occurs during falling stage, lowstand, and early were intermittently connected to the sea during seasonally high lake 152 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 levels or high bay water level associated with spring tides and lunar 8) This study underscores the importance of regionally developed perigee tides. Periodic variations in tropical precipitation patterns paleosols as a valuable tool in stratigraphic correlation and in se- could also have contributed to the cyclic development. Wanless and quence stratigraphic studies. Shepard (1936) were the first to propose more humid conditions ac- companied transgressions (now recognized as sea-level highstands), while drier conditions prevailed during repressions (i.e. lowstands) as- Acknowledgments sociated with glaciation. The drier climatic phases would be the expect- ed effect of the shift of mid latitude high pressure systems toward the Early phases of this study were supported by a grant from the Petro- equator during glacial periods, while wetter phases result from the ex- leum Research Fund (PRF 34516-B8). The author thanks Cortland Eble, pansion of the intertropical convergence zone (characterized by low Nick Fedorko, and Mitch Blake for helpful discussions, Patrick Foster pressure) during sea level highstands (Perlmutter and Matthews, and John Ferguson for field assistance, and Annalisha Johnson and Pat- 1989). A comparable wet/dry pattern linkage to glacioeustatic sea rick Foster for drafting support. Steve Greb and Jim Hower provided level operated during the Late Pleistocene and Holocene as well thorough and thoughtful reviews which significantly enhanced the (Crowley and North, 1991; Kershaw and Nanson, 1993). final version of this paper. Busch and West (1987) described climate change surfaces from the Glenshaw Formation in Ohio and Pennsylvanian. These surfaces sepa- rated well-drained paleosols often associated with arid conditions References from superjacent coals or limestones formed under more humid condi- tions. Nine to eleven paleosol-bounded allocycles are present in the Arkle Jr., T.A., 1974. Stratigraphy of the Pennsylvanian and Systems of the central Appalachians. In: Briggs, G. (Ed.), of the Southeastern United States. lower Conemaugh Group downdip to the north and west of the present Geological Society of America Special Paper 148, pp. 5–29. study area (Busch and Rollins, 1984; Martino, 2004). The dry glacial/wet Blakely, R.C., 2007. Carboniferous–Permian paleogeography of the assembly of Pangaea. interglacial viewpoint is widely held among many other recent In: Wong, T.E. (Ed.), Proceedings on the XVth International Congress on Carbonifer- ous and Permian Stratigraphy, Utrecht, 10–16 August 2003. Royal Dutch Academy researchers (e.g. Perlmutter and Matthews, 1989; Heckel, 1995; of Arts and Sciences, pp. 443–456. Tandon and Gibling, 1997; Driese and Ober, 2005; Catuneanu, 2006; Blum, M.D., Tornqvist, T.E., 2000. Fluvial responses to climate and sea level change: a re- Falcon-Lang et al., 2011). In contrast, a wet glacial/dry interglacial asso- view and a look forward. Sedimentology 47, 2–48. ciation has been advocated by some (e.g. Cecil et al., 2003; Rosenau Boardman II, D.R., Mapes, R.H., Yancey, T.E., Malikny, J.M., 1984. A new model for depth- related allogenic community succession within North American Pennsylvanian et al., 2013a). In this alternative scenario, peat accumulation and higher, cyclothems and implications on the black shale problem. In: Hyne, N.J. (Ed.)- less seasonal rainfall occurred during glacial maxima; vertisols, Limestones of the Mid-Continent: Tulsa Geological Society Special Publication 2, – aridosols, and lacustrine and marine limestones formed during drier pp. 141 182. Bragonier, W.A., 1989. Stratigraphy of flint clays of the Allegheny and Pottsville Groups, conditions associated with sea level highstands. This view is based western Pennsylvania. In: Harper, J.A. (Ed.), Geology in the Laurel Highlands of South- mainly on climate models that suggest that a stable low-pressure cell western Pennsylvania. Guidebook, 54th Annual Field Conference of Pennsylvanian would have been pinned to the equator by the development of high- Geologists. Pennsylvania Geologic Survey, pp. 69–88. Busch, R.M., Rollins, H.B., 1984. Correlation of Carboniferous strata using a hierarchy of pressure zones over the Gondwana ice sheets. However, this wet- transgressive regressive units. Geology 12, 471–474. glacial/dry interglacial model does not fit the succession of events re- Busch, R.M., West, R.R., 1987. Hierarchal genetic stratigraphy: a framework for corded in Conemaugh cyclothems, which would appear to require dry Paleoeceanography. Paleoceanography 2, 141–164. Catuneanu, O., 2006. Principles of Sequence Stratigraphy. Elsevier, Amsterdam, The seasonal conditions to be followed by wetter conditions associated Netherlands (375 pp.). with marine transgression. Cecil, C.B., 1990. Paleoclimate controls on stratigraphic repetition of chemical and siliciclastic rocks. Geology 18, 533–536. Cecil, C.B., Stanton, R.W., Neuzil, S.G., Dulong, F.T., Ruppert, L.F., Pierce, B.F., 1985. Paleoclimate controls on Late Paleozoic sedimentation and peat formation in the cen- 8. Conclusions tral Appalachian Basin, USA. Int. J. Coal Geol. 5, 195–230. Cecil, C.B., Dulong, F.T., Cobb, J.C., Supardi, M., 1993. Allogenic and autogenic controls on sedimentation in the Central Sumatra Basin as an analogue for Pennsylvanian coal- 1) Paleosol-bounded cyclothems of terrestrial strata in central West bearing strata in the Appalachian Basin. In: Cobb, J.C., Cecil, C.B. (Eds.), Modern and Virginia are correlated downdip to the north and west with Ancient Coal-forming Environments. Geological Society of America Special Paper 286, pp. 3–22. cyclothems containing marine units; Cecil, C.B., Dulong, F.T., West, R.R., Stamm, R., Wardlaw, B., Edgar, N.T., 2003. Climate con- trols on the stratigraphy of a Middle Pennsylvanian cyclothem in North America. In: 2) Initial stages of sequence-bounding paleosol development occurred Cecil, C.B., Edgar, N.T. (Eds.), Climate Controls on Stratigraphy. Society of Economic on interfluves during FSST and LST under well-drained oxidizing Paleontologists and Mineralogists Special Publication 77, pp. 151–180. conditions and strongly seasonal climate. Sediment bypassing Collinson, J.D., 1996. Alluvial sediments. In: Reading, H.G. (Ed.), Sedimentary Environ- ments: Processes, Facies and Stratigraphy. Blackwell Science Ltd., Oxford, UK, allowed for prolonged pedogenesis lasting 10s to 100s of thousand pp. 37–82. years; gleying of soils and local histosol development occurred dur- Crowley, T.J., Baum, S.K., 1991. Estimating Carboniferous sea level fluctuations from Gond- ing TST when local water tables rose due to rising sea level and wet- wana ice extent. Geology 19, 975–977. Crowley, T.J., North, G.R., 1991. Paleoclimatology. Oxford Monographs on Geology and ter, less seasonal climate; Geophysics 18. Oxford University Press, New York (339 pp.). 3) Microconchid limestones formed under lacustrine, palustrine, and Dalrymple, R.W., Boyd, R., Zaitlin, B.A., 1994. History of research, types and internal orga- possibly brackish intertidal conditions and represent maximum nization of incised-valley systems: introduction to the volume. In: Dalrymple, R., Boyd, R., Zaitlin, B. (Eds.), Incised-valley Systems: Origin and Sedimentary Sequences: flooding surfaces within predominantly terrestrial sequences; Society of Economic Paleontologists and Mineralogists Special Publication 51, 4) Coarsening-upward bay- and lake-fill sequences formed during the pp. 353–368. HST; Dolby, G., Falcon-Lang, H.J., Gibling, M.R., 2011. A conifer-dominated palynological assem- 5) IVFs occur in association with each paleosol-bounded cycle; the larg- blage from Pennsylvanian (Late Moscovian) alluvial drylands in Atlantic Canada: im- plications for the Vegetation of tropical lowlands during glacial phases. J. Geol. Soc. est ones correlate with the Mahoning, Saltsburg, and Grafton IVFs 168, 571–584. (Martino, 2004); Donaldson, A.C., 1979. Depositional environments of the Upper Pennsylvanian Series. In: 6) Sea-level/base level was the primary control on high-frequency Englund, K.J., Arndt, H.H., Henry, T.W. (Eds.), Proposed Pennsylvanian System fl Stratotype Virginia and West Virginia. American Geologic Institute, Selected Guide- variations in accommodation space within the coastal plain uvial book Series 1, pp. 123–132. systems during the development of the fourth-order cyclothems; Donaldson, A.C., Renton, J.J., Presley, M.W., 1985. Pennsylvanian deposystems and 7) Striking parallels in scale and internal architecture occur between paleoclimates of the Appalachians. Int. J. Coal Geol. 5, 167–193. ‘ ’ Driese, S.G., Ober, E.G., 2005. Paleopedologic and paleohydrologic records of precipitation downdip, marine-cored cyclothems and their updip nonmarine seasonality from Early Pennsylvanian “underclay” paleosols, U. S. A. J. Sediment. Res. cyclothem equivalents; 75, 997–1010. R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154 153

Eble, C.F., 1998. Paleobotanical considerations. Coal Geology, Paleobotany, and Regional Kosanke, R.M., 1988. Palynologic analyses of Upper Pennsylvanian coal beds and adjacent Stratigraphy of the Middle Part of the , Southern West Virginia. strata from the proposed Pennsylvanian System Stratotype in West Virginia. U. S. Geological Society of America Field Trip Guidebook, p. A56-F56. Geol. Surv. Prof. Pap. 1486, 1–24. Eble, C.F., Blake, B.M., Gillespie, W.H., Pfefferkorn, H.W., Chesnut, D.R.J., 2009. Appalachian Krebs, C.E., Teets Jr., D.D., 1914. Kanawha County. West Virginia Geologic Survey County Basin floras. In: Greb, S.F. (Ed.), Carboniferous of the Appalachian and Black Reports (679 pp.). Warrior Basins. Kentucky Geologic Survey Special Publication 10 Series XII, Leeder, M., 2011. Sedimentology and Sedimentary Basins. second ed. Wiley-Blackwell pp. 46–58. (768 pp.). Elliott, T., 1974. Interdistributary bay sequences and their genesis. Sedimentology 21, Mack, G.H., James, W.C., Monger, H.C., 1993. Classification of paleosols. Geol. Soc. Am. Bull. 611–622. 105, 129–136. Emery, D., Myers, K.J., 1996. Sequence Stratigraphy. Blackwell Science Ltd., London (297 Martino, R.L., 2004. Sequence stratigraphy of the Glenshaw Formation (Middle–Late pp.). Pennsylvanian) in the Central Appalachian Basin. In: Pashin, J.C., Gastaldo, R.A. Ettensohn, F., 2008. The Appalachian foreland basin in Eastern United States. In: Miall, (Eds.), Sequence Stratigraphy, Paleoclimate, and Tectonics of Coal-bearing Strata. A.D. (Ed.), The Sedimentary Basins of the United States and Canada. Elsevier, The American Association of Petroleum Geologists Studies in Geology 51, pp. 1–28. Netherlands, pp. 105–179. Martino, R.L., Greb, S., 2009. Walking trails of the giant terrestrial arthropod Arthropleura Falcon-Lang, H.J., 2004. Pennsylvanian tropical rainforests responded to glacial– from the Upper Carboniferous of Kentucky. J. Paleontol. 83, 140–146. interglacial rhythms. Geology 32, 689–692. McCarthy, P.J., 2002. Micromorphology and development of interfluve paleosols: a case Falcon-Lang, H.J., Heckel, P.E., Dimichele, W.A., Blake Jr., B.M., Easterday, C.R., Eble, C.F., study from the Cenomanian Dunvegan Formation, NE British Columbia, Canada. Elrick, S., Gastaldo, R.A., Greb, S.F., Martino, R.L., Nelson, W.J., Pfefferkorn, H.W., Can. Petrol. Geol. Bull. 50, 158–177. Phillips, T.L., Rosscoe, S.J., 2011. No major stratigraphic gap exists near the Middle- McCarthy, P.J., Plint, A.G., 1998. Recognition of interfluve sequence boundaries: integrat- Upper Pennsylvanian (Desmoinesian–Missourian) boundary in North America. ing paleopedology and sequence stratigraphy. Geology 26, 387–390. Palaios 26, 125–139. McCarthy, P.J., Faccini, U.F., Plint, A.G., 1999. Evolution of an ancient coastal plain: Fielding, C.R., Frank, T.D., Isabell, J.L., 2008. The Late Paleozoic ice age—a review of current palaeosols, interfluves and alluvial architecture in a sequence stratigraphic frame- understanding and synthesis of global climate patterns. In: Fielding, C.R., Frank, T.D., work, Cenomanian Dunvegan Formation, NE British Columbia, Canada. Sedimentolo- Isbell, J.L. (Eds.), Resolving the Late Paleozoic Ice Age in Time and Space. Geological gy 46, 861–891. Society of America Special Paper 441, pp. 343–354. Miall, A.D., 2010. The Geology of Stratigraphic Sequences. second ed. Springer-Verlag, Fonner, R.F., 1987. Geology along I-64 Putnam County, West Virginia. West Virginia Geo- Berlin Heidelberg (522 pp.). logical and Economic Survey Publication MAPWV-33 (35 pp.). Moon, C.F., Hurst, C.W., 1984. Fabric of muds and shales: an overview. In: Stowe, D.A.V., Fonner, R.F., Chappell, G.A., 1987. Geology along I-64 Cabell County, West Virginia. Piper, D.J.W. (Eds.), Fine-grained Sediments: Deep-water Processes and West Virginia Geological and Economic Survey Publication MAPWV-32 (59 FaciesGeological Society Special Publication 15. Blackwell ScientificPublications pp.). Limited, Oxford, pp. 579–593. Freylet, P., Plaziat, J.-C., 1982. Continental carbonate sedimentation and pedogenesis—Late Morris, D.A., 1967. Lower Conemaugh (Pennsylvanian) Depositional Environments and Cretaceous and Early Tertiary of Southern France. Contrib. Sedimentol. 12 (213 pp.). Paleogeography in the Appalachian Coal Basin. Unpublished PhD Dissertation, Uni- Gardner, T.W., Williams, E.G., Holbrook, P.W., 1988. Pedogenesis of some Pennsylvanian versity of Kansas, Lawrence (521 pp.) underclays; groundwater, topographic, and tectonic controls. Geol. Soc. Am. Spec. Olariu, C., Bhattacharya, J.P., 2006. Terminal distributary channels and delta front architec- Pap. 216, 81–101. ture of river-dominated delta systems. J. Sediment. Res. 76, 212–233. Gibling, M.R., Bird, D.J., 1994. Late Carboniferous cyclothems and alluvial paleovalleys in Outerbridge, W.F., 1989. Correlation of the Charleston Sandstone of the the Sydney Basin, Nova Scotia. Geol. Soc. Am. Bull. 106, 105–117. proposed Pennsylvanian stratotype with strata in eastern Kentucky, western West Gierlowski-Kordesch, E.H., Cassle, C.F., 2015. The ‘Spirorbis’ problem revisited: sedimen- Virginia, and southern Ohio. U.S. Geological Survey. Miscellaneous Field Studies Map tology and biology of microconchids in marine–nonmarine transitions. Earth Sci. MF — 2110. Rev. 148, 209–227. Peppers, R.A., 1996. Palynological correlation of major Pennsylvanian (Middle and Upper Greb, S.F., Martino, R.L., 2005. Fluvial–estuarine transitions in fluvial-dominant succes- Carboniferous) chronostratigraphic boundaries in the Illinois and other coal basins. sions; examples from the Lower Pennsylvanian of the Central Appalachian Basin. Geol. Soc. Am. Mem. 188 (111 pp.). In: Blum, M., Marriott, S.B., Leclair, S. (Eds.), Fluvial Sedimentology VIIInternational Perlmutter, M.A., Matthews, M.D., 1989. Global cyclostratigraphy—a model. In: Cross, T.A. Association of Sedimentologists, Special Publication 35. Blackwell Publishing, (Ed.), Quantitative Dynamic Stratigraphy. Prentice Hall, Englewood Cliffs, New Jersey, pp. 425–452. pp. 233–260. Greb, S.F., Pashin, J.C., Martino, R.L., Eble, C.F., 2008. Appalachian sedimentary cycles dur- Pfefferkorn, H.W., Gastaldo, R.A., DiMichele, W.A., Phillips, T.L., 2008. Pennsylvanian trop- ing the Pennsylvanian: changing influences of sea level, climate, and tectonics. In: ical floras from the United States as a record of changing climates. In: Fielding, C.R., Fielding, C.F., Frank, T.D., Isbell, J.L. (Eds.), Resolving the Late Paleozoic Gondwanan Frank, T.D., Isbell, J.L. (Eds.), Resolving the Late Paleozoic Ice Age in Time and Space. Ice Age in Time and Space. Geological Society of America Special Publication 441, Geological Society of America Special Paper 441, pp. 305–316. pp. 235–248. Platt, N.H., Wright, V.P., 1991. Lacustrine carbonates: facies models, facies distributions Hanneman, D.L., Wideman, C.J., 2010. Continental sequence stratigraphy and continental and hydrocarbon aspects. In: Anadon, P., Cabreraq, L.I., Kelts, K. (Eds.), Lacustrine carbonates. In: Alonso-Zara, A.M., Tanner, L.H. (Eds.), Carbonates in Continental Facies Analyses. International Association of Sedimentologists Special Publication SettingsDevelopments in Sedimentology 62. Elsevier, pp. 215–273. 13, pp. 55–73. Heckel, P.H., 1994. Evaluation of evidence for glacial-eustatic control over marine Plint, A.G., McCarthy, P.J., Faccini, U.F., 2001. Nonmarine sequence stratigraphy: updip ex- Pennsylvanian cyclothems in North America and consideration of possible tectonic pression of sequence boundaries and systems tracts in a high-resolution framework, effects. In: Dennsion, J.M., Ettensohn, F.R. (Eds.), Tectonic and Eustatic Controls on Cenomanian Dunvegan Formation, Alberta Foreland Basin, Canada. Am. Assoc. Pet. Sedimentary Cycles. Society of Economic Paleontologists and Mineralogists Concepts Geol. Bull. 85, 1967–2001. in Sedimentology and Paleontology 4, pp. 65–87. Potter, P.E., Maynard, J.B., Pryor, W.A., 1980. The Sedimentology of Shale. Springer-Verlag, Heckel, P.H., 1995. Glacial–eustatic base-level-climate model for Late Middle to Late New York (306 pp.). Pennsylvanian coal bed formation in the Appalachian Basin. J. Sediment. Res. B65, Pratt, B.R., 2010. Peritidal carbonates. In: James, N.P., Dalrymple, R.W. (Eds.), Facies 348–356. Models 4. Geological Association of Canada, pp. 401–420. Heckel, P.H., 2008. Pennsylvanian cyclothems in Midcontinent North America as far-field Quinlan, G.M., Beaumont, C., 1984. Appalachian thrusting, lithospheric flexure, and the effects of waxing and waning of Gondwana ice sheets. In: Fielding, C.R., Frank, T.D., Paleozoic stratigraphy of the Eastern Interior of North America. Can. J. Earth Sci. 21, Isbell, J.L. (Eds.), Resolving the Late Paleozoic Ice Age in Time and Space. Geological 973–996. Society of America Special Paper 441, pp. 275–289. Reading, H.G., Collinson, J.D., 1996. Clastic coasts. In: Reading, H.G. (Ed.), Sedimentary En- Hennen, R.V., Gawthrop, R.M., 1917. Braxton and Clay Counties. West Virginia Geological vironments: Processes, Facies, and Stratigraphy. Blackwell Science Ltd, Oxford, and Economic Survey (883 pp.). pp. 154–231. Henry, T.W., Lyons, P.C., Windolph Jr., J.F., 1979. Upper Pennsylvanian and Lower Retallack, G.J., 1988. Field recognition of paleosols. In: Reinhardt, J., Sigleo, W.R. (Eds.), Permian(?) Series in the area of the proposed Pennsylvanian System Stratotype. In: Paleosols and Weathering Through Geologic Time. Geological Society of America Englund, K.J., Arndt, H.H., Henry, T.W. (Eds.), Proposed Pennsylvanian System Special Paper 216, pp. 1–20. Stratotype, Virginia and West Virginia. American Geological Institute Selected Guide- Retallack, G.J., 2001. Soils of the Past. second ed. Blackwell Science, London (404 pp.). book Series, pp. 81–85. Rice, C.L., 1986. Pennsylvanian system (Kentucky). In: McDowell, R.C. (Ed.), The Geology Horne, J.C., Ferm, J.C., Caruccio, F.T., Baganz, B.P., 1978. Depositional models in coal explo- of Kentucky — A Text to Accompany the Geologic Map of Kentucky. U.S. Geological ration and mine planning. Am. Assoc. Pet. Geol. Bull. 62, 2379–2411. Survey Professional Paper 1151-H, pp. H31–H44. Isbell, J.L., Miller, M.F., Wolfe, K.L., Lenaker, P.A., 2003. Timing of the Late Paleozoic glaci- Rice, C.L., Hiett, J.K., Koozmin, E.D., 1994. Glossary of Pennsylvanian stratigraphic names, ation in Gondwana: was glaciation responsible for the development of northern central Appalachian basin. In: Rice, C.L. (Ed.), Elements of Pennsylvanian Stratigraphy, hemisphere cyclothems? In: Chan, M.A., Archer, A.W. (Eds.), Extreme Depositional Central Appalachian Basin. Geological Society of America Special Paper 294, Environments: Mega End Members in Geologic Time. Geological Society of America pp. 115–155. Special Publication 370, pp. 5–24 Rosenau, N.A., Tabor, N.J., Elrick, S.D., Nelson, W.J., 2013a. Polygenetic history of paleosols Keller, W.D., 1981. The sedimentology of flint clay. J. Sediment. Petrol. 51, 233–244. in Middle-Upper Pennsylvanian cyclothems of the Illinois Basin, U.S.A.: Part I. Charac- Keroher, G.C., Bedette, B., Eggleton, J.L., Hansen, B.E., Kopf, R.W., Mann, C., Melton Jr., W.G., terization of paleosol types and interpretation of pedogenic processes. J. Sediment. Nace, H.L., Palmer, K.D., Sando, W.J., Smedley, J.E., Stock, S.W., Swift, C.S., Toulmin, Res. 83, 606–636. M.S., Wilson, D., 1966. Lexicon of geologic names for the United States for Rosenau, N.A., Tabor, N.J., Elrick, S.D., Nelson, W.J., 2013b. Polygenetic history of 1936–1960. U.S. Geol. Surv. Bull. 1200 (4341 pp.). paleosols in Middle-Upper Pennsylvanian cyclothems of the Illinois Basin, Kershaw, A.P., Nanson, G.C., 1993. The last full glacial cycle in the Australian region. Glob. U.S.A.: Part II. Integrating geomorphology, climate, and glacioeustasy. Planet. Chang. 7, 1–9. J. Sediment. Res. 83, 637–668. 154 R.L. Martino / International Journal of Coal Geology 154–155 (2016) 136–154

Schultz, H.-P., 2009. Interpretation of marine and freshwater paleoenvironments in C.E. (Ed.), Seismic stratigraphy—Applications to Hydrocarbon Exploration. American Permo-Carboniferous deposits. Palaeogeogr. Palaeoclimatol. Palaeoecol. 281, Association of Petroleum Geologists Memoir 26, pp. 49–212. 126–136. Van Wagoner, J.C., Mitchum, R.M., Campion, K.M., Rahmanian, V.D., 1990. Siliciclastic se- Schumm, S.A., Ethridge, F.G., 1994. Origin, evolution and morphology of fluvial valleys. In: quence stratigraphy in well logs, cores, and outcrops: concepts for high-resolution Dalrymple, R., Boyd, R., Zaitlin, B. (Eds.), Incised-valley Systems: Origin and Sedimen- correlation of time and facies. American Association of Petroleum Geologists Methods tary Sequences. Society of Economic Paleontologists and Mineralogists Special Publi- in Exploration Series 7 (55 pp.). cation 51, pp. 11–27. Wanless, H.R., 1939. Pennsylvanian correlations in the eastern interior and Appalachian Shanley, K.W., McCabe, P.J., 1993. Alluvial architecture in a sequence stratigraphic frame- coal fields. Geol. Soc. Am. Spec. Pap. 17 (130 pp.). work: a case history from the Upper Cretaceous from southern Utah. In: Flint, S.S., Wanless, H.R., Shepard, F.P., 1936. Sea level and climatic changes related to late Paleozoic Bryant, I.D. (Eds.), The Geologic Modeling of Hydrocarbon Reservoirs and Outcrop cycles. Geol. Soc. Am. Bull. 47, 1177–1206. Analogues. International Association of Sedimentologists Special Publication 15, Wanless, H.R., Weller, J.M., 1932. Correlation and extent of Pennsylvanian cyclothems. pp. 21–56. Geol. Soc. Am. Bull. 43, 1003 –1016. Shanley, K.W., McCabe, P.J., 1994. Perspectives on the sequence stratigraphy of continen- Weller, M.J., 1930. Cyclic sedimentation of the Pennsylvanian Period and its significance. tal strata. Am. Assoc. Pet. Geol. Bull. 78, 544–568. J. Geol. 38, 97–135. Staub, J.R., Cohen, A.D., 1979. The Snuggedy Swamp of South Carolina: a back-barrier West Virginia Geological and Economic Survey Coal Bed Mapping Project; Pittsburgh estuarine coal-forming environment. J. Sediment. Res. 49, 133–143. Coal. http://www.wvgs.wvnet.edu/GIS/CBMP/Pittsburgh.html). Stout, W., 1947. Generalized geologic section of rocks in Ohio. Ohio Geologic Survey White, W.A., 1961. Colloid phenomena in sedimentation of argillaceous rocks. J. Sediment. Information Circular 4. Petrol. 31, 560–570. Sturgeon, M.T., Hoare, R.D., 1968. Pennsylvanian brachiopods of Ohio. Ohio Geol. Surv. Williams, E.G., Berbenback, R.F., Falla, W.S., Udagawa, S., 1968. Origin of some Bull. 63 (95 pp.). Pennsylvanian underclays in western Pennsylvania. J. Sediment. Petrol. 39, 1179–1193. Tandon, S.K., Gibling, M.R., 1997. Calcretes at sequence boundaries in Upper Carbonifer- Windolph Jr., J.F., 1987. Geologic Map of the Big Chimney Quadrangle, Kanawha County, ous cyclothems of the Sydney Basin, Atlantic Canada. Sediment. Geol. 112, 43–67. West Virginia. U. S. Geological Survey 1:24,000, Map GQ-1612. Tasch, P., 1980. Paleobiology of the Invertebrates, Data Retrieval From the Fossil Record. Woodland, B.G., Stenstrom, R.C., 1979. The occurrence and origin of siderite concretions in John Wiley and Sons, New York (975 pp.). the Francis Creek Shale (Pennsylvanian) of northeastern Illinois. In: Nitechi, M.H. Taylor, P.D., Vinn, O., 2006. Convergent morphology in small spiral worm tubes (Ed.), Mazon Creek Fossils. Academic Press, New York, pp. 69–103. (‘Spirorbis’) and its palaeoenvironmental implications. J. Geol. Soc. 163, 225–228. Wright, V.P., 1992. Paleopedology: stratigraphic relationships and empirical models. In: Thomas, R.G., Smith, D.G., Wood, J.M., Visser, J., Calverley-Range, E.A., Koster, E.H., 1987. Martini, I.P., Chesworth, W. (Eds.), Weathering, Soils, and Paleosols. Earth Surface Inclined heterolithic stratification—terminology, description, interpretation, and Processes 2. Elsevier, pp. 475–499. significance. Sediment. Geol. 53, 123–179. Wright, V.P., Marriott, S.B., 1993. The sequence stratigraphy of fluvial depositional Tucker, M.E., 2001. Sedimentary iron deposits. Sedimentary Petrology. Blackwell Science, systems: the role of floodplain sediment storage. Sediment. Geol. 86, 203–210. pp. 182–193. Zatoń, M., Peck, R.L., 2013. Morphology and palaeoecology of new, non-marine Vail, P.R., Mitchum, R.M., Todd, R.G., Widmier, J.M., Thompson, S., Sangree, J.B., Bubb, J.N., microconchid tubeworm from Lower Carboniferous (Upper ) of West Hatfield, W.G., 1977. Seismic stratigraphy and global changes in sea level. In: Payton, Virginia, USA. Ann. Soc. Geol. Pol. 83, 37–50.