You are on page 1of 22

Journal of Sedimentary Environments (2023) 8:261–282

https://doi.org/10.1007/s43217-023-00132-y

RESEARCH

Climate‑induced changes in fluvial ichnofossil assemblages


of the Pennsylvanian–Permian Appalachian Basin
Jennifer K. Crowell1 · Daniel I. Hembree2

Received: 31 January 2023 / Revised: 28 March 2023 / Accepted: 1 April 2023 / Published online: 21 April 2023
© The Author(s), under exclusive licence to Springer Nature Switzerland AG 2023

Abstract
The long-term response of riparian communities to shifting climatic conditions can be addressed by the ichnofossil record,
because organism behavior is typically altered in response to changes in environmental factors. During the late Paleozoic,
the Appalachian Basin experienced a shift from an ever-wet to wet–dry climate. Changes in the abundance, diversity, density,
and composition of ichnofossil assemblages were investigated in fluvial point bar sandstones from five roadside outcrops of
the Middle Pennsylvanian-to-early Permian Allegheny, Conemaugh, Monongahela, and Dunkard groups located in southeast
Ohio and northwest West Virginia. Ichnofossil data were collected using a 0.5 × 0.5-m grid placed on bedding plane surfaces
and from vertically oriented thin sections. Abundance, density, diversity, and burrow widths increased through the study
interval. Behaviors changed from stationary- to mobile-deposit feeding, while community composition shifted toward more
established, permanent generalists. These changes in ichnofossil assemblages suggest that the shift to a drier, more pro-
nounced seasonal climate made short- to long-term occupation of the point bar sands more advantageous as surface conditions
were more unfavorable and resources limited. This study helps us understand how terrestrial community composition and
ecosystem dynamics shift over long time intervals in response to environmental perturbations. By assessing these changes,
we can better predict what future impacts climatic shifts will have on continental ecosystems and terrestrial communities.

Keywords Continental · Point bar · Trace fossil · Behavior · Paleoecology

1 Introduction fundamental shifts in ecosystems due to significant environ-


mental changes including climate change (Bromley, 1996;
Ichnofossils are vital to understanding ancient ecosystems Buatois et al., 1998).
and environments as they are often the only evidence of life The terrestrial climate of the late Paleozoic has been
in some settings (Bromley, 1996). Ichnofossils have been characterized by the development of strong seasonality in
utilized to interpret various aspects of continental environ- the continental interior of Pangea (Cecil, 2013; Gastaldo
ments, including temperature, oxygenation, salinity, nutrient et al., 1996; Powell et al., 2009; Tabor & Poulsen, 2008;
content, and turbidity, because organism behavior is typi- Tabor et al., 2008). Several studies have been conducted on
cally altered in response to changes in environmental fac- the effect of the late Paleozoic climatic shift on paleosols of
tors (Gingras et al., 2007; Hasiotis, 2002; Hasiotis et al., the Appalachian Basin (Hembree, 2022; Hembree & Blair,
2007; Hembree, 2018; MacEachern et al., 2007; Minter 2016; Hembree & Bowen, 2017; Hembree & Carnes, 2018;
et al., 2016). Additionally, ichnofossils can help interpret Hembree & McFadden, 2020; Hembree & Nadon, 2011);
however, few have focused on changes in riparian ichno-
fossil assemblages. In this study, ichnofossil assemblages
Communicated by M. V. Alves Martins
preserved in point bar sandstones were studied through the
* Daniel I. Hembree strata deposited during the Pennsylvanian–Permian climatic
dhembre2@utk.edu transition in the Appalachian Basin to understand the long-
term impacts of climatic shifts on riparian ecosystems. Sec-
1
Department of Geology and Geological Engineering, tions of the Middle Pennsylvanian (Moscovian)-to-early
University of Mississippi, Oxford, MS, USA
Permian (Asselian) Allegheny, Conemaugh, Monongahela,
2
Department of Earth and Planetary Sciences, University and Dunkard groups in southeast Ohio and northwest West
of Tennessee Knoxville, Knoxville, TN, USA

13
Vol.:(0123456789)
262 J. K. Crowell, D. I. Hembree

Virginia were selected, because paleosols within these 2011). These paleosols record a regional shift to more vari-
groups indicate an up-section shift from a wet to a drier, able climatic conditions with fluctuations in paleoprecipita-
more seasonal, climate (Hembree, 2022). tion indicating a change from a humid ever-wet climate to
Investigating changes in terrestrial ichnofossil assem- a drier, more seasonal climate (Hembree & Carnes, 2018;
blages provides a broader understanding of the impact that Hembree & McFadden, 2020). A decrease in precipitation
climatic shifts have on environmentally sensitive compo- and drier seasons resulted in a shift away from plant com-
nents of ancient ecosystems. By assessing these changes, munities dominated by floodplain wetland flora assemblages
we can also better predict what future impacts climatic shifts with poorly rooted groundcover (Hembree, 2022). Abundant
will have on continental ecosystems and, specifically, ripar- and pervasive rhizoliths of the Middle-to-early Late Penn-
ian invertebrate communities (LaRoux and McGeoch, 2008; sylvanian indicate widespread surface vegetation of a high
Walther, 2010; Capon et al., 2013). moisture environment, whereas uncommon and clumped rhi-
zoliths of the Late Pennsylvanian to Early Permian indicate
sparse surface vegetation of a seasonally dry environment
2 Geologic setting (Hembree, 2022).

2.1 Middle Pennsylvanian‑to‑early Permian climate 2.2 Allegheny, Conemaugh, Monongahela,


transition and Dunkard groups

Widespread Late Devonian through Early Mississippian This study involved the investigation of fluvial facies asso-
glacial deposits are some of the earliest indicators of the ciations in the Middle Pennsylvanian (Moscovian)-to-early
initiation of the Late Paleozoic Ice Age (LPIA) (Montañez Permian (Asselian) Allegheny, Conemaugh, Monongahela,
& Poulsen, 2013). By the Late Mississippian-to-Middle and Dunkard groups in southeast Ohio and northwest West
Pennsylvanian, continental ice centers extended into areas Virginia (Fig. 1A). The Middle Pennsylvanian Allegheny
of southern Gondwana and the Earth was in an icehouse Group crops out in Ohio and Pennsylvania, is up to 300 m
state (Crowell, 1995; DiMichele, 2014; DiMichele et al., thick, and consists of shales, limestones, and sandstones as
2009; Montañez & Poulsen, 2013). The final assembly and well as six major coal zones (Ferm, 1970; Sturgeon & Mer-
northward drift of the supercontinent Pangea from the Late rill, 1949; Williams, 1960). During the deposition of the
Pennsylvanian to the early Permian had a long-lasting effect Allegheny, a narrow marine embayment existed in the region
on many Earth systems, including global atmospheric circu- and experienced periodic flooding during times of sea-level
lation patterns and glacial states (Blakey, 2008; Montañez rise (Fig. 1B) (Martin, 1998). The Late Pennsylvanian Cone-
& Poulsen, 2013; Tabor & Poulsen, 2008). Ice sheets waxed maugh Group of Ohio and Pennsylvania is between 325 and
and waned during this time, creating changes in the distri- 375 m thick and consists of fluvial and marine lithofacies,
bution of land and ocean as sea-level oscillated (Tabor & including shales, sandstones, limestones, and some coals
Poulsen, 2008). Low-latitude environmental changes, such divided into two formations, the Glenshaw and Casselman
as withdrawal of epeiric seas, increasing continentality, and (Condit, 1912; Hembree & Nadon, 2011; Sturgeon, 1958).
decreasing precipitation, likely resulted in the drying of The upper Allegheny and lower Conemaugh groups were
near-equatorial regions of Pangea (Tabor & Poulsen, 2008). deposited by a massive northwestwardly prograding flu-
In the Appalachian Basin, this resulted in a major shift from vial–deltaic system (DiMichele et al., 1996).
a humid, nonseasonal climate to a drier, subhumid, strongly Outcrops of the Monongahela and Dunkard groups can be
seasonal climate (Montañez & Cecil, 2013). Seasonality was found in Ohio, Pennsylvania, and West Virginia (Hembree &
especially stronger in the interior of Pangea as it was far McFadden, 2020; Martin, 1998). Both groups were deposited
from the moderating effects of the ocean. close to the equator within the central to northern portions of
Many studies have utilized paleosols and, to a lesser the Appalachian Basin and consist of fluvial and lacustrine
extent, ichnofossils, to document this climate transition in sediments (Fig. 1C) (Hembree & McFadden, 2020). The Mon-
the Appalachian Basin (Hembree, 2022; Hembree & Blair, ongahela Group is Late Pennsylvanian in age, 80–125 m thick,
2016; Hembree & Bowen, 2017; Hembree & Carnes, 2018; and is regionally divided into the Pittsburgh and Uniontown
Hembree & McFadden, 2020). Appalachian Basin paleosols formations (Hembree & McFadden, 2020; Martin, 1998; Stur-
of the Upper Pennsylvanian exhibit redox features, little-to- geon, 1958). These formations are composed of shales, mud-
no carbonate nodules, and minor vertic features in contrast to stones, sandstones, and limestones that were deposited in a
paleosols from the Lower Permian which exhibit oxidizing river-dominated deltaic system during a regression (Hembree
features, abundant carbonate nodules, and prominent ver- & McFadden, 2020; Martin, 1998). The Dunkard Group is
tic features (Hembree & Blair, 2016; Hembree & Carnes, Early Permian, up to 360 m thick, and regionally divided into
2018; Hembree & McFadden, 2020; Hembree & Nadon, the Waynesburg, Washington, and Greene formations which

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 263

Fig. 1  Stratigraphy and paleogeography of the study area. A Gen- vanian (B) and early Permian (C). The location of the study areas is
eral stratigraphic column showing the geologic age of the Allegheny, indicated by the white star. (Images by R. Blakey. Used with permis-
Conemaugh, Monongahela, and Dunkard groups. B, C Paleogeo- sion)
graphic reconstructions of North America during the Middle Pennsyl-

are mainly composed of shale, mudstone, and sandstone and Virginia, USA (Fig. 2). The studied sections included por-
minor coal and limestone (Hembree & Carnes, 2018; Hembree tions of the Middle Pennsylvanian Allegheny Group, Late
& McFadden, 2020; Lucas, 2013; Martin, 1998; Schneider Pennsylvanian Conemaugh and Monongahela groups, and
et al., 2013). Dunkard Group sediments were derived from the early Permian Dunkard Group. At each locality one to
the weathering of the Allegheny Mountains and deposited on two, 3–8 m-thick, fluvial facies associations comprised of
a broad fluvial plain (Martin, 1998). sandstones, shales, mudstones, siltstones, and coal were
studied. General stratigraphic sections of the facies associa-
tions were constructed to document major lithologies, sedi-
3 Materials and methods mentary structures, and ichnofossil occurrences.
A 0.5 × 0.5-m grid was used to assess the ichnofossil
The study area consisted of five roadside outcrops in south- assemblages along exposed bedding planes of point bar
east Ohio and northwest West Virginia. Field sites were sandstones at the top of each fluvial facies association. For
located in Athens County, Ohio and Wood County, West each facies association, eight grids were placed on bedding

13
264 J. K. Crowell, D. I. Hembree

Fig. 2  A Location of the study area within Ohio and West Virginia United States (at star). B Map showing the study sites (Sites 1–5) and
(within white square) relative to the Appalachian and Dunkard basins. the location of the stratigraphic sections
The inset map shows the location of the study region within the

plane surfaces with at least 0.5 m of exposure, spaced up differences between the various quantitative properties of
to 45 m apart. Ichnofossil data collected within the grids the ichnofossil assemblages existed. A Dunn’s post hoc
included abundance, density, diversity, and composition. comparison was used to indicate where these differences
Abundance was determined by counting the number of indi- occurred stratigraphically.
vidual ichnofossils identified within the grid. Density was
assessed by recording the concentration of ichnofossils in
the grid and the degree of sediment reworking using ich- 4 Results
nofabric indices (ranked 0–6) (Buatois & Mángano, 2011).
Diversity was assessed by counting the number of unique 4.1 Site descriptions
ichnogenera. Composition was assessed by (1) evaluating
the mean ichnofossil complexity by counting segments, 4.1.1 Allegheny Group—Site 1
branches, chambers, and openings to the surface of each
ichnofossil (Hembree, 2018); (2) measuring the widths of The section at Site 1 consisted of (base to top) (1) carbo-
each ichnofossil to determine a mean value; and (3) evalu- naceous, silty shale (240 cm thick), (2) green-gray massive
ating the number of different behaviors represented within claystone (30 cm thick) with root traces, (3) bituminous coal
the grid based on ichnofossil morphology (Bromley, 1996; (60 cm thick) with coarse plant fragments, (4) green-gray
Buatois & Mángano, 2011). micaceous shale (15 cm), and (5) medium- to coarse-grained
Sandstone samples were collected for thin-section prepa- (coarsening upward), massive, micaceous sandstone (wacke)
ration from each studied grid at the five field sites. Thin (80 cm thick) (Figs. 3A, 4A).
sections were vertically oriented to assess the vertical pen-
etration of the substrate by ichnofossils observed on bedding 4.1.2 Conemaugh Group—Site 2
planes, total levels of bioturbation, and to identify additional
ichnofossils in cross section that were not visible in the field. The section at Site 2 consisted of (base to top) (1) gray,
A total of 43 thin sections (35 1.5″ × 3.0″, 8 1.0″ × 2.0″ fine-grained, micaceous sandstone (wacke) (115 cm thick),
slides) were prepared by Texas Petrographic (Houston, TX, (2) gray, medium-grained, micaceous sandstone (wacke)
USA). Thin sections were examined using a Motic BA300 (90 cm thick), and (3) fine- to medium-grained (coarsen-
polarizing petrographic microscope. Mineralogy, grain size, ing upward), finely laminated to cross-laminated sandstone
grain sorting and rounding, sedimentary structures, and ich- (arenite) (150 cm thick) with iron concretions occurring near
nofossil content (i.e., burrows and rhizoliths) were described the base (Figs. 3B, 4B).
for each thin section.
The quantitative ichnofossil data from each grid and bed- 4.1.3 Monongahela Group—Sites 3A, 3B
ding plane were combined from each outcrop to compare
the ichnofossil assemblages between each unit. An analysis The section at Site 3A consisted of (base to top) (1) green-
of variance (ANOVA) was used to compare the different gray, blocky claystone (30 cm thick), (2) red blocky mud-
populations of abundance, density, diversity, and composi- stone (40 cm thick) with cutans and mottling, (3) green-gray
tion data between each of the five localities to test whether mudstone (50 cm thick) with cutans and mottling, (4) dark

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 265

Fig. 3  Stratigraphic sections from Sites 1–5. A Site 1 Allegheny Dunkard Group (east). F Site 4B Lower Dunkard Group (west). G
Group. B Site 2 Conemaugh Group. C Site 3A Monongahela Group Site 5 Upper Dunkard Group
(upper). D Site 3B Monongahela Group (lower). E Site 4A Lower

gray shale (30 cm thick) with carbonized plant fragments, The section at Site 3B consisted of (base to top) (1)
(5) green-gray blocky mudstone (90 cm thick) with cutans, red, silty mudstone (60 cm thick) with rhizoliths and mot-
slickensides, mottling, and iron nodules, (6) green-gray, tling, (2) green-gray, calcareous, silty mudstone (70 cm
platy mudstone (80 cm thick), (7) green-gray, laminated to thick), (3) red, blocky, silty mudstone (130 cm thick) with
cross-laminated shale (90 cm thick), and (8) four fining- rhizoliths and cutans, (4) variegated shale (180 cm thick),
upward successions (40, 70, 70, 80 cm thick) of fine-grained, and (5) fine- to medium-grained (coarsening upward),
cross-laminated to laminated, sandstone (wacke) (Figs. 3C). cross-laminated, micaceous sandstone (108 cm thick)

13
266 J. K. Crowell, D. I. Hembree

Fig. 4  Outcrop images and grid examples. A Allegheny Group (Site (Site 2). H Point bar sandstone surface with grid from Monongahela
1). B Conemaugh Group (Site 2). C Monongahela Group (Site 3b). Group (Site 3b). I Point bar sandstone surface with grid from Lower
D Lower Dunkard Group (Site 4a). E Upper Dunkard Group (Site 5). Dunkard Group (Site 4a). J Point bar sandstone surface with grid
F Point bar sandstone surface with grid from Allegheny Group (Site from Upper Dunkard Group (Site 5)
1). G Point bar sandstone surface with grid from Conemaugh Group

with interbedded sandstone, siltstone, and mudstone lenses micaceous sandy shale (60 cm thick), (3) green-gray platy
(Figs. 3D,4C). shale (65 cm thick), (4) fine-grained massive sandstone
(20 cm thick), (5) argillaceous shale (90 cm thick), (6) green,
4.1.4 Lower Dunkard Group—Sites 4A, 4B blocky argillaceous mudstone (20 cm thick), (7) red argil-
laceous mudstone (140 cm thick) with rhizoliths and cutans,
The section at Site 4A consisted of (base to top) (1) cross- (8) green-gray shale (90 cm thick), and (9) fine-grained
bedded, fine-grained sandstone (40 cm thick), (2) gray, sandstone (150 cm thick) (Figs. 3E,4D).

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 267

The section at Site 4B consisted of (base to top) (1) tabu- 4.2.4 Lower Dunkard Group—Site 4
lar, cross-bedded, fine-grained, micaceous sandstone (70 cm
thick), (2) green-gray sandy shale (60 cm thick), (3) green- Sandstones (n = 11) were medium–fine-grained to very-fine-
gray shale (60 cm thick), (4) light green-gray shale (10 cm grained, moderately sorted to well sorted, and composed
thick), (5) red blocky mudstone (130 cm thick), (6) gray of angular quartz, micas, iron-bearing minerals, rock frag-
shale (175 cm thick), and (7) fine-grained sandstone (160 cm ments, rare plagioclase feldspar, and a green matrix (quartz-
thick) (Figs. 3F). wacke) (Fig. 5). Lamination and bioturbation occurred in
several samples (Fig. 6E-H). Bioturbation included general
mixing of sediment grains as well as distinct burrows.
4.1.5 Upper Dunkard Group—Site 5
4.2.5 Upper Dunkard Group—Site 5
The section at Site 5 consisted of (base to top) (1) medium-
grained sandstone (70 cm thick) with trough crossbedding Sandstones (n = 3) were extremely fine-grained to
at the base and ripple cross lamination at the top, (2) thin medium–fine-grained, well sorted to moderately sorted,
sandstone (20 cm thick) with interbedded siltstone and mud- and composed of abundant angular quartz, common mica,
stone and iron nodules, medium-grained partially laminated rock fragments, and iron-bearing minerals, and a fine matrix
sandstone (35 cm thick), (3) reddish-brown, blocky, fining (quartz-wacke) (Fig. 5). Lamination and bioturbation were
upward, sandy shale (110 cm thick) with iron nodules, uncommon.
(4) green-gray argillaceous mudstone (30 cm thick), (5)
red argillaceous mudstone (180 cm thick) with rhizoliths, 4.3 Ichnofossils of the Allegheny, Conemaugh,
(6) gray argillaceous shale (40 cm thick), and (7) fine- to Monongahela, and Dunkard groups
medium-grained, bioturbated sandstone (180 cm thick)
(Figs. 3G,4E). Ichnofossils were described from sandstones at the top of
each stratigraphic sequence. Ichnofossils with distinct mor-
phologies were identified to ichnogenus level, as preserva-
4.2 Sandstone petrography tion prohibited confident identification to the ichnospecies
level. General bioturbation of some sites resulted in a rec-
4.2.1 Allegheny Group—Site 1 ognizable ichnofabric, however, it was not possible to assign
these to specific ichnotaxa. Occurrence is described as rare
Sandstones (n = 8) were medium–fine-grained, moderately (1–5), common (5–10), and abundant (> 10).
sorted, and composed of abundant angular quartz, common
micas, uncommon plagioclase feldspar and rock fragments, 4.3.1 
Skolithos
and a fine matrix (quartz-wacke) (Fig. 5A).
These were single, circular, horizontal cross sections of
vertical shafts found on sandstone bedding plane surfaces
4.2.2 Conemaugh Group—Site 2 (Fig. 7A, B). Shaft diameter ranged from 0.3 to 1.5 cm.
The shafts were unlined and had a fill similar to the matrix.
Sandstones (n = 7) were fine-grained, well-sorted, and com- These burrows were attributed to Skolithos, which is diag-
posed of abundant angular quartz, uncommon mica, rock nosed as a subcylindrical, unbranched, unlined-to-lined,
fragments, and plagioclase feldspar, a fine matrix (quartz- straight-to-curved structureless tube, perpendicular to the
wacke), and iron cement (Fig. 5B). bedding plane surface (Häntzschel, 1975). The burrows were
similar to Cylindrichum, but rounded terminations were not
observed, and Monocraterion, but lacked a funnel-shaped
4.2.3 Monongahela Group—Site 3 opening (Häntzschel, 1975). Skolithos were rare in the Cone-
maugh, common in the Monongahela, and abundant in the
Sandstones (n = 14) were very-fine-grained, well-sorted, and Upper and Lower Dunkard groups (Fig. 7C–D).
composed of abundant angular quartz, common mica and
rock fragments, uncommon iron-bearing minerals, rare pla- 4.3.2 
Arenicolites
gioclase feldspar, and a fine matrix (quartz-wacke) (Fig. 5C).
Cross lamination was observed in all samples. Bioturbation These were paired, circular, horizontal cross sec-
was noted in several thin sections, including mixing of sedi- tions of vertical, unlined shafts with a fill similar to the
ment grains and disturbance of primary sedimentary struc- matrix occurring on sandstone bedding plane surfaces
tures (Fig. 6A-D). (Fig. 7E–F). Individual shaft cross sections were similar

13
268 J. K. Crowell, D. I. Hembree

Fig. 5  Thin sections of point


bar sandstones. A Allegheny
Group (Site 1) sandstone con-
taining angular medium-grained
quartz, plagioclase feldspar, and
rock fragments. B Conemaugh
Group (Site 2) sandstone with
angular, fine-grained quartz,
mica, rock fragments, plagio-
clase feldspar, and iron cement.
C Monongahela Group (Site 3)
sandstone containing very-
fine-grained quartz, mica, rock
fragments, and a fine matrix. D
Lower Dunkard Group (Site 4)
sandstone containing very-fine-
grained quartz, mica, rock frag-
ments, and a fine green matrix.
E Upper Dunkard Group (Site
5) sandstone containing fine- to
medium-grained angular quartz,
mica, and rock fragments with a
fine matrix

in size and appearance to those assigned to Skolithos, 4.3.3 


Cruziana
however, the regular proximity (< 1 cm) of the two shafts
suggests they were likely connected in the subsurface. In These were straight, horizontal, bilobed trails, 1.0–1.5 cm
addition, samples from the Monongahela Group show the wide and 10–20 cm in length, with central furrows (Fig. 7G).
presence of the shafts on both the top and bottom surfaces Surficial features (i.e., bioglyphs) could not be assessed due
of the sample. Shafts were unlined and contained a fill to poor preservation. These trails were attributed to Cruzi-
like that of the surrounding matrix. These burrows were ana which is diagnosed as a trail with elongate, bilobed fur-
attributed to Arenicolites which is diagnosed as a simple, rows covered with herringbone-shaped, longitudinal stria-
vertical, cylindrical U-shaped tube with smooth sides, tions (Häntzschel, 1975). While the striations were missing
found perpendicular to the bedding plane (Häntzschel, from the observed specimens, the bilobed furrows are char-
1975). Tube openings can be flared or funnel-shaped acteristic of Cruziana. Sizes and appearance of these trails
(Häntzschel, 1975). These burrows are also similar to Dip- vary greatly due to the behavior of the trace makers and
locraterion but do not possess spreite between the two substrate consistency (Häntzschel, 1975). Cruziana were
shafts (Häntzschel, 1975). Arenicolites were rare in the typically found in assemblages with Skolithos and were rare
Monongahela and Dunkard groups. in the Dunkard Group.

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 269

Fig. 6  Thin sections from point


bar sandstones showing primary
sedimentary structures and
bioturbation. A Lamination and
general bioturbation (sediment
mixing) from the Monongahela
Group (Site 3B). B Cross lami-
nation and general bioturbation
(sediment mixing) from the
Monongahela Group (Site 3B).
C Root traces and sediment
mixing from the Monongahela
Group (Site 3B). D Disrupted
and deflected lamination from
the Monongahela Group (Site
3B). E General bioturbation
(sediment mixing) of the Lower
Dunkard Group (Site 4A).
F Burrows (Skolithos) and
disrupted clay laminae from the
Lower Dunkard Group (Site
4B). G Passively filled burrows
along the upper sandstone sur-
face from the Lower Dunkard
Group (Stie 4B). H General
bioturbation (sediment mixing
and disrupted laminae) from the
Lower Dunkard Group (Site 4B)

4.3.4 
Planolites surface with a fill darker than the matrix. The tunnel fill
consisted of irregular meniscate backfilling. Scratch marks
These were simple, unlined, and actively filled horizontal were observed along the sides of the burrows but were not
burrows (Fig. 8A, B). The burrows were generally straight, well preserved. These burrows were attributed to Scoyenia
less than 1 cm in width, and did not branch. The burrow which is diagnosed as a slender, straight to curved tunnel
fill was distinct from the matrix and was raised above the with rope-like sculpture produced by paired striations that
bedding plane. The tunnel surfaces were generally smooth cover the tunnel surface and active backfill (Häntzschel,
with no visible scratch marks or internal structure. These 1975). The tunnels commonly cross each other but do not
burrows were attribute to Planolites which is diagnosed as branch (Häntzschel, 1975). These burrows are also similar to
an unlined, subcylindrical to cylindrical, straight to curved, Taenidium and Beaconites but lack the well-defined menis-
unbranched but often overlapping, actively filled tunnel cate backfill of both and the lining of Beaconites, in addition
(Häntzschel, 1975). They are also similar to Palaeophy- to exhibiting striations on the tunnel surfaces which is not
cus, but lack a lining and do not branch (Häntzschel, 1975). present in either ichnogenus (Häntzschel, 1975). Scoyenia
Planolites were common in the Lower Dunkard Group and were found in assemblages with Skolithos and were common
rare in the Upper Dunkard Group and found in assemblages in the Lower Dunkard and Upper Dunkard groups.
with Skolithos and Scoyenia.
4.3.6 
Treptichnus
4.3.5 
Scoyenia
These were a series of short, horizontal, oval-shaped, con-
These were horizontal, straight to curved tunnels with a nected tunnel segments (Fig. 9). The segments were typi-
wrinkly surface texture (Fig. 8). The tunnels were non- cally connected in a linear pattern, although some were
branching but overlapping, 1 cm wide, and no longer than slightly curved. Segments within a series were usually
5 cm. The tunnels were raised above the bedding plane equal in size, however, some segments were elongated. The

13
270 J. K. Crowell, D. I. Hembree

Fig. 7  Ichnofossils. A, B Bed-


ding plane views of Skolithos
from the Monongahela Group
(Site 3). C, D Bedding plane
views of Arenicolites from the
Monongahela Group (Site 3). E,
F Bedding plane views of dense
concentrations of Skolithos and
Arenicolites from the Lower
Dunkard Group (Site 4). G Bed-
ding plane view of Cruziana
from the Upper Dunkard Group
(Site 5)

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 271

Fig. 8  Ichnofossils. A Bed-


ding plane view of Planolites
from the Upper Dunkard Group
(Site 5). B Bedding plane view
of Planolites from the Lower
Dunkard Group (Site 4). C
Bedding plane view of Scoyenia
from the Upper Dunkard Group
(Site 5). D Bedding plane view
of Scoyenia from the Lower
Dunkard Group (Site 4)

average segment length was 0.3–1.0 cm. Most commonly bioturbation) (Table 1). Ichnofossil assemblages of the Lower
each series consisted of three segments. Some bioglyphs, Dunkard Group were dominated by Skolithos but also included
including scratch marks, were present along the base of Arenicolites, Cruziana, Scoyenia, and Planolites. The mean
some segments. Burrow fill was similar to the matrix. These abundance was 188, diversity was 2.8, density was 3.6, com-
burrows were attributed to Treptichnus which is diagnosed plexity 1.1, mean width was 5.3 mm, and ichnofabric index
as a series of shallow, subhorizontal, J- or U-shaped tun- was 4 (moderate to high bioturbation) (Table 1). Most quanti-
nel segments joined together in irregular patterns near the tative assemblage properties increased from the Monongahela
tunnel ends (Häntzschel, 1975). They have low slopes and Group, except for mean width. Ichnofossil assemblages of the
a passive fill (Häntzschel, 1975). While individual tunnel Upper Dunkard Group included abundant Treptichnus as well
segments were similar to the basal expression of Arenicolites as Skolithos, Arenicolites, Cruziana, Scoyenia, and Planolites.
or Diplocraterion, their occurrence in repeated series make The mean abundance was 73, diversity was 2.4, density was
these diagnoses unlikely (Häntzschel, 1975). Treptichnus 1.9, complexity was 2, mean width was 5.7 mm, and ichnofab-
were abundant in the Upper Dunkard Group. ric index was 2 (moderate bioturbation) (Table 1). Abundance,
density, and ichnofabric index of the Upper Dunkard decreased
4.3.7 Analysis of ichnofossil assemblages relative to the Lower Dunkard, while the diversity, complexity,
and width increased.
No ichnofossils were observed in point bar sandstones of the Quantitative data was combined from the bedding planes
Middle Pennsylvanian Allegheny Group. Ichnofossil assem- of each outcrop to compare assemblages between the Alle-
blages in point bar sandstones of the Conemaugh Group gheny, Conemaugh, Monongahela, and Dunkard groups. An
were composed of Skolithos and had a mean abundance of analysis of variance (ANOVA) was performed to compare the
3.6, diversity of 1, density of 1, complexity of 1, mean width different populations of abundance, density, diversity, com-
of 5.2 mm, and an ichnofabric index of 1 (low bioturbation) plexity, and mean width data between each of the five groups.
(Table 1). The mean values from the Conemaugh were the P values for abundance, diversity, density, and complexity
lowest in the study interval. Ichnofossil assemblages of the were all less than 0.05 indicating that significant differences
Monongahela Group also consisted mostly of Skolithos, but existed between these aspects of the ichnofossil assemblages
also included rare Arenicolites. The mean abundance was (Table 2). Dunn’s post hoc comparison showed that the assem-
11.6, diversity was 1.2, density was 1, complexity was 1, blages of the Allegheny and Conemaugh groups were distinct
mean width was 6.3 mm, and ichnofabric index was 1.5 (low from those of the Monongahela and Dunkard groups (Table 2).

13
272 J. K. Crowell, D. I. Hembree

(Retallack et al., 2001). The upper sandstone coarsens


upwards indicating a change to a higher energy condition
(Miall, 2010).

5.1.2 Conemaugh Group—Site 2

Units 1–3 of Site 2 are interpreted as a series of stacked


point bar and channel deposits (Collinson, 1996; Miall,
2010). Units 1 and 2 coarsen upward and appear massive,
suggesting an upward increase in energy from the levee to
point bar (Collinson, 1996; Miall, 2010). Unit 3 contains a
succession of fine-grained sandstones with multiple, coarser-
grained, erosively based channel forms suggesting rapid lat-
eral movement of the channel (Miall, 2010).

5.1.3 Monongahela Group—Sites 3A, B

The section at Site 3A is interpreted as a transition from a


distal to proximal floodplain (Units 1–3, 5), floodplain pond
(Unit 4), levee (Units 6–7), to point bar (Units 8–11) (Collin-
son, 1996; Brierly et al., 1997; Aslan & Autin, 1998; Tabor
& Montañez, 2004; Stow, 2005; Miall, 2010). The prop-
erties of the paleosols indicate shifts from poorly drained
(Units 1, 3), well-drained (Unit 2), to moderately drained
(Unit 5) conditions, likely related to changes in proximity
to the active channel (Collinson, 1996; Tabor & Montañez,
2004). The properties of the succession of upper sandstones
(Units 8–11) indicate changes in energy from high to low as
the point bars migrated (Miall, 2010).
The section at Site 3B is interpreted as a transition from
moderately drained, distal to proximal floodplain (Units
1–3), levee (Unit 4), to point bar (Unit 5) (Collinson, 1996;
Brierly et al., 1997; Retallack et al., 2001; Tabor & Mon-
tañez, 2004; Stow, 2005; Miall, 2010). The interbedded
Fig. 9  Specimens of Treptichnus from the underside of sandstone sands, silts, and clays of Unit 5 indicate fluctuating energy
blocks of the Upper Dunkard Group (Site 5). A, B Cluster of long-to- conditions from high to low during periods of high and low
short, linked segments of Treptichus. C Close up of a single specimen discharge, respectively (Collinson, 1996; Miall, 2010).
of Treptichnus consisting of at least three, linked tunnel segments

5.1.4 Lower Dunkard Group—Sites 4A, B


5 Discussion
The section at Site 4A is interpreted as a transition from
5.1 Unit interpretations point bar (Unit 1), levee (Units 2–3), point bar (Unit 4),
levee (Unit 5), proximal to distal floodplain (Units 6–7),
5.1.1 Allegheny Group—Site 1 levee (Unit 8), to point bar (Unit 9) (Collinson, 1996; Bri-
erly et al., 1997; Aslan & Autin, 1998; Tabor & Montañez,
The section at Site 1 is interpreted as a transition from flood- 2004; Stow, 2005; Miall, 2010).
plain pond (Unit 1), proximal floodplain and mire (Units 2, The section at Site 4B is interpreted as a transition from
3), levee (Unit 4), to point bar (Unit 5) (Collinson, 1996; point bar (Unit 1), levee (Units 2–4), distal floodplain (Unit
Brierly et al., 1997; Aslan & Autin, 1998; Stow, 2005; 5), levee (Unit 6), to point bar (Unit 7) (Collinson, 1996; Bri-
Miall, 2010). The features of the paleosol (Unit 2) indicate erly et al., 1997; Retallack et al., 2001; Tabor & Montañez,
poor development and drainage as well as water saturation 2004; Stow, 2005; Miall, 2010).

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 273

Table 1  Quantitative properties of ichnofossil assemblages from Sites 2–5


Location Grid Abundance Diversity Density Complexity MW: Sk MW: Ar MW: Pl MW: Sc MW: Cr MW: Tr Behavior

Site 2 1
2 2 1 1 1 10 D, F
3 3 1 1 1 4 D, F
4 2 1 1 1 3 D, F
5
6 3 1 1 1 5 D, F
7
8 3 1 1 1 4 D, F
Site 3A 1 8 1 1 1 10 D, F
2 7 2 1 1 10 10 D, F
3 15 1 1 1 5 D, F
4 3 1 1 1 5 D, F
5 11 1 1 1 10 D, F
6 3 1 1 1 10 D, F
7 7 1 1 1 5 D, F
8 23 2 1 1 5 D, F
Site 3B 9 18 2 1 1 5 5 D, F
10 28 1 1 1 5 D, F
11 6 1 1 1 3 D, F
12 6 1 1 1 4 D, F
13 9 1 1 1 5 D, F
14 12 1 1 1 4 D, F
15 14 1 1 1 7 D, F
16 15 1 1 1 5 D, F
Site 4A 1 60 4 2 1 10 10 5 3 F
2 110 4 3 1 3 3 20 5 F
3 135 2 3 1 5 5 F
4 200 2 3 1 3 3 F
5 200 3 4 1 3 3 4 F
6 400 3 5 1 6 6 3 F
7 120 3 3 1 5 5 5 F
8 140 3 3 1 4 4 5 F
Site 4B 9 120 2 3 1 5 5 F
10 200 3 3 1 5 5 3 F
11 100 4 3 2 5 5 4 10 F, L
12 300 2 5 1 3 3 F
13 200 2 5 1 5 5 F
14 75 2 2 1 5 5 F
15 300 2 6 1 5 5 F
16 350 3 5 1 3 3 6 F
Site 5 1 150 1 2 3 3 F
2 130 1 2 2 3 F
3 110 2 2 4 4 3 D, F
4 75 2 2 3 2 4 F
5 30 3 1 1 15 5 10 F, L
6 20 3 1 1 10 5 3 F, L
7 50 4 4 1 3 3 10 5 F, L
8 15 3 1 1 10 5 5 F

MW mean width (measured in mm), Sk Skolithos, Ar Arenicolites, Pl Planolites, Cr Cruziana, Sc Scoyenia, Tr Treptichnus. In the Behavior col-
umn, D dwelling, F feeding, L locomotion

13
274 J. K. Crowell, D. I. Hembree

Table 2  ANOVA and Dunn’s post hoc results comparing abundance, diversity, density, complexity, and mean width values from the Cone-
maugh, Monongahela, and Dunkard group ichnofossil assemblages
ANOVA
Property p value

Abundance 9.917 x 10E-8


Diversity 0.0134
Density 1.246 x 10E-9
Complexity 0.0024
Width 0.2354
Dunn's Post Hoc

Abundance Complexity
Site 2 Site 3A Site 3B Site 4A Site 4B Site 5 Site 2 Site 3A Site 3B Site 4A Site 4B Site 5
Site 2 0.25 0.12 1.99E-05 7.34E-06 1.65E-03 Site 2 0.77 0.77 0.77 0.83 0.11
Site 3A 0.25 0.63 3.65E-04 1.38E-04 0.02 Site 3A 0.77 1 1 0.93 0.03
Site 3B 0.12 0.63 2.02E-03 8.51E-04 0.07 Site 3B 0.77 1 1 0.93 0.03
Site 4A 1.99E-05 3.65E-04 2.02E-03 0.8 0.2 Site 4A 0.77 1 1 0.93 0.03
Site 4B 7.34E-06 1.38E-04 8.51E-04 0.8 0.13 Site 4B 0.83 0.93 0.93 0.93 0.04
Site 5 1.65E-03 0.02 0.07 0.2 0.13 Site 5 0.11 0.03 0.03 0.03 0.04
Diversity Width
Site 2 Site 3A Site 3B Site 4A Site 4B Site 5 Site 2 Site 3A Site 3B Site 4A Site 4B Site 5
Site 2 0.76 0.84 0.04 0.33 0.03 Site 2 0.54 0.42 0.24 0.31 0.29
Site 3A 0.76 0.56 0.04 0.45 0.03 Site 3A 0.54 0.09 0.02 0.03 0.03
Site 3B 0.84 0.56 0.01 0.18 0.01 Site 3B 0.42 0.09 0.76 0.89 0.84
Site 4A 0.04 0.04 0.01 0.21 0.92 Site 4A 0.24 0.02 0.76 0.82 0.9
Site 4B 0.33 0.45 0.18 0.21 0.17 Site 4B 0.31 0.03 0.89 0.82 0.93
Site 5 0.03 0.03 0.01 0.92 0.17 Site 5 0.29 0.03 0.84 0.9 0.93
Density
Site 2 Site 3A Site 3B Site 4A Site 4B Site 5
Site 2 0.9 0.9 1.98E-03 5.12E-04 0.18
Site 3A 0.9 1 2.40E-04 4.02E-05 0.1
Site 3B 0.9 1 2.40E-04 4.02E-05 0.1
Site 4A 1.98E-03 2.40E-04 2.40E-04 0.66 0.04
Site 4B 5.12E-04 4.02E-05 4.02E-05 0.66 0.01
Site 5 0.18 0.1 0.1 0.04 0.01

Bold indicates significance

5.1.5 Upper Dunkard Group—Site 5 2004, 2011). In continental environments, Skolithos are


usually associated with active channels and sandbar depos-
The section at Site 5 is interpreted as a transition from point its (Buatois & Mángano, 2004, 2011; Hembree & Blair,
bar (Units 1–3), levee (Unit 4), proximal to distal floodplain 2016). Fluvial channels are high-energy environments
(Units 5–6), levee (Unit 7), to point bar (Unit 8) (Collinson, with rapid fluctuations in sedimentation and erosion rates,
1996; Brierly et al., 1997; Aslan & Autin, 1998; Tabor & resulting in very stressful and unstable conditions (Buat-
Montañez, 2004; Stow, 2005; Miall, 2010). ois & Mángano, 2004, 2011). As a result, Skolithos typi-
cally occur in low diversity assemblages and tracemakers
5.2 Ichnofossil interpretation were likely opportunistic generalists that rapidly colonized
exposed surfaces (Buatois & Mángano, 2004; Miller &
5.2.1 
Skolithos Collinson, 1994). Potential trace makers include various
arthropods, such as spiders, beetles, and numerous insect
Given the depositional setting, Skolithos most likely rep- larvae, however, the unstable nature of the environment
resents dwelling or deposit feeding (Buatois & Mángano,

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 275

would be unsuitable for the permanent dwellings of spi- 5.2.6 


Treptichnus
ders (Hembree, 2017; Hils & Hembree, 2015).
Treptichnus is associated with a variety of feeding activities
including detritus and deposit feeding and predatory activity
5.2.2 
Arenicolites (Getty et al., 2016). In the Upper Dunkard Group, deposit
feeding is the most probable given their depth of occur-
In this study area Arenicolites likely represent dwelling or rence and abundance. Treptichnus is typically associated
deposit feeding activities since point bar deposits are suba- with relatively diverse ichnofossil assemblages (Buatois &
erial environments (Hasiotis, 2002; Rindsberg & Kopaska- Mángano, 2011). Treptichnus is widely distributed and has
Merkel, 2005). Potential producers of Arenicolites include a broad stratigraphic range, with an equally broad range of
myriapods (millipedes and centipedes) and insect larvae potential tracemakers (Getty et al., 2016). Continental Trep-
such as those of chironamids, mayflies, and beetles (Hasio- tichnus are usually considered to be made by insect larvae
tis, 2002; Hembree, 2009, 2019; Mikuś & Uchman, 2013; (Getty et al., 2016).
Rindsberg & Kopaska-Merkel, 2005; Thacker & Hembree,
2021). 5.3 Changes in ichnofossil assemblages
from the Allegheny to Dunkard groups

5.2.3 Cruziana Ichnofossils can be used to reconstruct paleoenvironmen-


tal conditions at the time of trace production because they
Behaviors likely associated with Cruziana include loco- represent direct interactions of an organism with their envi-
motion and deposit feeding (Buatois & Mángano, 2004; ronment. Environmental conditions that can result in differ-
Häntzschel, 1975; Zonneveld et al., 2002). Cruziana have ent organism responses in continental settings may include
been previously described from fluvial settings, specifically temperature, precipitation, nutrient levels, water-table depth,
in desiccated overbank deposits (Buatois & Mángano, 2011). substrate moisture, and environmental stability (Hembree,
Scratch marks associated with Cruziana are attributed to 2018). Changes in these conditions may be reflected in ich-
active digging in a firm substrate (Buatois & Mángano, nofossils abundance, density, diversity, and composition.
2004, 2011; Zonneveld et al., 2002). Potential tracemak-
ers are isopods or beetles given the age and riparian setting 5.3.1 Allegheny Group
(Lavelle & Spain, 2005; Lawrence & Newton, 1982).
There are several explanations for the absence of ichnofos-
sils in the Allegheny Group. Destruction of traces can occur
5.2.4 
Planolites during high-energy events that mobilize the upper layer of
sediment (Gingras et al., 2007), high sedimentation rate
Activities associated with Planolites include mobile-deposit may limit time for the colonization of the surface (Gingras
feeding and are suggestive of sediment with high organic et al., 2007), or harsh environmental conditions within the
content (Häntzschel, 1975; Hasiotis, 2002). Many differ- substrate may inhibit its colonization (Beatty et al., 2008;
ent invertebrate tracemakers may produce Planolites, but in Morrissey & Braddy, 2004). None of these scenarios are
continental settings are most typically attributed to annelids supported by the sedimentological data from Site 1. Instead,
and insects (Buatois & Mángano, 2004; Hasiotis, 2002; Kale the absence of ichnofossils in the point bar sandstones of
et al., 1997). the Allegheny Group may have been related to the favorable
conditions above the substrate. The Allegheny Group was
deposited under an ever-wet climate (Garcés et al., 1997;
5.2.5 Scoyenia Hembree & Nadon, 2011; Milici, 2005) defined by high
annual rainfall, high humidity, and available surface water
Deposit feeding and dwelling are behaviors associated with throughout the year (Hasiotis et al., 2007; Underwood et al.,
Scoyenia (Häntzschel, 1975). Producers of Scoyenia are 2014). High soil moisture and water-table levels of ever-
indicative of moist to wet, compact, silty clay substrates wet climates result in low ichnodiversity because epigeal
and are typically associated with point bar, levee, and flood- and arboreal organism lifestyles are more dominant due to
plain deposits (Buatois & Mángano, 2011; Frey et al., 1984; water-saturated substrate conditions (Hasiotis et al., 2007).
Hasiotis, 2004; Hasiotis & Dubiel, 1993). Scratch marks on The abundance of land plants during this time would have
Scoyenia are indicative of an organism with appendages that contributed to a significant amount of detritus in continental
gripped the burrow walls including larvae of beetles or other settings (DiMichele, 2014). Detritus is a common source
insects (Hasiotis, 2004). of energy and nutrients for soil faunal communities; it can

13
276 J. K. Crowell, D. I. Hembree

also serve as a habitat (i.e., shelter, refugia), making it an ingested sediment usually consists of 95% inorganic material
integral part of ecosystems (Condron et al., 2010; Moore (Lopez & Levinton, 1987). Increased dependence on deposit
et al., 2004). The favorable surface conditions of an ever- feeding, which is also energetically expensive, suggests that
wet environment would have limited the need to colonize surface conditions were becoming less favorable, with an
point bar sands in search of resources or shelter. Readily overall trend of decreasing precipitation and increasing sea-
available detritus on the surface would have been selected sonality (Hembree & Bowen, 2017; Hembree & Carnes,
over resources found within the substrate that would have 2018; Hembree & McFadden, 2020).
required extra energy to acquire by burrowing (Jumars &
Wheatcroft, 1989). 5.3.4 Lower Dunkard Group

5.3.2 Conemaugh Group The ichnofossil assemblages of the Lower Dunkard Group


show a significant increase in their quantitative properties,
The appearance of Skolithos in the point bar deposits of the especially in abundance and diversity, from the Mononga-
Conemaugh Group suggests colonization by opportunistic hela Group. Multiple studies of the Lower Dunkard Group
arthropods engaged in deposit feeding and dwelling activi- have indicated that it was a dry-seasonal climate (Cecil,
ties (Buatois & Mángano, 2004, 2011; Miller & Collin- 2013; Fedorko & Skema, 2013; Hembree & Carnes, 2018;
son, 1994). Modern burrows similar to Skolithos are often Hembree & McFadden, 2020; Montañez & Cecil, 2013).
associated with opportunistic organisms that display a high Drier wet–dry climates like that of the Lower Dunkard have
tolerance to physiologically stressful conditions (Buatois & a strongly seasonal distribution and lower amounts of pre-
Mángano, 2011). Burrowing is an energy intensive form of cipitation (Hasiotis et al., 2007). These aspects of wet–dry
locomotion and is typically only utilized when there is a climates result in an increase in ichnodiversity in conti-
strong selective pressure to do so (Jumars & Wheatcroft, nental settings due to lower soil moisture levels (Hasiotis
1989). Likewise, soil fauna, such as arthropods, resort to et al., 2007). The major increase in abundance, diversity,
deposit feeding only if there is a significant selective pres- and density of ichnofossils in the Lower Dunkard shows a
sure to feed on nutrient-poor sediments (Jumars & Wheat- substantial increase in mobile-deposit feeding activities of
croft, 1989). The Conemaugh Skolithos may have also been animals to obtain buried resources from within the sediment.
produced as temporary refuges in response to seasonally Opportunistic animals are noted for their ability to vary their
stressed or unstable surface conditions, such as rapidly shift- feeding styles and sources based on food availability in their
ing water-table levels (Buatois & Mángano, 2004; Hembree environment (Cadée, 1984; Vossler & Pemberton, 1988). If
& Nadon, 2011). Overall, the transition from an absence of surface conditions were not optimal and contained minimal
ichnofossils in the Allegheny Group to their occurrence in resources, then these animals would utilize the best feeding
the Conemaugh Group suggests the onset of less optimal strategy to obtain sustenance, in this instance, deposit feed-
surface conditions as the region began its shift to a drier, ing. A diversity of arthropods like beetles, spiders, insect
more seasonal climate (Cecil et al., 1985; Joeckel, 1995; larvae, myriapods, and isopods were the probable trace
Hembree & Nadon, 2011; Catena & Hembree, 2012; Dze- makers of the ichnofossil assemblage of the Lower Dunkard
nowski and Hembree, 2012), resulting in the need for vari- (Buatois & Mángano, 2004; Hasiotis, 2002, 2004; Hembree
ous animals to occasionally engage in deposit feeding or cre- & Nadon, 2011).
ate temporary dwelling spaces to avoid stressful conditions.
5.3.5 Upper Dunkard Group
5.3.3 Monongahela Group
The Upper Dunkard Group represents a continuation of
Climate conditions continued to shift from the tropical- stressed environments in a dry-seasonal climate (Cecil,
seasonal, wet–dry climate of the Conemaugh to the drier, 2013; Fedorko & Skema, 2013; Hembree & Carnes, 2018;
subhumid climate of the Monongahela (Brezinski & Kollar, Hembree & McFadden, 2020; Montañez & Cecil, 2013).
2011; Cecil et al., 2004). Across this interval, most quantita- The ichnofossils of the Upper Dunkard point bar sandstones
tive aspects of the ichnofossil assemblages in the point bar are predominantly associated with mobile-deposit feeding
sandstone increased. The increased abundance of vertical behaviors. The mean complexity of the Upper Dunkard
burrows such as Skolithos and Arenicolites in low diver- assemblages was the highest of the studied units, due to the
sity assemblages suggests that this environment contained appearance of Treptichnus. Although not necessarily a new
soil arthropods engaging in deposit feeding and dwelling behavior, Treptichnus is significant, because it represents
behaviors in a stressed environment (Buatois & Mángano, an animal systemically moving through the substrate in a
2004; Vossler & Pemberton, 1988). Deposit feeders survive complex way not seen in the other studied units (e.g., Getty
on food sources with relatively poor nutritional value as et al., 2016; Miller, 2003).

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 277

5.4 Riparian community composition change wide environmental tolerances, short lifespans, high density
of individuals, low species diversity, and generalized feeding
Modern riparian habitats are the most dynamic, diverse, habits (Pianka, 1970; Ekdale, 1985; Buatois and Mangano,
complex, and ecologically productive terrestrial environ- 2011). Opportunistic organisms are known for their ability
ments on Earth (Kondolf et al., 1996; Naiman et al., 1993). to quickly colonize a habitat after a major environmental
For these reasons, it is particularly important to understand change and thrive in highly stressful and unstable environ-
how communities adapt to changing climatic and environ- ments with low resources (Ekdale, 1985; Vossler & Pem-
mental conditions in these settings. Ancient riparian com- berton, 1988; Zhao et al., 2020). Many insects fit into this
munities are beneficial to study, because they record changes life strategy and are the most likely producers of the ichno-
that occur through prolonged periods of time providing con- fossils found at the study sites (Pianka, 1970). Ichnofauna
text to evaluate the long-term responses of modern riparian produced by opportunistic organisms usually display low
communities to climatic shifts now and in the future. ichnodiversity, localized distributions, high abundance and
This study compared differences in the ichnofossil assem- density, and simple morphologies (Ekdale, 1985; Buatois
blages from the tropical (ever-wet) riparian environments and Mangano, 2011; Zhao et al., 2020). Based on the sub-
of the Allegheny Group to the semi-arid (wet–dry) ripar- tropical environments of the Conemaugh and Monongahela
ian environments of the Upper and Lower Dunkard groups. groups, it is likely that organisms in these groups only fed
Modern ever-wet tropical climates have abundant precipita- occasionally in point bar sands if the opportunity arose and
tion with a short dry season, constant high temperatures, likely tended to stay in the deposits of floodplains containing
and consistently wet soil conditions (Hasiotis et al., 2007; plentiful organic matter.
Walsh et al., 2009). This type of climate is most common The ichnofossil assemblages of the Monongahela and
in dense rainforests and swamps that receive over 240 rainy Dunkard groups were also produced by generalists; how-
days a year such as parts of the Amazon Basin, Indonesia, ever, the ichnofossils were indicative of a more permanent,
and Malaysia (Boyce & Lee, 2010; Hasiotis et al., 2007). established community. Such communities are mostly com-
Arthropod assemblages of ever-wet climates in modern set- posed of equilibrium, or K-selected, organisms (Ekdale,
tings typically require moist habitats with abundant leaf litter 1985). Equilibrium organisms are much slower to colonize
and include groups, such as millipedes, centipedes, amphi- a new environment compared to opportunistic organisms;
pods, and isopods (Nakamura et al., 2003). Point bar, levee, however, once established, they tend to be better adapted to
and floodplain deposits commonly contain terrestrial beetle the new environment (Ekdale, 1985; Pianka, 1970). Equilib-
traces (Hasiotis et al., 1998). rium organisms are typically specialized feeders that occupy
Modern semi-arid (wet–dry) climates experience high specific niches in a habitat and exhibit narrow environmental
seasonality and a range of precipitation where the amount of tolerances (Ekdale, 1985). They are usually part of high-
precipitation can be less, equal, or greater than evaporation diversity, persistent climax communities with maximized
rates (Hasiotis et al., 2007). Savannahs, grasslands, steppes, resource utilization occurring where the environment is sta-
and mixed savannah woodlands are typical of wet–dry cli- ble and steady (Ekdale, 1985; Whittaker, 1953). The ichno-
mates (Hasiotis et al., 2007). In modern wet–dry climates, fossil assemblages of the Monongahela and Dunkard groups
riparian zones are noted for their ecological importance as exhibited traits of a community that was more complex than
habitats, because they contain high biodiversity in compara- an r-selected community, but not a complex as a K-selected
tively small areas (Chan et al., 2008; Dimmitt, 2000; Kon- community. For instance, the Dunkard Group had ichnofos-
dolf et al., 1996). In wet–dry regions that are exposed to sil assemblages with high diversity, abundance, and density,
seasonal environmental extremes such as high temperatures but most ichnogenera were produced by simple dwelling,
and low humidity such as the subtropical Sonoran Desert, it locomotion, and feeding activities. Despite the simplicity
is common for surface dwelling animals to construct subsur- of the behaviors exhibited, the shift to an established com-
face shelters to avoid extreme surface conditions (Cloudsley- munity provides evidence that the long-term occupation of
Thompson, 1975; Dimmitt, 2000; Hembree et al., 2017). point bar sands had become more advantageous for perma-
Ancient semi-arid riparian environments also likely con- nent occupation by generalists as surface conditions became
tained a high biodiversity reflected in the ichnofossil record. drier and more seasonal.
Population strategies of trace-making organisms can be The shift from no ichnofossils in the Allegheny Group
interpreted using ichnofossil abundance, density, and diver- to hundreds of ichnofossils within the Dunkard Group sup-
sity (Ekdale, 1985). The ichnofossil assemblages of the ports the interpretation that as the climate changed, animals
Conemaugh and Monongahela groups were produced by expanded into less-suitable substrates to find available,
generalists within opportunistic, temporary communities. concentrated resources. Selective pressures on changing
Opportunistic, or r-selected, organisms are those that exhibit food quality and quantity greatly affect organism forag-
high reproductive rates, rapid growth rates, small body sizes, ing behavior (Jumars & Wheatcroft, 1989). The energy

13
278 J. K. Crowell, D. I. Hembree

spent searching, assessing, and exploiting a resource must The absence of ichnofossils in the Allegheny Group
be gained back by the obtained resource (Koy & Plotnick, sandstone was characteristic of an ever-wet climate in
2007). The transition to a drier, more seasonal climate which the environment was well vegetated with abundant
from the Late Pennsylvanian-to-early Permian would have resources resulting in no need for organisms to expend
resulted in changes in resource availability and, conse- energy on burrowing through point bar sands to deposit
quently, a shift in organism behavior to acquire food. Most feed. During the deposition of the Conemaugh Group,
generalist organisms can alter their feeding habits depending there was a shift toward a more seasonal, subtropical cli-
on resource availability in their environment, such as shift- mate. Surface conditions started to become less favorable
ing from detritus to deposit feeding if it was the most viable resulting in the appearance of ichnofossils produced by
option to obtain nutrients (Cadée, 1984; Vossler & Pember- dwelling and deposit feeding activities in the point bar
ton, 1988). Deposit feeding is an energy intensive form of sandstones. Surface conditions progressively declined as
foraging and is engaged in when there is a selective pressure the climate shifted to a drier, subhumid one during the
to consume nutrient-poor particles (Jumars & Wheatcroft, deposition of the Monongahela Group. The ichnofossils
1989). A shift to obligate rather than opportunistic deposit of the Conemaugh and Monongahela groups consisted
feeding in the Dunkard Group was likely driven by increas- of Skolithos and Arenicolites suggesting a community
ingly stressful surface conditions requiring the expenditure composed of opportunistic generalists that temporarily
of more energy to gain nutrients within the less ideal point occupied point bar sands. By the Lower Permian Dunkard
bar sands. Group, sedimentary units were representative of a dry-
Organisms respond to climatic changes in various ways seasonal climate. Ichnofossil abundance, diversity, and
depending mainly on the nature, rate, and length of the density increased significantly and more ichnofossil types
change as well as the range of biological responses available were observed including Cruziana, Planolites, Scoyenia,
to organisms (Erwin, 2009). Some modern biotic responses and Treptichnus. The community composition transitioned
to either short- or long-term climatic events suggest flex- to an established, permanent community of generalists that
ible community composition structure that allows many inhabited the substrate for long durations. Another trend
organisms to respond to rapid climatic shifts by migration in community composition observed was a shift from sta-
(Roy et al., 1996). In the case of the Dunkard Group, arthro- tionary detritus feeding and dwelling to mobile detritus
pods may have responded to shifting climatic conditions by and deposit feeding. This was indicated by an increase in
migrating into sandy, nutrient-poor substrates and adjust- the complexity of ichnofossil morphology as organisms
ing feeding behaviors to manage the demands of increas- engaged in slightly more sophisticated feeding styles to
ingly more arid and highly seasonal environments. maximize resources gained from the sediment.
Examining how continental trace-making commu-
nity composition and ecosystem dynamics shift over time
6 Conclusions is important, because it provides crucial details about
how communities responded to environmental perturbations
Continental ichnofossils provide important insight into in the past. With modern changes in climate and environ-
riparian ecosystems not revealed from body fossils alone ment becoming increasingly of great concern, particularly
and help provide a broader understanding of how long-term aridification, it is important to understand how ancient
climatic changes impact terrestrial communities and eco- organisms and environments responded to climatic shifts to
system dynamics. This study involved the investigation of more arid conditions (Capon et al., 2013; LeRoux & McGe-
changes in the abundance, density, diversity, and complexity och, 2008; Walther, 2010). The study revealed that drier,
of riparian ichnofossil assemblages in point bar sandstones highly seasonal climate conditions resulted in organisms
from the Middle Pennsylvanian-to-early Permian Allegheny, needing to colonize harsher, nutrient-poor substrates and
Conemaugh, Monongahela, and Dunkard groups of Ohio engage in energy-expensive feeding styles to have enough
and West Virginia. Sedimentological evidence from the resources to sustain themselves. By assessing how ancient
Allegheny to Dunkard groups record a climatic shift from communities adapted to changing environmental conditions,
the mid-Pennsylvanian to the early Permian from a tropi- more informed predictions can be made about what future
cal, ever-wet climate to a wet–dry, semi-arid, and highly impacts climatic shifts will have on continental ecosystems
seasonal climate, respectively. The riparian trace-making and terrestrial communities.
communities of the Appalachian Basin were comprised
Acknowledgements We would like to thank Joseph Wislocki for help
mostly of arthropods, such as isopods, beetles, and myri- with field work. This project would not have been possible without
apods whose interactions with the subaerial substrate pro- funding by the Ohio University Geological Sciences Graduate Student
vide salient information about how terrestrial invertebrate Alumni Grant (to JKC), the Geological Society of America Student
communities responded to these climatic changes. Research Grant (to JKC), the Society for Sedimentary Geology Student

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 279

Research Grant (to JKC), and the Paleontological Society Student Capon, S. J., Chambers, L. E., Mac Nally, R., Naiman, R. J., Davies,
Research Grant (to JKC). P., Marshall, N., Pittock, J., Reid, M., Capon, T., Douglas,
M., Catford, J., Baldwin, D. S., Stewardson, M., Roberts, J.,
Author contributions J.C. and D.H. wrote the manuscript text and Parsons, M., & Williams, S. E. (2013). Riparian ecosystems
prepared the figures and tables. All authors reviewed the manuscript. in the 21st century: Hotspots for climate change adaptation?
Ecosystems, 16(3), 359–381.
Funding Funding for this research was provided by the Ohio University Catena, A., & Hembree, D. (2012). Recognizing vertical and lat-
Geological Sciences Graduate Student Alumni Grant (to JKC), the eral variability in terrestrial landscapes: A case study from
Geological Society of America Student Research Grant (to JKC), the the paleosols of the Late Pennsylvanian Casselman Formation
Society for Sedimentary Geology Student Research Grant (to JKC), and (Conemaugh Group) southeast Ohio, USA. Geosciences, 2(4),
the Paleontological Society Student Research Grant (to JKC). 178–202.
Cecil, C. B. (2013). An overview and interpretation of autocyclic
Availability of data and materials All data collected are available in and allocyclic processes and the accumulation of strata during
the tables and appendices. the Pennsylvanian–Permian transition in the central Appala-
chian Basin, USA. International Journal of Coal Geology, 119,
Code availability Statistical analyses were performed using PAST ver- 21–31.
sion 3, free software available at https://​folk.​uio.​no/​ohamm​er/​past/. Cecil, C. B., Stanton, R. W., Neuzil, S. G., Dulong, F. T., Ruppert, L.
F., & Pierce, B. S. (1985). Paleoclimate controls on late Paleo-
Declarations zoic sedimentation and peat formation in the central Appalachian
Basin (USA). International Journal of Coal Geology, 5(1–2),
Conflict of interest The authors have no competing interests to declare 195–230.
that are relevant to the content of this article. Cecil, C. B., Brezinski, D. K., & Dulong, F. (2004). The Paleozoic
record of changes in global climate and sea level: Central Appa-
lachian Basin (Vol. 1264, pp. 77–133). Geology of the national
capital region—Field trip guidebook: US Geological Survey
Circular.
References Chan, E. K., Yu, Y. T., Zhang, Y., & Dudgeon, D. (2008). Distribution
patterns of birds and insect prey in a tropical riparian forest.
Aslan, A., & Autin, W. J. (1998). Holocene flood-plain soil forma- Biotropica, 40(5), 623–629.
tion in the southern lower Mississippi Valley: Implications for Cloudsley-Thompson, J. L. (1975). Adaptations of Arthropoda to arid
interpreting alluvial paleosols. Geological Society of America environments. Annual Review of Entomology, 20(1), 261–283.
Bulletin, 110(4), 433–449. Collinson, J. D. (1996). Alluvial sediments. In H. G. Reading (Ed.),
Beatty, T. W., Zonneveld, J. P., & Henderson, C. M. (2008). Anoma- Sedimentary environments: Processes, facies, and stratigraphy
lously diverse Early Triassic ichnofossil assemblages in north- (pp. 37–82). Wiley.
west Pangea: A case for a shallow-marine habitable zone. Geol- Condit, D. D. (1912). The Conemaugh formation in Ohio. Ohio Geo-
ogy, 36(10), 771–774. logical Survey 4th Series. Bulletin, 17, 363.
Blakey, R.C. (2008). Gondwana paleogeography from assembly to Condron, L., Stark, C., O’Callaghan, M., Clinton, P., & Huang, Z.
breakup—a 500 my odyssey. In C. R. Fielding, T. D. Frank, & (2010). The role of microbial communities in the formation and
J. L. Isbell (Eds.), Resolving the Late Paleozoic Ice Age in time decomposition of soil organic matter. In G. R. Dixon & E. L.
and space. Geological Society of America. pp. 1–28 Tilston (Eds.), Soil microbiology and sustainable crop produc-
Boyce, C. K., & Lee, J. E. (2010). An exceptional role for flower- tion (pp. 81–118). Springer.
ing plant physiology in the expansion of tropical rainforests Crowell, J. C. (1995). The ending of the late Paleozoic ice age during
and biodiversity. Proceedings. Biological Sciences, 277(1699), the Permian Period. In P. A. Scholle, T. M. Peryt, & D. S. Ulmer-
3437–3443. https://​doi.​org/​10.​1098/​rspb.​2010.​0485 Scholle (Eds.), The Permian of Northern Pangea (pp. 62–74).
Brezinski, D. K., & Kollar, A. D. (2011). Pennsylvanian climatic events Springer.
and their congruent biotic responses in the central Appalachian DiMichele, W. A. (2014). Wetland-dryland vegetational dynamics in
Basin. Field Guides, 20, 45–60. the Pennsylvanian ice age tropics. International Journal of Plant
Brierley, G. J., Ferguson, R. J., & Woolfe, K. J. (1997). What is a fluvial Sciences, 175(2), 123–164.
levee? Sedimentary Geology, 114(1–4), 1–9. DiMichele, W. A., Pfefferkorn, H. W., & Phillips, T. L. (1996). Persis-
Bromley, R. G. (1996). Trace fossils: Biology, taphonomy and applica- tence of Late Carboniferous tropical vegetation during glacially
tions. Chapman and Hall. driven climatic and sea-level fluctuations. Palaeogeography,
Buatois, L. A., & Mángano, M. A. (2004). Animal-substrate interac- Palaeoclimatology, Palaeoecology, 125, 105–128.
tions in freshwater environments: applications of ichnology in DiMichele, W. A., Montañez, I. P., Poulsen, C. J., & Tabor, N. J.
facies and sequence stratigraphic analysis of fluvio-lacustrine (2009). Climate and vegetational regime shifts in the late Paleo-
successions. In D. Mcllroy (Ed.), The application of ichnology zoic ice age earth. Geobiology, 7(2), 200–226.
to palaeoenvironmental and stratigraphic analysis (Vol. 228, pp. Dimmitt, M. A. (2000). Biomes and communities of the Sonoran
311–334). Geological Society Special Publication, Geological Desert Region. In S. J. Phillips & P. WentworthComus (Eds.), A
Society of London. natural history of the Sonoran Desert: Tucson, AZ, US, Arizona-
Buatois, L., & Mángano, M. G. (2011). Ichnology: organism-substrate Sonora Desert (pp. 3–18). Museum Press.
interactions in space and time. Cambridge University Press. Dunkard Group, Pennsylvania-West Virginia-Ohio, USA. Journal of
Buatois, L. A., Mángano, M. G., Genise, J. F., & Taylor, T. N. (1998). Coal Geology, 119, 79–87.
The ichnologic record of the continental invertebrate invasion; Dzenowski, N. D., & Hembree, D. I. (2014). The neoichnology of two
evolutionary trends in environmental expansion, ecospace uti- terrestrial ambystomatid salamanders: Quantifying amphibian
lization, and behavioral complexity. Palaios, 13(3), 217–240. burrows using modern analogs. In D. I. Hembree, B. F. Platt, &
Cadée, G. C. (1984). ‘Opportunistic feeding’, a serious pitfall in trophic J. J. Smith (Eds.), Experimental approaches to understanding
structure analysis of (paleo) faunas. Lethaia, 17(4), 289–292. fossil organisms (pp. 305–341). Springer.

13
280 J. K. Crowell, D. I. Hembree

Ekdale, A. A. (1985). Paleoecology of the marine endobenthos. Pal- paleoecology: Reconstructing Cenozoic terrestrial environ-
aeogeography, Palaeoclimatology, Palaeoecology, 50(1), 63–81. ments and ecological communities (pp. 185–214). Springer.
Erwin, D. H. (2009). Climate as a driver of evolutionary change. Cur- https://​doi.​org/​10.​1007/​978-3-​319-​94265-0
rent Biology, 19(14), R575–R583. Hembree, D. I. (2019). Burrows and ichnofabric produced by cen-
Fedorko, N., & Skema, V. (2013). A review of the stratigraphy and tipedes: modern and ancient examples. Palaios, 34, 468–489.
stratigraphic nomenclature of the Dunkard Group in West Vir- https://​doi.​org/​10.​2110/​palo.​2019.​059
ginia and Pennsylvania, USA. International Journal of Coal Hembree, D. I. (2022). Early effects of the Late Paleozoic climate
Geology, 119, 2–20. transition on soil ecosystems of the Appalachian Basin (Cone-
Ferm, J. C. (1970). Allegheny deltaic deposits. In J. P. Morgan & R. maugh, Monongahela, and Dunkard groups): Evidence from
H. Shaver (Eds.), Deltaic sedimentation, modern and ancient ichnofossils. Palaios, 37(11), 671–690. https://​d oi.​o rg/​1 0.​
(Vol. 15, pp. 246–255). SEPM Special Publication Society of 2110/​palo.​2021.​071
Economic Paleontologists and Mineralogists. https://​doi.​org/​10.​ Hembree, D. I., & Blair, M. G. (2016). A paleopedological and ich-
2110/​pec.​70.​11 nological approach to interpreting spatial and temporal vari-
Frey, R. W., Pemberton, S. G., & Fagerstrom, J. A. (1984). Morpho- ability in Early Permian fluvial deposits of the lower Dunkard
logical, ethological, and environmental significance of the ich- Group, West Virginia, USA. Palaeogeography, Palaeoclimatol-
nogenera Scoyenia and Ancorichnus. Journal of Paleontology, ogy, Palaeoecology, 454, 246–266.
58(2), 511–528. Hembree, D. I., & Bowen, J. J. (2017). Paleosols and ichnofossils
Garcés, B. L. V., Gierlowski-Kordesch, E., & Bragonier, W. A. (1997). of the Upper Pennsylvanian-Lower Permian Monongahela
Pennsylvanian continental cyclothem development: No evidence and Dunkard groups (Ohio, USA): A multi-proxy approach
of direct climatic control in the Upper Freeport Formation (Alle- to unraveling complex variability in ancient terrestrial land-
gheny Group) of Pennsylvania (northern Appalachian Basin). scapes. Palaios, 32, 295–320.
Sedimentary Geology, 109(3–4), 305–319. Hembree, D. I., & Carnes, J. L. (2018). Response of soils and soil
Gastaldo, R. A., DiMichele, W. A., & Pfefferkorn, H. W. (1996). Out ecosystems to the Pennsylvanian–Permian climate transition
of the icehouse into the greenhouse: a late Paleozoic analog for in the upper fluvial plain of the Dunkard Basin, southeastern
modern global vegetational change. GSA Today, 6, 1–7. Ohio, USA. Geosciences, 8, 203.
Getty, P. R., McCarthy, T. D., Hsieh, S., & Bush, A. M. (2016). A new Hembree, D. I., & McFadden, C. J. (2020). Analysis of climate and
reconstruction of continental Treptichnus based on exceptionally landscape change through the Pennsylvanian and Permian
preserved material from the Jurassic of Massachusetts. Journal Monongahela and Dunkard Groups, Southeastern Ohio, USA.
of Paleontology, 90(2), 269–278. Journal of Sedimentary Environments, 5, 321–353.
Gingras, M. K., Bann, K. L., MacEachern, J. A., Waldron, J., & Pem- Hembree, D. I., & Nadon, G. C. (2011). A paleopedologic and ich-
berton, S. G. (2007). A conceptual framework for the applica- nologic perspective of the terrestrial Pennsylvanian landscape
tion of trace fossils. In J. A. MacEachern, K. L. Bann, M. K. in the distal Appalachian Basin, USA. Palaeogeography, Pal-
Gingras, & S. G. Pemberton (Eds.), Applied ichnology. SEPM aeoclimatology, Palaeoecology, 312, 138–166.
short course notes (Vol. 52, pp. 1–26). Society for Sedimentary Hembree, D. I., Smith, J. J., Buynevich, I. V., & Platt, B. F. (2017).
Geology. Neoichnology of semiarid environments: Soils and burrowing
Häntzschel, W. (1975). Treatise on invertebrate paleontology. In P. W. animals of the Sonoran Desert, Arizona, USA. Palaios, 32(9),
Miscellanea (Ed.), Supplement 1. Trace fossils and problematica 620–638.
(2nd ed.). The Geological Society of America. Hils, J. M., & Hembree, D. I. (2015). Neoichnology of the burrowing
Hasiotis, S. T. (2002). Continental trace fossils. SEPM Short Course spiders Gorgyrella inermis (Mygalomorphae: Idiopidae) and
Notes (Vol. 51). Society for Sedimentary Geology. Hogna lenta (Araneomorphae: Lycosidae). Palaeontologica
Hasiotis, S. T. (2004). Reconnaissance of Upper Jurassic Morrison Electronica. https://​doi.​org/​10.​26879/​500
Formation ichnofossils, Rocky Mountain Region, USA: Paleoen- Joeckel, R. M. (1995). Paleosols below the Ames Marine Unit (Upper
vironmental, stratigraphic, and paleoclimatic significance of Pennsylvanian, Conemaugh Group) in the Appalachian Basin,
terrestrial and freshwater ichnocoenoses. Sedimentary Geology, USA: Variability on an ancient depositional landscape. Journal
167(3–4), 177–268. of Sedimentary Research, 65(2a), 393–407.
Hasiotis, S. T., & Dubiel, R. F. (1993). Continental trace fossils of the Jumars, P. A., & Wheatcroft, R. A. (1989). Responses of benthos
Upper Triassic Chinle Formation, Petrified Forest National Park, to changing food quality and quantity, with a focus on deposit
Arizona. The nonmarine Triassic. Bulletin of the New Mexico feeding and bioturbation. In W. H. Berger, V. S. Smetacek, &
Museum of Natural History and Science, 3, 175–178. G. Wefer (Eds.), Productivity of the ocean: Present and past
Hasiotis, S. T., Dubiel, R. F., & Demko, T. M. (1998). A holistic (pp. 235–253). Wiley.
approach to reconstructing Triassic paleoecosystems: Using Kale, V. S., Patil, S. S., Satoskar, V., & Kumar, P. (1997). Occurrence
ichnofossils and paleosols as a basic framework. National Park of Planolites from the Nagarjuna Sagar Area, Northwestern
Service Paleontological Research, 3, 122–124. Cuddapah Basin. Journal of Geological Society of India, 49(5),
Hasiotis, S. T., Kraus, M. J., & Demko, T. M. (2007). Climatic controls 589–596.
on continental trace fossils. In W. Miller (Ed.), Trace fossils: Kondolf, G. M., Kattelmann, R., Embury, M., & Erman, D. C.
Concepts, problems, prospects (pp. 172–195). Elsevier. (1996). Status of riparian habitat. In C. I. Millar (Ed.), Sierra
Hembree, D. I. (2009). Neoichnology of burrowing millipedes: link- Nevada Ecosystem Project: Final report to Congress (pp.
ing modern burrow morphology, organism behavior, and sedi- 1009–1030). University of California.
ment properties to interpret continental ichnofossils. Palaios, 24, Koy, K., & Plotnick, R. E. (2007). Theoretical and experimental
425–439. https://​doi.​org/​10.​2110/​palo.​2008.​p08-​098r ichnology of mobile foraging. In W. Miller (Ed.), Trace fossils:
Hembree, D. I. (2017). Neoichnology of tarantulas (Araneae: Ther- Concepts, problems, prospects (pp. 428–441). Elsevier.
aphosidae): criteria for recognizing spider burrows in the fossil Lavelle, P., & Spain, A. V. (2005). Soil ecology. Springer Publishing.
record. Palaeontologia Electronica. https://d​ oi.o​ rg/1​ 0.2​ 6879/7​ 80 Lawrence, J. F., & Newton, A. F., Jr. (1982). Evolution and classifi-
Hembree, D. (2018). The role of continental trace fossils in Ceno- cation of beetles. Annual Review of Ecology and Systematics,
zoic paleoenvironmental and paleoecological reconstructions. 13(1), 261–290.
In D. A. Croft, D. F. Su, & S. W. Simpson (Eds.), Methods in

13
Climate‑induced changes in fluvial ichnofossil assemblages of the Pennsylvanian–Permian… 281

LeRoux, P. C., & McGeoch, M. A. (2008). Rapid range expansion Powell, M. G., Schöne, B. R., & Jacob, D. E. (2009). Tropical marine
and community reorganization in response to warming. Global climate during the late Paleozoic ice age using trace element
Change Biology, 14(12), 2950–2962. analyses of brachiopods. Palaeogeography, Palaeoclimatology,
Lopez, G. R. & Levinton, J. S. (1987). Ecology of deposit-feeding Palaeoecology, 280(1–2), 143–149.
animals in marine sediments. The Quarterly Review of Biol- Retallack, G. J. (2001). Soils of the past: An introduction to paleopedol-
ogy, 62(3), 235–260. http://​www.​jstor.​org/​stable/​28289​74 ogy. John Wiley and Sons.
Lucas, S. G. (2013). Vertebrate biostratigraphy and biochronology of Rindsberg, A. K., & Kopaska-Merkel, D. C. (2005). Treptichnus and
the upper Paleozoic Dunkard Group, Pennsylvania-West Vir- Arenicolites from the Steven C. Minkin Paleozoic footprint site
ginia-Ohio, USA. Journal of Coal Geology, 119, 79–87. (Langsettian, Alabama, USA). In R. J. Buta (Ed.), Pennsylvanian
MacEachern, J. A., Pemberton, S. G., Bann, K. L., & Gingras, M. footprints in the Black Warrior Basin of Alabama (pp. 121–141).
K. (2007). Departures from the archetypal ichnofacies: effective University of Alabama.
recognition of physico-chemical stresses in the rock record. In J. Roy, K., Valentine, J. W., Jablonski, D., & Kidwell, S. M. (1996).
A. MacEachern, K. L. Bann, M. K. Gingras, & S. G. Pemberton Scales of climatic variability and time averaging in Pleistocene
(Eds.), Applied ichnology SEPM Short Course Notes (Vol. 52, biotas: Implications for ecology and evolution. Trends in Ecology
pp. 65–93). Society for Sedimentary Geology. & Evolution, 11(11), 458–463.
Martin, W. D. (1998). Geology of the Dunkard Group (Upper Penn- Schneider, R. J., Lucas, S. G., & Barrick, J. E. (2013). The Early Per-
sylvanian-Lower Permian) in Ohio, West Virginia, and Pennsyl- mian age of the Dunkard Group, Appalachian basin, U.S.A.,
vania. Ohio Division of Geological Survey Bulletin, 73, 1–49. based on spiloblattinid insect biostratigraphy. International
Miall, A. D. (2010). Alluvial deposits. In N. P. James & R. W. Dalrym- Journal of Coal Geology, 119, 88–92.
ple (Eds.), Facies models 4 (pp. 105–137). Geological Associa- Stow, D. A. V. (2005). Sedimentary rocks in the field: a color guide.
tion of Canada. Gulf Professional Publishing.
Mikuś, P., & Uchman, A. (2013). Beetle burrows with a terminal cham- Sturgeon, M. T. (1958). The geology and mineral resources of Athens
ber: a contribution to the knowledge of the trace fossil Macano- County, Ohio. Ohio Division of Geological Survey Bulletin, 57,
psis in continental sediments. Palaios, 28, 403–413. https://​doi.​ 1–600.
org/​10.​2110/​palo.​2012.​p12-​129r Sturgeon, M. T., & Merrill, W. M. (1949). An additional fossiliferous
Milici, R. C. (2005). Appalachian coal assessment: defining the coal member in the Allegheny formation (Pennsylvanian) of Ohio.
systems of the Appalachian basin. In P. D. Warwick (Ed.), Coal The Ohio Journal of Science, 49(1), 1–11.
systems (Vol. 387, pp. 9–30). Geological Society of America Tabor, N. J., & Montañez, I. P. (2004). Morphology and distribution
Special Paper. Geological Society of America. of fossil soils in the Permo-Pennsylvanian Wichita and Bowie
Miller, M. F. (2003). Styles of behavioral complexity recorded by Groups, north-central Texas, USA: Implications for western
selected trace fossils. Palaeogeography, Palaeoclimatology, equatorial Pangean palaeoclimate during icehouse–greenhouse
Palaeoecology, 192(1–4), 33–43. transition. Sedimentology, 51(4), 851–884.
Miller, M. F., & Collinson, J. W. (1994). Trace fossils from Permian Tabor, N. J., & Poulsen, C. J. (2008). Palaeoclimate across the Late
and Triassic sandy braided stream deposits, central Transantarc- Pennsylvanian-Early Permian tropical palaeolatitudes: A review
tic Mountains. Palaios, 9(6), 605–610. of climate indicators, their distribution, and relation to palaeo-
Minter, N. J., Buatois, L. A., Mángano, M. G., Davies, N. S., Gibling, physiographic climate factors. Palaeogeography, Palaeoclimatol-
M. R., & Labandeira, C. (2016). The establishment of continen- ogy, Palaeoecology, 268, 293–310.
tal ecosystems. In M. G. Mángano & L. A. Buatois (Eds.), The Tabor, N. J., Montañez, I. P., Scotese, C. R., Poulsen, C. J., & Mack,
trace-fossil record of major evolutionary events: Mesozoic and G. H. (2008). Paleosol archives of environmental and climatic
Cenozoic (Vol. 2, pp. 205–324). Springer. history in paleotropical western Pangea during the latest Penn-
Montañez, I. P., & Cecil, C. B. (2013). Paleoenvironmental clues sylvanian through Early Permian. In C. R. Fielding, T. D. Frank,
archived in non-marine Pennsylvanian–lower Permian limestones & J. L. Isbell (Eds.), Resolving the Late Paleozoic Ice Age in
of the Central Appalachian Basin, USA. International Journal of time and space (pp. 291–304). Geological Society of America.
Coal Geology, 119, 41–55. Thacker, H. A., & Hembree, D. I. (2021). Neoichnological study of
Montañez, I. P., & Poulsen, C. J. (2013). The Late Paleozoic ice age: burrowing darkling beetles (Coleoptera: Tenebrionidae) from
An evolving paradigm. Annual Review of Earth and Planetary larval to adult stages. Ichnos, 28(4), 290–308.
Sciences, 41, 629–656. Underwood, E. C., Olson, D., Hollander, A. D., & Quinn, J. F. (2014).
Moore, J. C., Berlow, E. L., Coleman, D. C., de Ruiter, P. C., Dong, Ever-wet tropical forests as biodiversity refuges. Nature Climate
Q., Hastings, A., Collins Johnson, N., McCann, K. S., Melville, Change, 4(9), 740–741.
K., Morin, P. J., Nadelhoffer, K., Rosemond, A. D., Post, D. M., Vossler, S. M., & Pemberton, S. G. (1988). Skolithos in the Upper Cre-
Sabo, J. L., Scow, K. M., Vanni, M. J., & Wall, D. H. (2004). taceous Cardium Formation: An ichnofossil example of oppor-
Detritus, trophic dynamics and biodiversity. Ecology Letters, tunistic ecology. Lethaia, 21(4), 351–362.
7(7), 584–600. Walsh, R. P., Blake, W. H., Slaymaker, O., & Spencer, T. (2009). Tropi-
Morrissey, L. B., & Braddy, S. J. (2004). Terrestrial trace fossils from cal rainforests. In O. Slaymaker, T. Spencer, & C. Embleton-
the Lower Old Red sandstone, southwest Wales. Geological Hamnn (Eds.), Geomorphology and global environmental change
Journal, 39(3–4), 315–336. (pp. 214–257). Cambridge University Press.
Naiman, R. J., Decamps, H., & Pollock, M. (1993). The Role of ripar- Walther, G. R. (2010). Community and ecosystem responses to recent
ian corridors in maintaining regional biodiversity. Ecological climate change. Philosophical Transactions of the Royal Soci-
Applications, 3(2), 209–212. https://​doi.​org/​10.​2307/​19418​22 ety of London, Series B, Biological Sciences, 365, 2019–2024.
Nakamura, A., Proctor, H., & Catterall, C. P. (2003). Using soil and https://​doi.​org/​10.​1098/​rstb.​2010.​0021
litter arthropods to assess the state of rainforest restoration. Eco- Whittaker, R. H. (1953). A consideration of climax theory: The cli-
logical Management & Restoration, 4, S20–S28. max as a population and pattern. Ecological Monographs, 23(1),
Pianka, E. R. (1970). On r-and K-selection. The American Naturalist, 41–78.
104(940), 592–597.

13
282 J. K. Crowell, D. I. Hembree

Williams, E. G. (1960). Marine and fresh water fossiliferous beds in Publisher's Note Springer Nature remains neutral with regard to
the Pottsville and Allegheny Groups of western Pennsylvania. jurisdictional claims in published maps and institutional affiliations.
Journal of Paleontology, 34, 908–922.
Zhao, Z., Fan, R. Y., Zhang, L. J., Rodríguez-Tovar, F. J., & Gong, Y. Springer Nature or its licensor (e.g. a society or other partner) holds
M. (2020). Behavioural responses of Rhizocorallium to storm exclusive rights to this article under a publishing agreement with the
events: Evidence from the Middle Triassic of SW China. Pal- author(s) or other rightsholder(s); author self-archiving of the accepted
aeogeography, Palaeoclimatology, Palaeoecology, 545, 109640. manuscript version of this article is solely governed by the terms of
Zonneveld, J. P., Pemberton, S. G., Saunders, T. D., & Pickerill, R. such publishing agreement and applicable law.
K. (2002). Large, robust Cruziana from the Middle Triassic of
northeastern British Columbia: Ethologic, biostratigraphic, and
paleobiologic significance. Palaios, 17(5), 435–448.

13

You might also like