You are on page 1of 239

Bulletin 49

New Mexico Museum of Natural History & Science

A Division of the

DEPARTMENT OF CULTURAL AFFAIRS

Carboniferous-Permian transition in
Cañon del Cobre, northern New Mexico

edited by
Spencer G. Lucas, Jörg W. Schneider
and Justin A. Spielmann

Albuquerque, 2010
Carboniferous-Permian
J.A.
Spielmann,
and
J.W.
Schneider,
S.G.,
Lucas,
49
Bulletin
NMMNH&S
Mexico
New
northern
Cobre,
del
Cañon
in
transition
Bulletin 49

New Mexico Museum of Natural History & Science

A Division of the
DEPARTMENT OF CULTURAL AFFAIRS

Carboniferous-Permian transition in
Cañon del Cobre, northern New Mexico

edited by
Spencer G. Lucas, Jörg W. Schneider
and Justin A. Spielmann

Albuquerque, 2010
Bulletin 49

New Mexico Museum of Natural History & Science

A Division of the
DEPARTMENT OF CULTURAL AFFAIRS

Carboniferous-Permian transition in
Cañon del Cobre, northern New Mexico

by
Spencer G. Lucas, Jörg W. Schneider
and Justin A. Spielmann

New Mexico Museum of Natural History & Science

Printed with the support of the


Vortman Fund

Albuquerque, 2010
STATE OF NEW MEXICO
Department of Cultural Affairs
Stuart Ashman, Secretary

NEW MEXICO MUSEUM OF NATURAL HISTORY AND SCIENCE


Hollis J. Gillespie, Executive Director

BOARD OF TRUSTEES
Bill Richardson, Governor, State of New Mexico, ex officio
Hollis J. Gillespie, Executive Director, ex officio
Gary Friedman, President
Peter F. Gerity, Ph.D.
Laurence Lattman, Ph.D.
Morton Lieberman, Ph. D.
Imogene Lindsay, Emerita
Viola Martinez
Nancy J. McMillan, Ph.D.
John Montgomery, Ph.D.
Joseph Powell, Ph.D.
Dennis P. Trujillo, Ph.D.
Steve West

Cover illustration: Cast of holotype skeleton and flesh restoration of the diadectomorph Limnoscelis
paludis from the Upper Pennsylvanian of Cañon del Cobre, New Mexico. Graphics and illustration by
Mark A. Klingler, Carnegie Museum of Natural History. See Bermanet al., this volume.

Original Printing

ISSN:1524-4156

Available from the New Mexico Museum of Natural History and Science, 1801 Mountain Road NW,
Albuquerque, NM87104; Telephone (505) 841-2800; Fax (505)841-2866; www.nmnaturalhistory.org

NMMNH Bulletinsonline at: http://econtent.unm.edu/cdm4/browse.php?CISOROOT=%2Fbulletins


BULLETIN OF THE NEW MEXICO MUSEUM
OF NATURAL HISTORY AND SCIENCE

EDITORIAL STAFF

Spencer G. Lucas, Editor-in-Chief


Justin A. Spielmann, Managing Editor

EDITORIAL BOARD

GuillermoAlvarado Asociación Costarricense de Geotecnica, San José, Costa Rica


AndreaArcucci Universidad Nacional de San Luis, San Luis, Argentina
Marco Avanzini Museo Tridentino di Scienze Naturali, Trento, Italy
David Berman Carnegie Museum of Natural History, Pittsburgh, PA, USA
Brent Breithaupt Laramie, WY, USA
John R. Foster Museum of Western Colorado, Grand Junction, CO, USA
Gerard Gierlinski Polish Geological Institute, Warsaw, Poland
Jean Guex University of Lausanne, Lausanne, Switzerland
Jerald D. Harris Dixie State College, St. George, UT, USA
Andrew B. Heckert Appalachian State University, Boone, NC, USA
Adrian P. Hunt Flying Heritage Collection, Everett, WA, USA
Hendrik Klein Neumarkt, Germany
Heinz Kozur Budapest, Hungary
Karl Krainer University of Innsbruck, Innsbruck,Austria
Martin G. Lockley University Colorado at Denver, Denver, CO, USA
of
Spencer G. Lucas New Mexico Museum of Natural History and Science,
Albuquerque, NM, USA
Claudia Marsicano Universidad de Buenos Aires, Buenos Aires,Argentina
Jesper Milàn Institute for Geography and Geology, Copenhagen, Denmark
Gary S. Morgan New Mexico Museum of Natural History and Science,
Albuquerque, NM, USA
Donald R. Prothero Occidental College, Los Angeles, CA, USA
Silvio Renesto Università degli Studidell’Insubria, Varese, Italy
Vincent L. Santucci National Park Service, McLean, VA, USA
Joerg W. Schneider Technical University BergAkademie of Freiberg, Freiberg, Germany
Jingeng Sha Nanjing Institute of Geology and Palaeontology, Nanjing, China
Edward Simpson Kutztown University, Kutztown, PA, USA
Justin A. Spielmann New Mexico Museum of Natural History and Science,
Albuquerque, NM, USA
Robert M. Sullivan State Museum of Pennsylvania, Harrisburg, PA, USA
Lawrence H. Tanner Le Moyne College, Syracuse, NY, USA
Dana Ulmer-Scholle New Mexico Tech, Socorro, NM, USA
Ralf Werneburg Naturhistorisches Museum Schloss Bertholdsburg, Schleusingen, Germany
Richard S. White, Jr. International Wildlife Museum, Tucson,AZ, USA
NEWMEXICO MUSEUM OF NATURAL HISTORYAND SCIENCE BULLETINS
7. The postcranial morphology of Paleocene Chriacus and Mixodectes and the phylogenetic relationships of archontan
mammals, 1996. Frederick S. Szalay and Spencer G. Lucas, 47 pp.
8. The beginning of the age of mammals in the San Juan Basin, New Mexico: Biostratigraphy and evolution of
Paleocene mammals of the Nacimiento Formation, 1996. Thomas E. Williamson, 141 pp.
9. Paleontology of the Greenhorn Cyclothem (Cretaceous: Late Cenomanian to middle Turonian) at Black Mesa,
northeastern Arizona, 1996. James Ian Kirkland, 131 pp. + 50 pl.
10. Land snails of New Mexico, 1997. edited by Artie L. Metcalf and Richard A. Smartt, 145 pp.
11. New Mexico’s fossil record I, 1997. edited by Spencer G. Lucas, John W. Estep, Thomas E. Williamson, and Gary S. Morgan,
143 pp. (Out of Print)
12. Permian stratigraphy and paleontology of the Robledo Mountains, New Mexico, 1998. edited by Spencer G. Lucas, John
W. Estep, and Jerry M. Hoffer 130 pp.
13. Reanalysis of Acrocanthosaurus atokensis, its phylogenetic status, and paleobiogeographic implications, based on a
new specimen from Texas, 1998. Jerald D. Harris, 75 pp.
14. Lower and middle Cretaceous terrestrial ecosystems, 1998. Spencer G. Lucas, James I. Kirkland, and John W. Estep (eds.),
330 pp. (Out of Print)
15. A new skull of Parasaurolophus (Dinosauria: Hadrosauridae) from the Kirtland Formation of New Mexico and a
revision of the genus, 1999. Robert M. Sullivan and Thomas E. Williamson, 49 pp.
16. New Mexico’s fossil record II, 1999. edited by Spencer G. Lucas, 284 pp.
17. Dinosaurs of New Mexico, 2000. edited by Spencer G. Lucas and Andrew B. Heckert, 230 pp.
18. Volcanology in New Mexico, 2001. edited by L. S. Crumpler and Spencer G. Lucas, 150 pp.
19. Walter Granger, 1872-1941, paleontologist, 2002. by Vincent L. Morgan and Spencer G. Lucas, 58 pp.
20. Dinosores: An annotated bibliography of dinosaur paleopathology and related topics—1838-2001, 2002. by Darren H.
Tanke and Bruce M. Rothschild, 96 pp.
21. Upper Triassic stratigraphy and paleontology, 2002. edited by Andrew B. Heckert and Spencer G. Lucas, 301 pp.
22. Notes from Diary-Fayum Trip, 1907, 2002. by Vincent L. Morgan and Spencer G. Lucas, 148 pp.
23. Paleoecological analysis of the vertebrate fauna of the Morrison Formation (Upper Jurassic), Rocky Mountain
Region, U.S.A., 2003. by John R. Foster, 95 pp.
24. Paleontology and geology of the Upper Triassic (Revueltian) Snyder quarry, New Mexico, U.S.A., 2003. edited by Kate
E. Zeigler, Andrew B. Heckert, and Spencer G. Lucas, 132 pp.
25. Carboniferous-Permian transition at Carrizo Arroyo, central New Mexico, 2004. edited by Spencer G. Lucas and Kate E.
Zeigler, 301 pp.
26. Paleogene Mammals, 2004. edited by Spencer G. Lucas, Kate E. Zeigler, and Peter E. Kondrashov, 230 pp.
27. Late Triassic microvertebrates from the lower Chinle Group (Otischalkian-Adamanian: Carnian), 2004. by Andrew
B. Heckert, 170 pp.
28. New Mexico’s Ice Ages, 2005. edited by Spencer G. Lucas, Gary S. Morgan, and Kate E. Zeigler, 300+ pp.
29. Vertebrate Paleontology in Arizona, 2005. edited by Andrew B. Heckert & Spencer G. Lucas, 210 pp.
30. The Nonmarine Permian, 2005. edited by Spencer G. Lucas & Kate E. Zeigler, 362 pp.
31. Permian of Central New Mexico, 2005. edited by Spencer G. Lucas, Kate E. Zeigler & Justin A Spielmann, 176 pp.
32. Genética y Mamíferos Mexicanos: Presente y Futuro, 2006. edited by Ella Vázquez-Domínguez and David J. Hafner, 73 pp.
33. Skeletal Impact of Disease, 2006. by Bruce M. Rothschild and Larry D. Martin, 226 pp.
34. America’s Antiquities: 100 Years of Managing Fossils on Federal Lands, 2006. edited by Spencer G. Lucas, Justin A.
Spielmann, Patricia M. Hester, Jason P. Kenworthy and Vincent L. Santucci, 185 pp.
35. Late Cretaceous Vertebrates from the Western Interior, 2006. edited by Spencer G. Lucas and Robert M. Sullivan, 410 pp.
36. Paleontology and Geology of the Upper Jurassic Morrison Formation, 2006. edited by John R. Foster and Spencer G.
Lucas, 249 pp.
37. The Triassic-Jurassic Terrestrial Transition, 2006. edited by Jerry D. Harris, Spencer G. Lucas, Justin A. Spielmann, Martin G.
Lockley, Andrew R.C. Milner and James I. Kirkland, 607 pp.
38. Pennsylvanian-Permian Fusulinaceans of the Big Hatchet Mountains, New Mexico, 2006. by Garner L. Wilde, 331 pp.
39. Upper Aptian-Albian Bivalves of Texas and Sonora: Biostratigraphic, Paleoecologic and Biogeographic Implications,
2007. edited by Robert W. Scott, 39 pp.
40. Triassic of the American West, 2007. edited by Spencer G. Lucas and Justin A. Spielmann, 247 pp.
41. The Global Triassic, 2007. edited by Spencer G. Lucas and Justin A. Spielmann, 415 pp.
42. Cenozoic Vertebrate Tracks and Traces, 2007. edited by Spencer G. Lucas, Justin A. Spielmann and Martin G. Lockley, 330
pp.
43. The Late Triassic archosauromorph Trilophosaurus, 2008. by Justin A. Spielmann, Spencer G. Lucas, Larry F. Rinehart and
Andrew B. Heckert, 177 pp.
44. Neogene Mammals, 2008. edited by Spencer G. Lucas, Gary S. Morgan, Justin A. Spielmann and Donald R. Prothero, 442 pp.
45. The Paleobiology of Coelophysis bauri (Cope) from the Upper Triassic (Apachean) Whitaker quarry, New Mexico, with
detailed analysis of a single quarry block, 2009. by Larry F. Rinehart, Spencer G. Lucas, Andrew B. Heckert, Justin A.
Spielmann and Matthew D. Celeskey, 260 pp.
46. The taxonomy and paleobiology of the Late Triassic (Carnian-Norian: Adamanian-Apachean) drepanosaurs
(Diapsida: Archosauromorpha: Drepanosauromorpha, 2010. by Silvio Renesto, Justin A. Spielmann, Spencer G. Lucas and
Giorgio Tarditi Spagnoli, 81 pp.
47. Ichnology of the Upper Triassic (Apachean) Redonda Formation, east-central New Mexico, 2010. by Spencer G. Lucas,
Justin A. Spielmann, Hendrik Klein and Allan J Lerner, 75 pp.
48. New Smithian (Early Triassic) ammonoids from Crittenden Springs, Elko County, Nevada: Implications for
taxonomy, biostratigraphy and biogeography, 2010. by James F. Jenks, Arnaud Brayard, Thomas Brühwiler and Hugo Bucher,
41 pp.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

i
TABLE OF CONTENTS

Introduction

Carboniferous-Permian transition in Cañon del Cobre, northern New Mexico:


An overview.......................................................Spencer G. Lucas, Jörg W. Schneider and JustinA. Spielmann 1

One hundred and fifty years of geological and paleontological research in Cañon del Cobre, northern New
Mexico..................................................................Spencer G. Lucas, Susan K. Harris and Justin A. Spielmann 7

Geology

Summary of geology of Cañon del Cobre, Rio Arriba County, New Mexico......................Spencer G. Lucas,
Karl Krainer and JustinA. Spielmann 15

Sedimentology of the Pennsylvanian-Permian Cutler Group and Lower Permian Abo Formation, northern
New Mexico..................................................................................................Karl Krainer and Spencer Lucas 25

Ichnology

Pennsylvanian lake-margin ichnoassemblage from Cañon del Cobre, Rio Arriba County,
New Mexico...........................................................................................Spencer G. Lucas and Allan JLerner 37

Euramerican Late Pennsylvanian/Early Permian arthropleurid/tetrapod associations - implications for the


habitat and paleobiology of the largest terrestrial arthropod...........Jörg W. Schneider, Spencer G. Lucas,
Ralf Werneburg and Ronny Rößler 49

Palynology/Paleobotany

Palynological investigation of the Upper Pennsylvanian (Carboniferous) El Cobre Canyon Formation, Cutler
Group, Cañon del Cobre, RioArriba County, New Mexico, U.S.A..........John Uttingand Spencer G. Lucas 71

Late Pennsylvanian floras in western equatorial Pangea, Cañon del Cobre,


New Mexico..................................WilliamA. Dimichele, Dan S. Chaney, Hans Kerp and Spencer G. Lucas 75

Vertebrate Paleontology

Vertebrate paleontology, biostratigraphy and biochronology of the Pennsylvanian-Permian Cutler Group,


Cañon del Cobre, northern New Mexico.............…...............Spencer G. Lucas, Susan K. Harris, Justin A.
Spielmann, Larry F. Rinehart, David SBerman,Amy C. Henriciand Karl Krainer 115

Dissorophoid record from the Upper Pennsylvanian of Cañon del Cobre, New Mexico............Ralf Werneburg,
David SBerman and Spencer G. Lucas 125

First Pennsylvanian Eryops (Temnospondyli) and its Permian recordfrom New Mexico............RalfWerneburg,
Spencer G. Lucas, Jörg W. Schneider and Larry F. Rinehart 129
ii
A partial skeleton of Ophiacodon navajovicus (Eupelycosauria: Ophiacodontidae) from the Upper
Pennsylvanian ofCañondel Cobre, New Mexico....Susan K. Harris,SpencerG.Lucas and JustinA. Spielmann 137

Re-evalutaion of Ruthiromia elcobriensis (Eupelycosauria: Ophiacodontidae?) from the Lower Permian


(Seymouran?) of Cañon del Cobre, north-central New Mexico....JustinA. Spielmann and Spencer G. Lucas 151

Redescription of the cranial anatomy of Sphenacodon ferox Marsh (Eupelycosauria: Sphenacodontidae)


from the Late Pennsylvanian-Early Permian ofNew Mexico....................JustinA. Spielmann,Larry F.Rinehart,
Spencer G. Lucas, David SBerman, Amy C. Henrici and Susan K. Harris 159

Redescription of the skull of Limnoscelis paludis Williston (Diadectomorpha:


Limnoscelidae)…..............................................................David SBerman, Robert R. Reisz and Diane Scott 185

Description of the postcranial skeleton of Limnoscelis (Diadectomorpha: Limnoscelidae) from the Upper
Pennsylvanian of northern New Mexico and central Colorado…….....................................Natalia K. Kennedy 211

Typothorax coccinarum (Archosauria: Stagonolepididae) from the Upper Triassic (Revueltian) Petrified
Forest Formation, El Puertocito, Cañon del Cobre, RioArriba County, New Mexico...........JustinA.Spielmann
and Spencer G. Lucas 221
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

1
CARBONIFEROUS-PERMIAN TRANSITION IN CAÑON DEL COBRE,
NORTHERN NEWMEXICO: ANOVERVIEW

SPENCER G. LUCAS1, JÖRGW. SCHNEIDER2 AND JUSTINA. SPIELMANN1


1New Mexico Museum of Natural History and Science, 1801 Mountain Road N.W., Albuquerque, New Mexico 87104;
2TU Bergakademie Freiberg, Institut für Geologie, B.v.Cotta-Strasse 2, D-09596 Freiberg, Germany

Abstract—Cañon del Cobre is a large box canyon in southeastern Rio Arriba County, New Mexico, along the
southeastern border of the Colorado Plateau. The canyon exposes an ~400 m thick section of Upper Pennsylva
nian-Lower Permian fluvial red beds (El Cobre Canyon and Arroyo del Agua formations of the Cutler Group) that
accumulated under wet-subhumid to semiarid climatic conditions. These beds yield palynomorphs, megafossil
plants, trace fossils and vertebrate fossils. The paleofloral and paleofaunal components are the main focus of the
articles in this volume. The red-bed succession across the Carboniferous-Permian boundary in Cañon del Cobre
documents increasing aridity that caused a slight drop in paleofloral diversity and a change in composition to
greater abudance of conifers in the paleoflora. These changes, however, are not mirrored by tetrapod assemblages,
which do not change substantially across the Carboniferous-Permian boundary.

INTRODUCTION GEOLOGY
Cañon del Cobre is a large box canyon in southeastern Rio Arriba Lucas, Spielmann and Krainer here review the geology of Cañon
County, New Mexico, along the southeastern border of the Colorado del Cobre. The canyon is traversed and bounded by several north-north
Plateau. The main portion of the canyon is a complexly faulted valley east-trending faults, and its eastern wall is along a faulted, north-north
with extensive exposures of the El Cobre Canyon and overlying Arroyo east-trending anticline, structures that were activated by Neogene exten
del Agua formations of the Carboniferous-Permian Cutler Group (Lucas sional tectonism of the nearby Rio Grande rift. The oldest rocks that
and Krainer, 2005; Lucas et al., 2005b; Kempter et al., 2007). These floor Cañon del Cobre are siliciclastic red beds of the El Cobre Canyon
Cutler Group strata were deposited by syntectonic fluvial systems in Formation of the Cutler Group (Fig. 2). An extensive unconformity
relatively upland and inland settings along the southern flank of the separates Cutler Group strata from overlying Upper Triassic strata of
Uncompahgre uplift of the ancestral Rocky Mountains (Fig.1). the Chinle Group around the rim of Cañon del Cobre (in this volume,
This volume documents geological and paleontological research in Spielmann and Lucas document aetosaur fossils collected from Chinle
Cañon del Cobre, most of which took place during the last decade as a strata near the mouth of Cañon del Cobre). Cenozoic bedrock of Eocene
collaborative project coordinated by the New Mexico Museum of Natu and Oligocene age is locally exposed around the rim of Cañon del Cobre.
ral History and Science. This research uncovered numerous new fossils Quaternary deposits – quartzite erratic boulders, alluvial terrace depos
in the canyon, especially invertebrate traces, palynomorphs, megafossil its, colluvium and alluvium – crop out in and around Cañon del Cobre.
plants and vertebrates. Some of these fossils have provided the basis for In this volume, Krainer and Lucas review the sedimentology of
reasonably certain (but not exact) placement of the Carboniferous-Per Cutler Group strata in Cañon del Cobre, confirming observations of
mian boundary in the red-bed section exposed at Cañon del Cobre. Also, earlier studies by Eberth (see especially Eberth and Miall, 1991). Thus,
these fossils and their enclosing sediments document relatively little the El Cobre Canyon and Arroyo del Agua formations are red beds that
biotic change across the system boundary. In this article, we provide an formed in river systems under semi-arid to arid climatic conditions. Sand
introduction to and overview of the articles in this volume and summa stones are dominantly arkoses to litharenites, and rare sublitharenites.
rize some of the trends across the Carboniferous-Permian boundary Sandstone sheets represent deposits of a braided river system; the inter
evident in Cañon del Cobre. calated sandstone beds and lenses are interpreted as sheet splays and
minor channel fills that formed during overbank flooding. Siltstone and
HISTORY OF STUDY mudstone units represent floodplain deposits. The exposed El Cobre
In this volume, Lucas, Harris and Spielmann recount the his- Canyon Formation is about 230 m thick (total thickness ~ 640 m) and is
tory of geological and paleontological studies in Canon del Cobre, which mostly siltstone and sandstone deposited in an ephemeral braided stream
began with J.S. Newberry in 1859, who discovered Triassic plant fossils environment. The overlying Early Permian Arroyo del Agua Formation
in the copper mines in the canyon (Newberry, 1876). David Baldwin is as much as 160 m thick, and also consists mostly of siltstone and
first discovered Carboniferous-Permian tetrapod fossils in Cañon del sandstone, but was deposited on laterally extensive floodplains that
Cobre in the late 1870s, and this led to subsequent searches for Carbon developed between relatively stable channels.
iferous-Permian tetrapod fossils. S.W. Williston and E.C. Case, working ICHNOLOGY
in 1911, first recognized that the oldest tetrapod fossils from Cañon del
Cobre are of Carboniferous age (Williston and Case, 1912), and subse Prior to the last decade, no trace fossils were reported from Cañon
quent research has supported this conclusion (e.g., Vaughn, 1963; Fracasso, del Cobre. Recent discoveries have highlighted two localized records in
1980; Hunt and Lucas, 1992; Lucas et al., 2005b) (Fig. 2). However, the canyon—a lake margin invertebrate and vertebrate ichnoassemblage
most of the bedrock strata in Cañon del Cobre were assigned a Triassic and a site with large Diplichnites, which are the trackways of the largest
age until 1961, when C. Smith, A. Budding and C.W. Pitratassigned them known terrestrial arthropod, Arthropleura.
to the Permian Cutler Formation. Collecting and research in Cañon del In Cañon del Cobre, rare lacustrine strata (thinly laminated silt
Cobre has now identified invertebrate and vertebrate traces, stones) of the El Cobre Canyon Formation represent shallow, localized
palynomorphs, plants and tetrapod bones and teeth of Carboniferous floodplain lakes. In this volume, Lucas and Lerner document the first
age, traces and tetrapods of Early Permian age and Late Triassic tetrapod ichnofossil assemblage known from such a Pennsylvanian intermontane
and plant fossils. Much of this volume focuses on documenting these lake deposit, and it consists of arthropod trackways, invertebrate graz
fossils. ing and feeding traces, tetrapod footprints and fish swimming trails. This
2
maturity and indicate a Late Pennsylvanian (Carboniferous) age not older
than Stephanian B. Fracasso (1980; also see Hunt and Lucas, 1992) also
reported a Pennsylvanian megafossil plant locality in the floor of Cañon
del Cobre.
During the last decade, an extensive search for fossil plants of
Carboniferous-Permian age in Cañon del Cobre identified localities at
five stratigraphic levels in the middle of the El Cobre Canyon Formation.
In this volume, DiMichele, Chaney, Kerp and Lucas document these
floras, which are compositionally typical of the Late Pennsylvanian,
specifically of the Virgilian (Gzhelian). Conspicuous elements are
Alethopteris zeilleri, Macroneuropteris scheuchzeri, Danaeites emersonii,
Sphenophyllum verticillatum, S. oblongifolium, and S. angustifolium,
Annularia carinata, Asterophyllites equisetiformis, Sigillaria brardii, and
other, less common forms typical of wetland floras, such as
Pseudomariopteris cordato-ovata. Walchian conifers and Taeniopteris
are rare. These floras mostly appear to have been growing on bars or
along the margins of braided streams, resulting in frequent burial of plants,
particularly calamite stems, in upright position, or the preservation of
tangled masses of sphenophyll stems and leaves, largely in siltstones.
Finer-grained deposits, seemingly representing small lakes, were strongly
dominated by pteridosperms.
VERTEBRATE PALEONTOLOGY
In the late 1870s, David Baldwin discovered fossil vertebrates in
the Carboniferous-Permian red beds of Cañon del Cobre, and since then
collecting has yielded unique assemblages of fossil vertebrates that are
among the oldest tetrapod fossils from North America. In this volume,
Lucas, Harris, Spielmann, Rinehart, Berman and Henrici review
and update Lucas et al. (2005b), which presented a detailed vertebrate
biostratigraphy in Cañon del Cobre constructed from all localities that
can be placed into precise lithostratigraphic position (Fig. 2). This bio
FIGURE 1. Principal uplifts (gray), basins and faults associated with the late stratigraphy identifies three, stratigraphically-successive vertebrate as
Paleozoic ancestral Rocky Mountains orogeny and location of Cañon del semblages that support recognition of three time-successive land verte
Cobre in New Mexico (after Woodward et al., 1999). Faults are: A = brate faunachrons (LVFs) across the Pennsylvanian-Permian boundary:
Albuquerque fault, E = Estancia fault, J = Joyita fault, P = Peñasco fault, PP Cobrean, Coyotean and Seymouran (also see Lucas, 2006). The Penn
= Pecos-Picuris fault, TC = Tijeras-Cañoncito fault and TP = Tusas-Picuris sylvanian-Permian boundary is in the Coyotean LVF.
fault. The other articles on vertebrate paleontology in this volume focus
on describing new fossils, documenting new anatomy and interpreting
ichnofossil assemblage is assigned to the Scoyenia ichnofacies and is these fossils within a taxonomic and/or biostratigraphic context. Thus,
composed of members of the Diplichnites and Mermia ichnoguilds, con Werneburg, Berman and Lucas describe the snout region of a large
sistent with a shallow lacustrine setting. The relatively low ichnodiversity dissorophoid skull from the Late Pennsylvanian interval of the El Cobre
of the Cañon del Cobre ichnofossil assemblage indicates a short-lived
Canyon Formation. Werneburg, Lucas, Schneider and Rinehart docu
water body colonized by a mobile invertebrate epifauna of detritus feed ment a well-preserved, partial skull of the temnospondyl amphibian
ers and predators that provided a food source for tetrapod predators.
Eryops from the Pennsylvanian strata of the El Cobre Canyon Forma
In this volume, Schneider, Lucas, Werneburg and Rössler re
tion. This Pennsylvanian Eryops is compared to Permian Eryops, includ
view the record of the giant arthropod Arthropleura, which was a com ing specimens from the Early Permian of New Mexico, but no species
mon member of the late Paleozoic continental biota of paleo-equatorial
assignment is possible.
biomes for more than 35 million years, from the middle Mississippian to
The most famous vertebrate fossil from Cañon del Cobre is the
the Early Permian. In Late Pennsylvanian red beds in Cañon del Cobre,
type specimen of Limnoscelis paludis, discovered by David Baldwin in
trackways of Arthropleura are present in strata that also yield body
1879 and first described by Williston (1911). Long considered pivotal to
fossils of the amphibian Eryops (Lucas et al., 2005a). A review of the understanding reptile (amniote) origins, Limnoscelis is now considered a
Arthropleura tracksite from Cañon del Cobre, as well as other tracksites diadectomorph. The original type specimen from Cañon del Cobre re
of this animal and arthropleurid/eryopid associations, supports the con
mains the best preserved and most complete fossil known of the taxon.
clusion that Arthropleura was well adapted to alluvial environments of In this volume, Berman provides a detailed redescription of the skull
ever wet humid to seasonally dry and semihumid climates. Preferred that corrects errors in previous descriptions and adds substantially to
habitats of semi-adult and adult Arthropleura were open, vegetated, river knowledge of the cranial anatomy of Limnoscelis. Kennedy provides a
landscapes, where these giant arthropods co-existed with semi-aquatic companion piece that redescribes the postcranial anatomy of Limnoscelis,
eryopid amphibians and terrestrial pelycosaurs. largely based on the superb skeleton that David Baldwin collected in
PALEOBOTANY Cañon del Cobre.
A recently discovered partial skeleton of the pelycosaurian-grade
Smith et al. (1961) first reported Carboniferous plants from Cañon synapsid Ophiacodon navajovicus, the most complete yet found, from
del Cobre. Their locality, which is in the middle part of the El Cobre the Late Pennsylvanian interval of the El Cobre Canyon Formation, is
Canyon Formation, also yields palynomorphs reported on by Utting described by Harris, Lucas and Spielmann. Its stratigraphic distribu
and Lucas in this volume. These are mostly miospores of low thermal tion in New Mexico, Utah and Colorado indicate that O. navajovicus
3

FIGURE 2. Composite lithostratigraphic section in Cañon del Cobre, showing distribution of fossil localities (numbers are localities of the New Mexico
Museum of Natural History and Science) and taxa.
4
characterizes the Cobrean LVF.
Ruthiromia elcobriensis is an eupelycosaur from the Permian strata
of the Cañon del Cobre known only from the holotype, a partial skel
eton. The type locality of R. elcobriensis has never been rediscovered.
However, given available information, Spielmann and Lucas conclude
it is in the lower part of the Arroyo del Agua Formation. Their reconsid
eration of its anatomy places R. elcobriensis outside of the Varanopsidae
and tentatively within the Ophiacodontidae.
Sphenacodon is a pelycosaurian-grade synapsid, best known from
the Early Permian of Rio Arriba County, northern New Mexico. Of the FIGURE 3. Summary of key trends (and non-trends) during the Carboniferous
two species (S. ferox and S. ferocior), S. ferox is known from compara Permian transition recorded in the Cutler Group strata of Cañon del Cobre
tively little skull material. Here, Spielmann, Rinehart, Lucas, Berman, and nearby outcrops in Rio Arriba County, New Mexico.
Henrici and Harris describe a newly collected, nearly complete skull of
S. ferox that demonstrates numerous cranial differences between the two Virgilian/Gzhelian). During the Late Pennsylvanian-Early Per
species of Sphenacodon. These features, and metric analysis, support mian, each successive paleoflora in the Euramerican paleo
the concept of two distinct species of Sphenacodon. S. ferox has a tem equatorial belt was increasingly dominated by mesophilous
poral range from the Late Pennsylvanian (Coyotean LVF) through the and xerophilous elements, especially by an increase in conifers
Early Permian (Seymouran LVF), whereas S. ferocior is restricted to the (“walchians”) (Fig. 3). The rarity of walchians in the red
Coyotean LVF, but does span the Pennsylvanian-Permian boundary. beds of Cañon del Cobre points to a relatively wet-subhumid
to perhaps dry-subhumid climate, most similar to the red
CARBONIFEROUS-PERMIAN TRANSITION beds of the European Stephanian B (Roscher and Schneider,
Megafossil plants, palynomorphs and fossil vertebrates indicate 2005, 2006). Nevertheless, it should be taken into account that
that the Pennsylvanian-Permian boundary (younger than the Virgilian regional orographic situations can modify this picture. Thus,
for example, at the southern border of the Late Pennsylvanian
Wolfcampian boundary) is stratigraphically high in the El Cobre Canyon
Formation in Cañon del Cobre, though the exact position of the bound Hercynian low mountain belt, in the Mauretanides of
ary is not certain (Fig. 2). Indeed, without biostratigraphic data based on Morocco, in addition to the typical hygrophilous Stephanian
phylogenetic lineages of successive zone-species, as described by paleofloral elements, the mesophilous to xerophilous walchians
and callipterids are very common as early as the Stephanian B
Werneburg and Schneider (2006) for amphibians, for insects as reported
by Schneider and Werneburg (2006) and supported by isotopic ages gray facies of the Moroccan Souss basin (Hmich et al. 2006).
Such orographic effects, however, apparently did not influ
(Roscher and Schneider 2005, Lützner et al. 2007), determining the exact
ence paleofloral composition across the Carboniferous-Per
placement of the Pennsylvanian/Permian boundary in nonmarine set
tings, such as Cañon del Cobre, will remain difficult. mian transition in northern New Mexico.
The fossil record of the Carboniferous-Permian transition in Cañon 4. The terrestrial tetrapod assemblages remain substantially
del Cobre is more extensive for the Pennsylvanian portion of the section unchanged from Late Pennsylvanian to Early Permian - they
than for the Early Permian (Fig. 2). Fortunately, a diverse Early Permian represent what can be termed the Coyotean chronofauna domi
record of plants and vertebrates is known from the upper part of the El nated by eryopid temnospondyls, diadectomorphs and
Cobre Canyon Formation at Arroyo del Agua, about 25 km southwest of sphenacodontid pelycosaurs (Lucas, 2006) (Fig. 3). In con
trast, for the environment of aquatic animals the cyclical in
Cañon del Cobre (e.g., DiMichele and Chaney, 2005; Lucas et al., 2005c).
Using the Arroyo del Agua record to enhance the Early Permian record, crease in aridity resulted in a directed development of lake and
the following trends (or lack of trends) are evident across the Carbonifer river types with associated aquatic invertebrate and vertebrate
ous-Permian boundary at Cañon del Cobre (Fig. 3): assemblages. The aquatic lake faunas of the European Penn
1. There is a change in the architecture of the fluvially-deposited sylvanian/Permian transition thus are characterized by a strong
red beds from braided systems of the Pennsylvanian to more decline in diversity caused by climate changes and tectonic
laterally extensive floodplains (with mature caliches) of the reorganization of basins and river systems (Franconian vol
Early Permian. Eberth and Miall (1991) attributed this change cano-tectonic event: Schneider et al. 1995, 2000). This signal
to increasing aridity. The presence of a thick section of eolianites has so far not been reported from the Ouachita-Appalachian
at the base of the Yeso Group (DeChelly Sandstone) immedi region, and is also not evident in the ancestral Rocky Moun
ately above the Arroyo del Agua Formation in the Arroyo del tain basin of northern New Mexico.
Agua area is evidence of even increased aridity by approxi The red-bed succession across the Carboniferous-Permian bound
mately the beginning of Leonardian time. ary in Cañon del Cobre thus documents increasing aridity that caused a
2. The paleoflora declines in diversity from the Pennsylvanian to slight drop in paleofloral diversity and a change in composition to a more
the Permian (though the species pool for both time periods is conifer-dominated paleoflora. These changes, however, are not mirrored
by tetrapod assemblages, which do not change substantially across the
substantially unchanged), a trend also seen in the co-eval north
Texas section (DiMichele and Chaney, 2005) (Fig. 3). Carboniferous-Permian boundary.
3. As shown by Roscher and Schneider (2006), the generally
ACKNOWLEDGMENTS
increasing aridity during the Permian reflects a cyclical
sequence of wet and dry phases in which each subsequent We are grateful to numerous colleagues and volunteers for collabo
phase progressively becomes dryer. Therefore, the paleoflora ration in the field in Cañon del Cobre. The U. S. Forest Service permitted
of the red beds in Cañon del Cobre is very similar in composi our fieldwork in Cañon del Cobre. William DiMichele and Karl Krainer
tion to the Late Pennsylvanian gray facies (particularly of the provided helpful comments on an earlier draft of the manuscript.
5
REFERENCES

DiMichele, W.A. and Chaney, D.S., 2005, Pennsylvanian-Permian fossil Lützner, H., Littmann, S., Mädler, J., Romer, R.L. and Schneider, J.W.,
floras from the Cutler Group, Cañon del Cobre and Arroyo del Agua 2007, Stratigraphic and radiometric age data for the continental
areas, in northern New Mexico: New Mexico Museum of Natural His- Permocarboniferous reference-section Thüringer-Wald, Germany: Pro
tory and Science, Bulletin 31, p. 26-33. ceedings XVth International Congress on Carboniferous and Permian
Eberth, D.A. and Miall, A.D., 1991, Stratigraphy, sedimentology and evolu Stratigraphy 2003, Utrecht, p. 161-174.
tion of a vertebrate-bearing, braided to anastomosed fluvial system, Newberry, J.S., 1876, Geological report; in Macomb, J.N., ed., Report of the
Cutler Formation (Permian-Pennsylvanian), north-central New Mexico: exploring expedition from Santa Fe…in 1859: U.S. Army Engineer
Sedimentary Geology, v. 72, p. 225-252. Department, Washington, D.C., p. 9-118.
Fracasso, M.A., 1980, Age of the Permo-Carboniferous Cutler Formation Roscher, M. and Schneider, J.W., 2005, An annotated correlation chart for
vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale continental Late Pennsylvanian and Permian basins and the marine
ontology, v. 54, p. 1237-1244. scale: New Mexico Museum of Natural History and Science, Bulletin 30,
Hunt, A.P. and Lucas, S.G., 1992, The paleoflora of the lower Cutler Forma p. 282-291
tion (Pennsylvanian, Desmoinesian?) in El Cobre Canyon, New Mexico, Roscher, M. and Schneider, J.W., 2006, Early Pennsylvanian to Late Per
and its biochronological significance: New Mexico Geological Society, mian climatic development of central Europe in a regional and global
Guidebook 43, p. 145-150. context; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds., Nonma
Hmich, D., Schneider, J.W., Saber, H., Voigt, S. and El Wartiti, M., 2006, rine Permian chronology and correlation: Geological Society, London,
New continental Carboniferous and Permian faunas of Morocco – im- Special Publication 265, p. 95-136.
plications for biostratigraphy, palaeobiogeography and palaeoclimate; Schneider, J. W. and Werneburg, R., 2006, Insect biostratigraphy of the
in Lucas, S.G., Cassinis, G. and Schneider J.W., eds., Non-marine Per European Late Carboniferous and Early Permian; in Lucas, S.G., Cassinis,
mian biostratigraphy and biochronology: Geological Society, London, G. and Schneider J.W., eds., Non-marine Permian biostratigraphy and
Special Publication 265, p. 297-324. biochronology: Geological Society, London, Special Publication 265, p.
Kempter, K., Zeigler, K., Koning, D. and Lucas, S., 2007, Preliminary 325-336.
geologic map of the Canjilon SE quadrangle, Rio Arriba County, New Schneider, J. W., Rössler, R. and Gaitzsch, B., 1995, Time lines of the late
Mexico: New Mexico Bureau of Geology and Mineral Resources, Open Variscan volcanism - holostratigraphic synthesis: Zentralblatt für
file Geological Map 150, 1:24,000. Geologie und Paläontologie Teil I, v. 5/6, 477-490.
Lucas, S.G., 2006, Global Permian tetrapod biostratigraphy and Schneider, J.W., Hampe, O. and Soler-Gijón, R., 2000, The Late Carbonif
biochronology; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds., erous and Permian: Aquatic vertebrate zonation in southern Spain and
Non-marine Permian biostratigraphy and biochronology: Geological German basins.-IGCP 328. Final Report: Courier Forschung-Institut
Society, London, Special Publication 265, p. 65-93. Senckenberg, p. 543-561.
Lucas, S.G. and Krainer, K., 2005, Stratigraphy and correlation of the Smith, C. T., Budding, A. J. and Pitrat, C. W., 1961, Geology of the south
Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New eastern part of the Chama Basin: New Mexico Bureau of Mines and
Mexico Geological Society, Guidebook 56, p. 145-159. Mineral Resources, Bulletin 75, 57 p
Lucas, S.G., Lerner, A.J, Hannibal, J.T., Hunt, A.P. and Schneider, J.W., Vaughn, P.P., 1963, The age and locality of the late Paleozoic vertebrates
2005a, Trackway of a giant Arthropleura from the Upper Pennsylva from El Cobre Canyon, Rio Arriba County, New Mexico: Journal of
nian of El Cobre Canyon, New Mexico: New Mexico Geological Society, Paleontology, v. 37, p. 283-296.
Guidebook 56, p. 279-282. Werneburg, R. and Schneider, J.W., 2006, Amphibian biostratigraphy of the
Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S and Henrici, A.C., European Permo-Carboniferous; in Lucas, S.G., Cassinis, G. and Schneider
2005b, Vertebrate biostratigraphy and biochronology of the Pennsylva J.W., eds., Non-marine Permian biostratigraphy and biochronology:
nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico: Geological Society, London, Special Publication 265, p. 201-215
New Mexico Museum of Natural History and Science, Bulletin 31, p. Williston, S.W., 1911, A new family of reptiles from the Permian of New
128-139. Mexico: American Journal of Science, v. 31, p. 378-398
Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S, Henrici, A.C., Williston, S.W. and Case, E.C., 1912, The Permo-Carboniferous of north
Heckert, A.B., Zeigler, K.E. and Rinehart, L.F., 2005c, Early Permian ern New Mexico: Journal of Geology, v. 20, p. 1-12.
vertebrate assemblage and its biostratigraphic significance, Arroyo del Woodward, L. A., Anderson, O. J. and Lucas, S. G., 1999, Late Paleozoic
Agua, Rio Arriba County, New Mexico: New Mexico Geological Society, right-slip faults in the ancestral Rocky Mountains: New Mexico Geo
Guidebook 56, p. 288-296. logical Society, Guidebook 50, p. 149-153.
6
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

7
ONE HUNDRED AND FIFTYYEARS OF GEOLOGICALAND PALEONTOLOGICAL RESEARCH
IN CAÑON DELCOBRE, NORTHERN NEWMEXICO

SPENCER G. LUCAS, SUSANK. HARRIS AND JUSTINA. SPIELMANN


New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, NM 87104-1375

Abstract—Scientific study of the geology and paleontology in Cañon del Cobre, northern New Mexico, began
with J.S. Newberry in 1859, who discovered Triassic plant fossils in the copper mines in the canyon. David
Baldwin first discovered Carboniferous-Permian tetrapod fossils in Cañon del Cobre in the late 1870s, and this led
to subsequent searches for Carboniferous-Permian tetrapod fossils. S.W. Williston and E.C. Case, in 1911, first
recognized that the oldest tetrapod fossils from Cañon del Cobre are of Carboniferous age, and subsequent research
has supported this conclusion. However, most of the bedrock strata in Cañon del Cobre were assigned a Triassic
age until 1961, when C. Smith, A. Budding and C.W. Pitrat assigned them to the Permian Cutler Formation.
Collecting and research in Cañon del Cobre has now identified invertebrate and vertebrate traces, palynomorphs,
plants and tetrapod bones and teeth of Carboniferous age, traces and tetrapods of Early Permian age and Late
Triassic tetrapod and plant fossils. Cañon del Cobre provides rare and important documentation of the Carbonif
erous-Permian transition, especially of tetrapods, in a nonmarine red-bed succession.

“From Abiquiu our entrance into the Red Beds was made in
the famous cañon known as El Cobre…” – Williston and
Case (1912, p. 4-5).
INTRODUCTION
Cañon del Cobre is a large box canyon located northwest of Abiquiu
in Rio Arriba County, New Mexico (Fig. 1). Also known by its English
name El Cobre Canyon, “copper canyon” (cobre is Spanish for copper)
is so named because of the copper mines located high on its eastern walls.
The Spanish first developed these mines, perhaps as early as the 1600s.
Newberry (1876), who briefly visited the canyon in 1859, provided the
first geological and paleontological information about Cañon del Cobre.
In the 150 years since Newberry’s visit, a variety of geologists and
paleontologists have studied Cañon del Cobre, and today it is a famous
locality for Carboniferous, Permian and Triassic fossils. Here, we review
the history of geological and paleontological research in Cañon del Cobre.
NEWBERRY (1859)
In 1859, John Strong Newberry (1822-1892) participated as the
geologist for the San Juan Exploring Expedition led by Captain J.N.
Macomb (often referred to as the “Macomb Expedition”). Newberry’s
(1876) report on the expedition contains the first descriptions of the
geology and paleontology of the Chama River basin, which includes the
drainage of Cañon del Cobre. FIGURE 1. Index map showing the location of Cañon del Cobre in northern
On 17 July 1859, Newberry (1876, p. 68-69) visited some of the New Mexico (modified from Ash, 1974).
copper mines (Minas de Pedro) developed in the Upper Triassic strata
exposed high on the eastern wall of Cañon del Cobre. Here, Newberry sandstones…here and there they showed large patches of
collected the first Triassic plant fossils with leaves discovered in the white, interstratified with green and brown shales…this
American West (Triassic fossil wood had been discovered in the 1840s). formation contains much saline matter…These cliffs [of
Newberry (1876) described these plants (Fig. 2), and his collection and Cañon del Cobre] are composed at base of the saliferous
additional material was published on by Fontaine and Knowlton (1890). sandstones and interstratified marls, some 250 feet in thick
Newberry (1876) believed that the red beds that floor Cañon del ness; above these blood-red marls and calcareous sand
Cobre and are exposed along most of its walls are of Triassic age: stones, 200 feet thick; the whole crowned by coarse yellow
sandstones having a thickness of about 150 feet…
This [Cañon del Cobre] is excavated in the Triassic series,
(Newberry, 1876, p. 68).
and its sides exhibit bands of brilliant color, red, orange,
blue, white, etc., as vivid as could be drawn from an artist’s Newberry (1876) correctly assigned a Triassic age to the fossil plants he
color-box…Ascending the arroyo, toward its head, we found collected from strata now assigned to the Shinarump Formation of the
the strata rising rapidly toward the north, and cut through Chinle Group (Lucas et al., 2005a). However, Newberry (1876) assigned
so as to expose the saliferous sandstones – the lowest the underlying red beds, Pennsylvanian-Permian strata of the Cutler
member of the Trias – in sections of at least 200 feet. The Group, a Triassic age (Fig. 3), and he also incorrectly assigned a Creta
group here consists of thick-bedded, chocolate ceous age to the sandstones above the plant-bearing horizon (Poleo For
8
mation of the Upper Triassic Chinle Group). Thus, Newberry (1876, p.
69) concluded “we have, therefore, in these plants evidence of the Trias
sic age of all the variegated gypsiferous rocks of northern New Mexico;
for the Lower Cretaceous sandstones immediately overlie the plant-beds
of the Cobre” (italics from Newberry, 1876).
BALDWIN (1877-1880)
David Baldwin was a fossil collector employed by O.C. Marsh
(1831-1899), the famous Yale University paleontologist, from 1876-
1880, to collect vertebrate fossils in northern New Mexico. According to
Williston (1911a, p. 378), Baldwin collected Carboniferous-Permian ver
tebrate fossils for Marsh from Rio Arriba County from November 1877
to December 1880, believing the fossils to be of Triassic age. Romer
(1960) states that Baldwin first discovered Carboniferous fossils in Cañon
del Cobre in 1877.
Marsh, however, did nothing with the bones Baldwin sent him
from Cañon del Cobre. In 1880, Marsh and Baldwin ended their working
relationship – Marsh was generally dissatisfied with the quality of the
fossils Baldwin was sending him, and Baldwin was not being compen
sated adequately or regularly by Marsh.
In 1880, Baldwin thus switched allegiance to Marsh’s scientific
rival, Philadelphia paleontologist Edward Drinker Cope (1840-1897).
However, Baldwin did no further collecting in Cañon del Cobre while
employed by Cope. Thus, Baldwin can be credited with discovering
Carboniferous vertebrate fossils in Cañon del Cobre, though it took
decades for his discovery to be recognized.
WILLISTON, CASE AND HUENE (1911)
Baldwin’s collection of Permian vertebrates from Rio Arriba
County, New Mexico, languished for nearly 30 years in the collections of
the Yale Peabody Museum. Then, about 1911, Samuel Wendell Williston
(1851-1918), a University of Chicago paleontologist, unpacked the
FIGURE 2. Plate 5 of Newberry (1876), illustrations of some of the Upper Baldwin collection, revealing the remarkable skeleton of a reptile from
Triassic plants he collected in the Shinarump Formation of the Chinle Cañon del Cobre that Williston (1911a) named Limnoscelis paludis (Fig.
Group along the eastern rim of Cañon del Cobre. 1, 1a and 2, Zamites 4). Williston (1911a, b, 1912) regarded Limnoscelis as a primitive reptile
occidentalis; 3, Otozamites macombi; and 4, 5, Pagiophyllum newberryi. (cotylosaur), and the exceptional specimen collected by Baldwin in Cañon
See Ash (1974) for further information. del Cobre has played a prominent role in subsequent discussions of
reptilian (amniote) origins (e.g., Romer, 1946; Carroll, 1969a, b, 1982;
Panchen, 1972; Lombard and Bolt, 1979; Heaton, 1980; Kemp, 1980;
Reisz and Heaton, 1980; Fracasso, 1987).
Indeed, Williston was so impressed by the Baldwin collection
that he resolved to visit Cañon del Cobre (and other nearby localities in
Rio Arriba County) to collect more fossils. Thus, in the summer of 1911,
Williston was joined by Ermine Cowles Case (1871-1953), a paleontol
ogy professor at the University of Michigan, Friedrich Freiherr von
Huene (1875-1969), a German paleontologist from Tübingen, and Paul
C. Miller, the fossil preparator and assistant curator working for Williston,
in an expedition to the red beds of Rio Arriba County (Williston and
Case, 1912; Case et al., 1913; Shor, 1971; Lucas, 2005).
No exact chronology of the expedition is available to us, but the
first stop was Cañon del Cobre, described by Williston (in Short, 1971,
p. 221) as “the most dreary and forsaken region I think I was ever in.”
Both a lack of water in the canyon and flash flooding of the canyon floor
proved a challenge to the expedition. Fieldwork in the canyon (Figs. 5-6)
proved relatively disappointing, with few and fragmentary vertebrate
fossils recovered. In Cañon del Cobre, the Williston team collected verte
brate fossils from the “lowermost beds” at “various places in the basin”

FIGURE 3. Evolution of stratigraphic nomenclature and age assignments


applied to Pennsylvanian, Permian and Triassic strata exposed in Cañon del FIGURE 4. Restoration of the skeleton of Limnoscelis paludis (original
Cobre. skeleton about 2 m long). From Williston (1912, fig. 32).
9

FIGURE 5. Selected photographs of the 1911 Williston expedition to northern New Mexico (from a photo album of E.C. Case donated to NMMNH by
Jiri Zidek). A-C, Wagon (A), team (B) and camp kitchen (C). D-E, S.W. Williston. F, Paul Miller.
10

FIGURE 6. Selected photographs of the 1911 Williston expedition to northern New Mexico (from a photo album of E.C. Case donated to NMMNH by
Jiri Zidek). A-C, First camp in the Cañon (at the Puertocito). D, Paul Miller (left) and E.C. Case (right).

(Williston and Case, 1913, p. 2), and also explored strata in the north discouraged extensive collecting efforts in the canyon for decades.
wall “somewhat below the middle of the Permo-Carboniferous strata of Harvard University paleontologistAlfred Sherwood Romer (1894
El Cobre canyon” (Case et al., 1913, p. 17), where a nearly complete 1973) noted that he collected Pennsylvanian-Permian vertebrate fossils
skeleton of the diadectid Diasparactus zenos was found (Case et al., in Cañon del Cobre “in 1926, 1931, 1946 and 1952, and other Harvard
1913). Williston and Case (1912), however, did provide a more detailed workers went there on three other occasions” (Romer, 1960, p. 48).
description of the geology and stratigraphy of Cañon del Cobre than had Although Romer reported fossils in the canyon to be rare, he was able to
Newberry. As they noted, “El Cobre Canyon or basin is formed by the describe the varanopseid Aerosaurus greenleeorum (Romer, 1937) from
erosion of an unsymmetrical dome-shaped anticline more or less faulted a “double handful of bones and bone fragments, not articulated but pre
on the northeastern and southeastern sides, the brim formed everywhere sumably from a single individual” (Romer and Price, 1940, p. 277).
by the massive sandstones of basal Upper Triassic age, the strata sloping Descriptions of the edaphosaurid Nitosaurus jacksonorum (Romer, 1937)
in all directions, but chiefly east and west” (Williston and Case, 1912, p. and the ophiacodontid Baldwinonus trux (Romer and Price, 1940) were
5-6). based on specimens collected by Baldwin. Romer (1950) also published
Williston and Case also presented a stratigraphic section of the a photograph taken by Williston in 1911 with a horizontal line indicating
western wall of Cañon del Cobre, assigning the strata to the Permo the vertebrate-producing horizon on the north wall of Cañon del Cobre,
Carboniferous, Lower Trias? and Upper Trias (Fig.2). Their work thus which provides the only precise information we have of where the
represented a major advance in our understanding of the paleontology Williston expedition collected its fossils (Fig. 8). Soon after, Langston
and age of strata exposed in Cañon del Cobre. Also, based on the work, (1953) provided a useful review of the Paleozoic vertebrate fauna of
Williston (1911b, fig. 1) published what can be considered the first Cañon del Cobre, but he did not collect there.
reconstruction of a Carboniferous landscape in Cañon del Cobre (Fig. 7).
SMITH, BUDDING AND PITRAT (1950S)
ROMER (1926-1952)
During the 1950s, the geological studies conducted by field classes
Although several descriptions of fossil vertebrates from Cañon from the New Mexico Institute of Mining and Mineral Technology
del Cobre were based on specimens collected by Baldwin and the (Socorro) led to the published geological mapping of Smith et al. (1961).
Williston expedition, their fieldwork did not inspire many followers. Thus, they mapped the geology of the U.S. Geological Survey Canjilon
The reputation of Cañon del Cobre as a remote, nearly inaccessible locale SE 7.5-minute topographic map, within which Cañon del Cobre is lo
whose main passageway, Arroyo del Cobre, is subject to flash floods cated. Most significantly, Smith et al. (1961; also see Budding et al.,
(one of which nearly drowned members of the Williston party of 1911), 1960) applied the lithostratigraphic name Cutler Formation to the Cañon
11

FIGURE 7. “A permocarboniferous [sic] landscape, with two figures of Eryops upon the land, each of about seven feet in length, and Limnoscelis, a
cotylosaur of the same length, in the water” (Williston, 1911b, p. 58, fig. 1).

del Cobre section, assigning most of the bedrock exposed in the canyon lected several incomplete vertebrae referable to Limnoscelis. Vaughn’s
to this unit, which they regarded as Early Permian in age (Fig. 3). They (1963) principal conclusion was to reaffirm a Carboniferous age for the
assigned overlying strata, which rim the canyon walls and contain the stratigraphically lowest vertebrate fossils in Cañon del Cobre.
copper mines and fossil plants first collected by Newberry, to the “lower During a brief trip to the canyon in 1965, A. Lewis and S. Olsen of
sandstone member” of the Chinle Formation (Fig. 3). They did not apply Harvard University collected a partial skeleton of a pelycosaur that was
the formal stratigraphic subdivisions of the Chinle introduced by Wood later described by Eberth and Brinkman (1983) as a new genus and
and Northrop (1946) in the Rio Puerco River Valley some 50 km to the species of varanopsid pelycosaur, Ruthiromia elcobriensis (also see
southwest. Near the northern terminus of Cañon del Cobre, Smith et al. Spielmann and Lucas, this volume). Although the precise locality is not
(1961, p. 5, pl. 6) located “thin-bedded carbonaceous, gypsiferous and known, field notes made by Lewis and Olsen (Eberth and Brinkman,
micaceous siltstone…exposed along a small fault” with “a well pre 1983) indicate that the material was collected from the western wall of
served flora, consisting predominantly of Alethopteris serlii…” They the canyon, strongly suggesting a stratigraphically high position in the
judged these strata to be of Pennsylvanian age and mapped them as a unit Arroyo del Agua Formation of Lucas and Krainer (2005), and thus a
separate from the Cutler Formation (Fig. 3). younger age for this specimen than that of fossils collected from the
classic sites on the canyon floor.
VAUGHN (1963) AND LEWIS AND OLSEN (1965)
FRACASSO (1978)
University of California at Los Angeles paleontologist Peter P.
Vaughn (1963) reported disarticulated, fragmentary vertebrate fossils Fracasso (1980) reported vertebrate fossils collected from five
that he collected from two sites on the floor of Cañon del Cobre. How sites on the canyon floor of the Cañon del Cobre collected during the
ever, he deemed the specimens sufficiently diagnostic to recognize three summer of 1978. The most significant find of this trip, made by one of us
additional taxa in the Cañon del Cobre vertebrate assemblage: the (SGL), as a member of the Fracasso reconnaissance team, was a partial
dissorophid Platyhystrix rugosus, based on a large number of neural jaw and cranial fragments of the primitive diadectomorph Desmatodon
spine fragments with surface ornamentation consistent with that of P. aff. D. hollandi, indicative of a Late Pennsylvanian age for the strata
rugosus from the Arroyo del Agua locality; the temnospondyl exposed on the canyon floor. It was described, but not illustrated, by
Chenoprosopus cf. C. milleri, recognized from a cranial fragment that Fracasso (1980) and Berman and Sumida (1995; but see Lucas et al.,
consists of the dermal roofing and palatal elements of the left narial 2005c, fig. 5f, i-j for an illustration). Fracasso (1980) also reported a
region; and the edaphosaurid Edaphosaurus cf. E. novomexicanus, based megafossil plant locality from the canyon floor that the late S.H. Mamay
on several incomplete neural spines bearing tubercles on both sides, with of the Smithsonian National Museum of Natural History deemed to be
a diameter of 8 mm, as in E. novomexicanus. In addition, Vaughn col of Late Pennsylvanian age. Fracasso (1987) subsequently presented an
12

FIGURE 8. Photograph by S.W. Williston (1911) looking approximately


northward of part of the western wall of Cañon del Cobre (view is primarily
of N½ secs. 17-18 and S1/4 secs. 7-8, T24N, R06E). The horizontal line
marks the vertebrate-producing horizon in the El Cobre Canyon Formation;
the “+” indicates the base of the Upper Triassic. From Romer (1950).

analysis of the fluvial sedimentology of the lower part of the Cutler


Group exposed in Cañon del Cobre.
BERMAN (1980s)
Collecting efforts of the last quarter century in Cañon del Cobre FIGURE 9. NMMNH fieldwork in Cañon del Cobre in 2005. A, Central mess
began with work by joint field parties from the Carnegie Museum of tent of base camp, manned by Chester Lawrence. B, Larry Rinehart
Natural History and the University of Toronto in the early 1980s led by excavating Eryops skull. C, Joseph Hannibal at giant Diplichnites trackway.
David S Berman from the former institution. As a result of this field D, Joshua Smith (above) and Matt Celeskey at Sphenacodon quarry.
work, three additional taxa were recognized in the Cañon del Cobre
vertebrate-fossil assemblage, and include the new genus and species of paleoflora, as well as the fossil-vertebrate fauna from “megasequence
trematopid temnospondyl, Anconastes vesperus, described by Berman one” of Eberth (1987) to be Pennsylvanian and suggested an age as old as
et al. (1987b) on the basis of two partial skeletons; a small, as yet Desmoinesian. Although the paleobotany of Cañon del Cobre remains an
undescribed embolomere known from a single multispecimen occurrence; active research pursuit of the NMMNH and its associates (see DiMichele
and the lungfish Gnathorhiza preserved in an aestivation burrow cast. In and Chaney, 2005), recent paleontological investigations (since 2000)
addition to these specimens, collected at three different intervals in the El have included the discovery and collection of numerous fossil verte
Cobre Canyon Formation section, a rich concentration of bones high in brates at several localities.
the section in the overlying Arroyo del Agua Formation on the eastern
A combined field effort of the New Mexico Museum of Natural
wall of the canyon was also discovered. Berman (1993) and Eberth and History, National Museum of Natural History (Smithsonian Institution)
Berman (1993) noted that at least three taxa, including Platyhystrix, and Carnegie Museum of Natural History in Cañon del Cobre began in
Diadectes and Sphenacodon ferox, are well represented at this site, which 2000 (Fig. 9). Particularly significant fieldwork took place in 2000-2007,
yields disarticulated, though well preserved bones, on which fine details
which was devoted to prospecting and collecting any and all fossil locali
are preserved. ties in the canyon and studying the stratigraphy and sedimentology of
In conjunction with paleontological studies, Eberth (1987; Eberth the Cutler Group strata in the canyon. Overlapping and in collaboration
and Miall, 1991; Eberth and Berman, 1983, 1993) studied the fluvial with the NMMNH, the New Mexico Bureau of Geology mapped the
sedimentology of the Cutler Group strata exposed in Cañon del Cobre. geology of the Canjilon SE quadrangle (Kempter et al., 2007). Also, K.
This work presented the first analysis of the depositional context of the Krainer of the University of Innsbruck (Austria) collaborated on the
fossils found in the Permo-Carboniferous Cutler Group strata in Cañon stratigraphy and sedimentology of the Cutler Group strata exposed in
del Cobre. Cañon del Cobre, and J. Hannibal of the Cleveland Museum of Natural
History and J. Schneider of Freiberg University (Germany) also worked
NEWMEXICO MUSEUM OF NATURALHISTORY (1991-2007)
on some aspects of Cutler Group paleontology. Initial results of the
Based on the field studies undertaken during the late 1980s, Lucas NMMNH project were published in 2005 (DiMichele and Chaney, 2005;
and Hunt (1992; also see Lucas et al., 2005b) presented a revised Triassic Hannibal et al., 2005; Lucas and Krainer, 2005; Lucas et al., 2005b, c, d).
stratigraphy in the Chama basin. Currently, based on this work, the Particularly significant results of the NMMNH collaborative re
Upper Triassic section exposed along the rim of the Cañon del Cobre is search in Cañon del Cobre include: (1) lithostratigraphic refinement by
assigned to the (ascending order) Shinarump, Salitral and Poleo forma recognition of two formations of the Cutler Group in the Chama basin,
tions of the Chinle Group (Fig. 3). the older El Cobre Canyon Formation overlain by the Arroyo del Agua
New Mexico Museum of Natural History field efforts in the Formation and studies of their petrography and sedimentology (Lucas
upper Paleozoic strata of Cañon del Cobre began in the summer of 1991 and Krainer, 2005; Kempter et al., 2007); (2) stratigraphic organization
with the collection of plant megafossils from two localities just below of the fossil vertebrate localities in the Cutler Group strata in Cañon del
the lowest tetrapod-bearing beds. Hunt and Lucas (1992) re-examined Cobre, allowing for the recognition of three temporally successive as
the fossil plant locality reported by Fracasso (1980) and determined the semblages that provided part of the basis for establishment of the Cobrean,
13
Coyotean and Seymouran land-vertebrate faunachrons (Lucas, 2006; ACKNOWLEDGMENTS
Lucas et al., 2005c, d); (3) discovery and collection of a substantial We are grateful to numerous NMMNH volunteers as well as
pteridosperm-dominated paleoflora (and associated palynomorphs) from colleagues from various institutions who collaborated with us in our
the El Cobre Canyon Formation (DiMichele and Chaney, 2005); and (4) work in Cañon del Cobre. Special thanks go to David Berman and Amy
study of the ichnology of the El Cobre Canyon Formation, including Henrici for sharing their knowledge of the paleontology of the canyon.
discovery of the walking traces of the giant Carboniferous arthropod David Berman, William DiMichele, Amy Henrici and Karl Krainer pro
Arthropleura (Lucas et al., 2005c). vided helpful reviews of the manuscript.

REFERENCES

Ash, S.R., 1974, Upper Triassic plants of Cañon del Cobre, New Mexico: vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale
New Mexico Geological Society, Guidebook 25, p. 179-184. ontology, v. 54, p. 1237-1244.
Berman, D.S, 1993, Lower Permian vertebrate localities of New Mexico Fracasso, M.A., 1983, Cranial osteology, functional morphology, system
and their assemblages: New Mexico Museum of Natural History and atics and paleoenvironment of Limnoscelis paludis Williston [Ph.D.
Science, Bulletin 2, p. 11-21. dissertation]: Yale University, New Haven, 624 p.
Berman, D.S, Reisz, R.R. and Eberth, D.A., 1987a, Seymouria sanjuanensis Fracasso, M.A., 1987a, Braincase of Limnoscelis paludis Williston: Postilla,
(Amphibia, Batrachosauria) from the Lower Permian Cutler Formation v. 201, p. 1-22.
of north-central New Mexico and the occurrence of sexual dimorphism Fracasso, M.A., 1987b, Fluvial deposition and paleoenvironment of the
in that genus questioned: Canadian Journal of Earth Science, v. 24, p. Cutler Formation red beds, El Cobre Canyon, New Mexico: New Mexico
1769-1784. Geology, v. 9, p. 14-18.
Berman, D.S, Reisz, R.R. and Eberth, D.A., 1987b, A new genus and species Hannibal, J.T., Lucas, S.G., Lerner, A.J and Chaney, D.S., 2005, An euryp
of trematopid amphibian from the Late Pennsylvanian of north-central terid (Adelophthalmus sp.) from a plant-rich lacustrine facies of Upper
New Mexico: Journal of Vertebrate Paleontology, v. 7, p. 252-269. Pennsylvanian strata in El Cobre Canyon, New Mexico: New Mexico
Berman, D.S and Sumida, S., 1995, New cranial material of the rare diadectid Museum of Natural History and Science, Bulletin 31, p. 34-38.
Desmatodon hesperis (Diadectomorpha) from the Late Pennsylvanian Heaton, M.J., 1980, The Cotylosauria: A reconsideration of a group of
of central Colorado: Annals of Carnegie Museum, v. 64, p. 315-336. archaic tetrapods; in Panchen, A.L., ed., The terrestrial environment
Budding, A.J., Pitrat, C.W. and Smith, C.I., 1960, Geology of the southern and origin of land vertebrates: Academic Press, New York, p. 497-551.
part of the Chama basin: New Mexico Geological Society, Guidebook 11, Hunt, A.P. and Lucas, S.G., 1992, The paleoflora of the lower Cutler Forma
p. 78-92. tion (Pennsylvanian, Desmoinesian?) in El Cobre Canyon, New Mexico,
Carroll, R.L., 1969a, Problems of the origin of the reptiles: Biological and its biochronological significance: New Mexico Geological Society,
Reviews, v. 44, p. 393-432. Guidebook 43, p. 145-150.
Carroll, R.L., 1969b, Origin of reptiles; in Gans, C., Bellairs, A. d’A. and Kemp, T.S., 1980, Origin of the mammal-like reptiles: Nature, v. 283, p.
Parsons, T.S., eds., Biology of the Reptilia, vol. 1 Morphology A: Aca 378-380.
demic Press, New York, p. 7-44. Kempter, K., Zeigler, K., Koning, D. and Lucas, S., 2007, Preliminary
Carroll, R.L., 1982, Early evolution of reptiles: Annual Review of Ecology geologic map of the Canjilon SE quadrangle, Rio Arriba County, New
and Systematics, v. 13, p. 87-109. Mexico: New Mexico Bureau of Geology and Mineral Resources, Open
Case, E.C. and Williston, S.W., 1913, Description of a nearly complete file Geological Map 150, 1:24,000.
skeleton of Diasparactus zenos Case: Carnegie Institution of Washing Langston, W., Jr., 1953, Permian amphibians from New Mexico: Univer
ton, Publication no. 181, p. 17-35. sity of California, Publications in Geological Sciences, v. 29, p. 349
Case, E.C., Williston, S.W. and Mehl, M.G., 1913, Permo-Carboniferous 416.
vertebrates from New Mexico Carnegie Institution of Washington, Pub Lombard, R.E. and Bolt, J.R., 1979, Evolution of the tetrapod ear: An
lication no. 181, p. 1-81. analysis and reinterpretation: Biological Journal of the Linnean Society,
DiMichele, W.A. and Chaney, D.S., 2005, Pennsylvanian-Permian fossil v. 11, p. 19-76.
floras from the Cutler Group, Cañon del Cobre and Arroyo del Agua Lucas, S.G., 2005, J.S. Newberry and the geology of the Chama basin: New
areas, in northern New Mexico: New Mexico Museum of Natural His- Mexico Geological Society, Guidebook 56, p. 53.
tory and Science, Bulletin 31, p. 26-33. Lucas, S.G., 2006, Global Permian tetrapod biostratigraphy and
Eberth, D.A., 1987, Stratigraphy, sedimentology, and paleoecology of Cut biochronology; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds.,
ler Formation redbeds (Permo-Pennsylvanian) in north-central New Non-marine Permian biostratigraphy and biochronology: Geological
Mexico [Ph.D. dissertation]: University of Toronto, Toronto, 264 p. Society, London, Special Publication 265, p. 65-93.
Eberth, D.A. and Berman, D.S, 1983, Sedimentology and paleontology of Lucas, S.G. and Hunt, A.P., 1992, Triassic stratigraphy and paleontology,
Lower Permian fluvial redbeds of north-central New Mexico – prelimi Chama basin and adjacent areas, north-central New Mexico: New Mexico
nary report: New Mexico Geology, v. 5, p. 21-25. Geological Society, Guidebook 43, p. 151-167.
Eberth, D.A. and Berman, D.S, 1993, Stratigraphy, sedimentology and ver Lucas, S.G. and Krainer, K., 2005, Stratigraphy and correlation of the
tebrate paleontology of the Cutler Formation redbeds (Pennsylvanian Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New
Permian) of north-central New Mexico: New Mexico Museum of Natu Mexico Geological Society, Guidebook 56, p. 145-159.
ral History and Science, Bulletin 2, p. 33-48. Lucas, S.G., Zeigler, K.E., Heckert, A.B. and Hunt, A.P., 2005a, Review of
Eberth, D.A. and Brinkman, D., 1983, Ruthiromia elcobriensis, a new Upper Triassic stratigraphy and biostratigraphy in the Chama basin,
pelycosaur from El Cobre Canyon, New Mexico: Breviora, v. 474, p. 1 northern New Mexico: New Mexico Geological Society, Guidebook 56,
26. p. 170-181.
Eberth, D.A. and Miall, A.D., 1991, Stratigraphy, sedimentology and evolu Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S and Henrici, A.C.,
tion of a vertebrate-bearing, braided to anastomosed fluvial system, 2005c, Vertebrate biostratigraphy and biochronology of the Pennsylva
Cutler Formation (Permian-Pennsylvanian), north-central New Mexico: nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico:
Sedimentary Geology, v. 72, p. 225-252. New Mexico Museum of Natural History and Science, Bulletin 31, p.
Fracasso, M.A., 1980, Age of the Permo-Carboniferous Cutler Formation 128-139.
14
Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S, Henrici, A.C., northwestern New Mexico: American Museum of Natural History, New
Heckert, A.B., Zeigler, K.E. and Rinehart, L.F., 2005d, Early Permian York, p. 48-55.
vertebrate assemblage and its biostratigraphic significance, Arroyo del Romer, A.S., 1960, The vertebrate fauna of the New Mexico Permian: New
Agua, Rio Arriba County, New Mexico: New Mexico Geological Society, Mexico Geological Society, Guidebook 11, p. 48-54.
Guidebook 56, p. 288-296. Romer, A.S. and Price, L.I., 1940, Review of the Pelycosauria: Geological
Lucas, S.G., Lerner, A.J, Hannibal, J.T., Hunt, A.P. and Schneider, J.W., Society of America, Special Paper No. 28, 538 p.
2005b, Trackway of a giant Arthropleura from the Upper Pennsylva Shor, E.N., 1971, Fossils and flies: The life of a complete scientist, Samuel
nian of El Cobre Canyon, New Mexico: New Mexico Geological Society, Wendell Williston (1851-1918): University of Oklahoma Press, Norman,
Guidebook 56, p. 279-282. 285 p.
Newberry, J.S., 1876, Geological report; in Macomb, J.N., ed., Report of the Smith, C. T., Budding, A. J. and Pitrat, C. W., 1961, Geology of the south
exploring expedition from Santa Fe…in 1859: U.S. Army Engineer eastern part of the Chama basin: New Mexico Bureau of Mines and
Department, Washington, D.C., p. 9-118. Mineral Resources, Bulletin 75, 57 p
Panchen, A.L., 1972, The interrelationships of the earliest tetrapods; in Vaughn, P.P., 1963, The age and locality of the late Paleozoic vertebrates
Joysey, K.A. and Kemp, T.S., eds., Studies in vertebrate evolution: Win from El Cobre Canyon, Rio Arriba County, New Mexico: Journal of
chester Press, New York, p. 65-87. Paleontology, v. 37, p. 283-296.
Reisz, R.R. and Heaton, M.J., 1980, Origin of mammal-like reptiles: A reply Williston, S.W., 1911a, A new family of reptiles from the Permian of New
to T.S. Kemp: Nature, v. 288, p. 193. Mexico: American Journal of Science, v. 31, p. 378-398
Romer, A.S., 1937, New genera and species of pelycosaurian reptiles: New Williston, S.W., 1911b, American Permian vertebrates: University of Chi
England Zoological Club, Proceedings, v. 16, p. 89-96. cago Press, Chicago, 145 p.
Romer, A.S., 1946, The primitive reptile Limnoscelis restudied: American Williston, S.W., 1912, Restoration of Limnoscelis, a cotylosaur reptile
Journal of Science, v. 244, p. 149-188. from New Mexico: American Journal of Science, v. 34, p. 457-468.
Romer, A.S., 1950, The upper Paleozoic Abo Formation and its vertebrate Williston, S.W., 1914, The osteology of some American Permian verte
fauna; in Colbert, E.H. and Northrop, S.A., eds., Guidebook for the brates: Journal of Geology, v. 22, p. 408-410.
fourth field conference of the Society of Vertebrate Paleontology in Williston, S.W. and Case, E.C., 1912, The Permo-Carboniferous of north
ern New Mexico: Journal of Geology, v. 20, p. 1-12.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

15
SUMMARY OF GEOLOGY OF CAÑON DELCOBRE,
RIOARRIBA COUNTY, NEWMEXICO

SPENCERG. LUCAS1, JUSTINA. SPIELMANN1 AND KARLKRAINER2


1New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, NM 87104-1375;
2 Institute of Geology and Paleontology, Innsbruck University, Innrain 52, Innsbruck, A-6020 AUSTRIA

Abstract—Cañon del Cobre is a large box canyon in southeastern Rio Arriba County, New Mexico, along the
southeastern border of the Colorado Plateau, that exposes strata of Pennsylvanian, Permian, Triassic and Cenozoic
age. The oldest rocks here are siliciclastic red beds of the El Cobre Canyon Formation of the Cutler Group, strata
of Late Pennsylvanian-Early Permian (Virgilian-Wolfcampian) age. The exposed El Cobre Canyon Formation is
about 230 m thick (total thickness ~ 640 m) and is mostly siltstone and sandstone deposited in an ephemeral
braided stream environment. The overlying Lower Permian Arroyo del Agua Formation of the Cutler Group is as
much as 160 m thick, and also consists mostly of siltstone and sandstone, but was deposited on laterally extensive
floodplains that developed between relatively stable channels. An extensive unconformity separates Cutler Group
strata from overlying Late Triassic strata of the Chinle Group around the rim of Cañon del Cobre. Chinle Group
strata are of fluvial origin and are assigned (in ascending order) to the Shinarump (up to 25 m of quartzose
sandstone, conglomeratic sandstone and siliceous, extraformational conglomerate), Salitral (up to 7 m of variegated
mudstone) and Poleo (up to 30 m of sandstone, conglomeratic sandstone and conglomerate) formations. The
Shinarump Formation hosts stratiform copper deposits along the eastern wall of the Cañon del Cobre that the
Spanish began mining during the 1600s.
Cenozoic strata locally exposed around the rim of Cañon del Cobre include the Eocene El Rito Formation (up
to 50 m of red-bed siltstones, sandstones and conglomerates) and Oligocene Ritito Conglomerate (up to 60 m of
conglomerate and arkosic sandstone). Quaternary deposits – quartzite erratic boulders, alluvial terrace deposits,
colluvium and alluvium – crop out in and around Cañon del Cobre. Cañon del Cobre is traversed and bounded by
several north-northeast-trending faults, and its eastern wall is along a faulted, north-northeast-trending anticline.
These structures were accomodated by Neogene extensional tectonism of the nearby Rio Grande rift.

INTRODUCTION
Cañon del Cobre (also known as El Cobre Canyon or as Copper
Canyon) is a large box canyon located in southeastern Rio Arriba County,
New Mexico (Fig. 1). This volume focuses on the geology and paleontol
ogy of the Carboniferous-Permian transition as recorded in Cutler Group
red beds exposed in Cañon del Cobre. Here, we present a brief summary
of the outcrop geology in and immediately around the canyon (Fig.2). In
this volume, Lucas et al. (2010) review the history of geological and
paleontological studies in Cañon del Cobre. Discussion of the Pennsyl
vanian, Permian and Triassic rocks is based primarily on our own work,
whereas observations on the Cenozoic geology and structure are summa
rized primarily from Kempter et al. (2007).
PHYSIOGRAPHY
Located in the southern Rocky Mountains along the southeastern
border of the Colorado Plateau, Cañon del Cobre extends from a narrow
set of water gaps at El Puertocito at its southern end to its northeastern
rim over a distance of about 10 km (Fig. 2). Throughout most of its
course, the canyon is 2-3 km wide, and its floor elevation ranges from
about 6300 to 7400 feet (1920 to 2255 m) above sea level. Maximum
depth of the canyon is about 305 m. The canyon has an almost north
south orientation for the southern 5 km, then bends to the northeast over
its northern 5 km. FIGURE 1. Index map showing location of Cañon del Cobre in Rio Arriba
The floor of the canyon is occupied by the Arroyo del Cobre, an County, northern New Mexico (after Ash, 1974).
intermittent tributary of the Rio Chama. A single, four-wheel drive dirt
road enters Cañon del Cobre through El Puertocito and extends through unnamed and unmarked mines (uranium or copper prospects) are lo
the southern half of the canyon. Old Spanish copper mines (first de cated in the floor of the canyon. An active artesian well is also present in
scribed by Newberry, 1876) – Las Minas Jimmie and Las Minas de the canyon (NW¼ SW¼ sec. 19, T24N, R06E), as are various seasonal
Pedro – are located high on the canyon’s eastern rim. Various other springs.
16

FIGURE 2. Geologic map of Cañon del Cobre (modified from Kempter et al., 2007). Square grid is km-spaced UTM (Universal Transverse Mercator) grid.
17
STRATIGRAPHY
We identify three groups of sedimentary deposits in Cañon del
Cobre: (1) Carboniferous-Permian Cutler Group; (2) Late Triassic Chinle
Group; and (3) Cenozoic strata, colluvium and alluvium (Fig. 3). Des
cription of the Cutler Group is based in part on Lucas and Krainer
(2005a, b), Chinle Group strata are described largely based on Lucas and
Hunt (1992), and Kempter et al. (2007) is our primary source for data on
the Cenozoic rocks. Jurassic strata (Entrada Formation) are also exposed
just SE of Cañon del Cobre (Fig. 2), but are not discussed here.

Carboniferous and Permian Systems


Cutler Group
The floor and canyon walls of Cañon del Cobre expose extensive
bedrock badlands, slopes and escarpments developed in an ~400 m thick
section of Cutler Group siliciclastic red beds (Smith et al., 1961; Lucas
and Krainer, 2005a, b; Kempter et al., 2007). These strata are of fluvial
origin (Eberth and Berman, 1983, 1993; Fracasso, 1987; Eberth and
Miall, 1991) and have been divided into a lower, El Cobre Canyon For
mation and an upper, Arroyo del Agua Formation (Lucas and Krainer,
2005a, b) (Figs. 3-6).

El Cobre Canyon Formation


In Cañon del Cobre, a total thickness of about 230 m of El Cobre
Canyon Formation is exposed, and it is mostly siltstone and sandstone
(Figs. 3-6; Appendix 1). Minor rock types are conglomerate/conglomer
atic sandstone, sandy shale and calcrete. These rocks are characteristi
cally “brown” (pale reddish brown) and are readily distinguished from
the overlying Arroyo del Agua Formation based on color and lithology.
Siltstones of the El Cobre Canyon Formation contain numerous rhizoliths,
and sandstones are typically coarse grained, arkosic, micaceous, trough
crossbedded and form laterally extensive, multistoried bodies that erode
to cliffs and ledges. Local shale beds represent lacustrine deposits. Con
glomerates characteristically have extraformational clasts of quartzite,
granite and gneiss (Kempter et al., 2007) or calcrete rip-ups (Appendix
1). The El Cobre Canyon Formation is interpreted to represent deposits
of a shallow, ephemeral braided stream environment according to Eberth
and Miall (1991).
The El Cobre Canyon Formation is best exposed in the floor of
Cañon del Cobre and in the Rio Puerco Valley near Arroyo del Agua,
about 40 km to the southwest. Its lower contact is not exposed, but
subsurface data published by Smith et al. (1961) indicate that ~410 m of
El Cobre Canyon Formation strata are present below the outcrop base in
Cañon del Cobre, and that the El Cobre Canyon Formation rests on
Proterozoic basement (Smith et al., 1961; Eberth and Miall, 1991). The
upper contact of the El Cobre Canyon Formation appears to be con
formable at the base of the first “orange” siltstone slope of the Arroyo
del Agua Formation. The El Cobre Canyon Formation approximately
corresponds to megasequence 1 of Eberth and Miall (1991).

Arroyo del Agua Formation


Conformably overlying the El Cobre Canyon, the Arroyo del
Agua Formation in Cañon del Cobre is as much as 120 m thick (Figs. 3
6) and mostly consists of sandy siltstone and sandstone. Minor rock
types are calcrete and conglomerate/conglomeratic sandstone. Siltstones
of the Arroyo del Agua Formation are characteristically “orange” (mod
erate reddish brown), locally contain rhizoliths and may contain abun
dant calcrete nodules. They underlie long slopes between thin sheets of
sandstone that are coarse grained, arkosic, trough crossbedded and, lo
cally, multistoried. Conglomerates are not as common in the Arroyo del
Agua Formation as in the underlying El Cobre Canyon Formation, and
most are intraformational (composed of calcrete clasts). However, in the
upper part of the Arroyo del Agua Formation extraformational conglom
erates (primarily composed of quartzite clasts) are present (this is FIGURE 3. Summary of stratigraphic units exposed in Cañon del Cobre.
18

FIGURE 4. View of part of western wall of Cañon del Cobre (principally secs. 24-25, T24N, R5E) with stratigraphic units labeled.

megasequence 3 of Eberth and Miall, 1991). merly Agua Zarca Member of the Chinle Formation: Wood and Northrop,
The Arroyo del Agua Formation is well exposed along the walls of 1946). The Shinarump is up to 25 m thick at Cañon del Cobre and
Cañon del Cobre and in the Rio Puerco Valley from Arroyo del Agua east consists of fluvially deposited quartzose and feldspathic litharenite sand
to Abiquiu Reservoir. At all outcrops in the Chama Basin, the Arroyo del stone, conglomeratic sandstone and siliceous (quartz, chert and/or quartz
Agua Formation conformably overlies the El Cobre Canyon Formation. ite), extraformational conglomerate. Colors are mostly greenish gray and
Gradation and interfingering characterizes this contact, so it is appar grayish yellow, and trough crossbeds are the dominant bedform. The
ently conformable. Throughout Cañon del Cobre, the Chinle Group rests Shinarump Formation forms a light-colored bench or cliff at the base of
disconformably on the Arroyo del Agua Formation (Lucas and Hunt, the Chinle Group. Its basal contact is usually a sharp, erosional scour on
1992; Lucas et al., 2003, 2005a; Kempter et al., 2007). (Figs. 4-5) underlying finer grained strata of the Arroyo del Agua Formation of the
The Arroyo del Agua Formation approximately encompasses Cutler Group. Locally, there are as much as 2 m of color-mottled strata
megasequences 2 and 3 of Eberth and Miall (1991). The facies shows between the Shinarump and Arroyo del Agua red beds – the “mottled
some significant differences compared to the El Cobre Canyon Forma strata” of Stewart et al. (1972) or Zuni Mountains Formation of Heckert
tion. The Arroyo del Agua Formation is characterized by locally abun and Lucas (2003) (Fig. 7). Local copper mineralization in the Shinarump
dant, major sandstone ribbons and u-shaped channel fills, separated by Formation was mined at Las Minas Jimmie and Las Minas de Pedro,
thick intervals of siltstone containing abundant calcrete nodules. When beginning during the 1600s (R. Eveleth, written commun., 2009).
compared to El Cobre Canyon Formation deposition, the fluvial style
had changed to laterally extensive floodplains that developed between Salitral Formation
relatively stable channels that included local, anastomosed channels. The The Shinarump Formation at Cañon del Cobre is sharply overlain
climate has been interpreted as more arid than during deposition of the El
by mudstone of the Salitral Formation. As much as 7 m thick, the Salitral
Cobre Canyon Formation (Eberth and Miall 1991). However, the change is color mottled (grayish red, purple, yellowish gray and light gray)
in the character of the fluvial deposits between the El Cobre Canyon
mudstone and sandy mudstone. It forms a notch or slope between the
Formation and the Arroyo del Agua Formation could have been caused
more resistant Shinarump Formation (below) and the Poleo Formation
by a shift in base level as the result of erosion and accumulation. In
(above).
Cañon del Cobre, all but two of the known Paleozoic vertebrate fossil
localities are in the El Cobre Canyon Formation (Lucas et al., 2005b). All Poleo Formation
known Paleozoic fossil plant and invertebrate trace fossil sites are also in
the El Cobre Canyon Formation (Fig. 4). The youngest Late Triassic strata exposed around the periphery
of Cañon del Cobre belong to the Poleo Formation. As much as 30 m
Triassic System thick locally, the Poleo Formation is fluvially-deposited sandstone, con
Chinle Group glomeratic sandstone and conglomerate. Sandstones are mostly yellow
ish gray micaceous litharenites and are typically crossbedded or laminar.
Late Triassic strata of the Chinle Group (Lucas, 1993) are ex Conglomerates have mixed populations of intrabasinal (siltstone and
posed across the Chama River basin and have recently been reviewed by nodular calcrete) and extrabasinal (chert and quartzite) clasts. The Poleo
Lucas and Hunt (1992) and Lucas et al. (2003, 2005a), who provide forms most of the cliffs that rim Cañon del Cobre (Fig. 4). Its base is an
detailed descriptions and discussions. Chinle strata crop out along the erosional unconformity on the Salitral Formation. Younger Chinle Group
upper rimrock (periphery) of the Cañon del Cobre and rest with pro strata (Petrified Forest and Rock Point formations) are exposed south of
found unconformity on the Arroyo del Agua Formation of the Cutler El Puertocito, just outside the southern periphery of Cañon del Cobre
Group (Figs. 4, 5, 7). (Kempter et al., 2007).

Shinarump Formation Cenozoic Units


The base of the Chinle section is the Shinarump Formation (for The oldest Cenozoic strata exposed around the rim of Cañon del
19

FIGURE 5. Measured stratigraphic section of Cutler Group strata in Cañon del Cobre, which is the type section of the El Cobre Canyon Formation of Lucas
and Krainer (2005a). See Appendix 1 for description of section.
20

FIGURE 6. Selected outcrops of the El Cobre Canyon Formation of the Cutler Group in Cañon del Cobre. A, C, Characteristic crossbedded sandstones. B,
Light-colored, thinly laminated siltstone of localized lacustrine deposit. D, Extabasinal matrix- to clast-supported conglomerate of (primarily) quartzite
pebbles. E, Rhizoliths. F, Bioturbation (cf. Planolites) in channel-margin sandstone.
21

FIGURE 8. Structural cross-section from northwest-to-southeast across


Cañon del Cobre. Modified from Kempter et al. (2007).

Cobre belong to the Eocene El Rito Formation. These are red-bed silt
stones, sandstones and conglomerates up to 50 m thick (Smith et al.,
1961; Kempter et al., 2007). They are fluvial deposits of late Laramide
(Eocene) age deposited at the northern end of the Eocene Galisteo-El
Rito basin (Logsdon, 1981; Chapin and Cather, 1981; Lucas, 1984).
The next Cenozoic unit exposed around the rim of the Cañon del
Cobre is the Ritito Conglomerate (formerly the lower member of the
Abiquiu Formation) (Kempter et al., 2007; Maldonado and Kelley, 2009).
These strata are gray to pink conglomerates and arkosic sandstones up to
60 m thick (Kempter et al., 2007).
Quaternary deposits in the Cañon del Cobre include valley-floor
alluvium, alluvial terrace deposits, high-level gravel deposits, colluvium,
landslides and quartzite erratic boulders, well described by Kempter et
al. (2007). Younger alluvium (late Pleistocene to Holocene) consists of
sands, silts and conglomerates that lie at or within 2 m above modern
drainage bottoms. Older alluvium found in the bottom of Cañon del
Cobre is composed of sand with minor gravel and silty-clayey beds
preserved as 0.2 to 2.0-m-thick terrace deposits on top of sandstone
benches. Late Pliocene-middle Pleistocene gravels may form relatively
thin veneers on high-level erosion surfaces west of the western escarp
ment of Cañon del Cobre (D. Koning, written commun., 2009). Collu
vium is mostly poorly sorted talus and rock debris, often as hillslope
deposits up to 8 m thick, and is of likely late Pleistocene to Holocene age.
Landslide deposits on steep slopes are thicker, some more than 20 m
thick, and composed of unsorted, chaotic debris that was emplaced dur
ing single, mass-wasting events, such as slump or block slides (Kempter
et al., 2007). Kempter et al. (2007) suggested a catastrophic flood event
emplaced Proterozoic quartzite boulders over 2 m in diameter, particu
larly on Poleo Formation dipslopes northeast of Las Minas Jimmie.

STRUCTURE
Cañon del Cobre is located on the southeastern edge of the Colo
rado Plateau, where the Plateau adjoins the Abiquiu embayment of the
Rio Grande rift. Westward-dipping strata on the west side of the canyon,
and eastward-dipping strata on its east side give Cañon del Cobre the
overall appearance of a large, breached anticlinal structure.
Cañon del Cobre is traversed and bounded by several east-north
east-trending, high angle normal faults, and its eastern wall is along a
faulted, north-northeast-trending anticline that plunges to the southwest
(Smith et al., 1961; Kempter et al., 2007) (Figs. 2, 8). The faults in Cañon
del Cobre are down to the north, east or west and most have less than 50
m of offset. These structures are part of the broad boundary between the
eastern Colorado Plateau and the Rio Grande rift. Most of them formed
during the Neogene extensional tectonism associated with the develop
ment of the rift.
ACKNOWLEDGMENTS
FIGURE 7. Measured stratigraphic section of the Chinle Group strata exposed
We thank several colleagues and volunteers for field assistance in
at Minas de Pedro on the eastern rim of Cañon del Cobre. See Appendix 2
for description of section. Modified from Lucas and Hunt (1992).
studying the geology of Cañon del Cobre. Dan Koning, John Nelson and
Joerg Schneider provided helpful reviews of the manuscript.
22
REFERENCES

Ash, S.R., 1974, Upper Triassic plants of Cañon del Cobre, New Mexico: Lucas, S.G. and Krainer, K., 2005a, Stratigraphy and correlation of the
New Mexico Geological Society, Guidebook 25, p. 179-184. Permo-Carboniferous Cutler Group, Chama basin, New Mexico: New
Chapin, C. E. and Cather, S. M., 1981, Eocene tectonics and sedimentation Mexico Geological Society, Guidebook 56, p. 145-159.
in the Colorado Plateau-Rocky Mountain area: Arizona Geological Soci Lucas, S.G. and Krainer, K., 2005b, Cutler Group (Permo-Carboniferous)
ety Digest, v. 14, p. 173-198. stratigraphy, Chama Basin, New Mexico: New Mexico Museum of Natu
Eberth, D.A. and Berman, D.S, 1983, Sedimentology and paleontology of ral History and Science, Bulletin 31, p. 90-100.
Lower Permian fluvial redbeds of north-central New Mexico – prelimi Lucas, S. G., Harris, S. K. and Spielmann, J. A., 2010, One hundred and fifty
nary report: New Mexico Geology, v. 5, p. 21-25. years of geological and paleontological research in Cañon del Cobre,
Eberth, D.A. and Berman, D.S, 1993, Stratigraphy, sedimentology and ver northern New Mexico: New Mexico Museum of Natural History and
tebrate paleontology of the Cutler Formation redbeds (Pennsylvanian Science Bulletin, this volume.
Permian) of north-central New Mexico: New Mexico Museum of Natu Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S and Henrici, A.C.,
ral History and Science, Bulletin 2, p. 33-48. 2005b, Vertebrate biostratigraphy and biochronology of the Pennsylva
Eberth, D.A. and Miall, A.D., 1991, Stratigraphy, sedimentology and evolu nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico:
tion of a vertebrate-bearing, braided to anastomosed fluvial system, New Mexico Museum of Natural History and Science, Bulletin 31, p.
Cutler Formation (Permian-Pennsylvanian), north-central New Mexico: 128-139.
Sedimentary Geology, v. 72, p. 225-252. Lucas, S.G., Zeigler, K.E., Heckert, A.B. and Hunt, A.P., 2003, Upper Tri
Fracasso, M.A., 1987, Fluvial deposition and paleoenvironment of the assic stratigraphy and biostratigraphy, Chama basin, north-central New
Cutler Formation red beds, El Cobre Canyon, New Mexico: New Mexico Mexico: New Mexico Museum of Natural History and Science, Bulletin
Geology, v. 9, p. 14-18. 24, p. 15-39.
Heckert, A. B. and Lucas, S. G., 2003, Triassic stratigraphy in the Zuni Lucas, S.G., Zeigler, K.E., Heckert, A.B. and Hunt, A.P., 2005a, Review of
Mountains, west-central New Mexico: New Mexico Geological Society, Upper Triassic stratigraphy and biostratigraphy in the Chama basin,
Guidebook 54, p. 245-262. northern New Mexico: New Mexico Geological Society, Guidebook 56,
Kempter, K., Zeigler, K., Koning, D. and Lucas, S., 2007, Preliminary p. 170-181.
geologic map of the Canjilon SE quadrangle, Rio Arriba County, New Maldonado, F. and Kelley, S.A., 2009, Revision to the stratigraphic nomen
Mexico: New Mexico Bureau of Geology and Mineral Resources, Open clature of the Abiquiu Formation, Abiquiu and contiguous areas, north
file Geological Map 150, 1:24,000. central New Mexico: New Mexico Geology, v. 31, p. 3-8.
Logsdon, M. J., 1981, A preliminary basin analysis of the El Rito Formation Newberry, J.S., 1876, Geological report; in Macomb, J.N., ed., Report of the
(Eocene), north-central New Mexico: Geological Society of America exploring expedition from Santa Fe…in 1859: U.S. Army Engineer
Bulletin, pt. I, v. 92, p. 968-975. Department, Washington, D.C., p. 9-118.
Lucas, S.G., 1984, Correlation of Eocene rocks of the northern Rio Grande Smith, C.T., Budding, A.J. and Pitrat, C.W., 1961, Geology of the southeast
rift and adjacent areas: Implications for Laramide tectonics: New Mexico ern part of the Chama basin: New Mexico Bureau of Mines and Mineral
Geological Society, Guidebook 35, p. 123-128. Resources, Bulletin 75, 57 p.
Lucas, S.G., 1993, The Chinle Group: Revised stratigraphy and biochronology Stewart, J.H., Poole, F.G. and Wilson, R.F., 1972, Stratigraphy and origin of
of Upper Triassic strata in the western United States: Museum of North the Chinle Formation and related Upper Triassic strata of the Colorado
ern Arizona, Bulletin 59, p. 27-50. Plateau region: U.S. Geological Survey, Professional Paper 690, 336 p.
Lucas, S.G. and Hunt, A.P., 1992, Triassic stratigraphy and paleontology, Wood, G.H. and Northrop, S.A., 1946, Geology of the Nacimiento Moun
Chama basin and adjacent areas, north-central New Mexico: New Mexico tains, San Pedro Mountain, and adjacent plateaus of Sandoval and Rio
Geological Society, Guidebook 43, p. 151-167. Arriba counties, New Mexico: U.S. Geological Survey, Oil and Gas Inves
tigations Map OM-57.

APPENDIX1
Cutler Group Measured Section, Cañon de Cobre Arroyo del Agua Formation:
Measured along the western floor and wall of El Cobre Canyon in 84.Sandy siltstone; mottled pale red (10R6/2) and light greenish
secs. 25-26, T24N, R5E. Base at UTM Zone 13, 378369E, 4016319 gray (5GY8/1); not calcareous; blocky. 1.5
(NAD27) and top at 376985E, 4016278N. Strata dip 8° to S40°W. This 83.Sandy siltstone; same as unit 75. 6.0
is section in Figure 5; also see Lucas and Krainer (2005a). Thicknesses 82. Sandstone; yellowish gray (5Y8/1) and moderate reddish brown
are in meters. (10R4/6); arkosic; coarse grained; not calcareous; trough
crossbedded; bench. 3.4
Late Triassic: 81.Sandy siltstone; moderate reddish brown (10R4/6); micaceous;
not calcareous. 1.3
Chinle Group: 80.Sandy siltstone; same as unit 75. 6.0
Shinarump Formation: 79.Sandstone; same as unit 76.3.3
78.Calcrete pebble conglomerate; pale yellowish brown (10YR6/
85.Conglomeratic sandstone; very pale orange (10YR8/2) and 2) and light olive gray (5Y6/1); bench. 0.7
moderate yellowish brown (10YR5/4); quartzose; coarse 77.Sandy siltstone; same as unit 75. 4.5
grained; not calcareous; clasts are chert, jasper and quartzite 76.Sandstone; moderate reddish brown (10R4/6) and pale reddish
pebbles; trough crossbedded. Not measured. brown (10R5/4); arkosic; coarse grained; not calcareous; trough
crossbedded; multistoried. 2.6
unconformity 75.Sandy siltstone; moderate reddish brown (10R4/6); slightly
Permo-Pennsylvanian: calcareous; some calcrete nodules; slope. 27.0
74.Sandstone; moderate reddish brown (10R4/6); arkosic; coarse
Cutler Group: grained; trough crossbeded. 1.5
23
73.Muddy siltstone; moderate reddish brown (10R4/6); calcrete 38.Sandstone; pinkish gray (5YR8/1) and light brownish gray
nodules. 14.0 (5YR6/1); arkosic; coarse grained; very calcareous; ripple
72.Conglomeratic sandstone; pale reddish brown (10R5/4) and laminated. 0.3
moderate reddish brown (10R4/6); arkosic; coarse grained; 37.Sandy siltstone; pale reddish brown (10R5/4); abundant
conglomerate is siltstone and calcrete rip-ups; trough calcrete nodules; blocky; slope. 2.7
crossbedded; scour base; multistoried bench. 5.7 36.Sandstone; same as unit 38; thin bioturbated ledge. 1.8
71.Muddy siltstone; same as unit 73. 5.2 35.Sandy siltstone; same as unit 37; some thin sandstone lenses
70. Sandstone; yellowish gray (5Y8/1) and moderate reddish brown like unit 38. 3.2
(10R4/6); arkosic; coarse grained; calcareous; trough 34.Sandstone; same as unit 32. 0.5
crossbedded; multistoried. 2.8 33.Sandy siltstone; same as unit 31.1.0
69.Sandstone; moderate reddish brown (10R4/6); arkosic; 32.Sandstone; pinkish gray (5YR8/1) and light olive gray (5Y6/1);
micaceous; coarse grained; slightly calcareous; in ledges (0.3-m arkosic; calcareous; coarse grained; bioturbated. 1.0
thick) separated by blocky siltstone like unit 68; some calcrete. 31.Sandy siltstone; pale reddish brown (10R5/4); micaceous;
6.0 numerouscalcrete nodules; blocky slope; a few lenses of
68.Sandy siltstone; same as unit 60; many calcrete nodules. 7.0 sandstone like unit 32.4.0
67.Sandstone; same as unit 63. 2.0 30.Sandstone; yellowish gray (5Y8/1) and light olive gray (5Y6/
66.Sandy siltstone; same as unit 60. 6.0 1); arkosic; coarse grained; calcareous; trough crossbedded. 3.5
65.Conglomeratic sandstone; sandstone is light greenish gray 29.Sandstone; same as unit 27, bench. 2.9
(5GY8/1) and moderate reddish brown (10R4/6); arkosic, coarse 28.Sandy siltstone; pale red (10R6/2) and pale reddish brown
grained and calcareous; clasts are calcrete pellets; trough (10R5/4); laminar; calcareous. 0.8
crossbedded. 1.2 27.Sandstone; yellowish gray (5Y8/1); arkosic; coarse grained;
64.Sandy siltstone; same as unit 60. 1.6 very calcareous; trough crossbedded; bench. 2.9
63.Sandstone; pale reddish brown (10R5/4) and moderate reddish 26.Sandy shale; light olive gray (5Y6/1); not calcareous; plant
brown (10R4/6); arkosic; coarse grained; trough crossbedded; debris. 2.5
scour base with 3 m of relief. 6.0 25.Conglomeratic sandstone; yellowish gray (5Y8/1) and moder
62.Sandy siltstone; same as unit 60. 0.8 ate orange pink (10R7/4); arkosic; coarse grained; very
61.Sandstone; same as unit 61. 0.7 calcareous; clasts are quartzite up to 6 cm diameter; trough
60.Sandy siltstone; moderate reddish brown (10R4/6); not crossbedded. 1.9
calcareous; blocky; calcrete nodules. 4.6 24.Sandy siltstone; same as unit 18.1.6
59.Sandy siltstone; same as unit 54; abundant calcrete. 4.0 23.Calcrete ledge; light olive gray (5Y6/1); some associated calcrete
58.Sandstone; moderate reddish brown (10R4/6); arkosic; coarse pebble conglomerate. 1.3
to very coarse grained; calcareous; some calcrete pebbles; trough 22.Sandy siltstone; same as unit 18; blocky; slope. 3.1
crossbedded; scour base locally cuts down to unit 55. 3.0 21.Calcrete ledge; same as unit 23.0.1
57.Sandy siltstone; grayish red (10R4/2); calcareous; blocky. 1.0 20.Sandy siltstone; same as unit 18. 0.7
56.Sandy siltstone; same as unit 54.6.2 19.Sandstone; pale red (10R6/2); arkosic; fine grained; very
55.Sandstone; pebbly at base; moderate reddish brown (10R4/6) calcareous; massive; cuesta. 0.6
and pale red (10R6/2); arkosic; coarse to very coarse grained; 18.Sandy siltstone; pale reddish brown (10R5/4); micaceous; not
calcareous; pebbles are siltstone and calcrete rip ups; trough calcareous; numerous rhizoliths. 7.5
crossbedded; bench. 1.7 17.Conglomerate; light olive gray (5Y6/1); calcrete pebbles in
54.Sandy siltstone; moderate reddish brown (10R4/6); calcare coarse-grained arkosic sandstone matrix; trough crossbedded;
ous; blocky. 8.7 bench. 1.6
16.Sandy siltstone; same as unit 18.3.3
El Cobre Canyon Formation: 15.Sandstone; same as unit 6. 0.2
53.Sandstone; pale reddish brown (10R5/4); fine grained; arkosic; 14.Sandy siltstone; same as unit 7.1.9
very calcareous; laminar and bioturbated. 1.7 13.Sandstone; yellowish gray (5Y8/1) and moderate orange pink
52.Sandy siltstone; same as unit 44. 1.6 (10R7/4); arkosic; coarse grained; calcareous; trough
51.Sandstone; same as unit 47. 0.6 crossbedded; multistoried bench. 2.4
50.Sandy siltstone; same as unit 44. 0.2 12.Sandy siltstone; same as unit 7 except trough crossbedded. 2.2
49.Sandstone; same as unit 47. 0.3 11.Sandy siltstone; same as unit 7.5.0
48.Sandy siltstone; same as unit 44.1.0 10.Sandstone; pale reddish brown (10R5/4); arkosic; coarse
47. Sandstone; pale reddish brown (10R5/4) and grayish red (10R4/ grained; not calcareous; trough crossbedded; top surface has
2); fine grained; arkosic; calcareous; massive. 0.8 desiccation tracks. 1.2
46.Sandy siltstone; same as unit 44. 1.5 9. Sandy siltstone; same as unit 7.2.2
45.Sandstone; pale reddish brown (10R5/4); arkosic; fine grained; 8. Sandstone; pale reddish brown (10R5/4); arkosic; medium
calcareous; bioturbated. 0.6 grained; very calcareous; some calcrete rip-ups; trough
44.Sandy siltstone; pale reddish brown (10R5/4); calcareous; crossbedded; multistoried bench. 2.0
calcrete; blocky. 2.0 7. Sandy siltstone; grayish red (10R4/2) and pale reddish brown
43.Conglomeratic sandstone; pale reddish brown (10R5/4); coarse (10R5/4); micaceous; not calcareous; abundant rhizoliths. 1.6
to very coarse grained; arkosic; clasts are quartzite and calcrete 6. Sandstone; light olive gray (5Y6/1); arkosic; micaceous;
pebbles; very calcareous; trough crossbedded; scour base; bench. medium grained; calcareous; ripple laminated. 0.4
2.7 5. Conglomerate; light olive gray (5Y6/1); calcrete pebbles and a
42.Sandy siltstone; same as unit 44 except laminar. 1.8 sparse matrix of arkosic sand grains; bench. 0.2
41.Sandy siltstone; same as unit 44. 7.3 4. Sandy siltstone; pale reddish brown (10R5/4) and grayish red
40.Sandstone; same as unit 38. 0.6 (10R4/2); micaceous; not calcareous; laminar. 5.5
39.Sandy siltstone; same as unit 44.3.8 3. Conglomeratic sandstone; same as unit 1.3.2
24
2. Sandy siltstone; pale reddish brown (10R5/4); micaceous;
unconformity (Tr-4)
calcareous; blocky. 7.7
1. Conglomeratic sandstone; pale reddish brown (10R5/4); arkosic; Salitral Formation:
micaceous; very calcareous; coarse grained; clasts are gray 8. Sandy mudstone and muddy sandstone: sandy mudstone is
quartzite; trough crossbedded; multistoried. 0.9 mottled grayish red purple (5RP4/2) and yellowish gray (5Y
APPENDIX2 7/2), bentonitic and very calcareous; muddy sandstone is
medium light gray (N6), fine grained, subangular, poorly sorted,
Chinle Group Section, El Cobre Canyon lithic wacke; trough crossbedded ledges with mudstone
partings. 5.5
Measured in the SE¼ SW¼ sec. 18 (unsurveyed), T24N, R06E,
Rio Arriba County. Modified Minas de Pedro section of Lucas and Hunt
Shinarump Formation:
(1992, p. 172). See Figure 7 for graphic illustration of section. Thick
nesses are in meters. 7. Sandstone: pale greenish yellow (10Y 8/2); fine-medium grained;
subrounded; moderately sorted; quartzose; slightly
Chinle Group: calcareous; trough crossbedded; some siliceous pebbles. 3.0
6. Sandstone: yellowish gray (5Y 7/2); very fine-medium grained;
Poleo Formation: subrounded; poorly sorted; feldspathic litharenitic;
15.Sandstone with conglomeratic beds: sandstone is light olive noncalcareous; trough crossbedded; contains some siliceous
brown (5Y 5/6), very fine grained, subangular-subrounded, pebbles; upper half of unit thinner bedded with some
moderately sorted, noncalcareous, micaceous litharenitic; mudstone partings. 3.1
conglomerate is light greenish gray (5GY 8/1) and composed of 5. Sandy mudstone: pale reddish purple (5RP6/2); noncalcareous.
siltstone and nodular calcrete clasts; laminar and planar 1.5
crossbedded with conglomerate concentrated at scour base. 4. Sandstone: yellowish gray (5Y 7/2); fine-medium grained;
11.3 subrounded; moderately sorted; micaceous feldspathic
14.Conglomerate: matrix is pale olive (10Y 6/2), very fine-coarse litharenitic; calcareous; trough crossbedded; some dispersed
grained, subangular quartzose sandstone; clasts are brown, siliceous pebbles. 2.8
3. Conglomerate: matrix is pale green (5G 7/2), fine-medium
white and red chert up to 1 cm in diameter; trough crossbedded.
1.6 grained, subangular, poorly sorted quartzose sandstone; clasts
are greenish black (5G 2/1) chert and white quartzite up to 15
13.Sandstone: yellowish gray (5Y 7/2); very fine-fine grained;
cm in diameter; trough crossbedded. 0.5
subrounded; moderately sorted; micaceous subarkosic; very
calcareous; laminar. 2.2
12.Sandstone: yellowish gray (5Y 7/2); very fine-fine grained; unconformity (Tr-3)
subangular-subrounded; moderately sorted; micaceous Mottled strata (Zuni Mountains Formation):
quartzarenitic; very calcareous; trough crossbedded;
multistoried with pebbly beds at scour bases. 10.3 2. Siltstone: mottled grayish yellow green (5GY 7/2) and dark
11.Sandstone: pale greenish yellow (10Y 8/2), weather moderate reddish brown (10R 3/4); bioturbated and massive. 1.2-1.4
yellowish brown (10Y 5/4); fine grained; subrounded; well
sorted; quartzarenitic; noncalcareous; trough crossbedded. 2.0 Cutler Group:
10.Sandstone:grayish orange (10R 7/4); very fine grained;
Arroyo del Agua Formation:
subrounded; well sorted; arkosic litharenite; calcareous; some
1. Sandstone: moderate reddish brown (10R4/6); very fine-fine
cover. 1.3
9. Conglomeratic sandstone: sandstone is dusky yellow (5Y 6/ grained; subangular; moderately sorted; micaceous lithic wacke;
4), very fine-fine grained, subangular, moderately sorted, noncalcareous; trough crossbedded. 2.0+
micaceous litharenitic, calcareous; pebbles are light gray, black,
white quartzite and petrified wood; trough crossbedded; scour
base with 0.3-0.6 m of relief. 3.0
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

25
SEDIMENTOLOGY OF THE PENNSYLVANIAN-PERMIAN CUTLER GROUP
AND LOWER PERMIANABO FORMATION, NORTHERN NEWMEXICO

KARL KRAINER1 AND SPENCERG. LUCAS2


of Geology and Paleontology, Innsbruck University, Innrain 52, Innsbruck, A-6020 AUSTRIA;
1Institute

2 New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, NM 87104-1375

Abstract—The Cutler Group at Cañon del Cobre, northern New Mexico, is divided into the El Cobre Canyon and
Arroyo del Agua formations, both composed of red beds that formed in a braided river system under semi-arid to
arid climatic conditions. Sandstones are dominantly arkoses to lithic arenites, and rarely sublitharenites. Towards
the south the Cutler red beds grade into the Abo Formation, which is well exposed along the Jemez River. Red beds
of the Abo Formation at the Jemez River and Gilman sections do not show any significant differences in facies and
mineralogical composition when compared to the Cutler red beds. Sandstone sheets represent deposits of a braided
river system; the intercalated sandstone beds and lenses are interpreted as sheet splays and minor channel fills
which formed during overbank flooding. Siltstone and mudstone units, which are the dominant lithofacies of the
Abo Formation, represent floodplain deposits. Whereas the Cutler Group includes red beds of Pennsylvanian age,
the Abo Formation is restricted to the Early Permian. As the red beds of the Cutler Group and Abo Formation are
very similar, a sharp boundary cannot be drawn, so we consider the Abo Formation to be part of the Cutler Group.

INTRODUCTION
Siliciclastic red beds of Late Pennsylvanian and Early Permian age
are exposed both north and south of the Jemez Mountains volcanic
edifice in northern New Mexico. North of the Jemez Mountains, in the
Chama River basin, these strata are assigned to the El Cobre Canyon and
the Arroyo del Agua formations of the Upper Pennsylvanian-Lower
Permian Cutler Group. However, to the south, the red beds are assigned
to the Lower Permian Abo Formation. A particularly significant outcrop
of these strata is in Cañon del Cobre (El Cobre Canyon) of Rio Arriba
County, where the type section of the El Cobre Canyon Formation is
overlain by a characteristic section of the Arroyo del Agua Formation.
The nonmarine red beds of the El Cobre Canyon and Arroyo del Agua
formations at Cañon del Cobre are particularly well known for their
fossil vertebrates, and also yield significant trace fossils, plant megafossils
and palynomorphs (see Lucas et al., this volume). Here, we present an
analysis of Abo-Cutler sedimentology in northern New Mexico to deter
mine depositional environments and to aid in the regional correlation of
these strata.
CUTLER GROUP, CAÑON DELCOBRE

Introduction
Red beds between the underlying Rico Formation and overlying
Triassic strata in southern Colorado, southeastern Utah, northeastern
Arizona and northwestern New Mexico were originally termed Cutler
Formation, named after Cutler Creek, west of Ouray, Colorado (see Jicha
and Lochman-Balk, 1958). Due to strong lateral and vertical facies changes
within the “Cutler Formation,” the succession was elevated to group
rank and divided into several formations of distinctive lateral and vertical
extent (Baars, 2000). In southeastern Utah, the Cutler Group FIGURE 1. Index map of part of northern New Mexico showing location of
unconformably overlies the Upper Pennsylvanian (Missourian) Honaker the measured sections at Cañon del Cobre (El Cobre Canyon), the Jemez
Trail Formation and is divided into Elephant Canyon Formation, Halgaito River and Gilman.
Shale, Cedar Mesa Sandstone, Organ Rock Shale and White Rim Sand
stone (e.g., Baars, 2003). In the Chama basin of northern New Mexico, includes strata of the El Cobre Canyon and Arroyo del Agua formations.
the Cutler Group consists of the El Cobre Canyon and overlying Arroyo Fracasso (1987) discussed the sedimentary environment of the
del Agua formations (Lucas and Krainer, 2005a, b). The most completely Cutler Formation at El Cobre Canyon, suggesting deposition of the red
exposed Cutler Group section in this area is at Cañon del Cobre (El beds in a single-channellow sinuosity stream system in an upper alluvial
Cobre Canyon), near Abiquiu (Fig. 1). At Cañon del Cobre (Fig. 1), we plain near a distal alluvial fan and braided-stream complex under semiarid
measured two sections: section A (Fig. 2) is 100 m thick and comprises climatic conditions. Fracasso (1987) did not subdivide the sedimentary
the El Cobre Canyon Formation, and section B (Fig. 3) is 255 m thick and succession.
26
and orange red (moderate reddish brown) in the Arroyo del Agua Forma
tion. Lithofacies are classified using the facies codes of Miall (1996). All
lithofacies are listed (Table 1) and were described in detail by Eberth and
Miall (1991).
The basal 11 m of section A (El Cobre Canyon Formation: Fig. 2)
is composed of horizontally-laminated greenish siltstone to fine-grained
sandstone (lithofacies Sh), overlain by greenish to dark gray shale with
plant fossils (lithofacies Fl), and two horizontally-laminated brownish
sandstone beds 0.3 and 0.2 m thick (Sh), separated by 0.6 m of greenish
gray shale (Fl). The sandstone is overlain by gray shale containing plant
fossils, grading into greenish siltstone (Fl) that is sharply overlain by
coarse-grained, red arkosic sandstone.
Conglomerate and sandstone occur as individual beds intercalated
in mudstone-siltstone or as thicker horizons that extend laterally over
more than 100 m, forming sandstone sheets. In the measured sections,
sandstone sheets are up to 10.5 m thick. Thicker sandstone sheets dis
play a fining-upward trend with a thin conglomerate at the base (Gcm,
Gt), grading into large-scale to small-scale trough cross-bedded sand
stone (St). Locally, medium- to fine-grained, horizontally-laminated sand
stone (Sh) and massive sandstone (Sm), and, rarely, ripple-cross-lami
nated sandstone (Sr), form the tops of sandstone sheets. Eberth and
Miall (1991) distinguished major and minor sandstone sheets, which
they described in detail. Locally, conglomerate and trough cross-bedded
sandstone are erosively cut into the underlying mudstone-siltstone, form
ing a relief up to 6.3 m deep and representing sandstone ribbons (see
Eberth and Miall, 1991 for details).
Conglomerate occurs as lithofacies Gcm (massive, clast supported;
lithofacies Gm of Eberth and Miall, 1991) and Gt (trough cross-bedded).
Maximum clast size is mostly 5 to 6 cm, rarely up to 15 cm. The
dominant clast type is carbonate derived from reworking of caliche hori
zons and caliche nodules within red mudstone-siltstone. Subordinately,
extraformational clasts are present. Conglomerate occurs at the base of
sheet sandstones and ribbon sandstones and as single conglomerate beds
intercalated in red mudstone-siltstone. Conglomerates are erosively cut
into underlying red mudstone-siltstone.
Trough cross-bedding (lithofacies St) is the most common sedi
mentary structure of sandstone. Large-scale trough cross-bedding sets
thicker than 1 m are rare and represent lateral accretion deposits (lithofacies
Sla of Eberth and Miall, 1991). Medium-scale (> 50 cm) and small-scale
(< 50 cm) trough cross-bedded sandstone represents channel-fill depos
its (Sti and Stii of Eberth and Miall, 1991). Coarse sandstone may be
pebbly.
Lithofacies Se contains mudstone-siltstone rip-up clasts and rep
resents erosional scours. Sandstone with low-angle cross-bedding is very
rare. Horizontally laminated sandstone (Sh) and massive sandstone (Sm)
occur as individual sandstone beds up to 0.5 m thick intercalated in
mudstone-siltstone and in the upper part of sandstone sheets associated
FIGURE 2. Measured section of the El Cobre Canyon Formation (section with trough cross-bedded sandstone. Ripple cross-lamination (Sr) is
A) at Cañon del Cobre. See Appendix for map coordinates of section. very rare. Medium- to fine-grained sandstone occasionally is bioturbated
(principally Planolites). Rarely, desiccation cracks are observed in sand
A detailed facies analysis of the Cutler Group at El Cobre Canyon stone.
was presented by Eberth and Miall (1991) and Eberth and Berman (1993). Mudstone-siltstone is more abundant in the Arroyo del Agua
The El Cobre Canyon Formation corresponds to megasequence 1 of
Formation than in the El Cobre Canyon Formation (Fig. 3). Mudstone
Eberth and Miall (1991), and the Arroyo del Agua Formation corre siltstone intervals are up to 7.5 m thick in the Arroyo del Agua Forma
sponds to megasequences 2 and 3 of Eberth and Miall (1991), although tion, and up to 27 m thick in its uppermost strata. The color is brownish
the boundary between El Cobre Canyon and Arroyo del Agua formations red in the El Cobre Canyon Formation and orange-red in the Arroyo del
does not correspond to the boundary between megasequence 1 and 2 but Agua Formation. Mudstone-siltstone is mostly massive (lithofacies Fsm),
lies slightly higher, within the lower part of megasequence 2 (Lucas and
rarely laminated (Fl) and locally contains caliche nodules that are concen
Krainer, 2005a, b). Here, we base our interpretation of deposition of the trated in thin layers (lithofacies Pc). A few mudstone-siltstone horizons
El Cobre Canyon Formation primarily on data gathered from the two
contain rhizoliths (lithofacies Fr), and rare desiccation cracks are ob
sections we measured (Fig. 2-3). served. Mudstone-siltstone is summarized under lithofacies Flin Eberth
Lithofacies and Miall (1991).
Besides pedogenic limestone (caliche, lithofacies Lp), Eberth and
Except for the basal 11 m of section A, the sediments are brown Miall (1991) distinguished other limestone types such as limestone nod
ish-red (pale reddish brown) colored in the El Cobre Canyon Formation ules of diagenetic origin (Ld), lacustrine marls (Lm) and limestone of
travertine/tufa origin (Lt).
27

FIGURE 3. Type section of the El Cobre Canyon Formation at Cañon del Cobre. See Appendix for map coordinates of section.
28
Table 1. Lithofacies. TABLE 2. Mineralogic composition of Cutler Group sandstones. For
acronyms (column heads), see captions to Figures 9 and 10.

(megasequence 3, uppermost part of the Arroyo del Agua Formation).


Climate changed from semi-arid to increasingly arid conditions during
sedimentation of the Cutler Group.
ABO FORMATION (JEMEZRIVERAND GILMAN)

Introduction
Along the Jemez River, the measured section of the Abo Forma
tion is 143 m thick (Fig. 4). The base is not exposed, and Abo red beds are
overlain by eolian sandstone of the “Meseta Blanca Sandstone Member
of the Yeso Formation” (DeChelly Formation of the Yeso Group: Lucas
et al., 2005). At Gilman, the Abo Formation is completely exposed, and
the measured thickness is 131 m (Figs. 5-6). The Abo Formation is
underlain by the Guadalupe Box Formation (Krainer et al., 2005) and
overlain by a few meters of Yeso red beds and lower Pleistocene Bandelier
Tuff.
At both localities, the Abo can be divided into lower and upper
members. The lower member is dominated by mudstone and siltstone
that constitute 69% of the Abo section at the Jemez River and 72% at
Gilman. In the upper member, mudstone and siltstone constitute much
Petrography less of the Abo section— 25% at the Jemez River and 41% at Gilman.
The lower member measures 83 m at the Jemez River and 90 m at
Sandstones of the Cutler Group are mostly medium- to coarse Gilman, and the upper member is 60 m at the Jemez River and 41 m at
grained, moderately to poorly sorted, angular arkoses to lithic arenites, Gilman. The lower member probably corresponds to the upper part of
and rarely sublitharenites (see details in Lucas and Krainer, 2005a, b; also the El Cobre Canyon Formation, and the upper member to the Arroyo
see Table 2). del Agua Formation of the Cutler Group (Lucas and Krainer, 2005a, b),
and, respectively, to the Scholle Member and Cañon de Espinoso Mem
Depositional Environments
ber of the Abo type section in central New Mexico (Lucas et al., 2005).
Our data are consistent with and support the interpretations of
Lithofacies
depositional environments presented by Eberth and Miall (1991) and
Eberth and Berman (1993). They recognized eight architectural elements Lithofacies were classified according to the scheme proposed by
in Cutler Group strata, including major sandstone sheets and ribbons, Miall (1978, 1981, 1996) using the facies codes of Miall (1996). In
minor sandstone sheets and lenses, stacked sheets and lenses, U-shaped general, in the Abo Formation the same lithofacies types as in the Cutler
mixed-fill, and two types of floodplain deposits. The dominant architec Group are present (lithofacies types are listed in Table 1).
tural element composed of coarse clastics is major sandstone sheets that Lithofacies Gcm and Gt at the Jemez River contain intraclasts,
are interpreted as braided, fluvial channel deposits. Major sandstone mostly carbonate clasts that are derived from caliche. Subordinate mud
ribbons are less common and interpreted to represent anastomosed flu stone clasts (rip-up clasts) are present. At Gilman, rare (unit 31 of
vial channels. The remaining architectural elements are interpreted as section A and unit 14 of section B) extraformational clasts including rare
overbank deposits, including crevasse channel and splay deposits, and granite and abundant well rounded quartzite clasts occur. Lithofacies Se
floodplain deposits. is rare and occurs as thin lenticular beds containing abundant red mud
During deposition of the El Cobre Canyon Formation (approxi stone rip-up clasts.
mately correlating to megasequence 1 of Eberth and Miall, 1991), broad Trough cross-bedding (St) is the most common sedimentary struc
and shallow channels of a braided river system (major sandstone sheets) ture, occurring at different scales (Fig. 7). Large-scale trough-crossbedding
developed under semiarid climatic conditions. Overbank areas were fre (St-l), which represents lateral accretion units, is rare and occurs in the
quently inundated by sheet floods and splays, causing deposition of lower part of sandstone sheets, grading upward into medium- to small
minor sandstone sheets and lenses, and stacked sheets and lenses. Small scale trough cross-bedded sandstones (St-m, St-s) and horizontally-lami
ponds filled with fine-grained lacustrine sediments were common (for nated sandstone (Sh). Sandstone displaying large-scale trough
example, the pond deposit at the base of section A). crossbedding may be pebbly, containing individual clasts up to 5 cm in
During deposition of theArroyo del Agua Formation (megasequence diameter.
2 of Eberth and Miall, 1991), channels became narrower and between the Medium- to small-scale trough cross-bedding (St-m, St-s) is the
channels extensive floodplains developed on which locally pedogenic most common sedimentary structure of sandstone sheets, forming stacked
carbonate horizons (caliche) formed. According to Eberth and Miall cosets of multistory channel fill successions with fining-upward trends
(1991), renewed uplift caused development of a poorly confined shallow and internal erosion surfaces. Low-angle cross-bedding (Sl) is rare and
braided river system and deposition of coarse clastic sediments occurs as thin beds associated with medium-scale trough cross-bedding
29

FIGURE 4. Measured section through the Abo Formation at Jemez River. See Appendix for map coordinates of section.
30
and horizontal lamination in the upper part of sandstone sheets.
Sandstone with horizontal lamination (Sh) occurs as thin beds
intercalated in mudstone or in the upper, finer-grained part of sandstone
sheets. Massive to crudely stratified sandstone (Sm) occurs as interca
lated beds up to 0.3 m thick within mudstone or associated with trough
cross-bedded (St-m, St-s) and horizontally laminated sandstone (Sh) in
the upper part of sandstone sheets. Asymmetric current ripples with
ripple cross-lamination are rare at the Jemez River where they occur in
units 30 and 32 in the fine-grained upper part of sandstone sheets asso
ciated with horizontally-laminated sandstone. Ripples were not observed
at Gilman.
Mudstone-siltstone units are up to 13.4 m thick, and partly cov
ered. The color is red to purple, and the beds are massive (Fsm) or rarely
laminated (Fl). At the Jemez River desiccation cracks occur in mudstone
of unit 18 (Fm). The red mudstone of unit 31 at the Jemez River is red
greenish mottled, containing rhizoliths (lithofacies Fr; Fig. 8). Large cal
cified root structures are also present in coarse, pebbly sandstone (unit
28) at the Jemez River.
Bioturbated sandstone is present at Gilman and occurs as a 0.7-m
thick, dark red, fine-grained sandstone unit intercalated in purple-red
mudstone in the lower member (unit 26 of section A) and in medium- to
fine-grained sandstone on top of sandstone sheets (units 2, 21, 30) of
section B. Soft sediment deformation (convolute bedding) was observed
in one fine-grained sandstone bed on top of a fining upward sequence
(unit 20) at the Jemez River.
Many of the mudstone-siltstone units contain abundant small
carbonate nodules (caliche nodules). Pedogenic carbonate horizons (cali
che, lithofacies P) occur in the lower member of both sections. At the
Jemez River a pedogenic horizon 1.2 m thick composed of pedogenic
limestone nodules in red mudstone occurs near the base (unit 2). Another
pedogenic horizon (unit 14) is 1.4 m thick and composed of dispersed
pedogenic limestone nodules in red mudstone, overlain by a nodular
caliche bed. At Gilman, two thin pedogenic carbonate horizons 0.3 and
0.5 m thick occur in the lower member in the red to purple mudstone of
unit 1 in section B. Another pedogenic carbonate horizon (nodular cali
che) 0.3 m thick is intercalated in red siltstone in the upper part of the
lower member (unit 18 of section B).
Architectural Elements
According to Miall (1985, 1992, 1996), architectural elements are
characterized by a distinctive facies assemblage, internal geometry, exter
nal form and, partly, by a distinct vertical profile. Miall proposed eight
basic types. Each type represents a particular process occurring within
a depositional system. Within the Abo Formation we distinguished three
architectural elements—sandstone sheets, intercalated sandstone beds
and lenses and siltstone-mudstone.

Sandstone sheets
The most characteristic and distinct facies assemblages in the Abo
Formation are sandstone sheets (Fig. 8), which correspond to the archi
tectural element CH (channel). Sandstone sheets are characterized by
width-to-depth ratios greater than 15, whereas sandstone ribbons also
corresponding to the element CH, have ratios smaller than 15.
Within the Abo Formation sandstone sheets form prominent, re
sistantledges that can be traced laterally over long (commonly more than
100 m, up to several hundred meters) distances. The base is generally
erosive, and lithofacies commonly show an upward decrease in bed thick
ness and grain size. Common lithofacies are Gcm and Gt at the base,
grading into multistoried trough cross-bedded sandstone (St), which in
the upper part may be associated with lithofacies Sh, Sm, Sl and Sr.
Rarely, fine-grained sandstone in the uppermost part is bioturbated.
FIGURE 5. Measured section through the Abo Formation (lower part) at Synsedimentary deformation structures are rarely observed in fine-grained
Gilman. See Appendix for map coordinates of section. Unit 25 = unit 1 in sandstone in the upper part. Sandstone sheets are very similar to the
section B (Fig. 6). “major sandstone sheets” that Eberth and Miall (1991) described from
31
the Cutler red beds. Sandstone ribbons have not been observed in the
studied sections.
Abundance of internal scours and erosional surfaces, presence of
conglomeratic lags at the base, variation in grain-size, and poorly- to
well-developed fining-upward sequences are indicative of ephemeral flu
vial regimes (e.g., Picard and High, 1973; Tunbridge, 1981, 1984; Miall,
1996). Sheet sandstones are interpreted as deposits of broad, shallow
channels of a probable braided river system. The lack of lateral accretion
deposits indicates that the channels were shallow, so that large macroform
bars could not develop.
The sandstone sheets are very similar to the “major sandstone
sheets” of the Cutler Group described in detail by Eberth and Miall
(1991). These authors compared the sheet sandstones to those of the
modern Platte River (model 9 of Miall, 1985).

Intercalated sandstone beds and lenses


Thin intercalated sandstone beds and lenses are mostly 10 to 30
cm, rarely up to 50 cm thick. They occur as single sandstone beds of
lithofacies St-s, St-m, Sh, Sl and Sm. Stacked sandstone units are up to 1
m thick. The sandstone beds occur as tabular or lens-shaped bodies; the
base may be erosive. Rarely, sandstone sheets are bioturbated. This
architectural element is similar to the “minor sandstone sheets and lenses”
that Eberth and Miall (1991) described from the Cutler red beds.
Intercalated, laterally restricted sheet sands and lenses are charac
teristic of overbank flooding, particularly in ephemeral systems (e.g.,
Williams, 1971; Picard and High, 1973; Tunbridge, 1981). The tabular
sandstone beds probably originated as sheet splays. Basal erosional
surfaces with mudstone rip-up clasts and reworked caliche clasts indi
cate high-energy conditions of rapid flooding and reworking of mudstone
and caliche beds on the floodplain. Lack of internal erosion and reactiva
tion surfaces within the intercalated sandstone beds is typical of sheet
like, non-channelized flow. Sandstone lenses represent minor channel
fills and may represent the feeder channels (crevasse channels) of the
sheet splays.

Siltstone-mudstone
Siltstone and mudstone units are the dominant lithofacies of the
Abo Formation, occurring as lithofacies Fsm, Fm, Fl and Fr (Fig. 8).
Rarely, pedogenic limestone is intercalated.
The siltstone-mudstone facies that forms sheet-like units extend
ing laterally over at least hundreds of meters belongs to element FF
(floodplain fine) of Miall (1996) and is interpreted as floodplain depos
its formed by settling from sheetfloods. The presence of lamination
(lithofacies Fl) indicates deposition from suspension. However, stratifi
cation is mostly absent (lithofacies Fsm), probably due to the presence
of organisms (bioturbation or pedoturbation). Desiccation cracks indi
cate periodic drying out; long-term drying out led to the formation of
calcic paleosols (pedogenic limestone)
Sedimentary Petrography
From both sections, petrographic thin sections of sandstones were
studied under the microscope (Jemez River 6, Gilman 14). Sandstones of
both sections are very similar in petrographic composition and textural
properties (see Table 2).
Coarse-grained sandstones are moderately to poorly sorted; finer
grained sandstones commonly are moderately to well sorted. Most grains
are angular to subangular, but subrounded to rounded grains also occur. In
particular, sedimentary rock fragments (reworked caliche grains) are
rounded to well rounded. In some sandstones, the original rounding is
diagenetically overprinted as grains are partly (randomly) replaced by
carbonate cement.
FIGURE 6. Measured section through the Abo Formation (upper part) at At the Jemez River all studied sandstones are calcite cemented
Gilman. See Appendix for map coordinates of section. Unit 1 = unit 25 in (Fig. 9), whereas at Gilman both carbonate-cemented sandstones and
section A (Fig. 5). sandstones in which carbonate cement is rare to absent occur (Fig. 10).
32

FIGURE 8. Sandstone sheets (interpreted as deposits of broad shallow


channels of a braided river system), intercalated in siltstone-mudstone
(floodplain deposits); Abo Formation at Jemez River. Rhizoliths are
FIGURE 7. Trough crossbedded sandstone (lithofacies St-m), the most developed in the lower siltstone-mudstone.
common lithofacies type of the sandstone facies within the Abo Formation
(Gilman).
cement, which is coarse, blocky and commonly poikilotopic (Figs. 9,
Monocrystalline quartz is a common constituent of all sandstones. Most 10A-B). Calcite cement randomly replaces detrital grains, particularly
monocrystalline quartz grains are slightly undulose, and subordinately feldspars, and (subordinately) also quartz. In the upper part of the Jemez
nonundulatory quartz grains are present. Riversection, subhedral dolomite rhombs replace calcite, and also replace
Polycrystalline quartz grains are mostly composed of only a few, feldspar and quartz subordinately.
large internal crystals of more equant shape with little or no intercrystalline Authigenic feldspar overgrowths were rarely observed in the Gilman
section. Replacement textures indicate that feldspar and quartz
suturing (Fig. 9). These grains are nonundulose to slightly undulose.
Stretched metamorphic rock fragments composed of abundant small, overgrowths were formed first, followed by calcite cement and finally by
elongate, intensely sutured and strongly undulose crystals are rarely dolomite, which replaced all other cements and detrital grains.
observed. Most of the grains (most mono- and polycrystalline quartz grains,
Detrital feldspars are abundant (Figs. 9 and 10). Most of the detritalfeldspars and magmatic rock fragments) are derived from granitic
feldspars are slightly to strongly altered, although fresh grains occur, too. source rocks. Only a small percentage of the detrital grains (some mono
Potassium feldspars dominate. They are mostly untwinned. Microcline and polycrystalline quartz grains and metamorphic rock fragments) are
is present throughout both sections, except in the lower part of the derived from metamorphic source rocks. Individual layers contain high
Gilman section. Perthitic feldspars are common. Plagioclase showing amounts of sedimentary rock fragments, particularly reworked pedogenic
polysynthetic albite twinning is rare. carbonate grains. The amount of individual grain types varies consider
The most abundant type of rock fragment consists of quartz and ably within one section, depending mainly on the grain size of the sedi
potassium feldspar and is of magmatic origin (Figs. 9A-C, E, 10A). ment. The coarser the sediment the higher is the amount of granitic rock
Rarely, magmatic rock fragments composed of several feldspar crystals fragments and polycrystalline quartz, and the lower is the amount of
occur. Fine-grained schistose metamorphic (phyllitic) rock fragments detrital feldspars, monocrystalline quartz and micas.
composed of elongate quartz and mica occur in very small quantities. The average petrographic composition of the Abo Formation at
Metamorphic rock fragments composed of several mica crystal units are the Jemez River and Gilman is very similar to that of the Cutler Group at
very rare. Cañon del Cobre and Arroyo del Agua (Table 2). Sandstones are domi
Some sandstone beds contain sedimentary rock fragments of re nantly arkoses and lithic arenites, subordinately subarkoses. Minor dif
crystallized micritic carbonate, commonly with smallangular quartz grains ferences exist in the amount of granitic rock fragments, carbonate cement,
(Figs. 9A and 10H). A few carbonate grains display fissure cracks filled and matrix. Slight differences also exist compared to the Abo type section
with calcite. Some carbonate grains display dark brown rims. in central New Mexico, where the amount of monocrystalline quartz is
Detrital micas include brown and greenish biotite, muscovite (Fig. higher, the amount of detrital feldspars is slightly higher and the amount
10F) and rare chlorite. They are more common in fine-grained sandstone of carbonate cement lower than at the Jemez River and at Gilman. These
and rare to almost absent in coarse-grained sandstone. Common acces differences are mainly caused by grain size and petrography of the source
sory minerals (heavy minerals) are zircon, tourmaline and apatite. Gar rocks.
net is present in a few samples from the lower part of both sections. The main source rock type is granite, from which the majority of
The sandstones contain very small amounts of matrix or are free mono- and polycrystalline quartz, almost all detritalfeldspars and micas,
of matrix, resulting in high original porosity and permeability. Bending of and all granitic rock fragments were derived. Metamorphic rocks pro
detrital micas indicates that original porosity was reduced by mechanical vided a few mono- and polycrystalline quartz, and metamorphic rock
compaction. Remaining pore space was filled with cement minerals such fragments. Most sedimentary rock fragments were derived from the re
as quartz, calcite and dolomite. working of caliche; subordinately, reworked mudstone and siltstone grains
Quartz cement occurs as authigenic overgrowths on detrital mono are present. At Gilman, coarse sandstone beds contain up to 58% carbon
and polycrystalline quartz grains. At Gilman, some of the sandstones ate grains (reworked caliche).
contain abundant quartz grains with authigenic quartz overgrowths (Figs. DISCUSSION
10D, G). These sandstones also contain small amounts of matrix, whereas
carbonate cement is rare, occurring as small patches. Most of the sand A comparison of the facies and mineralogical composition demon
stones of both sections are characterized by high amounts of calcite strates that no significant differences exist between the red beds of the
33

FIGURE 9. Thin section photographs (all taken under polarized light) of sandstones from the Abo Formation at Jemez River. Width of each photograph
is 6.3 mm. A, Coarse-grained arkosic sandstone composed of mono- (Qm) and polycrystalline quartz (Qp), detrital feldspars (F) including microcline (M),
sedimentary rock fragments (dark micritic carbonate grains - C) and rare granitic rock fragments (G), cemented by calcite. Sample JE 2. B, C, D, Arkosic
sandstone with mono- (Qm), polycrystalline quartz (Qp), perthitic feldspars (F), microcline (M) and granitic rock fragments (G) cemented by coarse,
poikilotopic calcite. B = sample JE 4, C and D = sample JE 6. E, Sandstone containing mono- (Qm) and polycrystalline quartz (Qp), detrital feldspars (F),
sedimentary rock fragments (dark micritic carbonate grains - C) and rare granitic rock fragments (G), cemented by calcite. Sample JE 2. F, H, Coarse
grained, arkosic sandstone, poorly sorted, angular-subangular, containing mono- (Qm), polycrystalline quartz (Qp) and detrital feldspars (F) cemented by
calcite. F = sample JE 5, H = sample JE 3. G, Arkosic sandstone containing abundant detrital feldspars, mostly perthitic potassium feldspars, microcline,
and detrital quartz grains. Sample JR 7.
34

FIGURE 10. Thin section photographs (all taken under polarized light) of sandstones from the Abo Formation at Gilman. Width of each photograph is
6.3 mm, except for F and G it is 3.2 mm. A, B, Coarse-grained, calcite-cemented arkosic sandstone containing mono- (Qm) and polycrystalline quartz
(Qp), detrital feldspars (F) including microcline (M) and a few granitic rock fragments (G). Feldspars are altered and partly replaced by calcite cement. A
= sample GI 13, B = sample GI 15b. C, Arkosic sandstone containing abundant feldspars (F) including microcline (M), mono- and polycrystalline quartz
and rare granitic rock fragments. Detrital grains are cemented by calcite. Sample GI 15a. D, Arkosic sandstone containing abundant monocrystalline quartz
grains, subordinate polycrystalline quartz grains, detrital feldspars and rare detrital micas (muscovite; yellow, lower right). Many quartz grains display
authigenic overgrowths (arrows). Sample GI 14. E, Arkosic sandstone, densely packed, composed of mono- and polycrystalline quartz, detrital feldspars,
some microcline and rare granitic rock fragments. Sample GI 23. F, Arkosic sandstone containing a large detrital muscovite grain. Sample GI 19. G, Arkosic
sandstone with quartz grains displaying well-developed authigenic quartz overgrowths (arrows). Qm monocrystalline, Qp polycrystalline quartz, F detrital
feldspars, Mi detrital mica. Sample GI 16. H, Coarse-grained sandstone containing abundant well rounded sedimentary rock fragments (micritic carbonate
grains), subordinate granitic rock fragments (G) and detrital feldspars (F). Detrital grains are cemented by calcite. Sample GI 17.
35
Cutler Group and the Abo Formation in northern New Mexico. Differ According to Kues and Giles (2004), the Wolfcampian of northern
ences exist only in age. Whereas the Cutler Group also includes red beds and central New Mexico is represented by nonmarine red beds of the
of Pennsylvanian age, the Abo Formation is restricted to the Early Per Abo Formation and equivalent units. Towards the south, in southern
mian and either overlies Late Pennsylvanian marine strata (e.g. Guadalupe New Mexico (Sierra and Doña Ana counties), the Abo red beds interfinger
Box Formation, Oso Ridge Member of Bursum Formation) or directly with marine carbonate rocks of the Hueco Group. To the west, Abo red
overlies Precambrian basement rocks, as in the southern Jemez Moun beds grade into Cutler red beds in the subsurface northwest of a line that
runs roughly from Cuba to Gallup.
tains east of Cuba (e.g., Woodward, 1996; Krainer et al., 2005).
Due to the similar facies and composition it is impossible to draw At many places in central New Mexico, red beds of the Abo
a boundary between the Permian part of the Cutler Group and the Abo Formation disconformably overlie the Bursum Formation. Locally, as in
Formation in northwestern New Mexico. Indeed, Baars (1983) stated the hills east of Socorro, the Abo Formation base is a conglomerate that
that the arkosic red beds of the Abo Formation are similar in physical erosively overlies the Bursum Formation. Limestone clasts in this con
aspect and age to the lower Cutler beds of the Defiance uplift and Monu glomerate indicate that, locally, substantial parts of the Bursum Forma
tion were eroded prior to deposition of the Abo red beds, probably
ment Valley region. We therefore include theAbo Formation in the Cutler
Group. forming deeply incised valleys (Krainer and Lucas, 2009). This
In the Paradox Basin, Campbell (1980) interpreted the red beds of unconformity is also recognized at other locations and indicates a major
the Cutler Group as alluvial fan deposits that were shed southward from tectonic pulse near the Virgilian/Wolfcampian boundary (Krainer and
the Uncompahgre uplift and southwestward from the San Luis uplift. Lucas, 2009).
Like the Abo Formation in New Mexico, the Cutler Group also displays There is much evidence that the unconformity on top of the Bursum
a facies change from proximal to distal fluvial facies from the Uncompahgre Formation resulted from a major tectonic pulse of the ancestral Rocky
uplift towards the south and southwest (see also Eberth and Miall, 1991; Mountain deformation, causing a significant rejuvenation of basement
Huffman and Condon, 1993). For both the Cutler Group and Abo For uplifts that resulted in increased siliciclastic influx and deposition of
mation of northern New Mexico, the Uncompahgre uplift composed of nonmarine red beds of the Abo Formation in central New Mexico. This
Precambrian rocks was the main source area; both display a very similar unconformity near the Virgilian/Wolfcampian boundary is not developed
within the Cutler Group, in the upper part of the El Cobre Canyon
mineralogical composition. Local sources for the Abo Formation were
the Zuni and Peñasco uplifts. Both the Zuni and the Peñasco uplifts Formation. This underlines the fact that tectonic activity of the ancestral
were already buried during mid- to late Wolfcampian time (Woodward, Rocky Mountain deformation occurred at different times at different
1987). places during the Late Pennsylvanian-Early Permian.
In the northern Nacimiento Mountains, Wood and Northrop (1946)
arbitrarily separated the Abo from the Cutler red beds along the border ACKNOWLEDGMENTS
between Rio Arriba and Sandoval counties. Woodward (1987) called all Jess Hunley, Peter Reser and Justin Spielmann provided assis
of these red beds Abo and reported thicknesses of 76 m at the southern tance in the field. The U.S. Forest Service granted access to land. Felix
end of the Nacimiento Mountains to as much as 884 m along the northern Heller prepared thin sections. We thank Giuseppe Cassinis (Pavia) and
side of San Pedro Mountain. Jörg W. Schneider (Freiberg) for reviewing the manuscript and helpful
suggestions.

REFERENCES

Baars, D.L., 1983, The Colorado Plateau. A geologic history: University of Jicha, H. L. Jr. and Lochman-Balk, C., 1958, Lexicon of New Mexico
New Mexico Press, Albuquerque, 279 p. geologic names: Precambrian through Paleozoic: New Mexico Bureau of
Baars, D.L., 2000, The Colorado Plateau. A geologic history: University of Mines and Mineral Resources, Bulletin 61, 137 p.
New Mexico Press, Albuquerque, 254 p. (revised and updated edition) Krainer, K. and Lucas, S.G., 2009, Cyclic sedimentation of the Upper
Baars, D.L., 2003, Geology of Canyonlands National Park, Utah; in Sprinkel, Carboniferous Bursum Formation, central New Mexico: Tectonics ver
D.A., Chidsey, T.C. Jr. and Anderson, P.B., eds, Geology of Utah´s Parks sus glacioeustasy: New Mexico Geological Society, Guidebook 60, p.
and Monuments (second edition): Utah Geological Association Publica 167-182.
tion 28, p. 61-83. Krainer, K., Vachard, D. and Lucas, S.G., 2005, Lithostratigraphy and bio
Campbell, J.A., 1980, Lower Permian depositional systems and Wolfcampian stratigraphy of the Pennsylvanian-Permian transition in the Jemez
paleogeography, Uncompahgre basin, eastern Utah and southwestern Mountains, north-central New Mexico: New Mexico Museum of Natu
Colorado; in Fouch, T.D. and Magathan, E.R., eds., Paleozoic paleo ral History and Science, Bulletin 31, p. 74-89.
geography of the west-central United States: Society of Paleontologists Kues, B.S. and Giles, K.A., 2004, The late Paleozoic ancestral Rocky Moun
and Mineralogists Rocky Mountain Section, Rocky Mountain Paleo tains system in New Mexico; in Mack, G.H. and Giles, K.A., eds., The
geography Symposium 1, p. 327-340. geology of New Mexico, a geologic history: New Mexico Geological
Eberth, D.A, and Berman, D.S., 1993, Stratigraphy, sedimentology and Society, Special Publication 11, p. 95-136.
vertebrate paleoecology of the Cutler Formation red beds (Pennsylva Lucas, S.G. and Krainer, K., 2005a, Cutler Group (Permo-Carboniferous)
nian-Permian) of north-central New Mexico: New Mexico Museum of stratigraphy, Chama Basin, New Mexico: New Mexico Museum of Natu
Natural History and Science, Bulletin 2, p. 33-47. ral History and Science, Bulletin 31, p. 90-100.
Eberth, D.A. and Miall, A.D., 1991, Stratigraphy, sedimentology and evolu Lucas, S.G. and Krainer, K., 2005b, Stratigraphy and correlation of the
tion of a vertebrate-bearing, braided to anastomosed fluvial system, Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New
Cutler Formation (Permian-Pennsylvanian), north-central New Mexico: Mexico Geological Society, Guidebook 56, p. 145-159.
Sedimentary Geology, v. 72, p. 225-252. Lucas, S.G., Krainer, K. and Colpitts, R.M., 2005, Abo-Yeso (Lower Per
Fracasso, M.A., 1987, Fluvial deposition and paleoenvironment of the mian) stratigraphy in central New Mexico: New Mexico Museum of
Cutler Formation red beds, El Cobre Canyon, New Mexico: New Mexico Natural History and Science, Bulletin 31, p. 101-117.
Geology, v. 9, p.14-25. Miall, A.D., 1978, Lithofacies types and vertical profile models in braided
Huffman, A.C. and Condon, S.M., 1993, Stratigraphy, structure, and paleo rivers: A summary; in Miall A.D., ed., Fluvial sedimentology: Canadian
geography of Pennsylvanian and Permian rocks, San Juan Basin and Society of Petroleum Geologists, Memoir 5, p. 597-604.
adjacent areas, Utah, Colorado, Arizona, and New Mexico: U.S. Geologi Miall, A.D., 1981, Analysis of fluvial depositional systems: AAPG Educa
cal Survey, Bulletin 1808, p. 1-44. tion Course Note Series 20, 75 p.
36
Miall, A. D., 1985, Architectural-element analysis: A new method of facies playa complex: The Middle Devonian Trentishoe Formation of North
analysis applied to fluvial deposits: Earth Science Reviews, v. 22, p. 261 Devon, UK: Sedimentology, v. 31, p. 697-716.
308. Williams, G. E., 1971, Flood deposits of the sandbed ephemeral streams of
Miall, A.D., 1992, Alluvial deposits; in Walker, R.G. and James, N.P., eds., central Australia: Sedimentology, v. 17, p. 1-40.
Facies models. Response to sea level change: Geological Association of Wood, G.H. and Northrop, S.A., 1946, Geology of the Nacimiento Moun
Canada, Special Publication, p. 119-142. tains, San Pedro Mountain, and adjacent plateaus in parts of Sandoval
Miall, A.D., 1996, The geology of fluvial deposits. Berlin, Springer, 582 p. and Rio Arriba Counties, New Mexico: U.S. Geological Survey, Oil and
Picard, M. D. and High, L.R., 1973, Sedimentary structures of ephemeral Gas Investigations Map OM-57.
streams: Developments in Sedimentology, no. 17, 223 p. Woodward, L., 1987, Geology and mineral resources of Sierra Nacimiento
Tunbridge, I. P., 1981, Sandy high-energy flood sedimentation – some and vicinity, New Mexico: New Mexico Bureau of Mines and Mineral
criteria for recognition, with an example from the Devonian of SW Resources, Memoir 42, 84 p.
England: Sedimentary Geology, v. 28, p. 79-96. Woodward, L., 1996, Paleotectonics of the late Paleozoic Peñasco uplift,
Tunbridge, I. P., 1984, Facies models for a sandy ephemeral stream and clay Nacimiento region, northern New Mexico: New Mexico Geological So
ciety, Guidebook 47, p. 107-113.

APPENDIX

Location of measured stratigraphic sections


All coordinates reported here are UTM’s in zone 13, datum NAD 27.

Cañon del Cobre: base of section at 378369E, 4016319N; top of section at 376985E, 4016278N.

Cañon del Cobre A: section at 382564E, 4021118N and immediate vicinity.

Gilman: base of segment A at 341134E, 3952490N, top at 341224E, 3552720N; base of segment B at 341320E, 3953055N,
top at 341432E, 3953124N.

Jemez River: base of section at 343754E, 3952554; top at 344598E, 3952583N.


Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

37
PENNSYLVANIANLAKE-MARGINICHNOASSEMBLAGE
FROM CAÑON DELCOBRE, RIO ARRIBA COUNTY, NEWMEXICO

SPENCER G. LUCAS AND ALLAN JLERNER


New Mexico Museum of Natural History and Science, 1801 Mountain Road N. W., Albuquerque, NM 87104

Abstract—In Cañon del Cobre of Rio Arriba County, northern New Mexico, red-bed siliciclastic sediments of the
Late Pennsylvanian-Early Permian Cutler Group were deposited by synorogenic fluvial systems in upland and
inland settings along the flank of the Uncompahgre uplift of the ancestral Rocky Mountains. Rare lacustrine strata
(thinly laminated siltstones) of the El Cobre Canyon Formation of the Cutler Group represent shallow, localized
floodplain lakes of the late Paleozoic Pangean tropics. We document the first ichnofossil assemblage known from
such an intermontane lake deposit, in strata of Late Pennsylvanian age. This ichnofossil assemblage consists of
arthropod trackways (Diplichnites x Diplopodichnus, Diplichnites gouldi, cf. Paleohelcura tridactyla, Protichnites
isp.), invertebrate grazing and feeding traces (Gordia indianensis and Helminthoidichnites tenuis), tetrapod foot
prints (Batrachichnus salamandroides and aff. Amphisauropus isp.) and fish swimming trails (Undichna britannica).
The numerically dominant ichnofossils are Helminthoidichnites and Diplichnites; the other ichnotaxa are repre
sented by few specimens. The Cañon del Cobre ichnofossil assemblage is readily assigned to the Scoyenia
ichnofacies and is composed of members of the Diplichnites and Mermia ichnoguilds, consistent with a shallow
lacustrine setting. The relatively low ichnodiversity of the Cañon del Cobre ichnofossil assemblage indicates a
short-lived water body colonized by a mobile invertebrate epifauna of detritus feeders and predators that provided
a food source for tetrapod predators.

INTRODUCTION
Cañon del Cobre (El Cobre Canyon) is a large box canyon in Rio
Arriba County, northern New Mexico (Fig. 1) that exposes extensive
bedrock badlands, slopes and escarpments developed in an ~400 m thick
section of Pennsylvanian-Permian Cutler Group siliciclastic red beds
(Smith et al., 1961; Eberth and Miall, 1991; Eberth and Berman, 1993;
Lucas and Krainer, 2005a, b). These strata are primarily of fluvial origin
divided into a lower, El Cobre Canyon Formation, and an upper, Arroyo
del Agua Formation. Fossils described from these strata, most of them
from the El Cobre Canyon Formation, include palynomorphs, megafossil
plants and vertebrates. Indeed, Cañon del Cobre is most famous
paleontologically for its vertebrate fossil assemblages that span the Penn
sylvanian-Permian transition (Vaughn, 1963; Fracasso, 1980; Lucas et
al., 2005a; Lucas, 2006).
The fluvial strata in Cañon del Cobre also contain ichnofossils,
mostly a low diversity assemblage (dominated by Palaeophycus) in chan
nel-margin deposits. One of these deposits also includes trackways of
the giant millipede-like Arthropleura, assigned to the ichnogenus
Diplichnites (Lucas et al., 2005a; Schneider et al., this volume). Also,
among the fluvial red beds of the El Cobre Canyon Formation in Cañon
del Cobre are some rare and localized lacustrine deposits that yield all the
fossil leaves known from these strata. One of these lake deposits (Figs.
2-3) also yields the ichnofossil assemblage described here. This assem
FIGURE 1. Location of Cañon del Cobre in northern New Mexico (after
blage is a rare example of a Carboniferous ichnoassemblage of a clastic
Ash, 1974).
lake margin in the ancestral Rocky Mountain basins of western North
America. extraformational clasts of quartzite, granite and gneiss. The El Cobre
Institutional abbreviation: NMMNH = New Mexico Museum Canyon Formation approximately corresponds to megasequence 1 of
of Natural History and Science, Albuquerque.
Eberth and Miall (1991).
GEOLOGY The ichnoassemblage described here is from NMMNH locality
5874 in the floor of Cañon del Cobre. This outcrop encompasses an
In Cañon del Cobre, a total thickness of about 230 m of El Cobre abandoned copper mine developed in a localized deposit that we inter
Canyon Formation is exposed, and it is mostly siltstone and sandstone. pret as representative of a clastic lake (Figs. 2-3). The section begins
Minor rock types are conglomerate/conglomeratic sandstone, shale and with a sandstone bench that is mostly laminated and contains an exten
calcrete. Siltstones of the El Cobre Canyon Formation contain numerous sive paleoflora on at least two bedding planes. Strata above this sand
rhizoliths, and sandstones are typically coarse grained, arkosic, mica stone bench are a 4.7-m-thick unit of thinly laminated purple and green
ceous, trough crossbedded and form multistoried bodies that erode to siltstone/shale that contains some isolated limestone (micrite) concre
thick cliffs and benches. Conglomerates characteristically have tions. A crossbedded sandstone bench caps this succession. The trace
38

FIGURE 3. Measured stratigraphic section at locality 5874.

FIGURE 2. Photographs of NMMNH locality 5874. A, Overview of the Ichnogenus Diplichnites Dawson, 1873
locality, which is just above a mine adit, and below a prominent, bench Diplichnites gouldi Gevers, in Gevers et al., 1971
forming crossbedded sandstone (see Fig. 3). The locality is at the base of a
Fig. 4B
slope formed by thinly laminated siltstone. B, The lowest bed, marked with
5874, is the trace-fossil-bearing bed, immediately below thinly-laminated Referred Specimens: NMMNH P-56250 (Fig. 4B), P-56262,
siltstone. Scale in cm. P-56269,
Characteristics: P-56250 shows a single trackway preserved in
fossils reported here (NMMNH locality 5874) occur in the upper few part and counterpart. Symmetrical track rows consist of closely-spaced
mm of the sandstone bench, just below the siltstone/shale (Figs. 2-3). tracks that are 1 mm or slightly less in width and rounded to elongate in
Deposition of the El Cobre Canyon Formation took place in an shape. Elongate tracks are oblique to the midline. The internal width
ephemeral braided stream environment (Eberth and Miall, 1991; Krainer between track rows is 3 mm, and the external width is 5 mm. There are no
and Lucas, this volume). Localized (less than 100 m on strike) deposits medial impressions. The preservation is too indistinct to accurately de
of thinly laminated siltstone/shale are rare in this depositional system, termine the number of tracks in a series. Trackway length is 236 mm, and
and the strata that encompass locality 5874 (Figs. 2-3) are interpreted by the track course is mostly straight with slight curves.
us as the deposits of a clastic lake because: (1) they have a limited extent Specimen P-56262 is preserved in part and counterpart. Sym
and lenticular geometry, pinching out laterally into overbank floodplain metrical track rows consist of closely-spaced, elongate tracks that are 1
mudrock; (2) the lithofacies of the deposit, especially the thinly lami mm wide or slightly less and oblique to the midline. The internal width
nated siltstone/shale interval, are those often associated with clastic lakes between track rows is 4 mm, and the external width is 7 mm. There are no
(e.g., Picard and High, 1981); and (3) the stratigraphic succession is from medial impressions, and the trackway is short and only 12 mm long.
laminar sandstone to thinly laminated siltstone/shale with micritic lime Indistinct undertracks of a similar trackway are also seen on the slab
stone nodules, readily interpreted as an upward-fining succession from surface.
lake margin sandstone to lake bottom siltstone/shale (e.g., Picard and Specimen P-56269 shows a single trackway preserved in positive
High, 1981). hyporelief. Symmetrical track rows consist of closely-spaced tracks that
are 1 mm wide or slightly less and rounded to elongate in shape. Elongate
ARTHROPOD TRACKWAYS
tracks are oblique to the midline, and the number of tracks in a series is
The arthropod trackway terminology that we use conforms to six. The internal width between track rows is 3 mm, and the external
that of Minter et al. (2007a), which was modified from that of Anderson width is 6 mm. The trackway length is 23 mm, and there are no medial
(1975), Trewin (1994), and Braddy (2001). impressions.
Discussion: Specimens are assigned to Diplichnites gouldi based
39

FIGURE 4. Diplichnites x Diplopodichnus (A, NMMNH P-56252) and Diplichnites gouldi (B, NMMNH P-56250) from locality 5874.

on their close similarity to the emended diagnosis of the ichnospecies by aged individuals of one type of animal.
al.,
and
Buatois
widths (e.g.,
exhibitedetdifferent
2007a).
morphology,
These
al. which
Trewinfew
(1998a). may
Some
Cañon and
Cobre
indicate
morphotypes
and del
McNamara, that of were
Smith
trackways
occurrences
a1995;they
relatively wide al.,
etare range
all
produced
Diplichnites 2003; by
ofMinter
ofgouldi
similar et
size
same-
external
have Diplichnites is generally thought to have been produced by myri
apods or arthropleurids (Weber and Braddy, 2004), although crustaceans
and euthycarcinoids are also possible track makers (Minteret al., 2007a).
D. gouldi ranges from the Ordovician to Permian and is found in a variety
of settings, including transitional terrestrial to subaqueous freshwater
40
environments (Buatois et al., 1998a). one end. This series shows three round tracks with 1 mm diameters that
This is the second occurrence of Diplichnites from Cañon del are arranged into a linear group. It also has a series width of 9 mm,
Cobre. An arthropleurid trackway assigned to D. cuithensis has previ measured from inner to outer track. It constitutes a set with the series
ously been reported (Hunt et al., 2004; Lucas et al., 2005b). This track that occurs on the end of the opposite track row. The external width
way was the first report of this ichnospecies from the American West between track rows is 30 mm, as measured by the single set. The internal
and is one of the largest examples of D. cuithensis known (see Schneider width is 15 mm. A medial impression is not seen. The series appear to be
et al., this volume). obliquely arranged with staggered symmetry, although both cannot be
confidently determined due to the limits of preservation. Preservation is
Diplichnites x Diplopodichnus in positive hyporelief.
Fig. 4A P-56641 consists of a part and counterpart that shows multiple
Referred specimen: NMMNH P-56252 (Fig. 4A). circular tracks distributed across the surface in a confused (“over-tracked”)
Characteristics: P-56252 is preserved in part and counterpart. arrangement. Linear groups of three tracks are recognizable within this
The slab surface shows three similar trackways with external widths overall distribution, but none can be placed confidently into a recogniz
between 5 and 7 mm. Each trackway shows two track rows that consist able trackway pattern. Typical width of these groups of three tracks is 9
of small rounded tracks that, in places, become continuous ridges in mm, which is the same as that seen for the series in P-56640. The
which individual tracks are no longer distinguishable. Two of the trackways remaining three specimens (P-56642, P-56643, P-56644) show surfaces
overcross and have mostly straight courses with some gentle curves. with a few circular tracks that are sometimes arranged in groups of three.
Both of these trackways have similar lengths of between 105 and 115 Single tracks in isolation and groups of two also occur. Medial impres-
mm. The third trackway shows a curved course, with one abrupt change sions are absent, and preservation is in positive hyporelief.
in the opposite direction. It is the longest trackway on the slab, with a Discussion: These specimens are poorly preserved and are most
length of 185 mm. The internal margin between inner track rows on all likely undertracks. As best seen on P-56640, the appearance of these
three trackways displays either ridges or furrows along some of the mostly circular tracks and their oblique arrangement into linear groups of
three is most typical of Paleohelcura tridactyla, to which these speci
course. Small grazing trails (cf. Helminthoidichnites) occur in association
with the trackways, together with impressions of sparse, probable plant mens are provisionally referred. There is one group of tracks on P-56640
fragments. that resembles Octopodichnus by showing a faintly impressed fourth
imprint.
Discussion: The trackways on this slab show characteristics that
intergrade between Diplichnites and Diplopodichnus Brady, 1947. Minter These tracks show some superficial resemblance to the circular
et al. (2007b) discussed the different trackway morphologies that result surface expression of vertical burrows such as Treptichnus or Arenicolites.
from the varying moisture content of “soft” substrates (soup, soft and However, a track preserved on P-56641 occurs on the very edge of part
firmgrounds), and made specific reference to the classic example of the and counterpart slabs, and examination of both slabs on-edge revealed no
intergradation of Diplichnites, Diplopodichnus and Dendroidichnites. indication of vertical burrowing beneath the surface imprints. Indeed,
Rather than arguing for synonomy Minter et al. recommended retention vertical burrows of any sort are absent within the Cañon del Cobre
of each ichnogenus when seen in isolation but not for when intergrada ichnoassemblage.
tion occurs with other ichnotaxa. They recommended that these Another indication that these specimens are locomotion traces
rather than burrows is the presence of a bifid track on P-56640.
intergrading examples should be regarded as hybrids. Accordingly, we
note P-56252 as a hybrid form (Diplichnites x Diplopodichnus), which Neoichnological experiments (Brady, 1947; Sadler, 1993) have shown
resulted from the track maker having traversed a moist substrate that that scorpions, which possess bifid extremities, are capable of producing
varied in its consistency. Paleohelcura trackways. While it is likely that the Cañon del Cobre
Diplopodichnus is a facies-crossing ichnotaxon that ranges in age specimens were produced by scorpionids or arachnids (Davis et al.,
from the Ordovician to Permian (Buatois et al., 1998a). Previous reports 2007), it is possible, too, that they were made by eurypterids.
of Diplopodichnus include those from a Carboniferous lacustrine facies Paleohelcura is similar to Palmichnium, which is generally attributed to
in Argentina (Aceñolaza and Buatois, 1991, 1993). Davis et al. (2007) eurypterid producers. The main feature distinguishing these two
have recently reported experimentally producing traces similar to ichnogenera is different gait patterns (Trewin and McNamara, 1995).
Diplopodichnus and Dendroidichnites in substrates approximating tran Unfortunately, the gait pattern of the referred Cañon del Cobre material
sitional subaqueous-subaerial conditions using an African giant black cannot be determined due to poor preservation.
millipede. Eurypterids and scorpions are known to have co-existed during
the middle to late Paleozoic (Selden, 1985). Late Carboniferous euryp
Ichnogenus Paleohelcura Gilmore, 1926 terids were mostly amphibious and largely represented by the genus
cf. Paleohelcura tridactyla Gilmore, 1926 Adelophthalmus (Selden and Nudds, 2004), a form typically found with
Fig. 5B freshwater and terrestrial animals and plants (Plotnick, 1997). Terrestrial
excursions by eurypterids were unlikely to be the result of food scaveng
Referred Specimens: NMMNH P-56640 (Fig. 5B), P-56641, ing, as they lacked suitable adaptations for terrestrial feeding (Braddy,
P-56642, P-56643, P-56644. 2004). A more likely, albeit speculative, explanation is that these excur
Characteristics: P-56640 shows a partial trackway through which sions were driven by environmental factors, such as the drying up of a
some of the slab surface has flaked off and left a few remaining series of formerly inhabited body of water (Donovan, 2001) or to mate (Braddy,
tracks contained in two track rows. The best preserved row contains 1999). A eurypterid prosoma assigned to Adelophthalmus sp. was dis
three complete series consisting of three tracks each. Two of these series covered at Cañon del Cobre from an areally restricted lacustrine deposit
are linear in arrangement, and one is triangular. There is a single track at (Hannibal et al., 2005). It comes from a succession that is stratigraphically
one end of the track row, which appears to have been part of an incom above locality 5874, establishing the presence of these animals in the
plete series that flaked off. The series widths are 8-9 mm, measured from Cañon del Cobre biota and indicating their potential as track makers.
inner to outer tracks. Inter-series distances are between 2 and 3 mm. The Late Carboniferous to Early Permian Carrizo Arroyo
Tracks are round, except for one in the triangular group, which is bifid. Lagerstätte in central New Mexico also yields a significant Adelophthalmus
The tracks are typically 1 mm in diameter. The stride is 9 mm, but pace record, due in part to the unusually large number of specimens (Kues and
cannot be accurately determined due to the incomplete preservation. The Kietzke, 1981). These eurypterids inhabited shallow ponds or enclosed
opposite track row shows a single series of tracks that are preserved at lagoons on the deltaic plain adjacent to the ancestral Cañon del Cobre
41

FIGURE 5. Protichnites ichnosp. (A, NMMNH P-56248) and cf. Paleohelcura tridactyla (B, NMMNH P-56640) from locality 5874.

area (Kues and Kietzke, 1981). The presumed leg span of the Carrizo Ichnogenus Protichnites Owen, 1852
Arroyo eurypterids during locomotion was about 100 mm (Briggs and
Protichnites isp.
Rolfe, 1983). In comparison, the partial trackway seen on specimen P Fig. 5A
56640 indicates a leg span of about 30 mm, which would indicate a
juvenile, if the Cañon del Cobre track producer were a eurypterid. Referred Specimen: NMMNH P-56248 (Fig. 5A).
Paleohelcura is primarily known from Permian eolian deposits in Characteristics: The specimen is preserved in part and counter
the American Southwest, including the Coconino Sandstone of Arizona part. The slab surfaces show a single trackway consisting of two rows of
(Gilmore, 1926, 1927; Brady, 1947; Brand and Kramer, 1996), the De variably-shaped tracks that are separated by a medial impression. One
Chelly Sandstone of Arizona (Sadler, 1993), and the Lyons Sandstone of track row on the part is missing approximately half of its length from the
Colorado (Toepelman and Rodeck, 1936). There is also a record from the slab surface. The counterpart track row that corresponds to the missing
eolian Jurassic Navajo Sandstone of Colorado (Faul and Roberts, 1951). section is obscured by a thin covering of matrix. It probably appears
There is a single prior Pennsylvanian record, which comes from a tidal there as a result of differential splitting that occurred along a plane of
flat setting in the McAlester Formation of Oklahoma (Lucas et al., 2004). weakness created by a medial impression between the track rows. The
There are no previous North American records of Paleohelcura from more complete, and therefore longer, track row shows one end that is
lacustrine settings. narrow, with a width that varies between 1 and 2 mm. The tracks vary in
42
size and are irregularly rounded to ellipsoidal in shape. The larger tracks
sembles Gordia indianaensis, which is characterized by having sharply
measure 2 mm at their longest axis, and the smaller tracks are 1 mm or
angled turns (Buatois et al., 1998b). Records of Gordia indianaensis
less. The tracks seen along this section of the row are generally disorga
have primarily come from Carboniferous lacustrine to marginal-marine
nized but sometimes form groups of two or three that are oblique to the
settings (Buatois et al., 1998b). There is a single Early Permian report
midline. Individual series cannot be determined. The track row gently
from a playa lake ichnoassemblage at Castle Peak, Texas (Minter et al.,
curves midway through the course and widens to 3 mm. The tracks along
this section are more regularly rounded and crowded into irregular group 2007a).
ings of 2-3 that are oblique to the midline. Ichnogenus Helminthoidichnites Fitch, 1850
The shorter, incomplete track row consists of tracks that are
Helminthoidichnites tenuis Fitch, 1850
similar to those seen in the same section of the opposite row. They are
Fig. 6A, C
irregularly arranged, although some occur in groups of two and three that
vary in their orientation to the midline. Individual series cannot be deter Referred specimens: NMMNH P-56251, P-56265, P-56271,
mined. The track row varies in width between 1and 2 mm, and it termi P-56273 (Fig. 6C), P-56274, P-56283 (Fig. 6A).
nates where the course is seen to curve. The external width between the Characteristics: Slab surfaces show multiple, unbranched, hori
two track rows averages 7 mm. The internal width averages 3 mm. The zontal burrows preserved in positive hyporelief. These simple burrows
length of the longer track row is 45 mm, and that of the shorter track row are thin and have diameters of 1 mm or less. Burrow courses are varied
is 22 mm. A centrally-placed medial impression is present in the portion from straight to curving, although circular courses are occasionally seen.
of the trackway where both track rows are seen. It is less than 1 mm Overcrossing among different specimens commonly occurs, although
wide, and a small amount of splay appears along its course. A secondary self-overcrossing does not. Burrow fill is structureless and identical to
impression branches from the centrally placed one at the point where the the surrounding slab matrix. Burrow lengths are relatively short, although
trackway course turns and the shorter track row terminates. This some rare examples extend to 180 mm.
what thinner, secondary impression is faintly impressed onto the cen Discussion: These simple grazing traces are readily assigned to
tralized one where the branching occurs. It is laterally offset to a high Helminthoidichnites tenuis based on their simple morphology, generally
degree and makes contact with some of the inner tracks in the longer track straight to curving courses, and lack of self-overcrossing. Since the early
row. It then becomes discontinuous and curves away from the track row 1990’s there has been an increase in reports of H. tenuis from lacustrine
and back towards the midline. settings (Metz, 1992, 1996; Pickerill, 1992; Buatois and Mángano, 1993;
Discussion: This short trackway is referred to Protichnites based Aceñolaza and Buatois, 1993; Buatois et al., 1996, 1997; Keighley and
on its simple rounded tracks shown in similar series on either side of a Pickerill, 1997, Minter et al., 2007a), although some of these are now
medial impression (after Keighley and Pickerill, 1998). The Cañon del regarded as Helminthopsis following a review of that ichnogenus by Han
Cobre specimen most resembles P. multinotatus Owen in its track ar and Pickerill (1995).
rangement, but the material is not well enough preserved to make an
ichnospecific assignment, so it is left in open nomenclature. The trace TETRAPOD TRACES
producer’s terminal appendage may have had to have been lifted when it
was turned, as indicated by the secondary medial impression. Walker Ichnogenus Batrachichnus Woodworth, 1900
(1985) referred to this as “offsetting,” which reflects a degree of lateral Batrachichnus salamandroides Geinitz, 1861
inflexibility in the tail. Owen (1852) similarly noted in his original de Fig. 7A
scription of P. multinotatus that the medial impression was formed by an Referred Specimens: NMMNH P-56261, P-56284, P-56635,
appendage that continued to incline to one side after the body had begun P-56636 (Fig. 7A).
to turn in another direction. Characteristics: Slab surfaces show similar, small, tracks of a
Possible producers of Protichnites in terrestrial freshwater envi quadruped that are preserved in positive hyporelief. P-56636 is the best
ronments include small crustaceans or euthycarcinoids (Savage, 1971; preserved specimen and has two track sets, each containing a manus and
Smith et al., 2003). North American Carboniferous records of Protichnites pes. One set has a semi-plantigrade manus with four thin, curved digits
are primarily from the Mabou and Cumberland groups of eastern Canada, that increase in length from digit I to III, with digit IV being the longest.
where it occurs in various floodplain-sheetflood depositional settings The length of this manus is 10 mm, and the width is 8 mm. The accom
including ephemeral ponds (Keighley and Pickerill, 1998, 2003).
panying pes slightly oversteps the manus’ posterior and is slightly far
Protichnites has the distinction of having been the first fossilized arthro ther from the midline. This pes is plantigrade and somewhat deeply
pod trackway that was given a Linnaean name and formal description
(MacNaughton and Hagadorn, 2006). This is the first record of impressed, which obscures other details. The second track set show a
manus that is semi-plantigrade and contains four, thin, straight digits that
Protichnites from New Mexico.
increase in length from digits I to IV, with digit IV being the longest. The
GRAZING AND FEEDING TRACES length of this manus is 10 mm, and the width is 7 mm. The accompanying
pes is set closely posterior to the manus. It is semi-plantigrade, and its
Ichnogenus Gordia Emmons, 1844 details are somewhat obscured. It is slightly out of line with the accom
Gordia indianaensis Miller, 1889 panying manus, being closer to the midline. The outer width between
Fig. 6B opposite, corresponding tracks is 27 mm. There is no medial impression
between the track sets.
Referred Specimen: NMMNH P-56268 (Fig. 6B). Discussion: Batrachichnus is a commonly occurring Carbonifer
Characteristics: This single specimen consists of a horizontal, ous-Early Permian track that is attributed to small temnospondyl am
unbranched, burrow with a diameter of 1 mm. The burrow fill is struc phibian producers. The wide range of extramorphological variation that
tureless and identical to the surrounding slab matrix. The burrow course occurs in the preservation of Batrachichnus tracks has been well docu
is mostly straight but contains a tight turn that over-crosses the straight mented (e.g., Haubold, 1996, 2000; Melchor and Sarjeant, 2004; Lucas,
segment. Total burrow length is 80 mm, and preservation is in part and 2005). The elongate digit impressions seen in the Cañon del Cobre speci
counterpart. mens are characteristic of a type of extramorphologic variation referred
Discussion: This specimen displays self-overcrossing, so we sepa to as “Gracilichnium” preservation. This morphotype is typically asso
rate it from the similar burrow Helminthoidichnites, which also appears ciated with depositional environments that had substrates with high
in the Cañon del Cobre ichnofauna. This specimen most closely re moisture content (e.g., tidal flats, fluvial and lacustrine settings).
43

FIGURE 6. Helminthoidichnites tenuis (A, NMMNH P-56283, C, NMMNH P-56273) and Gordia indianensis (B, NMMNH P-56268) from locality
5874.

Ichnogenus Amphisauropus Haubold, 1970 posterior to the manus. It is 30 mm wide and 20 mm long. The second
aff. Amphisauropus isp. manus also shows a rounded posterior that is 30 mm wide. The digits are
Fig. 7B not distinguishable; although three anterior drag marks are seen. The
longest one is 40 mm long and is on line with the outside margin of the
Referred Specimens: NMMNH P-56637 (Fig. 7B). track. Directly posterior to this manus is the faint outline of an undertrack
Characteristics: The slab shows two manus tracks, preserved in that is probably from the accompanying pes. The external width be
negative epirelief, and a partial pes track, preserved in positive epirelief. tween both pes tracks is 150 mm.
One manus is plantigrade, with a broad, rounded posterior that is 30 mm Discussion: The best preserved tracks on this slab show a broad,
wide. This track shows three digits that increase in length from digits I to inwardly directed manus and an apparently forward directed pes, which
III; with digit I being 13 mm, digit II being 18 mm, and digit III being 30 are characteristic of Amphisauropus (Haubold, 1970, 1971). We cau
mm long. These digits are tapered and directed inward, and may partially tiously leave this single specimen in open nomenclature due to its rela
include drag marks in their length. The overall length of this track is 70 tively poor preservation.
mm. A single, partially preserved pes with a rounded heel is seen directly Amphisauropus is a commonly occurring and widely distributed
44

FIGURE 7. Batrachichnus salamandroides (A, NMMNH P-56636), aff. Amphisauropus ichnosp. (B, NMMNH P-56637) and Undichna britannica (C,
NMMNH P-56263) from locality 5874.

track in the Early Permian interval of the European Rotliegend (Lucas et Characteristics: P-56254 is a slab that shows several sets of
al., 2001; Voigt, 2005). The first United States record of Amphisauropus thin, paired trails on its surface. Preservation is in positive hyporelief.
came from sheetflood deposits in the Early Permian Abo Formation of Two sets of trails are relatively well preserved compared to the others.
central New Mexico (Lerner et al., 2001; Lucas et al., 2001, 2009). There The first set shows a pair of thin trails with similar courses, each of
have since followed other U.S. reports, all of which are from the Early which consists of a single wave-like curve. The longer trail crosses over
Permian (Wolfcampian-Leonardian). These reports are from the the shorter one once. The trails are of different lengths, with the longer
Hennessey Formation of central Oklahoma (Lucas and Suneson, 2002), being 70 mm and the shorter being 50 mm. The trails are separated by a
the Clear Fork Formation of north-central Texas (Lucas et al., 2003), the distance of 6 mm at their farthest extent. The second set shows a pair of
trails with similar courses, each of which consists of a single wave-like
Yeso Group of central New Mexico (Smith et al. 2005; Lucas et al.,
2005c) and the Choza Formation of north-central Texas (Minter et al., curve. They overcross once midway through their course. The trails are
2007a). Amphisauropus has also been reported from Early Permian red of equal length, which is 65 mm. The trails are separated by a distance of
bed sequences in eastern Canada (Mossman and Place, 1989; Van Allen 16 mm at their farthest extent. The wavelengths and amplitudes of both
and Calder, 2003; Van Allen et al., 2005). The presumed producers of sets of trails cannot accurately be determined because each set abruptly
Amphisauropus were seymouriamorph amphibians. ends at the edge of the slab, leaving these traces partially preserved. The
slab also contains two similar sets of thin, paired trails that are of rela
Ichnogenus Undichna Anderson, 1976 tively short length, which is about 20 mm. The courses of these trails are
Undichna britannica Higgs, 1988 gently curved and roughly parallel, and they are separated by a distance
Fig. 7C of 6 mm at their greatest extant.
Discussion: These simple, intertwined sinusoidal trails are as
Referred Specimens: NMMNH P-56263 (Fig. 7C), P-56275. signed to Undichna britannica. They are somewhat similar in appear
45
ance to some of the invertebrate feeding burrows commonly seen at this et al. (1998c). Thus, the Diplichnites ichnoguild is represented by surface
locality. The Undichna traces, however, appear more sharply incised to shallow subsurface walking traces of arthropod predators and detritus
and fluidly produced than the invertebrate traces, which lends support feeders (Diplichnites, Diplopodichnus, Paleohelcura, Protichnites). The
to their interpretation as locomotion trails produced by fishes. The sev Mermia ichnoguild comprises surface to shallow subsurface traces of
eral pairs of trails seen on P-56255 are not aligned to any one direction detritus feeders (Helminthoidichnites, Gordia). One (Diplichnites
along the slab surface, which might otherwise suggest that they were tool ichnoguild) typically represents lake margins, whereas the other (Mermia
marks produced by currents. ichnoguild) represents lake bottoms, and their co-occurrence in a clastic
A review of Undichna, with examples from the Early Permian of lake margin is not surprising. Both ichnoguilds are represented by transi
New Mexico, was recently provided by Minter and Braddy (2006b). tory structures made by mobile predators and detritus feeders.
The majority of Undichna occurrences are from the Carboniferous, in Ichnofossil assemblages of clastic freshwater lakes of Carbonifer
cluding several from various lacustrine depositional settings (Walter, ous age are not common (Parker and Gierlowski-Kordesch, 2007), and
1982; Turek, 1989, 1996; Melchor et al., 1993; Buatois and Mángano, we know of no such record from the ancestral Rocky Mountains of the
1993, 1994; Keighley and Pickerill, 1997; Higgs, 1998; Cameron et al., western United States. The Cañon del Cobre ichnofossil assemblage thus
1998; Pazos, 2002; Netto et al., 2003). adds to a limited fossil record, and shows successful colonization of a
clastic lake margin in a relatively inland and upland setting of the ances
DISCUSSION tral Rocky Mountain foreland. The presence of mobile detritus feeders
The ichnofossil assemblage described here consists of arthropod and arthropod predators confirms a pattern already well documented in
trackways (Diplichnites, Diplopodichnus, cf. Paleohelcura, Protichnites), other Carboniferous terrestrial ichnofaunas (Buatois et al., 1998c; Parker
grazing and feeding traces (Gordia, Helminthoidichnites), tetrapod tracks and Gierlowski-Kordesch, 2007; Buatois and Mángano, 2007). The Cañon
(Batrachichnus, aff. Amphisauropus) and fish-swimming trails del Cobre ichnofossil assemblage thus provides evidence of a community
(Undichna). Arthropod walking traces and grazing/feeding traces domi of detritus feeding and predatory arthropods, likely preyed upon by
nate the assemblage; all other kinds of traces are rare, being represented tetrapod predators, on the Late Pennsylvanian braided river plains of the
by one or a few specimens. This assemblage is readily assigned to the ancestral Rocky Mountain foreland.
Scoyenia ichnofacies (cf. Buatois and Mángano, 2007), though note that ACKNOWLEDGMENTS
it lacks horizontal, meniscate back-filled traces.
The Cañon del Cobre ichnofossil assemblage consists of ichnotaxa Dan Chaney and Josh Smith provided valuable field assistance
characteristic of both the Diplichnites and Mermia ichnoguilds of Buatois collecting NMMNH locality 5874. Elizabeth Gierlowski-Kordesch, Joerg
Schneider and Nicholas J. Minter reviewed the manuscript.

REFERENCES

Aceñolaza, F. G. and Buatois, L. A., 1991, Tazas fósiles de la Formación Buatois, L. A. and Mángano, M. G., 1994, Pistas de peces en el Carbonifero
Patquía en el Bordo Atravesado, Sierra de Famatina, La Rioja: Acta de la cuenca Paganzo (Argentina): Su significado estragráfico y
Geológica Lilloana, v. 15, p. 19-29. paleoambiental: Ameghiniana, v. 31, p. 33-40.
Aceñolaza, F. G. and Buatois, L. A. 1993, Nonmarine perigondwanic trace Buatois, L. A. and Mángano, M. G., 2007, Invertebrate ichnology of conti
fossils from the late Paleozoic of Argentina: Ichnos, v. 2, p. 183-201. nental freshwater environments; in Miller, W., III, ed., Trace fossils:
Anderson, A. M., 1975, The ‘trilobite’ trackways in the Table Mountain Concepts, problems, prospects: Amsterdam, Elsevier, p. 285-323.
Group (Ordovician) of South Africa: Palaeontologia Africana, v. 18, p. Buatois, L. A., Jalfin, G. and Aceñolaza, F. G., 1997, Permian nonmarine
35-45. invertebrate trace fossils from the southern Patagonia, Argentina:
Anderson, A. M., 1976, Fish trails from the Early Permian of South Africa: Ichnologic signatures of substrate consolidation and colonization se
Palaeontology, v. 19, p. 397-409. quences: Journal of Paleontology, v. 71, p. 324-336.
Ash, S.R., 1974, Upper Triassic plants of Cañon del Cobre, New Mexico: Buatois, L. A., Mángano, M. G., Maples, C. G. and Lanier, W. P., 1998a,
New Mexico Geological Society, Guidebook 25, p. 179-184. Taxonomic reassessment of the ichnogenus Beaconichnus and addi
Braddy, S. J., 1995, A new arthropod trackway and associated invertebrate tional examples from the Carboniferous of Kansas, U.S.A.: Ichnos, v. 5,
ichnofauna from the Lower Permian Hueco Formation of the Robledo p. 287-302.
Mountains, southern New Mexico: New Mexico Museum of Natural Buatois, L. A., Mángano, M. G., Maples, C. G. and Lanier, W. P., 1998b.
History and Science, Bulletin 6, p. 101-105. Ichnology of an Upper Carboniferous fluvio-estuarine paleovalley: The
Braddy, S. J., 1999. Eurypterid palaeoecology: Palaeobiological, ichnological Tonganoxie Sandstone, Buildex Quarry, Eastern Kansas, USA: Journal
and comparative evidence for a ‘mass-moult-mate’ hypothesis: of Paleontology, v. 72, p. 152-180.
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 172, p. 115 Buatois, L. A., Mángano, M. G., Wu, X. and Zhang, G., 1996, Trace fossils
132. from Jurassic lacustrine turbites of the Anyao Formation (central China)
Braddy, S. J., 2001, Trackways-arthropod locomotion; in Briggs, D.E.G. and their environmental and evolutionary significance: Ichnos, v.4, p.
and Crowther, P. R., eds., Palaeobiology II, p. 389-393. 287-303.
Braddy, S. J., 2004, Ichnological evidence for the arthropod invasion of Buatois, LA., Mángano, M. G., Genise, J. F. and Taylor, T. N., 1998c, The
land: Fossils and Strata, v. 51, p. 136-140. ichnologic record of the invertebrate invasion of nonmarine ecosys
Brady, L. F., 1947, Invertebrate tracks from the Coconino Sandstone of tems: Evolutionary trends in ecospace utilization, environmental ex
northern Arizona: Journal of Paleontology, v. 21, p. 466-472. pansion, and behavioral complexity: Palaios, v. 13, p. 217-240.
Brand, L. R. and Kramer, J., 1996, Underprints of vertebrate trackways in Cameron, B., Wood, D. and van Dommelen, R., 1998, Vertebrate and inver
the Permian Coconino Sandstone in Arizona: Ichnos, v. 4, p. 225-230. tebrate trace fossils from the Horton Bluff Formation (Lower Carbonif
Briggs, D. E. G., and Rolfe, I., 1983, A giant arthropod trackway from the erous) of Nova Scotia: Mining Matters for Nova Scotia, v. 98, p. 6.
Lower Mississippian of Pennsylvania: Journal of Paleontology, v. 57, p. Davis, R. B., Minter, N. J. and Braddy, S. J., 2007. The neoichnology of
377-390. terrestrial arthropods: Palaeogeography, Palaeoclimatology, Palaeo
Buatois, L. A. and Mángano, M. G., 1993, Trace fossils from a Carbonifer ecology, v. 255, p. 284-307.
ous turbidtic lake: Implications for the recognition of additional nonma Dawson, J. W., 1873, Impressions and footprints of aquatic animals and
rine ichnofacies: Ichnos, v. 2, p. 237-258. imitative markings on Carboniferous rocks: American Journal of Sci
46
ence and Arts, v. 105, p. 16-24. 83-112.
Donovan, S. K., 2001, Fossils explained 37, eurypterids: Geology Today, v. Keighley, D. G. and Pickerill, R. K., 2003, Ichnocoenoses from the Carbon
17, p. 195-199. iferous of eastern Canada and their implications for the recognition of
Eberth, D. S. and Berman D. S, 1993, Stratigraphy, sedimentology and ichnofacies in nonmarine strata: Atlantic Geology, v. 39, p. 1-22.
vertebrate paleontology of the Cutler Formation redbeds (Pennsylva Kues, B. S. and Kietzke, K., 1981, A large assemblage of a new eurypterid
nian-Permian) of north-central New Mexico: New Mexico Museum of from the Red Tanks Member, Madera Formation (Late Pennsylvanian
Natural History and Science, Bulletin 2, p. 33-48. Early Permian) of New Mexico: Journal of Paleontology, v. 55, p. 709
Eberth, D. A., and Miall, A. D., 1991, Stratigraphy, sedimentology and 729.
vertebrate paleoecology of the Cutler Formation redbeds (Pennsylva Lerner, A. J, Lucas, S. G. and Haubold, H., 2001, The footprint ichnogenera
nian-Permian) of north-central New Mexico: Sedimentary Geology, v. Amphisauropus and Varanopus in the Lower Permian of New Mexico:
72, p. 225-252. Journal of Vertebrate Paleontology, v. 21, no. 3, p. 72A.
Emmons, E., 1844, The Taconic System: Based on observations in New Lucas, S. G., 2005, Tetrapod ichnofacies and ichnotaxonomy: Quo vadis?
York, Massachusetts, Maine, Vermont and Rhode Island. Caroll and Ichnos, v. 12, p. 1-6
Cook, Albany, 68 p. Lucas, S. G., 2006, Global Permian tetrapod biostratigraphy and
Faul, H., and Roberts, W. A., 1951, New fossil footprints from the Na biochronology; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds.,
vajo(?) Sandstone of Colorado: Journal of Paleontology, v. 25, p. 266 Non-marine Permian biostratigraphy and biochronology. Geological
274. Society, London, Special Publication 265, p. 65-93.
Fitch, A., 1850, A historical, topographical and agricultural survey of the Lucas, S. G. and Krainer, K., 2005a, Stratigraphy and correlation of the
County of Washington. Part 2-5: Transactions of the New York Agri Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New
cultural Society, v. 9, p. 753-944. Mexico Geological Society, Guidebook 56, p. 145-159.
Eberth, D. S. and Miall, A. D., 1991, Stratigraphy, sedimentology and evo Lucas, S. G. and Krainer, K., 2005b, Cutler Group (Permo-Carboniferous)
lution of a vertebrate-bearing, braided to anastomosed fluvial system, stratigraphy, Chama basin, New Mexico: New Mexico Museum of Natu
Cutler Formation (Permian-Pennsylvanian), north-central New Mexico: ral History and Science, Bulletin 31, p. 90-100.
Sedimentary Geology, v. 72, p. 225-252. Lucas, S. G. and Suneson, N. H., 2002, Amphibian and reptile tracks from
Fracasso, M. A., 1980, Age of the Permo-Carboniferous Cutler Formation the Hennessey Formation (Leonardian, Permian), Oklahoma County,
vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale Oklahoma: Oklahoma Geology Notes, v. 62, no. 2., p. 56-62.
ontology, v. 54, p. 1237-1244. Lucas, S. G., Lerner, A. J, Brunner, M. and Shipman, P., 2004, Middle
Geinitz, H. B., 1861. Dyas, I, 130 p. Pennsylvanian ichnofauna from eastern Oklahoma, USA: Ichnos, v. 11,
Gevers, T. W., Frakes, L. A., Edwards, N. and Marzolf, J. E., 1971, Trace p. 45-55.
fossils in the Lower Beacon sediments (Devonian), Darwin Mountains, Lucas, S. G., Lerner, A. J, Hannibal, J. T., Hunt, A. P. and Schneider, J. W.,
southern Victoria Land, Antarctica: Journal of Paleontology, v. 45, p. 2005a, Trackway of a giant Arthropleura from the Upper Pennsylva
81-94. nian of El Cobre Canyon, New Mexico: New Mexico Geological Society,
Gilmore, C. W., 1926, Fossil footprints from the Grand Canyon: Smithsonian Guidebook 56, p. 279-282.
Miscellaneous Collections, v. 77, p. 1-41. Lucas, S. G., Minter, N. J., Spielmann, J. A., Hunt, A. P. and Braddy, S.J.,
Gilmore, C. W., 1927, Fossil footprints from the Grand Canyon: Smithsonian 2005b, Early Permian ichnofossil assemblage from the Fra Cristobal
Miscellaneous Collections, v. 80, p. 1-78. Mountains, southern New Mexico: New Mexico Museum of Natural
Han, Y. and Pickerill, R. K., 1995, Taxonomic review of the ichnogenus History and Science, Bulletin 31, p. 140-150.
Helminthopsis Heer 1877 with a statistical analysis of selected Lucas, S. G., Lerner, A. J, and Haubold, H., 2001, First record of
ichnospecies: Ichnos, v. 4, p. 83-118. Amphisauropus and Varanopus in the Lower Permian Abo Formation,
Hannibal, J. T., Lucas, S. G., Lerner, A. J and Chaney, D. S., 2005, An central New Mexico: Hallesches Jarhbuch für Geowissenschaften, v. B23,
eurypterid (Adelophthalmus sp.) from a plant-rich lacustrine facies of p. 69-78.
Upper Pennsylvanian strata in El Cobre Canyon, New Mexico: New Lucas, S. G., Spielmann, J. A. and Lerner, A. J, 2009, The Abo Pass tracksite:
Mexico Museum of Natural History and Science, Bulletin 31, p. 34-38. A Lower Permian tetrapod footprint assemblage from central New
Haubold, H., 1970, Vesuch einer Revision der Amphibien-Fährten de Karbon Mexico: New Mexico Geological Society, Guidebook 60, p. 285-290.
und Perm: Freiberger Forschungs Hefte, C 260, p. 83-117. Lucas, S. G., Lerner, A. J, Nelson, Hunt, A. P., and DiMichele, W. A., 2003.
Haubold, H., 1971, Ichnia amphibiorum et reptiliorum fossilium: Encyclo Early Permian (Leonardian) tetrapod footprints from Lake Kemp, north
pedia of Paleoherpetology, v. 18, 124 p. central Texas: Geological Society of America, Abstracts with Programs,
Haubold, H., 1996, Ichnotaxonomie und Klassifikation von v. 35, no. 6, p. 499.
Tetrapodenfährten aus dem Perm: Hallesches Jahrbuch Lucas, S. G., Smith, J. A., and Hunt, A. P., 2005c. Tetrapod tracks from the
Geowissenschaften, v. B18 p. 23-88. Lower Permian Yeso Group: New Mexico Museum of Natural History
Haubold, H., 2000, Tetrapodenfährten aus dem Perm-Kenntnisstand und and Science, Bulletin 31, p. 119-124.
Progress 2000: Hallesches Jahrbuch Geowissenschaften, v. B22 p. 1-16. MacNaughton, R. B., and Hagadorn, J. W., 2006. Report on plaster casts of
Higgs, R., 1988, Fish trails in the Upper Carboniferous of south-west En arthropod-produced trace fossils (Protichnites) figured in W.E. Logan’s
gland: Palaeontology, v. 31, 255-272. Geology of Canada (1863), and recently copied from material in the
Hunt, A. P., Lucas, S. G. and Huber, P., 1990, Early Permian footprint fauna Amherst College Museum of Natural History, Amherst, MA: Report
from the Sangre de Cristo Formation of northeastern New Mexico: New 001-RBM-2006/1, Geological Survey of Canada (Calgary).
Mexico Geological Society, Guidebook 41, p. 291-303. Melchor, R. N. and Sarjeant, W. A. S., 2004, Small amphibian and reptile
Hunt, A. P., Lucas, S. G., Lerner, A. J and Hannibal, J. T., 2004, The giant footprints from the Permian Carapacha basin, Argentina: Ichnos, v. 11,
Arthropleura trackway Diplichnites cuithensis from the Cutler Group p. 57-78.
(Upper Pennsylvanian) of New Mexico: Geological Society of America, Melchor, R. N., Bazán, J., and Fernández, M. A., 1993. Asociación de trazas
Abstracts with Programs, v. 36, no. 5, p. 66. fósiles de la facies peliticas de la Formación Agua Escondida (Carbonifero
Keighley, D. G. and Pickerill, R. K., 1997, Systematic ichnology of the Superior?), sureste de Mendoza, Argentina: Primera Reunión Argentina
Mabou and Cumberland groups (Carboniferous) of western Cape Breton de Icnologia, Resúmenes y Conferencias Invitadas, p. 15.
Island, eastern Canada, 1: Burrows, pits, trails, and coprolites: Atlantic Metz, R., 1992. Trace fossils from the Lower Jurassic nonmarine Towaco
Geology, v. 33, p. 181-215. Formation, New Jersey: Northeastern Geology, v. 14, p. 29-34.
Keighley, D. G. and Pickerill, R. K., 1998, Systematic ichnology of the Metz. R, 1996. Newark Basin ichnology: the Late Triassic Perkasie Mem
Mabou and Cumberland groups (Carboniferous) of western Cape Breton ber of the Passaic Formation, Sanatoga, Pennsylvania: Northeastern
Island, eastern Canada, 2: Surface markings: Atlantic Geology, v. 34, p. Geology and Environmental Sciences, v. 18, p. 118-129.
47
Miller, S.A., 1889. North American geology and palaeontology for the use Smith, A., Braddy, S.J., Marriott, S. B., and Briggs, D. E. G., 2003. Arthro
of amateurs, students and scientists. Western Methodist Book Concern, pod trackways from the Early Devonian of South Wales: a functional
Cincinnati, Ohio. 664 p. analysis of producers and their behaviour: Geological Magazine, v. 140,
Minter, N.J., and Braddy, S.J., 2006a. Walking and jumping with Palaeozoic p. 63-72.
apterygote insects: Palaeontology, v. 49, p. 827-835. Smith, C. T., Budding, A. J. and Pitrat, C. W., 1961, Geology of the south
Minter, N. J., and Braddy, S. J., 2006b. The fish and amphibian swimming eastern part of the Chama Basin: New Mexico Bureau of Mines and
traces Undichna and Lunichnium, with examples from the Lower Per Mineral Resources, Bulletin 75, 57 p.
mian of New Mexico, USA: Palaeontology, v. 49, p. 1123-1142. Smith, J., Lucas, S. G., Hunt, A. P., and Schneider, J., 2005. First report of
Minter, N. J., Braddy, S. J., and Davis, R. B., 2007b. Between a rock and a tetrapod tracks from the Meseta Blanca Member of the Lower Permian
hard place: Arthropod trackways and ichnotaxonomy: Lethaia, v. 40, p. Yeso Formation, central New Mexico: Journal of Vertebrate Paleontol
365-375. ogy, v. 25, no. 3, p. 116.
Minter, N. J., Krainer, K., Lucas, S. G., Braddy, S.J., and Hunt, A. P., 2007a. Toepelman, W. C., and Rodeck, H. G., 1936. Footprints in late Paleozoic
Palaeoecology of an Early Permian playa lake trace fossil assemblage red beds near Boulder, Colorado: Journal of Paleontology, v. 10, p. 660
from Castle Peak, Texas, USA: Palaeogeography, Palaeoclimatology, 662.
Palaeoecology, v. 246, p. 390-423. Trewin, N. H., 1994. A draft system for the identification and description of
Mossman, D. J., and Place, C. H., 1989, Early Permian fossil vertebrate arthropod trackways: Palaeontology, v. 37, p. 811-823.
footprints and their stratigraphic setting in megacycle II red beds, Prim Trewin, N. H., and McNamara, K. J., 1995. Arthropods invade the land:
Point, Prince Edward Island: Canadian Journal of Earth Science, v. 26, p. Trace fossils and palaeoenvironments of the Tumblagooda Sandstone
591-605. (?late Silurian) of Kalbarri, Western Australia: Transactions of the Royal
Netto, R. G., Balistieri, P., Paz, C. P., and Lavina, E., 2003. Upper Carbon Society of Edinburgh: Earth Sciences, v. 85, p. 177-210.
iferous-Lower Permian rhythmites from Itararé Group (Paraná Basin, Turek, V., 1989. Fish and amphibian trace fossils from Westphalian sedi
Brazil): Are they all varvites?: III Simposio Argentino del Paleozoico ments of Bohemia: Palaeontology, v. 32, p. 623-643.
Superior y Reunión Anual de Grupo de Trabajo Argentino del Proyecto Turek, V., 1996. Fish trace fossil interpreted as a food gathering swimming
IGGP 471, La Plata, p. 4. trail from the Upper Carboniferous (Westphalian) of Bohemia: Casopis
Owen, R., 1852. Description of the impressions and footprints of the Narodniho Muzea Rada Prirodovedna, v. 165, p. 5-8.
Protichnites from the Potsdam sandstone of Canada: Quarterly Journal Van Allen, H. E. K., and Calder, J. H., 2003. Brule; the earliest record of
of the Geological Society [London], v. 8, p. 214-225. gregarious tetrapod behaviour in the world’s only known walchian coni
Parker, L. E. and Gierlowski-Kordesch, E. H., 2007, Paleozoic lake faunas: fer fossil forest: Geological Society of America, Abstracts with Pro
Establishing aquatic life on land: Paleogeography, Palaeoclimatology, grams, v. 35, no. 3, p. 25.
Palaeoecology, v. 249, p. 160-179. Van Allen, H. E. K., Calder, J. H. and Hunt, A. P., 2005, The trackway record
Pazos, P. J., 2002. Palaeoenvironmental framework of the glacial-postgla of a tetrapod community in a walchian conifer forest from the Permo
cial transition (Late Paleozoic) in the Paganzo-Calingasta Basin (south Carboniferous of Nova Scotia: New Mexico Museum of Natural History
ern South America) and the Great Karoo-Kalahari Basin (southern Af and Science, Bulletin 30, p. 322-332.
rica): Ichnological implications: Gondwana Research, v. 5, p. 619-640. Vaughn, P. P., 1963, The age and locality of the Paleozoic vertebrates from
Picard, M. D. and High, L. R., Jr., 1981, Physical stratigraphy of ancient El Cobre Canyon, Rio Arriba County, New Mexico: Journal of Paleon
lacustrine deposits: SEPM Special Publication, no. 31, p. 233-259. tology, v. 37, p. 283-296.
Pickerill, R. K., 1992. Carboniferous nonmarine invertebrate ichnocoenoses Voigt, S., 2005. Die Tetrapodenichnofauna des kontinentalen Oberkarbon
from southern New Brunswick, eastern Canada: Ichnos, v. 2, p. 21-35. und Perm im Thüringer Wald – Ichnotaxonomie, Paläoökologie und
Plotnick, R. E., 1997. Eurypterida; in Shabica, C., and Hay, A., eds., Biostratigraphie. Cuvillier Verlag, Göttingen, Germany, 179 p.
Richardson’s Guide to the Fossil Fauna of Mazon Creek, p. 1-308. Walker, E. F., 1985. Arthropod ichnofauna of the Old Red Sandstone at
Pollard, J. E., and Walker, E., 1984. Reassessment of sediments and trace Dunure and Montrose, Scotland: Transactions of the Royal Society of
fossils from the Old Red Sandstone (Lower Devonia) of Dunure, Scot Edinburgh: Earth Sciences, v. 76, p. 287-297.
land, described by John Smith (1909): Geobios, v. 17, p. 567-576. Walter, H., 1982. Neue Arthropodenfährthen aus den Oberhöfer Schichten
Sadler, C. J., 1993. Arthropod trace fossils from the Permian De Chelly (Rotliegendes, Thüringer Wald) mit Bemerkungen über Ichnia limnisch
Sandstone, northeastern Arizona: Journal of Paleontology, v. 67, p. terrestricher Tuffite innerhalb der varistischen Molasse: Freiberger
240-249. Forschungschefte, C 382, p. 87-100.
Savage, N. M., 1971. A varvite ichnocoenosis from the Dwyka Series of Weber, B, and Braddy, S. J., 2004. A marginal marine ichnofauna from the
Natal: Lethaia, v. 4, p. 217-233. Blaiklock Glacier Group (?Lower Ordovician) of the Shackleton Range,
Selden, P. A., 1985. Eurypterid respiration: Philosophical Transactions of Antarctica: Transactions of the Royal Society of Edinburgh: Earth Sci
the Royal Society, London, B, v. 309, p. 219-226. ences, v. 94, p. 1-20.
Selden, P. A. and Nudds, J., 2004. Evolution of fossil ecosystems: Manson Woodworth, J. B., 1900. Vertebrate footprints on Carboniferous shales of
Publishing Ltd, p. 1-160. Plainville, Massachusetts: Bulletin of the Geological Society of America,
v. 11, p. 449-454.
48

|
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

49
EURAMERICAN LATE PENNSYLVANIAN / EARLY PERMIAN
ARTHROPLEURID/TETRAPOD ASSOCIATIONS – IMPLICATIONS FOR THE
HABITAT AND PALEOBIOLOGY OF THE LARGEST TERRESTRIALARTHROPOD

JÖRGW. SCHNEIDER1, SPENCERG. LUCAS2, RALFWERNEBURG3 AND RONNYRÖßLER4


1TU Bergakademie Freiberg, Institut für Geologie, B.v.Cotta-Strasse 2, D-09596 Freiberg, Germany;

2 New Mexico Museum of Natural History and Science, 1801 Mountain Road N.W., Albuquerque, New Mexico 87104;
3 Naturhistorisches Museum Schloss Bertholdsburg, Burgstrasse 6, D-98553 Schleusingen, Germany;

4 Museum für Naturkunde, Moritzstrasse 20, D-09111 Chemnitz, Germany

Abstract—The giant arthropod Arthropleura was a common member of the late Paleozoic continental biota of
paleo-equatorial biomes for more than 35 million years, from the Early Carboniferous late Visean (FOD; Middle
Mississippian, Asbian/Brigantian) up to the Early Permian lower Rotliegend (LOD = ?LAD; Asselian). In Upper
Pennsylvanian red beds in Cañon del Cobre of northern New Mexico, trackways of Arthropleura are present in
strata that also yield body fossils of the amphibian Eryops. We review the Arthropleura tracksite from Cañon del
Cobre, New Mexico, as well as other tracksites of this animal and arthropleurid/eryopid associations in order to
better interpret the paleoenvironmental preference and the paleobiology of Arthropleura. This review supports
the conclusion that Arthropleura was well adapted to alluvial environments of ever wet humid to seasonally dry
and semihumid climates. Preferred habitats of semi-adult and adult Arthropleura were open, vegetated, river
landscapes. They co-occurred in these habitats with semi-aquatic eryopid amphibians and terrestrial pelycosaurs.

INTRODUCTION ments in order to better interpret the paleoenvironmental preference and


the paleobiology of Arthropleura.
Since the first description of Arthropleura (von Meyer, 1853;
Institutional abbreviations: FG – Technische Universität
Jordan and v. Meyer, 1854; cf. Kraus and Brauckmann, 2003), most Bergakademie Freiberg, Paleontological Collection, Germany; MfNC–
remains of this enigmatic giant arthropod were sampled by plant collec Museum für Naturkunde, Chemnitz, Germany; MHNA – Museé
tors and palaeobotanists in plant-rich roof shales of coal seams from
d´Histoire naturelle, Autun, France; MTB - private collection M. Thiele
mine dumps. This led to the assumption that this animal preferably lived Bourcier, Kirkel-Neuhäusel, Germany; NHMS - Naturhistorisches Mu
in the tight hygrophilous to hydrophilous vegetation of Carboniferous
seum Schloss Bertholdsburg, Schleusingen, Germany; NMMNH – New
coal-forming tropical forests (e.g., Rolfe and Ingham, 1967; Rolfe, 1969, Mexico Museum of Natural History and Science, Albuquerque, New
1980). However, an interesting observation, which would have modified Mexico, USA; SSB – Sammlungen Schloß Burgk, Feital, Germany.
this picture very early, has been overlooked: Guthörl (1940) reported
from his detailed paleontological core investigation of the deep drilling LITHO- AND BIOFACIES OF THE DIPLICHNITES
Hangard No. 38 the discovery of Arthropleura remains at six levels in the CUITHENSIS OCCURRENCE IN THE EL COBRE CANYON
depth interval between 50 to 350 m in the Westphalian D Saarbrücken FORMATION, LATEPENNSYLVANIAN/EARLYPERMIAN,
Group of the Saar-Nahe basin. Remarkably, those finds do not come CHAMA BASIN, NEWMEXICO
from roof shales of coal seams but from a 235-m-thick sequence of
alluvial plain deposits without any indication of coal-forming forests and Nonmarine siliciclastic red beds at the base of the Phanerozoic
mires. Obviously, Arthropleura body remains are as common in alluvial section across most of the Chama Basin of northern New Mexico are
plain deposits as they are in deposits of peat environments. assigned to the Pennsylvanian-Permian Cutler Group. These strata are
In contrast to body remains, which could be transported over divided into two mappable lithostratigraphic units, the El Cobre Canyon
large distances, tracks of animals are absolutely autochthonous. Since the and overlying Arroyo del Agua formations (Krainer and Lucas, 2001;
first discovery of Arthropleura tracks in the Westphalian B (Bashkirian/ Lucas and Krainer, 2005). The El Cobre Canyon Formation is up to 500
Moscovian) of Nova Scotia, Canada (Ferguson, 1966), at least 11 more m of brown siltstone, sandstone and extraformational conglomerate of an
tracksites became known in North America and Europe, ranging from the ephemeral braided and anastomosed stream environment (Eberth and
Visean of Scotland (Pearson, 1992) up to the Stephanian/Lower Rotliegend Miall, 1991) that overlies Proterozoic basement in the subsurface and is
transition of Germany (Early Permian, Asselian: Walter and Gaitzsch, conformably overlain by the Arroyo del Agua Formation. Siltstone beds
1988) as well as the assumed Early Permian of Nova Scotia (Ryan, of the El Cobre Canyon Formation contain numerous rhizoliths (Fig.
1986). These track records considerably widen our knowledge of the 2E-F) and comprise relatively thin, slope-forming units between
locomotion and paleoenvironments of Arthropleura (Briggs et al., 1979, multistoried sandstone beds that are arkosic, micaceous, coarse grained
1984; Wilson, 2003). The most recent description of an Arthropleura and trough crossbedded (Fig. 2A-C). Scoyenia traces are typical of hori
tracksite reports the co-occurrence with the eryopid amphibian track zontal bedded, mica-rich siltstones (Fig. 2 D). Complete bioturbation of
Limnopus in the Upper Pennsylvanian of Kentucky (Martino and Greb, the Planolites montanus type is rare.
2009). It provides additional strong evidence of the co-existence of The El Cobre Canyon Formation approximately corresponds to
arthropleurids and eryopid amphibians in alluvial environments outside megasequence 1 of Eberth and Miall (1991). According to Eberth and
of coal-forming forests and swamps, which was discussed earlier by Miall (1991), during deposition of megasequence 1 the climate was semi
Schneider and Barthel (1997), and by Schneider and Werneburg (1998). arid, major channels were broad and shallow, interfluve areas were fre
Here, the Arthropleura tracksite from Cañon del Cobre, New Mexico quently inundated by sheet floods and splays, and floodbasin ponds
(Lucas et al., 2005c) (Fig. 1) will be discussed in comparison with other were relatively common. These sediments are devoid of coal, but do
tracksites of this animal as well as further arthropleurid/eryopid environ contain leaf floras dominated by seed ferns (DiMichele and Chaney,
2005).
50

FIGURE 1. Location map of Cañon del Cobre in northern New Mexico and stratigraphy and fossil locality distribution (numbers are fossil localities of the
New Mexico Museum of Natural History and Science) of the Upper Pennsylvanian and Early Permian strata in Cañon del Cobre (after Lucas et al., 2005b).
51

FIGURE 2. Facies pattern and facies-typical fossils of wet red beds of the El Cobre Canyon Formation (Late Pennsylvanian/Early Permian) in Cañon del
Cobre. A, Typical alluvial plain multistoried sandstones and pedogenically-overprinted siltstones, which make up the whole profile of the formation. B,
One of the rare, horizontal-bedded, shallow lacustrine mudstone and fine sandstone horizons, which is erosively overlain by trough-cross bedded channel
sandstones; at channel bottoms are large mudstone intraclasts (arrows). C, Interbedding of lenticular sandstones of minor channels with overbank
siltstones, covered by a major channel sandstone. Siltstones altered to calcic soil with bright red to whitish-bluish color mottling and horizons and layers
of carbonate nodules. D, Fine sandstone with cf. Scoyenia burrows. E, Densely arranged vertical rhizoliths in siltstone. F, Horizontal rhizoliths weathered
out from siltstone.
52
to 1.5 to 2 cm deep (Fig. 4C-E). The external width of the trackway is 32
The age of the El Cobre Canyon Formation is Late Pennsylva
nian-Early Permian (Lucas et al., 2005b; Lucas, 2006). Fossils from the 38 cm. Based on the body width/length ratio of 3.6 to 4.4 (Kraus, 1993;
lower part of the formation in the floor of Cañon del Cobre indicate a Schneider and Werneburg, 1998), an Arthropleura of 1.37 to 1.67 m body
Late Pennsylvanian age: palynomorphs (Utting and Lucas, 2010, this length was the trail maker.
The track-bearing sandstone horizon is interpreted as the result of
volume), megafossil plants (Alethopteris flora: Smith et al., 1961; Fracasso, lateral avulsion on the floodplain of minor side channels of a major
1980; Hunt and Lucas, 1992; DiMichele and Chaney, 2005; DiMichele
et al., 2010, this volume) and fossil vertebrates such as Desmatodon and channel in a braided river system. Larger conglomeratic channels are
Limnoscelis (Fracasso, 1980; Lucas et al., 2005b). exposed 1.95 m above the track-bearing horizon (Fig. 3), which could
In general, the red-bed sequences of the El Cobre Canyon Forma easily be regarded as lateral equivalents of the minor channels. A part of
the tracks is preserved as more than 2 cm deep, impressed, crescent
tion belong to the type called “wet red beds” (Schneider et al., 2006;
Roscher and Schneider, 2006). Wet red beds are characterized by the shaped undertracks, well visible where the higher parts of the sandstone
following litho- and biofacies markers, which are observable infine clastics with the surface, on which the animal walked, were weathered away (Fig.
of overbank deposits in channel-sandstone-dominated alluvial plains as 4C-D). The remarkable depth of the tracks as well as the soft sediment
well as in fine, clastic-dominated floodplain deposits: deformation with indications of plastic to fluid behavior of the river sand
1. Horizontal, planar to weakly flaser bedded, fine sandy to clayey, during trackmaking indicate that the animal moved on wet sand (Fig. 4B
E). Identical observations were made on several other Arthropleura
often mica rich, red-brown to red siltstones with common Scoyenia bur
rows and/or completely bioturbated by Planolites montanus burrows tracksites (see below). The sandstone is strongly carbonate cemented,
(Scoyenia ichnofacies). In places, laminated siltstone and claystone of which may indicate high primary porosity as a prerequisite for fluidiza
shallow pools and lakes (Fig. 2B) preserve traces of the Mermia tion. Bright red to whitish colors and greenish patches (Fig. 4B-D) result
from leaching by paleo-groundwater flow, which is responsible for the
ichnoguild.
2. Bed-scale, pedogenically overprinted, in situ micro-brecciation synsedimentary to early diagenetic carbonate cementation as well.
(mm- to cm-scale), pedogenic slickensides (vertisols) and color mottling. About 15 m above the Arthropleura tracksite (NMMNH locality
3. In places, mm-thin, mainly horizontal arranged and branched 6037), another NMMNH locality (6121) contained a well preserved
root systems, preserved as red clay inundations or leached whitish. Eryops skull together with bones of an embolomere, of the dissorophid
4. Calcic soils of different maturity, ranging from mm- to dm-sized Platyhystrix and pelycosaurs in muddy sandstones (Lucas et al., 2005a;
calcareous nodules and rhizoconcretions up to massive calcrete horizons Werneburg et al., 2010, this volume).
(Fig. 2C, E, F). A rare ichnofossil assemblage of the nonmarine Late Carbonifer
5. Rarely, dm-thick lacustrine micritic limestones with ostracods, ous was recently discovered at NNMNH locality 5874 about 35 m
gastropods and characeans. below the Arthropleura tracksite (Lucas and Lerner, 2010, this volume).
6. Channel sandstones and conglomerates that may be leached to The assemblage comes from a clastic lake-margin facies and consists of
bright reddish and whitish-greenish because of paleo-groundwater flow. arthropod trackways (Diplichnites x Diplopodichnus, Diplichnites gouldi,
7. Rare desiccation cracks and raindrop imprints. cf. Paleohelcura tridactyla, Protichnites isp.), invertebrate grazing and
In contrast to the semiarid to arid dry red beds of playa and feeding traces (Gordia indianensis and Helminthoidichnites tenuis), tet
sabkha environments, wet red beds are indicative of a semihumid to rapod footprints (Batrachichnus salamandroides and aff. Amphisauropus
semiarid climate with seasonal wet and dry conditions. Evaporation is isp.) and fish swim traces (Undichna isp.) The numerically dominant
higher than precipitation, allowing the formation of calcic soils. Seasonal ichnofossils are Helminthoidichnites and Diplichnites; the otherichnotaxa
high groundwater levels allow for sparse vegetation only, which is merely are represented by a few specimens. Overall, the wet red beds of the El
documented by the root structures. Above ground (surface) macro-re Cobre Canyon Formation belong to the Scoyenia ichnofacies (cf. Buatois
mains are mainly rare or missing. In the paleo-equatorial belt, these kind and Mángano, 2007) and it is composed of members of the Diplichnites
of red beds were first widespread in the Westphalian D after the and Mermia ichnoguilds, typical of floodplains and shallow lacustrine
Westphalian C/D wet phase A of Roscher and Schneider (2006) and can settings (Buatois et al., 1998; Buatois and Mángano, 2009).
be observed in all the succeeding dry phases up to the intensive aridization LITHO- AND BIOFACIES OF THEARTHROPLEURA OCCUR
during the late Early Permian (Kungurian: cf. Roscher and Schneider,
RENCE IN THE MANEBACHFORMATION, EARLYASSELIAN,
2006, fig. 15a-b; Schneider et al., 2006), when they are increasingly
THURINGIAN FOREST BASIN, GERMANY
replaced by playa red beds.
The El Cobre arthropleurid trackways (Lucas et al., 2005c) occur The Thuringian Forest Basin (formerly SW-Saale basin), an ap
in the lower part of the exposed section of the El Cobre Canyon Forma proximately 40 to 60 km wide NW-SE oriented depression, is largely
tion on the top surface of a 13- to 15-cm-thick bed of mainly fine- to exposed in the horst structure of the Thuringian Forest Mountains in
medium-, slightly coarse-grained, small scale trough-cross bedded to ripple East Germany. The basin is situated at the northern border of the Variscan
laminated, micaceous, arkosic channel sandstone at NMMNH locality orogen above deeply-eroded Variscan methamorphites and granites. The
6037 (Figs. 3-4). The channel sandstone is erosively cut into small scale, Gzhelian/Asselian (Late Stephanian) to Guadalupian/Lopingian basin
trough-cross bedded, dark reddish-brown-violet, fine-sandy siltstone and fill of about 6000 m thickness is subdivided into nine formations (Lützner
it is overlain by grayish-red, plant-bearing muddy siltstone beds with et al., 2007). The coal-bearing (in the type area up to 180 m thick)
dm-wide fine sandstone lenses. The siltstone beds are indistinctly hori completely gray Manebach Formation was deposited in a low-relief
zontal bedded; common cm-size hematitic nodules as well as roots point landscape by forested swamps, local lakes, and fine-clastic-dominated
to a low pedogenic overprint. They can be regarded as an immature fluvial deposits rich in organic matter. Volcanic rocks in the Manebach
ferralatic paleosol. Formation are restricted to mm to cm thick ash layers within lacustrine
Two trails with the typical features of Diplichnites cuithensis black shales. This formation is famous for its characteristic and well
were found (Hunt et al., 2004; Lucas et al., 2005c), both running nearly investigated Euramerican Stephanian/Lower Rotliegend (Gzhelian/
to the west about 5 m distant from one another. The best preserved track Asselian) flora (e.g., Barthel, 2001, 2003-2008).
is stored now in the New Mexico Museum of Natural History and Based on both the lithofacies and fossil content the following sub
Science (NMMNH P-45287). It consists of two track rows of mostly environments can be distinguished in the Manebach Formation (Figs.
crescentic imprints oriented perpendicular to the midline (Fig. 4B; for 5A, 6; for details see Schneider, 1996; Werneburg, 1997; Barthel, 2001,
further details see Lucas et al., 2005c). Imprints are 5-7 cm wide and up 2003-2008; Lützner, 2001):
53

FIGURE 3. Sketch of the Arthropleura tracksite at NMMNH locality 6037 (compare Fig. 4) in alluvial plain deposits of the El Cobre Canyon Formation
(Late Pennsylvanian/Early Permian) in Cañon del Cobre. 1, 30 cm thick (base not exposed), silty fine sandstone, dark reddish-brown-violet, mica rich,
trough-cross bedded, intercalated with up-to-10-cm-thick lenticular fine sandstone. 2, 13-15 cm thick, bright red to whitish, fine- to medium-grained, small
scale trough-cross bedded to ripple laminated, micaceous arkosic channel sandstone; at the top is the up-to-38-cm-wide Diplichnites cuithensis trail,
NMMNH P-45287, extending W-E. 3, 45 cm thick, silty fine sandstone, mica rich, grayish-red-brown, small scale trough-cross bedded, and two
intercalated, 5-cm-thick, discontinuous layers of fine-grained sandstone lenses. 4, 40 cm thick, silty fine sandstone, dark brown, structureless/massive. 5,
10 cm thick, silty fine sandstone, greenish leached (immature soil). 6, 100 cm thick, fine sandy siltstone, dark brown, at mm-scale indistinctly horizontal
bedded, containing plant fragments as well as cm-size hematitic nodules, some with about 1 mm wide central root channels (tubes) (immature ferralitic soil).
7, up to 45 cm thick, pebbly coarse channel sandstone, whitish, trough cross-bedded. 8, 70 cm thick (top eroded), fine conglomeratic coarse channel
sandstone, dark brown, mainly rounded to sub-rounded quartz pebbles.

1. Medium- to coarse-grained, pebbly, trough cross-bedded chan homogenized by bioturbation (Pelecypodichnus).


nel sandstones; common are stem and strobili remains of Calamites gi 5. Rooted siltstones and claystones of very wet floodplains, and
gas, twigs of meso- to xerophilous conifers (“walchians“) and skeletal in places pure hydromorphic to subhydric cordaitalean root horizons of
remains of the eryopid amphibian Onchiodon (Fig. 5A) as well as iso coal-forming forest mires. In the roof shales of seams, the autochthonous
lated bones of a pelycosaur (?Haptodus) (Werneburg, 2007). In the im swamp forest communities are preserved; most common are Psaronius
mediate neighborhood of channels as well as in point bar sandstones and ferns with their fronds (Pecopteris, Scolecopteris) and pteridosperms,
channel sandstones, C. gigas has been found buried in an upright posi like Odontopteris schlotheimii, Dicksonites pluckenetii, Taeniopteris
tion (Fig. 5E). At the Manebach localities, this unique succulent calamite jejunata, different neuropterids and others, the hygrophilous Calamites
forms (nearly) monotypic stands with about 1 m distance between the multiramis and C. undulates as well as the coal-forming cordaitaleans
stems (Barthel and Rößler, 1996; Barthel, 2001). (Barthel, 2001); insect remains are present (mostly cockroaches).
2. Fine- to coarse-grained sheet sands generated during flood events 6. Above the coal-seam-containing part of the profile there appear
as overbank and crevasse splay deposits; autochthonous Calamites gi lacustrine, carbon-rich siltstones and claystones with the typical Early
gas stands are present as well as allochthonous remains of meso- to Permian palaeoniscid-xenacanthid-fish association of smaller lakes
xerophilous plants from different growth sites above the groundwater (Schneider et al., 2000). Branchiosaur amphibians are very rare; common
level, such as sand bars along river courses and from drier, elevated areas lacustrine invertebrates are ostracods, and, in layers, conchostracans; the
inside the basin and along the basin borders (callipterids, different coni terrestrial biota is represented by diverse plant fragments and blattid
fers such as walchians, the coniferophyte Dicranophyllum and insects (most common, as in many Euramerican lake sediments, is
Odontopteris lingulata, etc.). Anthracoblattina).
3. Fine, sandy siltstones to clayey floodplain siltstones deposited The Manebach locality has been sampled by private plant collec
during waning stages of flood events with layers of species-rich tors and palaeobotanists for about 300(!) years (Barthel and Rößler,
parautochthonous (often well-preserved large fern fronds), hygro- to 1995). Collecting has focused on the plant-rich roof shales of the coal
mesophilous plant remains, representative of the fern-pteridosperm seams. Arthropleura and tetrapod skeletal remains were not discovered
calamite vegetation of floodplain areas outside the peat-forming mires; before the first paleontological research in the fluvial deposits of this
Arthropleura remains are not rare. profile commenced (Werneburg, 1987; Schneider and Werneburg, 1998).
4. Laminated claystones and siltstones of floodplain pools be Obviously, they are restricted to those fluvial deposits and their deposi
tween the channels, with layers of Anthraconaia (Fig. 5D) in places tional environments. From reconstructed leg length and the size of
54

FIGURE 4. Overview of the Arthropleura tracksite, NMMNH locality 6037, and the preservation of Diplichnites cuithensis in alluvial plain deposits of
the El Cobre Canyon Formation (Late Pennsylvanian/Early Permian) in Cañon del Cobre. A, D. cuithensis trail on top of a sandstone horizon originated
from channel avulsion; both the trail-bearing, bright red to whitish channel sandstone and the dark reddish-brown, silty fine sandstone below are trough
cross-bedded. B, Up to 38 cm wide D. cuithensis trail (NMMNH P-45287), which is weathered down to different levels of undertracks; the black star marks
the track shown in figs. D and E, double arrow indicates the inside/outside orientation of the trail and tracks in figs. D and E. C, Deepest undertracks in
front; tracks become shallower and sharper in outline in the direction of the hammer. D, Closeup of the upper track row in B; the transition from simple
crescent undertracks (right) to increasingly shallow but more detailed tracks (left) is not caused just by the varying depth of modern erosion of the trail but
also by the primary depth of imprints, which is 1.5 cm at the track marked by the black star and more than 2 cm (measured from the level of the modern
erosion surface) for the track above the scale bar; the depth of undertracks obviously depends on changing wetness of the channel sand. E, Closeup of tracks
in B (for orientation see black star); the form of the imprints indicates fluid behavior of the sand during trackmaking. F, Thin section of the track-bearing
sandstone; grain size ranges from dominantly medium to fine sand with some coarse grains, mainly angular, mostly quartz followed by plagioclase, rare
lithoclasts, strongly carbonate cemented; strong cementation point on high primary porosity and water fill, which was responsible for the fluid behavior
of the still uncemented sand during trackmaking.
55

FIGURE 5. Exposure and fossil content of the Manebach Formation (Early Permian) at the type locality south of Manebach village, slope at the B4 road,
Thuringian Mountains, Germany. A, Excavation site 2005; numbers 2 to 8 refer to the meter-scale in Figure 6; remains of the eryopid Onchiodon
thuringiensis Werneburg, 2007 were found in the pebbly channel sandstones at 5; Arthropleura remains come from plant-bearing overbank siltstones
between 7 and 8 as well as from loose blocks. B, ca. 30-cm-long skull of Onchiodon thuringiensis; NHMS-WP 2140a. C, Cardiocarpus fructifications
washed together in fluvial fine sandstone; NHMS-WP 3310. D, Freshwater pelecypod Anthraconaia in floodplain pool siltstones; NHMS-WP 3350. E,
Trunk base of Calamites gigas buried upright in fluvial sandstone; NHMS-WP 3407.

paratergits (“pleurites”), body lengths of 0.85 m to 2.25 m have been dosperm-calamite vegetation on floodplain areas between river courses.
calculated for the individuals from Manebach. Based on the above char
acterized litho- and biofacies types, the biotope of Arthropleura and LITHO- AND BIOFACIES OF THEARTHROPLEURA OCCUR
Onchiodon during Manebach time could be described as follows. RENCE IN THE DÖHLEN FORMATION, EARLYASSELIAN,
The Manebach Formation belongs to the Late Gzhelian/Early
DÖHLEN BASIN, GERMANY
Asselian wet phase C of Roscher and Schneider (2006), which is repre
sented, e.g., in the Saar-Nahe basin of western Germany, by the This Late Pennsylvanian to Early Permian basin forms a small
Breitenbach to Altenglan formations and in the French Massif Central by half graben of 22 km by 6 km with its primary extent in the Elbe zone,
the Molloy and Igornay formations (Roscher and Schneider, 2006, fig. which is part of the NW-SE striking Elbe lineament (Schneider, 1994).
15 a-b). Red sediment colors are almost completely absent; only in the The basement and the border of the basin are formed by different Variscan
alluvial fan conglomerates of the basin border facies do violet-colored rock units such as the Meißen Intrusive Complex, metamorphosed and
coarse clastics appear in places. Characteristically, coarse channel clastics folded early Paleozoic sediments as well as the Ore Mountain (Erzgebirge)
are of whitish-yellowish color, which is interpreted as the result of gneisses. Basin topology and the high number of clastic dikes (formed by
leaching by paleo-groundwater flows. This is supported by the presence paleo-earthquakes) indicate that basin development and tectonic activity
of pale gray leached, primary dark violet to reddish brown rhyolite clasts along the Elbe zone were coeval. Preserved basin fill amounts to about
of the fan and channel conglomerates (Lützner, 2001; Lützner et al., 800 m of thickness and is subdivided into four fining-up megacycles or
2007). Lamination as an indication of seasonality is not really well ex formations. In contrast to other Permocarboniferous basins, pyroclastic
pressed in the lake deposits. Therefore, nearly year-around high precipi rocks dominate, comprising up to 50% of the deposits of this basin
tation as well as high groundwater levels can be inferred. In this way, the (Schneider and Hoffmann, 2001). Generally, sedimentation and facies
Manebach Formation is climatically very close to the Westphalian C pattern are governed by higher subsidence along the main fault at the
(Bolsovian) and early Westphalian D, from which most Arthropleura southwestern graben border as well as by strong volcano-tectonic activ
remains were discovered in Europe. Both Arthropleura and Onchiodon ity, including strong seismicity.
lived outside the swamp areas in a river landscape that was dominated by The Döhlen Formation (as much as 100 m thick) comprises two,
Calamites gigas stands along the river banks as well as by fern-pteri approximately 50-m-thick fining-up mesocycles. The mainly red-col
56
ored basal conglomerates are overlain by gray sandstones and siltstones, crystal tuff that overlies with sharp contact and load casts the third coal
locally with seams of carbonaceous shale. The second mesocycle con seam in the mining field Bannewitz of the Döhlen basin. Both pyroclas
sists of grayish to yellowish-white fluvial pebbly arkoses, pyroclastic tic horizons contain the above-characterized mesophilous pteridosperm
rocks, and five seams of carbonaceous shale and coal beds up to 6 m, and fern community (sub-environment 5). Compared to the mining field
locally 12 m, thick. As in the Manebach Formation of the Thuringian Gittersee, 3.5 km away, the discovery horizon in the mining field
Forest basin, red beds are completely missing in the coal-bearing upper Bannewitz was situated on an elevation generated by the ash falls, as
mesocycles. Primary dark violet rhyolite clasts of channel conglomerates shown by Schneider and Barthel (1997, fig. 3). Immediately overlying
are leached to pale gray. The presence of subaerial to subaquatically strata are 3-4 m of sandy tuffites containing tight Calamites gigas stands.
deposited air fall pyroclastics and fluvially-reworked pyroclastics be The occurrence of eryopids in the Döhlen Formation is demon
tween the coal seams indicates that nearly continuous peat formation strated by the discovery of Limnopus tracks in pyroclastics between the
occasionally was interrupted by strong ash falls. These ash falls are first and second coal seams of the Döhlen Formation (Hausse, 1910, and
responsible for in situ preservation of plant communities at their growth later discoveries). The track-bearing horizons are cm-thick, silty to fine
sites (Figs. 7F, 8; Barthel, 1976; Schneider and Barthel, 1997; Rößler and sandy pyroclastics with desiccation cracks. Additionally, in sediments
Barthel, 1998). Upright-standing (up to 3 m tall) calamite trunks at the just above the uppermost or first seam, which is about 6.5 m above the
tops of seams indicate that they were buried during catastrophic ash-fall track-bearing horizon, the famous concentration of six skeletons of
events. Haptodus (Pantelosaurus) saxonicus was discovered (Hausse, 1902; v.
Because of the absence of surface outcrops, the lithofacies and Huene 1925). Thus, Arthropleura, eryopids and haptodonts (pelyco
biofacies pattern of the Döhlen Formation are known only from subsur saurs) can be regarded as quasi-contemporaneous in the above-character
face coal mining and well cores in the coal-seam-containing facies in front ized drier environments with hygro- to mesophilous plant communities
of the alluvial fans that frame the northeastern basin border. Generally, of the coal-forming Döhlen Formation.
facies architectures are governed by volcanotectonic subsidence and re
gional ash falls. Characteristic are vertical and lateral changes between COMPARISON OF PARAUTOCHTHONOUS ARTHROPLEURA
fluvial tuffitic pebbly arkoses, fine-grained ash tuffs and coarse-grained BODY REMAIN OCCURENCES AND AUTOCHTHONOUS
crystal tuffs, shallow lacustrine tuffitic pelites and marlstones as well as TRAIL OCCURRENCES
dirty coals and pure coals. Peat formation was interrupted several times
by ash falls on the swamps, which form basin wide, cm-to-dm-thick The oldest true Arthropleura body fragments (legs, paratergits),
which mark the oldest occurrence of this animal, are known from the
pyroclastic layers (tonsteins) within the coal seams as well as the sharp
Visean (late Asbian to early Brigantian, Middle Mississippian) in the
upper boundaries of the main seams. The ash falls have buried plant
Hainichen basin of East Germany (Rößler and Schneider, 1997). Leg
communities in situ, often as monospecific stands (sinuses) (Figs. 7F, 8),
lengths of about 6 cm (Fig. 9A) indicate individuals of 18 cm body width
and very often the trunks of calamites in upright position.
and 0.8 m body length; lengths of up to 1 m are indicated by the size of
Based on lithofacies and fossil content the following sub-environ
ments could be distinguished in the Döhlen Formation (for details see the paratergits (Fig. 9B) (calculation based on Rolfe, 1969; Schneider and
Schneider and Barthel, 1997; Rößler and Barthel, 1998; Schneider and Werneburg, 1998), which is in good agreement with the body sizes calcu
lated from tracks from the Visean (see below). The Arthropleura remains
Hoffmann, 2001):
1. Well-drained channel arkoses, sand flats of coarse grained tuffs/ occur together with other terrestrial arthropods such as arachnids (e.g.,
the phalangiotarbid Bornatarbus mayasi and the trigonotarbid
tuffites and elevations inside the mire bear monospecific, in situ stands
of the xerophytic Calamites gigas, rarely together with single C. undulatus. Aphantomartus areolatus) and scorpionids as well as shark egg capsules
2. Inside the mire, pure stands of the hydrophilous to hygrophil of the Fayolia type. They are preserved in fluvial sandstones of an
ous, peat-forming, rhizome-bearing Calamites multiramis (=Calamites alluvial braid plain/floodplain environment with upright, buried stumps
striata in anatomical preservation) are present on wet organic substrates of arborescent lycopsids and sphenopsids of the Archaeocalamites
(mostly buried in upright position). radiatus type. These trees colonized channel banks, mires and open
3. Hygrophilous forest swamp communities in permanently wa forested areas, the latter containing higher amounts of mesophilous plants
terlogged areas with broad-leafed, coal-forming cordaitaleans and their (Schneider et al., 2005).
The next oldest record of Arthropleura body remains comes from
typical roots (Fig. 7G-H), psaroneaceous tree ferns, Calamites multiramis, the Westphalian A (Langsettian, Middle Bashkirian, Early Pennsylva
Sphenophyllum oblongifolium and rare Nemejcopteris feminaeformis as
well as other ferns; fluent lateral transitions into pure cordaitalean swamps nian) of the Ostrava-Karvina basin in the Czech Republic (eho and
as well as into C. multiramis stands. ehoova, 1972). In younger strata, the finds increase in abundance,
4. Together with 3, locally monotonous scrambler communities with a maximum in the Westphalian C (Bolsovian) and D (Moscovian,
(sinuses) of S. oblongifolium and N. feminaeformis in places with Middle Pennsylvanian), which is caused simply by widespread coal
Botryopteris, Dactylotheca, Taeniopteris and Calamites multiramis. mining in deposits of this time slice in both Europe and North America
5. Mesophilous pteridosperm-fern-communities (open pioneer (Hannibal, 1997). Guthörl (1936) counted 105 Arthropleura fragments
communities) of Oligocarpia leptophylla, Senftenbergia, Dactylotheca, in the Westphalian B and C (Duckmantian and Bolsovian) of the Saar
different pecopterid ferns and pteridosperms like Dicksonites, basin (Guthörl, 1936, fig. 1 and p.189-195), which led him to the conclu
Neuropteris and Barthelopteris, and rarely, Autunia and Subsigillaria sion that this arthropod “is the most common animal fossil of the
brardii, on well-drained elevated drier areas inside and at the borders of Saarbrücken beds.”
the swamps; the border facies of the swamp is characterized by shallow The vast majority of these finds come from roof shales of coal
but wide fluvial channels filled with reworked pyroclastics. seams, stored on the mine dumps. (Since the 1980s, a German private
6. Mesophilous to xerophilous conifers on slopes and hills sur collector, Michael Thiele-Boucier [personal comm.], has sampled more
rounding the basin and possibly on well-drained elevations and swells than 100 Arthropleura remains from coal mine dumps of the Westphalian
inside the basin, too. C and D in the Saar basin). Roof shales represent the silting-up of swamps
The articulated ventral part with sternites and legs of a recon by fluvial deposits in intracontinental basins and by tidalites in paralic
structed, approximately 0.85-m-long Arthropleura was found in a 6-cm basins (Gastaldo et al., 2004). Floras and faunas of roof shales are,
thick, slightly fluvially-reworked pyroclastic horizon. This horizon is therefore, a nearly indistinguishable admixture of terrestrial and aquatic
situated at the top of a 0.6-m- to 0.8-m-thick, medium to coarse, air fall organisms of swamps, floodplain environments and shallow marine en
57

FIGURE 6. Sedimentology and fossil content of the Manebach Formation (Early Permian) at the type locality south of Manebach village, slope at the B4
road, Thuringian Mountains, Germany; excavation site 2005, compare Figure 5A. Channel lag deposits at profile-meter 5 contain calamite trunks as well
as the Onchiodon remains – skull about 30 cm long with the lower jaw as well as femur (Fem) and ilium (Il) (from Gebhardt et al., 1995; completed after
the 2005 excavation by Werneburg).
58

FIGURE 7. Arthropleura and selected fossils from the Döhlen Formation (Early Permian) of the Döhlen basin in Saxony, Germany; (D from the Saar basin,
Germany, see below). A-B, Arthropleura in partially fluvially redeposited pyroclastics above the 3rd coal seam in the mining field Bannewitz-North of the
Döhlen basin, ca. 24 cm long portion of ventral side showing the well preserved articulated limbs as well as the elements of the ventral exoskeleton, E1
– 1st limb segment, E8 – 8th (last) limb segment with the terminal claw, Sp – movable articulated unpaired spines at each limb segment, R – rosette plate,
PM – ventral part of the flexible pleural membrane on which the sclerites S and K are fixed, S – sternites, K – K-plate, B – non-sclerotised B-element (?a
kind of tracheal pocket); apodema of the limbs in black; FG Thümmel collection. C, Closeup from A, intensely wrinkled B-element and the strongly
sclerotised undeformed sternite and K-plate; sternite and K-plate show on the outer (ventral) surface fine dimples with increasing density towards the
posterior borders of these sclerites. D, Ventral sclerites; note the fine dimples on the outer surface of the sternite and the K-plate as in C and the smooth
surface of the B-element apart from some dimples at the base only; early Westphalian D, pit 4 at Göttelborn, Saar basin, Germany; MTB 1215. E,
Limnopus, tracks of an eryopid amphibian in pyroclastics between the 1st and 2nd coal seams of the Döhlen Formation; SSB. F, Apical portion of a Calamites
multiramis shoot with Annularia spinulosa leaves, buried in situ by volcanic ash fall, pyroclastics above the 2nd coal seam, Döhlen Formation, Zauckerode;
FG 15/98. G, Root horizon in fine-grained volcanic ashes with three-dimensionally preserved cordaitalean roots, Döhlen Formation, Zauckerode; FG 609.
H, Cordaites principalis, pyroclastics above the 3rd coal seam, Döhlen Formation; private collection Hertl, Freital, Germany.
59

FIGURE 8. Reconstruction of the floral associations buried in situ above the 3rd coal seam of the Döhlen Formation (Early Permian) in the Döhlen basin
based on plant collections from the mining fields Gittersee and Bannewitz (from Schneider and Barthel, 1997). The Arthropleura ventral side, shown in
Figure 7A-C, comes from the mining field Bannewitz. Floras of the Gittersee type belong to the hygro- to hydrophile peat-forming communities of poorly
drained environments close to or below the groundwater level. The Bannewitz flora represents the pioneering mesophile associations on well-drained
mineral stands above groundwater level. The open vegetation of comparable environments on sandy substrates of alluvial plains and deltaic settings seems
to have been the preferred habitat of semi-adult and adult Arthropleura.

vironments. Additionally, roof shale samples from mine dumps lack have a relatively narrow range between 21-24 cm. The longest preserved
sedimentological context. They are not suited to answer the question of trail is 5 m, forming an S-bend. Some trails cross one another; at one place
the environmental preference of Arthropleura. As shown by Guthörl the sediment is pitted by deeply impressed, crescent-shaped and ran
(1940) for the deep drilling Hangard No. 38 (see above), Arthropleura domly oriented marks, which may represent hundreds of individual foot
finds are as common in coal-seam-free alluvial plain deposits as they are falls. Generally, individual tracks often appear crescent shaped and deeply
in roof shales. This fits with the still scattered but nearly continuous impressed (1-2 cm estimated depth from the photographs in Pearson,
records of Arthropleura tracks from the Namurian up to the Early Per 1992), comparable to the El Cobre trails. Not excluding the possibility
mian of Europe and North America in alluvial environments outside that the trails are made by two or three individuals only, this mass
swamp areas. Those occurrences will be characterized in the following occurrence of Arthropleura trails did not originate by chance. Animals of
sections based on literature data and personal observations. 0.80 m to 1.32 m body length (maximally ?2.11 m for the uncertain, 46
cm wide trail) were the trail makers.
Strathclyde Group, Visean (Asbian, Mississippian), Fife,
Scotland (Pearson, 1992) Namurian Limestone Coal Group (Pendleian, Early
Serpukhovian), Arran, Scotland (Briggs et al., 1979)
The sediments of the Strathclyde Group of Fife are interpreted as
deltaic in origin with occasional marine incursions. Nonmarine facies In general, the Limestone Coal Group was deposited in a proximal
include channel sandstones and sheet-flood deposits, thin coals, seatearths, deltaic environment. The trail-bearing horizon is situated 6 mm below
lagoonal mudstones and laminated siltstones and shales as well as red or the top of a 6-m-thick white sandstone unit, which formed the roof of
black ironstones and dolomites with nonmarine bivalves. The track coal seams. The trail is preserved on the surface of a bedding plane of
bearing part of the section consists of meter-thick dark, organic-rich and heterogeneous sandstone ranging in grain size from quartz arenite to fine
often finely laminated mudstones and siltstones interspersed with ~ 1.5- sandstone and containing discontinuous layers of shale. Bedding struc
m-thick, white, well-sorted, medium-grained channel-fill sandstones. The tures indicate deposition in a variable low-velocity flow regime in a
sandstones display multi-storied trough-cross bedded units with undu gradually silted-up fluvial channel. Rootstructures suggest that the veg
lating surfaces and, in places, current ripples. Parts of the track-bearing etation was penecontemporaneous with the formation of the trail. A
sandstone bed are spotted with small mineralized rootlets; also present subaqueous origin of the trail is regarded as unlikely because of the
are poorly-preserved Stigmaria root traces. Minimally, 13 individual clarity of the imprints in such a coarse lithology. The width of the
trails occur, ranging in width between 18 and 30 (?46) cm; most of them trackway is 36 cm, suggesting a 1.6-m-long individual as the trackmaker.
60

FIGURE 9. Oldest known Arthropleura remains, from the Visean (late Asbian to early Brigantian, Middle Mississippian) in the Hainichen basin, Borna near
Chemnitz, Germany (Rößler and Schneider, 1997). A, Counterpart of a nearly complete, 7.6 cm long limb, R – rosette plate, E1 – 1st limb segment, E8 –
8th (last) limb segment; MfNC F 439b. B, Counterpart of a fragmentary paratergite (“pleurite”) with the typical sculpture of A. armata from the
Pennsylvanian; MfNC F 1125.

Little River Formation, Cumberland Group, ?Late Namurian to drained coastal plain assemblages (pPDF) and the 31% of the Joggins
basal Westphalian, (?Kinderscoutian/Langsettian, Bashkirian, Formation comprising well-drained alluvial plain red-bed assemblages
(WDF). Arthropleura trails were reported from the last two assem
Early Pennsylvanian), Lower Cove, Nova Scotia, Canada (Calder
blages. In the “grey mudstone with channel bodies facies” of the pPDF,
et al., 2005)
a few D. cuithensis trails (Falcon-Lang et al., 2006) and Dromillopus
The Little River Formation of the Lower Cumberland Group was microsaur trackways (Cotton et. al., 1995) occur on top of some channel
deposited in the Cumberland sub-basin as part of the late Paleozoic bodies.
Maritimes Basin complex of Atlantic Canada (Falcon-Lang et al., 2006). The trails reported by Briggs et al. (1979) occur in the “red mud
The formation consist of a wet red bed-succession dominated by stone with channel bodies facies” of the WDF (Falcon-Lang, et al., 2006).
mudrocks, which exhibit pervasive mottling from root traces and local This facies comprises red mudstone successions with scattered pedogenic
pedogenic carbonate, with channel sandstone bodies typically 3-6 m carbonate nodules, sandstone sheets, and small, ribbon-like channel sand
thick. It represents the deposits of a well-drained alluvial plain dissected stone bodies. They contain common upright calamiteans, and a few
by shallow rivers characterized by flashy flow under a seasonally dry upright lycopsids with attached Stigmaria roots. Plant assemblages are
climate. Coals seams and brackish, marine-bivalve-bearing limestone beds, dominated by cordaitaleans (Dadoxylon, Mesoxylon, Cordaites, Artisia)
which are present in the overlying Joggins Formation, are missing. A 23 as well as a few lycopsids and calamiteans. The trails occur in sheet
cm wide Diplichnites cuithensis is reported from inside a body of stacked sands thickening into a channel sand bed. Ranging in total width from 20
channels by Calder et al. (2005). The tracemaker may have had a length to 26 cm, they indicate an animal about 0.88 to 1.15 m long. The preser
of about 1 m. This is the stratigraphically oldest record of Arthropleura vation of the largest trail, described as a paired series of regularly-spaced,
in wet red beds of seasonally dry environments. oval depressions, elongate normal to the axis, resemble the preservation
of the El Cobre trackways in wet sandstone. The smaller trail shows
Joggins Formation, Cumberland Group, Westphalian A individual imprints, apparently arranged in closely spaced diagonals.
(Langsettian; Bashkirian, Early Pennsylvanian), Joggins, Nova These differences in preservation are explained by decreasing water con
Scotia, Canada (Ferguson, 1966, 1975; Briggs et al., 1979; tent of the sediment, i.e., increased cohesiveness when the smaller trail
Falcon-Lang et al., 2006) was formed.
The Early Pennsylvanian Joggins Formation, containing the fa Tynemouth Creek Formation, Westphalian A to ?B (Langsettian -
mous Joggins paleoecosystem, overlies the Little River Formation (see ?Duckmantian; Bashkirian/?Moscovian, Pennsylvanian), New
above). At that time, the Cumberland sub-basin was connected to the Brunswick, Canada (Briggs et al., 1984)
open sea during sea-level highstands, as indicated by brackish incursions,
and was more restricted and intra-continental during sea-levellowstands The Tynemouth Creek Formation is regarded as partially coeval
to the Joggins Formation of the late Paleozoic Maritimes Basin complex
(Falcon-Lang et al., 2006). The latter authors revised the fossil content of
the Joggins Formation in a precise facies context, recognizing three main of Atlantic Canada (see above and Falcon-Lang, 2006). The formation is
communities: open water brackish sea fossil assemblages (OW), poorly built up mainly of red siltstones, red and gray sandstones, and coarse
61
conglomerates in an overall coarsening upward sequence (Briggs et al., closely spaced, elongated and sigmoidal imprints oriented normal to the
1984). Rare freshwater limestones are locally present. The sedimentary axis of the trackway. Details of the imprints are not preserved because of
environment is interpreted as a relatively dry alluvial fan environment the superimposition of imprints and soft-sediment deformation, which
characterized by periodic sheetfloods across an otherwise quiescent area is visible in figure 32C of Mángano et al. (2002). Obviously, the sandy
of relatively slow deposition. Red/green mottled paleosols with slicken sediment was wet and plastically deformable when the animal was walk
sides and carbonate nodules are interpreted as vertisols of a seasonally ing on it, as was the case with the El Cobre trackway.
dry climate (Falcon-Lang, 2006). The trail-bearing section consists of
red, fine, slightly silty, tabular sandstones, interbedded with red and Lower Conemaugh Formation, Virgilian (Gzhelian, Late
green siltstones. The up-to-2-m-thick sandstones are dominantly mas Pennsylvanian), Boyd County, Kentucky (Martino and Greb,
sive, but include planar and cross-laminated units. Upright buried calam 2009)
ite stems, up to 10 cm in diameter, are common.
The best-preserved trails occur at Gardner Creek in the top of a The lower Conemaugh Formation of the Appalachian basin in
Kentucky is regarded as an equivalent of the Glenshaw Formation of the
40-cm-thick, fine sandstone with numerous in situ calamitaleans rooted
in the 12-cm-thick siltstone horizon below. The fine sandstone grades up Conemaugh Group in neighboring West Virginia, which consists pre
into a few mm of siltstone, and it is on this surface that the Arthropleura dominantly of coastal plain fluvial sandstones and red to olive mudrocks
trail is preserved. The sandstones are interpreted as deposits of major deposited in a tropical flood-basin setting. Martino and Greb (2009)
sheet floods. The trail-bearing siltstone accumulated during the waning reported new direct evidence for the co-occurrence of Arthropleura with
eryopid amphibians, indicated by the tracks of both on one-and-the
phase of the flood, and the trail was produced after subaerial emergence,
same bedding surface from the lower Conemaugh Formation, 6.5 m be
when the Arthropleura was walking along a sinuous course through the
low the marine Ames Limestone. The roof shales of the Ames Limestone
calamite forest. The width of the trail in the straight part is 29.5 cm,
suggesting an animal about 1.30 m long. Two further trails are mentioned can be directly correlated with the early Stephanian of Central Europe
in the top of channel sandbodies beneath overbank sediments at approxi based on fossil insects (Schneider and Werneburg, 2006). The tracks
mately the same stratigraphic level as the Gardner Creek trails. Their were found in float from a roadcut.
width of 27 cm and 30 cm implies similar large animals. The trackways were derived from a 4-m-thick interbedded sand
stone and shale interval 4.3 m below the base of the Harlem coal. The
Cumberland Group, Westphalian B-C (Duckmantian-Bolsovian; tracks are preserved as convex hyporeliefs in fine to very fine sandstone
Moscovian, Middle Pennsylvanian), Smith Point and Pugwash, that is interbedded with silty shale. They were mainly produced on top
Cumberland County, Nova Scotia, Canada (Ryan, 1986) of the silty shale. In one case (specimen PC-6 of Martino and Greb,
2009, p. 142, fig. 4), an Arthropleura trail is preserved as undertracks in
The Westphalian Cumberland Group of Cumberland County con the sandstone together with a Limnopus manus cast (?undertrack). The
sists, according to Ryan (1986), of continental gray to red-gray calcare track-bearing sandstones occur as three broad, thin sheets and a small
ous mud-chip conglomerate, coarse- to fine-grained sandstone, siltstone channel fill. This, together with graded bedding, parallel and current
and mudstone with minor amounts of coal, limestone and shale. Sand ripple cross lamination and desiccation cracks leads to the interpretation
stones and conglomerates are trough cross-stratified and were deposited of crevasse splay deposits. Besides sparse root traces, no plant remains
in low sinuosity streams. At Smith Point, 25 trails were found on two were reported. The Pittsburgh red shale, which directly underlies the
different surfaces that are separated by a 1.3-m-thick, cross-stratified, trackway interval, is indicative of the climate conditions during deposi
medium-grained sandstone. Crossover of trails is common; up to 300 tion. These red beds consist of regionally-developed calcic vertisols,
degrees of bending is observed. The average width of the trails on the which classify them as wet red beds in a seasonal dry/wet semihumid to
upper bedding surface is 37 cm, and it is 32 cm on the lower bedding semiarid climate (see above and Greb et al., 2006). Open woodland
surface, which gives a body length of 1.63 and 1.40 m, respectively, for vegetation is assumed. Based on the width of the Arthropleura tracks of
the producer. The tracks are closely spaced, about 24 per meter; maxi 20 and 30 cm, we infer a body length of the tracemaker of 0.88 m for one
mum depth of the tracks is 3 cm. At the Pugwash locality, 12 trails of the tracks and 1.32 m for two other tracks. The Limnopus traces
preserved mainly as undertracks were found. The trail width is approxi consist of 75 to 80 mm long pes and 53? to 57 mm wide manus imprints.
mately 38 to 41 cm; depth of tracks varies between 2 and 3 cm. The An additional tracksite from the Conemaugh Group, Casselman Forma
length of the trail-producing Arthropleura is calculated as about 1.8 m. tion, in Cambria County, Pennsylvania, was briefly reported by Marks
At both localities, the tracks occur in pebbly arkosic sandstones et al. (1998). The 15 cm wide by 3 m long trackway resembles D.
that are overlain by thin veneers of mudstone, preserved on some of the cuithensis as far as can be determined from the published photographs.
trail surfaces. The depositional environment is interpreted as mid-chan
Montceau-les-Mines, latest Stephanian C/earliest Autunian
nel dunes prograding during flood stages, which were subsequently sub
aerially exposed as channel bars during a dry period, in which the trails (Asselian, Cisuralian, Early Permian), French Massif Central
were produced. (Briggs, 1986; Briggs and Almond, 1994)

Stull Shale Member, Kanwaka Formation, Shawnee Group, The Blanzy-Montceau basin is one of the typical Late Pennsyl
Virgilian (Gzhelian, Late Pennsylvanian), Waverly, Eastern vanian/Early Permian continental coal basins of the French Massif Cen
Kansas (Mángano et al., 2002) tral. It is famous for its fossiliferous nodules similar to Mazon Creek in
Illinois. Arthropleurid trails were found in a unit of alternating pale gray
The Stull Shale Member of the Forest City basin in eastern Kan mudstone and fine sandstones at the level of the first coal seam of the
sas represents tidal flat deposits along a microtidal shoreline directly Assise de Montceau (Langiaux, 1984). The trails are preserved at the top
connected with the open sea. The arthropleurid tracks occur at the top of of whitish-gray, silty, fine sandstone that contains numerous plant frag
a 30-cm-thick, medium- to fine-grained channelized sandstone body 66 ments (Fig. 10C; Briggs, 1986). The width of the trail of 10.8 cm indi
m wide, which cuts erosively into an intertidal runoff channel. Both cates a length of the producing arthropleurid of 0.47 m. About 15 m
channel bodies are encased by progressively shallowing intertidal mixed above the trail-bearing level, three-dimensionally preserved tiny
to mudflat deposits. The track-bearing sandstone is interpreted as a arthropleurids of 2.7 to 4.5 cm length have been found in sideritic nod
channel fill in a coastal fluvial system. The width of the trackways is ules (Fig. 10A-B; Secretan, 1980; Almond, 1985; Briggs and Almond,
23.3 to 30.2 cm, which indicates body lengths of about 1 to 1.3 m. The 1994). The width of these dwarf arthropleurids ranges from 1.1 to 1.7
preservation style resembles very closely the El Cobre trackways with cm. They differ from the smallest true Arthropleura (A. moyseyi Calman,
62

FIGURE 10. The tiny arthropleurids of Montceau-les Mines, France, and trails of juvenile arthropleurids from Flechting High, Germany. A-B, Two
approximately 1.6 to 1.7 cm wide, three-dimensional arthropleurids preserved in sideritic nodules, Assise de Montceau, latest Stephanian C/earliest
Autunian (Asselian, Cisuralian, Early Permian), Montceau-les-Mines, French Massif Central; A, 3.7 cm long natural cast of the ventral (inner) side of the
lateral paratergits and the dorsal side of the body as well as two incomplete ventral molds of syntergites that are about half of the length of the specimen;
anteriormost plate of the ventral mold of a collum or cephalic shield (cf. Briggs and Almond, 1994; Kraus et al., 2003); MHNA 002122b. B, 3.6 cm long
natural cast of the ventral (inner) side of the lateral paratergits and the dorsal side of the body as well as on the right distal side of the specimen imprints
and counterprints of closely arranged limbs; note the strongly tapered posterior region of the trunk; MHNA 002123b. C-D, D. cuithensis trails of juvenile
Arthropleura preserved as convex hyporelief in silty-clayey pyroclastics, Eiche Member, Altmark Subgroup, Stephanian ?C (Late Pennsylvanian/Early
Permian, Ghzelian/Asselian), Flechting High, Germany (see Walter and Gaitzsch, 1988); lower trail 2.0 to 2.2 cm wide; note the faint striae (especially in
the bent part of the trail) made by the movable spines of the limbs during forward stroke; this trail was made in wet plastic mud; upper trail 3.2 cm wide;
tracks have a lesser width despite the larger trail width and show more distinct limb end claw imprints of repeated foot falls; this trail was possibly made in
a somewhat drier substrate; note the very common small cf. Batrachichnus tetrapod tracks; FG 607.
63
1915) of about 6 cm body length by a distally very strong, tapering body long to the same Arthropleura species as the D. cuithensis trail makers.
and possible diplopody (Briggs and Almond, 1994) as well as very long The only way to produce trails of D. cuithensis type by the tiny
and thin, spike-like sculpture cones on the paratergits (JWS, personal Montceau-les-Mines arthropleurids would be if the posteriormost limbs
observation on silicon rubber casts). They may belong to Arthropleura were held clear of the ground (Briggs and Almond, 1994, p.132).
or represent a new minute form of arthropleurid — for discussion see The Eiche Member D. cuithensis occur on different bedding planes.
Briggs and Almond (1994) and below. In places, normal surface trails (epichnia) are crossed by undertracks of
The depositional environment is described as a fault-bounded trails produced later after a new layer of sediment was laid down. They
graben with rapid lateral and vertical transitions of alluvial fan deposits, are commonly associated with small cf. Batrachichnus trails of 3 to 7 mm
debris flows and braided river clastics interfingering with palustrine (lo track length (Fig. 10D). The wide range of preservation forms of the
cally up to 40 m thick coal seams!) and lacustrine sediments. Siltstones, arthropod and tetrapod trails indicate different amounts of wetness of
intercalated into the first coal seam have yielded a diverse tetrapod track the substrate, ranging from possibly thin water cover over muddy to
association (Langiaux, 1984; Gand, 1994), produced mainly by nearly stiff. Sparse remains of macroflora, mainly calamite trunks (up to
temnospondyl amphibians (e.g., Batrachichnus, Limnopus) and pelyco 5 cm diameter) and Annularia leaves as well as rare fern leaves, indicate
saurs (Dimetropus). The fossiliferous nodules and track-bearing Assise scattered vegetation in surrounding areas of small ponds. The high fre
de Montceau is commonly attributed to the Stephanian B/C based on quency of arthropleurid trails (laterally and vertically) and in places the
macroflora. Spiloblattinid cockroach wings indicate a latest Stephanian common association with trails of small tetrapods demonstrates that this
C/earliest Autunian, i.e., Early Permian (Asselian) age (Schneider and environment was settled by juveniles of Arthropleura. The lack of larger
Werneburg, in prep.). A transitional Stephanian/Autunian age had been trails may indicate a different biotope preference of juveniles and adults.
advocated by Doubinger (1994) based on palynomorphs, but was never
taken into account. Pictou Group, Upper Red Beds, Early Permian (Cisuralian), Cape
John, Pictou County, Nova Scotia, Canada (Ryan, 1986)
Eiche Member, Altmark Subgroup, Stephanian ?C (Late
Pennsylvanian/Early Permian, Gzhelian/Asselian), Flechting The Early Permian Upper Red Beds, now called the Cape John
High, northeast Germany (Walter and Gaitzsch, 1988) Formation (Ryan et al., 1991), of the Pictou Group are composed of
fluvial cycles of red to gray conglomerates and arkosic sandstones as well
The Eiche Member is exposed at the top of subeffusive basaltic as siltstones and mudstones. The trail-bearing pebbly arkosic sandstones
volcanites situated near the base of the Late Pennsylvanian/Early Per are interpreted as anastomosing river deposits, with sediment transport
mian Altmark Volcanite Complex of northeast Germany (Gaitzsch et al., and accumulation primarily as the result of dune progradation at channel
1995). It consists mainly of about 10 m of light-gray fluvial sandstones bottoms. At Cape John, on the northern coast of Nova Scotia, three trails
at the base followed by several meters offine ash tuffs covered by debris were found, and two of them cross each other. The width of the trails is
flow fanglomerates with an ignimbrite at the top. The track-bearing unit on average 47 cm, giving a body length of the arthropleurid of about 2.07
is composed of (?water laid) grayish-green air fall tuffs as well as fluvially m. They are preserved as two parallel rows of tracks, each 13 cm wide,
reworked and lacustrine, redeposited fine ash tuffs. Shallow channels elongate and nearly normal to the trail axis with 28 tracks per meter. The
with trough-cross bedding have a lateral width of only about one meter. depth of the tracks is about 1.3 cm.
Sedimentation took place under low energy conditions from suspension In the Cape John Formation as well, about 8 km southwest of the
as indicated by the silty grain size and normal grading. Raindrop marks, Cape John Arthropleura tracksite, a spectacular tetrapod tracksite to
shallow wave ripples and rare “anthracosian” (?Anthraconaia) imprints gether with in-situ stumps of a walchian conifer forest is exposed at
as well as Pelecypodichnus traces indicate the presence of shallow pools. Brule Harbour, Colchester County (Van Allen et al., 2005). The tetrapod
The track-bearing layers are horizontally bedded at a 1 to 2 cm scale. tracks represent a typical cosmopolitan Euramerican association with
Tens of arthropleurid trails per square meter are preserved on the bed strong European affinity, consisting of Limnopus, Amphisauropus,
ding planes, in places as undertracks. Crossover and bending is common Dimetropus, Varanopus and Dromopus and others. Absolutely domi
(Fig. 10D). nant is Limnopus, followed closely by Amphisauropus. Provisionally
Because of their small size, ranging from 2.0 to 4.0 cm width, the determined invertebrate traces, as Gordia, Beaconichnus, and
trails have been described as Diplichnites minimus Walter and Gaitzsch, Onisciodichnus, are typical representatives of the shallow freshwater
1988. Apart from bundles of striae situated in front of the single tracks in Mermia ichnoguild of the Scoyenia ichnofacies, an interpretation that is
both track rows of some trails (Fig. 10D, lower track) no real differences strongly supported by the co-occurrence with the ostracod Carbonita
exist when compared to D. cuithensis. The striae result from the spines and leaiaid conchostracans. Besides Paleohelcura, no arthropod walking
on the limb segments during forward stroke. Therefore, D. minimus trails are reported. Most tetrapod tracks are preserved in thin, silty
should be regarded as a junior subjective synonym of D. cuithensis. sandstone beds commonly draped by very thin mud layers. Desiccation
Based on the width of the trails, a length of 8.8 to 17.6 cm of the cracks are typical. The depositional environment of the Brule Harbour
trailmakers can be calculated. This corresponds well to the smallest tetrapod tracksite is described as an abandoned distributary channel
known, unquestionable Arthropleura (A. moyseyi Calman, 1915) with a hollow that was repeatedly flooded and exposed. Red beds of Prince
body length of about 6 cm. There are no indications of the strong poste Edward Island, adjacent and contiguous to Brule Harbour, have yielded
rior tapering of the body as observed for the tiny “juvenile” arthropleurids skeletal remains of eryopid and brachyopoid amphibians and the
from Montceau-les-Mines (see Briggs and Almond, 1994, figs. 1 and 2; reptiliomorphs Seymouria and Diadectes as well as of different pelyco
here Fig. 10D). As discussed by Briggs and Almond (1994), the posterior saurs (Van Allen et al., 2005).
tapering of the trunk and the closely-spaced rear limbs on either side
THE HABITAT PREFERENCE OFARTHROPLEURA
should result in a trail in which the imprints occupy over 50% of the total
width. The trail width of these small forms should be about 1.6 cm, Based on the foregoing, Arthropleura trail-bearing paleoenviron
corresponding to the largest width in the middle part of the body. This is ments can be characterized as follows:
very close to the smallest trails of 2 cm width from the Eiche Member. 1. Arthropleurid trails of dm width are preferably preserved both
But, the trails from this member do not show any indication of a trailmaker in nearshore transitional terrestrial/marine deltaic and tidal flat sediments
with a posteriorly tapering body – they display the normal picture of as well as in intracontinental river braid plains on sandflats and on bed
two rows of tracks, as do the larger D. cuithensis. It is therefore very ding planes at the top of channel sandstones. Smaller trails up to about 5
unlikely that the minute arthropleurids from Montceau-les-Mines be cm wide are preserved in silty mud of temporary pools.
64
2. As a prerequisite for the preservation of tracks, the substrate wet sands. Alternatively, the absence of tracks produced by arthropleurids
must be wet (of course) when the animal was walking on it; some of the larger than 0.18 m in the mass occurrence at Flechting High may indicate
trails show indications of plastic deformation/fluidization of the sand different habitat preferences of juveniles and adults. Nevertheless, it
and mud caused by deeply impressed tracks (Fig. 4D-E). could be stated that arthropleurids, starting no later than 0.8 m body
3. Vegetation of the tracksites is scattered, consisting mainly of length, walked preferably on sandy substrates. This confirms Mángano
calamitaleans, and during the Westphalian, additionally of arborescent et al. (2002), who described the depositional environment of D. cuithensis
lycopsids, as indicated by their upright buried stumps and Stigmaria as “typical subaerial, commonly exposed fluvial bars, silted channels,
rhizomes/rootlets. and desiccated sheetflood deposits.” More generally, the preferred habi
4. Associated with the track-bearing sandstones in coastal marine tat of adult arthropleurids could be characterized as loosely vegetated
environments are gray lagoonal and fluvial overbank mudstones and silt sandy areas in open river landscapes under a year-round wet climate in
stones, more or less rooted. Additionally, tracks are present since the coastal environments to seasonally wet, semihumid climates in continen
middle Westphalian in alluvial plain and floodplain red siltstones and tal settings. As indicated by skeletal remains and tracks, the most com
mudstones of wet red-bed facies type in continental settings. mon tetrapods in these habitats since the start of the Late Pennsylvanian
5. In rare cases only, coal seams or swamp deposits are associated were eryopid amphibians and pelycosaurid reptiles (haptodonts).
with the Arthropleura track-bearing facies architectures.
6. The most common tetrapods in those settings are terrestrially IMPLICATIONS FOR THE
adapted eryopid amphibians (Figs. 5B, 7E) and pelycosaurid reptiles. PALEOBIOLOGY OFARTHROPLEURA
The first question to be answered is: do the trail occurrences
The most frequently discussed issues concerning the anatomy
reflect the biotope preference of the trail makers or is it biased by sam
and physiology of Arthropleura concern the up-to-now-missing struc
pling or by preservation potential? Of course, the discovery of large tures of the head, especially the mandibles and the missing respiratory
trails of 20 cm and more width and several meters long, and the chance to
organs as well as the arrangement of the limbs (Brauckmann et al., 1997;
recognize them as trails, obviously depends on the dimensions of the
Kraus and Brauckmann, 2003). Kraus and Brauckmann (2003) and Kraus
exposed bedding planes. They are largest in areas of high recenterosional
(2005) discussed the phylogenetic relationships of Arthropleura as well
rates such as at modern sea coasts and in sandstone quarries. There, the as the anatomy and biology of this animal based on the reinvestigation of
fine clastics covering track-bearing sandstones are quickly eroded on
the tiny Montceau-les-Mines arthropleurids, the holotype of A. moyseyi,
large surfaces or – in the case of the quarries – removed. This is clearly the large, well-preserved, so-called Maybach-specimen (Guthörl 1935;
shown, for example, in the case of the tracksites along modern coasts of Hahn et al., 1986), and the Arthropleura ventral side from the Döhlen
Scotland, by the exceptional exposures along the coast of Nova Scotia basin (Schneider and Barthel, 1997) as well as several new body frag
and by the channel sandstones exposed in the modern, non-vegetated and ments, including a presumed cephalic shield and presumed head cap
seasonally dry river bed of Cañon del Cobre as well. The apparent
sules, from the Saar basin. Some of their observations and inferences as
absence of these large trails in Central Europe, where Arthropleura body well as those of other authors will be discussed here with regard to our
remains are very common, is simply caused by the lack of such large conclusions on the paleobiology of Arthropleura.
exposure surfaces in upper Paleozoic fluvial sandstone units. As shown by Schneider and Barthel (1997), based on a well
The preservation potential for animal tracks (invertebrates and
preserved, 42-cm-long part of a ventral side with about 25 limb seg
vertebrates as well) is generally highest in fine clastics of river flood ments, there are no indications that succeeding pairs of legs form groups
plains and playas (e.g., Haubold and Katzung, 1978; Haubold and Lucas, of two pairs. In contrast, the legs had an almost regular arrangement
2001; Voigt, 2005; Lucas, 2005). Tetrapod tracks are very common on (their pl. 7; here Fig. 7A-B). Whether or not one pair of legs corresponds
floodplain mudflats and around temporary pools on floodplains that to one of the dorsal syntergites still remains open (cf. Schneider and
became dry after floods. But, those tetrapod track-bearing sediments Barthel, 1979, fig. 5). Diplopody, as shown by the tiny Montceau-les
have not delivered any associated arthropleurid tracks on the same bed Mines specimens (Briggs and Almond, 1994), can be excluded for
ding planes. Even so, they are broadly contemporaneous with Arthropleura Arthropleura of more than 80 cm body length. A change from diplopody
tracksites, as in the case of Cañon del Cobre (Lucas and Lerner, 2010) in early juvenile stages to a regular arrangement of single limb pairs in late
and the case of the spectacular tetrapod tracksite of Brule Harbor, Nova juvenile and adult stages can also be excluded. Therefore, the inferences
Scotia (Van Allen et al., 2005). The only two exceptions are the Limnopus of Briggs et al. (1979) regarding the interpretation of Arthropleura traces
tracks together with 20 to 30 cm wide Arthropleura trails in the Upper still appear to be correct. However, the assumption of Kraus (2005) that
Pennsylvanian of Kentucky (Martino and Greb, 2009; see above) as well the tiny Montceau-les-Mines arthropleurids are early juveniles of the
as the minute Batrachichnus tracks together with 2-4 cm wide
giant Arthropleura as well as his conclusions on phylogenetic relation
Arthropleura trails in the Upper Pennsylvanian/Lower Permian of the ships to the Penicillata clade of Diplopoda are not well supported.
Flechting High, Germany (Walter and Gaitzsch, 1988; see above). The respiratory structures of Arthropleura remain unknown. Rolfe
As shown before, the vast majority of Arthropleura trails are
and Ingham (1967) and Rolfe (1985) argued that the K- and B-plates may
preserved in sandy substrates. Additionally, those trails are observed have covered a respiratory chamber. Schneider and Barthel (1997) have
(depending on exposure conditions) mostly as mass occurrences (e.g., up shown that the K-plate was as strongly sclerotised as is the sternite plate
to 25 trails at Smith Point, Nova Scotia – see above). Besides the above (Fig. 7A-C). Unlike the K-plates and the sternites, the so called B
cited 10 cm wide Montceau-les-Mines trail and the small Flechting High
“plates” are often wrinkled (Schneider and Barthel, 1997, pl. 8, figs. 1-3;
trails, the width of the trails has a range of 18 to 47 cm, which indicates
here Fig. 7C). Possibly, this B-element, which is situated close to the
body lengths ranging from about 0.80 to 2 m. The variety of body lengths base of each walking limb and linked to apodema, had a respiratory
for particular sites, calculated from the width of the trails (e.g., Fife,
function as a kind of tracheal pocket. The “spongy internal structures”
Scotland: 18-30 cm; Kentucky 20-30 cm; see above), indicates animals of (Kraus and Brauckmann, 2003, p. 47) of K-plates are fine dimples on the
different growth stages as tracemakers (despite variations in trail width
outer surface, which give the inner surface a verrucose appearance. Not
caused by locomotory behavior). only the K-plate shows this structures but the sternites do, too (Fig.
This leads to the conclusion that arthropleurids avoided muddy
7D), both dorsally and ventrally. Similar pits are situated on parts of the
substrates, except for early juveniles of up to about 0.15 m body length, posterior surfaces of the limbs. Remarkably, the B-elements show only
as at Flechting High (Fig. 10D). The preservation in sandy substrates
very weak pits on their median ends (Fig. 7D). A semiaquatic life and
begins by about 0.8 m body length. This may be caused simply by the “plastron-breathing via series of paired ventral K-plates” (ibid, p. 48)
fact that only animals of this size were heavy enough to leave tracks in
65

FIGURE 11. Reconstruction of the Arthropleura habitat in well-drained areas of alluvial environments with calamitaleans stands (here Calamites gigas)
under ever wet humid to seasonally dry semihumid climate conditions; in the background an eryopid amphibian, often associated with arthropleurids as
indicated by tracks and/or skeletal remains of both animals (modified after Schneider and Werneburg, 1998, fig. 15).

can be excluded because of the nature of the K-plates and the definite should therefore be “an evolutionary answer to the availability of an
subaerial habitats indicated by the above discussed trails of late juveniles unusually plentiful food niche: masses of spores and pteridosperm pol
and adults. However, it cannot be excluded that a part of the small len, perhaps also prothallia.” However, feeding on spores and pollen
Flechting High trails was produced under thin water cover, but this is appears unlikely for an animal of this size, even if they are concentrated
irrelevant to the mode of respiration. in lacustrine deposits such as cannel coals. If Arthropleura had used this
Kraus and Brauckmann (2003) interpreted Arthropleura as a thin nutrient source, lakeshores should have been most frequented by this
skinned, caterpillar-like animal, stabilized by means of antagonistic hy animal. Again, this is very improbable because Arthropleura remains are
draulics of body fluids only. This interpretation is based on fern leaves completely missing in lake deposits thus far (see Hannibal, 1997, for the
and other plant fragments impressed from below through paraterga (ibid, supposed occurrence in the lacustrine black shales of Linton, Ohio).
fig. 6). This interpretation has to be handled with care, because most if Alternatively, fructifications, megasporophylls and cm-size seeds
not all Arthropleura remains are exuvians. They exhibit unquestionably (ovules) such as Cardiocarpus and Trigonocarpus could have been an
thin cuticles – but this could be simply a result of resorption processes energy-rich food for large arthropods. The question is whether or not
during ecdysis. The common preservation of arthropleurid trails with up this food source was available year-round. The production of fructifica
to 3 cm deep impressions of the limbs as well as the commonly deep tions and seeds could have been very seasonal as the climate became
undertracks in wet sand indicate a heavy animal with a well sclerotised increasingly seasonal after the late Westphalian (Late Moscovian). Spo
exoskeleton similar to functionally comparable modern counterparts such radic mass occurrences of gymnosperm seeds such as Samaropsis and
as some chilopods (scolopenders). Cardiocarpus in distinct layers (as in the Manebach Arthropleura local
A herbivorous diet of Arthropleura seemed to be demonstrated by ity – see Barthel, 2001; here Fig. 5C), are interpreted here as the result of
the description of gut contents by Rolfe and Ingham (1967). Kraus seasonal seed production and taphonomic effects (washed together).
(2005, p. 18-20) has reinvestigated this specimen with the result that Assumed carnivory and interpretation of Arthropleura as a chilopod in
“the wooden parts were accidentally fossilized together with the Barthel and Schneider (1997) is simply an incorrect citation by Kraus
Arthropleura specimen and do not constitute any gut content.” How (2005, p. 20). Barthel and Schneider (1997, p. 198) solely compared
ever, assuming that Arthropleura was a penicillate diplodod, Kraus (2005, functionally the construction of the ventral side of Arthropleura with
p. 20) concluded in a perfectly circular argument that Arthropleura was chilopods but did not do so in terms of any phylogenetic relationship.
herbivorous, as are (nearly) all diplopods. The gigantism of arthropleurids Nevertheless, it could not be excluded that Arthropleura was predatory
66
FIGURE 12. Stratigraphic distribution of the arthropleurid track Diplichnites
cuithensis (1-6) in different environments (gr – gray facies, wr – wet red
beds) and the occurrence of Arthropleura body remains (a - ?d). Occurrences
of D. cuithensis: 1 - Visean (Asbian, Mississippian), Fife, Scotland (Pearson,
1992); 2 - Namurian (Pendleian, Early Serpukhovian), Arran, Scotland
(Briggs et al., 1979); 3 - ?Late Namurian to basal Westphalian
(?Kinderscoutian/Langsettian, Bashkirian, Early Pennsylvanian), Nova
Scotia (Calder et al., 2005); 4 - Westphalian A (Langsettian; Bashkirian,
Early Pennsylvanian), Nova Scotia (Ferguson, 1966, 1975; Briggs et al.,
1979; Falcon-Lang et al., 2006); Westphalian A to ?B (Langsettian
?Duckmantian; Bashkirian/?Moscovian, Pennsylvanian), New Brunswick
(Briggs et al., 1984); Westphalian B-C (Duckmantian-Bolsovian; Moscovian,
Middle Pennsylvanian), Nova Scotia (Ryan, 1986); 5 - Virgilian (Gzhelian,
Late Pennsylvanian), eastern Kansas (Mángano et al., 2002); Virgilian
(Gzhelian, Late Pennsylvanian), Kentucky (Martino and Greb, 2009);
Virgilian (Late Pennsylvanian), New Mexico (Lucas et al., 2005c); 6 - latest
Stephanian C/earliest Autunian (Asselian, Cisuralian, Early Permian), French
Massif Central (Briggs, 1986; Briggs and Almond 1994); Stephanian ?C
(Late Pennsylvanian/Early Permian, Gzhelian/Asselian), Northeast Germany
(Walter and Gaitzsch, 1988); Early Permian (Cisuralian), Cape John, Pictou
County, Nova Scotia, Canada (Ryan, 1986). Occurrences of Arthropleura
body remains: a - Visean (late Asbian to early Brigantian, Middle Mississippian),
Hainichen basin, Germany (Rössler and Schneider, 1997); b - Westphalian A
(Langsettian, Middle Bashkirian, Early Pennsylvanian), Ostrava-Karvina
basin, Czech Republic (eho and ehoova, 1972); c – Westphalian B
(Bashkirian, Pennsylvanian) to Lower Rotliegend (Early Asselian, Cisuralian)
of the European basins (e.g., Hahn et al., 1986; for the youngest occurrences
in the Permian (see Schneider and Barthel, 1997; Schneider and Werneburg,
1998); in North America (Hannibal, 1997): Lower Conemaugh, ?Missourian,
Kasimovian, Ohio; Desmoinesian, Moscovian, Pennsylvania; Desmoinesian,
Moscovian, Mazon Creek, Illinois; Springhill, Atokan, Bashkirian/
Moscovian, Nova Scotia; d – upper Lower Rotliegend (Sakmarian, Cisuralian),
Döhlen basin, Germany, large chitinous cuticle fragments that may belong
to Arthropleura (Schneider and Barthel, 1997).

Rotliegend (LOD = ?LAD; Asselian) (Rössler and Schneider, 1997;


Schneider and Barthel, 1997; Schneider and Werneburg, 1998). Trails are
as common in the gray facies as in the facies of wet red beds. Preferred
habitats of semi-adults and adults were open-vegetated river landscapes.
They co-occurred in these habitats since the Late Pennsylvanian with
semi-aquatic eryopid amphibians and terrestrial pelycosaur reptilians.
But, the large size of 1 m to more than 2 m length in adult stages of
Arthropleura protected it for a long time against natural enemies. Why
did it suddenly become extinct in the Early Permian?
The changes from arborescent lycopsid- to tree fern- and
cordaitalean-dominated wetlands during the Pennsylvanian through the
Early Permian (Kerp, 2000; DiMichele et al., 2001, 2009; Falcon-Lang,
2004, Pfefferkorn, 2008) had no influence on Arthropleura as shown by
the occurrences of (parautochthonous) body remains in the cordaitalean
dominated mires of the Early Permian in the Döhlen basin and Thuringian
Forest basin of Germany (Schneider and Barthel, 1997; Schneider and
Werneburg, 1998). Obviously, Arthropleura was not conditioned to any
particular kind of mire and wetland vegetation.
It thus seems more likely that environmental changes driven by
and could overcome tetrapods as large as itself, as is the case with mod broadscale climate developments may have been one of the causes of the
ern predatory chilopods such as Scolopendra. It can be speculated only disappearance of Arthropleura. The last known occurrence of
that amphibians such as the eryopids, which lived in the same environ
Arthropleura is in the “wet phase C” of Roscher and Schneider (2006)
ment, could have easily been preyed upon by this giant arthropod. How
around the Gzhelian/Asselian transition. As shown by them, during the
ever, the problem of diet remains open as long as mouthparts or true gut cyclically increasing aridization towards the Late Permian, each wet
contents of Arthropleura have not been found. phase was drier than the preceding one. Wet phase C represents the last
CONCLUSIONS extensive gray facies with coal formation in most of the European basins
(Roscher and Schneider, 2006, fig. 51a-b). The wet red beds of the fore
Well adapted to alluvial environments of every wet humid to going dry phase of the Stephanian B to early C (Early Gzhelian) as well
seasonal dry semihumid climates, Arthropleura was a common member as of the succeeding dry phase around the Middle Asselian are character
of the late Paleozoic continental biota of the paleo-equatorial biomes for ized by common calcic soils and calcretes. Arthropleura had been adapted
more than 35 Ma, from the Early Carboniferous Late Visean (FOD; to these seasonally dry environments since the Early Westphalian (see
Middle Mississippian, Asbian/Brigantian) up to the Early Permian Lower above, Little River Formation, Nova Scotia).
67
But, the intensity of seasonality increased as is indicated by in and the Döhlen basin and provided the photograph in Figure 7G. Michael
tensely laminated (varved) lake sediments, which are very typical of all Magnus (Freiberg) made the interpretation and photographs of thin sec
perennial lakes in the European basins after wet phase C (Clausing and tions (Fig. 4F). Thanks to Jean-Michel Pacaud (Paris) for casts of the
Boy, 2000; Roscher and Schneider, 2006). Intensified seasonal dryness Montceau-les-Mines arthropleurids and to Dominique Chabard (Autun)
might have contributed to the demise of Arthropleura. Local populations for the permission to investigate the originals in the collection of the
may only have survived up to wet phase D in the Sakmarian, if the up to Autun Museum. Cordial thanks to Thilo Thümmel (Paulsdorf), who
2 cm size cuticle fragments, described from the Döhlen basin by Schneider provided the Döhlen Arthropleura specimen on permanent loan to the
and Barthel (1997, p. 200, pl. 9, figs. 1 and 2), belong to this animal. TU Bergakademie Freiberg. We thank Carsten Brauckmann (Clausthal
Possibly, the Döhlen basin during wet phase D was one of the “wet Zellerfeld) and Otto Kraus (Hamburg) for long discussions on the
spots,” shrinking refuges of a formerly widespread wetland biota paleobiology of arthropleurids; for discussions on sedimentology,
(DiMichele et al., 2006). Increasingly critical prey-predator relation Sebastian Voigt, Olaf Elicki, Birgit Gaitzsch (all Freiberg), Ute Gebhardt
ships between Arthropleura and terrestrially-adapted amphibians and (Karlsruhe) and Berit Legler (London), and for discussions on paleocli
reptiles may have been an additional trigger to the demise of this giant matology as well as for drawing Figure 12, Marco Roscher (Oslo).
animal, which was the largest terrestrial arthropod of earth history. Cristopher J. Cleal (Cardiff) gave information on the Fife locality in
ACKNOWLEDGMENTS Scotland, and Jan Fischer (Freiberg) helped with literature. Thanks to
Dieter Waloszek (Ulm) for critical hints concerning the anatomy and life
For collaboration in the fieldwork in Cañon del Cobre we thank reconstruction of Arthropleura. Photographs for Figure 5C-D were de
Larry Rinehart and Josh Smith (both Albuquerque), at the Manebach livered by Silke Sekora (Freiberg), for Figure 7E by Harald Walter
locality in Germany Berthold Lugert (Manebach), Georg Sommer (Freiberg) and for Figure 7D by Michael Thiele-Bourcier (Kirkel
(Schlesuingen), Silke Sekora (Freiberg) and Harald Lützner (Jena) as well Neuhäusel). The reviewers, Bill DiMichele and Carsten Brauckmann,
as at Flechting High, Birgit Gaitzsch (Freiberg). Manfred Barthel (Ber provided helpful comments. This publication is part of the German
lin) gave valuable information on floras of the Thuringian Forest basin Science Foundation projects of J.W.S. and R.W., grants DFG SCHN408/
12 and DFG WE 2833/3.

REFERENCES

Almond, J. E., 1985, Les arthropleurides du Stephanian de Montceau-les Bulletin de la Société d´Histoire Naturelle d´Autun, v. 117, p. 141-148.
Mines, France: Bulletin de la Societe d’Histoire Naturelle Autun, v. 115, Briggs, D. E. G. and Almond, J. E., 1994, The arthropleurids from the
p. 59-60. Stephanian (Late Carboniferous) of Montceau-les-Mines (Massif Cen
Barthel, M., 1976, Die Rotliegendflora Sachsens: Abhandlungen Staatliches tral - France); in Poplin, C. and Heyler, D., eds., Quand le Massif Central
Museum für Mineralogie und Geologie Dresden, v. 24, p. 1-190. était sous l’équateur: Un écosystème Carbonifère à Montceau-les-Mines:
Barthel, M., 2001, Pflanzengruppen und Vegetationseinheiten der Manebach Mémoires de la Section des Sciences, v. 12, p. 127-135.
Formation: Beiträge zur Geologie von Thüringen, Neue Folge, v. 8, p. Briggs, D. E. G., Plint, A. G. and Pickerill, R. K., 1984, Arthropleura trails
93-123. from the Westphalian of eastern Canada: Palaeontology, v. 27, p. 843
Barthel, M., 2003, Die Rotliegendflora des Thüringer Waldes. Teil 1: 855.
Einführung und Keilblattpflanzen (Sphenophyllales): Veröffentlichungen Briggs, D. E. G., Rolfe, W. D. I. and Brannan, J., 1979, A giant myriapod
Naturhistorisches Museum Schleusingen, v. 18, p. 3-16. trail from the Namurian of Arran, Scotland: Palaeontology, v. 22, p.
Barthel, M., 2004, Die Rotliegendflora des Thüringer Waldes. Teil 2: 273-291.
Calamiten und Lepidophyten: Veröffentlichungen Naturhistorisches Buatois, L. A. and Mángano, M. G., 2007, Invertebrate ichnology of conti
Museum Schleusingen, v. 19, p. 19-48. nental freshwater environments; in Miller, W., III, ed., Trace fossils:
Barthel, M., 2005, Die Rotliegendflora des Thüringer Waldes. Teil 3: Farne: Concepts, problems, prospects: Amsterdam, Elsevier, p. 285-323.
Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 20, p. Buatois, L. A. and Mángano, M. G., 2009, Applications of ichnology in
27-56. lacustrine sequence stratigraphy: Potential and limitations:
Barthel, M., 2006, Die Rotliegendflora des Thüringer Waldes. Teil 4: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 272, p. 127
Farnsamer und Farnlaub unbekannter taxonomischer Stellung: 142.
Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 21, p. Buatois, L. A., Mángano, M. G., Genise, J. F. and Taylor, T. N., 1998, The
33-72. ichnologic record of the invertebrate invasion of nonmarine ecosys
Barthel, M., 2007, Die Rotliegendflora des Thüringer Waldes Teil 5: tems: Evolutionary trends in ecospace utilization, environmental ex
Gingkophyten, Coniferophyten: Veröffentlichungen Naturhistorisches pansion, and behavioral complexity: Palaios, v. 13, p. 217-240.
Museum Schleusingen, v. 22, p. 17-43. Calder, J. H., Rygel, M. C., Ryan, R. J., Falcon-Lang, H. J. and Herbert, B.
Barthel, M., 2008, Die Rotliegendflora des Thüringer Waldes Teil 6: Wurzeln L., 2005, Stratigraphy and sedimentology of Early Pennsylvanian red
und fertile Organe. Algen und Bakterien. Vegetation: Veröffentlichungen beds at Lower Cove, Nova Scotia, Canada: The Little River Formation
Naturhistorisches Museum Schleusingen, v. 23, p. 39-62. with redefinition of the Joggins Formation: Atlantic Geology, v. 41, p.
Barthel, M. and Rößler, R., 1995, “Eine ganz unbekanndte Frucht.” 300 143-167.
Jahre paläobotanisches Sammeln und Beobachten in Manebach: Calman, W. T., 1915, On Arthropleura moyseyin. sp., from the Coal
Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 10, p. Measures of Derbyshire: Geological Magazine, v. 51, p. 541-544.
49-56. Clausing, A. and Boy, J. A., 2000, Lamination and primary production in
Barthel, M. and Rößler, R., 1996, Paläontologische Fundschichten im fossil lakes: Relationship to palaeoclimate in the Carboniferous-Per
Rotliegend von Manebach (Thür. Wald) mit Calamites gigas mian transition; in Hart, M.B., ed., Climates: Past and present. Geologi
(Sphenophyta): Veröffentlichungen Naturhistorisches Museum cal Society, London, Special Publication 181, p. 5-16.
Schleusingen, v. 11, p. 3-21. Cotton, W. D., Hunt, A. P. and Cotton, J. E., 1995, Paleozoic vertebrate
Brauckmann, C., Gröning, E. and Thiele-Bourcier, M., 1997, Kopf- und tracksites in eastern North America: New Mexico Museum of Natural
Schwanz-Region von Arthropleura armata Jordan, 1854 (Arthropoda; History and Science, Bulletin 6, p. 189-211.
Ober-Karbon): Geologica et Palaeontologica, v. 31, p. 179-192. DiMichele, W.A. and Chaney, D.S., 2005, Pennsylvanian-Permian fossil
Briggs, D. E. G., 1986, Walking trails of the giant arthropod Arthropleura: floras from the Cutler Group, Cañon del Cobre and Arroyo del Agua
68
areas, in northern New Mexico: New Mexico Museum of Natural His- und seine Verbreitung in den europäischen Steinkohlenbecken: Glückauf,
tory and Science, Bulletin 31, p. 27-33. v. 72, p. 965-975.
DiMichele, W. A., Pfefferkorn, H. W. and Gastaldo, R. A., 2001, Response Guthörl, P., 1940, Zur Arthropoden-Fauna des Karbons und Perms 12.
of Late Carboniferous and Early Permian plant communities to climate Insekten- und Arthropleura-Reste aus der Tiefbohrung 38 (Hangard) bei
change: Annual Review of Earth and Planetary Sciences, v. 29, p. 461- Neunkirchen-Saar: Paläontologische Zeitschrift, v. 22, p. 109-119.
487. Hahn, G., Hahn, R. and Brauckmann, C., 1986, Zur Kenntnis von
DiMichele, W. A., Chaney, D. S., Kerp, H. and Lucas, S. G., 2010, Late Arthropleura (Myriapoda; Ober-Karbon): Geologica et Palaeontologica,
Pennsylvanian floras in western equatorial Pangaea, Cañon del Cobre, v. 20, p. 125-137.
New Mexico: New Mexico Museum of Natural History and Science, Hannibal, J.T., 1997, Remains of Arthropleura, a gigantic myriapod arthro
Bulletin, this volume. pod, from the Pennsylvanian of Ohio and Pennsylvania: Kirtlandia, v.
DiMichele, W.A., Tabor, N.J., Chaney, D.S. and Nelson, W.J., 2006, From 50, p. 1-9.
wetlands to wet spots: Environmental tracking and the fate of Carbon Haubold, H. and Katzung, G., 1978, Palaeoecology and palaeoenvironments
iferous elements in Early Permian tropical floras: Geological Society of of tetrapod footprints from the Rotliegend (Lower Permian) of Central
America, Special Paper 399, p. 223-248. Europe: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 23, p.
DiMichele, W.A., Montañez, I. P., Poulsen, C. J. and Tabor, N. J., 2009, 307-323.
Climate and vegetational regime shifts in the late Paleozoic ice age Haubold, H. and Lucas, S. G., 2001, Early Permian tetrapod tracks - preser
earth: Geobiology, v. 7, p. 200-226. vation, taxonomy, and Euramerian distribution: Natura Bresciana, v. 25,
Doubinger, J., 1994, Spores et pollen du basin Carbonifère et Permien de p. 347-354.
Blanzy-Montceau (Massif Central – France); in Poplin, C. and Heyler, Hausse, R., 1902, Ein Massengrab von Sauriern im Unter-Rotliegenden des
D., eds., Quand le Massif Central était sous l’équateur: Un écosystème Döhlener Kohlenbecken im Plauenschen Grunde bei Dresden: Jahrbuch
Carbonifère à Montceau-les-Mines: Mémoires de la Section des Sciences, für das Berg- und Hüttenwesen in Sachsen, v. 1902, p. 25-50.
v. 12, p. 61-72. Hausse, R., ed., 1910, Fossile Tierfährten im Unterrotliegenden des
Eberth, D. A. and Miall, A. D., 1991, Stratigraphy, sedimentology and Steinkohlenbecken im Plauenschen Grunde (des Döhlener Beckens) bei
evolution of a vertebrate-bearing, braided to anastomosed fluvial sys Dresden: Jahrbuch für das Berg- und Hüttenwesen in Sachsen, v. 1910, p.
tem, Cutler Formation (Permian-Pennsylvanian), north-central New 3-19.
Mexico: Sedimentary Geology, v. 72, p. 225-252. Huene, v., F., 1925, Ein neuer Pelycosaurier aus der unteren Permformation
Falcon-Lang, H. J., 2004, Pennsylvanian tropical rain forests respond to Sachsens: Geologische und Paläontologische Abhandlungen, N.F., v. 14,
glacial-interglacial rhythms: Geology, v. 32, p. 689-692. p. 5-52.
Falcon-Lang, H. J., 2006, Vegetation ecology of Early Pennsylvanian allu Hunt, A. P. and Lucas, S. G., 1992, The paleoflora of the lower Cutler
vial fan and piedmont environments in southern New Brunswick, Canada: Formation (Pennsylvanian, Desmoinesian?) in El Cobre Canyon, New
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 233, p. 34-50. Mexico, and its biochronological significance: New Mexico Geological
Falcon-Lang, H. J., Benton, M. J., Braddy, S. J. and Davies, S.J., 2006, The Society, Guidebook 43, p. 145-150.
Pennsylvanian tropical biome reconstructed from the Joggins Forma Hunt, A. P., Lucas, S. G., Lerner, A. J and Hannibal, J. T., 2004, The giant
tion of Nova Scotia: Journal of the Geological Society of London, v. Arthropleura trackway Diplichnites cuithensis from the Cutler Group
163, p. 561-576. (Upper Pennsylvanian) of New Mexico: Geological Society of America,
Ferguson, L., 1966, The recovery of some large track-bearing slabs from Abstracts with Programs, v. 36, no. 5, p. 66.
Joggins, Nova Scotia: Maritime Sediments and Atlantic Geology, v. 2, p. Jordan, F. W. H. and von Meyer, H., 1854, Ueber die Crustaceen der
128-130. Steinkohlenformation von Saarbrücken: Palaeontographica, v.4, p. 1
Ferguson, L., 1975, The Joggins section: Maritime Sediments and Atlantic 15.
Geology, v. 11, p. 69-76. Kerp, H., 2000, The modernization of landscapes during the late Paleozoic
Fracasso, M. A., 1980, Age of the Permo-Carboniferous Cutler Formation early Mesozoic: Paleontological Society Papers, v. 6, p. 79-113.
vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale Krainer, K. and Lucas, S. G., 2001, Pennsylvanian-Permian depositional
ontology, v. 54, p. 1237-1244. break in the Cutler Formation, El Cobre Canyon, New Mexico: Geologi
Gaitzsch, B., Ellenberg, J., Lützner, H. and Benek, R., 1995, Flechtinger cal Society of America, Abstracts with Programs, v. 33, no. 5, p. A5-A6.
Scholle; in Plein, E., ed., Stratigraphie von Deutschland I. Norddeutsches Kraus, O., 2005, On the structure and biology of Arthropleura species
Rotliegendbecken. Rotliegend-Monographie Teil II: Courier (Atelocerata, Diplopoda; Upper Carboniferous/Lower Permian):
Forschungsinstitut Senckenberg, v. 183, p. 84-96. Verhandlungen des naturwissenschaftlichen Vereins Hamburg, Neue Folge,
Gand, G., 1994, Les traces de vertébrés tétrapodes du Carbonifère superieur v. 41, p. 5-23.
et du Permien du bassin de Montceau-les-Mines (Massif Central – France); Kraus, O. and Brauckmann, C. 2003, Fossil giants and surviving dwarfs.
in Poplin, C. and Heyler, D., eds., Quand le Massif Central était sous Arthropleurida and Pselaphognatha (Atelocerata, Diplopoda): Charac
l’équateur: Un écosystème Carbonifère à Montceau-les-Mines. Mémoires ters, phylogenetic relationships and construction: Verhandlungen des
de la Section des Sciences, v. 12, p. 269-281. naturwissenschaftlichen Vereins Hamburg, Neue Folge, v. 40, p. 5-50.
Gastaldo, R.A., Stevanovic-Walls, I., Ware, W.N. and Greb, S.F., 2004, Kraus, W., 1993, Eine 2 m lange semiflexible Rekonstruktion von
Community heterogeneity of Early Pennsylvanian peat mires: Geology, Arthropleura, dem größten bekannten Myriapoden aus dem Oberkarbon
v. 32, p. 693-696. Mitteleuropas: Der Präparator, v. 39, p. 141-158.
Gebhardt, U., Gaitzsch, B., Schneider, J. and Rössler, R., 1995, Stratigraphy Langiaux, J., 1984, Flores et faunes des formations supérieures du Stéphanien
and facies of the Middle European continental Carboniferous and Per de Blanz-Montceau (Massif Central Francais) stratigraphie et
mian. Guide to Excursion A5: 13. International Congress on Carbonifer paléoécologie: La Physiophile, suppl. no. 100, 270 p.
ous-Permian, Krakow 1995, Supplement, 48 p. Lucas, S. G., 2005, Tetrapod ichnofacies and ichnotaxonomy: Quo vadis?
Greb, S. F., DiMichele, W. A. and Gastaldo, R. A., 2006, Evolution and Ichnos, v. 12, p. 1-6.
importance of wetlands in earth history; in Greb S.F. and DiMichele, Lucas, S. G., 2006, Global Permian tetrapod biostratigraphy and
W.A., eds., Wetlands through time: Geological Society of America, Spe biochronology; in Lucas, S. G., Cassinis, G. and Schneider, J. W., eds.,
cial Paper 399, p. 1-40. Non-marine Permian biostratigraphy and biochronology. Geological
Guthörl, P., 1935, Entdeckung und Bergung des größten, bis jetzt bekannten Society, London, Special Publication 265, p. 65-93.
Exemplars von Arthropleura armata Jordan & v. Meyer in Grube Lucas, S. G. and Krainer, K., 2005, Cutler Group (Permo-Carboniferous)
Maybach-Saar: Zeitschrift der Deutschen Geologischen Gesellschaft, v. stratigraphy, Chama basin, New Mexico: New Mexico Museum of Natu
87, p. 686-692. ral History and Science, Bulletin, v. 31, p. 90-100.
Guthörl, P., 1936, Arthropleura, der Riesengliederfüßler des Obercarbons Lucas, S. G. and Lerner, A. J., 2010, Pennsylvanian lake-margin
69
ichnoassemblage from Cañon del Cobre, Rio Arriba County, New Mexico: Permian climatic development of central Europe in a regional and glo
New Mexico Museum of Natural History and Science Bulletin, this vol bal context; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds., Non
ume. marine Permian chronology and correlation: Geological Society, Lon
Lucas, S. G., Rinehart, L. F. and Hunt, A. P., 2005a, Eryops (Amphibia, don, Special Publication 265, p. 95-136.
Temnospondyli) from the Upper Pennsylvanian of El Cobre Canyon, Rößler, R. and Barthel, M., 1998, Rotliegend taphocoenoses preservation
New Mexico: New Mexico Museum of Natural History and Science, favoured by rhyolithic explosive volcanism: Freiberger Forschungshefte,
Bulletin 31, p. 118-120. v. C474, p. 59-101.
Lucas, S. G., Harris, S. K., Spielmann, J. A., Berman, D. S. and Henrici, A. C., Rößler, R. and Schneider, J., 1997, Eine bemerkenswerte Paläobiocoenose
2005b, Vertebrate biostratigraphy and biochronology of the Pennsylva im Unterkarbon Mitteleuropas - Fossilführung und Paläoenvironment
nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico: der Hainichen-Subgruppe (Erzgebirge-Becken): Veröffentlichungen des
New Mexico Museum of Natural History and Science, Bulletin 31, p. Museum für Naturkunde Chemnitz, v. 20, p. 5-44.
128-139. Ryan, R. J., 1986, Fossil myriapod trails in the Permo-Carboniferous strata
Lucas, S. G., Lerner, A. J., Hannibal, J. T., Hunt, A. P. and Schneider, J. W., of northern Nova Scotia, Canada: Maritime Sediments and Atlantic
2005c, Trackway of a giant Arthropleura from the Upper Pennsylva Geology, v. 22, p. 156-161.
nian of El Cobre Canyon, New Mexico: New Mexico Geological Society, Ryan, R. J., Boehner, R. C. and Calder, J. H., 1991, Lithostratigraphic
Guidebook 56, p. 279-282. revision of the Upper Carboniferous to Lower Permian strata in the
Lützner, H., 2001, Sedimentologie der Manebach-Formation in den Cumberland basin, Nova Scotia and the regional implications for the
fossilführenden Aufschlüssen bei Manebach: Beiträge zur Geologie von Maritime Basin in Atlantic Canada: Bulletin of the Canadian Society of
Thüringen, N.F., v. 8, p. 67-91. Petroleum Geology, v. 39, p. 289-314.
Lützner, H., Littmann, S., Mädler, J., Romer, R. L. and Schneider, J. W., Schneider, J., 1994, Environment, biotas and taphonomy of the lacustrine
2007, Stratigraphic and radiometric age data for the continental Niederhäslich Limestone, Döhlen Basin, Germany: Transactions of the
Permocarboniferous reference-section Thüringer-Wald, Germany: Royal Society of Edinburgh, Earth Science, v. 84, p. 453-464.
Procedings XVth International Congress on Carboniferous and Permian Schneider, J., 1996. Biostratigraphie des kontinentalen Oberkarbon und
Stratigraphy, 2003, p. 161-174. Perm im Thüringer Wald, SW-Saale-Senke - Stand und Probleme: Beiträge
Mángano, M. G., Buatois, L. A., West, R. R. and Maples, C. G., 2002, zur Geologie von Thüringen, N.F., v. 3, p. 121-151.
Ichnology of a Pennsylvanian equatorial tidal flat - the Stull Shale Schneider, J. and Barthel, M., 1997, Eine Taphocoenose mit Arthropleura
Member at Waverly, eastern Kansas: Kansas Geological Survey, Bulletin (Arthropoda) aus dem Rotliegend (Unterperm) des Döhlen-Becken (Elbe
245, p. 1-133. Zone, Sachsen): Freiberger Forschungsheft, v. C466, p. 183-223.
Marks, W. J., Marks, R. I. and Pompa, A. M., 1998, Problematic tracks in Schneider, J. W. and Hoffman, U., 2001, Jungpaläozoikum der Döhlener
the Casselman Formation of Cambria County: Pennsylvania Geology, v. Senke; in Alexovsky, W., Schneider, J. W., Tröger, K. A. and Wolf, L.,
29, no. 2-3, p. 2-6. eds., Geologische Karte des Freistaates Sachsen 1:25.000, Erläuterungen
Martino, L. R. and Greb, S. G., 2009, Walking trails of the giant terrestrial zu Blatt 4948 Dresden, p. 15-40.
arthropod Arthropleura from the Upper Carboniferous of Kentucky: Schneider, J. W. and Werneburg, R., 1998, Arthropleura und Diplopoda
Journal of Paleontology, v. 83, p. 140-146. (Arthropoda) aus dem Unter-Rotliegend (Unter-Perm, Assel) des
Menning, M., Aleseev, A. S., Chuvashov, B. I., Favydov, V.I., Devuyst, F. Thüringer Waldes (Südwest-Saale-Senke): Veröffentlichungen
X., Forke, H. C., Grunt, T. A., Hance, L., Heckel, P. H., Izokh, Jin, Y. Naturhistorisches Museum Schleusingen, v. 13, p. 19-36.
G., Jones, P., Kotlyar, G. V., Kozur, H. W., Nemyrovska, T.I., Schneider, Schneider, J. W. and Werneburg, R., 2006, Insect biostratigraphy of the
J. W., Wang, X.-D., Weddige, K., Weyer, D. and Work, D. M., 2006, European Late Carboniferous and Early Permian; in Lucas, S.G., Cassinis,
Global time scale and regional stratigraphic reference scales of Central G. and Schneider J.W. eds., Non-marine Permian biostratigraphy and
and West Europe, East Europe, Tethys, South China, and North America biochronology. Geological Society, London, Special Publication 265, p.
as used in the Devonian-Carboniferous-Permian Correlation Chart 2003 325-336.
(DCP 2003): Palaeogeography, Palaeoclimatology, Palaeoecology, v. Schneider, J. W., Hampe, O. and Soler-Gijon, R., 2000, The Late Carbonif
240, p. 318-372. erous and Permian aquatic vertebrate zonation in Southern Spain and
Montañez, I. P., Tabor, N. J., Niemeier, D., DiMichele, W.A., Frank, T. D., German basins. IGCP 328. Final Report: Courier Forschungsinstitut
Fielding, C. R. and Isbell, J. L., 2007, CO2-forced climate and vegetation Senckenberg, p. 543-561.
instability during late Paleozoic deglaciations: Science, v. 315, p. 87-91. Schneider, J. W., Hoth, K., Gaitzsch, B. G., Berger, H.-J., Steinborn, H.,
Pearson, P. N., 1992, Walking traces of the giant myriapod Arthropleura Walter, H. and Zeidler, M., 2005, Carboniferous stratigraphy and devel
from the Strathclyde Group (Lower Carboniferous) of Fife: Scottish opment of the Erzgebirge Basin, East Germany: Zeitschrift der Deutschen
Journal of Geology, v. 28, p. 127-133. Gesellschaft für Geowissenschaften, v. 156, p. 431-466.
Pfefferkorn, H.W., Gastaldo, R.A., DiMichele, W.A. and Phillips, T.L., Schneider, J. W., Körner, F., Roscher, M. and Kroner, U., 2006, Permian
2008, Pennsylvanian tropical floras of the United States as a record of climate development in the northern peri-Tethys area – the Lodève
changing climate: Geological Society of America, Special Paper 441, p. basin, French Massif Central, compared in a European and global con
305-316. text: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 240, p.
eho, F. and ehoova, M., 1972, Makrofauna uhlonosného karbonu 161-183.
eskoslovenské asti hornoslezské pánve: Profil, v. 1972, p. 1-130. Secretan, S., 1980. Les arthropodes du Stéphanien de Montceau-les-Mines:
Rolfe, W. D. I., 1969, Arthropleurida; in Moore, R. C., ed., Treatise on Bulletin Trimestriel de la Société d’Histoire Naturelle et des Amis du
invertebrate paleontology: Geological Society of America, University Muséum d’Autun, v. 94, p. 23-35.
of Kansas Press., Part R, Arthropoda, v.4, p. 607-620. Smith, C. T., Budding, A. J. and Pitrat, C. W., 1961, Geology of the south
Rolfe, W.D.I., 1980, Early invertebrate terrestrial faunas; in Panchen, A.L., eastern part of the Chama Basin: New Mexico Bureau of Mines and
ed., The terrestrial environment and the origin of land vertebrates: New Mineral Resources, Bulletin 75, 57 p.
York, Academic Press, p. 117-157. Van Allen, H. E. K., Calder, J. H. and Hunt, A. P., 2005, The trackway record
Rolfe, W. D. I., 1985, Aspects of the Carboniferous terrestrial arthropod of a tetrapod community in a walchian conifer forest from the Permo
community: Comptes Rendus IX Congrès International de Stratigraphie Carboniferous of Nova Scotia: New Mexico Museum of Natural History
et de Géologie de Carbonifère, p. 303-316. and Science, Bulletin 30, p. 322-332.
Rolfe, W. D. I. and Ingham, J. K., 1967, Limb structure, affinity and diet of Voigt, S., 2005, Die Tetrapodenichnofauna des kontinentalen Oberkarbon
the Carboniferous “centipede” Arthropleura: Scottish Journal of Geol und Perm im Thüringer Wald - Ichnotaxonomie, Paläoökologie und
ogy, v. 3, p. 118-124. Biostratigraphie: Cuvillier Verlag, Göttingen, 305 p.
Roscher, M. and Schneider, J. W., 2006, Early Pennsylvanian to Late Walter, H. and Gaitzsch, B., 1988, Beiträge zur Ichnologie limnisch
70
terrestrischer Sedimentationsräume. Teil II: Diplichnites minimus n. (Onchiodon) aus dem Rotliegend (Unter-Perm) des Thüringer Waldes:
ichnosp. aus dem Permosiles des Flechtinger Höhenzuges: Freiberger Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 22, p. 3
Forschungsheft, v. C427, p. 73-84. 40.
Werneburg, R., 1997, Ein See-Profil aus dem Unter-Rotliegend (Unter Werneburg, R., Lucas, S. G., Schneider, J. W. and Rinehart, L.F., 2010, First
Perm) von Manebach (Thüringer Wald): Veröffentlichungen Pennsylvanian Eryops (Temnospondyli) and its Permian records from
Naturhistorisches Museum Schleusingen, v. 12, p. 63-67. New Mexico: New Mexico Museum of Natural History and Science,
Werneburg, R., 1987, Schädelrest eines sehr großwüchsigen Eryopiden (Am Bulletin, this volume.
phibia) aus dem Unterrotliegenden (Unterperm) des Thüringer Waldes: Wilson, H. M., 2003, Functional morphology of locomotion in the giant
Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 2, p. 52 Paleozoic millipede Arthropleura: Insights from trace fossils and kine
56. matics of locomotion in extant millipedes: Geological Society of America
Werneburg, R., 2007, Der “Manebacher Saurier” – ein neuer großer Eryopide Abstracts with Programs, v. 35, no. 6, p. 538.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

71
PALYNOLOGICAL INVESTIGATION OF THE UPPER PENNSYLVANIAN
(CARBONIFEROUS) ELCOBRE CANYONFORMATION, CUTLERGROUP,
CAÑON DEL COBRE, RIO ARRIBA COUNTY, NEWMEXICO, U.S.A.

JOHNUTTING1 AND SPENCER G. LUCAS2


1GeologicalSurvey of Canada, Natural Resources Canada, 3303 33rd St. NW, Calgary, Alberta, T2L 2A7, Canada;
2 New Mexico Museum of Natural History, 1801 Mountain Road NW, Albuquerque, New Mexico 87104-1375, U.S.A.

Abstract—Palynomorphs from the middle part of the El Cobre Canyon Formation in Cañon del Cobre of Rio
Arriba County, New Mexico, come from a 2 m interval of lacustrine shale of Late Pennsylvanian (Carboniferous)
age, not older than Stephanian B. The thermal maturity is low.

INTRODUCTION LOCALITIES
The Pennsylvanian (Carboniferous)-Permian Cutler Group red In the following text, the abbreviation GSC refers to sample num
beds in the Chama basin of northern New Mexico (Fig. 1) are fluvial bers of the Geological Survey of Canada, Calgary, and NMMNH refers
strata that were deposited along the flanks of the Uncompaghre uplift, to sample numbers of the New Mexico Museum of Natural History,
one of the basement-cored positive areas of the ancestral Rocky Moun Albuquerque. For the location of the latter and further details on the
tains. The Cutler Group was deposited in western Pangea in the equato paleofloras the reader is referred to Hunt and Lucas (1992) and DiMichele
rial tropics under a relatively wet, but seasonally dry paleoclimate and Chaney (2005). Four samples were collected for palynomorphs
(Schneider et al., this volume). It yields a diversity of nonmarine fossils, from lacustrine shale in the middle part of the El Cobre Canyon Forma
including vertebrates and plants. DiMichele and Chaney (2005) sug tion in Cañon del Cobre.
gested that the latter were typical of wetland floodplains of the Late From lower to upper the samples are: GSC, C-430711 (100 m),
Pennsylvanian (Carboniferous). The Cutler Group strata are divided NMMNH locality 7827 at UTM zone 13,377751E, 4016075N (NAD
into two mappable lithostratigraphic units, the El Cobre Canyon and 27): GSC, C-430708 (135 m), NMMNH locality 4563, at UTM zone
overlying Arroyo del Agua formations (Lucas and Krainer, 2005). The El 13,382564E, 4021118N (NAD27):GSC, C-430709 (136 m) and GSC,C
Cobre Canyon Formation is up to 500 m thick, but only 213 m are 430710 (137 m), NMMNH locality 7826 at UTM zone 13, 378840E,
exposed in the study area, where they comprise brown siltstone, sand 4016443N (NAD27). Samples GSC, C-430708, GSC,C-430709, GSC,C
stone and extraformational conglomerate of an ephemeral braided stream 430710 from a 2 m shale bed were productive and contained well-pre
environment. The conglomerate overlies Proterozoic basement in the served miospores, whereas a stratigraphically lower sample (GSC, C
subsurface and is conformably overlain by 143 m of the Arroyo del Agua 430711) contained only woody and coaly fragments. The good preserva
Formation. tion of miospores in the productive samples suggests conditions at the
The aim of the study was to determine the age and thermal matu sediment/ water interface were neither severely oxic nor anoxic. The
rity of the El Cobre Canyon Formation based on palynomorphs col Thermal Alteration Index of the samples is TAI 2, approximately equiva
lected in the Cañon del Cobre of Rio Arriba County, New Mexico. lent to a vitrinite reflectance of Ro 0.55%, which is in the “oil window”
(Utting et al., 1989).

IDENTIFICATIONS
Because the three productive samples come from a 2 m shale bed
in the middle part of the El Cobre Canyon Formation, identifications are
given as a single composite list. Selected specimens are illustrated on
Figure 2.
Triletes
Anacanthotriletes paucispinosus Ravn 1986, Apiculatisporis
variocorneus Sullivan 1964, Calamospora parva Guennel 1958,
Colatisporites sp. cf. C. decorus Bharadwaj and Venkatachala comb. nov.
emend. Williams 1973, Convolutispora jugosa Smith and Butterworth
1967, C. varicosa Butterworth and Williams 1958 (abundant),
Cyclogranisporites aureus (Loose) Potonié and Kremp 1955,
Dictyotriletes sp., Lophotriletes commissuralis (Kosanke) Potonié and
Kremp 1955, Microreticulatisporites nobilis (Wicher) Knox 1950,
Pilosisporites williamsii Ravn 1986, Punctatisporites glaber (Naumova)
Playford 1962, Reticulatisporites carnosus (Knox) Neves 1964,
Reticulatisporites reticulatus (Ibrahim) Ibrahim 1933, Savitrisporites nux
(Butterworth and Williams) Sullivan emend. Smith and Butterworth 1967,
and Verrucisporites morulatus (Knox) emend. Smith and Butterworth
1967.

FIGURE 1. Location map of Cañon del Cobre in Rio Arriba County, New Monoletes
Mexico. Latosporites globosus (Schemel) Potonié and Kremp 1955.
73
Saccites Red Tanks Member of the Bursum Formation of Carrizo Arroyo, New
Mexico, dated as Virgilian by Utting et al. (2004). Both units contain
Monosaccites: Florinites pumicosus (Ibrahim) Schopf, Wilson trilete and monolete spores, monosaccate pollen, taeniate bisaccate and
and Bentall 1944, Florinites similis Kosanke 1950, Florinites sp.,
non-taeniate bisaccate pollen. For example, taxa that occur in both the El
Potonieisporites bhardwajii Remy and Remy, Potonieisporites elegans Cobre Canyon Formation and the Red Tanks Member include
(Wilson and Kosanke) Wilson and Venkatachala 1964, and Potonieisporites Colatisporites sp. cf. C. decorus, Florinites pumicosus, Florinites similis,
bhardwajii/novicus complex. Illinites unicus, Schopfipollenites ellipsoides, Potonieisporites elegans
Bisaccites (taeniate): Illinites unicus Kosanke emend. and Potonieisporites bhardwajii/novicus complex. However, apparently
Helby1966, and Protohaploxypinus sp. lacking in the El Cobre Canyon assemblage are Alisporites zapfei,
Bisaccites (non-taeniate): Pityosporites sp. Grandispora sp. A and Grandispora sp. B (illustrated but not described
Plicates by Utting et al., 2004), Vesicaspora schaubergeri, Platysaccus saarensi,
Striatoabieites richteri, S. elongatus, Striatopodocarpites sp.,
Schopfipollenites ellipsoides (Ibrahim) Potonié and Kremp 1954. Limitisporites monstruosus and Vittatina costabilis. Based on the occur
No evidence was found of any marine palynomorphs such as rence of Vittatina costabilis the El Cobre Canyon assemblage is slightly
older than the Red Tanks assemblage dated as Virgilian.
acanthomorph acritarchs or scolecodonts. Precise comparison of the El Cobre Canyon assemblage with the
AGE spore zones of Western Europe (Clayton et al. 1977; Owens, 1996) and
the North Sea (McLean et al., 2005) is difficult because many of the taxa
Although age determination based on a few productive samples on which these zones are based are apparently not present in the New
from a single 2 m thick unit may be hazardous, it is nevertheless possible Mexico material, although further work may indicate further taxa in
to make some general observations. The El Cobre Canyon assemblage common. However, they may prove to have different vertical distribu
contains common trilete spores, but also contains monosaccate, taeniate tions. For example, Savitrisporites nux, common in the El Cobre Canyon
bisaccate and non-taeniate bisaccate pollen. Based on palynological events Formation, has its range top in the late Bolsovian (Westphalian C = late
in western Europe and North America (Clayton et al., 1977; Owens, Atokan). It is not clear why there is such a discrepancy in the range top
1996) the appearance of taeniate pollen indicates an age no older than of Savitrisporites nux, although it could well be due to climatic differ
Stephanian B (mid-Missourian). The assemblage is similar to that of the ences.

REFERENCES

Clayton, G., Coquel, R., Doubinger, J., Gueinn, K.J., Loboziak, S., Owens, B. tratigraphy of the North Sea; in Collinson, J. D., Evans, D. J. and
and Streel, M., 1977, Carboniferous miospores of western Europe: Illus Holliday, N. S., eds., Occasional Publications Series of the Yorkshire
tration and zonation: Report of Commission Internationale de Microflore Geological Society, v. 7, p. 13-24.
du Paléozoïque/Working Group on Carboniferous Stratigraphical Pa Owens, B., 1996, Chapter 18D. Upper Carboniferous spores and pollen; in
lynology; Mededelingen Rocks Geologische Dienst, v. 29, 71 p. Jansonius, J. and McGregor, D.C., eds., Palynology principles and appli
DiMichele, W.A. and Chaney, D.C., 2005, Pennsylvanian-Permian fossil cations: American Association of Stratigraphic Palynologists Founda
floras from the Cutler Group, Cañon del Cobre and Arroyo del Agua tion, v. 2, p. 597-606.
areas, in northern New Mexico: New Mexico Museum of Natural His- Smith, A. H. V. and Butterworth, M. A., 1967, Miospores in the coal seams
tory and Science, Bulletin 31, p. 26-33. of the Carboniferous of Great Britain: Palaeontology, Special Paper 1,
Hunt, A.P. and Lucas, S.G., 1992, The paleoflora of the lower Cutler Forma 324 p.
tion (Pennsylvanian,Desmoinesian?) in El Cobre Canyon, New Mexico, Utting, J., Hartkopf-Fröder, C., Lucas, S. G. and Traverse, A., 2004, Pa
and its biochronological significance: New Mexico Geological Society, lynological investigation of the Upper Pennsylvanian Red Tanks Mem
Guidebook 43, p. 145-150. ber, Bursum Formation, Carrizo Arroyo, New Mexico, U.S.A: New Mexico
Lucas, S. G. and Krainer, K., 2005, Stratigraphy and correlation of the Museum of Natural History and Science, Bulletin 25, p. 89-96.
Permo-Carboniferous Cutler Group, Chama basin, New Mexico: New Utting, J., Goodarzi, F., Dougherty, B. J. and Henderson, C. M., 1989,
Mexico Geological Society, Guidebook 56, p. 145-159. Thermal maturity of Carboniferous and Permian rocks of the Sverdrup
McLean, D, Owens, B. and Neves, R., 2005, Carboniferous miospore bios Basin, Canadian Arctic Archipelago: Geological Survey of Canada, Paper
89-19, 20 p.

FIGURE 2. Selected palynomorphs from the El Cobre Canyon Formation, Cañon del Cobre, New Mexico. The magnification of all illustrated specimens
is x 500, unless otherwise stated. In the explanation of figures, the species name is followed by the GSC locality number (preceded by the letter C), the slide
number, stage co-ordinates (in mm), and the GSC type number. All specimens are in the collections of the Geological Survey of Canada, 601 Booth St.,
Ottawa, Ontario, Canada. The stage co-ordinates (in mm) are from a Sensor Control Display digital readout with the zero point of both X and Y axes set
at the upper right hand corner of a 76 x 25 mm glass microscope slide (label on left hand side). 1. Anacanthotriletes paucispinosus C-430710,3,
3,41.1x13.5,GSC 133030. 2. Apiculatisporis variocorneus C-430710,3,-38.1x15.8, GSC 133031. 3. Convolutispora jugosa C-430710,3,-3,27.7x18.9,
GSC 133032. 4. Convolutispora varicosa C-430710,3,-20.7x6.9, GSC 133033. 5. Cyclogranisporites aureus C-430710,3,-30.0x9.2, GSC 133034. 6.
Lophotriletes commissuralis C-430708,3, -39.4x8.4, GSC 133035.7. Microreticulatisporites nobilis C-439710,3,-38.3x8.4, GSC 133036. 8. Pilosisporites
williamsii C-439708,3,-15.1x7.3, GSC 133037.9. Punctatisporites glaber C-430708,3,-17x5.5, GSC133038. 10. Pustulatisporites pustulatus C-430710,3,
28.7x9.3, GSC 133039.11. Raistrickia aculeata C-430708,3, -28.7x7.2, GSC 133040.12. Reticulatisporites carnosus C-430710,3, -23.2x9.25, GSC
133041. 13. Reticulatisporites reticulatus C-430710,4,-18.3x15.9,GSC 133042. 14. Dictyotriletes sp. C-430708,4,-25.9x4.7 (x250), GSC 133043. 15.
Savitrisporites nux C-430708,3,-24.0x7.09,GSC 133044. 16.Verrucosisporites morulatus C-430708,4,-17.9x7.9, GSC 133045. 17. Schopfipollenites
ellipsoides C-43010,4,-40.4x12.1,(x250),GSC 133046. 18. Latosporites globosus C-430708,3,-29.2x7.4, GSC 133047. 19. Florinites pumicosus C
430708,(4)-24.9x4.6,(x 250), GSC 133048. 20. Florinites similis C-430710,4,-19.7x15.1,(x250), GSC 133049. 21. Florinites sp. C-430710,4,-15.7x14.7,
(x250) GSC 133050. 22. Potonieisporites elegans C-430710,3,-13.9x18.6, GSC 133051. 23. Pityosporites sp. C-430710, 3,-14.7x18.5, GSC 133052. 24.
Pityosporites sp. C-430710,4,-36.5x13.7,(x250), GSC 133053. 25. Platysaccus sp. C-430708,3,-44.6x6.6, GSC 133054. 26. Protohaploxypinus sp. C
430710,3,-12.9x4.8,(x250), GSC 133055. 27. Illinites unicus C-430710, 3,-13.2x6.39, GSC 133056.
74
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

75
LATE PENNSYLVANIANFLORASIN WESTERN
EQUATORIALPANGEA, CAÑON DELCOBRE, NEWMEXICO

WILLIAMA. DIMICHELE1, DANS. CHANEY1, HANS KERP2 AND SPENCERG. LUCAS3


1 Department of Paleobiology, NMNH Smithsonian Institution, Washington, DC 20560;

2 Forschungsstelle für Paläobotanik, Institut für Geologie und Paläontologie, Westfälische Wilhelms-Universität, Münster, GERMANY;
3 New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, NM 87104

Abstract—Fossil plants were collected from five stratigraphic levels in Cañon del Cobre, New Mexico. These
floras are compositionally typical of the Late Pennsylvanian, particularly of the later part, the Virgilian (Gzhelian).
Conspicuous elements are Alethopteris zeilleri, Macroneuropteris scheuchzeri, Danaeites emersonii, Sphenophyllum
verticillatum, Sphenophyllum oblongifolium, and Sphenophyllum angustifolium, Annularia carinata, Asterophyllites
equisetiformis, Sigillaria brardii, and other, less common forms typical of wetland floras, such as Pseudomariopteris
cordato-ovata. Walchian conifers and Taeniopteris were encountered rarely. For the most part, these floras appear
to have been growing on bars or along the margins of braided streams, resulting in frequent burial of plants,
particularly calamite stems, in upright position, or preservation of tangled masses of sphenophyll stems and
leaves, largely in siltstones. Finer-grained deposits, seemingly representing small lakes, were strongly dominated
by pteridosperms.

Paleozoic equatorial region in North America and Europe (Euramerica).


INTRODUCTION In most of them, the exact Carboniferous-Permian systemic boundary
The Pennsylvanian-Permian transition in the Pangean low lati cannot be located precisely due to the lack of marine beds that provide
tudes records a general change from wetland floras dominated by primi the indices needed to correlate any particular section to the global stan
tive seed plants, marattialean tree ferns, and lycopsids, to floras of sea dards and with equivalent sections in the other basins. However, each
sonally dry environments, dominated by derived seed plants, including basin records the same basic pattern of transition from dominance of the
conifers, callipterids, and other groups not represented or poorly repre wetland flora to dominance, or at least commonness, of the flora typical
sented in the wetlands (Potonié, 1893; Kerp and Fichter, 1985). These of subhumid to semi-arid climates. From west-to-east, the transition
two floras can be considered, broadly, to be distinct biomes (DiMichele generally is marked by increasingly longer persistence of wetlands, marked
et al., 2005). These distinctions reflect different environmental condi by coal beds and the relative proportions of the two biomes.
tions in the lowlands caused by differences in ambient climate (Montañez In New Mexico, this transition includes the typical change from
et al., 2007; DiMichele et al., 2008, 2009a). Oscillations between the two predominance of wetland floras to more xeromorphic floras typical of
biomes began during the Middle Pennsylvanian (e.g., Falcon-Lang et al., seasonal drought. There are, however, no indicators of persistent wet
2009), with both wetland and seasonally dry floras becoming common, climatic intervals, such as coals (lacking), and a variety of geological
at different times, in basinal lowlands during the Late Pennsylvanian indicators suggest that periods of low moisture availablility were the
(DiMichele and Aronson, 1992). The seasonally dry flora came to pre most common conditions.
dominate in the Permian equatorial lowlands. This paper focuses on floras from Cañon del Cobre, in north
The Pennsylvanian Subperiod has long been recognized as a time central New Mexico (Fig. 1). It details the composition of these floras,
of major ice accumulations in at least the southern polar regions, resulting their environmental and paleoecological significance, and their biostrati
in a pattern of regular glacial-interglacial oscillations (Wanless and Weller, graphic implications. Plant fossils have been known from Cañon del
1932; Wanless and Shepard, 1936). Recent work suggests that ice vol Cobre since the late 1800s (Fontaine, 1890), and have been recognized
ume declined significantly during the late Middle Pennsylvanian (Field explicitly as of Pennsylvanian age since the 1960s (see Hunt and Lucas,
ing et al., 2008). The Late Pennsylvanian appears to have been a period 1992). For the most part, the plants have served as somewhat ambiguous
of tropical warmth, especially the early part when derived, xeromorphic biostratigraphic markers, in some conflict with vertebrate age estimates
floras first become common in lowlands (Cridland and Morris, 1963; (Hunt and Lucas, 1992), suggesting ages from late Middle Pennsylvanian
Feldman et al., 2005). The tropical regions appear to have become gener (Desmoinesian), through earliest Permian. The purported Middle Penn
ally wetter in the later Late Pennsylvanian, at least in central and western sylvanian age has been reassessed (DiMichele and Chaney, 2005) and
Pangaea. Ice again became prominentator just before the Pennsylvanian will be reconsidered here in more detail.
Permian transition (Fielding et al., 2008; Rygel et al., 2008). These changes
in ice volume, possibly attributable to changes in atmospheric CO2 GEOLOGY
(Montañez et al., 2007), may have been major drivers of the climate The upper Paleozoic rocks of Cañon del Cobre encompass the
changes that affected terrestrial edaphic conditions and, thence, vegeta Upper Pennsylvanian to Lower Permian El Cobre Canyon Formation
tional distributions during the Pennsylvanian-Permian transition. and the Lower Permian Arroyo del Agua Formation, together making up
Because of the distinct environmental preferences of the two prin the Cutler Group (Lucas and Krainer, 2005). Eberth and Miall (1991)
cipal floras/biomes, the change from one to the other is not locally or described the El Cobre Canyon Formation as consisting of sediments
regionally useful as a biostratigraphic marker. Temporal internal changes deposited in shallow, ephemeral braided streams. Similarly, Lucas et al.
within each biome may record useful evolutionary trends. The periodic (2005) described these rocks as having formed from a “low sinuosity
(or episodic) cycling between these biomes records areal and temporal river system with locally vegetated interfluves and floodplain surfaces.”
changes in environmental conditions that favor one of these biomes over They described trace fossils, including Scoyenia burrows and trackways
the other. of arthropleurid arthropods, the latter consistent with open vegetation
Basins preserving this transition are strung out along the late on river floodplains. Also common are ferric soils, immature calicisols,
76

FIGURE 1. Index map and stratigraphic column. United States with the state of New Mexico shaded. Cañon del Cobre area enlarged to show spatial and
elevational relationship between various plant localities labeled A-E, oldest to youngest. Stratigraphic column with stratigraphic position of plant localities
(names to left of rock column and associated locality numbers, both USNM and NMMNH, to the right).

and calcareous rhizoconcretions in paleosols, in combination indicative Permian (Wolfcampian) depending on the fossil group. Palynology
of strongly seasonal climates with periodic moisture limitation (Lucas et (Utting, 2001), based on limited sampling, provides a very broad range of
al., 2005). possible ages, ranging from late Middle to Late Pennsylvanian. How
Our own observations suggest that, in the El Cobre Canyon For
ever, as the author notes, the limited numbers of samples require that
mation, paleosols commonly have petrocalcic concretions, and that the these estimates be treated with caution. Plants are quite sensitive to
calcic content increases, on average, upward through the section. In some facies-environmental variables, so attribution of age based on floristic
places, the carbonate nodules are intergrown into thick beds of nodular data can be misleading if there are strong environmental signals in the
pedogenic carbonate, several decimeters thick. These observations sug rock sequence from which the samples were drawn. Often, environment
gest that climate was strongly seasonal with evapotranspiration exceed will be the major controlling variable, and can greatly confound interpre
ing rainfall for much of the year, at least during the times of paleosol tations of age (see, for example, the debate of the age of the Dunkard
formation. The climatic signatures of these paleosols often contrasts Formation of the Appalachian region: Barlow, 1975).
strongly with that inferred from the megafossil flora preserved in chan Macrofossil plant remains from the El Cobre Canyon Formation
nel-fill deposits cut into them. Such contrasts suggest longer-term pat have been interpreted to indicate an age as old as late Desmoinesian (late
terns of climatic oscillation, with plant preservation occurring differen Middle Pennsylvanian) and as young as Virgilian (late Late Pennsylva
tially during the wetter periods.
nian) (Hunt and Lucas, 1992, and reports of other investigators noted
AGE OF THE EL COBRE CANYON FORMATION therein). The most recent analysis (DiMichele and Chaney, 2005) esti
mated the age to be late Late Pennsylvanian, probably later Virgilian. The
Biochronological age estimates of the El Cobre Canyon Formation flora is broadly consistent with European (e.g., Wagner, 1985; Doubinger
range from late Middle Pennsylvanian (Desmoinesian) through Early et al., 1995; Laveine, 1989) and American (Blake et al., 2002; DiMichele
77
et al., 2004, 2005) floras of this age in containing such long-ranging, particular taxa). A quantification can be accomplished rapidly, even with
mainly Late Pennsylvanian species as Alethopteris zeilleri, large collections. It overcomes problems of “what to count,” which is
Pseudomariopteris cordato-ovata, Sphenophyllum oblongifolium, especially problematic in Pennsylvanian collections, where some taxa
Annularia carinata, and Sigillaria brardii. It also contains more typi are known mainly as stems of various sizes, others as large leaves, others
cally later Late Pennsylvanian species, such as, Sphenophyllum mainly as isolated pinnules of larger fronds; if sample size is not too
angustifolium, Sphenophyllum verticillatum, and Odontopteris lingulata large, thus the “hand sample” prescription, the potential bias of counting
(Wagner, 1984; Castro, 2005). As such, the flora falls broadly into one large stem or leaf fragment as equal to one tiny pinnule is minimized,
Wagner’s (1984) Sphenophyllum angustifolium or Autunia conferta due both to the large number of samples and to the relatively small size
biozones, which correlate with the Stephanian C. However, it should be of a “quadrat.” In addition, empirical studies show that this method
understood that the stratigraphy of Late Pennsylvanian terrestrial de preserves the basic dominance-diversity structure of the assemblage
posits is in the process of revision at this time, particularly as delimited (Lamboy and Lesnikowska, 1988; DiMichele et al., 1991).
by plant fossils (Doubinger et al., 1995; Wagner, 1998). Given the rela Specimens were photographed digitally, mostly under natural light.
tively small, and depauperate flora described from Cañon del Cobre, a All photographs were processed in Adobe Photoshop.
well-constrained age determination cannot be made with confidence, al
though the flora, to the degree presently known, can be placed in the Late FLORA
Pennsylvanian, probably the Virgilian in American terminology (Blake et Lycopsids
al., 2002), likely equivalent to the Gzehlian and within the Stephanian B
or C in international terminology (Wagner, 1984, 1988; Heckel and Lycopsid stem remains identified as Sigillaria brardii (Fig. 4)
Clayton, 2006). consist of leaf scars with rounded upper and lower edges, sharply pointed
The vertebrate chronology of the Cutler Group is discussed by lateral angles, and distinct foliar parichnos below a small leaf trace. Ligule
Lucas et al. (2005). They note the occurrence of Desmatodon, Limnoscelis pits are not well marked. The scars are not part of well-developed leaf
and a primitive Eryops in the lower El Cobre Canyon Formation, which cushions. Scars are separated by interareas with vertically oriented lines
is consistent with a latest Pennsylvanian to earliest Permian age, and in that weakly converged around the scar areas.
conformance with earlier interpretations of the El Cobre Canyon Forma Sigillaria remains were rare in the fossil assemblages; only two
tion vertebrates as of latest Pennsylvanian age (Fracasso, 1980; Lucas, stem specimens were found. These remains were associated with cones
2002). Vertebrates from the Arroyo del Agua Formation, better known in identified as Sigillariostrobus and ribbon-like axes with central thicken
nearby Arroyo del Agua, indicate an unambiguous Early Permian age ings (midveins?) that are interpreted either as leaves or rootlets of sigillarian
(Langston, 1953; Lucas, 2002: Lucas and Krainer, 2005; Lucas et al., stigmaria (Fig. 5). We lean toward the identification of these structures as
2005). leaves, given the overall parautochthonous nature of this flora, the sedi
mentary environment (active stream channels or lake deposits) and their
MATERIALS AND METHODS relative abundance.
Fossil plants were collected from 21 sites representing five strati The presence of S. brardii is an indication of the persistence of
graphic horizons from the El Cobre Canyon Formation exposed within wetter areas within these channel settings, though sigillarians, or at least
Cañon del Cobre (Figs. 1-3). Collections were made in as unbiased a some of them, appear to have been tolerant of somewhat drier conditions
than those typical of lepidodendrids (Phillips and DiMichele, 1992;
manner as possible, meaning that all hand samples that showed any
indications of fossil plant debris were collected. The collection sizes Pfefferkorn and Wang, 2009).
from any individual excavation varied from more than 100 specimens to
Calamitales
only a few, with most sites yielding between 25 and 95 hand samples.
All collections are housed in the Paleobotanical Collections of the The calamitalean sphenopsids are represented by stem, foliage
National Museum of Natural History, Smithsonian Institution in Wash and reproductive remains (Figs. 6-10). Stem remains consist mainly of
ington, DC. Each collecting locality has been given a unique locality Calamites sp., consisting of typically ribbed stems lacking branch scars
number. All photographed specimens have been given unique collection at the nodes (Figs. 6, 7; see Fig. 6.4). These kinds of specimens generally
numbers. are described as “pith casts,” though we have challenged this conven
Specimens were identified using a hand-lens and microscope. Cu tional wisdom (DiMichele et al., 2009b), arguing instead that many ca
ticles were not prepared. During the process of identification, each col lamite stems preserved in compression preservation are, in fact, stem
lection was quantified to establish the relative abundances of the compo casts or compressed stems, comparable to modern Equisetum. Following
nent species. This was carried out using the quadrat method of Pfefferkorn Rössler and Noll (2006), we see the Calamitales as a much more diverse
et al. (1975), in which each hand specimen is treated as a sampling group structurally and taxonomically than has been appreciated, at least
quadrat, and taxa are identified as having occurred once if present on a in recent work.
quadrat/hand-sample, regardless of how large the remains or how numer Calamite foliage is attributable to Annularia carinata (Kerp, 1984;
ous (e.g., one pinnule counts the same as more than one, if all occur on the Kerp and Fichter, 1985) (Figs. 8, 9.2), possibly Annularia spicata (Fig.
same quadrat). Both sides of each specimen were considered. Counter 9.1) and Asterophyllites equisetiformis.
parts were not counted twice. This kind of analysis produces a fre Annularia carinata is characterized by leaves of unequal length,
quency distribution in which the relative abundance of a taxon is re forming ovoid whorls of varying diameter. The individual leaves are
ported as a percentage of the sample quadrats on which it was identified, spatulate in form and have mucronate tips (these specimens may also be
i.e., all taxa could, hypothetically, occur at 100% abundance. This method referred to Annularia mucronata; the taxonomy of Annularia is in need
tends to reduce the relative abundance of the dominant forms and raise of revision). Populations of this species frequently are confused for
the abundance of the rare forms, compared to point-count methods or mixed assemblages of Annularia sphenophylloides (smaller whorls) and
counts of individual plant parts (Wing and DiMichele, 1995). This hand Annularia stellata (larger whorls). Each of these latter species has sym
sample/quadrat method was chosen for several reasons. It can be applied metrical whorls formed by leaves of approximately equal length. Leaves
easily to museum collections (the basis of this study), assuming those of A. sphenophylloides are spatulate and mucronate. Large mucronate,
collections were made in an unbiased manner (meaning that an attempt spatulate leaves often have been attributed to A. stellata, but this appears
was made while collecting to preserve the proportions of specimens as to be incorrect given that the type of A. stellata has lanceolate leaves.
they were found in the excavation, not concentrating “better” material or Three forms of calamite cones occur in the El Cobre assemblage;
78

FIGURE 2. Field photos. 1, Copper mine locality with arrows indicating the east-west extent of the outcrop. 2, Closeup of the finely-laminated sediment
of USNM 42293 at the Copper Mine locality from which sparse plant remains were collected. 3, Arroyo wall with the myriapod trackway-producing
sandstone bed low along right side of wash. Arrow indicates finely laminated red micaceous beds from which fossil plants (USNM 42125 and 42126) were
recovered. 4, Closeup of plant-bearing beds above trackway-producing sandstone. 5, Large sandstone ridge into which “uranium” adits were dug. The right
arrow indicates position of the north adit (USNM 42296); left arrow indicates location of south adit (USNM 42297 and 422598). 6, Thick sandstone ledge
at the south adit; there is a small clay lens from which USNM 42297 was collected. Note quarter sheet of newspaper to the left of lense for scale.
79
along the lateral margins and curved distal margin of the leaf, so some
traverse a relatively short, arched path. Along the leaf margin, the vein
density appears to be about 35.
Pteridosperms
The most commonly encountered pteridosperm foliar remains are
Alethopteris zeilleri and Macroneuropteris scheuchzeri. All other pteri
dosperm foliage is rare, and includes Odontopteris cf. lingulata,
Odontopteris cf. brardii, Pseudomariopteris cordato-ovata, and isolated
pinnules, possibly of neuropteroid affinity.
A few specimens in the collection were identified tentatively as
Odontopteris cf. lingulata (Figs. 14.1, 14.4 & 14.5, perhaps Figures 14.1,
14.2 and along the edge of Fig. 25.5) based on pinnule shape and vena
tion. Odontopteris lingulata has been designated Mixoneuralingulata by
Wagner and Castro (1998). The specimens from Cañon del Cobre are
quite similar to Macroneuropteris scheuchzeri, and some of the ones
shown here may be. They are large, tongue-shaped with undulatory
sides and a bluntly rounded to broadly pointed apex. The midvein is well
developed, but appears to consist of a bundle of veins rather than a single
vein; it extends about two-thirds the length of the pinnule before splay
FIGURE 3. Field photo. Copper mine outcrop, view from east to west. The ing broadly into secondary veins. The secondaries are relatively widely
finely laminated beds, seen in Figure 2.2, are at the distant end of the out spaced and strong, arching steeply before curving to meet the pinnule
crop in the image. In the near ground are specimens, from the sandier more margin at about 45o. Lateral veins branch three or more times before
fluvial facies, laid out on a fallen sandstone block. reaching the pinnule margin. Like M. scheuchzeri, O. lingulata may have
smaller, rounded pinnules positioned basipetally to the large, linguloid
all are rare. The three include Palaeostachya sp. (Fig. 10), which is forms. These smaller pinnules vary in number from one to many, gener
locally common, Calamostachys sp., which is rare, and Macrostachya ally lack midveins, and have a rounded shape. We found only one pos
sp. (Figs. 9.3, 9.4), represented by two specimens. sible example of this configuration.
Sphenophylls Pinnae bearing small, triangular pinnules with broad attachment to
the rachis, multiple veins enter the base of the pinnule, and no obvious
Three species of Sphenophyllum were identified in the collections midvein were encounted rarely and always in fragmentary preservation
from Cañon del Cobre plus one additional potentially distinct form. (Figs. 14.2, 14.6). The identification of such pinnae is difficult and can
Sphenophyllum angustifolium (Figs. 11.1, 11.2, 11.4, 11.5, 11.7) not be made with confidence. The specimens in the collection resemble
consists of whorls of six narrow leaves, each with four to six veins, taxa such as Lescuropteris genuina or Odontopteris brardii.
terminating in long pointed teeth. Two veins enter the base of a leaf and Alethopteris zeilleri (Figs. 15-17) pinnules are somewhat variable
fork near the base, rarely again prior to termination. The edges of the in shape, degree of confluence with adjacent pinnules, and the degree of
leaves also can be deeply lacerate. Two pairs of leaves project at sub-90o angularity with which the pinnules join the pinna rachis. Generally
angles to the stem, and two additional leaves, one on each side, project at straight-sided with bluntly rounded apices, some pinnules may be en
nearly a 90o angle to the stem, or are slightly deflected toward the base of larged in the middle and taper gradually to a broadly triangular apex.
the stem. Pinnules may be nearly free to the base and only weakly confluent with
Sphenophyllum verticillatum (Figs. 11.3, 11.6, 12.3, 12.4, 12.5) adjacent pinnules, or the confluence may be well developed. The pinnule
leaf whorls are nearly round, containing six leaves. The leaves are wedge surface is strongly vaulted, leading to a sunken midvein, which extends
shaped and broaden greatly from a narrow base. They have arcuate distal about two-thirds of the distance from the base to the pinnule apex.
margins, with rounded, but mucronate, teeth. Generally, two veins enter Venation is dense, with veins forking first near the midvein of the pinnule
the base of the leaf and fork two to three times before terminating in a and generally once more before terminating at the margin, which they
terminal tooth. Because of the rounded distal leaf margin, the veins termi meet nearly at right angles. Pinnule shape and size can be variable within
nate at slightly different distances from the base. Along the margin of a pinna. Pinnules decrease in size, become more triangular in shape and
larger leaves there can be as many as 20 veins. R.H. Wagner suggests increasingly confluent toward the terminus of a pinna, which ends in an
comparing the specimens illustrated in Figures 11.3 and 11.6 with elongate, narrow, terminal pinnule. Note, R.H. Wagner suggests that
Parasphenophyllum Asama (Asama, 1970), in that the teeth are some these specimens bear comparison with Alethopteris virginiana Fontaine
what elongate and bluntly pointed and the distal margins are not clearly and White.
set off from the lateral margins. Macroneuropteris scheuchzeri (Figs. 19, 20, possibly Figs. 14.1
Sphenophyllum oblongifolium (Figs. 12.1, 12.6, 13.1, 13.2, 13.4, and 14.3) pinnules are generally found in isolation. They are straight
13.5) is characterized by whorls consisting of six leaves. Two opposite sided to slightly falcate, with a thin lamina and little vaulting. The pin
pairs of leaves project at sub-90o angles to the stem. Two additional nule base is shallowly cordate, with a single point of attachment to the
leaves form a basal bib, directed roughly parallel to the stem. Leaves have pinna rachis. Large pinnules may have two lobes or free smaller, round
eight to ten subparallel veins, forking generally one to two times before pinnules at their bases. These smaller pinnules also may be found in
terminating in small, pointed teeth at the leaf margin. The distal margins isolation. The pinnule midvein is strong and extends about 80% of the
of the leaves can be lacerate, in addition to toothed. distance from the pinnule base to the tip. Lateral veins are fine, dense and
A single isolated specimen is herein identified as Sphenophyllum broadly arched, reaching the pinnule margin at an angle of about 30o.
(Parasphenophyllum) cf. thonii (Fig. 13.3). The small leaf is approxi Lateral veins may fork three or more times before reaching the margin. In
mately 5 mm wide and has a distinctly rounded distal margin with small, many pinnules, small, stiff, apically directed hairs are present, most
shallow, blunt teeth. Two veins enter the base and may branch as often as commonly on the abaxial surface, and mostly along the midvein or in the
five to six times before terminating at the margin. Veins terminate all basal third of the pinna; such small, adpressed hairs are diagnostic of this
80

FIGURE 4. Sigillaria brardii. USNM 528655 from loc 42122; 1a, part; 1b, counterpart. Both 2 x.

species, in combination with its shape, size, and venation. alethopteroid in character. The midvein is straight and extends through
Pseudomariopteris cordato-ovata (Fig. 21, possibly Fig. 33.2) about three-quarters of the lamina, becoming more narrow apically
occurs rarely in the collection. Specimens consist of isolated pinnae throughout its length. The dense lateral veins arch upward from the
bearing small, rounded to triangular pinnules with acuminate apices, and midvein but shortly become subhorizontal, forking two to three times
constricted, but still broadly attached bases. The pinnules are distinctly before reaching the margin at a high angle, nearly perpendicularly in some
inclined toward the pinna apex. Some pinnules may have basal lobes. specimens.
Venation is obscure, probably due to a thick pinnule lamina, which ap Axes most likely attributable to pteridosperms were found at
pears to be slightly vaulted, with a thickened or inrolled margin. A midvein most collecting localities. The axes varied from striate to striate and
is present, mainly in the lower part of the pinnule. Secondary veins are punctuate (Fig. 33.3) to variously smooth. Others were ribbon like and
widely spaced and thin, arching steeply through the pinnule lamina. resembled stigmarian rootlets (Figs. 33.1, 33.4, 33.5), though their affin
Importantly, the terminal portion of the pinnule ends in a spine-like ity could not be established. Seeds of indeterminate affinity were com
prolongation, characteristic of plants in this genus, which is thought be a mon at certain sites. These large, platyspermic, ovoid to cordate seeds
vine (Krings and Kerp, 2000; Krings et al., 2003). (Figs. 34.1-34.7) are of indeterminate affinity; they may be cordaitalean
A single, isolated pinnule was encounted that may be assignable to or attributable to one of the pteridosperms, though the latter option
Neurocallipteris cf. planchardii (Fig. 19.3) (see Cleal et al., 1990). The seems unlikely. Only a single ribbed, likely pteridosperm seed was found
specimen is auriculate, otherwise straight sided, with a broad, rounded (Fig. 34.8)
apex, wide lamina relative to length, and a relatively flat base. The midvein
is thin and extends through two-thirds of the length of the lamina. Lateral Other Seed Plants
veins are fine, dense, and arch steeply through the lamina, meeting the Cordaitalean leaves (Fig. 22) are characterized by broad, flatlami
margin at an angle of 30o-45o. Clearly, a positive identification is not
nae, 2-3 cm wide, that have fine, parallel veins that branch occasionally.
possible. The lamina enlarges gradually from the base, is straight-sided throughout
Small isolated pinnules of ambiguous, but likely pteridospermous most of its length, and tapers toward a bluntly pointed apex.
affinity were found. These pinnules have a neuropteroid shape, with a Fragmentary Taeniopteris, similar to T. abnormis (T. multinervis
broadly rounded apex, straight to rounded sides, and a tapering ovate has been placed in synonymy with T. abnormis by Barthel, 1976; see
base (Fig. 18). The venation, however, is not neuropteroid, but more also Wagner and Martínez Garcia, 1982), were found rarely (Fig. 23).
81

FIGURE 5. Lycopsid leaves and cones. 1, Lycopsid leaves, USNM 539039 from loc 42123 x 2. 2, Lycopsid leaf, USNM 539031 from loc 42121 x 1. 3,
Lycopsid cone, USNM 539038 from loc 42122, x 2.

These leaves have wide laminae, nearly 4 cm. Midveins also are wide, common element of the collections (Figs. 25, 26). Sterile foliage is similar
nearly 0.5 cm. Lateral veins depart from the midvein at right angles and to Polymorphopteris (Acitheca) polymorpha (Zodrow et al., 2006). Pin
proceed without inflection to the leaf margin. Lateral veins branch imme nae bear large subopposite pinnules with a constricted but broadly at
diately adjacent to the midvein and may branch one or more times before tached base, sides that can be slightly convex or acropetally convex and
reaching the margin. acropetally weakly concave, with a bluntly pointed tip and vaulted
Walchian conifers (Fig. 24) were preserved only as small frag laminae. Terminal pinnae are small, and rounded, with the pinnules get
ments. Most of these specimens have leaves that appear, probably due ting rapidly smaller basipetally; those adjacent to the terminal pinnule
to the nature of their exposure in the rock matrix, as narrow and needle may have odontopterid venation, with several veins entering the base
like, with an upward curving aspect, not adpressed to the stem. How without a well developed midvein. The midvein of more proximal pin
ever, when exposed by breaks paralleling the outer (abaxial) surfaces, nules is sunken and extends two-thirds or less of the distance from the
along the edge of a shoot, these leaves are revealed as 3-4 mm long and base to the tip of the pinnule. Venation is “polymorphopterid” (sensu
1.0-1.5 mm wide with acuminate apices. They resemble Walchia Wagner 1958b), meaning that lateral veins initially arch steeply upward
pinniformis. A single conifer cone scale was found (Fig. 24.2). Preserva and branch close to their point of insertion into the midvein; each deriva
tion was poor but the scale appears to be multilobed, consistent with a tive vein then tracks at increasing inflexion from the midvein and branches
walchian affinity. again, within one-third of the distance to the leaf margin, the final branched
segments extending the long distance to the leaf margin, which they
Ferns contact at a high angle of 60o or more, being nearly horizontal with the
Sterile and fertile foliage attributable to Danaeites emersonii is a rachis in the lower, basipetal portions of a pinnule. Veins can be dense.

FIGURE 6 (Next page). Calamites sp. 1, Calamites sp., 3D cast, 1a and 1b, different views of same specimen, USNM 538814 from loc 42128 x 0.5. 2,
Calamites sp., 3D cast, USNM 528487 from loc 42128, x 1.3, Calamites sp., 3D compacted cast, USNM 536769 from 42994 x 1. 4, Calamites sp., branch
scars, USNM 539022 from loc 42120 x 1.5, Calamites sp., USNM 539023 from loc 42120 x 1.
FIGURE 7 (Following page). Calamites sp. 1, Calamites sp., compression, USNM 536777 from loc 42995 x 0.5. 2, Calamites sp., USNM 536771 from
loc 42294 x 0.5. 3, Calamites sp., USNM 536775 from loc 40621 x 2.4, Calamites sp., USNM 538803 from loc 42128 x 2.5, Calamites sp., ribs not
visible, USNM 536772 from loc 42994 x 1.
FIGURE 8 (Following page). Annularia carinata. 1, Annularia carinata, USNM 536801 from loc 42293 x 2. 2, Annularia carinata, USNM 536770 from
loc 42293 x 2. 3, Annularia carinata, USNM 536768 from loc 42294 x 2. 4, Annularia carinata, USNM 536770 from loc 42294 x 2.
85

FIGURE 9. Annularia cf. spicata, Annularia carinata and Macrostachya sp. 1, Annularia cf. spicata, USNM 528479 from loc 42128 x 4. 2, Annularia
carinata, USNM 528642 from loc 41878 x 2. 3, Macrostachya sp., USNM 539021 from loc 42120 x 1. 4, Macrostachya sp., USNM 536810 from loc
42298 x 2.
86

FIGURE 10. Palaeostachya sp. 1, Palaeostachya sp., USNM 536778 from loc 42295; shaded polygon in 1a indicates area of enlargement shown in 1b;
1a x 1, 1b x 2. 2, Palaeostsachya sp., USNM 536791 from loc 42295 x 2. 3, Palaeostsachya sp., USNM 536787a from loc 42295 X2. 4, Palaeostachya
sp., USNM 536767 from loc 42994 x 2.5, Palaeostsachya sp., USNM 528460 from loc 42128 x 2. 6, Palaeostachya sp., USNM 528462 from loc 42128
x 2. 7, Palaeostachya sp., USNM 528463 from loc 42128 x 2.
87

FIGURE 11. Sphenophyllum angustifolium, Sphenophyllum verticillatum. 1, Sphenophyllum angustifolium, USNM 528658 from loc 42120 x 2. 2,
Sphenophyllum angustifolium, USNM 539019 from loc 42120 x 2. 3, Sphenophyllum verticillatum, USNM 539027 from 42121 x 2.4, Sphenophyllum
angustifolium, USNM 539018 from loc 42120 x 2.5, Sphenophyllum angustifolium, USNM 539016 from loc 42120 x 2. 6, Sphenophyllum verticillatum,
USNM 539028 from loc 42121 x 4. 7, Sphenophyllum angustifolium, USNM 539029 from loc 42121 x 2.
88

FIGURE 12. Sphenophyllum. 1, Sphenopyllum oblongifolium, USNM 528637 and Pecopteris unita pinnules from loc 41875 x 2. 2, Sphenophyllum sp.,
stem, USNM 538821 from loc 42126 x 2. 3, Sphenophyllum verticillatum, USNM 528640 from loc 41875 x 1. 4, Sphenophyllum verticillatum, USNM
539017 from loc 42120 x 4, 5, Sphenophyllum oblongifolium, USNM 528636 from loc 41876 x 1.6, Sphenophyllum verticullatum, USNM 528654 from
loc 42121 x 2. 7, Sphenophyllum sp., stem, USNM 538821 from loc 42126 x 1.8, Sphenophyllum sp., stem, USNM 538820 from loc 42126 x 1.
89

FIGURE 13. Sphenophyllum sp., Sphenophyllum oblongifolium. 1, Sphenophyllum sp., USNM 536773 from loc 40621 x 2. 2, Sphenophyllum sp.,
USNM 536805 from loc 42299 x 1.3, Sphenophyllum cf. thonii, USNM 538602 from loc 42293 x 4. 4, Sphenophyllum oblongifolium, USNM 536773
from loc 40621 x 0.5. 5, Sphenophyllum oblongifolium, USNM 538805 from loc 42128 x 4.
90

FIGURE 14. Odontopteris cf. lingulata/Macroneuropteris scheuchzeri, Odontopteris cf. brardii. 1, Odontopteris cf. lingulata or Macroneuropteris
scheuchzeri, USNM 536794 from loc 42296 x 2. 2, Odontopteris cf. brardii, USNM 528477 from loc 42178 x 4. 3, Odontopteris cf. lingulata or
Macroneuropteris scheuchzeri, USNM 536788 from loc 42295 x 2.4, Odontopteris cf. lingulata, USNM 528466 from loc 42128 x 2.5, Odontopteris cf.
lingulata, USNM 528464 from loc 42128 x 2. 6, Odontopteris cf. brardii, USNM 536781 from loc 42295 x 2.
FIGURE 15 (Facing page). Alethopteris zeilleri. 1, Alethopteris zeilleri, USNM 528644 from loc 41877 x 2. 2, Alethopteris zeilleri, USNM 528657 from
loc 42120 2a with shaded polygon indicating area of enlargement seen in 2b; 2a x 12b x 2. 3, Alethopteris zeilleri, USNM 538842 from loc 42120 x 2.
4, Alethopteris zeilleri, USNM 538809 from loc 42128 x 2.5, Alethopteris zeilleri, USNM 536804 from loc. 42299 x 2. 6, Alethopteris zeilleri, USNM
538810 from loc 42128 x 2. 7, Alethopteris zeilleri, USNM 538898 from loc 42298 x 2.
FIGURE 16 (Following page). Alethopteris zeilleri. 1, Alethopteris zeilleri, USNM 538840 from loc 42120 x 2. 2, Alethopteris zeilleri, USNM 538841 from
loc 42120 x 2. 3, Alethopteris zeilleri, USNM 538844 from loc 42120. 4, Alethopteris zeilleri, USNM 539042 from loc 42123 x 2.5, Alethopteris zeilleri,
USNM 538845 from loc 42120 x 2. 6, Alethopteris zeilleri, USNM 539030 from loc 42121 x 1.7, Alethopteris zeilleri, USNM 539037 from loc 42122
x 2.
94

FIGURE 18. Indeterminate pteridosperm pinnules. All specimens illustrated here are of uncertain affinity. They may be referred to a Neurocallipteris
species such as N. planchardii, an undescribed form of Neuropteris, or even to Alethopteris (though their venation is not like that of A. zeilleri). 1, USNM
536813 from loc 42297 x 2. 2, USNM 536783 from loc 42295 x 2. 3, USNM 536816 from loc 42297 x 2. 4, USNM 536796 from loc 42296 x 4.

Fertile foliage is of the same size and shape as the sterile, but lacks clearly between two-thirds and completely to the pinnule margin. Lateral veins
defined veins, which are obscured by reproductive organs. The sporangia arch steeply upward in a concave path from their point of midvein
are arranged circular sori that are closely packed, extending to the margin insertion to the margin, often forming what can be described as a fascicle.
of the pinnule in elongate clusters at right angles to the midvein. In the most extreme cases nearly all lateral veins reach the margin in the
A number of forms attributable to Pecopteris accounted for a few upper 25% of the pinnule.
percent of most collections, and rarely are common to abundant. Most Nemejcopteris feminaeformis (Figs. 30.1, 30.3, 30.4) is a filicalean
are characterized by small to medium sized pinnules with broad to slightly fern of zygopterid affinity. It is easily recognizable by the form of its
constricted basal attachments to the subtending rachis (Figs. 27.1-27.4, pinnules. These are small, broadly attached, narrow, and with serrate
27.6, 28, 30.2). Many of these pecopterid pinnules or partial pinnae pinnule margins that extend onto a blunt tip. The pinnule midvein is
appear to be fragmentary bits of large fronds with lobate pinnule-pinna weak and follows a somewhat irregular path. Lateral veins are high angle
architecture, which, where the pattern of lobing and associated venation and terminate in the marginal serrations; they may fork one time, about
could be seen more fully, conform broadly to the concept of Lobatopteris half-way between midvein and margin.
(Fig. 29) (Wagner, 1958a). Venation in the lobatopterid forms varies in a Sphenopteris biturica was identified from fragmentary pinnae bear
regular manner. In some pinnules venation is simple throughout a pin ing small pinnules with sharply pointed, shallow teeth (Figs. 31.1, 31.4
nule. This grades into pinnules in which some veins are once-branched, 31.8). Larger pinnules may have six or more teeth per side, borne on
which then grades further into pinnules with more complex vein branch larger shallow, somewhat rounded lobes. Pinnules shrink gradually in
ing architectures approximating those typical of lobatopterid pecopterids size in the vicinity of the pinna apex, creating an elongated appearance,
(Wagner, 1958a). Lobatopterid venation stereotypically is characterized with the terminal pinnule being small and slightly fused to the nearest
by veins that branch in a fasciculate candelabra-like manner, each succes lateral pinnules. Venation is difficult to observe. It appears to be sparse,
sive forking occurring toward the interior of the fascicle. Terminal pin with a thin midvein that persists to near the pinnule apex, the lateral
nules, where seen, are small, terminating in a gradually tapering, elongate veins terminating in the pinnule teeth. These specimens look similar to
pinna tip, with pinnules that become progressively smaller in size, with those described as S. biturica by Wagner (1985).
out fusion. The specimens illustrated in Figure 29 can be compared Commonly encountered at some sampling horizons are partial
tentatively with Lobatopteris (Pecopteris) tenuinervis of Fontaine and fronds of sphenopterid aspect, similar to Oligocarpia gutberi or O.
White (1880). leptophylla (Fig. 32). On the largest specimens, the ultimate pinnae are
Pinnae of Pecopteris (Ptychocarpus) unita (Figs. 12.1, 27.5) con borne at nearly rightangles to the next higher order rachis, in alternate and
sist of small pinnules of rectangular aspect, with a broad basal attach opposite arrangement. These pinnae shrink in length, gradually approach
ment to the rachis and slight confluence between adjacent pinnules, straight ing the apex of the next higher order pinna, creating an overall, somewhat
parallel sides, and a blunt, often nearly flat tip. In other instances, these pointed aspect to the higher-order pinna. Ultimate pinnae reach lengths
pinnules may be fused laterally for nearly their full length, leaving just a of about 2 cm. The pinnules are free up to the apical area of the pinna,
free scalloped margin to the pinna. The midvein of a pinnule varies from where they fuse laterally, forming an enlarged laminar, apical pinnule,
acropetally curving at the base, straightening out about one-third from which is of rounded form. There may be a slight enlargement of the basal
the base of the pinnule lamina, to straight from the rachis to somewhere basiscopic pinnule, though this could not be determined with certainly

FIGURE 17 (Previous page). Alethopteris zeilleri. 1, Alethopteris zeilleri, USNM 538827 from loc 42125; 1b with shaded polygon indicating area of
enlargement seen in 1a; 1a x 2, 1b x 0.5. 2, Alethopteris zeilleri, USNM 538824 from loc 42125 x 2. 3, Alethopteris zeilleri, pinna tip, USNM 538832 from
loc 42125 x 4. 4, Alethopteris zeilleri, USNM 538823 from loc 42125. 5, Alethopteris zeilleri, USNM 538822 from loc 4215 x 1.6, Alethopteris zeilleri,
USNM 538830 from loc 42125 x 2. 7, Alethopteris zeilleri, USNM 538831 from loc 42125 x 2.
FIGURE 19 (Facing page). Macroneuropteris scheuchzeri, Neurocallipteris cf. planchardii, Macroneuropteris/Odontopteris lingulata. 1, Macroneuropteris
scheuchzeri, USNM 539034 from loc 41221 x 1. 2, Macroneuropteris scheuchzeri, enlargment of 19-1, x 1.5. 3, Neurocallipteris cf. planchardii, USNM
528461 from loc 42128 x 2.4, Macroneuropteris scheuchzeri, USNM 539025 from loc 42120 x 2.5, Macroneuropteris scheuchzeri, USNM 538843 from
loc 42120 x 1. 6, Macroneuropteris scheuchzeri or Odontopteris lingulata, USNM 536807 from loc 42299 x 2. 7, Macroneuropteris scheuchzeri, USNM
536814 from loc 42297 x 2. 8, Macroneuropteris scheuchzeri, USNM 538812 from loc 42128 x 4.9, Macroneuropteris scheuchzeri, USNM 539033 from
loc 41422 x 4.
95
97

FIGURE 21. Pseudomariopteris cordato-ovata. 1, Pseudomariopteris cordato-ovata, USNM 528638 from loc 41878 x 2. 2, Pseudomariopteris cordato
ovata, USNM 539040 42123 x 4.

FIGURE 20 (Facing page). Macroneuropteris scheuchzeri. 1, Macroneuropteris scheuchzeri, USNM 538816 from loc 42126 x 1. 2, Macroneuropteris
scheuchzeri, USNM 538818 from loc 42126 x 2. 3, Macroneuropteris scheuchzeri, USNM 538817 from loc 42126 x 2.4, Macroneuropteris scheuchzeri,
USNM 539035A from loc 42125 x 2.5, Macroneuropteris scheuchzeri, USNM 538828 from loc 42125 x 2. 6, Macroneuropteris scheuchzeri, USNM
538829 from loc 42125 x 2. 7, Macroneuropteris scheuchzeri, USNM 538825 from loc 42125 x 1.8, Macroneuropteris scheuchzeri, USNM 528641 from
loc 41878 x 1. 9, Macroneuropteris scheuchzeri, USNM 538826 from loc 42125 x 2.
98
99

FIGURE 23. Taeniopteris sp. 1a, Taeniopteris sp., USNM 536812a from loc 42298 x 2.1b, Taeniopteris sp., USNM 536812b from loc 42298 x 2. 2,
Taeniopteris sp., USNM 536815 from loc 42297 x 2.

due to the quality of preservation. Pinnules are shallowly lobed. The


Hunt and Lucas (1992). Collections were obtained from two beds of
pinnule lamina appears to be thin, given its weak carbonization and the micaceous, finely to thickly laminated red siltstone with a clay matrix.
clarity of the veins. The midvein is flexuous and could be argued to
The flora from the lower of the two beds, from which 90 hand samples
extend nearly to the tip of the pinnule in some instances; in most cases were collected, contains only Alethopteris zeilleri and Macroneuropteris
the midvein forks and becomes indistinguishable from the lateral veins
scheuchzeri. The specimens are large, though fragmentary, and appear to
somewhere between one-half and two-thirds of the distance from the have been deposited in standing quiet water; the sediments are laminated
base of the pinnule. Lateral veins are sparse and mostly extend straight to claystones to fine siltstones and the plant fossils are undistorted and
the margin at about a 30-45o angle; the basalmost veins, particularly on
flat-lying on all lamina/bedding surfaces. A considerably smaller collec
the basiscopic side, tend to curve upward strongly as they approach the tion was obtained from the upper bed in this interval. This collection was
pinnule margin. Veins may be unforked or may fork widely, one or more
dominantly Macroneuropteris scheuchzeri, with common indeterminate
times before reaching the margin of the pinnule.
sphenophyll stems and sparse Annularia carinata whorls. This flora
FLORISTIC PATTERNS also appears to have been deposited in a lake or quiet body of water.
Three collections were obtained from the third stratigraphic hori
Flora was sampled at 21 locations from five stratigraphic hori zon, designated “El Cobre Canyon 3”. Two of these, both very small
zons. There are compositional and taphonomic differences among these voucher collections, come from organic-rich, fine sandstone with
sampling levels (Fig. 1; Table 1). interlaminated layers of organics, consisting of plant material of various
The lowest horizon, designated “Copper Mine,” consists of coarse sizes, including comminuted plant debris. The plant material is some
siltstone and fine sandstone that appears to have been deposited in an what jumbled and appears to have been deposited rapidly during a pe
active channel setting with flashy discharge, perhaps briefly abandoned riod of waning flow, carried partly in bedload. Most noteworthy is the
parts of a braided stream, where plants were growing on bars in the occurrence of a single specimen of Walchia. The third collection, consist
stream or along the channel banks. Several examples of upright calamite ing of nearly 100 hand samples, was obtained from a gray siltstone with
stems were noted and collected. These were up to 9 cm in diameter and fine-sand stringers. No species dominates this diverse flora. The most
appeared to have been buried rapidly, in place. Some flattened calamite common elements, represented on more than 10% of the hand samples,
stems were over 25 cm in diameter. Large mats of Sphenophyllum also include the tree-fern foliage Danaeites, calamite stems, an indeterminate,
were common at one site, suggesting burial of mats or thickets of low small-pinnuled Pecopteris, and Oligocarpia gutberi. Some of the calam
growing habit, close to the site of growth. Calamite remains dominated ite remains were preserved in situ, upright, buried in place. As with most
most of the collections from this horizon. Sphenophylls, cordaitalean of the previously discussed collections, this flora appears to have been
foliage (and large, possibly cordaitalean seeds), and Oligocarpia each deposited under slackening flow in an active stream channel deposit,
were common at one of the collecting sites. perhaps in a braided channel with variable, possibly flashy or seasonal
The second stratigraphic horizon from the bottom, designated changes in water volume. Walchia is represented in this third collection
“Myriopod trackway” because of the large arthropleurid track found at by a single specimen.
that level (Lucas et al., 2005), contains a flora previously described by The fourth collecting horizon, designated “El Cobre Canyon 1,”

FIGURE 22. Cordaites sp. 1, Cordaites sp., USNM 536803 from loc 42293; 1a with shaded polygon indicating area of enlargement seen in 1b; 1a x 1, 1b
x 4. 2, Cordaites sp., USNM 528480 from loc 42128 x 1.3, Cordaites sp., USNM 528484 from loc 42128 x 1. 4, Cordaites sp., USNM 528481b from loc
42128 x 1.5, Cordaites sp., USNM 536790 from loc 42295 x 1.
100

FIGURE 24. Walchia sp. 1, Walchia sp., USNM 538799 from loc 42127 x 4. 2, Walchia sp., USNM 538806b cone scale from loc 42298 x 4.3, Walchia
sp., USNM 528482a from loc 42128 x 4. 4, Walchia sp., USNM 536627a from loc 42298 x 4.5, Walchia sp., USNM 528482b from loc 42128 x 4, 6,
Walchia sp., USNM 536811 from loc 42298 x 4.

FIGURE 25 (Facing page). Danaeites emersonii. 1, Danaeites emersonii, USNM 538808 from loc 42128 x 2. 2, Danaeites emersonii, USNM 538804 from
loc 42128 x 2. 3, Odontopteris lingulata, USNM 538919 from loc 42296 x 2. 4, Danaeites emersonii, sterile foliage, USNM 536780 from loc 42295 x 1.
5, Danaeites emersonii with tips of Odontopteris cf. lingulata, USNM 536795A from loc 42296 x 1. 6, Danaeites emersonii, USNM 536798 from loc
42296 x 4. 7, Danaeites emersonii, USNM 536784 from loc 42295; 7a x 1, 7b x 4.
101
102

FIGURE 26. Danaeites emersonii fertile foliage. 1, Danaeites emersonii, USNM 536779 from loc 42295 x 2. 2, Danaeites emersonii, USNM 536786 from
loc 42295 x 1. 3, Danaeites emersonii, USNM 536799 from loc 42296 x 3. 4, Danaeites emersonii, USNM 539036 from loc 42122 x 3.
103

FIGURE 27. Pecopteris sp., Pecopteris unita. 1, Pecopteris sp., USNM 528469 from loc 42128 x 1. 2, Pecopteris sp., USNM 528649 from loc 41877 x
2. 3, Pecopteris sp., USNM 536789 from loc 42295 x 2.4, Pecopteris sp., USNM 528468 from loc 42128 x 2.5, Pecopteris unita, USNM 528637 from
loc 41875; 5a with shaded polygon indicating area of enlargement seen in 5b; 5a x 2, 5b x 8. 6, Pecopteris sp., USNM 528478 from loc 42128 x 3.
104

FIGURE 28. Pecopteris sp. 1, Pecopteris sp., USNM 539015 from loc 42120 x 4. 2, Pecopteris sp., USNM 538811 from loc 42128 x 4. 3, Pecopteris sp.,
large USNM 539032 from loc 42121 x 5.4, Pecopteris sp., USNM 528467 from loc 42128 x 2.5, Pecopteris sp., USNM 528470 from loc 42128 x 2. 6,
Pecopteris sp., USNM 539026 from loc 42120 x 2. 7, Pecopteris sp., USNM 538815 from loc 42128 x 4.
FIGURE 29 (Facing Page). Lobatopteris sp., cf. L. tenuinervis. 1, Lobatopteris sp., USNM 536793 from loc 42295 x 2. 2, Lobatopteris sp., USNM 538837
from loc 42294 x 4.3, Lobatopteris sp., USNM 538807 from loc 42128 x 2.4, Lobatopteris sp., USNM 536792 from loc 42295 x 2.
105
106

FIGURE 30. Nemejcopteris feminaeformis, fertile Pecopteris sp. 1, Nemejcopteris feminaeformis, USNM 536806 from loc 42299 x 4. 2, Fertile Pecopteris
sp., USNM 538802 from loc 42128 x 2. 3, Nemejcopteris feminaeformis, USNM 539012 from loc 42120 x 4.4, Nemejcopteris feminaeformis, USNM
528653 from loc 42120 x 2.
107

FIGURE 31. Sphenopteris biturica. 1, Sphenopteris biturica, USNM 539061 from loc 42299 x 4. 2, Sphenopteris sp., USNM 536797 from loc 42296 x
2. 3, Sphenopteris sp., USNM 539011 from loc 42120 x 4. 4, Sphenopteris biturica, USNM 538848 from loc 42120 x 2.5, Sphenopteris biturica, USNM
539024 from loc 42120 x 5.6, Sphenopteris biturica, USNM 538847 from loc 42120 x 4. 7, Sphenopteris biturica, USNM 538846 from loc 42120 x 2.
8, Sphenopteris biturica, USNM 539014 from loc 42120 x 2.
108

FIGURE 32. Oligocarpia gutbieri/leptophylla. 1, Oligocarpia gutbieri/leptophylla, USNM 528656 from loc 42120; 1a with shaded polygon indicating
area of enlargement seen in 1b; 1a x 2, 1b x 4. 2, Oligocarpia gutbieri/leptophylla, USNM 536782b from loc 42295 x 2.
109

FIGURE 33. Miscellaneous unknowns. 1, Smooth axes, possibly lycopsid rootlets, USNM 528472 from loc 42128 x 1. 2, Unidentified foliage, USNM
528639 from loc 41878 x 4. 3, Punctate axis, USNM 538839 from loc 42120 x 2. 4, Ribbon-like axes, probably lycopsid rootlet, USNM 538801 from
loc 42128 x 2.5, Ribbon-like axis, probably lycopsid rootlet, USNM 536785 from loc 42295 x 2.
110

FIGURE 34. Seeds. 1, Seed, USNM 538813 from loc 42128 x 1. 2, Seed, USNM 538833 from loc 42295 x 1. 3, Seed, USNM 538834 from loc 42294 x
1. 4, Seed, USNM 538835 from loc 42294 x1.5, Seed, USNM 538836 from loc 42294 x 1. 6, Seed, USNM 538838 from loc 40621 x 1. 7, Seed, USNM
538800 from loc 40621 x 1. 8, Seed, USNM 539041 from loc 42123 x 1.

TABLE 1. El Cobre Canyon flora: Composition by stratigraphic level. was the bed on which we focused most of our collecting activity. The
eight separate USNM collections consist of dark gray siltstones and
sandstones with abundant mica, variably iron-rich, poorly bedded, bear
ing poorly sorted, often jumbled plant material. Preservation of the plant
material is variable, although frequently relatively good. The plants ap
pear to have been deposited rapidly, suggesting local transport and depo
sition in active stream channels. One collection (USNM 41875) a
heterolithic bed of thinly bedded finer and coarser siltstones and fine
sandstones, suggesting transition from channel to still-water conditions.
These collections contain a diverse flora typical of a Late Pennsylvanian
wetland. Most are dominated by or rich in Alethopteris zeilleri with
significant amounts of Macroneuropteris scheuchzeri, Pecopteris
(Pecopteris unita and Pecopteris spp.), Sphenophyllum (S. oblongifolium,
S. verticillatum, and S. angustifolium), and Sigillaria brardii, including
its cones and foliage. With the exception of a single, questionable conifer
cone scale, this flora appears to be drawn from a wet substrate environ
ment probably growing under a relatively seasonally wet, perhaps hu
mid-to-subhumid climate regime.
The final, and stratigraphically highest (youngest), collecting ho
rizon is designated “Uranium Mine.” The matrix of all four Uranium
Mine collecting sites consists of micaceous coarse siltstone and fine
sandstone with considerable iron staining. The plant fossils vary from
jumbled in the sediment to flat lying, and appear to have been trans
ported locally and probably deposited rapidly from flood and during
post-flood slack water conditions, much like the other deposits of silt
stone and fine sandstone described above. The assemblage includes large
axes as well as finer plant material. There is no segregation of plant taxa
within the assemblage. Most noteworthy about the composition of these
assemblages is the mixture of the tree fern foliage Danaeites and plants
more typical of seasonally moisture limited conditions, perhaps
subhumid-to-semiarid conditions: Walchia pinniformis, Taeniopteris sp.,
and Odontopteris cf. lingulata. More generally “wet” environment plants
are rare (Alethopteris zeilleri and Macroneuropteris scheuchzeri, calam
ite remains, and sphenophylls).

DISCUSSION
The flora of the El Cobre Canyon Formation is heterogeneous
among the 21 collecting localities from five distinct horizons, but ap
111
pears to be drawn largely from a common regional species pool, allowing dry conditions (Fielding et al., 2009). Under such conditions, the land
for the vagaries of sampling and small specimen numbers. It consists scape around the channel systems would be expected to be far more
mainly of plants typical of wet substrates, characteristic of Pennsylva heterogeneous than under more uniformly wet climatic conditions. In
nian wetland vegetation from coal basins of central and eastern North such conditions, it would be possible to have typically “wetland” plants
American through central Europe. Only at the youngest horizon are growing in close proximity to xeromorphic species typical of seasonally
plants typical of drier conditions a conspicuous element of the flora. dry landscapes. Prolonged drought increases the microhabitat heteroge
The floral composition, particularly the combined presence of neity of a landscape and encourages local variability, resulting in strong
Alethopteris zeilleri, Sphenophyllum verticillatum, Sphenophyllum ecological contrasts between wetter stream corridors and drier inter
angustifolium, Annularia carinata, Pseudomariopteris cordato-ovata, fluves.
Oligocarpia gutberi/leptophylla, and Sphenopteris biturica strongly sug The paleobotanical patterns at Cañon del Cobre suggest an in
gests a later Late Pennsylvanian age, Virgilian in regional terminology, crease in seasonal moisture limitation in the youngest deposits. This is
Stephanian Cin global terrestrial terminology. Sigillaria brardii, Danaeites consistent with patterns identified in other Late Pennsylvanian deposits
emersonii, Nemejcopteris feminaeformis, Odontopteris cf. brardii and from western America (e.g., Mapes and Gastaldo, 1987; Rothwell and
Sphenophyllum oblongifolium are all consistent with a Late Pennsylva Mapes, 1988; Mamay and Mapes, 1992; DiMichele et al., 2005). Cañon
nian (Stephanian) age. del Cobre floras are significantly different from younger deposits to the
The occurrence of small numbers of conifers (Walchia) and south in New Mexico. Those from Carrizo Arroyo, which span the
Taeniopteris is consistent with landscape heterogeneity created by cli Pennsylvanian-Permian boundary (Lucas and Zeigler, 2004), and thus
matic seasonality, and with the sedimentary style. The presence of these are likely slightly younger than El Cobre, are dominated by plants more
plants requires a somewhat more extended discussion. typical of seasonally dry settings (Tidwell et al., 1999), including coni
It is clear that floras dominated by xeromorphic plants are fers, callipterids and the probable seed fern Sphenopteridium
stratigraphically intercalated throughout the Late Pennsylvanian with manzanitanum (DiMichele et al., 2004). Floras from the Bursum Forma
typically wetland floras throughout the Euramerican equatorial region tion have been identified still further to the south, in Socorro County, at
(Broutin et al., 1990; DiMichele and Aronson, 1992; Kerp, 1996, 2000; the Pennsylvanian-Permian boundary (DiMichele et al., unpublished),
DiMichele et al., 2008). Such stratigraphic, bed- or formational-scale dominated by a seasonally dry flora of callipterids, conifers and
intercalation likely reflects climate oscillation between seasonally dry, cordaitaleans. This pattern of the progressive appearance of ever drier
semi-arid to subhumid climate regimes and humid climates on time scales floras is one generally identified for the transition between the Pennsyl
of tens of thousands of years. The combination of apparently vanian and Permian, though it is not certain that it occurs in a time
parautochthonous, environmentally contrasting mixed floras at some of equivalent manner across the Euramerican portion of Pangaea. What is
the Cañon del Cobre collecting sites requires the existence of a mixture of certain is that at least two, if not more, distinct biomes, characterized by
physical conditions on the landscape from which the flora was drawn. different dominant taxa, coexisted in tropical and paratropical latitudes
In the case of the El Cobre Canyon Formation, there is strong during the late Paleozoic. The origin and dynamics of the distinct biomes
evidence that most of the plant remains typical of wetter substrates that developed across the Carboniferous-Permian boundary has been
represent mainly plants growing within and along channels or on the difficult to interpret for many years. However, information is slowly
margins of lakes. For example, there are numerous examples of in situ emerging that sheds light on this enigmatic period. We have every reason
calamite stems buried upright, in place. And, in several collections, to believe that further field studies (e.g. Looy, 2007) will continue to
Sphenophyllum specimens formed masses, as if buried in place or trans reveal the details of the links between the ecological processes and evo
ported only short distances from stream banks or exposed intra-channel lutionary origins of the assemblages and their constituent lineages.
bars. These deposits also typically have abundant preserved organic
ACKNOWLEDGMENTS
matter and plant remains that cross bedding surfaces, indicating that such
organic material was likely carried in bed load, with periodic, rapid depo The authors thank the Smithsonian Institution for support of
sition. In the few instances of lake-like deposits, the floras were of field work in New Mexico. We are grateful to various NMMNH volun
monotonously low diversity, and appeared to consist of elements floated teers and staff, who helped collect the Cañon del Cobre paleofloras. We
from the margin of the water body into a quiet depositional setting, thank the late Richard Ballsard who contributed useful comments on
largely the seed ferns Alethopteris zeilleri and Macroneuropteris earlier versions of this manuscript. We also wish to thank reviewers
scheuchzeri. Arden Bashforth, B. Mitchel Blake, and Robert H. Wagner who made
Evidence of plants growing within the stream channel itself, of extensive and valuable comments on the paper, and Jack Wittry for
flashy discharge, and rapid burial of upright plants, all suggest seasonally comments on some of the identifications.

REFERENCES

Asama, K. 1970. Evolution and classification of Sphenophyllales in Broutin, J., Doubinger, J., Farjanel, G., Freytet, P., Kerp, H., Langiaux, J.,
Cathaysia land: Bulletin National Science Museum, v. 13, p. 291-317. Lebreton, M.L., Sebban, S. and Satta, S., 1990, Le renouvellement des
Barlow, J.A., ed., 1975, Proceedings of the First I.C. White Memorial flores au passage Carbonifére-Permian: approche stratigraphique,
Symposium, “The Age of the Dunkard” (September 25-29, 1972). West biologique, sédimentologique: Comptes Rendus de l’Académie des Sci
Virginia Geological and Economic Survey, Morgantown, WV, 352 p. ences Paris II, v. 311, p. 1563-1569.
Barthel, M., 1976, Die Rotliegendflora Sachsens: Abhandlungen Staatliches Castro, M.P., 2005, La flora estefaniense B de La Magdalena (Léon, España),
Museum für Mineralogie und Geologie (Dresden), v. 24, p. 1-190. un referente europeo. Tomo I: Antecedentes y análisis floristico:
Blake, B.M., Jr., Cross, A.T., Eble, C.F., Gillespie, W.H. and Pfefferkorn, Publicaciones del Instituto Geológico y Minero de España, Cuadernos
H.W., 1999, Selected plant megafossils from the Carboniferous of the del Museo GeoMinero, v.4, p. 1-251.
Appalachian region, eastern United States: Geographic and stratigraphic Cleal, C.J., Shute, C.H. and Zodrow, E.L., 1990, A revised taxonomy for
distribution; in Hills, L.V., Henderson, C.M. and Bamber, E.W., eds., Palaeozoic neuropterid foliage: Taxon 39: 486-529.
Carboniferous and Permian of the world: Canadian Society of Petroleum Cridland, A.A. and Morris, J.E., 1963, Taeniopteris, Walchia, and
Geologists Memoir, v. 19, p. 259-335. Dichophyllum in the Pennsylvanian System in Kansas: University of
112
Kansas Science, Bulletin 44, p. 71-85. Kerp, H. and Fichter, J., 1985, Die Macrofloren des saarpfälzischen
DiMichele, W.A. and Aronson, R.B., 1992, The Pennsylvanian-Permian Rotliegenden (?Ober-Karbon – Unter-Perm; SW-Deutschland): Mainzer
vegetational transition: A terrestrial analogue to the onshore-offshore geowissschaftliche Mitteilungen, v. 14, p. 159-286.
hypothesis: Evolution, v. 46, p. 807-824. Kerp, J.H.F., 1984, Aspects of Permian palaeobotany and palynology. V.
DiMichele, W.A., Kerp, H. and Chaney, D.S., 2004, Tropical floras of the On the nature of Asterophyllites dumasii Zeiller, its correlation with
Late Pennsylvanian-Early Permian transition: Carrizo Arroyo in con Calamites gigas Brongniart and the problem concerning its sterile foli
text: New Mexico Museum of Natural History, Bulletin 25, p. 105-110. age: Review of Palaeobotany and Palynology, v. 41, p. 310-317.
DiMichele, W.A. and Chaney, D.S., 2005, Pennsylvanian-Permian fossil Krings, M. and Kerp, H., 2000, A contribution to the knowledge of the
floras from the Cutler Group, Cañon del Cobre and Arroyo del Agua pteridosperm genera Pseudomariopteris Danzé-Corsin nov. emend. and
areas, in northern New Mexico: New Mexico Museum of Natural His- Helenopteris nov. gen.: Review of Palaeobotany and Palynology, v. 111,
tory and Science, Bulletin 31, p. 26-33. p. 145-195.
DiMichele, W.A., Gastaldo, R.A., and Pfefferkorn, H.W., 2005, Plant Krings, M., Kerp, H., Taylor, T.N. and Taylor, E.L., 2003, How Paleozoic
biodiversity partitioning in the Late Carboniferous and Early Permian vines and lianas got off the ground: On scrambling and climbing Carbon
and its implications for ecosystem assembly: Proceedings of the Califor iferous–Early Permian pteridosperms: Botanical Review, v. 69, p. 204
nia Academy of Sciences, v. 56 (supplement 1), p. 32-49. 224.
DiMichele, W.A., Kerp, H., Tabor, N.J., and Looy, C.V., 2008, Revisiting Lamboy, W. and Lesnikowska, A.D., 1988, Some statistical methods useful
the so-called “Paleophytic-Mesophytic” transition in equatorial Pangea: in the analysis of plant paleoecological data: Palaios, v. 3, p. 86-94.
Vegetational integrity and climatic tracking: Palaeoecology, Langston, W., Jr., 1953, Permian amphibians from New Mexico: Univer
Palaeclimatology, Palaeogeography, v. 268, p. 152-163. sity of California, Publications in Geological Sciences, v. 29, p. 349
DiMichele, W.A., Montañez, I.P., Poulsen, C.J. and Tabor, N.J., 2009a, 416.
Climate and vegetational regime shifts in the late Paleozoic ice age Laveine, J.-P. 1989, Guide Paleobotanique dan le Terrain Houiller Sarro
earth: Geobiology, v. 7, p. 200-226. Lorrain: Documents des Houilleres du Bassin de Lorraine, v. 1, p. 1-154.
DiMichele, W.A., Nelson, W.J., Elrick, S. and Ames, P.R., 2009b, Cata Looy, C.V., 2007, Extending the range of derived late Paleozoic conifers:
strophically buried Middle Pennsylvanian Sigillaria and calamitean sphe Lebowskia gen. nov. (Majonicaceae): International Journal of Plant
nopsids from Indiana, USA: What kind of vegetation was this?: Palaios, Sciences, v. 168, p. 957-972.
v. 24, p. 159-166. Lucas, S.G., 2002, Tetrapods and the subdivision of Permian time; in Hills,
Doubinger, J., Vetter, P., Langiaux, J., Galtier, J., and Broutin, J. 1995, La L.V., Henderson, C.M. and Bamber E.W., eds., Carboniferous and Per
flore fossile du basin houiller de Saint-Étienne: Mémoires du Muséum mian of the world: Canadian Society of Petroleum Geologists Memoir,
National d’Histoire Naturelle, v. 164, p. 1-355. v. 19, p. 479-491.
Eberth, D.A. and Miall, A.D., 1991, Stratigraphy, sedimentology and evolu Lucas, S.G. and Krainer, K., 2005, Cutler Group (Permo-Carboniferous)
tion of a vertebrate-bearing, braided to anastomosed fluvial system, stratigraphy, Chama Basin, New Mexico: New Mexico Museum of Natu
Cutler Formation (Permian-Pennsylvanian), north-central New Mexico: ral History and Science, Bulletin 31, p. 90-100.
Sedimentary Geology, v. 72, p. 225-252. Lucas, S.G. and Zeigler, K.E., eds., 2004, Carboniferous-Permian transition
Falcon-Lang, H.J., Nelson, W.J., Elrick, S., Looy, C.V., Ames, P.R. and at Carrizo Arroyo, central New Mexico: New Mexico Museum of Natu
DiMichele, W.A., 2009, Incised channel-fills containing conifers indi ral History and Science, Bulletin 25, 300 p.
cate that seasonally dry vegetation dominated Pennsylvanian tropical Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S. and Henrici, A.C.,
lowlands: Geology, v. 37, p. 923-926. 2005, Vertebrate biostratigraphy and biochronology of the Pennsylva
Feldman, H.R., Franseen, E.K., Joeckel, R.M., and Heckel, P.H., 2005, nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico:
Impact of longer-term modest climate shifts on architecture of high New Mexico Museum of Natural History and Science, Bulletin 31, p.
frequency sequences (cyclothems), Pennsylvanian of Midcontinent, 128-139.
U.S.A.: Journal of Sedimentary Research, v. 75, p. 350-368. Mamay, S.H. and Mapes, G., 1992, Virgilian plant megafossils from the
Fielding, C.R., Frank, T.D., Birgenheier, L., Rygel, M.C., Jones, A.T., and Kinney Brick Company Quarry, Manzanita Mountains, New Mexico:
Roberts, J., 2008, Stratigraphic imprint of the late Palaeozoic ice age in New Mexico Bureau of Mines and Mineral Resources, Bulletin 138, p.
eastern Australia: A record of alternating glacial and nonglacial climate 61-85.
regime: Journal of the Geological Society, v. 165, p. 129-140. Mapes G. and Gastaldo, R.A., 1987, Late Paleozoic non-peat accumulating
Fielding, C.R., Allen, J.P., Alexander, J. and Gibling, M.R., 2009, Facies floras: University of Tennessee Studies in Geology, v. 2, p. 115–127.
models for fluvial systems in the seasonal tropics and subtropics: Geol Montañez, I. P., Tabor, N. J., Niemeier, D., DiMichele, W. A., Frank, T. D.,
ogy, v. 37, p. 623-626. Fielding, C. R., and Isbell, J. L. 2007. CO2-forced climate and vegetation
Fontaine, W.M., 1890, Description of the species: U.S. National Museum instability during late Paleozoic deglaciation: Science, v. 315. p. 87-91.
Proceedings, v. 13, p. 225-252. Pfefferkorn, H.W. and Wang, J., 2009, Stigmariopsis, Stigmaria asiatica,
Fontaine, W.M. and White, I.C., 1880, The Permian or Upper Carbonifer and the survival of the Sigillaria brardii-ichthyolepis group in the trop
ous flora of West Virginia and S.W. Pennsylvania: Second Geological ics of the Late Pennsylvanian and Early Permian: PalaeoWorld, v. 18,
Survey of Pennsylvania, Report of Progress PP, p. 1-143. p. 130-135.
Fracasso, M.A., 1980, Age of the Permo-Carboniferous Cutler Formation Pfefferkorn, H.W., Mustafa, H., and Hass, H., 1975, Quantitative
vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale Characterisierung ober-karboner Abdruckfloren: Neues Jahrbuch für
ontology, v. 54, p. 1237-1244. Geologie und Paläontologie Abhandlung, v. 150, p. 253-269.
Heckel, P.H. and Clayton, G., 2006, The Carboniferous System. Use of the Phillips, T.L. and DiMichele, W.A., 1992, Comparative ecology and
new official names for the subsystems, series and stages: Geologica Acta, life-history biology of arborescent lycopods in Late Carboniferous swamps
v.4, p. 403-407. of Euramerica: Annals of the Missouri Botanical Garden, v. 79, p. 560-588.
Hunt, A.P. and Lucas, S.G., 1992, The paleoflora of the Lower Cutler Potonié, R., 1893, Die Flora des Rothliegenden von Thüringen: Abhandlungen
Formation (Pennsylvanian, Desmoinesian?) in El Cobre Canyon, New der königlich-preussischen geologischen Landesanstalt N.F., v. 9, p. 1
Mexico, and its biochronological significance: New Mexico Geological 185.
Society, Guidebook 43, p. 145-150. Rössler, R., and Noll, R., 2006, Sphenopsids of the Permian: Pt. 1, The
Kerp, H., 1996, Post-Variscan late Paleozoic northern hemisphere gymno largest known anatomically preserved calamite, an exceptional find
sperms: the onset to the Mesozoic: Review of Palaeobotany and Pa from the petrified forest of Chemnitz, Germany: Review of Palaeobotany
lynology, v. 90, p. 263-285. and Palynology, v. 140, p. 145–162.
Kerp, H., 2000, The modernization of landscapes during the late Paleozoic Rothwell, G.W. and Mapes, G., 1988, Vegetation of a Paleozoic conifer
early Mesozoic: Paleontological Society Papers, v. 6, p. 80-113. community; in Mapes, G. and Mapes, R.H., eds., Regional geology and
paleontology of upper Paleozoic Hamilton quarry area in southeastern
113
Kansas: Kansas Geological Survey Guidebook Series, v. 6, p. 213-224. the Puertollano Basin, Ciudad Real, Spain: Annais de Faculdade de
Rygel, M.C., Fielding, C.R., Frank, T.D., and Birgenheier, L.P., 2008, The Ciências, Universidade do Porto, v. 64, supplement, p. 171-231.
magnitude of late Paleozoic glacioeustatic fluctuations: A synthesis: Wagner, R.H., 1998, Consideraciones sobre los pisos de la Serie Estafaniense:
Journal of Sedimentary Research, v. 78, p. 500-511. Mongrafias de la Academia de Ciencias de Zaragoza, v. 13, p. 9-19.
Tidwell, W.D., Ash, S. R, Kues, B.S., Kietzke, K. K. and Lucas S. G., 1999, Wagner, R.H. and Castro, M.P., 1998, Neuropteris obtusa, a rare but wide
Early Permian megafossils from Carrizo Arroyo, central New Mexico: spread Late Carboniferous pteridosperm: Palaeontology, v. 41, p. 1-22.
New Mexico Geological Society, Guidebook 50, p. 297-304. Wagner, R.H. and Martínez Garcia, E., 1982, Description of an Early Per
Utting, J., 2001, Palynological study of three samples from the Late Car mian flora from Asturias and comments on similar occurrences in the
boniferous Cutler Formation, El Cobre Canyon, Rio Arriba County, New Iberian Peninsula: Trabajos de Geologia (Universidad de Oviedo), v. 12,
Mexico submitted by Spencer Lucas, New Mexico Museum of Natural p. 273-287.
History and Science (UTM Zone 13 NAD 27): Geological Survey of Wanless, H.R. and Weller, J.M., 1932, Correlation and extent of Pennsyl
Canada, Paleontological Report 1-JU-2001, 3 p. vanian cyclothems: Geological Society of America Bulletin, v. 43, p.
Wagner, R.H., 1958a, Some Stephanian pecopterids from NW Spain: 1003-1016.
Mededelingen van de Geologische Stichting New Series, v. 12, p. 5–23. Wanless, H.R. and Shepard, F.P., 1936, Sea level and climate changes related
Wagner, R.H., 1958b, Pecopteris pseudobucklandi Andrae and its general to late Paleozoic cycles: Geological Society of America Bulletin, v. 47,
affinities: Mededelingen van de Geologische Stichting New Series, v. 12, p. 1177-1206.
p. 25-31. Wing, S.L. and DiMichele, W.A., 1995, Conflict between local and global
Wagner, R.H., 1984, Megafloral zones of the Carboniferous: Compte Rendu changes in plant diversity through geological time: Palaios, v. 10, p.
Neuvième Congrès International de Stratigraphie et de Géologie du 551-564.
Carbonifère (Washington and Urbana, 1979), v. 2, p. 109-134. Zodrow, E.L., Šimnek, Z., Cleal, C.J., Bek, J. and Pšenika, J., 2006,
Wagner, R.H., 1985, Upper Stephanian stratigraphy and paleaeontology of Taxonomic revision of the Palaeozoic marattialean fern Acitheca
Schimper: Review of Palaeobotany and Palynology, v. 138, p. 239–280.
114
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

115
VERTEBRATE PALEONTOLOGY, BIOSTRATIGRAPHYAND
BIOCHRONOLOGYOFTHEPENNSYLVANIAN-PERMIAN
CUTLER GROUP, CAÑON DELCOBRE, NORTHERN NEWMEXICO

SPENCER G. LUCAS1, SUSANK. HARRIS1, JUSTINA. SPIELMANN1,


LARRY F. RINEHART1, DAVIDS BERMAN2,AMYC. HENRICI2 AND KARLKRAINER3
1 New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, NM 87104;

2 Section of Vertebrate Paleontology, Carnegie Museum of Natural History, 4400 Forbes Avenue, Pittsburgh, PA 15213;
3 Institute of Geology and Paleontology, Innsbruck University, Innrain 52, Innsbruck, A-6020 AUSTRIA

Abstracts—For more than a century, fossil vertebrates have been collected from Pennsylvanian-Permian nonma
rine redbeds of the Cutler Group (El Cobre Canyon and Arroyo del Agua formations) in Cañon del Cobre (El Cobre
Canyon), Rio Arriba County, New Mexico. Megafossil plants, palynomorphs and fossil vertebrates indicate that
the Pennsylvanian-Permian (~Virgilian-Wolfcampian) boundary is stratigraphically high in the El Cobre Canyon
Formation in Cañon del Cobre, though the exact position of the boundary is not certain. A detailed vertebrate
biostratigraphy in Cañon del Cobre constructed from all localities that can be placed into precise lithostratigraphic
position identifies three, temporally successive and distinct vertebrate fossil assemblages that represent three land
vertebrate faunachrons (LVFs). The middle and upper assemblages belong to the Coyotean and Seymouran LVFs,
respectively, and the lower assemblage is the characteristic vertebrate fossil assemblage of the Cobrean LVF, an
older, entirely Late Pennsylvanian (Virgilian) LVF. The Pennsylvanian-Permian boundary is in the Coyotean LVF.
The Cañon del Cobre vertebrate fossil record thus provides three, stratigraphically-successive vertebrate assem
blages that support recognition of three time-successive LVFs across the Pennsylvanian-Permian boundary.

INTRODUCTION
Cañon del Cobre (also known as El Cobre Canyon) is a large box
canyon about 8 km northwest of the village of Abiquiu in Rio Arriba
County, north-central New Mexico (Fig. 1). The main portion of the
canyon is a complexly faulted valley with extensive exposures of the El
Cobre Canyon and overlying Arroyo del Agua formations of the Cutler
Group (Lucas and Krainer, 2005a, b; Lucas et al., 2005c, this volume;
Kempter et al., 2007). Sediments of the Chinle Group rest disconform
ably on the Arroyo del Agua Formation along the upper 30 m of the steep
canyon walls (Lucas and Hunt, 1992; Lucas et al., 2005b).
The classic vertebrate-fossil collecting area on the canyon floor, as
well as sites in the canyon walls, have yielded unique assemblages of
fossil vertebrates that are among the oldest tetrapod fossils from North
America. Here, we update the composition of these fossil assemblages as
they were reviewed by Lucas et al. (2005c). Stratigraphic organization of
the Cañon del Cobre vertebrate fossil localities indicates there are three
superposed, biostratigraphically distinct vertebrate fossil assemblages
that correspond to three land vertebrate faunachrons (LVFs) across the
Pennsylvanian-Permian boundary (Fig. 2).
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; CM = Carnegie Museum of Natural His
tory, Pittsburgh; MCZ = Museum of Comparative Zoology, Harvard;
NMMNH = New Mexico Museum of Natural History and Science,
FIGURE 1. Index map showing the location of Cañon del Cobre (after Ash,
Albuquerque; UCLA = University of California, Los Angeles; and YPM 1974).
= Yale Peabody Museum, New Haven.
GEOLOGICAL SETTING have been divided into a lower, El Cobre Canyon Formation and an
upper, Arroyo del Agua Formation (Lucas and Krainer, 2005a, b) (Fig.
Cañon del Cobre is eroded along a series of southwest-northeast 2).
trending normal faults that are downthrown to the northwest (Smith et In Cañon del Cobre, a total thickness of about 230 m of El Cobre
al., 1961, pl. 6; Kempter et al., 2007; Lucas et al., this volume). The floor Canyon Formation is exposed, and composed of mostly siltstone and
and canyon walls expose extensive bedrock badlands, slopes and escarp sandstone (Fig. 2). Minor rock types are conglomerate/conglomeratic
ments developed in an ~400 m thick section of Cutler Group siliciclastic sandstone, sandy shale and calcrete. These rocks are characteristically
red beds (Smith et al., 1961; Lucas and Krainer, 2005a, b; Kempter et al., “brown” (pale reddish brown) and are readily distinguishable on the
2007). These strata of fluvial origin (Eberth and Berman, 1983, 1993; basis of color and lithology from the overlying Arroyo del Agua Forma
Fracasso, 1987; Eberth and Miall, 1991; Krainer and Lucas, this volume) tion. Siltstones of the El Cobre Canyon Formation contain numerous
116

FIGURE 2. Composite lithostratigraphic section in Cañon del Cobre, showing distribution of NMMNH fossil localities (numbered) and vertebrate taxa.
117
incomplete for a species level identification (Schultze and Chorn, 1997).
rhizoliths, and sandstones are typically coarse-grained, arkosic, mica
Both dipnoan specimens were collected from strata of the lower part of
ceous, trough cross-bedded and form multistoried bodies that erode to the El Cobre Canyon Formation section that yield the characteristic
thick cliffs and benches. Conglomerates characteristically have
tetrapod assemblage of the Cobrean land vertebrate faunachron (LVF)
extraformational clasts of quartzite, granite and gneiss.
The El Cobre Canyon Formation is best exposed in the floor of (Fig.2).
Cañon del Cobre. Its lower contact is not exposed, and subsurface data Chenoprosopus cf. C. milleri Mehl, 1913
indicate it rests on Proterozoic basement (Smith et al., 1961; Eberth and
Miall, 1991). The upper contact of the El Cobre Canyon Formation The cochleosaurid amphibian Chenoprosopus milleri was estab
appears to be conformable at the base of the first “orange” siltstone lished as a new genus and species based on an incomplete skull collected
slope of the Arroyo del Agua Formation. The El Cobre Canyon Forma from the upper part of the El Cobre Canyon Formation in the classic
tion approximately corresponds to megasequence 1 of Eberth and Miall collecting locality of Arroyo del Agua, Rio Arriba County, New Mexico
(Mehl, 1913). Vaughn (1963) described a cranial fragment, CM 47798
(1991). According to Eberth and Miall (1991), sediments of the El Cobre
(formerly UCLA VP 1640), that he assigned to Chenoprosopus cf. C.
Canyon Formation were deposited in a shallow, ephemeral braided stream
environment. milleri (Lucas et al., 2005c, fig5C-D). Collected from a site in the floor of
Conformably overlying the El Cobre Canyon Formation, the Ar Cañon del Cobre, the specimen consists of the dermal roof and palatal
royo del Agua Formation in Cañon del Cobre is as much as 160 m thick elements of the left narial region. Vaughn (1963) noted that the posterior
and consists mostly of siltstone and sandstone (Fig. 2). Minor rock position of the external naris, as well as the pitting of the dermal roof
types are calcrete and conglomerate/conglomeratic sandstone. Siltstones elements, are especially characteristic of Chenoprosopus.
of the Arroyo del Agua Formation are characteristically “orange” (mod
Eryops sp.
erate reddish brown) and contain abundant calcrete nodules. They form
relatively thick slopes between thin sheets of sandstone that are coarse Ubiquitous in Lower Permian tetrapod assemblages, the
grained, arkosic and trough crossbedded. Conglomerates are not as com temnospondyl amphibian genus Eryops was first established by Cope
mon in the Arroyo del Agua Formation as in the underlying El Cobre (1878) based on a skull and postcranial elements from Texas. Lucas et al.
Canyon Formation, and most are intraformational (composed of calcrete (2005a, fig. 1A-B) reported a rare Upper Pennsylvanian record of the
clasts). However, in the upper part of the Arroyo del Agua Formation genus, the first from the Cañon del Cobre, represented by a nearly com
extraformational conglomerates (primarily composed of quartzite clasts) plete skull (NMMNH P-46379), for which specific assignment awaits
are present (this is megasequence 3 of Eberth and Miall, 1991). publication of a species-level revision of Eryops (Milner, 1996). In this
The Arroyo del Agua Formation is well exposed along the walls of volume, Werneburg et al. describe this skull in detail and discuss its
Cañon del Cobre, and it conformably overlies the El Cobre Canyon taxonomic position.
Formation. Gradation and interfingering characterize this contact, so it is
apparently conformable. Throughout Cañon del Cobre, the Upper Tri Broiliellus novomexicanus (Williston, 1911b)
assic Chinle Group rests disconformably on the Arroyo del Agua Forma Williston (1911b, plate 38, fig. 1) described the dissorophidam
tion (Lucas and Hunt, 1992; Lucas et al., 2003, 2005b; Kempter et al., phibian Aspidosaurus novomexicanus on the basis of a skull and articu
2007). lated presacral vertebral column (YPM 810) that may have been col
The Arroyo del Agua Formation approximately encompasses
lected near the “type” locality of Limnoscelis paludis in Cañon del Cobre
megasequences 2 and 3 of Eberth and Miall (1991). The facies contrasts (Langston, 1953). However, the exact locality is unknown.
with those of the El Cobre Canyon Formation. Thus, the Arroyo del
Fracasso (1980) reported A. novomexicanus to be endemic to
Agua Formation is characterized by locally abundant major sandstone
Cañon del Cobre at the specific level, apparently unaware of Langston’s
ribbons and U-shaped channel fills, separated by thick intervals of silt (1953) reassignment of the specimen to the genus Broiliellus under the
stone containing abundant calcrete nodules. When compared to El Cobre new combination Broiliellus novomexicanus. In addition, Langston pro
Canyon Formation deposition, the fluvial style had changed to laterally vided a description of a well-preserved skull of B. novomexicanus from
extensive floodplains that developed between relatively stable channels, the upper part of the El Cobre Canyon Formation (Lower Permian) in
and locally anastomosed channels formed. The climate was more arid the nearby locality of Arroyo del Agua. Langston’s reassignment was
than during deposition of the El Cobre Canyon Formation (Eberth and followed by Berman and Lucas (2003), despite Carroll’s (1964) reassign
Miall 1991). ment to Aspidosaurus (which was followed by others, including Bolt
In Cañon del Cobre all but two of the known Paleozoic vertebrate and DeMar in various articles). The description of a new species of
fossil localities are in the El Cobre Canyon Formation (Fig.2). All Paleo Aspidosaurus, A. binasser, based on the most complete material of the
zoic palynomorph, fossil plant and invertebrate trace fossil sites are also genus known to date, allowed Berman and Lucas (2003) to state un
in the El Cobre Canyon Formation. equivocally that “Aspidosaurus obviously does not include A.
novomexicanus.” This exclusion of the species from the genus
VERTEBRATE FOSSILTAXA
Aspidosaurus is followed here, though we acknowledge, as noted by A.
The most recent list of fossil taxa comprising the late Paleozoic El Huttenlocker (written commun., 2009), that the generic assignment of A.
Cobre Canyon vertebrate assemblages was provided by Lucas et al. novomexicanus merits further study. In this volume, Werneburg et al.
(2005c) (Table 1). Here, we review the taxa and add some new records. describe a dissorophoid snout from the type locality of Anconastes
vesperus.
Gnathorhiza sp. and Sagenodus sp.
Platyhystrix rugosus (Case, 1910)
Until very recently, the only record of the dipnoan Gnathorhiza
from strata of the El Cobre Canyon Formation at Cañon del Cobre Williston (1911a,b) recognized the generic distinctiveness of elon
pertained to bone preserved within a fossil burrow cast identified as gate neural spines with small, irregular bosses assigned to the pelycosaur
Gnathorhiza sp. (Berman, 1993). Whereas this estivation chamber was Ctenosaurus rugosus (Case, 1910) and consequently erected the new
found lying free in the classic collecting area on the canyon floor, a combination, Platyhystrix rugosus, for their reception. Vaughn (1963)
lungfish tooth plate (Fig. 3J), collected by one of us (JAS) in the summer reported the presence of P. rugosus in the Cañon del Cobre vertebrate
of 2007, was recovered from strata of the north canyon wall. The tooth assemblage based on a large number of neural spine fragments with sur
plate displays characters diagnostic of the genus Sagenodus but is too face ornamentation consistent with that of P. rugosus from the Arroyo
118
TABLE 1. Fossil vertebrate taxa from the Cutler Group in Cañon del Cobre.
Seymouran (Fig.2).
Anconastes vesperus Berman, Reisz and Eberth, 1987
Berman et al. (1987, figs. 5-11) described a new genus and species
of trematopid amphibian, Anconastes vesperus on the basis of an incom
plete, articulated skeleton with skull and mandibles (holotype, CM 41711)
and the right margin of a skull with mandible (paratype, CM 28590).
With the possible exception of Ecolsonia cutlerensis, this is the only
definitive trematopid known from New Mexico, and represents a rare
Upper Pennsylvanian occurrence of this otherwise mainly Lower Per
mian family. The A. vesperus fossils were recovered from near the base
of the locally exposed section of the El Cobre Canyon Formation and
thus represent the stratigraphically lowest vertebrate taxon of the char
acteristic assemblage of the Cobrean LVF (Fig. 2).
Sumida et al. (1996) reported the possible occurrence of this
trematopid genus from the Lower Permian Lower Tambach Formation of
central Germany (Bromacker locality). However, the material, which
consists of a partial skull and postcrania, was subsequently described as
a new trematopid genus and species, Tambachia trogallas (Sumida et al.,
1998).
Embolomeres
A multispecimen occurrence of an undetermined embolomerous
anthracosaur was reported by Berman (1993) from strata of the upper
part of the El Cobre Canyon Formation that yields tetrapods character
istic of the Coyotean LVF (Lucas et al., 2005c). The presence of a much
larger embolomere within the stratigraphically lower Cobrean assem
blage is suggested by a single, isolated embolomerous pleurocentrum
with a diameter, if complete, that would be three times those reported by
Berman (1993) for the smaller embolomere (Lucas et al., 2005c, fig. 7G).
Diasparactus zenos Case, 1910
Case (1910, 1911) described the diadectid species Diasparactus
zenos from material collected in the 1870s by D. Baldwin (precise local
ity unknown) in Cañon del Cobre. The holotype material consists of a
short sequence of articulated dorsal and sacral vertebrae characterized by
the small size of the centra relative to the height and width of the neural
arch and zygapophyses. During the 1911 Williston expedition to Cañon
del Cobre, Case and Williston collected a nearly complete Diasparactus
zenos skeleton that includes the lower portion of a skull, jaws and a well
preserved, articulated postcranial skeleton (Case and Williston, 1913,
figs. 8-22).
A little-known photograph taken by Williston in 1911 in Cañon
del Cobre (Romer, 1950) shows strata of the north wall on which a
superimposed line delineates the vertebrate-producing horizon (see Lucas
et al., this volume). Although Romer did not mention the material Williston
collected from the north wall, it most likely included the D. zenos skel
eton. Case and Williston (1913, p. 17) stated it was collected from
“reddish, clayey sandstone somewhat below the middle of the Permo
Carboniferous strata of El Cobre Canyon,” which is consistent with the
line in the photograph, and obviously not from the canyon floor. Thus,
D. zenos is a likely member of the Coyotean-age tetrapod assemblage
from Cañon del Cobre.
Long considered to be endemic to Cañon del Cobre, D. zenos was
recently recognized from a second locality in the Upper Pennsylvanian
(Virgilian) Ada Formation of Oklahoma (Kissel and Lehman, 2002). The
del Agua locality. The largest of these neural spines (CM 47797, for material includes a left dentary and two dorsal vertebrae that display
merly UCLA VP 1638) has a height of 24 cm and an anteroposterior several characters diagnostic of Diasparactus.
length of 55 mm (Lucas et al., 2005c, fig.5B). Isolated neural spines with Desmatodon aff. D. hollandi Case, 1908
similar ornamentation have also been collected by NMMNH field par
ties (Lucas et al., 2005c, fig. 7A). The monospecific genus Platyhystrix is The holotype of the type species of Desmatodon, D. hollandi,
notable among other fossil taxa of the Cañon del Cobre vertebrate assem was described by Case (1908) on the basis of a tooth-bearing fragment of
blages in that it is the only taxon with a stratigraphic range that spans all a left maxilla from Pitcairn, near Pittsburgh, Pennsylvania. Fracasso (1980)
three LVFs represented in the canyon, from the Cobrean through the reported a partial jaw and cranial fragments of the primitive diadectomorph
119

FIGURE 3. Selected vertebrate fossils from the El Cobre Canyon Formation in Cañon del Cobre. A-D, Sphenacodon sp., NMMNH P-46534. A-D,
complete left femur in A, anterior, B, posterior, C, proximal and D, distal views. E-G, Ophiacodon sp., NMMNH P-58279, distal right femur in E,
anterior, F, posterior and G, distal views. H-I, Edaphosaurus sp., NMMNH P-43101, neural spine fragment in H, anteroposterior and I, lateral views. J,
Sagenodus sp., NMMNH P-58277, incomplete tooth plate in occlusal view. K-M, Desmatodon aff. D. hollandi, YPM 7397. K-M, incomplete maxilla
with natural mold; K, natural mold of “cheek” teeth, L, a pair of anterior teeth, and M, rubber cast of natural mold. Upper left scale applies to A-G, middle
scale applies to H-I, and lower right scale applies to J.
120
Desmatodon aff. D. hollandi (Lucas et al., 2005c, fig. 5F-J) (Fig. 3K-M) in size between the very small cranial fragments, which resemble the
from strata of the canyon floor that yield a Cobrean age assemblage. eothyridid Oedaleops from the Arroyo del Agua locality, and the much
These specimens were also described by Berman and Sumida (1995), larger postcranial elements that are consistent in size with the varanopid
who noted that, whereas the size, spacing and cusp development of the
Aerosaurus. Also noted as problematic is the completely unassociated
cheek teeth are consistent with those of D. hollandi, the size and mor nature of the skeletal elements, strongly suggesting the holotype of
phology of the anterior two maxillary teeth more closely match those of Nitosaurus is a composite. Consequently, we regard Nitosaurus as a
Diadectes lentus. However, they tentatively concurred with Fracasso dubious taxon and do not include it in the Cañon del Cobre vertebrate
(1980) in the assignment of the specimen to Desmatodon aff. D. hollandi. assemblage list.
Limnoscelis paludis Williston, 1911b Ruthiromia elcobriensis Eberth and Brinkman, 1983
Perhaps the best known taxon of the Cañon del Cobre Carbonifer The only vertebrate fossils known to have been recovered from
ous-Permian vertebrate assemblages is Limnoscelis paludis, long consid the western wall of Cañon del Cobre (exact locality unknown) pertain to
ered the oldest reptile, but more recently reassigned to the order the varanopid eupelycosaur Ruthiromia elcobriensis. Collected in 1965
Diadectomorpha (Heaton, 1980). Originally described by Williston
by A. Lewis and S. Olsen, and later described by Eberth and Brinkman
(1911a, b, 1912) and later redescribed by Romer (1946), the holotype (1983, figs. 2-15), the holotype (MCZ 3150) consists of an incomplete,
material (YPM 811) consists of an articulated, nearly complete skeleton articulated vertebral column, pelvis and right hind limb, as well as several
(Williston, 1911a, figs. 1-6; 1911b, figs. 3-16; 1912, figs. 10-11, figs. 27- cranial fragments. Spielmann and Lucas (this volume) redescribe the
29; Romer, 1946, figs. 1-4, 5C, 6A, 7A, 8A, 10). Two articulated post
holotype of R. elcobriensis, assign it to the Ophiacodontidae and suggest
cranial skeletons (MCZ 1947 and MCZ 1948, formerly YPM 819 and
that its holotype came from the Arroyo del Agua Formation, so it is of
YPM 809, respectively), one of which is complete except for the left and
Seymouran age (Table 1).
right manus, were recovered from the same locality (Williston, 1911a,
fig. 7; 1912, figs. 12-14, figs. 24-26, fig. 30). Williston (1912, figs. 1-9, Ophiacodon navajovicus Case, 1907
figs.15-23, fig. 31) reported additional articulated and associated material
with features diagnostic of L. paludis. Vaughn (1963) also collected The numerical dominance of fossils pertaining to the ophiacodontid
Limnoscelis material from Cañon del Cobre that includes one nearly Ophiacodon navajovicus in strata of the canyon floor at Cañon del
Cobre led Romer and Price (1940, p. 234) to speculate that “O.
complete vertebra and several vertebral fragments (Lucas et al. 2005c,
fig. 3E-F). navajovicus was the most common pelycosaur in its area at the time of
Fracasso (1980) reported through personal communication with deposition of the El Cobre beds.” Its complicated taxonomic history
P. P. Vaughn that material identified as Limnoscelis cf. L. paludis had began with its description by Case (1907, pl. 5, fig. 6; pl. 27, figs. 1-5, 7)
recently been discovered in the “Missourian” interval of the Sangre de as a species of the genus Dimetrodon, followed by transfer to the genus
Cristo Formation near Howard, Colorado. Consequently, he removed L. Arribasaurus (Williston, 1914, fig. 11, nos. 1-14), culminating in reas
paludis from the list of taxa endemic to the Cañon del Cobre both at the signment to the genus Ophiacodon under the new combination Ophiacodon
generic and specific levels. However, the material from Howard, Colo navajovicus (Romer and Price, 1940, figs. 22, 25, 30, 33, 34, 36, 38-40).
rado has since been identified as a new limnoscelid species, L. dynatis Although Reisz (1986) noted that known specimens of Ophiacodon
navajovicus are restricted to Cañon del Cobre, Vaughn (1962) reported
(Berman and Sumida, 1990), which is distinguished from L. paludis by the occurrence of Ophiacodon cf. O. navajovicus in the Halgaito Forma
10 characters, most of which suggest L. paludis is much more derived
tion of the Cutler Group in San Juan County, Utah. He noted that the
relative to L. dynatis. Thus, only the Cañon del Cobre locality has yielded
vertebrae are indistinguishable from those of O. navajovicus, and that,
material attributable to the species L. paludis (Berman and Sumida, 1990).
among other characters, the centra display the unusual flattened strip or
Chamasaurus dolichognathus Williston, 1915 double keeled structure characteristic of this species. Another likely
occurrence of O. navajovicus outside of Cañon del Cobre, in the Cutler
It should be noted that the holotype and only known specimen of Group of southwestern Colorado, was reported by Lewis and Vaughn
the possible captorhinomorph Chamasaurus dolichognathus described
(1965). They described and illustrated an articulated string offive small
by Williston (1915) and consisting of the anterior portion of a very small
vertebrae with centra that have a mid-ventral keel characteristic of O.
left mandible, has been lost. Because the identity of this specimen cannot
navajovicus.
be confirmed, we have removed this taxon from the Cañon del Cobre In the summer of 2004, a NMMNH field party collected numer
assemblage list (Table 1). ous associated axial and appendicular elements of a single individual that
Aerosaurus greenleeorum Romer, 1937 exhibit characters diagnostic of Ophiacodon navajovicus. These elements,
some of which were briefly mentioned and illustrated by Lucas et al.
Romer (1937) and Romer and Price (1940, figs. 22,56) described (2005c, fig. 6A-K), are described in this volume by Harris et al.
the varanopid eupelycosaur Aerosaurus greenleeorum from an articu
lated portion of the occipital region of the skull and a fragmentary post Baldwinonus trux Romer and Price, 1940
cranial skeleton. A. greenleeorum was long assumed to occur at a second The validity of the ophiacodontid Baldwinonus trux, described by
locality (Arroyo del Agua) outside of Cañon del Cobre (Langston, 1953; Romer and Price (1940, fig. 54), was questioned by Reisz (1980) and
Vaughn, 1963; Fracasso, 1980). However, the Arroyo del Agua material, Langston and Reisz (1981) based on their assumption that the holotype
which consists of two nearly complete skeletons, was distinguished material (AMNH 4780) represents two different pelycosaurs, a
specifically from A. greenleeorum as A. wellesi (Langston and Reisz, sphenacodontid (maxilla and neural spines) and an ophiacodontid (verte
1981). Therefore, A. greenleeorum is endemic to Cañon del Cobre at the brae). However, restudy and further preparation of the specimen by
specific level. Brinkman and Eberth (1986) showed that one of the neural spines fits
Nitosaurus jacksonorum Romer, 1937 together with one of the centra, demonstrating that the material does
indeed represent a single taxon. They concluded that Baldwinonus is a
The validity of the type specimen of the edaphosaurid Nitosaurus valid genus, morphologically similar to Stereophallodon, and that both
jacksonorum, briefly described by Romer (1937) and redescribed by genera are primitive members of a clade that includes Ophiacodon.
Romer and Price (1940, figs. 24, 28, 32, 33, 37, 70), was questioned by It should be noted that Romer (1952) described a partial maxilla
Langston and Reisz (1981) and Reisz (1986). They noted a discrepancy from the Lower Permian interval of the Dunkard Group of Ohio as
121

Baldwinonus? dunkardensis. However, because of the very tentative VERTEBRATE BIOSTRATIGRAPHYAND BIOCHRONOLOGY
referral of this specimen to Baldwinonus, the occurrence of the genus Vertebrate biostratigraphy and biochronology of the Lower Per
outside of Cañon del Cobre was never widely accepted. Hence, most
mian was long based on the classic redbed section in north-central Texas
workers (Langston, 1953; Vaughn, 1963; Hunt and Lucas, 1992; Eberth
(see review by Lucas, 1998, 2002, 2005, 2006). Because very few Upper
and Berman, 1993) have regarded Baldwinonus as endemic to Cañon del
Pennsylvanian vertebrates are known from Texas, understanding of the
Cobre. biostratigraphic succession of vertebrates across the Pennsylvanian-Per
Edaphosaurus novomexicanus Williston and Case, 1913 mian boundary in North America long relied on comparing disparate
Pennsylvanian sites (such as Howard in Colorado or Linton in Ohio) to
Edaphosaurus novomexicanus was described as a new species by the Early Permian record in Texas or elsewhere. Cañon del Cobre is the
Williston and Case (1913a) on the basis of a skull and incomplete, articu only location we know of in North America where substantial assem
lated postcranial elements from the nearby locality of Arroyo del Agua. blages of vertebrate fossils closely bracket the Pennsylvanian-Permian
Its presence in the lower vertebrate assemblage of the canyon floor of boundary (Fig. 2). We base this conclusion on several lines of evidence
Cañon del Cobre was first recognized by Vaughn (1963) when he col that support placement of the Pennsylvanian-Permian boundary in the
lected several incomplete neural spines with characteristic bilateral tu upper part of the El Cobre Canyon Formation at Cañon del Cobre:
bercles. The longest of these neural spine fragments (CM 47800, for 1. Megafossil plants from the lower part of the exposed section of
merly UCLA VP1641) was illustrated by Lucas et al. (2005c, fig. 5E). El Cobre Canyon Formation in the Cañon del Cobre are an Alethopteris-
dominated paleoflora of Late Pennsylvanian age (Fig.2). Although previ
Sphenacodon ferox Marsh, 1878 ous workers disagreed over a precise age in the Pennsylvanian
Marsh (1878) described the sphenacodontid Sphenacodon ferox (Desmoinesian, Missourian or Virgilian: Read in Smith et al., 1961; Mamay
on the basis of a left dentary collected from the Baldwin bonebed locality in Fracasso, 1980; Hunt and Lucas, 1992; DiMichele and Chaney, 2005),
near Arroyo del Agua. Williston (1911b, plate 34, fig. 1) redescribed and in this volume DiMichele et al. argue for a Virgilian age of the extensive
illustrated the holotype specimen in addition to establishing two lecto megaflora they collected from the El Cobre Canyon Formation.
types based on left and right maxillae that pertain to the same individual. 2. Utting and Lucas (this volume) assign a Late Pennsylvanian age
A rich concentration of fossil bones, stratigraphically high in the to palynomorphs from the El Cobre Canyon Formation at NMMNH
Arroyo del Agua Formation (Lower Permian, Seymouran LVF) of the plant localities 4563 and 4564 (Fig. 2).
eastern wall of Cañon del Cobre was first reported by Berman (1993) 3. Several tetrapod taxa from the lower part of the Cañon del
and Eberth and Berman (1993). Although the bonebed yields several Cobre section, including Desmatodon, Limnoscelis and Diasparactus,
taxa, including Platyhystrix and Diadectes, a preponderance of material are known elsewhere only from Pennsylvanian strata. For example, cor
pertains to Sphenacodon ferox, of which the greatest proportion is disar relation with the Pennsylvanian portion of the Halgaito Formation based
ticulated cranial and mandible elements (Lucas et al., 2005c, fig. 4A-G; on vertebrate taxa is well supported (Lucas, 2006). Some of the other
Rinehart et al. 2007; Spielmann et al., this volume). taxa from the lower assemblage at Cañon del Cobre, such as Edaphosaurus
A complete femur diagnostic of the genus Sphenacodon (NMMNH novomexicanus, are more primitive than Early Permian close relatives,
P-46534, Fig. 3A-D) (the distal portion of the specimen was illustrated suggestive of a Late Pennsylvanian age.
by Lucas et al., 2005c, fig. 7H-I) was collected from strata in the upper 5. At Arroyo del Agua, about 30 km southwest of Cañon del
third of the El Cobre Canyon Formation approximately 60 m below the Cobre, extensive vertebrate fossil assemblages from the upper 80 m of
overlying contact with the Arroyo del Agua Formation (Fig. 2). This site the El Cobre Canyon Formation have long been (justifiably) correlated to
represents the lowest known interval in the El Cobre Canyon Formation the lower part of the Texas Wichita Group (primarily Archer City For
section to yield Sphenacodon fossils. mation), which indicates a middle Wolfcampian (Early Permian) age for
Although Fracasso (1980) assumed that Sphenacodon material the upper part of the El Cobre Canyon Formation (Langston, 1953;
collected by Baldwin was from the floor of the canyon, Baldwin’s field Romer, 1960; Lucas et al., 2005d; Lucas, 2006).
notes regarding the collection of this material, as recounted by Romer and 6. A locality about 13 km northeast of Arroyo del Agua yielded six
Price (1940), state that the Sphenacodon fossils were collected from the partial to complete specimens of Seymouria sanjuanensis (Eberth and
“highest horizon up to date in El Cobre.” Therefore, Baldwin’s Berman, 1983; Berman et al., 1987). This locality is stratigraphically
Sphenacodon site is significantly higher in the section than the “classic” high in the Arroyo del Agua Formation. The LO (lowest occurrence) of
sites on the canyon floor, and may be as stratigraphically high as the Seymouria in the Texas section is in the upper Wolfcampian “Nocona”
Sphenacodon site in the overlying Arroyo del Agua Formation men Formation (= upper Archer City Formation: Lucas, 2006) (Hook, 1989),
tioned above. To date, the oldest known occurrence of the genus so this suggests an age of late Wolfcampian for part of the Arroyo del
Sphenacodon is from the Virgilian Red Tanks Member of the Bursum Agua Formation.
Formation in south-central New Mexico (Harris et al., 2004). It is thus reasonable to conclude that the Virgilian-Wolfcampian
boundary is stratigraphically high in the El Cobre Canyon Formation in
Summary
Cañon del Cobre, though the exact position of the boundary is not certain
Three vertebrate-fossil taxa (Anconastes vesperus, Baldwinonus (Fig. 4). A detailed vertebrate biostratigraphy in Cañon del Cobre can be
constructed from all localities that can be placed into precise
trux and Ruthiromia elcobriensis) are endemic to Cañon del Cobre at the
genus level, whereas two taxa (Limnoscelis paludis and Aerosaurus lithostratigraphic position (Fig. 2) and suggests the presence of three
greenleeorum) are endemic at the species level. Five taxa from Cañon del temporally successive and distinct vertebrate fossil assemblages (Fig. 4):
Cobre, Chamasaurus dolicognathus (sic), Limnoscelis paludis, 1. A lower assemblage of the canyon floor and lower canyon walls,
Diasparactus xenos (sic), Baldwinonus trux and Nitosaurus jacksonorum, which occurs through a stratigraphic thickness of about 150 m. This
were listed by Langston (1953) as taxa “unknown elsewhere.” Vaughn assemblage yields the characteristically Pennsylvanian tetrapod fossils
in association with Pennsylvanian plants and palynomorphs. This is the
(1963) concurred, listing the same five taxa as “peculiar to El Cobre
Canyon.” However, mostly as a result of putative occurrences of Cañon characteristic vertebrate assemblage of the Cobrean LVF of Lucas (2006).
del Cobre taxa in other localities, some variation in the list of taxa consid- 2. A middle assemblage from the LO of Sphenacodon to about the
ered endemic to Cañon del Cobre is found among different authors. top of the El Cobre Canyon Formation, a stratigraphic interval about 60
m thick. By correlation to the Arroyo del Agua bonebeds, this assem
122
nonmarine redbeds at Cañon del Cobre to marine chronostratigraphic
concepts such as Virgilian and Wolfcampian. Instead, it is more useful to
place the three successive assemblages of vertebrate fossils at Cañon del
Cobre into the framework of vertebrate biostratigraphy/biochronology
developed by Lucas (2002, 2005, 2006). This framework indicates that
the three vertebrate fossil assemblages at Cañon del Cobre represent
three land vertebrate faunachrons (LVFs). The middle and upper assem
blages belong to the Coyotean and Seymouran LVFs of Lucas (2006),
and the lower assemblage at Cañon del Cobre represents the Cobrean
LVF, an older, entirely Pennsylvanian LVF (Lucas et al., 2005c).
In biochronological terms, the beginning of the Cobrean is the
FAD (first appearance datum) of Eryops, and its end is the beginning of
the Coyotean, which is the FAD of Sphenacodon. Index taxa of Cobrean
time include Diasparactus, Desmatodon and Limnoscelis. The LO of
Eryops is now within the Cobrean characteristic assemblage, but we
believe the actual FAD of the genus (which is known from Pennsylva
nian strataelsewhere: Case, 1908; Vaughn, 1958; Milner, 1996; Harris et
al., 2004) may be older than its currently known LO in Cañon del Cobre.
FIGURE 4. Summary of vertebrate biostratigraphy and biochronology across The base of the Permian is now within the lower Wolfcampian, so the
the Pennsylvanian-Permian boundary in Cañon del Cobre. Permian base is within the Coyotean LVF (Lucas, 2006). The Cañon del
Cobre tetrapod record thus provides a standard of three, stratigraphically
blage is of Coyotean age and most likely of early-middle Wolfcampian successive vertebrate assemblages (and their corresponding time-succes
age, and encompasses the Virgilian-Wolfcampian boundary. sive LVFs) that encompass the Pennsylvanian-Permian boundary (Fig.
3. A third assemblage is the single, Sphenacodon-dominated local 4).
ity in the upper part of the Arroyo del Agua Formation. By correlation to
ACKNOWLEDGMENTS
Arroyo del Agua, where Seymouria occurs in a similar stratigraphic po
sition, this uppermost assemblage at Cañon del Cobre is of Seymouran The U. S. Forest Service permitted work in Cañon del Cobre.
age, approximately late Wolfcampian. Numerous colleagues and volunteers assisted in the fieldwork. Adam
It is extremely difficult to correlate vertebrate fossils in the wholly Huttenlocker, Joerg Schneider and Ralf Werneburg reviewed the manu
script.

REFERENCES

Ash, S. R., 1974, Upper Triassic plants of Cañon del Cobre, New Mexico: Institution of Washington, Publication no. 145, p. 1-121.
New Mexico Geological Society, Guidebook 25, p. 179-184. Case, E. C. and Williston, S. W., 1913, Description of a nearly complete
Berman, D. S, 1993, Lower Permian vertebrate localities of New Mexico skeleton of Diasparactus zenos Case: Carnegie Institution of Washing
and their assemblages: New Mexico Museum of Natural History and ton, Publication no. 181, p. 17-35.
Science, Bulletin 2, p. 11-21. Cope, E. D., 1878, Descriptions of extinct Batrachia and Reptilia from the
Berman, D. S and Lucas, S. G., 2003, Aspidosaurus binasser (Amphibia, Permian Formation of Texas: Proceedings of the American Philosophi
Temnospondyli) a new species of Dissorophidae from the Lower Per cal Society, v. 17, p. 505-530.
mian of Texas: Annals of Carnegie Museum, v. 72, p. 241-262. DiMichele, W. A. and Chaney, D. S., 2005, Pennsylvanian-Permian fossil
Berman, D. S and Sumida, S. S., 1990, A new species of Limnoscelis (Am floras from two areas of northern New Mexico: New Mexico Museum of
phibia,Diadectomorpha) from the Late Pennsylvanian Sangre de Cristo Natural History and Science, Bulletin 31, p. 26-33.
Formation of central Colorado: Annals of Carnegie Museum, v. 59, p. Eberth, D. A. and Berman, D. S, 1983, Sedimentology and paleontology of
303-341. Lower Permian fluvial redbeds of north-central New Mexico – prelimi
Berman, D. S and Sumida, S. S., 1995, New cranial material of the rare nary report: New Mexico Geology, v. 5, p. 21-25.
diadectid Desmatodon hesperis (Diadectomorpha) from the Late Penn Eberth, D. S. and Berman D. S, 1993, Stratigraphy, sedimentology and
sylvanian of central Colorado: Annals of Carnegie Museum, v. 64, p. vertebrate paleontology of the Cutler Formation redbeds (Pennsylva
481-499. nian-Permian) of north-central New Mexico: New Mexico Museum of
Berman, D. S, Reisz, R. R. and Eberth, D. A., 1987, A new genus and species Natural History and Science, Bulletin 2, p. 33-48.
of trematopid amphibian from the Late Pennsylvanian of north-central Eberth, D. S. and Brinkman D., 1983, Ruthiromia elcobriensis, a new pely
New Mexico: Journal of Vertebrate Paleontology, v, 7, p. 252-269. cosaur from El Cobre Canyon, New Mexico: Breviora, no. 474, p. 1-26.
Brinkman, D. and Eberth, D. A., 1986, The anatomy and relationships of Eberth, D. S. and Miall, A. D., 1991, Stratigraphy, sedimentology and evo
Stereophallodon and Baldwinonus (Reptilia, Pelycosauria): Breviora, lution of a vertebrate-bearing, braided to anastomosed fluvial system,
no. 485, p. 1-34. Cutler Formation (Permian-Pennsylvanian), north-central New Mexico:
Carroll, R. L., 1964, Early evolution of the dissorophid amphibians: Mu Sedimentary Geology, v. 72, p. 225-252.
seum of Compartive Zoology, Bulletin, v. 131, p. 161-250. Fracasso, M. A., 1980, Age of the Permo-Carboniferous Cutler Formation
Case, E. C., 1907, Revision of the Pelycosauria of North America: Carnegie vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale
Institution of Washington, Publication no. 55, p. 1-176. ontology, v. 54, p. 1237-1244.
Case, E. C., 1908, Description of vertebrate fossils from the vicinity of Fracasso, M.A., 1987, Braincase of Limnoscelis paludis Williston: Postilla,
Pittsburgh, Pennsylvania: Annals of the Carnegie Museum, v.4, p. 234- v. 201, p. 1-22.
241. Harris, S. K., Lucas, S. G., Berman, D. S and Henrici, A. C., 2004, Vertebrate
Case, E. C., 1910, New or little known reptiles and amphibians from the fossil assemblage from the Upper Pennsylvanian Red Tanks Member of
Permian(?) of Texas: Bulletin of the American Museum of Natural the Bursum Formation, Lucero uplift, central New Mexico: New Mexico
History, v. 28, p. 168-181. Museum of Natural History and Science, Bulletin 25, p. 267-283.
Case, E. C., 1911, A revision of the Cotylosauria of North America: Carnegie Heaton, M. J., 1980, The Cotylosauria: A reconsideration of a group of
123
archaic tetrapods; in Panchen, A. L., eds., The terrestrial environment Agua, Rio Arriba County, New Mexico: New Mexico Geological Society,
and the origin of land vertebrates: New York, Academic Press, p. 497 Guidebook 56, p. 288-296.
551. Lucas, S. G., Lerner, A. J, Hannibal, J. T., Hunt, A. P. and Schneider, J. W.,
Hook, R. W. 1989. Stratigraphic distribution of tetrapods in the Bowie and 2005e, Trackway of a giant Arthropleura from the Upper Pennsylva
Wichita Groups, Permo-Carboniferous of north-central Texas; in Hook, nian of El Cobre Canyon, New Mexico: New Mexico Geological Society,
R.W., ed., Permo-Carboniferous vertebrate paleontology, litho Guidebook 56, p. 279-282.
stratigraphy and depositional environments of north-central Texas: Marsh, O. C., 1878, Notice of new fossil reptiles: American Journal of
Austin, Society of Vertebrate Paleontology, p. 47-53. Science, v. 15, p. 409-411.
Hunt, A. P. and Lucas, S. G., 1992, The paleoflora of the lower Cutler Mehl, M. G., 1913, A description of Chenoprosopus milleri, gen. et sp. nov:
Formation (Pennsylvanian, Desmoinesian?) in El Cobre Canyon, New Publications of the Carnegie Institute of Washington, v. 181, p. 11-16.
Mexico, and its biochronological significance: New Mexico Geological Milner, A. R., 1996, Systematics of the genus Eryops (Amphibia:
Society, Guidebook 43, p. 145-150. Temnospondyli) and its possible biostratigraphical value: Journal of
Kemp, T. S., 1980, Origin of mammal-like reptiles: Nature, v. 5745, p. 378 Vertebrate Paleontology, v. 16, supplement to no. 3, p. 53A.
380. Reisz, R. R., 1980, The Pelycosauria: a review of phylogenetic relation
Kempter, K., Zeigler, K., Koning, D. and Lucas, S., 2007, Preliminary ships; in Panchen, A. L., ed., The terrestrial environment and the origin
geologic map of the Canjilon SE quadrangle, Rio Arriba County, New of land vertebrates: London and New York, Academic Press, p. 553-592.
Mexico: New Mexico Bureau of Geology and Mineral Resources, Open Reisz, R. R., 1986, Pelycosauria: Encyclopedia of Paleoherpetology, Part
file Geological Map 150, 1:24,000. 17A, 102 p.
Kissel, R. A. and Lehman, T. M., 2002, Upper Pennsylvanian tetrapods Rinehart, L. F., Lucas, S. G. and Harris, S. K., 2007, Lithology and taphonomy
from the Ada Formation of Seminole County, Oklahoma: Journal of of an Early Permian Sphenacodon bonebed in Cañon del Cobre, north
Paleontology, v. 76, p. 529-545. central New Mexico: New Mexico Geology, v. 29, p. 63.
Langston, W. Jr., 1953, Permian amphibians from New Mexico: University Romer, A. S., 1937, New genera and species of pelycosaurian reptiles:
of California, Publications in Geological Sciences, v. 29, p. 349-416. Proceedings of the New England Zoological Club, v. 16, p. 89-96.
Langston, W, Jr. and Reisz, R. R., 1981, Aerosaurus wellesi, new species, a Romer, A. S., 1946, The primitive reptile Limnoscelis restudied: American
varanopseid mammal-like reptile (Synapsida: Pelycosauria) from the Journal of Science, v. 244, p. 149-188.
Lower Permian of New Mexico: Journal of Vertebrate Paleontology, v. Romer, A. S., 1950, The upper Paleozoic Abo Formation and its vertebrate
1, p. 73-96. fauna; in Colbert, E.H. and Northrop, S.A., eds., Guidebook for the 4th
Lewis, G. E. and Vaughn, P. P., 1965, Early Permian vertebrates from the Field Conferenceof the Society of Vertebrate Paleontology in north
Cutler Formation of the Placerville area, Colorado: U. S. Geological western New Mexico: New York, American Museum of Natural History,
Survey, Professional Paper,503-C, 50 p. p. 48-55.
Lucas, S. G., 1998, Toward a tetrapod biochronology of the Permian: New Romer, A. S., 1952, Late Pennsylvanian and Early Permian vertebrates of
Mexico Museum of Natural History and Science, Bulletin 12, p. 71-91. the Pittsburgh-West Virginia region: Annals of Carnegie Museum, v. 33,
Lucas, S. G., 2002, Tetrapods and the subdivision of Permian time; in Hills, p. 47-112.
L.V., Henderson, C. M., and Bamber, E. W., eds., Carboniferous and Romer, A. S., 1960, The vertebrate fauna of the New Mexico Permian: New
Permian of the world: Canadian Society of Petroleum Geologist Memoir Mexico Geological Society, Guidebook 11, p. 48-54.
19, p. 479-491. Romer, A. S. and Price, L. I., 1940, Review of the Pelycosauria: Geological
Lucas, S.G., 2005, Permian tetrapod faunachrons: New Mexico Museum of Society of America Special Paper, v. 28, p. 1-538.
Natural History and Science, Bulletin 30, p. 197-201. Schultze, H.-P. and Chorn, J., 1997, The Permo-Carboniferous genus
Lucas, S.G., 2006, Global Permian tetrapod biostratigraphy and Sagenodus and the beginning of modern lungfish: Contributions to Zo
biochronology; in Lucas, S. G., Cassinis, G. and Schneider, J. W., eds., ology, v. 67, p. 9-70.
Non-marine Permian biostratigraphy and biochronology: Geological Smith, C. T., Budding, A. J. and Pitrat, C. W., 1961, Geology of the south
Society, London, Special Publication 265, p. 65-93. eastern part of the Chama Basin: New Mexico Bureau of Mines and
Lucas, S.G., and Hunt, A.P., 1992, Triassic stratigraphy and paleontology, Mineral Resources, Bulletin 75, 57 p.
Chama Basin and adjacent areas, north-central New Mexico: New Mexico Sumida, S. S., Berman, D. S and Martens T., 1996, Biostratigraphic correla
Geological Society, Guidebook 43, p. 151-167 tions between the Lower Permian of North America and central Europe
Lucas, S.G. and Krainer, K., 2005a, Stratigraphy and correlation of the using the first record of an assemblage of terrestrial tetrapods from
Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New Germany: PaleoBios, v. 17, p. 1-12.
Mexico Geological Society, Guidebook 56, p. 145-159. Sumida, S. S., Berman, D. S and Martens T., 1998, A new trematopid am
Lucas, S.G. and Krainer, K., 2005b, Cutler Group (Permo-Carboniferous) phibian from the Lower Permian of central Germany: Palaeontology, v.
stratigraphy, Chama basin, New Mexico: New Mexico Museum of Natu 41, p. 605-629.
ral History and Science, Bulletin 31, p. 90-100. Vaughn, P. P., 1958, On the geologic range of the labyrinthodont amphibian
Lucas, S. G., Rinehart, L. F. and Hunt, A. P., 2005a, Eryops (Amphibia, Eryops: Journal of Paleontology, v. 32, p. 918-922.
Temnospondyli) from the Upper Pennsylvanian of El Cobre Canyon, Vaughn, P. P., 1962, Vertebrates from the Halgaito Tongue of the Cutler
New Mexico: New Mexico Museum of Natural History and Science, Formation, Permian of San Juan County, Utah: Journal of Paleontology,
Bulletin 31, p. 118-120. v. 36, p. 529-539.
Lucas, S. G., Zeigler, K. E., Heckert, A. B., and Hunt, A. P., 2003, Upper Vaughn, P. P., 1963, The age and locality of the Paleozoic vertebrates from
Triassic stratigraphy and biostratigraphy, Chama Basin, north-central El Cobre Canyon, Rio Arriba County, New Mexico: Journal of Paleon
New Mexico: New Mexico Museum of Natural History and Science tology, v. 37, p 283-296.
Bulletin, v. 24, p. 15-39. Williston, S. W., 1911a, A new family of reptiles from the Permian of New
Lucas, S. G., Zeigler, K. E., Heckert, A. B. and Hunt, A. P., 2005b, Review of Mexico: American Journal of Science, v. 31, p. 378-398.
Upper Triassic stratigraphy and biostratigraphy in the Chama basin, Williston, S. W., 1911b, American Permian vertebrates: Chicago, Univer
northern New Mexico: New Mexico Geological Society, Guidebook 56, sity ofChicago Press, 145 p.
p. 170-181. Williston, S.W., 1912, Restoration of Limnoscelis, a cotylosaur reptile
Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S and Henrici, A.C., from New Mexico: American Journal of Science, v. 34, p. 457-468.
2005c, Vertebrate biostratigraphy and biochronology of the Pennsylva Williston, S. W., 1914, The osteology of some American Permian verte
nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico: brates: Journal of Geology, v. 22, p. 364-419.
New Mexico Museum of Natural History and Science, Bulletin 31, p. Williston, S. W., 1915, New genera of Permian reptiles: American Journal
128-139. of Science, v. 39, p. 575-579.
Lucas, S. G., Harris, S. K., Spielmann, J. A., Berman, D. S, Henrici, A. C., Williston, S. W. and Case, E. C., 1913, A description of Edaphosaurus
Heckert, A. B., Zeigler, K. E. and Rinehart, L. F., 2005d, Early Permian Cope: Carnegie Institution of Washington, Publication no. 181, p. 71
vertebrate assemblage and its biostratigraphic significance, Arroyo del 81.
124
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

125
DISSOROPHOID RECORD FROM THE UPPER PENNSYLVANIAN
OF CAÑON DEL COBRE, NEWMEXICO

RALFWERNEBURG1, DAVID S BERMAN2 AND SPENCERG. LUCAS3


1 Naturhistorisches Museum Schloss Bertholdsburg, Burgstr. 6, D-98553 Schleusingen, Germany;

2 Carnegie Museum of Natural History, 4400 Forbes Avenue, Pittsburgh, Pennsylvania 15213;

3 New Mexico Museum of Natural History, 1801 Mountain Road N. W., Albuquerque, New Mexico 87104

Abstract—The snout region of a large dissorophoid skull (CM 76873) is described from the Upper Pennsylvanian
El Cobre Canyon Formation of Cañon del Cobre, north-central New Mexico. The following characters are variably
attributable to a wide variety of dissorophoids: 1) large size (skull length about 20 cm); 2) premaxilla massive with
large dorsal exposure; 3) large internarial fontanel; 4) nasals extend well beyond the level of the anterior margin of
the external nares; and 5) lacrimal has a wide marginal entrance into the septomaxilla-naris-complex. Features that
appear characteristic of CM 76873 include: 1) broadly convex snout; 2) a radial pattern of ornamentation on the
dorsal roofing bones; 3) an oval, relatively short naris; 4) paired vomerine depressions; 5) small vomerine fangs; 6)
a narrow, shortanterior embayment of the choana lateral to the paired vomerine fangs; and 7) relatively short, wide
choanae.

INTRODUCTION
has obviously been shortened by weathering. The only indication of
Cañon del Cobre has been an important and prolific source of their truncation is an impression on the nasal. The nasals narrow slightly
vertebrates, including the lungfishes Gnathorhiza and Sagenodus, six as they extend deeply between the alary processes of the premaxillae and
taxa of temnospondyl amphibians, two embolomerous anthracosaurs, well beyond the level of the anterior margin of the external nares to form
three diadectomorphs, a probable captorhinomorph reptile, and seven the posterior half of the internarial fontanel. Only the anterior half of the
pelycosaurian-grade synapsids (e.g., Eberth and Brinkman, 1983; Berman, right lacrimal is preserved, which ends anteriorly in a wide marginal
1993; Lucas et al., 2005, this volume). In 1983, two specimens of entrance into the septomaxilla-naris-complex. The naris is broad but
trematopid, including the greater portion of a skeleton with skull and a elongated slightly anteroposteriorly, giving it an oval outline.
partial skull and portion of the postcranium, were collected from the In the ventral view of the anterior snout region of CM 76873, the
Upper Pennsylvanian strata in Cañon del Cobre and were subsequently premaxilla forms a narrow, ventral rim of the skull, which is occupied
described as a new genus and species, Anconastes vesperus (Berman et mainly by the alveolar shelf. The shelf is presumably continued posteri
al., 1987). Closely associated with these specimens and from the same orly by the maxilla, and the broad area between the lateral shelves is
bed was a poorly-preserved snout portion of a moderate-sized skull occupied by the vomers. Paired vomerine depressions, narrowly sepa
(CM 76873), which, despite extensive preparation, could not be as rated from one another at the midline, lie anterior and posterior to the
signed to an existing taxon. Numerous poorly preserved, fragmentary vomer-premaxilla contact. Small, paired fangs are located in circular de
bones were found associated with the snout, but offered no clues to its pressions at the anteromedial corner of the choana. Lateral to the paired
identity. Although the snout obviously represents a new taxon, at least fangs there is a narrow, short anterior embayment of the choana.
for the Cañon del Cobre locality, there was a reluctance to describe it, Between the paired fangs and the embayment of the right vomer is a
with the hope that future discoveries would clarify its identity. Whereas small cluster of denticles. A pronounced, rounded ridge without teeth or
this has not happened, the specimen seems important enough to warrant denticles borders the medial margin of the choana. The choana appears to
description. be short and wide, with a width equalling about one fourth the length.
Institutional Abbreviations: CM = Carnegie Museum of Natu
ral History, Pittsburgh; FMNH = Field Museum of Natural History, DISCUSSION
Chicago; MCZ = Museum of Comparative Zoology, Harvard Univer The snout fragment, CM 76873, from Cañon del Cobre exhibits a
sity, Cambridge; NMMNH = New Mexico Museum of Natural History, mosaic of features that are either variably attributable to a wide variety
Albuquerque; SMNS = Staatliches Museum für Naturkunde, Stuttgart; of dissorophoids and temnospondyls of close affinity (e.g., zatrachydids
TUBAF=Technische Universität Bergakademie Freiberg; UBA= Uni and dvinosaurs) or are otherwise unique. Shared features most promi
versity of Buenos Aires; UW = University of Washington. nent among the former group include: 1) large size (skull length about 20
DESCRIPTION (FIGS. 1-2) cm); 2) premaxillae massive and have a relatively large dorsal exposure;
3) large internarial fontanel; 4) nasals extending well beyond the level of
The anterior snout region of CM 76873, which is exposed in both the anterior margin of the external nares; and 5) lacrimal has a wide
dorsal and palatal views, has a maximum length of 50 mm and a maximum marginal entrance into the septomaxilla-naris-complex.
width of 62 mm. These proportions suggest an original skull length of The anterior skull fragment of the Cañon del Cobre specimen CM
about 180-200 mm. Although the skull surfaces exhibit extensive weath 76873 is compared with different temnospondyls in dorsal view (Fig.
ering, the dorsal roofing bones present an ornamentation pattern of radi 3A-H). It differs strongly from the amphibamid, micromelerpetontid,
ating ridges or striations. This is unusual for most adult temnospondyl and branchiosaurid dissorophoids in its much greater size, but is well
skulls of comparable size, in which the sculpturing pattern typically within the maximum ranges of trematopids and dissorophids. The mas
consists of low ridges surrounding small, subcircular depressions (Fig. sive and relatively large dorsal exposure of the premaxilla is a feature seen
3). A large internarial fontanel near the tip of the snout is positioned on only in the trematopids Phonerpeton (Dilkes, 1990) and Acheloma (Olson,
the midline suture of the premaxillae and nasals. The premaxilla is mas 1941; Bolt, 1974; Dilkes and Reisz, 1987). Furthermore, in CM 76873
sive and has a relatively large dorsal exposure. The posterior extent of the the nasals extend well beyond the level of the anterior margin of the
alary processes bordering the anterodorsal margin of the external naris external nares, a feature shared in the same two trematopids (Olson,
126

FIGURE 1. Snout of partial dissorophoid skull, CM 76873, from the Upper Pennsylvanian of the El Cobre Canyon Formation in Cañon del Cobre, New
Mexico, dorsal view. Abbreviations: inf, internarial fontanel; l, lachrymal; mx, maxilla; n, nasal; na, naris; pm, premaxilla; sm, septomaxilla.

FIGURE 2. Snout of partial dissorophoid skull, CM 76873, from the Upper Pennsylvanian of the El Cobre Canyon Formation in Cañon del Cobre, New
Mexico, palatal view. Abbreviations: ch, choane; mx, maxilla; pm, premaxilla; v, vomer; vd, vomerine depression.

1941; Bolt, 1974; Dilkes and Reisz, 1987; Dilkes, 1990). The possession snout, as in Aspidosaurus; 2) a radial pattern of ornamentation on the
of an internarial fontanel in CM 76873 also occurs in many dissorophoids, dorsal roofing bones; 3) an oval, relatively short naris; 4) paired vomerine
including: amphibamids, such as Doleserpeton (Bolt, 1969), Eoscopus depressions; 5) small vomerine fangs; 6) a narrow, short anterior
(Daly, 1994; Werneburg, 1993), Tersomius (Carroll, 1964), Plemmyradytes embayment of the choana lateral to the paired vomerine fangs; and 7)
(Huttenlocker et al., 2007) and Georgenthalia (Anderson et al., 2008); relatively short, wide choanae. The broadly convex snout in CM 76873
branchiosaurids and some micromelerpetontids (Boy, 1972; Werneburg, appears to set it apart from nearly all dissorophoids. Moustafa’s (1955)
1989; Schoch and Milner, 2008); some trematopids, such as Phonerpeton reconstruction of the skull in Parioxys depicts the proportions of the
(Dilkes, 1990) and Acheloma (Olson, 1941; Bolt, 1974; Dilkes and Reisz, snout as appearing to be very similar to those in CM 76873. This is in
1987); and some dissorophids such as Conjunctio (Carroll, 1964) and sharp contrast, however, to the Parioxys ferricolus specimen MCZ 1162
Aspidosaurus (Berman and Lucas, 2003). But, in contrast to these ex shown in Figure 3D, which has a much narrower snout. Despite the
amples, the internarial opening in CM 76873 is much larger. An internarial strong probability that CM 76873 represents an adult specimen because
fontanel is absent in eryopids and Parioxys. An “internarial pit” was of its large size and heavy ossification, its skull roof sculpturing is inter
described in Parioxys by Moustafa (1955), but it is not clearly defined esting in having a radial pattern of ridges typically seen in juvenile
and appears to be sculptured, possibly due to weathering. In CM 76873 temnospondyls. In contrast, in similarly sized, presumably adult skulls
the presence of a wide marginal entrance of the lacrimal into the of dissorophoids and Parioxys, the sculpturing pattern, although varying
septomaxilla-naris-complex is shared by members of all the dissorophoid in intensity among taxa, consists of a dense network of pits surrounded
families, as well as by intasuchids, eryopids, and Parioxys. by low, rounded ridges.
On the basis of the same comparative taxa as above, those features Also not seen among the dissorophoids is the relatively short,
which appear characteristic to CM 76873 include: 1) broadly convex slightly oval naris in CM 76873. The nares of Parioxys are triangular and
127

FIGURE 3. Comparison of CM 76873 in dorsal view with selected other Permo-Carboniferous temnospondyls. A, Dissorophoid, CM 76873, from the
Upper Pennsylvanian El Cobre Canyon Formation in Cañon del Cobre, New Mexico; length of the figured snout region = 5.0 cm; B, Eryops sp., NMMNH
P-46379, from the Upper Pennsylvanian El Cobre Canyon Formation in Cañon del Cobre, New Mexico (see this volume); length of the figured snout
region = 15.9 cm; C, Anconastes vesperus, CM 41711, from the Upper Pennsylvanian El Cobre Canyon Formation in Cañon del Cobre, New Mexico
(Berman et al., 1987); length of the figured snout region = 4.1 cm; D, Parioxys ferricolus, MCZ 1162; length of the figured snout region = 7.4 cm; E,
Acheloma cuminsi, UC 481; length of the figured snout region = 7.7 cm; F, Tersomius texensis, MCZ 1912; length of the figured snout region = 2.2 cm;
G, Conjunctio multidens, UC 673; length of the figured snout region = 4.3 cm; H, Dissorophus multicinctus, MCZ 4168; length of the figured snout region
= 5.9 cm.
128
closer together than in CM 76873. The paired vomerine depressions are amphibamids Amphibamus, Platyrhinops and Eoscopus, as well as the
also considered unique to CM 76873. Although vomerine depressions trematopids Ecolsonia, Actiobates and Anconastes, a dissorophoid with
are sometimes observed in dissorophoids (as well as zatrachydids), they probable affinities to Dissorophidae, Platyhystrix, is also known from
are usually present in the form of a single, deep depression on the midline the Pennsylvanian and was found together with Eryops in Cañon del
of the vomers, as in Amphibamus (Milner, 1982; Schoch, 2001), Tersomius Cobre of New Mexico (compare Werneburg et al., in this volume). In
(T. texensis: MCZ 1912, personal observation of RW), Micropholis (Boy, contrast to CM 76873, Platyhystrix rugosus from the Lower Permian of
1985), Pasawiops (Fröbisch and Reisz, 2008) and trematopids such as Arroyo del Agua in New Mexico (Berman et al., 1981) has a very narrow
Acheloma (Olson, 1941; Bolt, 1974) and Phonerpeton (Dilkes, 1990). skull with a pointed snout region and probably lacks an internarial fon
The small vomerine fangs in CM 76873 are apparently also absent in tanel.
some dissorophoids, Parioxys, and eryopids. Also, in contrast to the
latter group, in CM 76873 the choanae are relatively short and wide. The ACKNOWLEDGMENTS
paired vomerine fangs produce a narrow, short anterior embayment of Support for the field work that yielded the dissorophoid snout
the choana, which is similar to broad-skulled morphs of Micropholis (see CM 76873 was provided by a grant from the New Mexico Bureau of
Schoch and Rubidge, 2005, figs. 1D, 2D). Mines and Mineral Resources (to DSB). Thanks are also extended to
Given the present knowledge of CM 76873, it can be tentatively Amy C. Henrici (CM) and Larry Rinehart (NMMNH) for the prepara
considered a member of the Dissorophoidea but without familial or lower tion of CM 76873. We thank our colleagues J.R. Bolt (FMNH), F.A.
level taxonomic assignment. This specimen may extend the knowledge of Jenkins Jr. (MCZ), C. Schaff (MCZ), J.W. Schneider (TU BAF) and
the Pennsylvanian dissorophoids, which had already diversified into the R.R. Schoch (SMNS) for the opportunity to study specimens in their
trematopids, dissorophids, amphibamids, micromelerpetontids and care and for helpful references. This work was supported by a grant of
branchiosaurids. the Deutsche Forschungsgemeinschaft (to RW: DFG-We 2833/2-1, DFG
From the stratigraphic point of view all Pennsylvanian
We 2833/3-1). Reviews by A. Huttenlocker (UW) and C. A. Marsicano
dissorophoids are interesting for further comparison. Beside the small (UBA) improved the manuscript.

REFERENCES

Anderson, J. A., Henrici, A. C., Sumida, S. S., Martens, T. and Berman, D. S., identified as a junior synonym of Acheloma cumminsi Cope, 1882, with
2008, Georgenthalia clavinasica, a new genus and species of dissorophoid a revision of the genus: American Museum Novitates, no. 2902, p. 1-12.
temnospondyl from the Early Permian of Germany, and the relation Eberth, D. S. and Brinkman, D., 1983, Ruthiromia elcobriensis, a new
ships of the Family Amphibamidae: Journal of Vertebrate Paleontology, pelycosaur from El Cobre Canyon, New Mexico: Breviora, no. 474, p.
v. 28, p. 61-75. 1-26.
Berman, D. S, 1993, Lower Permian vertebrate localities of New Mexico Fröbisch, N. B. and Reisz, R. R., 2008, A new Lower Permian amphibamid
and their assemblages: New Mexico Museum of Natural History and (Dissorophoidea, Temnospondyli) from the fissure fill deposits near
Science, Bulletin 2, p. 11-21. Richards Spur, Oklahoma: Journal of Vertebrate Paleontology, v. 28, p.
Berman, D. S, Reisz, R. R. and Fracasso, M, 1981, Skull of the Lower 1015-1030.
Permian dissorophid amphibian Platyhystrix rugosus: Annals of Carnegie Huttenlocker, A. et
shintoni, gen. K.,sp.Pardo,
nov., an
J. D.
Early
andPermian
Small, amphibamid
B. J., 2007, (Temnospondyli:
Plemmyradytes
Museum, v. 50, p. 391-416.
Berman, D. S, Reisz, R. R. and Eberth, D. A., 1987, A new genus and species Dissorophoidea) from the Eskridge Formation, Nebraska: Journal of
of trematopid amphibian from the Late Pennsylvanian of north-central Vertebrate Paleontology, v. 27, p. 316-328.
New Mexico: Journal of Vertebrate Paleontology, v. 7, p. 252-269. Lucas, S. G., Rinehart, L. F. and Hunt, A. P., 2005, Eryops (Amphibia,
Berman, D.S, and Lucas, S.G., 2003. Aspidosaurus binasser (Amphibia, Temnospondyli) from the Upper Pennsylvanian of El Cobre Canyon,
Temnospondyli), a new species of Dissorophidae from the Lower Per New Mexico: New Mexico Museum of Natural History and Science,
mian of Texas: Annals of Carnegie Museum, v. 72, p. 241-262. Bulletin 31, p. 118-120.
Bolt, J. R., 1969, Lissamphibian origins: Possible protolissamphibian from Milner, A. R., 1982, Small temnospondyl amphibians from the Middle
the Lower Permian of Oklahoma: Science, v. 166, p. 888-891. Pennsylvanian of Illinois: Palaeontology, v. 25, p. 635-664.
Bolt, J. R., 1974, Evolution and functional interpretation of some suture Moustafa, Y. S., 1955, The affinities of Parioxys ferricolus and the phylog
patterns in Paleozoic labyrinthodont amphibians and lower tetrapods: eny of the “eryopsid” amphibians: Bulletin Institute d’Egypte, v. 36, p.
Journal of Paleontology, v. 48, p. 434-458. 77-104.
Boy, J. A., 1972, Die Branchiosaurier (Amphibia) des saarpfälzischen Olson, E. C., 1941, The Family Trematopsidae: Journal of Geology, v. 49,
Rotliegenden (Perm, SW-Deutschland): Hessisches Landesamt für p. 149-176.
Bodenforschung, v. 65, p. 1-137. Schoch, R. R., 2001, Can metamorphosis be recognized in Palaeozoic am
Boy, J. A., 1985, Über Micropholis, den letzten Überlebenden der phibians?: Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen,
Dissorophoidea (Amphibia, Temnospondyli; Unter-Trias): Neues v. 220, p. 335-367.
Jahrbuch für Geologie und Paläontologie, Monatshefte, v.1985, p. 29 Schoch, R. R. and Milner, A. R., 2008, The intrarelationships and evolu
45. tionary history of the temnospondyl family Branchiosauridae: Journal
Carroll, R. L., 1964, Early evolution of the dissorophoid amphibians: Bul of Systematic Palaeontology, v. 2008, p. 1-23.
letin of the Museum of Comparative Zoology, Harvard University, v. Schoch, R. R. and Rubidge, B.S. 2005. The amphibamid Micropholis stowi
131, p. 161-250. from the Lystrosaurus Assemblage Zone of South Africa: Journal of
Daly, E., 1994, The Amphibamidae (Amphibia: Temnospondyli), with a Vertebrate Paleontology, v. 25, p. 502-522.
description of a new genus from the Upper Pennsylvanian of Kansas: Werneburg, R., 1989, Labyrinthodontier (Amphibia) aus dem Oberkarbon
The University of Kansas, Miscellaneous Publications, v. 85, p. 1-59. und Unterperm Mitteleuropas - Systematik, Phylogenie und
Dilkes, D.W., 1990. A new trematopsid amphibian (Temnospondyli: Biostratigraphie: Freiberger Forschungshefte, v. C 436, p. 7-57.
Dissorophoidea) from the Lower Permian of Texas: Journal of Verte Werneburg, R., 1993, Ein Schädelrest von Eoscopus lockardi Daly (Am
brate Paleontology, v. 10, p. 222-243. phibia: Dissorophoidea) aus dem Oberkarbon von Kansas:
Dilkes, D.W., and R.R. Reisz, 1987. Trematops milleri Williston, 1909 Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 7/8, p.
147-149.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

129
FIRST PENNSYLVANIAN ERYOPS (TEMNOSPONDYLI)AND
ITS PERMIAN RECORD FROM NEW MEXICO

RALFWERNEBURG1, SPENCERG. LUCAS2, JÖRGW. SCHNEIDER3 AND LARRYF.RINEHART2


1Naturhistorisches Museum Schloss Bertholdsburg, Burgstr. 6, D-98553 Schleusingen, Germany;

2 New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, New Mexico 87104;
3 TU Bergakademie Freiberg, Institut für Geologie, B.v.Cotta-Str. 2, D-09596 Freiberg, Germany

Abstract—A well-preserved, partial skull of the temnospondyl amphibian Eryops is described from the Pennsyl
vanian strata of the El Cobre Canyon Formation of Cañon del Cobre, Rio Arriba County, New Mexico. The
presence of the autapomorphic interfrontal presents convincing evidence of its assignment to Eryops. The two
other known Pennsylvanian eryopid skulls, Glaukerpeton avinoffi from Pittsburgh, Pennsylvania, and Eryops cf.
E. avinoffi from Sutton, West Virginia, belong to the valid, more aquatically-adapted eryopid genus Glaukerpeton.
The Pennsylvanian Eryops of New Mexico is compared with distinctly defined types of Permian Eryops, includ
ing specimens from the Lower Permian of New Mexico, but at present no species distinction is recognized. The
Lower Permian specimens of Eryops from New Mexico and Utah are tentatively assigned to E. grandis, despite
having a more slender skull with a narrow jugal and relatively fine dermal sculpturing.

The dorsal skull table has a relatively fine and shallow sculpture
INTRODUCTION of low, narrow small ridges and shallow pits (Fig. 1), which is compa
The temnospondyl Eryops, one of the best known Permo-Car rable to that of Eryops sp. (MCZ 1914) from the lower part of the
boniferous amphibian fossils, occurs at widely distributed sites in the Archer City Formation of Texas (Werneburg, 2007, fig. 7a). The dermal
USA. The general morphology of the eryopids Eryops (Sawin, 1941; sculpturing of stratigraphically younger Eryops specimens, for example
Moulton, 1974) and Onchiodon (Boy, 1990; Werneburg, 2007) is well the holotype (AMNH 4189) of E. megacephalus (Werneburg, 2007, fig.
known, but the specific taxonomy of the rich Eryops material is unsolved 6), is much more pronounced, with higher ridges and deeper pits.
at present. Therefore, a detailed description of the Eryops skull from the The entire New Mexican skull exhibits a dorsal strutting pattern
Pennsylvanian of New Mexico (Lucas et al., 2005a) and comparison with large ridges for better mechanical stability of the skull as is already
with Permian Eryops would be one step toward a solution of the species known in Eryops and Onchiodon (Sawin, 1941; Boy, 1990; Werneburg,
level taxonomic problems of the genus Eryops. 2007). A well-developed, large longitudinal ridge extends from the
The Eryops skull from the Cañon del Cobre was discovered in postfrontal to the naris (Fig. 1B). Additional ridges run diagonally across
June, 2005, as part of the NMMNH collecting expeditions to Cañon del the jugal and lachrymal and as transverse ridges occur between the longi
Cobre led by SGL. The locality is NMMNH 6121, sec. 18, T24N, R6E. tudinal ridges in the intraorbital region, the midlength of the nasals and
It is approximately 125 m below the contact of the El Cobre Canyon the premaxillaries. Depressions are present between these ridges.
Formation and the overlying Arroyo del Agua Formation (Lucas et al., The skull is dorsoventrally crushed, as is clearly indicated by the
2005b). The skull was collected from a muddy sandstone together with medial displacement of the right jugal below the postorbital, and the
bones of an embolomere and Platyhystrix, as well as pelycosaurian-grade displacement of the left side of the skull roof bones slightly anteriorly.
synapsids. The skull is part of the characteristic assemblage of the Cobrean However, the mid and posterior widths of the skull are affected little by
land-vertebrate faunachron of Lucas (2006), and all biostratigraphic data the distortion, as evidenced by the intact palate (Fig. 2). Therefore, the
(palynostratigraphy, megafossil plants and vertebrate biostratigraphy) preserved shape of the skull is generally narrow with a relative short,
indicate it is of Late Pennsylvanian age (Lucas et al., 2005b; and this blunt snout region that accurately portrays its true outline. As such, it
volume). has an intermediate shape between the narrow skull morphotype with a
Institutional abbreviations: AMNH = American Museum of more pointed muzzle of the Eryops megacephalus holotype (AMNH
Natural History, New York; BMNH = British Museum of Natural His 4189) and the broad snouted “Onchiodon-morphotype” of Eryops sp.
tory, London; CM = Carnegie Museum of Natural History, Pittsburgh; (MCZ 1914, cf. Werneburg, 2007). However, the interorbital region in
CMNH = Cleveland Museum of Natural History; FMNH = Field Mu NMMNH P-46379 is narrower than in both mentioned morphotypes.
seum of Natural History, Chicago; MCZ = Museum of Comparative
Skull Roof (Figs. 1-2)
Zoology, Harvard University, Cambridge; NMMNH = New Mexico
Museum of Natural History, Albuquerque; SMNS = Staatliches Mu The premaxillaries are relatively short and form a rounded, blunt
seum für Naturkunde, Stuttgart; UBA = University of Buenos Aires. snout region, as in Eryops sp. (MCZ 1914). The alary process overlies
the nasal. The premaxilla, together with the maxilla, forms a shelf-like
DESCRIPTION AND COMPARISONS
margin of the naris. The unsculptured septomaxilla is oriented ventrolat
General Morphology (Figs. 1-2) erally into the naris. The “angle” in the middle part of the right maxilla is
due to the distortion of the jugal. A small internasal seems to be situated
The skull of NMMNH P-46379 consists of the anterior half and between the premaxillaries and the nasals. Sometimes such a median
the left cheek region. The marginal and palatal dentitions are exception bone is intercalated in the anterior skull table and reflects the ancestral
ally well preserved. The length from the anterior end of the premaxilla to rostral elements. Werneburg (2007) described a slightly larger bone in the
the level of the most posterior end of the quadrate is 41.5 cm. The same position in the eryopid Onchiodon thuringiensis. The posterior
midline length is about 33-34 cm, comparable to measurements of other portions of the nasals are greatly widened, so as to greatly restrict the
Eryops skulls (Lucas et al., 2005a, table 1). entrance of the lachrymal into the posteroventral margin of the naris. The
130

FIGURE 1. Eryops sp., NMMNH P-46379, from the Upper Pennsylvanian of Cañon del Cobre in New Mexico; dorsal view. Abbreviations: f = frontal, if
= interfrontal, in = internasal, j = jugal, l = lachrymal, mx = maxilla, n = nasal, pm = premaxilla, po = postorbital, pof = postfrontal, prf = praefrontal,
pt = pterygoid, q = quadrate, qj = quadratojugal, sm = septomaxilla, sq = squamosal.

lachrymal has a wide middle portion. The presence of the interfrontal is is mediolaterally expanded in the marginal dentition (Fig. 2B), as illus
clearly discernible by its radial sculpturing and its anterior and posterior trated by Stickler (1899, fig. 1A). All marginal teeth of the maxillae and
sutures. It is slightly shorter than the interfrontal in the Eryops the palatal tusks are slightly less recurved posteromedially. The teeth of
megacephalus holotype. The frontals and therefore the interorbital re the premaxillaries are not recurved (with one exception, the tooth in
gion are narrow, whereas the orbital rim of the postfrontal is relatively
position 11 of the right premaxilla is more incurved). The premaxilla has
wide. Prefrontals, frontals and jugals extend to roughly the same anterior 11 to 12 tooth positions; five or six are represented by empty alveoli.
level. The postorbital appears to be short, but its posterior portion is not The largest teeth of the premaxilla occupy tooth positions 7 to 11, with
preserved. The squamosal is the largest bone of the cheek and is posteri 8 and 9 being the largest. The maxilla has 34 tooth positions, 11 being
orly and posteromedially bordered by the quadrate and pterygoid, re
represented by empty alveoli. The largest teeth in the maxilla are mar
spectively. The quadratojugal is narrow and subequal in length to the ginal to the choanae and the tusk and socket pair of the palatine; these
squamosal. The dorsal portion of the quadrate has a narrow exposure, occupy tooth positions 4 to 8 of the right maxilla and 3 to 10 of the left
but is widened, especially in the anteromedial direction.
maxilla. These larger tooth positions did not cause a lateral swelling of
Palate (Figs. 2-3) the skull margin as is known from other Eryops skulls (e.g., Sawin,
1941). The largest maxillary teeth are smaller than those in the premax
The anterior and marginal palatal bones are well preserved. The illa.
parasphenoid and mid-region of the pterygoids are missing. The entire A depression on the ventral surface of the premaxilla lies close to
surface of the palate is covered with small, tooth-like denticles. the midline. The elongated vomers have a step-like elevation just poste
The marginal and palatal dentitions of NMMNH P-46379 are rior to the level of the vomerine tusk-and-socket pair. The paired fangs
better preserved than in most other Eryops skulls. All the marginal teeth of the vomers have a maximum length of about 14 mm and a basal
have sharp cutting edges on their distal halves (Fig. 3). Stickler (1899, fig. diameter of 6-7 mm (Fig. 3B). A raised, denticulated ridge surrounding
1) and Williston (1899, pl. 27) had illustrated the tooth morphology of the paired fangs continues along or has a posterior extension along the
Eryops with blades. The proximal halves of the larger teeth show the margin of the short and wide choanae. The palatine is wide anteriorly,
characteristic ridges of infolded dentine. The same morphology of the whereas posteriorly it quickly narrows to a splint-like process that
teeth has been described in the eryopid Onchiodon and the archegosauroids wedges between the ectopterygoid and pterygoid. The palatine fangs are
Sclerocephalus and “Cheliderpeton” (Werneburg, 1996). The tooth base the largest teeth of the skull, which have a length of about 25 mm and a
131

FIGURE 2. Eryops sp. NMMNH P-46379, from the Upper Pennsylvanian of Cañon del Cobre in New Mexico; palatal view. Abbreviations: ch = choane,
ec = ectopterygoid, f.pq = paraquadrate foramen, j = jugal, mx = maxilla, pl = palatine, pm = premaxilla, pt = pterygoid, q = quadrate, qj = quadratojugal,
sph = sphenethmoid, sq = squamosal, vd = vomerine depth.

basal diameter of 10 mm. The basal ectopterygoid also has a large tusk the USA, which have been assigned to Eryops: Glaukerpeton avinoffi
and socket pair, whose size is about 18 mm in length and 9 mm in basal from the Upper Conemough Group at Pittsburgh in Pennsylvania (Romer,
diameter. The ectopterygoid fangs are larger than those of the vomer and 1952), and Eryops cf. E. avinoffi from the Middle Conemough Group of
equal in size to the largest teeth of the premaxilla. This is a clear differ Sutton in West Virginia (Murphy, 1971). Vaughn (1958) reported an
ence from the eryopid Onchiodon, which has extremely small interfrontal in “Glaukerpeton” avinoffi (contrary to Romer, 1952) and
ectopterygoid fangs (Werneburg, 2007). concluded that Glaukerpeton is a synonym of Eryops. However, we
A large portion of the sphenethmoid is preserved, lying posterior have restudied both specimens, and they lack an interfrontal and should
to the vomer and below the interorbital region. The left pterygoid has an
be reassigned to Glaukerpeton. The valid genus Glaukerpeton Romer,
elongated, narrow process lacking denticles, as does the anterior end of 1952, may be a more aquatically adapted eryopid of the newly defined
the palatal ramus, and it was probably overlain ventrally by the vomer. “Actinodon–complex” (Werneburg, 2007). The members of this eryopid
The transverse flange is preserved on the right pterygoid. The right
complex were formerly attributed to Onchiodon, and they are distin
quadrate ramus, forming the medial wall of the adductor fossa, has a guished from Eryops by: (1) a narrow preorbital region of the skull, (2) a
narrow ventral contact with the quadrate. The quadrate is wide and short narrow interorbital region, (3) a narrow jugal, (4) a relatively elongate
and forms the quadrate condyle. The ventral portion of the quadratojugal postorbital portion of the skull with elongated supratemporals, (5) an
is expanded in an anteromedial direction. Two small paraquadrate fo-
elongated interclavicle, and (6) an elongated ventral clavicle.
ramina are situated at the posterolateral suture with the quadrate. A larger
The type species of Eryops, E. megacephalus Cope, 1877, is well
third foramen lies anteromedial to these foramina around a ridge in a small known from specimen MCZ 1129 (Sawin, 1941) and from the holotype
depression on the quadratojugal. specimen AMNH 4189 (Werneburg, 2007). Both skulls are from the
DISCUSSION Permian Petrolia Formation (Wichita Group) of Texas. However, about
100 skulls of Eryops are known from the Lower Permian (Cisuralian) of
The skull, NMMNH P-46379, from the Pennsylvanian of Cañon Texas, New Mexico, Utah, Oklahoma, and the Tristate Area (West Vir
del Cobre in New Mexico is doubtless a member of the genus Eryops ginia, Ohio and Pennsylvania). Milner (1996) provided a brief statement
(Lucas et al., 2005a). The presence of an interfrontal, which is an about a new systematic concept for the genus Eryops, containing four
autapomorphic feature of Eryops (Werneburg, 2007), confirms this as species based on cranial and dental features, but a detailed analysis has
signment. never been published. Therefore, it is difficult to determine the species
Two other eryopid skulls are known from the Pennsylvanian of level identity of the New Mexican skull NMMNH P-46379. At this
132

FIGURE 3. Eryops sp., NMMNH P-46379; median incurved labyrinthodont teeth with sharp lateral blades. A, anterior maxilla at margin of choanae; B,
fangs of the right vomer with denticulation of the socket.

stage of knowledge it is appropriate to compare it with distinctly defined This specimen was found in the Halgaito Shale of the Cutler Group
types, such as the holotype of E. megacephalus or Eryops sp., MCZ (Wolfcampian) in southeastern Utah. The right jugal is narrow, and an
1914 (see description above), as well as the known Eryops skulls from interfrontal is visible. This suggests that the species Eryops grandis may
the Lower Permian of New Mexico. be valid for a slender skull, if it is not really an effect of preservation,
Initially, it is necessary to compare the skull NMMNH P-46379 with relatively fine dermal sculpture of the same size as E. megacephalus.
with the known species of Eryops from the Cutler Group and the Abo On the other hand, the differences in skull morphology between the
Formation of New Mexico. Langston (1953) discussed the history of broad Pennsylvanian skull from Cañon del Cobre, NMMNH P-46379,
two species. E. reticulates Cope (1881) is a nomen vanum according to and the newly characterized E. grandis are clearcut. Therefore, the spe
Case (1911) and Langston (1953). Some postcranial material from the cies E. grandis cannot be applied to all Eryops specimens from New
University of California, Berkeley, localities in New Mexico (Spanish Mexico.
Queen Mine, Quarry butte, Camp quarry and Anderson quarry) is grouped Surprisingly, another Eryops skull is known from the Cañon del
under the species E. grandis by Langston (1953), which may be smaller Cobre in New Mexico. It was collected by P. C. Miller in 1911 and is
than the other Eryops specimens. According to Douthitt (1917), the housed in the collection of the Field Museum of Natural History, Chi
holotype specimen of E. grandis is lost. Therefore, he erected the new cago (FMNH-UC 751; Fig. 7). The skull is crushed, but it exhibits a fine
genus Eryopsoides for two skulls from the Baldwin bone bed, the type dermal sculpturing and the general dentition pattern. In FMNH-UC 751
locality and stratum typicum of Eryops grandis. But, the holotype of E. at the level of the ectopterygian fangs, there are three large-sized maxil
grandis exists in the MCZ collection (a mandible, sacral rib and centra: lary teeth (Fig. 7A). At this level in NMMNH P-46379, the teeth of the
MCZ 2731). Romer (1947) did not recognize a basis for the new genus maxillae are clearly smaller (Fig.2). The snout region of FMNH-UC 751
Eryopsoides and considered it to be a junior synonym of Eryops. seems to be narrow, as seen in palatal view (similar to E. grandis and
Unfortunately the available holotype material of Eryops grandis skulls from Texas), but the jugal region is wider (Fig. 7B). Therefore, the
lacks a skull, though the two skulls described by Douthitt (1917) are two skulls from Cañon del Cobre are not identical.
from the same locality and were considered as plesiotypes (topotypes) Another skull with an approximately 35 cm length is known from
of E. grandis. Both skulls are illustrated here (Figs. 4-5). MCZ 2733 the Cutler Group (Wolfcampian) at CM locality 1070, 1.6 km northwest
(Fig. 4) is a crushed skull with a right mandible, but with the same fine of Arroyo del Agua, on the west side of the Rio Puerco (CM 34906;
sculpture pattern as in MCZ 2732 (Fig. 5). This skull, MCZ 2732, is Lucas et al., 2005c, fig 4E, I; Fig. 8). This skull clearly belongs to the
partially reconstructed with a skull length of about 34 cm. The skull genus Eryops based on the following features: (1) interfrontal present,
shape in the middle and anterior portions is very slender, which is un (2) characteristic stabilization system of different sculpture ridges on
usual for nearly all other Eryops specimens. The jugal seems also to be dorsal skull roof, (3) coronoids with full denticulation as reported by
narrow. This skull is the same size as, for example, the holotype speci Langston (1953) for E. grandis, but also known from Texas Eryops
men of E. megacephalus (Werneburg, 2007). It is possible that the MCZ material (Broom, 1913; Werneburg, 2007), and (4) broad jugal. Other
2732 skull was incorrectly reconstructed. However, a very similar skull features in common for both skulls are the fine dermal sculpturing of the
shape is known from one Eryops skull (CM47817) from Utah (Fig. 6). dorsal skull roof, the outline shape of the skull, similar skull lengths, and
133

FIGURE 4. Eryops grandis; plesiotype MCZ 2733, dorsal skull with right
mandible, from the “Baldwin bonebed” of the Lower Permian El Cobre FIGURE 6. Eryops grandis, CM 47817, dorsal skull, from the Early Permian
Canyon Formation of Arroyo del Agua in New Mexico. Cutler Formation (Halgaito Shale) of the Valley of the Gods in Utah.

FIGURE 7. Eryops sp., FMNH UC 751; skull from the El Cobre Canyon
Formation of the Cañon del Cobre in New Mexico; A, ventral view of
narrow snout region with palatal and maxillary dentition; B, lateral view of
wide left jugal with fine dermal sculpture.

FIGURE 5. Eryops grandis, plesiotype MCZ 2732, reconstructed skull with


original parts in dorsal view, from the “Baldwin bonebed” of the Lower
Permian El Cobre Canyon Formation of Arroyo del Agua in New Mexico.
134

FIGURE 8. Eryops sp., CMNH 34906; skull with left mandible; from the Lower Permian El Cobre Canyon Formation of Arroyo del Agua (1.6 km to NW,
west side of Rio Puerco, CM locality 1070) in New Mexico; A, dorsal skull roof; B, lateral view with mandible.

(probably) not very enlarged maxillary teeth, near the level of the Diplichnites tracks made by arthropleurids in the top of fluvial channel
ectopterygoid fangs. However, some differences in the proportions of sandstones in the El Cobre Canyon Formation. The co-occurrence of
the dorsal skull roof are present. The Permian skull (CMNH 34906, Fig. eryopid amphibians with the giant terrestrial arthropod Arthropleura
8) differs from the Pennsylvanian skull of similar size (NMMNH P was reported and discussed for the first time by Schneider and Werneburg
46379) in the following features: (1) a much wider interorbital region, (2) (1998) based on the example of the Early Permian (Asselian) Manebach
a more elongated snout region (premaxilla), which is slightly more pointed, Formation of the Thuringian Forest basin and the Late Pennsylvanian/
(3) a longer distance from the orbit to the naris, and (4) a more elongated Early Permian (Gzhelian/Asselian) Döhlen Formation of the Döhlen
interfrontal. Therefore, both skulls are different at the species level. basin, both in Germany. In the Thuringian Forest basin, the eryopid
In summary, it is possible to consider that the skull NMMNH P Onchiodon as well as the pelycosaurid reptile Haptodus have been dis
46379 from the Late Pennsylvanian of Cañon del Cobre clearly belongs covered in pebbly channel sandstones of an anastomosing to braided
to the genus Eryops, but the species assignment is uncertain. It is not river system in a floodplain environment (Werneburg, 2007). The vegeta
identical to the Early Permian Eryops specimens from New Mexico. The tion of this landscape consists of gallery forests of calamites along river
Pennsylvanian Eryops of New Mexico thus may represent a new spe courses, as well as swampy areas with coal-forming cordaite forests.
cies. This conclusion will be clarified by a better knowledge of the basal Arthropleura remains occur in shales and sandy siltstones together with
Eryops specimens from the Permian of Texas. plant remains from different growth sites, which were washed together
The discovery of Eryops in the El Cobre Canyon Formation also during flood events.
sheds new light on Late Pennsylvanian/Early Permian tetrapod/arthro In the Döhlen basin, Arthropleura has been discovered in
pod relationships. Lucas et al. (2005d) reported the occurrence of large pyroclastics with mesophilous vegetation between two coal seams
135
(Schneider and Barthel, 1997). Several meters below, another volcanic vironments include pelycosaurian-grade synapsids, such as Haptodus in
the Thuringian Forest basin and the Döhlen basin, as well as
ash horizon contains Limnopus tracks, which probably were produced
by eryopid amphibians (Hausse, 1910). A further occurrence of Limnopus undeterminated pelycosaurian-grade synapsid remains in the El Cobre
has been reported together with the arthropleurid track Diplichnites on Canyon Formation. Further synecological relationships, such as pos
the same bedding plane in coastal plain fluvial sandstones of the lower sible prey-predator relationships, are discussed by Schneider, Lucas,
Conemaugh Formation of Kentucky (Martino and Greb, 2009). The Werneburg and Rößler (2010, this volume).
repeated co-occurrence of eryopid skeletal fossils (Eryops in North
ACKNOWLEDGMENTS
America and its ecological counterpart Onchiodon in Europe) and the
Limnopus tracks, which were likely made by eryopid amphibians, with We thank our colleagues D. S Berman (CM), J. R. Bolt (FMNH
Arthropleura body remains or the arthropleurid track Diplichnites, can Chicago), F. A. Jenkins Jr. (MCZ), C. Mehling (AMNH), A. R. Milner
be regarded as indications of a characteristic biotope preference of eryopids (BMNH), C. Norris (AMNH), M. J. Ryan (CMNH Cleveland), C.
and arthropleurids. Obviously, eryopids spent most of their time in Schaff (MCZ) and R. R. Schoch (SMN Stuttgart) for the opportunity to
water and only occasionally ventured on land in fluvial environments. study specimens and helpful references. Reviews by D. S Berman (CM),
The sparsely vegetated landscape along the river courses, mainly calam A. Henrici (CM) and C. A. Marsicano (UBA) improved the manuscript.
ite-dominated gallery forests, were the preferred biotope of the giant This work was supported by a grant of the Deutsche Forschungs
arthropleurids. Rarely, fully terrestrially-adapted tetrapods in those en gemeinschaft (DFG-We 2833/2-1, DFG-We 2833/3-1, DFG-Schn 408/
12-1 and 2).

REFERENCES

Berman, D. S, 1993, Lower Permian vertebrate localities of New Mexico Lucas, S. G., Lerner, A. J., Hannibal, J. T., Hunt, A. P. and Schneider, J. W.,
and their assemblages: New Mexico Museum of Natural History and 2005d, Trackway of a giant Arthropleura from the Upper Pennsylva
Science, Bulletin 2, p. 11-21. nian of El Cobre Canyon, New Mexico: New Mexico Geological Society,
Boy, J. A., 1990, Über einige Vertreter Eryopoidea (Amphibia: Guidebook 56, p. 279-282.
Temnospondyli) aus dem europäischen Rotliegend (?höchstes Karbon – Martino, L.R. and Greb, S. G., 2009, Walking trails of the giant terrestrial
Perm). 3. Onchiodon: Paläontologische Zeitschrift, v. 64, p. 287-312. arthropod Arthropleura from the Upper Carboniferous of Kentucky:
Broom, R., 1913, Studies on the Permian temnospondylous stegocephalians Journal of Paleontology, v. 83, p. 140-146.
of North America: American Museum of Natural History Bulletin, v. 32, Milner, A. R., 1996, Systematics of the genus Eryops (Amphibia:
p. 563-595. Temnospondyli) and its possible biostratigraphical value: Journal of
Case, E. C., 1911, Revision of the Amphibia and Pisces of the Permian of Vertebrate Paleontology, v. 16, supplement to no. 3, p. 53A.
North America: Carnegie Publications, v. 146, p. 1-148. Moulton, J. M., 1974, A description of the vertebral column of Eryops based
Cope, E. D., 1877, Descriptions of extinct Vertebrata from the Permian on the notes and drawings of A.S. Romer: Breviora, v. 428, p. 1-44.
and Triassic formations of the United States: Proceedings of American Murphy, J. L., 1971, Eryopsid remains from the Conemaugh Group, Braxton
Philosophical Society, v. 17, p. 182-193. County, West Virginia: Southeastern Geology, v. 13, p. 265-273.
Cope, E. D., 1881, The Permian formation of New Mexico: American Romer, A. S., 1947, Review of the Labyrinthodontia: Bulletin Museum
Naturalist, v. 15, p. 1020. Comparative Zoology Harvard, v. 99, p. 1-368.
Douthitt, H., 1917, Eryops; Eryopsoides, gen nov. from the New Mexico Romer, A. S., 1952, Late Pennsylvanian and Early Permian vertebrates of
Permian: The Kansas University Science Bulletin, v. 10, p. 237-242. the Pittsburgh West Virginia region: Annals of the Carnegie Museum, v.
Harris, S. K., Lucas, S. G., Berman, D. S and Henrici, A. C., 2004, Vertebrate 33, p. 47-110.
fossil assemblage from the Upper Pennsylvanian Red Tanks Member of Sawin, H. J., 1941, The cranial anatomy of Eryops megacephalus: Bulletin
the Bursum Formation, Lucero uplift, central New Mexico: New Mexico of Museum of Comparative Zoology Harvard College, v. 88, p. 407
Museum of Natural History and Science, Bulletin 25, p. 267-283. 463.
Hausse, R., 1910, Fossile Tierfährten im Unterrotliegenden des Schneider, J. and Barthel, M., 1997, Eine Biocoenose mit Arthropleura im
Steinkohlenbecken im Plauenschen Grunde (des Döhlener Beckens) bei Rotliegend (?Unterperm) des Döhlen-Becken (Elbezone, Sachsen):
Dresden: Jahrbuch für das Berg- und Hüttenwesen in Sachsen 1910, p. 3 Freiberger Forschungsheft C , v. 466, p. 183-223.
19. Schneider, J. W. and Werneburg, R., 1998, Arthropleura und Diplopoda
Langston, W., Jr., 1953, Permian amphibians from New Mexico: Univer (Arthropoda) aus dem Unter-Rotliegend (Unter-Perm, Assel) des
sity of California Publications in Geological Sciences, v. 29, p. 349-416. Thüringer Waldes (Südwest-Saale-Senke): Veröffentlichungen
Lucas, S. G., 2006, Global Permian tetrapod biostratigraphy and Naturhistorisches Museum Schleusingen, v. 13, p. 19-36.
biochronology; in Lucas, S. G., Cassinis, G. and Schneider, J. W. (eds.), Stickler, L., 1899, Über den mikroskopischen Bau der Faltenzähne in Eryops:
Non-marine Permian biostratigraphy and biochronology. Geological Paläontographica, v. 46, p. 85-94.
Society, London, Special Publication 265, p. 65-93. Vaughn, P. P., 1958, On the geologic range of the labyrinthodont amphibian
Lucas, S. G., Rinehart, L. F. and Hunt, A. P., 2005a, Eryops (Amphibia, Eryops: Journal of Paleontology, v. 32, p. 918-922.
Temnospondyli) from the Upper Pennsylvanian of El Cobre Canyon, Werneburg, R., 1993, Onchiodon (Eryopidae, Amphibia) aus dem Rotliegend
New Mexico: New Mexico Museum of Natural History and Science, des Innersudetischen Beckens (Böhmen): Paläontologische Zeitschrift,
Bulletin 31, p. 118-120. v. 67, p. 343-355.
Lucas, S. G., Harris, S. K., Spielmann, J. A., Berman, D. S. and Henrici, A. C., Werneburg, R., 1996, Temnospondyle Amphibien aus dem Karbon
2005b, Vertebrate biostratigraphy and biochronology of the Pennsylva Mitteldeutschlands: Veröffentlichungen Naturhistorisches Museum
nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico: Schleusingen, v. 11, p. 23-64.
New Mexico Museum of Natural History and Science, Bulletin 31, p. Werneburg, R., 2007, Der “Manebacher Saurier” – ein neuer großer Eryopide
128-139. (Onchiodon) aus dem Rotliegend (Unter-Perm) des Thüringer Waldes:
Lucas, S. G., Harris, S. K., Spielmann, J. A., Berman, D. S, Henrici, A. C., Veröffentlichungen Naturhistorisches Museum Schleusingen, v. 22, p. 3
Heckert, A. B., Zeigler, K. E. and Rinehart, L. F., 2005c, Early Permian 40.
vertebrate biostratigraphy at Arroyo del Agua, Rio Arriba County, New Williston, S. W., 1899, Notes on the coraco-scapula of Eryops COPE: The
Mexico: New Mexico Museum of Natural History and Science, Bulletin Kansas University Quarterly, v. 8, p. 185-186.
31, p. 163-169.
136

Eryops attacks a xenacanth shark. Artwork by Matt Celeskey.


Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

137
A PARTIAL SKELETON OF OPHIACODON NAVAJOVICUS
(EUPELYCOSAURIA: OPHIACODONTIDAE) FROM THE
UPPER PENNSYLVANIAN OF CAÑON DELCOBRE, NEWMEXICO

SUSANK. HARRIS, SPENCERG. LUCASAND JUSTINA. SPIELMANN


New Mexico Museum of Natural History and Science, 1801 Mountain Road NW, Albuquerque, NM 87104

Abstract—A recently discovered partial skeleton of the pelycosaurian-grade synapsid Ophiacodon navajovicus,
the most complete yet found, is described from the Upper Pennsylvanian interval of the El Cobre Canyon
Formation (Cutler Group) at Cañon del Cobre, Rio Arriba County, New Mexico. The postcranial elements include
vertebrae representative of all regions of the vertebral column, a nearly complete pelvis, a complete right femur and
portions of other limb bones, as well as bones of the right pes. In addition, fragments of a right maxilla and dentary
were also found. Its stratigraphic distribution in New Mexico, Utah and Colorado indicate that O. navajovicus
characterizes the Cobrean land-vertebrate faunachron (lvf).

INTRODUCTION the holotype of a new species, Dimetrodon navajovicus. Ironically, an


other humerus, AMNH 4793 from lot No. 2285, was chosen for descrip
Although no longer recognized as the oldest species of the genus tion and illustration instead of the holotypic humerus, which is illus
Ophiacodon, O. navajovicus is distinctive in its diminutive size and trated here for the first time (Fig. 7A-B).
remarkable in its abundance in sediments of the El Cobre Canyon Forma The Yale Peabody Museum also received a large number of similar
tion in Cañon del Cobre, northern New Mexico. Since its initial descrip
tion as a new species by Case in 1907, only a few key skeletal elements specimens, sent by Baldwin to O. C. Marsh in the 1880s. After returning
from the Chicago expedition to Cañon del Cobre in 1911, Williston
have been adequately described or illustrated, although four institutions
(1914) compared material he had collected for the University of Chicago
hold large quantities of specimens identified as O. navajovicus. Here, we from the surface of the canyon floor to the Baldwin material. The similar
document the most complete specimen of O. navajovicus known (Figs. ity of the fossils was so great that it led him to speculate that all of it had
1-6) and discuss its utility as an index taxon of the Cobrean land-verte
brate faunachron (lvf). perhaps been recovered from the same bone bed. Consequently, Williston
In this paper, AMNH = American Museum of Natural History, (1914) erected the new genus Arribasaurus, not only for the two similar
New York; MCZ = Museum of Comparative Zoology, Cambridge, Mas groups of fossils in Chicago and New Haven, but also for the American
sachusetts; NMMNH = New Mexico Museum of Natural History and Museum material described and named by Case (1907). As justification
for removal of the material from the genus Dimetrodon, Williston (1914,
Science, Albuquerque; and YPM = Yale Peabody Museum, New Haven. p. 409) merely stated that “the teeth will exclude it from the
LOCALITY Sphenacodontidae.”
Romer and Price (1940) compared directly the Yale material with
The Ophiacodon navajovicus specimen described here, NMMNH that in the American Museum and concurred with Williston regarding the
P-43121, is from Cañon del Cobre at NMMNH locality 5660 in section conspecific nature of the two groups of fossils. However, they con
25, T24N, R6E. This locality is in the middle part of the locally exposed
cluded that no morphological differences separate “Arribasaurus” from
section of the El Cobre Canyon Formation (Cutler Group). It was pre Ophiacodon that are not arguably the result of ontogenetic changes,
served in brownish-red mudstone with blue concretions located approxi thereby relegating the former taxon to the status of junior subjective
mately 80 m below the contact of the El Cobre Canyon Formation and
synonym. As a result, Romer and Price (1940) transferred all the material
the overlying Arroyo del Agua Formation (Lucas and Krainer, 2005). A to the genus Ophiacodon, under the new combination Ophiacodon
majority of the O. navajovicus elements were found in close association
navajovicus.
in a small, shallow “pocket,” whereas some elements were weathered
Ophiacodon navajovicus was thought to be endemic to Cañon del
free and collected over an area of approximately 4 by 5 m. Diagnostic,
Cobre until a wider geographical range for the species was recognized by
but isolated, elements of O. navajovicus were abundant at two additional Vaughn (1962) when he identified material as Ophiacodon cf. O.
sites in the same stratigraphic interval.
navajovicus from the Halgaito “tongue” (raised to formation status by
PREVIOUS STUDIES Wengerd and Matheny, 1958) of the Cutler Group in San Juan County,
Utah. The most compelling evidence on which Vaughn (1962, p. 537)
In the early 1880s, David Baldwin collected numerous Ophiacodon based his identification was the distal portion of a humerus possessing a
fossils from Cañon del Cobre that were later recognized as representative “broad entepicondyle” that “looks exactly like that of AMNH 4777, the
of a single species. Part of this material reposited in the American Mu holotype of O. navajovicus.” In addition, Vaughn (1962, p. 537) based
seum of Natural History was assigned several Cope collection lot num the probable presence of O. navajovicus in the Halgaito Formation on
bers (No. 2285, No. 2298 and No. 2299), each of which included mul isolated vertebrae characterized by “the typically Ophiacodon pattern,”
tiple individuals (Figs. 7A-E, 8). Although Cope (1888) briefly described combined with “a flattened strip along the ventral midline of the cen
and illustrated a “pes” (later identified as a manus by Romer and Price, trum” (Vaughn, 1962, p. 537; Sumida et al., 1999).
[1940]) from this material without naming a new species, he apparently Another possible occurrence of Ophiacodon navajovicus outside
wrote on the label of a box containing a characteristic humerus the word of Cañon del Cobre was reported by Lewis and Vaughn (1965) from the
“navajovicus” (Case, 1907, p. 56). Cutler Group of southwestern Colorado. They described and illustrated
The suggested specific epithet was later read and adopted by Case an articulated string of five small vertebrae whose centra are character
(1907) when he selected a humerus, AMNH 4777 from lot No. 2298, as ized by “a flattened area bounded by a pair of longitudinal ridges” (Lewis
138
and Vaughn, 1965, p. C26). Although the linear measurements of the dentary fragment is 25 mm. The teeth are oval in horizontal section, with
vertebrae fall within the ranges listed by Romer and Price (1940, table 3) the long axis oriented mesiodistally.
for O. navajovicus, Lewis and Vaughn were undecided as to whether
specific assignment should be made to O. navajovicus or Ophiacodon Axial Skeleton
mirus. However, an illustrated transverse section through one of the Although Romer and Price (1940) state that vertebrae from all
vertebrae (Lewis and Vaughn, 1965, fig. 10A) clearly shows the narrow, regions of the column of Ophiacodon navajovicus are present in the
flat, ventral surface of the centrum diagnostic of O. navajovicus.
material collected by Baldwin, only dorsal and lumbar vertebrae were
In the summer of 2004, a NMMNH field party led by one of us
selected for description or illustration by previous workers (see Case,
(SGL) collected numerous associated axial and appendicular elements
1907, pl. 27, figs. 3-5; Williston, 1914, fig. 11, nos. 8 and 12). However,
from the floor of Cañon del Cobre of a single individual of Ophiacodon
Romer and Price (1940, table 3, p. 438) provide a table of measurements
navajovicus (NMMNHP-43121, Fig. 1). These elements, some of which for selected cervical, dorsal, lumbar and caudal vertebrae of O. navajovicus
were briefly mentioned and illustrated by Lucas et al. (2005), represent
specimens reposited in the AMNH and YPM collections. Measure
the single most complete skeleton of O. navajovicus known, and are the
ments of all vertebrae of NMMNH-43121 fall within the ranges given in
subject of this paper. this table (measurements for the atlas, axis and sacrals were not given).
NMMNH P-43121 includes a total of 34 vertebrae, some of
DESCRIPTION
which are articulated in short strings, and are representative of all regions
As discussed above, all the skeletal material described below was of the vertebral column (Figs. 1, 2-4). Among these are the atlantal
collected from a small, restricted area of a few square meters. Given the centrum, axis, articulated cervicals and the two sacrals that are described
close association of these fossils, their uniform presentation, the lack of and illustrated here for the first time.
duplicative elements, the extent to which many elements are readily Atlas-axis complex - The atlantal centrum is a small, cuboidal
articulated to each other and the consistency of their relative sizes, we element (Fig. 2G-K). It has an anteroposterior length of 10 mm, little
interpret these remains as pertaining to a single individual. more than half that of other preserved presacral vertebrae. The upper
half of its convex anterior face (width=10 mm) encloses a small foramen
Cranial Skeleton for passage of the notochord, whereas its posterior face (width=15 mm)
To date, cranial material of Ophiacodon navajovicus is largely bears a deep notochordal funnel. A shallow groove along the length of its
unknown and consists of isolated maxillae, quadrates and braincase ele dorsal surface forms the floor of the neural canal. Highly characteristic of
ments (Romer and Price, 1940) that have never been described or illus Ophiacodontidae (Romer and Price, 1940) is the ventral surface of the
trated. Recovered cranial elements of NMMNH P-43121 are limited to a atlantal centrum, which is firmly fused to the axial intercentrum and
fragmentary maxilla and dentary (Fig. 2A-F). exhibits a round facet on each side of its posterior edge for the reception
Maxilla - A fragmentary right maxilla contains two partial, seri of the axial rib capitulum. The ventral surface of the axial intercentrum
ally adjacent caniniform teeth: anteriorly the basal half of a large canini bears a prominent, rounded, longitudinal ridge. Total height of the com
form tooth is preserved, followed posteriorly by the anterior half of a bined elements is 17 mm.
hollow-cored tooth, which may represent a resorption pit. Posterior to The greatly enlarged axial neural spine has a maximum anteropos
the caniniform teeth are two very small alveoli (Fig. 2A-C). In many terior length of 29 mm at its distal margin (Fig. 2L). The straight poste
rior margin is vertical, whereas the anterior margin is angled 45 degrees
pelycosaurian-grade synapsids, the two largest teeth of the maxilla are
alternately replaced (Reisz, 1986), so that only one is functional at any anteriorly from the vertical so that it projects anteriorly 8 mm beyond
given time. The robustly constructed maxillary shelf (preserved length = the level of the anterior rim of the centrum. The thin anterior edge of the
26 mm) is enlarged above the caniniform teeth and is supported dorsally blade thickens posteriorly to within a very short distance of the poste
by a narrow buttress attached to the medial surface of the maxilla. rior margin, where it again thins. The prezygapophyses are very small
Dentary - A small, right dentary fragment preserves the striated and project slightly anterolaterally from the surface of the neural arch, in
bases of three marginal teeth, the fractured pulp cavity of a fourth tooth contrast to the robust, elongate postzygapophyses that project
and an empty alveolus (Fig. 2D-F). As preserved, the length of the posterolaterally 4 mm beyond the posterior rim of the centrum. All that

FIGURE 1. Ophiacodon navajovicus, NMMNH P-43121, incomplete skeleton laid out in plan view. Anterior is to the right.
139

FIGURE 2. Ophiacodon navajovicus, A-R, NMMNH P-43121. A-C, right maxilla fragment in A, lateral, B, medial and C, occlusal views. D-F, right
dentary fragment in D, lateral, E, medial and F, occlusal views. G-K, fused atlas centrum and axial intercentrum in G, posterior, H, anterior, I, right lateral,
J, dorsal and K, ventral views. L-N, axis in L, left lateral, M, ventral and N, anterior views. O-R, three cervical vertebrae in O, anterior, P, posterior, Q,
right lateral and R, ventral views.
140
remains of the transverse processes is a raised rugose, circular outline on are wedge-shaped in end view, and characterized by a narrow, flat ventral
either side of the dorsal margin of the centrum near its anterior rim (Fig. surface bordered by well-defined edges (Fig. 3D, H).
2L-N). A sharp, midventral keel extends the full length of the 21 mm long Four similar, but fragmentary vertebrae are also likely from the
centrum (Fig. 2M). dorsal region of the column, although their exact serial position cannot be
Postaxial Cervical Vertebrae - Three articulated postaxial cer determined due to the absence of their neural spines. Two dorsal proxi
vical vertebrae, probably representing positions 3-5 or 4-6, are nearly mal rib fragments are anteroposteriorly flattened and have a dorsoventral
complete, with the posteriormost one lacking the posterior portion of height of 17 mm. Anteriorly, the dichocephalus head is triangular in
the centrum and a right transverse process (Fig. 2O-R). The neural spines outline. A small indentation separates the tubercular and capitular sur
are inclined anteriorly 50 degrees from the vertical (Fig. 2Q), an indica faces, which are connected by a thin web of bone.
tion of the upward curvature of the column in the neck region (Reisz, Lumbar Vertebrae – Two lumbar vertebrae, the better preserved
1986). In lateral view the cervicals of Ophiacodon navajovicus closely of which is illustrated in Fig. 3I-J, consist of the centrum and the anterior
resemble those of Ophiacodon mirus illustrated by Williston and Case portion of the neural arch. The prezygapophyses are very small and the
(1913, fig. 24), but with the oblique orientation of the neural spine being facets are inclined medially 20 degrees from the horizontal, and posi
more pronounced in the four cervicals immediately following the axis. tioned close to the midline, which is very different from the condition of
The neural spines are stout, with a height of 20 mm and a maxi the zygapophyses in the preceding presacral series. In anterior view the
mum transverse width of 8 mm at the distal terminus, where there is a centra are smaller in diameter and rounder than those of the dorsal verte
lateral boss on both sides of the spine (Fig. 2N-Q). They are oval to brae. The narrow midventral surface of the centrum, though flattened, is
diamond-shaped in sagittal section. The medial margins of the pre and very gently concave between two slightly raised parallel edges.
postzygapophyses are widely separated and their facets are inclined 20 Sacral Vertebrae - The sacrum (Fig. 3K-N) consists of only two
degrees from the horizontal. Unlike the postzygapophyses of the axis, vertebrae, as is the case in all ophiacodontids for which this region of the
those of the postaxial cervicals do not extend beyond the level of the vertebral column is known (Romer and Price, 1940; Romer, 1956). The
posterior border of the centrum. first sacral vertebra in NMMNH P-43121, though missing most of its
The transverse processes originate high on the dorsolateral sur neural arch and spine, has a well-preserved centrum with a length of 14
face of the centrum, almost at the level of the zygapophyses (Fig. 2O-P). mm. In anterior view (Fig. 3L) the height of its lateral walls is enclosed by
They possess the typical ophiacodontid wing-shaped ventral flange or the fused, bulbous transverse processes, which form a continuous sur
web that extends from the diapophysis to the parapophysis at the ante face with the lateral border of the prezygapophyses, giving the centrum
rior edge of the centrum. Compared to transverse processes of the dorsal a massive, “swollen” appearance. It is nearly identical to the first sacral
vertebrae of this specimen, they have a slight posteroventral orientation vertebra included in AMNH 4793 (formerly cataloged in specimen lot
and are significantly longer (length = 11 mm) and more robust. AMNH No. 2285) (Fig. 7D). Near the ventral surface of the centrum, at
In contrast to the sharp, midventral keel of the axis, the centra of the terminus of the transverse process, there is a small, ventrolaterally
the postaxial cervical vertebrae possess a narrow ventral surface that is directed rib facet. The preserved distal end of the first sacral rib (Fig. 3O
flat and bordered by straight longitudinal edges (Fig. 2R). The centra are P) forms a cupped, fan-like expansion with a thickened anterior border,
wedge-shaped in cross section. Large, well-ossified intercentra with a broadly similar to that of Ophiacodon illustrated by Romer (1956, fig.
midventral length of 5 mm are articulated between all three centra. 133A). The proximal shaft of the rib is not preserved. As in the lumbar
Dorsal Vertebrae – Eleven dorsal vertebrae were recovered. Three vertebrae, the preserved anterior zygapophyses are very small and posi
of them are articulated though fragmentary, but the neural spine of the tioned close together at the midline.
first is complete and well-preserved (Fig. 3A-C). This spine has a height The tranverse processes of the second sacral vertebra (Fig. 3K,
of 27 mm, an anteroposterior length of 10 mm, and a transverse width of M-N) are also bulbous and fused to the lateral walls of the centrum, but
4 mm, making it one of the tallest and broadest neural spines of the are smaller than those of the first sacral, with an anteroventrally-directed
preserved column. Williston and Case (1913) divided the 20 dorsal ver rib facet that allowed the second sacral rib to support that of the first
tebrae of Ophiacodon mirus into three regions based on the dimensions sacral, instead of contacting the ilium (Reisz, 1986; see also illustration in
of the neural spines. By analogy with their model, the relative height, Romer, 1956, fig. 133A). The postzygapophyses are small, positioned
length and width of this neural spine would place these three articulated close to the midline and extend beyond the posterior border of the cen
vertebrae in the mid-dorsal region among serial positions 14-20, with trum by 4 mm. The midventral surface of the centrum of both sacrals, as
other dorsal vertebrae that have relatively tall, broad neural spines of in the lumbar vertebrae, is very gently ventrolaterally concave with slightly
moderate thickness. An isolated vertebra has a neural spine of similar raised, paralleledges (Fig. 3N).
proportions and thus most likely represents a mid-dorsal (Fig. 3E-H). Caudal Vertebrae – The first two caudal vertebrae, one of which
However, it differs from the three articulated dorsal vertebrae in its is illustrated in Figure 4A-C, have a length of 17 mm with short, moder
posteriorly positioned neural arch that overhangs the posterior border of ately stout rectangular neural spines. The neural arches are high and
the centrum by 5 mm (Fig. 3E). perpendicular to the long axis of the centrum. The one transverse process
A single, isolated neural spine is distinctively tall, narrow and thin is small (transverse length 9 mm), but robust and has a short ventral
(height = 30 mm; anteroposterior length = 8 mm; transverse width = 2 extension of the diapophysis. The zygapophyses are very widely sepa
mm), so its probable serial position was more posterior within the range rated, and their articular facets are gently inclined medioventrally. Similar
of vertebral positions 21-23, as these vertebrae possess gracile neural to the ventral surface of the axis, the anterior caudals have a sharp
spines. midventral keel (Fig. 4C).
Because the orientation of the neural arches of the dorsal vertebrae Positioned somewhat posteriorly to the first two caudal vertebrae
of NMMNH P-43121 is not inclined anteriorly like that of the cervicals, are caudals with a length of 14 mm, in which the transverse processes
but is instead normal to the long axis of the centrum, the dorsal neural extend halfway down the lateral walls of the centra (Fig. 4D-G). Their
arches are higher and form a less constricted neural canal. The inclination diapophyses are oval in cross section, with the long axis oriented longi
and spacing of the articular facets of the zygapophyses remain the same tudinally. Just ventral to the diapophyses are ridges that extend length
as in the cervicals, however, the postzygapophyses are not as robust and wise along the lateral centrum walls. The midventral surface of the cen
elongate. The single preserved transverse process is short and originates trum is v-shaped in cross section, but not as sharp as those of the axis
somewhat lower on the dorsal margin of the centrum than those of the and anterior caudals (Fig. 4E). Positioned somewhat posteriorly are three
cervicals (Fig. 3C). The maximum lateral length of its horizontal dorsal additional caudal vertebrae that are slightly smaller (length = 12 mm)
surface is 7 mm. Similar to the cervicals, the centra of the dorsal vertebrae with faint longitudinal ridges on the lateral walls of the centrum. The
141

FIGURE 3. Ophiacodon navajovicus, A-P, NMMNH P-43121. A-D, three articulated dorsal vertebrae in A, left lateral, B, anterior, C, posterior and D,
ventral views. E-H, isolated dorsal vertebra in E, left lateral, F, anterior, G, posterior and H, ventral views. I-J, lumbar vertebra in I, right lateral and J,
anterior views. K-N, articulated sacrum in K, left lateral, L, anterior, M, posterior and N, ventral views. O-P, right sacral rib fragment in O, dorsal and P,
ventral views.
142
ventral surface of the centrum is keeled and v-shaped in cross section. trough for the axial musculature. (Dorsoventral crushing of the dorsal
Six of the most posteriorly positioned preserved caudals decrease trough of the right iliac blade has resulted in widening of its proximal
in length from 12 mm to 9 mm. Their centra are laterally compressed,
portion). Ventral to the iliac blade, a short neck is formed with a width of
simple cylinders (Fig. 4H-K). Notably, the ventral surface of even the
20 mm. The thickened ventral base of the ilium has a width of 31 mm and
smallest distal caudal is flat with straight edges. contributes to more than one third of the acetabular surface. The sutural
Pectoral Girdle contact of the ilium with the pubis and ischium form a y-shaped pattern
within the acetabulum.
Elements of the pectoral girdle of Ophiacodon navajovicus are Pubis – The right and left pubic portions of the median symphy
rarely preserved. However, Romer and Price (1940, fig. 22) illustrated a sis are preserved and meet ventrally along deeply concave, unfinished
reconstruction of the scapulocoracoid of O. navajovicus based on frag surfaces that in life may have been filled with cartilage or connective
mentary portions of the dorsal scapular blade and the glenoid area of the tissue. The stout dorsal ridge of the pubis bears a prominent pubic
coracoid. NMMNH P-43121 includes a fragment of the ventral portion tubercle more than midway along its length of 70 mm. Its anterior margin
of the left scapular blade (Fig. 4L-M). With the exception of a short has a truncated appearance terminating in a blunt, unfinished surface that
length of the thickened anterior border just above the supraglenoid but was likely continued in cartilage. Below the dorsal pubic ridge the bone
tress, all edges are broken. Its maximum preserved anteroposterior length thins rapidly, with only fragments of its ventral portion being preserved.
is 59 mm. Immediately anteroventral to the anterior edge of the acetabulum the
pubis is pierced by the large, ovoid obturator foramen.
Forelimb Ischium - The preserved anterodorsal portion of the ischium is
Humerus – Recovered elements of NMMNH P-43121 only in wedge-shaped with a preserved length of 30 mm. Its wide, flat dorsal
clude the proximal portion of a right humerus. It has a maximum width of margin forms a crescentic ventral acetabular surface. The dorsomedial
41 mm (Fig. 4N-P), the same value as that listed for a proximal part of the edge of the ischium meets the ventral margin of the ilium in an open
humerus, AMNH 4784, by Romer and Price (1940, table 4, p. 156). In suture.
dorsal or ventral view, its outline is broadly triangular with a slightly Hind Limb
rounded, spiral articular surface. The proximal thickened ridge of the well
developed deltopectoral crest broadens distally to a triangular terminus. Femur - A complete, well-preserved right femur has a length of
Only a short portion of the proximal end of the shaft is preserved, but 107 mm (Fig. 5D-G). In ventral view the proximal head is rectangular
indicates that it was slender and round in cross section. These features with a slightly rounded articular surface with a width of 34 mm. A
compare closely with those of the humerus designated as the holotype of smooth, shallow intertrochanteric fossa bears a well-developed internal
Ophiacodon navajovicus, AMNH 4777 (Fig. 7A-B), as well as to an trochanter on its anterior border. In these features NMMNH P-43121
other characteristic humerus illustrated by Case (1907, plate 27, fig 2). conforms closely to the femur of Ophiacodon navajovicus illustrated by
Radius - The radius of NMMNH P-43121 was not recovered. Case (1907, pl. 27, fig. 1).
Ulna – The ulna is represented only by the proximal and distal The shaft is narrow and subcircular in cross section. A weakly
ends of the right ulna. The proximal head (Fig. 4 Q-R) possesses a developed adductor ridge arises just below the internal trochanter before
moderately developed, well-ossified olecranon process. Its transverse it continues somewhat diagonally across the shaft, with its most promi
width, measured proximal to the well-defined sigmoid notch, is 24 mm. nent portion assuming a nearly mid-ventral position. Romer and Price
The preserved portion of the ulnar shaft is somewhat mediolaterally (1940) illustrated an Ophiacodon navajovicus femur with the typical
compressed and narrows proximally from a width of 18 mm at its convex Ophiacodon adductor ridge pattern in which the ridge makes an abrupt
distal articular end to 10 mm at its preserved distal end. Measurements posterior turn below the intertrochanteric fossa, after which it extends
for both proximal and distal ends of the ulna of NMMNH P-43121 fall distally in a linear path along the posterior edge of the shaft to the
within the respective ranges listed for the ulna of Ophiacodon navajovicus posterior condyle. However, Romer and Price (p. 234) acknowledged
by Romer and Price (1940, table 4). that the O. navajovicus femur available to them was “imperfectly pre
Manus - The recovered elements of NMMNH P-43121 do not served” and that they reconstructed the ventral ridge system from “only
include bones of the manus. … a line of rugosities.”
In ventral or dorsal view both distal condyles are convex and fully
Pelvic Girdle ossified, an indication of the maturity of the specimen, with the poste
rior condyle extending much farther distally than that of the anterior
To date, known material of the mature pelvic girdle of Ophiacodon
navajovicus is limited to one specimen, MCZ 1595, which consists of an condyle. In direct contrast is a femur illustrated in dorsal view by Williston
incomplete right ilium and complete right ischium, though lack of diag (1914, fig. 11, no. 4) in which the distal margin of the bone is almost
nostic features in these elements makes their assignment to O. navajovicus straight with no visible development of the articular condyles, an obvi
questionable (see Appendix). These elements formed the basis of a re ously juvenile condition.
Tibia – In anterior or posterior view the proximal end of the tibia
construction of the pelvic girdle illustrated by Romer and Price (1940,
fig. 25). Williston (1914, fig. 11, no. 5) illustrated a small, round pubis (Fig. 5H-I) is roughly triangular with a maximum width of 31 mm, which
that is notched by the obturator foramen as tentatively diagnostic of the slightly exceeds the range of tibial widths listed by Romer and Price
genus “Arribasaurus.” These characters were later interpreted by Romer (1940, table 4) for Ophiacodon navajovicus. The convex articular sur
and Price (1940) as ontogenetically, but not taxonomically relevant within face (Fig.5J) bears a poorly-preserved cnemial crest. A wide, deep groove
the genus Ophiacodon. separates the larger anteromedial articular condyle from that of the some
NMMNH P-43121 includes a nearly complete and articulated what smaller posterolateral articular condyle. A missing segment sepa
rates the proximal head of the tibia from its shaft, which bears a sharp
pelvis (Fig. 5B). Unless otherwise stated, the following description is
ridge along its medial edge and is wedge-shaped in cross section. The
based on the more fully preserved right side (Fig. 5A-C) that notably
includes a nearly complete pubis of a mature individual. distal margin of the articular surface is convex in dorsal view, but concave
Ilium – In lateral or medial view the preserved anterior portion of in ventral view, with a width of 21 mm.
the iliac blade is narrow dorsoventrally and tapers posteriorly to the
Pes
level of breakage. An isolated distal end of the posterior process confirms
that the iliac blade ends in a very pointed posterior process. Medial and With the exception of the proximal tarsals, the pes of Ophiacodon
lateral walls of the iliac blade form a deep, triangular, dorsally-facing navajovicus is largely unknown. Case (1907) illustrates an astragalus in
143

FIGURE 4. Ophiacodon navajovicus, A-P, NMMNH P-43121. A-C, anterior caudal vertebra in A, left lateral, B, anterior and C, ventral views. D-G, mid
caudal vertebra in D, left lateral, E, ventral, F, anterior and G, posterior views. H-K, posterior caudal vertebra in H, left lateral, I, anterior, J, posterior and
K, ventral views. L-M, left scapular blade fragment in L, lateral and M, medial views. N-P, proximal right humerus in N, ventral, O, dorsal and P, proximal
views. Q-R, right ulna in two fragments in Q, anterior and R, posterior views.
144

FIGURE 5. Ophiacodon navajovicus, A-J, NMMNH P-43121. A, right lateral view of pelvis, C, medial view of right half of pelvis, and B, pelvis in dorsal
view. D-G, right femur in D, dorsal, E, ventral, F, distal and G, proximal views. H-J, right tibia in H, posterior, I, anterior and J, proximal views.

ventral view, whereas Williston (1914, fig. 11, nos. 2 and 6) illustrates roughened peripheral articular margins. Its width is 31 mm. Its lateral
both a juvenile and mature astragalus and calcaneum preserved in articu half is thin and flat, whereas its medial portion provides a thickened
lation. The pes of NMMNH P-43121 includes an astragalus, calcaneum, wedge for articulation with the astragalus. A small notch for the perforat
two articulated tarsals, two complete and several broken metatarsals, ing artery is positioned near the distal end of its contact with the astraga
and five partially articulated digits (Fig. 6). These elements were found lus.
closely associated, but not articulated; a reconstruction of the pedal Tarsals - Two tarsals are preserved in near articulation with one
elements (Fig. 6A) is based on published reconstructions of Ophiacodon another (Fig.6F-G). One, probably the 3rd distal tarsal, is pentangular in
pedes in the literature (e.g. Williston and Case, 1913, fig. 34C-D; Romer outline with slightly concave dorsal and ventral surfaces. The second is
and Price, 1940, fig. 41; Romer, 1956, fig. 195A). subequal, but much thicker and quadrangular in outline. The ventral
Astragalus - The astragalus is a broad, L-shaped element with an surface is flat and contrasts with a somewhat rounded dorsal surface that
elongate neck (Fig. 6D-E). The articular facet for the calcaneum is deeply bears several angular articular surfaces. These elements likely represent
concave in transverse section and bears a small notch near the distal end distal tarsals 3 and 4, respectively.
of the contact for passage of the perforating artery. In dorsal view the Metatarsals - Metatarsal I, preserved in articulation with its
distal surface of the horizontal medial process bears a large, rugose facet proximal phalanx, is short with a length of 14 mm and has a greatly
for the tibia. widened proximal articular surface (Fig. 6A, J-K). Another complete
Calcaneum - The calcaneum (Fig. 6B-C) is a wide, ovoid bone metatarsal with a length of 35 mm most likely represents metatarsal IV
with a broadly concave dorsal surface that nearly reaches the thickened, (Fig. 6A, H-I). Portions of three additional metatarsals have broad, flat
145

FIGURE 6. Ophiacodon navajovicus, A-O, NMMNH P-43121. A, reconstructed right pes in dorsal view. B-C, right calcaneum in B, dorsal and C, ventral
views. D-E, right astragalus in D, dorsal and E, ventral views. F-G, third and fourth tarsals in F, dorsal and G, ventral views. H-I, fourth metatarsal in H,
dorsal and I, ventral views. J-K, articulated pedal digits (from left to right digits 5 to 1) in J, dorsal and K, ventral views. L-O, ungual in L, dorsal, M,
ventral, N, lateral and O, proximal views. Upper left scale bar applies to A, upper right scale bar applies to B-E and middle right scale bar applies to F-O.
146

FIGURE 7. Ophiacodon navajovicus, type and referred specimens. A-E, AMNH 4777, holotype left humerus in A, ventral and B, dorsal views, C, proximal
tibia in anterior view, D, first sacral vertebra in anterior view, E, radius in anterior view. F-G, MCZ 1595(?) incomplete right pelvis in F, lateral(?) and G,
medial(?) views. Left scale bar applies to A-B, middle scale bar applies to C-E and right scale bar applies to F-G.

tened shafts with moderately expanded proximal and distal ends (Fig. the iliac blade”); (3) pubic tubercle better developed than in other species
6A). of Ophiacodon in which it is very small or represented only by a rugos
Phalanges - Articulated phalanges form segments offive digits of ity; and (4) ventral ridge of femur represented only by a line of rugosities
the right pes (Fig. 6A, J-K). Digit 1 is represented by a single phalanx
(but see above description of femur). In addition to these features, how
preserved in articulation with a short metatarsal described above. A prob ever, the postcranial skeleton of O. navajovicus can be readily distin
able digit 2 includes two slender, articulated phalanges, whereas digit 3 guished from that of O. mirus on the basis of the ventral surface of the
consists of three articulated phalanges that collectively curve slightly to centrum.
the left. Digit 4 includes three articulated phalanges and the distal portion In Ophiacodon mirus variation of the ventral surface of the cen
of a third. As articulated the digit is arched slightly dorsally and strongly trum within a single vertebral column typically progresses from a sharp
curved medially. Digit 5 includes the proximal two articulated phalanges midventral keel in the cervicals to a thicker, rounded midventral keel in
in which the first is twice the length of the second. Three disarticulated the dorsals, lumbars and sacrals, followed by anterior caudals with a
ungual phalanges, one of which is illustrated in Figure 6L-O, are some sharply keeled, midventral ridge, which becomes smaller in the mid
what more laterally compressed than in other species of Ophiacodon. caudals, where it lies between parallel grooves, and is replaced in the
Their positions shown in the reconstruction (Fig. 6A) are based on size. distal caudals by a rounded ventral surface (Williston, 1914; Kissel and
DISCUSSION Lehman, 2002). Thus, the ventral surface of the centrum in O. mirus is
always keeled, but varies from sharp to rounded.
As currently understood, the postcranial skeleton of Ophiacodon Ophiacodon navajovicus is characterized by a significantly dif
navajovicus is virtually indistinguishable from that of Ophiacodon mirus, ferent morphological pattern of the ventral surface of the centrum (Figs.
the only two species of Ophiacodon known from Cutler Group strata of 2M and R; 3D, H and N; 4C, E and K). In this species the ventral surface
northern New Mexico (Romer and Price, 1940; Lewis and Vaughn, 1965; of the axis typically possesses a sharp midventral keel, which is fol
Berman, 1993). Romer and Price (1940) listed only four characters by lowed abruptly in the postaxial cervicals by a distinctively flat midventral
which O. navajovicus can be specifically separated from O. mirus: (1) ridge that is defined by straight longitudinal edges. The narrow, flattened,
linear dimensions 4/5 that of O. mirus; (2) ilium more pointed posteri midventral ridge is retained posteriorly through the presacral column to
orly (restated by Reisz [1986, p. 69] as “pointed posterior process of the sacrals. The anterior caudals possess a sharply keeled, midventral
147

FIGURE 8. Ophiacodon navajovicus, referred specimens. A-I, AMNH 4793. A-B, right scapular blade in A, lateral and B, medial views. C-D, left humerus
in C, ventral and D, distal dorsal views. E-F, left acetabulum in E, lateral and F, medial views. G-H, proximal femur in G, ventral and H, dorsal views. I, two
dorsal vertebrae in right lateral view.

ridge, but in the mid-caudals the midventral ridge varies from a somewhat with well-defined lateral margins.
rounded to a v-shaped keel, which is followed in the most distal caudals The tendency toward a greater ventral flattening of the centra in
by a flattened ventral surface. Ophiacodon navajovicus when compared with other species of
Thus, the ventral surface of the centrum in Ophiacodon navajovicus Ophiacodon is present also in two genera considered to be primitive
is similar to that of Ophiacodon mirus only in the axis and anterior and members of a clade that includes Ophiacodon (Brinkman and Eberth,
middle sections of the caudal region. All other regions of the vertebral 1986). However, in these two taxa, Baldwinonus, collected from the
column in O. navajovicus, including most notably the postaxial cervicals Upper Pennsylvanian El Cobre Canyon Formation of Cañon del Cobre
and the distal caudals, are characterized by a narrow, flat, midventral keel (Romer and Price, 1940), and Stereophallodon, from the Lower Permian
148
Markley Formation (Wichita Group) of Texas (Romer, 1937), the pre temnospondyl Anconastes vesperus, the dissorophid Platyhystrix rugusus,
the edaphosaurid Edaphosaurus cf. E. novomexicanus and indetermi
sacral centra can be distinguished from those of O. navajovicus by their
greater size and subrectangular cross-sectional outline. nate eupelycosaurs.
An additional primitive character-state shared by Ophiacodon To date, only two potential records of Ophiacodon navajovicus
navajovicus and Stereophallodon is the mid-ventral position of the ad are known outside of Cañon del Cobre, both of which are in Cutler
Group strata of the Four Corners area. The first of these is from sedi
ductor ridge of the femur. Although the adductor ridge is positioned
similarly in two other ophiacodontid genera, Ruthiromia elcobriensis ments of the Halgaito Formation near Mexican Hat, Utah (Vaughn, 1962).
from Lower Permian (Wolfcampian) strata of Cañon del Cobre, (Eberth Arguments regarding the age of the Halgaito Formation are complex,
and Brinkman, 1983; Brinkman and Eberth, 1986; Spielmann and Lucas, leaving the position of the Virgilian-Wolfcampian boundary in that unit
this volume) and an ophiacodontid assigned to “Ophiacodon cf. O. mirus unclear. In addition to postcranial material identified as Ophiacodon cf.
(sic)” from the Upper Pennsylvanian (middle Virgilian) Ada Formation O. navajovicus, these strata yield a Coyotean tetrapod assemblage that
of Oklahoma (Kissel and Lehman, 2002), it is developed in these taxa to includes Diplocaulus, Phlegethonia?, a trimerorhachid, Eryops,
a much greater degree, so that it forms a well-ossified, prominent crest. Platyhystrix, Archeria, a limnoscelid, Edaphosaurus, Sphenacodon and
This condition is in sharp contrast to that of O. navajovicus and an araeoscelid? (Lucas, 2006).
In the Placerville area of southwestern Colorado the upper part of
Stereophallodon in which a weakly developed ridge is formed from a line
of rugosities (Brinkman and Eberth, 1986). the Cutler Formation yielded vertebrae that we believe can be assigned to
The morphology of the ventral surface of the centrum in Ophiacodon navajovicus based on the description and illustration pro
Ophiacodon navajovicus is potentially advantageous in the identifica vided by Lewis and Vaughn (1965, fig. 10A). (See discussion above in
tion of isolated postaxial through precaudal centra when present in com Previous Studies section.) Because the Cutler Formation in this area
bination with the typical ophiacodontid wing-shaped transverse process remains undivided, the exact age of the strata is unknown. However,
in which a web of bone extends from the diapophyses to the parapophysis strata of the Placerville area have yielded a tetrapod assemblage that
and the centrum has a wedge-shaped cross-sectional outline. Therefore, includes Eryops, Platyhystrix, a seymouriid, Diadectes, a captorhinid?
O. navajovicus can be considered of potential use in Late Pennsylvanian and a haptodontid (Lewis and Vaughn, 1965; Wideman et al., 2005), taxa
tetrapod biostratigraphy, provided its stratigraphic range can be well characteristic of the Coyotean lvf (Lucas, 2006), strongly suggesting the
established. section from which these fossils were recovered encompasses the Virgilian
Lucas et al. (2005) constructed a detailed vertebrate biostratigra Wolfcampian boundary.
phy in the Pennsylvanian-Permian redbeds of the El Cobre Canyon Clearly, detailed stratigraphic and biochronological resolution,
Formation in Cañon del Cobre that placed into precise lithostratigraphic currently lacking for both the Halgaito Formation, as well as strata of the
position three, temporally successive and distinct vertebrate fossil as Placerville area in the upper part of the Cutler Formation, is needed so
semblages that represent three land-vertebrate faunachrons (lvfs). The that all localities yielding Ophiacodon navajovicus fossils can be placed
into precise lithostratographic order enabling the lowest occurrence of
middle and upper assemblages belong to the Coyotean and Seymouran
lvfs that are of Virgilian-Wolfcampian and late Wolfcampian age, respec Sphenacodon relative to that of O. navajovicus to be determined. Until
tively. Ophiacodon navajovicus is restricted to the lower assemblage at such time, O. navajovicus is best considered to be restricted to the
Cañon del Cobre that represents an older, entirely Pennsylvanian lvf, the Cobrean lvf, as it is in the El Cobre Canyon Formation in Cañon del
Cobrean. Cobre, the only location in North America where substantial assem
The Cobrean lvf is defined as beginning with the FAD (first blages of vertebrate fossils are documented as closely bracketing the
appearance datum) of the temnospondyl Eryops, and ending with the Pennsylvanian-Permian boundary.
beginning of the Coyotean, defined by the FAD of the pelycosaurian ACKNOWLEDGMENTS
grade synapsid Sphenacodon. Index taxa of Cobrean time currently in
clude the diadectids Diasparactus and Desmatodon and the limnoscelid We thank AMNH and MCZ for collection access. Al Lerner,
Limnoscelis. Other Pennsylvanian tetrapod fossils of this assemblage, Sterling Nesbitt, Randy Pence and Larry Rinehart assisted in fieldwork.
according to Lucas (2005), include Eryops sp., the trematopid Careful reviews by Dave Berman, Amy Henrici and Sean Modesto greatly
improved the manuscript.

REFERENCES

Brinkman, D. and Eberth, D.A., 1986, The anatomy and relationships of biochronology; in Lucas, S. G., Cassinis, G. and Schneider, J. W., eds.,
Stereophallodon and Baldwinonus (Reptilia, Pelycosauria): Breviora, Non-marine Permian biostratigraphy and biochronology: London, Geo
no. 485, p. 1-34. logical Society, Special Publications, v. 265, p. 65-93.
Case, E. C., 1907, Revision of the Pelycosauria of North America: Carnegie Lucas, S. G., Harris, S. K., Spielmann, J. A., Berman, D. S and Henrici, A. C.,
Institute of Washington, Publication 55, 176 p. 2005, Vertebrate biostratigraphy and biochronology of the Pennsylva
Cope, E. D., 1888, Systematic catalogue of the species of Vertebrata found nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico:
in the beds of the Permian epoch in North America, with notes and New Mexico Museum of Natural History and Science, Bulletin 31, p.
descriptions: American Philosophical Society Transactions, v. 16, p. 128-139.
285-297. Lucas, S. G. and Krainer, K., 2005, Stratigraphy and correlation of the
Eberth, D.A. and Brinkman, D., 1983, Ruthiromia elcobriensis, a new Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New
pelycosaur from El Cobre Canyon, New Mexico: Breviora, no. 474, p. Mexico Geological Society, Guidebook 56, p. 145-159.
1-26. Reisz, R. R., 1986, Pelycosauria: Encyclopedia of Paleoherpetology, Part
Kissel, R. A. and Lehman, T. M., 2002, Upper Pennsylvanian tetrapods 17A, 102 p.
from the Ada Formation of Seminole County, Oklahoma: Journal of Romer, A.S., 1937, New genera and species of pelycosaurian reptiles: Pro
Paleontology, v. 76, p. 529-545. ceedings of the New England Zoological Club, 16, p. 89-96.
Lewis, G. E. and Vaughn, V. V., 1965, Early Permian vertebrates from the Romer, A. S., 1956, Osteology of the reptiles: The University of Chicago
Cutler Formation of the Placerville area, Colorado: U. S. Geological Press, 772 p.
Survey, Professional Paper, 503-C, 50 p. Romer, A. S. and Price, L. I., 1940, Review of the Pelycosauria: Geological
Lucas, S. G., 2006, Global Permian tetrapod biostratigraphy and Society of America Special Paper, v. 28, p. 1-538.
149
Sumida, S. S., Lombard, R. E., Berman, D. S and Henrici, A. C., 1999, Late Corners region: Bulletin of the American Association of Petroleum
Paleozoic amniotes and their near relatives from Utah and northeastern Geologists, v. 42, p. 2048-2106.
Arizona, with comments on the Permian-Pennsylvanian boundary in Wideman, N. K., Sumida, S. S. and O’Neil, M., 2005, A reassessment of the
Utah and northern Arizona; in Gillette, D., ed., Vertebrate paleontology taxonomic status of the materials assigned to the Early Permian tetra
in Utah: Utah Geological Survey, Miscellaneous Publication, 99-1, p. pod genera Limnosceloides and Limnoscelops: New Mexico Museum of
31-43. Natural History and Science, Bulletin 31, p. 358-362.
Vaughn, P. P., 1962, Vertebrates from the Halgaito Tongue of the Cutler Williston, S. W., 1914, The osteology of some American Permian verte
Formation, Permian of San Juan County, Utah: Journal of Paleontology, brates: Journal of Geology, v. 22, p. 364-419.
v. 36, p. 529-539. Williston, S. W. and Case, E. C., 1913, Description of a nearly complete
Wengerd, S. A. and Matheny, M. L., 1958, Pennsylvanian System of Four skeleton of Ophiacodon Marsh: Carnegie Institution of Washington,
Publication no. 181, p. 37-59.

APPENDIX
Referred specimens of Ophiacodon navajovicus

AMNH
FR 4776: Forefoot, mostly articulated.
FR 4777: Partial skeleton, vertebrae and forelimbs, humerus, proximal and distal ulna, parts of femur, distal ?tibia and fibula, astragalus
(Fig. 7A-E).
FR 4781: Hindfoot, tibia and parts of other limb bones and several vertebrae.
FR 4783: Fragmentary skeleton (with no. 4782) chiefly vertebrae and foot bones, proximal and distal ulna, proximal tibia and fibula.
FR4784: Fragmentary skeleton including dorsal, sacral and caudal vertebrae, parts of various limb and foot bones, right and left proximal
femora, etc.
FR4789: Skeleton fragments, vertebrae, limb and foot bones.
FR 4791: Skeleton fragments, vertebrae, limb bone fragments, femur, tibia and fibula, etc.
FR 4793: Humerus, half of femur, partial pelvis, vertebrae and foot bones, astragalus and 3 podials and phalanges (Fig. 8A-I).
FR4794: Skull and jaw fragments, articular distal tibia, ulna?, phalanges and scrap.
FR 4799: vertebrae.
FR 4801: Scapulocoracoid fragment, partial ilium, 4 proximal ends and distal end of femora, vertebra and bone fragment.

MCZ
MCZ 1595: An incomplete right pelvic girdle, consisting of a complete ischium and incomplete ilium. Collected by White and Price in 1935
from El Cobre Canyon. They originally identified it as a left pelvic girdle and incorrectly interpreted the ischium as the ilium.
Though originally assigned to Ophiacodon navajovicus by Romer and Price (1940), given the lack of any diagnostic features of
the specimen, its referral to O. navajovicus is questionable (Fig. 7F-G).
MCZ 2007: A large lot of numerous boxes of postcranial elements collected from Cañon del Cobre. Elements include jaw fragments,
fragmentary pectoral girdles, femora, tibiae and fibulae. At least one femur fragment from these lots appears to have the ventral
ridge characteristic of Ophiacodon navajovicus, much of the other material is probably not diagnostic.
MCZ 8079: Nine vertebral centra, some have square ventral keels, while others have triangular ventral keels. The centra with the square
ventral keels can be referred with confidence to Ophiacodon navajovicus. No locality information listed beyond “From El Cobre
Canyon.”
150

Arthropleura against a Late Pennsylvanian moon, 300 year million years ago in Cañon del Cobre, New Mexico. Artwork by Matt Celeskey.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

151
RE-EVALUATION OF RUTHIROMIAELCOBRIENSIS (EUPELYCOSAURIA:
OPHIACODONTIDAE?) FROM THE LOWER PERMIAN
(SEYMOURAN?) OF CAÑONDELCOBRE, NORTHERN NEWMEXICO

JUSTINA. SPIELMANN AND SPENCER G. LUCAS


New Mexico Museum of Natural History and Science, 1801 Mountain Rd NW, Albuquerque, NM 87104-1375

Abstract—Ruthiromia elcobriensis is a eupelycosaur from the Lower Permian strata of the Cañon del Cobre, Rio
Arriba County, New Mexico. The taxon is known only from the holotype, a partial skeleton consisting of skull
fragments, isolated dorsal vertebrae, articulated vertebral series of “lumbar,” sacral and anterior caudals, incom
plete pelvic girdle, left humerus and much of the right hindlimb. The type locality of R. elcobriensis is imprecisely
known, but available data suggest it was collected from the lower part of the Arroyo del Agua Formation,
considered Seymouran in age and equivalent to the late Wolfcampian marine stage. Ruthiromia elcobriensis remains
a problematic taxon within the Eupelycosauria. Re-evaluation identifies only two diagnostic characters of the
taxon, the pinched-in presacral centra and distally expansive deltopectoral crest. Comparisons to othereupelycosaurs
indicate that R. elcobriensis displays morphologic features that are present in both varanopids and ophiacodontids,
and given this combination of characteristics R. elcobriensis is assigned tentatively to the Ophiacodontidae, and
likely a close relative of Ophiacodon navajovicus.

INTRODUCTION STRATIGRAPHIC POSITION AND AGE


Eupelycosaurs were the top terrestrial predators of the Early Eberth and Brinkman (1983, p. 1-2) provided the following sum
Permian and significant collections of various eupelycosaur taxa have mary of the stratigraphic position of Ruthiromia elcobriensis: “In 1965
been made from Texas and New Mexico. Many eupelycosaurs are known A. Lewis and S. Olsen collected the larger part of the postcranial skeleton
from numerous specimens (e.g., Sphenacodon from the Arroyo del Agua and disarticulated cranial fragments of a pelycosaur from the west wall of
bonebeds of north-central New Mexico: Lucas et al., 2005b) and thus the canyon.” The canyon they refer to is Cañon del Cobre (= El Cobre or
provide an extensive comparative record. Unfortunately, no recent at Copper Canyon), Rio Arriba County, New Mexico. Most of the strata
tempts have been made to significantly revise eupelycosaurian relation exposed in Cañon del Cobre are Late Pennsylvanian to Early Permian in
ships, leaving many species with very general diagnoses that are anec age (Lucas and Krainer, 2005; Lucas et al., 2005a, this volume), including
dotal in nature and genera that are taxonomically oversplit (e.g., 11 recog the west wall where the holotype of R. elcobriensis was collected. We
nized species of Dimetrodon characterized mostly by roughly defined attempted to relocate the excavation site of the holotype of R. elcobriensis,
size grades: Reisz, 1986). As a result, taxonomic uncertainty makes but failed because of the lack of specific information about the site’s
incomplete eupelycosaurian specimens difficult to diagnose, especially position.
when lacking a complete skull or dorsal vertebrae (e.g., the non-axial Based on the detailed stratigraphy of Lucas and Krainer (2005)
postcrania of Dimetrodon and Sphenacodon are identical). and mapping of Kempter et al. (2007: though see modifications by Lucas
Unfortunately, this is the case with Ruthiromia elcobriensis, a et al., this volume) we can now tentatively indicate the stratigraphic
taxon named by Eberth and Brinkman (1983) based on an incomplete position of the site where the Ruthiromia elcobriensis holotype was
skeleton, consisting of the occipital and quadrate fragments, posterior collected. Along the west wall of the Cañon del Cobre the Arroyo del
left mandible, atlantal centrum and axial intercentrum, cervical rib, four Agua Formation is exposed, with the underlying El Cobre Canyon For
disarticulated dorsal vertebrae, a disarticulated neural arch and partial mation forming the canyon floor (Kempter et al., 2007; Lucas et al., this
neural spine, an incomplete, articulated vertebral column and pelvis, and volume). The lithology of these formations is distinct, with the El Cobre
an incomplete distal portion of the right hindlimb. Eberth and Brinkman Canyon Formation being brown in color with numerous rhizolith hori
(1983, p. 22-23) assigned Ruthiromia elcobriensis to the Varanopidae zons and the Arroyo del Agua Formation being orange with abundant
based on it possessing “excavated neural arches and the presence of a calcrete nodules. The “orange” (moderately reddish brown) siltstone
shallow fossa on the proximal surface of the deltopectoral crest,” fea matrix of the partially prepared vertebral column of R. elcobriensis,
tures that in combination are unique to Varanopidae (Langston and Reisz, (Lucas and Krainer, 2005, table 1) indicates the holotype was collected
1981). from the Arroyo del Agua Formation. This then means that it is from the
One of us (JAS) recently re-examined the holotype and only uppermost tetrapod assemblage zone in Cañon del Cobre (Lucas et al.,
known material of Ruthiromia elcobriensis and found discrepancies in 2005a), which includes the diadectomorph Diadectes, the sphenacodontid
the original description of Eberth and Brinkman (1983). Here, we discuss eupelycosaur Sphenacodon ferox and the dissorophid amphibian
the stratigraphic position and probable age of Ruthiromia elcobriensis, Platyhystrix. Lucas et al. (2005a) and Lucas (2005, 2006) assigned this
modify its diagnosis, reevaluate its taxonomic position within upper assemblage zone to the Early Permian Seymouran land-vertebrate
Eupelycosauria and photographically document portions of the holo faunachron, based on correlations with the type section of the Arroyo
type for the first time. del Agua Formation at Mesa Montosa, where Seymouria, the principal
Institutional abbreviations: MCZ, Museum of Comparative index taxon of the Seymouran, is present. Lucas (2006) indicates that the
Zoology, Harvard University, Cambridge, MA; and NMMNH, New Seymouran lvf is equivalent to the late Wolfcampian marine stage.
Mexico Museum of Natural History and Science, Albuquerque NM.
152
SYSTEMATIC PALEONTOLOGY the astragalus represents 35 percent of the total height of that element.”
Our reexamination of the holotype and relevant literature allows an evalu
Suborder EUPELYCOSAURIA ation of these characteristics:
Family ?OPHIACODONTIDAE [1] Linear measurements approximately 30 percent greater than
Genus Ruthiromia Eberth and Brinkman, 1983 those of Varanodon: Eberth and Brinkman (1983) interpreted Ruthiromia
Ruthiromia elcobriensis Eberth and Brinkman, 1983 elcobriensis as a varanopid and used its large size to distinguish it from
all other varanopids. However, this was prior to the recognition of
Figs. 1-5
Watongia meireri as a varanopid and collection of additional larger speci
1983 Ruthiromia elcobriensis: Eberth and Brinkman, figs. 2-8, 10 mens of Varanodon agilis, which have larger humeri than R. elcobriensis
15. (humeral length: R. elcobriensis = 105.6 mm; Watongia meieri = 130 mm;
Varanodon agilis = 124 mm; data from Eberth and Brinkman, 1983, table
Holotype: MCZ 3150, partial skeleton consisting of occiput and
1 and Reisz and Laurin, 2004, table 1). Thus, Ruthiromia is no longer the
quadrate fragments (Fig. 1B-C), a posterior portion of the left mandible, largest varanopid. We interpret Ruthiromia as demonstrating characteris
atlantal centrum and axial intercentrum, a cervical rib, four disarticulated tics of both varanopids and ophiacodontids, and cannot assign it to
dorsal vertebrae (e.g., Fig. 1A); an isolated disarticulated neural arch and either family with confidence (see below for further discussion). The
partial neural spine, an incomplete, articulated vertebral column and
size of R. elcobriensis is not considered a useful diagnostic feature, as its
pelvis (Fig. 2), left humerus (Fig. 3) and an incomplete, distal portion of size, based on femur length, is considerably smaller than the largest
a right hindlimb (Figs. 4-5).
species of Ophiacodon (femoral length of R. elcobriensis: 116.5 mm
Referred specimens: Only known from the holotype. from Eberth and Brinkman, 1983, p. 9, table 1 and Ophiacodon
Type locality and horizon: Arroyo del Agua Formation, Cutler
retroversus: 140-192 mm from Romer and Price, 1940, p. 465, table 4).
Group in Cañon del Cobre, Rio Arriba County, New Mexico. See above [2] Dorsal and lumbar centra deeply pinched-in laterally, just
for interpretation of the stratigraphic position of the type locality, exact
below the level of the notochordal pit, giving the centra an hour-glass
collection site unknown.
Revised diagnosis (modified from Eberth and Brinkman, appearance: this is the most distinctive feature of Ruthiromia elcobriensis
and contrasts with the morphology of the centra of all other non-varanopid
1983): A large eupelycosaur distinguished from all others by the follow eupelycosaurs (Fig. 6). We have modified the original character as listed
ing features: dorsal and lumbar centra pinched-in laterally, just below the
by Eberth and Brinkman (1983) because the three isolated dorsal centra
level of the notochordal pit, giving the centra an hourglass shape in cross
described also possess this character, indicating this feature may be
section; the deltopectoral crest of the humerus is massively developed
present in all presacral centra, not just the centra immediately anterior to
and continues distally as a sharp ridge below the area of the pectoralis the sacrum.
muscle insertion. [3] Each of the eight [posteriormost presacral] centra displays a
Description: Eberth and Brinkman (1983) described the holo broad rounded ventral surface: we have removed this feature from the
type Ruthiromia elcobriensis in detail, obviating the need for further
of diagnosis because it is also present in various ophiacodontids (e.g.,
description here (other than the emendations noted below). Stereophallodon and Clepsydrops: Romer and Price, 1940, fig. 17) and
Discussion: Eberth and Brinkman (1983, p. 2) diagnosed
thus cannot be used to distinguish Ruthiromia elcobriensis from all other
Ruthiromia elcobriensis as a varanopid on the basis of the five following pelycosaurian-grade synapsids. However, it can be noted that these
characters “[1] linear measurements approximately 30 percent greater
than those of Varanodon…[2] the first eight centra directly anterior to ophiacodontids lack the pinched-in presacral centra of R. elcobriensis.
[4] The deltopectoral crest of the humerus is massively developed
the pelvis deeply pinched-in laterally, just below the level of the noto
and continues distally as a sharp ridge below the area of pectoralis muscle
chordal pit, giving the centra an hour-glass appearance; [3] each of the
insertion: this prominent crest is retained in the diagnosis of R. elcobriensis,
eight centra displays a broad rounded ventral surface; [4] the deltopectoral
as it distinguishes this taxon from all other pelycosaurian-grade synapsids.
crest of the humerus is massively developed and continues distally as a
This feature is also independent of size, as Dimetrodon limbatus, which
sharp ridge below the area of pectoralis muscle insertion; [5] the neck of is larger than R. elcobriensis (humerus length of R. elcobriensis: 105.6
mm from Eberth and Brinkman, p. 9, table 1 and D. limbatus: 180-208
mm from Romer and Price, 1940, p. 456, table 4) lacks such a distal
extension of the deltopectoral crest (Romer and Price, 1940, pl. 29C and
Brinkman, 1988, fig. 3).
[5] The neck of the astragalus represents 35 percent of the total
height of that element: we have removed this feature from the diagnosis
because a long neck of the astragalus is also present in Ophiacodon
navajovicus (Harris et al., this volume, fig. 9a, d-e). In O. navajovicus,
NMMNH P-43121, the height of the neck of the astragalus (13.3 mm) is
38.8 percent of the total height of the astragalus (34.3 mm). Thus, a
relatively tall astragular neck cannot be used to separate Ruthiromia
elcobriensis from all other pelycosaurian-grade synapsids.
Overall, only two of the five characteristics originally diagnosed
by Eberth and Brinkman (1983) are still useful in identifying Ruthiromia
elcobriensis. However, the pinched-in dorsal and lumbar centra and dis
tally expansive deltopectoral crest are unique among eupelycosaurs and
allow R. elcobriensis to be recognized from isolated postcranial fossils.
TAXONOMIC PLACEMENT
FIGURE 1. Ruthiromia elcobriensis, MCZ 3150, holotype. A, isolated dorsal
vertebra in left lateral view, note pinched-in centrum and broad rounded Removal from the Varanopidae
ventral centrum surface (illustrated in Eberth and Brinkman, 1983, fig.
7[left]). B-C, opisthotic/supraoccipital fragment in B, posterior and C, As noted above, Eberth and Brinkman (1983) identified two fea
anterior views. Abbreviations: op, opisthotic; st, supratemporal. tures – the excavated neural arches and the presence of a shallow fossa on
153

FIGURE 2. Ruthiromia elcobriensis, MCZ 3150, holotype. A-C, vertebral column and pelvis. A-B, vertebral column in A, left lateral and B, dorsal views.
C, Close-up of dorsal neural spine and arch in dorsolateral view. Vertebrae are numbered in series, with 1 being anterior. Abbreviations: c, centrum; ex,
excavation in neural spines; is, ischium; n, neural spines: na, neural arch; p, pubis.
154

FIGURE 3. Ruthiromia elcobriensis, MCZ 3150, A-F, holotype left humerus in A, dorsal, B, ventral, C, posterior, D, anterior, E, proximal and F, distal
views. Abbreviations: dpc, deltopectoral crest; ent, entepicondyle.

the proximal deltopectoral crest–that identified Ruthiromia elcobriensis brae 5-8 of the presacral series) possess enough neural that are complete
as a varanopid, following Langston and Reisz (1981), who noted that enough for identification of excavations. Of these four, only two are
these two features in combination are unique to the Varanopidae. Our completely unobstructed, whereas the other two have matrix filling the
reexamination of the holotype of R. elcobriensis has called these two neural arch excavations. Thus, only four “lumbar” vertebrae demonstra
features into question, and in so doing modified the taxonomic placement bly possess the excavated neural arches (Fig. 2), and this is the only
of R. elcobriensis. character shared between varanopids and R. elcobriensis, compared to
Most of the vertebrae of Ruthiromia elcobriensis are represented the centra and femoral morphology common to both ophiacodontids and
by incomplete centra. Only four of the eight “lumbar” vertebrae (verte R. elcobriensis.
155

FIGURE 4. Ruthiromia elcobriensis, MCZ 3150, A-F, holotype right femur in A, dorsal, B, posterior, C, anterior, D, ventral, E, proximal and F, distal
views. G-J, holotype left femur fragment in G, ventral, H, anterior, I, dorsal and J, proximal views. K-M, holotype fibula in K, lateral, L, medial and M,
proximal views. Abbreviations: ar, adductor ridge.
156

FIGURE 5. Ruthiromia elcobriensis, MCZ 3150, A-B, incomplete right tibia and pes in A, dorsal and B, dorsolateral views.

The shallow fossa on the deltopectoral crest of the humerus could


elcobriensis in the Varanopidae, our re-evaluation of the taxon has led us
not be confirmed in our examination of the holotype. It appears to be the to place it outside of this family and place it tentatively within the
result of damage to the specimen and is far less prominent than the Ophiacodontidae. R. elcobriensis possesses excavated neural arches, but
deltopectoral crest fossa in some varanopids (compare Fig. 3B-E and lacks a deltopectoral crest fossa, precluding it from inclusion in the
Eberth and Brinkman, 1983, fig. 11B-C, with the humerus of Aerosaurus Varanopidae. Berman et al. (1995) provided a diagnosis for the
greenleeorum illustrated by Langston and Reisz, 1981, fig. 15B-D and Ophiacodontidae that included ilium possesses dorsal groove or trough
the Richards Spur varanodontine illustrated by Maddin et al., 2006, fig.
and posterior process of ilium is slender and extends close to the level of
4A-C). the posterior limit of the ischium, two features present in R. elcobriensis
Placement within the Ophiacodontidae (compare R. elcobriensis in Eberth and Brinkman, 1983, fig. 12a-b to
Ophiacodon in Reisz, 1986, fig. 22b-c). In addition, both R. elcobriensis
Whereas Reisz (1986) only tentatively placed Ruthiromia and Ophiacodon navajovicus share an adductor ridge that extends across
157
eliminates one of the diagnostic features of O. navajovicus listed by
Romer and Price (1940) and Reisz (1986), leaving the pointed posterior
process of the ilium and the well-developed pubic tubercle as the diag
nostic characteristics of the taxon. Also, R. elcobriensis possesses centra
morphology similar to other ophiacodontids and astragalus proportions
comparable to O. navajovicus, to the exclusion of all othereupelycosaurs.
Nevertheless, O. navajovicus cannot be synonymous with R. elcobriensis,
as it lacks excavated neural arches, pinched-in dorsal and lumber centra
and a prominent distal expansion of the deltopecotral crest.
Eberth and Brinkman (1983, fig. 16C) provided four possible
taxonomic positions for Ruthiromia elcobriensis based on the derived
feature of excavated neural arches: as the sister taxon to the Varanopidae;
the sister taxon to (Edaphosauridae + Sphenacodontidae); within
Edaphosauridae as the sister taxon to an undescribed taxon from the
Pennsylvanian of Kansas; and the sister taxon to the Sphenacodontidae.
It should be noted that Eberth and Brinkman (1983, fig. 16C) placed the
Varanopidae as the sister taxon to the Caseidae, a phylogenetic position
not followed by later studies (e.g., Reisz, 1986, fig. 41). Reisz (1986, fig.
41) placed Varanopidae as the sister taxon to (Ophiacodontidae +
(Edaphosauridae + (Sphenacodontidae + Therapsida))). This study fol
lows most other classifications that place Varanopidae and
Ophiacodontidae as closely related, with either Varanopidae (Reisz, 1986,
fig. 41) or Ophiacodontidae (Reisz, 1986, fig. 40[lower] modified from
Reisz, 1980, fig. 17) as the more inclusive taxon. Unfortunately, Reisz
(1986, table 3) did not provide any postcranial characteristics to identify
the various eupelycosaurian families in his analysis, though R. elcobriensis
does possess a disproportionately large entepicondyle (Fig. 3), a feature
FIGURE 6. Dorsal vertebrae of Ruthiromia elcobriensis compared to various used by Reisz (1986, p. 66) to identify the Ophiacodontidae. Given the
other eupelycosaurs in cross section. Images from Romer and Price (1940, similarities in the femoral morphology and proportions of the astragalus
fig. 17) and Eberth and Brinkman (1983, figs. 7-8). Higher level taxonomy between R. elcobriensis and Ophiacodon navajovicus we tentatively
after Reisz (1986). Note pinched-in centra of R. elcobriensis. place R. elcobriensis in the Ophiacodontidae, noting that it is likely most
closely related to O. navajovicus.
the shaft with its most prominent portion assuming a nearly mid-ventral
ACKNOWLEDGMENTS
position (Harris et al., this volume), though the adductor ridge of R.
elcobriensis is more well developed, extending away from the femoral Chuck Schaff provided collection access for JAS at the MCZ.
shaft a much greater distance than the adductor ridge of O. navajovicus Susan Harris, David Berman and Stuart Sumida provided helpful reviews
(Fig. 4). The presence of adductor ridges on the femora of both these taxa that improved the manuscript.

REFERENCES

Berman, D.S, Reisz, R.R., Bolt, J.R. and Scott, D., 1995, The cranial anatomy Langston, W., Jr. and Reisz, R.R., 1981, Aerosaurus wellesi, new species, a
and relationships of the synapsid Varanosaurus (Eupelycosauria: varanopseid mammal-like reptile (Synapsida: Pelycosauria) from the
Ophiacodontidae) from the Early Permian of Texas and Oklahoma: Lower Permian of New Mexico: Journal of Vertebrate Paleontology, v.
Annals of Carnegie Museum, v. 8, p. 172-180. 1, p. 73-96.
Brinkman, D., 1988, Size-independent criteria for estimating relative age in Lucas, S.G., 2005, Permian tetrapod faunachrons: New Mexico Museum of
Ophiacodon and Dimetrodon (Reptilia, Pelycosauria) from the Admiral Natural History and Science, Bulletin 30, p. 197-201.
and lower Belle Plains formations of west-central Texas: Journal of Lucas, S.G., 2006, Global Permian tetrapod biostratigraphy and
Vertebrate Paleontology, v. 8, p. 172-180. biochronology; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds.,
Eberth, D.A. and Brinkman, D., 1983, Ruthiromia elcobriensis, a new Non-marine Permian biostratigraphy and biochronology: Geological
pelycosaur from El Cobre Canyon, New Mexico: Breviora, no. 474, 26 Society, London, Special Publication 265, p. 65-93.
p. Lucas, S.G. and Krainer, K., 2005, Stratigraphy and correlation of the
Evans, D.C. and Reisz, R.R., 2006, An Early Permian varanodontine Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New
varanopid (Synapsida: Eupelycosauria) from the Richards Spur locality, Mexico Geological Society, Guidebook 56, p. 145-159.
Oklahoma: Journal of Vertebrate Paleontology, v. 26, p. 957-966. Lucas S.G., Harris, S.K., Spielmann, J.A., Berman, D.S and Henrici, A.C.,
Harris, S.K., Lucas, S.G. and Spielmann, J.A., 2010, The postcranial skel 2005a, Vertebrate biostratigraphy and biochronology of the Pennsylva
eton of Ophiacodon navajovicus (Eupelycosauria: Ophiacodontidae) nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico:
from the Upper Pennsylvanian of Cañon del Cobre, New Mexico: New New Mexico Museum of Natural History and Science, Bulletin 31, p.
Mexico Museum of Natural History and Science, Bulletin 49. 128-139.
Kempter, K., Zeigler, K., Koning, D. and Lucas, S., 2007, Preliminary Lucas, S.G., Harris, S.K., Spielmann, J.A., Berman, D.S, Henrici, A.C.,
geologic map of the Canjilon SE quadrangle, Rio Arriba County, New Heckert, A.B., Zeigler, K.E. and Rinehart, L.F., 2005b, Early Permian
Mexico: New Mexico Bureau of Geology and Mineral Resources, Open vertebrate biostratigraphy at Arroyo del Agua, Rio Arriba County, New
file Geological Map 150, 1:24,000. Mexico: New Mexico Museum of Natural History and Science, Bulletin
158
31, p. 163-169. Reisz, R.R., 1986, Pelycosauria: Encyclopedia of Paleoherpetology, Part
Lucas, S.G., Spielmann, J.A. and Krainer, K., 2010, Summary of geology of 17A, 102 p.
Cañon del Cobre, Rio Arriba County, New Mexico: New Mexico Museum Reisz, R.R., 2004, A reevaluation of the enigmatic Permian synapsid Watongia
of Natural History and Science, Bulletin 49, this volume. and its stratigraphic significance: Canadian Journal of Earth Science, v.
Reisz, R.R., 1980, The Pelycosauria: A review of phlogenetic relationships; 41, p. 377-386.
in Panchen, A.L., ed., The terrestrial environment and the origin of land Romer, A.S. and Price, L.I., 1940, Review of the Pelycosauria: Geological
vertebrates: Academic Press, London, p. 553-592. Society of America, Special Paper 28, p. 1-538.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

159
REDESCRIPTION OF THE CRANIALANATOMY OF SPHENACODON FEROX
MARSH (EUPELYCOSAURIA: SPHENACODONTIDAE) FROM THE LATE
PENNSYLVANIAN-EARLY PERMIAN OF NEW MEXICO

JUSTINA. SPIELMANN1, LARRYF. RINEHART1, SPENCERG. LUCAS1,


DAVID S BERMAN2,AMYC. HENRICI2 AND SUSANK. HARRIS1
1 New Mexico Museum of Natural History and Science, 1801 Mountain Rd. NW, Albuquerque, NM 87104-1375;

2 Section of Vertebrate Paleontology, Carnegie Museum of Natural History, 4400 Forbes Ave., Pittsburgh, PA 15213

Abstract—Sphenacodon is a pelycosaurian-grade synapsid, best known from the Lower Permian of Rio Arriba
County, northern New Mexico. Of the two species (S. ferox and S. ferocior), S. ferox is known from comparatively
little skull material, and the skulls historically assigned to the taxon are heavily reconstructed and composed offew
actual cranial elements. Here, we report on a new, nearly complete skull of S. ferox that demonstrates numerous
cranial differences between the two species of Sphenacodon. S. ferox has a temporal range from the Late Pennsyl
vanian (late Virigilian: Coyotean lvf) through the Early Permian (late Wolfcampian: Seymouran lvf), whereas S.
ferocior is restricted to the Coyotean lvf, but does span the Pennsylvanian-Permian boundary.

INTRODUCTION
Sphenacodon ferox is a pelycosaurian-grade synapsid
(Eupelycosauria: Sphenacodontidae: Sphenacodontinae) best known from
the Lower Permian of Rio Arriba County, northern New Mexico (Fig. 1).
S. ferox is known from numerous, incomplete skulls, most of which have
been extensively reconstructed. Here, we review all known S. ferox skulls
(Figs. 2-9, 13) and contrast their cranial morphology with that of S.
ferocior, whose skull morphology is well documented (Eberth, 1985),
and describe the most complete skull of S. ferox known, NMMNH P
55367 (Figs. 13-14), recently collected from the eastern wall of Cañon
del Cobre (El Cobre Canyon), a classic Permo-Carboniferous collecting
area (Lucas et al., this volume). This results in a revision of the diagnoses
of the two species of Sphenacodon.
Institutional abbreviations: AMNH, American Museum of
Natural History, New York; CM, Carnegie Museum of Natural History,
Pittsburgh; FMNH, Field Museum of Natural History, Chicago;
NMMNH, New Mexico Museum of Natural History and Science, Albu FIGURE 1. Pennsylvanian-Permian fossil localities in northern New Mexico,
querque; UCMP, University of California Museum of Paleontology, Rio Arriba County, that yield Sphenacodon fossils (Arroyo del Agua area
Berkeley; UMMP, University of Michigan Museum of Paleontology, and Cañon del Cobre).
Ann Arbor; YPM, Yale Peabody Museum, New Haven.
goid. Unfortunately, the reconstruction of this skull followed the com
mon practice in the early 20th century of trying to blend the recon
SKULLS OF SPHENACODON FEROX
structed elements with the actual skeletal materialso that they are indis
Several incomplete skulls of Sphenacodon ferox have been de tinguishable from each other. We have summarized what is clearly visible
scribed since the taxon was erected by Marsh (1878), and were the basis on the specimen, though the only way to determine the full extent of the
of several reconstructions. Here, we first review the most significant of
reconstruction would be to reprepare the specimen. UCMP 34226 can
these skulls listing the elements used in the reconstruction and discuss be referred to Sphenacodon ferox on the basis of its nearly linear ventral
the basis of their referral to S. ferox. This is followed by description of
maxillary margin, overall small size and reduced maxillary step (see be
the much more complete skulls NMMNH P-58231 and P-55367. low).
YPM 806 (Holotype) AMNH 4779
The holotype of Sphenacodon ferox is a left mandible and associ AMNH 4779 is an extensively reconstructed skull in the region
ated skull fragments (Fig.2). These additional cranial fragments (a right
anterior to the temporal fenestra (Fig. 4). The elements used in the
premaxilla, left and right maxilla and a right mandible fragment) clearly reconstruction include fragments of the left maxilla and jugal, and the
demonstrate features that identify the specimen as S. ferox, most nota right nasal, frontal, maxilla, lacrimal and jugal. Some of the inferred su
bly the linear ventral maxillary margin (see below). tures have been incorrectly traced on the reconstructed portions of the
UCMP 34226 skull (e.g., the frontal-nasal suture), and there is no indication of the
anterior portion of the maxilla-lacrimal suture. Based on examination of
UCMP 34226 is an extensively reconstructed skull (Fig. 3) with the curvature of the ventral maxillary margin, AMNH 4779 may be a
an associated mandible (not illustrated). The skull elements used in the composite specimen composed of bones of both Sphenacodon ferocior
reconstruction are the right premaxilla, maxilla, jugal and possible ptery and S. ferox. The left maxilla is linear (Fig. 4C) and thus conforms to S.
160

FIGURE 2. Sphenacodon ferox, YPM 806, holotype of S. ferox. A, right maxilla in lateral view. B, right mandible fragment in lateral view. C, left
premaxilla in lateral view. D, left maxilla in lateral view. E, left mandible in lateral view.
161

FIGURE 3. Sphenacodon ferox, UCMP 34226, reconstructed skull in lateral view.

ferox, whereas the right maxilla is slightly convex ventrally (Fig. 4A) and more rigorous differentiation between the two species of Sphenacodon.
thus closer to the definition of S. ferocior, though given its relatively
LOCALITY 5379:AGE, LITHOLOGYAND TAPHONOMY
small size it likely pertains to S. ferox.
NMMNH P-55367 was collected from NMMNH locality 5379,
FMNH 35 a site originally found by David Eberth and described by Berman (1993),
Originally described by Williston (1918), this skull is the basis for who reviewed the vertebrate assemblages of Cañon del Cobre and deter
the most influential skull reconstruction of Sphenacodon ferox, as it was mined that the assemblage containing locality 5379 is Wolfcampian in
considered the best known skull of the species at the time (Figs. 5B-C, age. The NMMNH locality 5379 site is stratigraphically high in the
6). Unfortunately, much of the skull has been reconstructed on the basis Lower Permian Arroyo del Agua Formation (Lucas and Krainer, 2005),
of fragments of the maxillae, nasals, prefrontals, frontals, postfrontals, ~50 m below the base of the overlying Upper Triassic Chinle Group.
right lacrimal, jugals and right squamosal (Fig. 7). The nearly linear ven Topographically, the quarry is located on the eastern wall of Cañon del
tral maxillary margin clearly indicates assignment to Sphenacodon ferox. Cobre approximately 120 m (400 feet) above the canyon floor (Fig. 10).
The bonebed has yielded both cranial and postcranial elements of
FMNH 1218 Sphenacodon ferox, in addition to fossils of the diadectomorph Diadectes
FMNH 1218 consists of associated skull fragments, including and the temnospondyl Platyhystrix (Lucas et al., 2005). NMMNH P
55367 was located approximately 0.6 m above the main bonebed level (in
portions of the right nasal, lacrimal and maxilla, and a nearly complete
Figure 10F the main bonebed is at the feet of the worker, and the skull is
right mandible (Fig. 8). The maxilla preserves portions of the two canini in the highwall at the level of the chisel).
form teeth and eight postcaniniforms. Based on its small size and the
nearly linear ventral maxillary margin, FMNH 1218 is referable to The stratigraphic position of NMMNH locality 5379 places it
within the upper assemblage zone of tetrapod fossils found in Cañon del
Sphenacodon ferox (see below).
Cobre (Lucas et al., 2005). Lucas et al. (2005) assigned this upper assem
NMMNH P-58231 blage zone to the Early Permian Seymouran land-vertebrate faunachron
of Lucas (2005, 2006), based on correlations with the Arroyo del Agua
NMMNH P-58231 consists of three right maxillae, one left man Formation near Arroyo del Agua in Rio Arriba County, New Mexico
dible and two skull fragments (Fig. 9), thus representing at least three (Fig. 1), where Seymouria, the principal index taxon of the Seymouran, is
different skulls. The maxillae exhibit nearly linear ventral margins with a present. Lucas (2006) indicates that the Seymouran lvf is equivalent to
greatly reduced maxillary step, so they are referable to Sphenacodon the late Wolfcampian marine stage.
ferox. Lithology and taphonomy: The lithology and taphonomy of the
NMMNH locality 5379 quarry were first reported by Rinehart et al.
Summary
(2007). The quarry facies comprise poorly sorted (vfL to cL), finely
As demonstrated above, all of the skulls of Sphenacodon ferox, laminated, dark reddish brown (10R 3/4), immature arkosic sandstone
other than NMMNH P-55367 (described below), are very incomplete, with pale yellowish-green (10GY 7/2) banding at ~10 cm intervals (Fig.
generally represented by the maxillae and other fragmentary elements. 10B-D, F). The yellowish green bands are similar in composition to the
Hence, the longstanding diagnosis based solely on maxillae characters reddish brown rock and apparently represent chemically reduced layers.
and overall size. While the maxillae characters can still be used to diag The quarry site lithology is typical of the Cutler Group throughout
nose S. ferox, newly recognized skull characters based on NMMNH P northern New Mexico. In situ rhizoliths are common in the quarry, and
55367 and described below provide additional features that allow for a occasionally disrupt the fossil bones, including NMMNH P-55367 (see
162

FIGURE 4. Sphenacodon ferox, AMNH 4779, heavily reconstructed skull anterior to the temporal fenesta in A, right lateral, B, dorsal and C, left lateral
views. Note incorrect and incomplete tracing of sutures on the reconstructed portions of the skull.
163

FIGURE 5. Reconstructions of the skulls of A, Sphenacodon. ferocior in lateral view, from Romer and Price (1940), and B-C, Sphenacodon ferox in B,
lateral and C, dorsal views, from Williston (1918).
164

FIGURE 6. Reconstructions of the skeleton of A-B, Sphenacodon ferox in lateral view, A, from Williston (1918) and B, from Romer and Price (1940),
and C, Sphenacodon ferocior, in lateral view from Romer and Price (1940).

below for details). traction on the bottom and are more difficult to transport. They require
Field jackets were marked with a north-south arrow during collec higher flow rates than Group I, and Group III bones lag and are very
tion, and this index was used during preparation to ascertain the direc difficult to transport. The proportions of the Voorhies groups relative to
tional orientation of the bones. A small sample of elongate (length greater the proportions of each category of bone in a complete skeleton thus
than or equal to four times width) bones was found to trend N40°E with indicates the extent of hydraulic sorting (Fig. 12). In unsorted deposits,
reasonably close alignment (± 35°) (Fig. 11). Such alignment probably the bones are found in the same proportions as in a complete skeleton. If
indicates gentle to moderate flow. Additionally, the fossil bones lie flat on the deposit is impoverished in Group I elements, winnowing is indi
the bedding planes, which indicates a low energy depositional environ cated, and if both Group I and Group II elements are impoverished
ment (Voorhies, 1969). relative to Group III, a lag deposit is indicated (Shipman, 1981; Lyman,
Hydraulic sorting of the bones is evident. Using the Voorhies 1994). Our ternary plot (Fig. 12) shows Group I to be severely depleted
group definitions of Behrensmeyer (1975) (Appendix 1), which are based (representing 20%) and Group II to be considerably impoverished (rep
on mammal bones, we categorized and plotted all the prepared bones resenting 33%), whereas Group III is greatly enriched (47%) over the
from the quarry (Fig. 12). The three Voorhies groups are based on diffi expected proportion. Thus, the ternary plot indicates a heavily win
culty of transport by flowing water. Group I comprises bones that are nowed or lag deposit.
easily transported by floating or saltation, and Group II bones have In summary, the microstratigraphy of the quarry, the alignment of
165

FIGURE 7. Sphenacodon ferox, NMMNH C-4771, cast of FMNH 35, reconstructed skull in A, left lateral and B, right lateral views. Labeled elements
indicate actual skeletal material. Abbreviations: d, dentary; f, frontal; j, jugal; max, maxilla; na, nasal; pf, postfrontal; po, postorbital; prf, prefrontal;
sq, squamosal.
166

FIGURE 8. Sphenacodon ferox, FMNH 1218, portions of the right nasal, lacrimal, maxilla and nearly complete right mandible in A, medial and B, lateral
views.
167

FIGURE 9. Sphenacodon ferox, NMMNH P-58231, A, associated maxillae and mandible fragments in medial and lateral views. B-C, isolated right maxilla
in B, lateral and C, medial views.
168

FIGURE 10. NMMNH locality 5379 where NMMNH P-55367 was collected. A-C, Photographs of the quarry at various points during the 2005
excavation. D, Small jackets collected from the quarry; NMMNH P-55367 is in the jacket at the bottom center of the photograph. E, NMMNH P-55367
as it was exposed in the field. F, Collecting of NMMNH P-55367; note that the primary quarry level is at the feet of the person in the photograph and
NMMNH P-55367 was collected from approximately 0.6 m above this horizon at the position of the chisel.
169
tics is difficult given how similar the two North American species of
Sphenacodon, S. ferox and S. ferocior, are to one another and to the
various species of Dimetrodon and Ctenospondylus. Sphenacodon fos
sils are found predominantly in New Mexico, whereas Dimetrodon fos
sils are found predominantly in Texas and rarely in New Mexico (e.g.,
Berman, 1977; Lucas et al., 2009).
We identify NMMNH P-55367 as Sphenacodon ferox based on
the following features: the specimen was found in a nearly monotaxic
assemblage that included postcrania identifiable as Sphenacodon (Lucas
et al., 2005); overall size of the skull, i.e., the skull metrics are virtually
identical to those of S. ferox and dissimilar from S. ferocior (discussed
below); lesser development of maxillary step; lesser curvature of the
ventral margin of the maxilla; and additional cranial features that are
slightly varied and more primitive than those found in S. ferocior. The
mediolaterally compressed, blade-like neural spines of the dorsal verte
brae differentiate Sphenacodon from Dimetrodon, Ctenorhachis,
Ctenospondylus and all other sphenacodontids, and only this morphol
ogy of dorsal vertebra is present in the locality 5379 bonebed (Lucas et
al., 2005).
The reduced maxillary step and less convex ventral maxillary mar
gin of NMMNH P-55367 identify it as Sphenacodon ferox (Figs. 13
14). These features contrast with a large maxillary notch and convex
ventral maxillary margin in S. ferocior and have been used by Romer and
Price (1940) and Reisz (1986, p. 77-78) to distinguish the two species
from one another.
FIGURE 11. Rose plot of elongate bone orientations from L-5379 showing Based on the closed sutures of the skull we interpret NMMNH
NE-SW alignment. P-55367 as representing an adult individual. Its overall size indicates that
the skull conforms more closely to the size range of Sphenacodon ferox
than to that of S. ferocior, which can be up to 40% larger than S. ferox
(Figs. 13-14). The skull metrics of NMMNH P-55367 compared to
average skull sizes of Sphenacodon ferox and S. ferocior, as reported by
Romer and Price (1940), are summarized in Table 1.

SYSTEMATIC PALEONTOLOGY

Suborder EUPELYCOSAURIA Kemp, 1982


Family SPHENACODONTIDAE Marsh, 1878
Subfamily SPHENACODONTINAE Romer and Price, 1940
Genus Sphenacodon Marsh, 1878
Type species: Sphenacodon ferox Marsh, 1878.
Included species: The type species and Sphenacodon ferocior
Romer and Price, 1940.
Diagnosis: Differs from other sphenacodontines by possessing
tall, blade-like neural spines that are shorter than the neural spines of
Dimetrodon, Ctenospondylus and Ctenorhachis.
Distribution: Upper Pennsylvanian-Lower Permian (Coyotean
and Seymouran lvfs) of New Mexico and Utah.
Discussion: “Oxydon” brittanicus von Huene (1908) is a
sphenacodontine known only from the holotype, which consists of a
part and counterpart of a maxilla, from the Kenilworth Sandstone of
England, originally named by von Huene (1908). This taxon was as
signed by Paton (1974) to Sphenacodon based on its similar size, pro
FIGURE 12. A ternary plot of the relative abundances (%) of the Voorhies portions and relative size of the caniniform tooth, who retained the
groups at L-5379. The uppermost data point represents the expected
specific name on the basis of its geographic isolation from all other
proportions if the bones were unsorted. The lower point represents the
occurrences of Sphenacodon. Reisz (1986, p. 78) followed this assign
measured proportions in the quarry. A heavily winnowed or lag deposit is
ment, but noted that it may prove to be invalid. Eberth (1985) pointed
indicated. out that based on the criteria provided by Paton (1974), “Oxydon”
brittanicus could also be assigned to Dimetrodon booneorum and is best
the bones and the preponderance of Voorhies Group III elements indicate
designated as Sphenacodontine incertae sedis until additional material
that the bonebed was probably accumulated as a lag deposit on a flood
suggests otherwise. We follow Eberth (1985) and retain only two species
plain that was subject to periodic sheetfloods of low to medium velocity of Sphenacodon, S. ferox and S. ferocior.
and lacking turbulent flow. An argument could be made for synonymizing the two species of
NMMNH P-55367: IDENTIFICATION Sphenacodon, given their morphologically identical postcrania and the
few, subtle cranial differences, which some workers (e.g., Eberth, 1985)
Identifying sphenacodontines based solely on skull characteris have interpreted as merely reflecting size and/or ontogeny. However,
170

FIGURE 13. Sphenacodon ferox, NMMNH P-55367, nearly complete skull in A, left lateral and B, right ventrolateral views.
171

FIGURE 14. Sphenacodon ferox, NMMNH P-55367, line drawing of nearly complete skull in A, left lateral and B, right ventrolateral views. Skeletal
elements labeled. Stippled shading is exfoliated bone, black shading is openings in the skull, light gray shading is reconstruction and dark gray shading is
adhered matrix. Abbreviations: bs, basisphenoid; ec, ectopterygoid; in, internal narial opening; f, frontal; j, jugal; la, lacrimal; ltf, lateral temporal
fenestra; max, maxilla; na, nasal; p, parietal; pal, palatine; pf, postfrontal; pmx, premaxilla; po, postorbital; pp, postparietal; prf, prefrontal; ps,
parasphenoid; pt, pterygoid; q, quadrate; qj, quadratojugal; s, stapes; sm, septomaxilla; sq, squamosal; srf, sclerotic ring fragments; so, supraoccipital; st?,
supratemporal; t?, tabular?; v, vomer.
172
TABLE 1. Comparative skull metrics of S. ferox, S. ferocior (Romer and
Price, 1940) and P-55367. All measurements in mm. Sphenacodon ferocior Romer and Price, 1940
1940 Sphenacodon ferocior: Romer and Price, p. 327-329.
1985 Sphenacodon ferocior: Eberth, p. 1-40, figs. 1-39.
1986 Sphenacodon ferocior: Reisz, p. 78, figs. 21i, 23j, 30j, 38.
Holotype: MCZ 1489, skull and anterior vertebrae.
Type locality and horizon: San Diego Canyon, Sandoval County,
New Mexico, precise locality unknown. Abo Formation, Lower Per
mian.
Principal referred specimens: YPM 818, partial skull and nearly
complete presacral vertebral column, pectoral and pelvic girdles, hu
merus and femur from “Rito Puerco,” probably a misreading of Rio
NMMNH P-55367, being the most complete S. ferox skull known, Puerco, by David Baldwin, late 19th century, in the Arroyo del Agua area,
contains new information and its description below identifies numerous Rio Arriba County, NM. Various UCMP and MCZ specimens dis
characters that distinguish S. ferox from S. ferocior. The NMMNH speci cussed and illustrated by Eberth (1985) including: UCMP 34184, left
men is especially important because it is an adult skull, and thus elimi postfrontal; UCMP 34192, partial lacrimal; UCMP 34194, incomplete
nates the problem of comparing different ontogenetic stages of growth parietal; UCMP34196, partial skull; UCMP34201, left splenial; UCMP
by allowing for comparisons between fully adult individuals of the two 34202, left articular; UCMP 34218, prearticular, occiput and braincase;
species. UCMP 34219, left quadrate and quadratojugal; UCMP 83047, right
premaxilla; UCMP 83185, right prefrontal; UCMP 81393, left frontal;
Sphenacodon ferox Marsh, 1878 UCMP 83060, right maxilla; UCMP 83203, right nasal; UCMP 83208,
1878 Sphenacodon ferox Marsh, p. 410 partial skull; UCMP 83251, supraoccipital; UCMP 83285, left palatine;
1907 Elcabrosaurus baldwini Case, p. 28 UCMP 83378, right surangular; UCMP 83397, right angular; UCMP
1913 Scoliomus puercensis Williston and Case, p. 60, fig. 37 83459, right coronoid; UCMP 83478, right dentary; UCMP 123288,
1940 Sphenacodon ferox: Romer and Price, p. 326-327 partial stapes; UCMP 125583, partial stapes.
1986 Sphenacodon ferox: Reisz, p. 77, figs. 4, 18c, 25g, 26e, 27g Revised diagnosis: Aspecies of Sphenacodon distinguished from
S. ferox by the following features: overall large size, up to 40% larger
Holotype: YPM 806, left mandible (Fig. 2E). Clearly associated than the largest individuals of S. ferox; possessing three premaxillary
elements of YPM 806 include the left maxilla, a left mandibular fragment, teeth; one or two precaniniform maxillary teeth in adult specimens; con
right premaxilla and maxilla (Fig. 2A-D). Given their association these vex ventral maxillary margin; prominent maxillary step; dorsal maxillary
elements should be considered elements of the holotype, due to all the process that results in a more rounded anterior end of the lacrimal and a
skull elements being similarly proportioned and likely derived from a longer maxilla-nasal suture; orbital process of the frontal thickened; tem
single skull. See above under Skulls of Sphenacodon ferox for addi
poral fenestra with a rounded anteroventral margin; a squamosal-jugal
tional comments on this specimen. suture that extends posteriorly beyond the posterior temporal fenestra;
Type locality and horizon: Baldwin bonebed (Fig. 15), near anterior parietal has shallow, elliptical sutural surface, no embayments;
Arroyo del Agua, Rio Arriba County, New Mexico, from the El Cobre
and tooth field of the pterygoid extends to its lateral margin. See Figure 4
Canyon Formation, Cutler Group (Berman, 1993). for graphic illustration of these characteristics.
Principal referred material: Skulls of Sphenacodon ferox in
clude:UCMP 34226, reconstructed skull based on partial right premax NMMNH P-55367: DESCRIPTION AND COMPARISON
illa, and left maxilla and left jugal (Fig. 3); NMMNH P-55367, nearly
NMMNH P-55367 is a nearly complete skull of Sphenacodon
complete skull, crushed mediolaterally; NMMNH P-58231, maxillae,
mandibles and skull fragments (Fig. 9); AMNH 4779, mostly recon ferox that has been obliquely crushed mediolaterally to the extent that
structed skull based on right nasal, frontal, maxilla, lacrimal, jugal, pos the left lateral side and the skull table are visible in the same plane (Figs.
torbital, postfrontal and left maxilla and jugal, all elements in various 13-14). The following elements are not preserved in NMMNHP-55367:
states of completeness (Fig. 4); FMNH 35, mostly reconstructed skull tabulars, epipterygoid, basioccipital, exoccipital, supraoccipital,
using various skull fragments incorporated in the reconstruction (Fig. 7); opisthotic, sphenethmoid. Portions of left preorbital region of the skull
FMNH 1218, incomplete skull and mandible (Fig. 8). and cheek region are patched with epoxy putty (for strength and stabili
Revised diagnosis: Aspecies of Sphenacodon distinguished from zation). In addition, left caniniform 2 has been reconstructed based on an
S. ferocior by the following characters: overall smaller size, as much as impression in the surrounding matrix. Some of the cranial elements ante
rior to the antorbital reconstruction are partially exfoliated, further com
40% smaller than the largest individuals of S. ferocior; four premaxillary
teeth; three precaniniform maxillary teeth in adult specimens; nearly plicating suture identification.
linear ventral maxillary margin; greatly reduced maxillary step; dorsal Oblique crushing has resulted in most of the right side of the skull,
lamina of maxilla less developed, resulting in a wedge-shaped lacrimal the palate and braincase being visible in a single plane. The right lacrimal
is missing part of its central portion. The posteroventral margin of the
and a shorter sutural contact between the maxilla and nasal; orbital pro
maxilla is missing. The posterior part of the skull and braincase are
cess of the frontal not as greatly thickened; anteroventral margin of the
temporal fenestra forms a right angle; squamosal-jugal suture does not missing, including most of the squamosal and basioccipital.
Here, we describe the skull of Sphenacodon ferox based on
extend posteriorly beyond the posterior extent of the temporal fenestra; NMMNH P-55367, the most complete skull known for this taxon, and
parietals with circular, concave embayments on anterior margin; and
compare it to descriptions of S. ferocior (Eberth, 1985) and of other
denticle field on pterygoid flange does not extend to its lateral margin.
Discussion: A summary of the diagnostic features of S. ferox is sphenacodontines.
provided below under the heading Diagnostic cranial features of Premaxilla
Sphenacodon ferox.
Description: Description of NMMNH P-55367 is provided be Both premaxillae are present, complete and articulated, with the
low; all other type and referred material is discussed above under the right slightly offset dorsally from the left, due to oblique crushing (Figs.
heading Skulls of Sphenacodon ferox. 13-14). The dorsal process is posterodorsally curved and ends in a point
173

FIGURE 15. Photograph of the Baldwin bonebed taken by E.C. Case. White arrow highlights the level of the quarry.

medial to the anterior end of the nasal. Due to the disarticulation of the septomaxilla, which is only visible in lateral view. The septomaxilla
nasals (see below) the exact suture between these elements and the pre appears as a crescent-shaped element in the posterior half of the external
maxillae is not clear. Four tooth sockets are present (Fig. 16A-B). In naris. It extends for the full height of the external naris and contacts
ventral view the sinusoidal suture with the anterior margin of the maxilla anteriorly the premaxilla. The medial shelf appears to be high on the
is angled posteromedially, so that in lateral view the fourth (posteriormost) septomaxilla and more similar to the left septomaxilla illustrated by
premaxillary tooth is positioned directly below the premaxilla-maxilla Eberth (1985, fig. 5C) in MCZ 1489, than the right illustrated by Eberth
suture. Thus, the premaxillary process of the maxilla overlaps laterally (1985, fig. 5A) in UCMP34196. The anterior foramen is located directly
the maxillary process of the premaxilla. The anterior two premaxillary under the medial shelf and is directed ventromedially. A small, posteri
teeth are conical and not recurved, whereas the posterior two premaxil orly-directed, triangular notch is present at the posteroventral margin of
lary teeth are much smaller, approximately half the size of the anterior the septomaxilla that leads to the septomaxillary foramen. The
teeth, and are recurved. Due to crushing, the suture between the premax septomaxilla of NMMNHP-55367 differs from that of S. ferocior (Eberth,
illae and vomers is not completely clear. Crushing also prevents a clear 1985, fig. 5) in the following ways: has an overall crescent-shaped out
view of the septomaxilla-maxilla suture and the posterior premaxillary line that is not taller than wide (Eberth, 1985, fig. 5A) or polygonal
and prepalatal foramina, which are well described in Sphenacodon ferocior (Eberth, 1985, fig. 5C) and is more recessed into the naris with a promi
by Eberth (1985, p. 9, fig. 4). nent posteroventral rim.
Septomaxilla Nasal
The left septomaxilla is complete, whereas the right is incomplete Both nasals are present (Figs. 13A, 14A). The nasal contacts the
and disarticulated (Figs. 13A, 14A). The description is based on the left lacrimal ventrolaterally with a finely digitate suture anteriorly, whereas
174

FIGURE 16. Closeup of Sphenacodon ferox skull, NMMNH P-55367, highlighting diagnostic features of S. ferox. A-B, premaxillae and anterior maxillae
in ventrolateral views; four premaxillary teeth indicated by arrows, premaxilla-maxilla suture and fourth premaxillary tooth socket outlined. C-D,
maxillae in ventrolateral view; three maxillary precaniniform teeth indicated by arrows, premaxilla-maxilla suture and first caniniform tooth socket
outlined. Abbreviations: mx pc1-pc3, maxillary precaniniform teeth one through three; mx 1-2, maxillary caniniform teeth one and two; pmx t1-t4,
premaxillary teeth one through four.
175
posteriorly the suture is obscured by exfoliated bone. The turbinal ridge postfrontal. Also, it forms the anterodorsal margin of the lateral temporal
is clearly visible on the ventral surface of the nasal, also noted in fenestra. The ventral process of the postorbital overlaps the postorbital
Sphenacodon ferocior by Eberth (1985, p. 12-13, fig. 6A). process of the jugal and contacts the parietal with a distinct suture. The
thin posterior process borders the entire lateral length of the wing-like
Frontal
posterior extension of the parietals. There are no differences in this
Both frontals are present (Figs. 13A, 14A). The anterior process element between the two species of Sphenacodon.
of the frontal tapers anteriorly to a triangular point that fits in the poste
rior margin of the nasal and is only clearly visible on the right side, due to Maxilla
reconstruction. The frontal, like the nasal, possesses a turbinal ridge. The Both maxillae are present, though the left is reconstructed par
orbital process of the frontal does not appear to be as thickened in tially and the right is incomplete. There are three precaniniform, two
Sphenacodon ferox as in S. ferocior (Eberth, 1985, fig. 7A), yet the caniniform and 12 postcaniniform tooth positions, for a total of 17 (Figs.
orbital processes of both the prefrontal and postfrontal are thickened in 13-14, 16C-D). The left maxilla preserves teeth in precaniniforms 2 and
S. ferox.
3, caniniform 2 and postcaniniforms 1, 2, 9 and 10. The right maxilla
Parietal preserves teeth in precaniniform 1-3, caniniform 2 and postcaniniform 1,
2 and 10-12. Three precaniniform teeth in an adult Sphenacodon ferox
Both parietals are present and essentially complete, although the distinguishes it from S. ferocior, which generally has one, and rarely two,
right parietal is missing the supratemporal sutural surface (Figs. 13A, precaniniform teeth in adults, though juveniles can have up to four.
14A). The midline suture is indistinct and the pineal foramen was dam Overall, the ventral margin of the maxilla is only slightly convex, com
aged by a fossilized root (rhizolith). On the anterior margin of the pari pared to being strongly convex in S. ferocior; this feature has long been
etal the embayment for the postfrontal is deep and semicircular (Fig. used to distinguish between S. ferox and S. ferocior (Romer and Price,
17C-D), whereas in Sphenacodon ferocior (Eberth, 1985, fig. 8A) it is 1940; Reisz, 1986). The dorsal maxillary process is present on the ante
represented by a shallow, concave sutural scar on the dorsal surface of rior maxilla, resulting in maxillae and lacrimals that are shaped differently
the anterior margin (Fig. 17E-F). Also, no features mark the tabular than in S. ferocior.
sutural surface, which is distinguished in S. ferocior by a shallow ellipti
cal embayment, similar to the embayment for the postfrontal. Lacrimal

Postparietal As noted above, both lacrimals are present but incomplete (Figs.
13-14), with the right partially disarticulated and missing its central
An unpaired postparietal is present, an element not known in portion, whereas the left is partially exfoliated, further confirming Eberth’s
Sphenacodon ferocior (Eberth, 1985, p. 15) (Figs. 13A, 14A). As hy (1985, p. 18) observation that they are thin elements. The ventral margin
pothesized by Eberth (1985), the postparietal has a low, squat, triangu of the anterior lacrimal is bounded by a strongly digitate suture with the
lar outline, with the apex directed ventrally. The postparietal of NMMNH maxilla and dorsally by a finer digitate suture with the nasal. The anterior
P-55367 is bowed anteriorly and is bordered dorsally and laterally by the portion of the lacrimal is wedge-shaped, rather than rounded, as in S.
parietal, which overlap it marginally. A faint, radial, surface texture is ferocior. This difference may be the result of a less well-developed dorsal
present on the margins of the postparietal. Although obscured by crush maxillary lamina that borders the lacrimal anteriorly in S. ferocior.
ing, the postparietal was likely bordered ventrally by the supraoccipital
Jugal
and ventrolaterally by the tabulars, as in other sphenacodontines (e.g.,
Dimetrodon limbatus: Reisz, 1986, fig. 15). Both jugals are present, with the left being complete, whereas the
Prefrontal right is missing its anterior margin (Figs. 13-14). Eberth (1985, p. 18, fig.
15) noted two distinct morphologies of the posterior process of the jugal
The left prefrontal is nearly complete (except for some minor in Sphenacodon ferocior: one with only a slight downward angulation of
areas of reconstruction) and in articulation with neighboring elements, the posterior process and the other with a strong downward angulation.
whereas the right is incomplete, missing the margin bordering the lacrimal The posterior process in NMMNHP-55367 is significantly downturned
(Figs. 13-14). The thickening of the orbital margin is prominent, espe and is much shorter than in S. ferocior. However, in NMMNH P-55367
cially compared to the unthickened orbital process of the frontal (see the angle between the body of the jugal and the posterior process ap
above). The depression immediately anterior to the orbital deepens to pears to be more acute.
ward the well-developed orbital ridge, more so than illustrated in
Squamosal
Sphenacodon ferocior (Eberth, 1985, fig. 9A). Also, the thickening of the
prefrontal continues anteriorly and dorsal to the orbital process, a fea Both squamosals are very incomplete (Figs. 13-14). The left squa
ture not seen in S. ferocior. The area of the suture with the lacrimal has
mosal is represented by a small triangular fragment sutured to the poste
been lost, but the prefrontal-frontal suture is preserved. rior process of the jugal, which together form the anterior portion of the
Postfrontal zygomatic arch. A small portion of the right squamosal is preserved at
the posterior margin of the temporal fenestra. Based on these incomplete
Both postfrontals are present, complete and in place (Figs. 13 elements, there is no obvious differences between the two species of
14). The postfrontal forms the posterodorsal margin of the orbit, where, Sphenacodon.
like the prefrontal, it is greatly thickened. Overall, the element is qua
drangular in dorsal view. There are no differences in this element between Supratemporal
the two species of Sphenacodon. An incomplete element is identified tentatively as the left su
Postorbital pratemporal (Figs. 13A, 14A). It is articulated with the lateral process of
the parietal and is curved posteroventrally. In sphenacodontids the su
Both postorbitals are present, though the left is more complete pratemporal is a splint-like bone whose anterior portion is sutured to a
and well articulated (Figs. 13-14). The postorbital forms the posterior shallow groove in the posterolateral wing of the postparietal, whereas
margin of the orbit, but it is not thickened as are the prefrontal and the medial margin of the posterior portion contacts the lateral margin of
the tabular.
176

FIGURE 17. Closeup of Sphenacodon ferox skull, NMMNH P-55367, highlighting diagnostic features of S. ferox. A-B, dorsal orbital margin in lateral view;
note the unthickened orbital process of the frontal, sutures between the prefrontal-frontal, frontal-postfrontal and postfrontal-postorbital are outlined.
C-D, parietals in dorsal view; parietals are outlined and the deep, circular embayments are indicated by arrows. E-F, comparison of the parietals of E,
Sphenacodon ferox (based on NMMNH P-55367) and F, S. ferocior (based on UCMP 34194 and MCZ 1489, from Eberth, 1985, fig. 8A). Abbreviations:
emb, parietal embayments; op f, orbital process of the frontal; pf, postfrontal; po, postorbital; prf, prefrontal.
177
Vomer
footplate, which is articulated with the posterior braincase (Figs. 13B,
The paired vomers are present and visible in ventral view. They 14B, 18C-D).
form a prominent ridge at the midline, which borders the medial margin of
the internal narial opening (Figs. 13B, 14B). Posteriorly, the vomers may Sclerotic Ring
be unique in wedging deeply between the anterior ends of the palatines.
A sclerotic ring is present in the left orbit of P-55367 (Fig. 19A).
The posterior suture with the palatine is V-shaped with the paired vomers The ring is apparently folded and flattened (forming a short, flattened,
wedging between the anterior ends of the palatine, whereas the anterior hollow cylinder exposed in lateral view). Fourteen ossicles are visible,
suture with the premaxilla is obscured. Posteriorly, the vomers expand represented as either complete or partial elements, or as impressions in
slightly before contacting the anteromedial corners of the palatine, then
the matrix. Space exists for four or five additional ossicles, which may be
narrow abruptly to short processes that overlap the distal ends of the present, but obscured by other elements of the skull. Walls (1942) desig
pterygoids. nated sclerotic ossicles as plus (+) or minus (-) for those that overlap, or
Palatine are overlapped by, both of their neighboring ossicles, respectively and (i)
for those that show an imbricated pattern in which they both overlap and
The palatine is represented by only the incompletely preserved are overlapped by each of their respective neighboring ossicles. The
left. It is subtriangular, and the tooth field abuts the vomer-palatine sclerotic ossicles in NMMNH P-55367 are imbricated (i) with the ex
suture (Figs. 13B, 14B). The palatine contribution to the posterior mar ception of a single anteriormost (+) ossicle in the nasal position of the
gin of the internal naris is not visible. A narrow rectangular denticle field ring (Fig. 19B). The outer, orbital edge of the ring is paper-thin and
covers the medial half of the palatine and is continuous with that of the translucent, whereas the inner, corneal boundary is much thicker and
pterygoid. opaque.
The visible portion of the ring is flattened into a nearly straight
Ectopterygoid structure that is 47 mm long. If more ossicles are present in the covered
Both ectopterygoids are present and have a narrowly portion of the ring, it could have a circumference of up to 70 mm. The
subrectangular outline with a slightly convex ventral surface. The radial distance from the corneal aperture to the orbital edge of the ring is
retroangular posterior portion of the ectopterygoid is incised into the 11 mm. Individual imbricated ossicles are pear-shaped. The wider, cor
anterolateral area of the transverse flange of the pterygoid. It appears to neal portion forms the corneal aperture and overlaps half of each of the
be greatly reduced compared to the ectopterygoid in other sphenacodontids adjacent ossicles, whereas the narrower, orbital portion abuts its adja
(Figs. 13B, 14B). Unfortunately, examples of this element are unknown cent ossicles (Fig. 19B). In the best-exposed ossicle, the orbital margin is
for S. ferocior (Eberth, 1985, p. 22). 2.9 mm wide and the corneal margin is 6.5 mm wide.

Pterygoid DIAGNOSTIC CRANIAL


FEATURES OF SPHENACODON FEROX
Both pterygoids are presumably complete, but only their poste
rior portions are visible in ventral view (Figs. 13B, 14B, 18C-D). The left Figure 20 provides graphic illustration of the diagnostic character
and right transverse flanges possess six (two are missing) and seven istics of the two Sphenacodon species. Here, we briefly discuss the
teeth, respectively, which are aligned in a single row along their posterior diagnostic cranial features that distinguish S. ferox from S. ferocior and
margin. The anterior margin of the transverse flange has a triangular other sphenacodontines.
concavity, similar to the condition in S. ferocior (Eberth, 1985, fig. 21B). 1. The overall smaller size of Sphenacodon ferox: The skull of S.
Anterior to the transverse row of teeth, the triangular dental field nar ferox is as much as 40% smaller than that of S. ferocior (Figs. 20-21).
rows anteriorly to an apex and closely borders the medial margin of the This is a well-established characteristic that has been used to distinguish
pterygoid and does not reach its lateral margin (Fig. 18A-B). The den the two species since Romer and Price (1940) originally described and
ticles increase in size toward the center of the field. Eberth (1985, fig. named S. ferocior.
21B) illustrated the tooth field of S. ferocior as more polygonal, extend One possibility to explain the size difference is ontogeny, with
ing to the lateral margins of the pterygoid, with no concentration of larger the smaller S. ferox being the juvenile of the larger S. ferocior. In order to
denticles. test this idea we used a statistical approach to determine if there are two
significantly different size classes representing the two species of
Quadrate Sphenacodon, S. ferox and S. ferocior. Scapulocoracoid data from Romer
Both quadrates are complete, but visible only in ventral view, and Price (1940) were used because they provide the largest sample size
although the right has been damaged by a rhizolith that extended through available in the literature (Table 2). Our first approach was to apply
the pineal foramen and formed a circular hole in its posterolateral margin ANOVA testing using PAST software (Hammer et al., 2001) to deter
mine if two size classes are present, but the tests failed to run because of
(Figs. 13B, 14B, 18C-D). As noted by Eberth (1985, p. 24), the dorsal
process of the quadrate is essentially diamond-shaped and exceedingly inadequate sample size in the S. ferocior data. Likewise, the sample size
thin. was inadequate for assessment of shape through PC analysis. Our next
approach was to use probability plotting. In a probability plot of scapu
Parabasisphenoid lae lengths (Fig. 22) the five data points representing S. ferox appear to
be distributed normally as indicated by the straight line fit on the normal
The parabasisphenoid complex of the braincase is preserved in
probability scale and show a mean value and standard deviation (sigma)
ventral view (Figs. 13B, 14B, 18C-D). In overall morphology this ele of 134 mm and 9.3 mm, respectively. The calculated coefficient of varia
ment appears similar to the parabasisphenoid of Sphenacodon ferocior tion (CV = (sigma/mean value) *100) is a very reasonable 6.9%. CVs in
(Eberth, 1985, p. 26-27, fig. 26) except that the fossa between the basal this range (between 4% and 10%, and most often 5% or 6%) usually
tubers of NMMNH P-55367 is more pronounced and lacks the small, u
indicate that the data are relatively “pure,” that is, they include only
shaped depression just posterior to the cartoid foramina found in S. animals of similar age class and taxonomic assignment (Simpson et al.,
ferocior. 1960).
Stapes Admittedly, the number of data points, which is well below an
optimal sample size, does not inspire confidence in the result. However,
The right stapes is represented in NMMNH P-55367 only by the low scatter about the regression line (R2 = 0.99) indicates that the stan
178

FIGURE 18. Closeup of Sphenacodon ferox skull, NMMNH P-55367, highlighting diagnostic features of S. ferox. A-B, pterygoid in ventral view;
pterygoid-maxilla suture outlined and maximum extent of pterygoid tooth field outlined by dotted line. C-D, posterior skull/braincase in ventral view.
Abbreviations: bs, basisphenoid; ic, internal carotid foramen; mx, maxilla; ps, parasphenoid; pt, pterygoid; q, quadrate; Q. pr, quadrate process of the
pterygoid; qj, quadratojugal; s, stapes.

dard error of the estimate is low, thus increasing confidence (Glantz, data points represent extreme outliers of the S. ferox distribution thus is
2005). The value of the CV, which implies that the data derive from a vanishingly small. We conclude that within the limits of a very small
single taxon and age class, also raises confidence in the result. sample size, we have shown with reasonable confidence that S. ferox and
On the normal probability scale (Fig. 22) the +1, +2, and +3 sigma S. ferocior embody two distinct size classes that represent two species
probability levels occur at 84%, 97.5%, and 99.85%, respectively. If the and are not part of a single ontogenetic trajectory.
two Sphenacodon ferocior data points are translated horizontally to the 2. Possession of four premaxillary teeth: Eberth (1985) noted that
S. ferox distribution (the curve fit line), they intersect the distribution at only three premaxillary teeth occur in Sphenacodon ferocior, based on
approximately the +4 sigma level (off the plot to the right). Thus, there 14 specimens, thus clearly distinguishing it from S. ferox, which has four
is a less than one in 10,000 probability (P<0.0001) that these data points premaxillary teeth (Fig. 16A-B). Although Eberth (1985, p. 34), follow
actually belong to the same distribution. Forcing the regression line to the ing Langston (1952), noted that there are “three premaxillary teeth in all
extreme limits of the error bars does not change this result significantly. known specimens of S. ferox,” the only skull with a complete premaxil
Haldane (1948) asserted that a four sigma difference in the metric charac lary dentition is NMMNH P-55367; all other skulls of S. ferox are
ters of two populations of fossils alone was sufficient to establish them extremely incomplete and/or heavily reconstructed (see above for our
as “almost certainly” different species. The likelihood that both S. ferocior summary of known S. ferox skulls). In general, the various species of
179

FIGURE 19. Sphenacodon ferox, NMMNH P-55367. A, Closeup of sclerotic ring in lateral view. The leftmost ossicle in the photo is the single visible (+)
ossicle. B, Schematic sketch of the imbricated ossicles that comprise the bulk of the sclerotic ring.
180

FIGURE 20. Comparison of Sphenacodon ferox and S. ferocior skulls. Skulls reproduced at the same size for comparison, not to scale, except in inset (1),
reproduced at the same size for comparison. Diagnostic features are numbered: (1) overall size; (2) number of premaxillary teeth; (3) number of
precaniniform maxillary teeth in adult specimens; (4) ventral maxillary margin shape; (5) development of maxillary step; (6) possession/absence of dorsal
maxillary process and shape of anterior lacrimal; (7) anteroventral temporal fenestra shape; (8) position of the squamosal-jugal suture; and (9) extent of
the tooth field on the pterygoid. S. ferocior redrawn from Eberth (1985, fig. 3a) and S. ferox based on NMMNH P-55367.

Dimetrodon possess three premaxillary teeth, with D. grandis having illustrated by Eberth (1985, p. 33) are isolated, so their taxonomic iden
two and D. natalis having five (Romer and Price, 1940, p. 332). Thus, tification is questionable. NMMNH P-55367, unquestionably an adult,
neither S. ferocior nor any species of Dimetrodon possess four premax possesses a linear ventral maxillary margin that can be distinguished
illary teeth. easily from the convex maxillary margin of S. ferocior. Thus, a linear
3. Three precaniniform maxillary teeth in adults: Injuvenile speci ventral maxillary margin remains a useful feature in distinguishing adult
mens of Sphenacodon ferocior, the precaniniform maxillary tooth count skulls of S. ferox from those of S. ferocior.
can range from 1 to 4 teeth. However, adult specimens of S. ferocior 5. Greatly reduced maxillary step: Eberth (1985, p. 33) noted that
possess only one or two (Eberth, 1985, table 1). This is clearly distinct the maxillary step is well developed in all specimens of Sphenacodon
from the three precaniniform maxillary teeth of adult S. ferox (Fig. 16C ferocior except the smallest (based on 7 specimens of S. ferox and 13
D). specimens of S. ferocior). When comparing adult skulls of Sphenacodon,
4. Linear ventral maxillary margin: Eberth (1985, p. 33) noted that S. ferocior always has a larger, more developed maxillary notch than S.
the degree of curvature of the ventral margin of the maxilla in both ferox. Thus, this feature, originally used by Romer and Price (1940),
Sphenacodon ferox and S. ferocior increases during ontogeny. He based remains useful to distinguish the two species of Sphenacodon.
this on seven specimens of S. ferox and 13 specimens of S. ferocior, 6. Dorsal lamina of maxilla less developed, resulting in a wedge
though no specimen numbers for the illustrated material are provided, shaped lacrimal and a shorter sutural contact between the maxilla and
making identification of the specimens used and thus reproducibility of nasal: Comparing the skulls of the two species of Sphenacodon (Fig. 20),
this observation exceedingly difficult. It is also unclear why Eberth (1985) the lacrimal of S. ferocior is bordered anteriorly by the maxillary dorsal
identified maxillae with curved ventral margins as S. ferox, with no addi process, a small triangular extension of the maxilla, just behind the exter
tional characters to support his identification. In addition, the maxillae nal narial opening. This process is not present in S. ferox, resulting in an
181

FIGURE 21. Comparing the size of Sphenacodon ferox and S. ferocior skulls. All skulls are to the same scale. Skulls aligned by the anterior margin of their
orbits (gray bar). From top to bottom: UCMP 34226 (S. ferox); NMMNH P-55367 (S. ferox) (photo reversed horizontally for comparison); AMNH 4779
(S. ferox); and UCMP 34196 (S. ferocior). Black bars approximate length of the S. ferocior skull for comparison. Note that NMMNH P-55367 is similar
in overall size to S. ferox and considerably smaller than S. ferocior.
182
TABLE 2. Sphenacodon scapulocoracoid data from Romer and Price (1940).
Scapulocoracoid measurements are used because they produce the largest
statistical sample. Length refers to distance from “top of anterior end of
glenoid to top of scapula.” WM, Walker Museum, now part of the Field
Museum of Natural History.

FIGURE 23. Temporal distribution of Sphenacodon.

7. Orbital process of the frontal not thickened: Sphenacodon ferox


is the only sphenacodontid that does not possess a thickened orbital
process of the frontal (Fig. 17A-B), including S. ferocior (Eberth, 1985,
p. 13-14) and Dimetrodon. This is especially interesting given that both
the prefrontal and postfrontals that bracket the orbital process are thick
ened in S. ferox.
8. Temporal fenestra with a right angle at the anteroventral corner:
In Sphenacodon ferocior, the anteroventral margin of the temporal fenes
tra is rounded, giving the fenestra a subelliptical shape. In S. ferox, the
anteroventral margin of the temporal fenestra forms a right angle (Fig.
FIGURE 22. Probability plot of scapulae lengths in Sphenacodon ferox and 20), resulting in a reversed D-shaped outline.
S. ferocior. Error bars represent the standard error. Both S. ferocior data 9. The squamosal-jugal suture does not extend posteriorly be
points occur at the ~+4 sigma probability level on the S. ferox line indicating yond the temporal fenestra: Eberth (1985, fig. 3) described and illus
extremely low probability that they belong to the same distribution trated two distinct morphs of Sphenacodon ferocior, based primarily on
(P<0.0001). the shape of the jugal. In both, the squamosal-jugal suture extends be
yond the posterior margin of the temporal fenestra. In contrast, this
anterior lacrimal that is more wedge-shaped and extends farther anteri suture in S. ferox is not as long and does not extend beyond the posterior
orly than in S. ferocior. This can be seen most clearly if one compares the margin of the fenestra (Fig. 20).
respective lengths of the nasal-maxilla suture in the two species. 10. Parietals with circular embayments: The parietals of
Characters 4 through 6 all relate to the maxilla, and thus are likely Sphenacodon ferox have a shape distinct from those of S. ferocior (Eberth,
part of a single larger functional complex. In a phylogenetic analysis all 1985, fig. 8). In S. ferox the parietals have hemispherical embayments on
three of these characters would have identical distributions and our dis their anterior margins (Fig. 17C-E), whereas in S. ferocior only a broadly
cussion of these characteristics is merely a means of detailing the features concave sutural scar for the postfrontals is present (Fig. 17F).
of the maxilla. 11. Dental field on the palatal ramus of the pterygoid does not
183
TABLE 3. Stratigraphic distribution of key Sphenacodon cranial specimens extend to its lateral margin: Eberth (1985, p. 22, fig. 21) described and
discussed in the text. illustrated the dental field on the palatal ramus of the pterygoid of
Sphenacodon ferocior as extending to the element’s lateral margin, whereas
in NMMNH P-55367 it does not (Fig. 18A-B).

STRATIGRAPHIC DISTRIBUTION
Fossils of Sphenacodon are known principally from New Mexico
(Fig. 1). The only non-New Mexican records appear to be from the
Halgaito Formation, Organ Rock Shale and Cedar Mesa Sandstone of the
Arizona-Utah borderland (Monument Valley) (Vaughn, 1962, 1964;
Sumida et al., 1999). In New Mexico, Sphenacodon fossils are known
from the: (1) Red Tanks Member of the Bursum Formation in the Lucero
uplift (Harris et al., 2004); (2) upper part of the El Cobre Canyon
Formation in the Arroyo del Agua area of Rio Arriba County (Berman,
1993; Lucas et al., 2005); (3) El Cobre Canyon and Arroyo del Agua
formations in Cañon del Cobre, Rio Arriba County (Lucas et al., 2005);
(4) Abo Formation in the vicinity of Jemez Springs (San Diego Canyon);
(5) Sangre de Cristo Formation near Villaneuva (Vaughn, 1964; Berman,
1993); (6) Abo Formation near Tularosa, Otero County (Vaughn, 1969);
and (7) the Abo Formation at Gallina Well in Socorro County (Berman,
1993; Spielmann et al., 2009). The stratigraphic distribution of key
Sphenacodon fossils is summarized in Table 3.
Lucas (2006) used the first appearance datum (FAD) of
Sphenacodon to define the beginning of the Coyotean land-vertebrate
faunachron, an interval of time that encompasses the Pennsylvanian
Permian boundary (Fig. 23). Thus, the lowest occurrence (LO) of
Sphenacodon in New Mexico (and in the Halgaito Formation) is in strata
of demonstrable Virgilian (Late Pennsylvanian) age (Red Tanks Member
of the Bursum Formation). Its youngest New Mexican record is in strata
of Seymouran (late Wolfcampian) age – Arroyo del Agua Formation.
Both species – S. ferocior and S. ferox – co-occur in Coyotean age strata,
but only S. ferox is known from the Seymouran, so it is the youngest
known Sphenacodon species.
ACKNOWLEDGMENTS
Pat Holroyd (UCMP), Bill Simpson (FMNH), Walter Joyce and
Dan Brinkman (YPM), Carl Mehling (AMNH) and Chuck Schaff (MCZ)
provided collection access. Sean Modesto and Stuart Sumida provided
helpful reviews that improved the manuscript.
184
REFERENCES

Behrensmeyer, A.K., 1975, The taphonomy and paleoecology of Plio Permian Abo Formation, Socorro County, New Mexico: New Mexico
Pleistocene vertebrate assemblages east of Lake Rudolf, Kenya: Bulletin Geological Society, Guidebook 60, p. 281-284.
of the Museum of Comparative Zoology, v. 146, p. 474-578. Lyman, R.L., 1994, Vertebrate taphonomy, Cambridge Manuals in Archae
Berman, D.S, 1977, A new species of Dimetrodon (Reptilia: Pelycosauria) ology: Cambridge, Cambridge University Press, 524 p.
from the Lower Permian of north-central New Mexico: Journal of Marsh, O.C., 1878, Notice of new fossil reptiles: American Journal of
Paleontology, v. 51, p. 108-115. Science, v. 15, p. 409-411.
Berman, D.S, 1993, Lower Permian vertebrate localities of New Mexico Paton, R.L., 1974, Lower Permian pelycosaurs from the English midlands:
and their assemblages: New Mexico Museum of Natural History and Palaeontology, v. 17, p. 541-552.
Science, Bulletin 2, p. 11-21. Reisz, R.R., 1986, Pelycosauria: Encyclopedia of Paleoherpetology, Part
Case, E.C., 1907, Revision of the Pelycosauria of North America: Carnegie 17A, 102 p.
Institution of Washington, Publication 55, 176 p. Rinehart, L.F., Lucas, S.G. and Harris, S.K., 2007, Lithology and taphonomy
Eberth, D.A., 1985, The skull of Sphenacodon ferocior, and comparisons of an early Permian Sphenacodon bonebed in Canon del Cobre, north
with other sphenacodontines (Reptilia: Pelycosauria): New Mexico Bu central New Mexico: New Mexico Geology, v. 29, no. 2, p. 63.
reau of Mines and Mineral Resources, Circular 190, 40 p. Romer, A.S. and Price, L.I., 1940, Review of the Pelycosauria: Geological
Glantz, S.A., 2005, Primer of Biostatistics: New York, McGraw-Hill Inc., Society of America, Special Paper 28, p. 1-538.
520 p. Shipman, P., 1981, Life history of a fossil: An introduction to taphonomy
Haldane, J.B.S., 1948, Suggestions as to quantitative measurement of rates and paleoecology: Cambridge, Harvard University Press, 222 p.
of evolution: Evolution, v. 3, p. 51-56. Simpson, G.G., Roe, A. and Lewontin, R.C., 1960, Quantitative Zoology:
Hammer, O., Harper, D.A.T. and Ryan, P.D., 2001, PAST: Paleontological New York, Harcourt, Brace and World Inc., 440 p.
statistics software package for education and data analysis, v.4, p. 9. Spielmann, J.A., Lucas, S.G., Berman, D.S and Henrici, A.C., 2009, An Early
Harris, S.K., Lucas, S.G., Berman, D.S and Henrici, A.C., 2004, Vertebrate Permian (Wolfcampian-Seymouran) vertebrate fauna from the Abo
fossil assemblage from the Upper Pennsylvanian Red Tanks Member of Formation (Scholle Member), Gallina Well, Socorro County, NM: New
the Bursum Formation, Lucero uplift, central New Mexico: New Mexico Mexico Geological Society, Guidebook 60, p. 69-70.
Museum of Natural History and Science, Bulletin 25, p. 267-283. Sumida, S. S., Lombard, R. E., Berman, D. S and Henrici, A. C., 1999, Late
Huene, F.v., 1908, Neue und verkannte Pelycosaurier: Reste aus Europe: Paleozoic amniotes and their near relatives from Utah and northeastern
Centralblatt fur Mineralogie, Geologie und Palaontologie, v. 14, p. 431- Arizona, with comments on the Permian-Pennsylvanian boundary in
434. Utah and northern Arizona: Utah Geological Survey, Miscellaneous Pub
Langston, W., Jr., 1952, Permian vertebrates of New Mexico [Ph.D. the lication, 99-1, p. 31-43.
sis]: University of California, Berkeley, p. 141-153. Vaughn, P.P., 1963, The age and locality of the Late Paleozoic vertebrates
Lucas, S.G., 2005, Permian tetrapod faunachrons: New Mexico Museum of from El Cobre Canyon, Rio Arriba County, New Mexico: Journal of
Natural History and Science, Bulletin 30, p. 197-201. Paleontology, v. 37, p. 283-286.
Lucas, S.G., 2006, Global Permian tetrapod biostratigraphy and Vaughn, P.P., 1964, Vertebrates from the Organ Rock Shale of the Cutler
biochronology; in Lucas, S.G., Cassinis, G. and Schneider, J.W., eds., Group, Permian of Monument Valley and vicinity, Utah and Arizona:
Non-marine Permian biostratigraphy and biochronology: Geological Journal of Paleontology, v. 38, p. 567-583.
Society, London, Special Publication 265, p. 65-93. Vaughn, P.P., 1969, Early Permian vertebrates from southern New Mexico
Lucas, S.G. and Krainer, K., 2005, Stratigraphy and correlation of the and their paleozoogeographic significance: Los Angeles County Mu
Permo-Carboniferous Cutler Group, Chama Basin, New Mexico: New seum Contributions in Science, v. 166, p. 1-22.
Mexico Geological Society, Guidebook 56, p. 145-159. Voorhies, M.R., 1969, Taphonomy and population dynamics of the early
Lucas S.G., Harris, S.K., Spielmann, J.A., Berman, D.S and Henrici, A.C., Pliocene vertebrate fauna, Knox County, Nebraska: Contributions to
2005, Vertebrate biostratigraphy and biochronlogy of the Pennsylva Geology, Special Paper No. 1.
nian-Permian Cutler Group, El Cobre Canyon, northern New Mexico: Walls, G.L., 1942, The vertebrate eye and its adaptive radiation: Cranbrook
New Mexico Museum of Natural History and Science, Bulletin 31, p. Institute of Science Bulletin 19, p. 1-785.
128-139. Williston, S.W., 1917, The phylogeny and classification of reptiles: Contri
Lucas, S. G., Spielmann, J. A., Rinehart, L. F., and Martens, T., 2009, butions of the Walker Museum, v. 2, p. 61-71.
Dimetrodon (Amniota: Synapsida: Sphenacodontidae) from the Lower Williston, S.W. and Case, E.C., 1913, A description of Edaphosaurus Cope:
Carnegie Institution of Washington, Publication 181, p. 71-81.

APPENDIX
Voorhies Group definitions used in this study; after Behrensmeyer (1975).
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

185
REDESCRIPTION OFTHE SKULL OF LIMMOSCELIS PALUDIS
WILLISTON(DIADECTOMORPHA: LIMNOSCELIDAE) FROM THE
PENNSYLVANIAN OF CAÑON DELCOBRE, NORTHERN NEWMEXICO

DAVID S BERMAN1, ROBERTR. REISZ2 AND DIANESCOTT2


1Carnegie Museum of Natural History, Section of Vertebrate Paleontology, Pittsburgh, Pennsylvania, USA 15213,
bermand@carnegiemnh.org; 2 Department of Biology, University of Toronto at Mississauga, 3359 Mississauga Rd.,
Mississauga, Ontario, Canada L5L 1C6, rreisz@utm.utoronto.ca, dscott@utm.utoronto.ca

Abstract—The skull of the Late Pennsylvanian diadectomorph Limnoscelis paludis Williston, 1911, is redescribed
in detail on the basis of the holotype YPM 811, the only one of three known specimens to include the skull.
Although originally collected by David Baldwin in 1879 in Cañon del Cobre, Rio Arriba County, north-central New
Mexico, the anatomy of the skull has been the subject of only two, relatively recent, comprehensive, detailed
descriptions. In addition to inaccuracies, these two studies contain numerous, contradictory interpretations. A
primary goal of the description presented here is an attempt to resolve these deficiencies. In the more recent of the
previous descriptions the skull had been separated into 17 pieces along pre-existent fractures and more extensively
prepared to facilitate study. Most were prepared further and illustrated for the present study, permitting descrip
tion of internal structures typically not available in intact specimens. On the basis of new anatomical information
a revised reconstruction of the cranium and mandible is presented. In addition, a reconstruction of the entire
skeleton of L. paludis in lateral and dorsal views and a whole-body restoration in lateral view is presented, but
without discussion. A more accurate, detailed description of the skull of L. paludis allows comparisons with other
diatectomorphs in light of a greatly expanded knowledge of their anatomy.

INTRODUCTION David Baldwin in 1879 in Cañon del Cobre, Rio Arriba County, north
central New Mexico, while employed by O. C. Marsh of the Yale Peabody
The Late Pennsylvanian Limnoscelis paludis Williston, 1911, has
Museum. These included an essentially complete, articulated skeleton
always been considered as one of the most important late Paleozoic (YPM 811) and two partial postcranial skeletons (MCZ 1947 and 1948,
tetrapods because of the widely accepted view of its close phylogenetic formerly YPM819 and 809, respectively). According to Fracasso (1980,
association with the origin of amniotes. Since its original description by 1983), the discovery of the specimen now believed to be the holotype,
S. W. Williston in a series of publications (Williston, 1911a, b, 1912),
YPM 811, which included the only known skull of L. paludis, was first
there have been only two detailed studies of the anatomy of the cranium
reported to Marsh by Baldwin in a letter dated September 19, 1879.
and mandible. The first was by A. S. Romer (1946), who reinterpreted However, as is so common in vertebrate paleontology, its significance
the anatomy of the cranium (defined here as the skull without the lower
was overlooked, probably due to the distraction of the discovery at that
jaw) on the basis of the growing knowledge of early reptiles. His recon time of the much more spectacular Como Bluff dinosaurs by another of
struction of the cranium contained only a few major errors, mainly in the Marsh’s collecting teams and his rival E. D. Cope. As a consequence, the
temporal and occipital regions (Berman, 2000). Subsequently, M. A.
L. paludis specimens were not studied until 30 years later by Williston
Fracasso (1987) described the skull (defined here as including the cra (1911a, b, 1912).
nium and mandible) in great detail in his Ph.D. thesis, but published only As the type species and best known member of Limnoscelidae, a
a redescription of the braincase (Fracasso, 1983), although accompanied detailed study of the skull morphology of Limnoscelis paludis will per
also by reconstructions of the cranium in ventral and occipital views mit a more rigorous evaluation of its relationships with Diadectidae and
without comments. More recent restudies (Berman and Sumida, 1990; Tseajaiidae, which together with Limnoscelidae constitute
Berman et al., 1992; Berman, 2000) of selective areas of the cranium have Diadectormorpha. Recent phylogenetic analyses (Berman et al., 1992;
noted numerous, important errors in the descriptions by Romer (1946),
Lee and Spencer, 1997; Berman, 2000; Kissel and Reisz, 2004) strongly
but most notably by Fracasso (1983, 1987). A reinterpretation of the
support the hypothesis that the three families form a natural group, the
occipital region of the cranium by Berman (2000) corrected some of the Diadectomorpha. Most importantly, however, the present study will
errors in their descriptions and reconstructions. provide a greater anatomical basis to help resolve the controversy of the
The continued interest in the skull anatomy of Limnoscelis paludis relationship of diadectomorphs to the amniotes. Amniota is divisible
since Williston’s studies (1911a, b, 1912) has resulted in several at into two universally accepted clades, Synapsida (extinct mammal-like
tempts at further preparation. Fracasso (1987) facilitated his study by forms and their mammalian descendants) and Reptilia (all amniotes ex
separating the skull into 17 pieces along existing, natural breaks, which clusive of Synapsida), and two hypotheses have been proposed relating
allowed an unprecedented access to internal structures. The skull was Diadectomopha to these two clades: firstly, the diadectomorphs com
still in a disassembled state at the initiation of the present study, but for
prise a nonamniote sister clade to all amniotes, with Diadectomorpha
purposes of illustration was partially reassembled into seven pieces.
and Amniota forming a monophyletic clade which constitutes the most
Unfortunately, some important features of the skull were lost or com recent definition of Cotylosauria (Laurin and Reisz, 1995, 1997; Reisz,
promised during preparation by previous investigators. As an example, 2007); and secondly, the Diadectomorpha and Synapsida comprise an
clearly the dermal sculpturing must have been much more extensive and unnamed, primitive sister clade to Reptilia (Berman et al., 1992; Lee and
pronounced than the current state of the skull indicates, but most impor Spencer, 1997; Berman, 2000). The latter hypothesis would mandate the
tantly, some details of the occiput and dentition were lost. assignment of Diadectomorpha to Amniota. However, in order to avoid
Although named and initially described by Williston in 1911, the any confusion in the description and comparisons of L. paludis pre
materials pertaining to Limnoscelis paludis were originally collected by sented below, the more widely accepted former hypothesis that
186
diadectomorphs comprise a nonamniote sister clade to all amniotes is tinguished from non-limnoscelid diadectomorphs by the following char
followed. acters: elongated snout in which the nasal is twice the length of the
Regarding the intrarelationships of the Diadectomorpha, there is parietal and has a greatest length equal to 50% of the midline skull length;
wide consensus that Diadectes Cope, 1878, and presumably also the the premaxillae united externally by a strongly interdigitating suture to
diadectids Desmatodon Case, 1908, and Orobates Berman, Henrici, Kissel, form a massive, bluntly rounded snout tip in which the ventral margin is
Sumida, and Martens, 2004, on the one hand, and Tseajaia Vaughn, 1969, inclined anteroventrally; long vomerine process of premaxilla extends to
on the other, share a more recent common ancestor than either does with
a level at least equal to that of the maxillary process of the premaxilla;
Limnoscelis (Berman et al., 1992; Kissel and Reisz, 2004; Reisz, 2007). anterior two teeth of premaxilla and second and third teeth of dentary
A thorough reconsideration of the phylogenetic relationships of the dominate the upper and lower marginal series as long, massive canines;
Diadectomorpha to Amniota, however, must await further study and long anterior and posterior processes of nasals taper laterally from mid
will be presented at a future date.
Despite a long history of extensive handling and repeated at line to extend far beyond their midline union; anterior portions of frontals
converge strongly on the midline to wedge deeply between posterior
tempts at preparation and description, the skull of Limnoscelis paludis
remains inadequately described. We have therefore undertaken that task processes of the nasals; broad, dorsally arching suture separates the
postorbital and supratemporal from the squamosal in which the lateral
here and present a revised, detailed description accompanied by a series margin of postorbital and supratemporal project ventrally and slightly
of illustrations of disassociated portions of the cranium and mandible, as laterally about 3 to 5 mm beyond lateral surface of squamosal in a struc
well as reconstructions of the cranium and mandible in several views. To ture likened to the eave of a house roof; lacrimal long and narrow, with a
this is added a revised reconstruction of the entire skeleton in lateral and height only slightly exceeding that of the maxilla; posterior process of
dorsal views, as well as a whole-body restoration (Fig. 1), but without jugal along ventral margin of skull roof extends much farther as it narrows
discussion. in a wedge-like terminus below the anterior end of quadratojugal; absence
In the description that follows, comparisons of Limnoscelis paludis
of a continuous single row of well-developed teeth that extends along the
with other diadectomorphs is limited to a few taxa whose cranial and
medial margin of premaxillary process of vomer and palatal process of
mandible anatomies are adequately known: the Late Pennsylvanian L. pterygoid; narrow, arcuate flange projects medially and slightly dorsally
dynatis Berman and Sumida, 1990, the only other valid member of the from a level just above the ventral edge of the quadrate ramus; lateral
genus (Berman and Sumida, 1990); among the Diadectidae the Early
surface of dentary at the symphysis of mandibles swollen into a mas
Permian type genus Diadectes, which has recently been redescribed in sive, hemispherical structure; prearticular and articular form a large, mas
detail (Berman et al., 1992; Berman et al., 1998; Berman, 2000), and the sive, anteriorly inflected pterygoideus process; a low, broadly convex
closely related, primitive diadectids Desmatodon hesperis Vaughn, 1969 coronoid eminence of mandible formed by the surangular.
(Vaughn, 1969, 1972; Berman and Sumida, 1995) and Orobatis pabsti Comment - Although several limnoscelids have been named and
Berman, Henrici, Kissel, Sumida, and Martens, 2004 (Berman et al.,
described on the basis of fragmentary materials, a recent taxonomic re
2004) from the Late Pennsylvanian and Early Permian, respectively; and view of the family (Wideman, 2002, Wideman et al., 2005) persuasively
the Early Permian Tseajaia campi Vaughn, 1964, sole member of
demonstrates that only two taxa are valid, Limnoscelis paludis and L.
Tseajaiidae, the holotype of which was described by Vaughn (1964) and
Moss (1972), but whose cranial anatomy description was more recently dynatis.
updated on the basis of a subsequently discovered specimen (Berman et DESCRIPTION AND COMPARISONS
al., 1992). In the description that follows the repetitive citation of these
references has been eliminated as an unnecessary redundancy. However, General -The skull of YPM 811 is essentially complete, missing
only the hyoid elements and stapes (Fig. 2). Although there has been
whenever a partial list of the comparative diatectomorphs is given, it
should be understood that the relevant morphological data are either some loss of portions of the roofing bones and postmortem compression
lacking or inconclusive for the excluded taxa. of the cranium, this has been easily compensated for in producing reli
ably accurate reconstructions in several views (Figs. 3-4). Modest over
ABBREVIATIONS all dorsoventral crushing with a slight oblique component directed vent
rolaterally to the right side has caused the most obvious distortions of
The following acronyms refer to institutional repositories of speci the cranium. This has distorted the entire cranium in several ways (Figs.
mens: CM, Carnegie Museum of Natural History, Pittsburgh, Pennsyl 2, 5-7, 12): 1) overall ventral depression of the skull table by approxi
vania; MCZ, Museum of Comparative Zoology, Harvard University, mately 2 cm; 2) minor ventral vaulting of the snout along the midline
Cambridge, Massachusetts; MNG, Museum der Natur, Gotha, Ger
union of the nasals and frontals; 3) inward buckling of the postorbital
many; YPM, Yale Peabody Museum, New Haven, Connecticut.
cheeks, resulting in a strong ventrolateral splaying of the cheeks from the
Key to abbreviations used in figures for anatomical structures: an,
angular; ar, articular; bo, basioccipital; c1, c2, anterior, posterior skull table; 4) breakage and displacement of the strut-like dorsal pro
coronoids; cp, cultriform process; d, dentary; e, epipterygoid; ec, cesses of the epipterygoids; and 5) some telescoping of the postorbital
bones of the right temporal region. Little distortion of the mandibles has
ectopterygoid; ex, exoccipital; f, frontal; j, jugal; l, lacrimal; m, maxilla;
n, nasal; op, opisthotic; p, parietal; pal, palatine; pf, postfrontal; pm, occurred, and they are tightly attached to the skull without any displace
ment at the jaw articulation. However, as a consequence of their tight
premaxilla; po, postorbital; pp, postparietal; pra, prearticular; prf, pre
frontal; ps, parasphenoid; pt, pterygoid; q, quadrate; qj, quadratojugal; attachment to the skull, particularly at the jaw articulation, the man
sm, septomaxillae; sp, splenials; sph, sphenethmoid; sq, squamosal; so, dibles have been displaced laterally due to the outward splaying of the
postorbital cheek regions of the cranium. Concomitant with this dis
supraoccipital; st, supratemporal; su, surangular; t, tabular; tt, tooth; v, placement the jaw has shifted slightly to the left side of the cranium, the
vomer. angle between the mandibles has increased, and the left mandible has
SYSTEMATIC PALEONTOLOGY been displaced dorsally to lie against the left side of the snout and slightly
lateral to the upper tooth row.
COTYLOSAURIA Cope, 1880 The cranium and mandibles are massively built but low in lateral
DIADECTOMORPHAWatson, 1917 profile, with a robust snout region that appears slender only because of
the elongation of the antorbital region (Figs. 2-4). The massiveness and
LIMNOSCELIDAE Williston, 1911
Limnoscelis paludis Williston, 1911 structural strength of the snout region is most prominently exemplified
by the large size of the premaxilla, but the entire cranium is also robustly
Revised Diagnosis - Diadectomorph cotylosaur that can be dis constructed, as indicated by the thickness of the skull roof elements,
187
188

FIGURE 2. Limnoscelis paludis, holotype, YPM 811. A, dorsal, B, ventral (palatal), and C, left lateral views of skull.
189

FIGURE 3. Reconstruction of Limnoscelis paludis, holotype, YPM 811. Cranium in A, dorsal, and B, lateral views; C, left mandible in lateral and medial
views.
190
particularly the circumorbital and ventral marginal bones (Fig. 9). The illa, as in Tseajaia, Desmatodon, Orobates, and Diadectes, forms nearly
interorbital and intertemporal regions of the cranium contribute to an the anterior two-thirds of the ventral border of the external naris. In
unusually broad, essentially flat skull table, whereas the dorsal elements Limnoscelis paludis it extends posteriorly as a vertical, wing-like blade
of the snout are slightly domed in transverse section. In lateral profile the
that contacts the medial surface of the premaxillary process of the max
nearly vertical postorbital cheek region is low and only slightly taller illa to about the level of its second tooth. Although Fracasso (1983) did
than the snout. The external nares and orbits face laterally, being visible not recognize a palatal, vomerine process of the premaxilla in L. paludis,
only slightly in dorsal view. The pineal foramen is transversely oval in
additional preparation has revealed a long, narrow pair of vomerine pro
outline. The sculpturing of the skull table ends posteriorly in a very low, cesses with rounded ventral margins exposed in a narrow, midline gap
occipital ridge that extends transversely just anterior to the posterior between the anterior ends of the vomers that clasped them together. The
margin of the parietal, then curves posterolaterally for a short distance exposed ventral margins of the vomerine processes reveal a posterior
along the anterolateral margin of the tabular. The ridge presumably delin extent equal to at least that of the maxillary processes of the premaxillae,
eates the boundary of the posterior extent of the direct skin attachment but their true posterior extent may be greater. Although the vomerine
to the skull roof. Posterior to the occipital ridge the skull table slopes
process of an isolated premaxilla of L. dynatis was described as short,
posteroventrally to occupy a plane somewhat intermediate between the reexamination indicates that it is incomplete distally, and its length is
horizontal plane of the skull table and the nearly vertical occiput. This
therefore indeterminate. The vomerine processes are extremely short in
narrow, smoothly finished, posterior portion of the skull table is bowed Tseajaia, Orobates, and Diadectes.
slightly anteriorly in dorsal view and includes a very thin strip of the Each premaxilla has four tooth positions, with the first two being
posterior margin of the parietal, the entire, single, median postparietal, represented by a tooth and socket pair but with reversed serial replace
and a posterolaterally directed, horn-like portion of the tabular. ment patterns. The first two teeth greatly dominate the series in mas
In palatal view of the cranium (Fig. 4A) the long anteroposterior
siveness and length, although exhibiting a marked decrease in size poste
length of the internal nares, which occupies nearly 30 % of the skull
riorly. The massiveness of the premaxilla is undoubtedly related to the
length, reflects the great elongation of the snout. A moderately deep, large size and deep implantation of this pair of teeth. The much smaller
anteroposteriorly elongate, narrow, oval, basin-like depression straddles posterior two premaxillary teeth exhibit a dramatic decrease in size pos
the contact of the medial margin of the ectopterygoid with the pterygoid
teriorly. The premaxillary teeth are similar, however, in being essentially
just anterior to the transverse flange. An identical depression is seen in
subcircular in cross-section with the basal halves maintaining a constant
Diadectes (Olson, 1947), whereas in Orobates there is a less well-de diameter and the distal halves or crowns tapering to a slightly recurved
fined, shallow depression of the ectopterygoid adjacent to its sutures point. In addition, they exhibit very fine, narrowly spaced, groove-like
with the pterygoid and palatine. The narrow interpterygoid vacuity
striations that are most noticeable at their bases. Cross-sections of the
occupies half the length of the palate anterior to the basicranial articula teeth reveal the striations to be indicative of a labyrinthine infolding of
tion. The subtemporal adductor fossae, the largest paired openings on the surface enamel that appears to penetrate the full thickness of the
the ventral surface of the cranium, are located between the quadrate
dentine the entire length of the root and crown. Loss of the surface bone
ramus of the pterygoid and the contributions of the quadratojugal and of the premaxillary teeth precludes a description of their superficial
jugal to the ventral rim of the cranium. Their large size undoubtedly
features. The marginal teeth in Limnoscelis paludis are similar to those of
reflects a massive development of the adductor musculature. other diadectomorphs in their deep implantation and apparent infolding
The occipital region of the skull (Figs. 2, 3A, 4B) is broad and low of the surface enamel deeply into the dentine.
and dominated by a large occipital plate that is inclined slightly
Berman and Sumida (1990) noted one possible noteworthy differ
anterodorsally, which is barely sufficient to expose almost its entire ence between the premaxillae in Limnoscelis paludis and L. dynatis. In
surface, as well as the foramen magnum, in dorsal view. Vaughn’s (1969) description of the right premaxilla in L. dynatis he recog
The mandibles are massively built despite having a relatively low
nized only two teeth. However, as noted by Berman and Sumida (1990),
lateral profile (Figs. 2C, 13-14). They are markedly shorter than the what Vaughn (1969) described as the base of the larger and more anterior
skull, suggesting that when tightly closed the large anterior mandibular of the two, actually represents two teeth. But, as they cautioned, the
teeth must have fitted into the anterior ends of the internal nares. The two teeth may represent a stage of one tooth replacing the other in a
symphyseal area is massive, the coronoid prominence is low and ex single tooth socket and therefore should be counted as a single functional
ceeded in height by the largest marginal teeth at the symphysis, and the tooth.
medial fenestra is narrow but long, with a length equaling 28% of that of
Septomaxilla - The septomaxillae (Figs. 3B, 6, 8) are nearly
the mandible. The sculpturing pattern of the lateral surface mandible at
complete, but have been distorted slightly by the dorsoventral crushing
its massive anterior end is dominated by coarse pitting, whereas most of
of the cranium, but together they permit a reliable reconstruction of its
the posterior half of the lateral surface exhibits a very fine pattern of
anatomy and relationship with neighboring elements. It is a squat ele
alternating grooves and ridges radiating from a central area of minute pits ment that has in general an hourglass outline with large, flared ventral and
on the surangular.
dorsal ends, the former being substantially larger, which are separated by
Skull Roof a short, narrow, pillar-like dorsal process. Ventrally it contacts mainly
the premaxilla and very narrowly the maxilla and lacrimal along the ven
Premaxilla -The premaxillae (Figs. 3A, B, 4A, 6, 8, 11) are united tral margin of the naris, whereas dorsally it appears to have ended short
along the dorsal midline in a strongly interdigitating suture and form a of a contact with the nasal border of the naris. The dorsal process is
massive, bluntly rounded snout tip in which the ventral margin in lateral pierced anteroposteriorly by a septomaxillary canal (not visible in illus
view is inclined anteroventrally. The surface sculpturing of the premax trations). In all these features it is closely comparable to the septomaxillae
illa is more pronounced than on other parts of the skull roof, despite in Orobates and Diadectes. The septomaxilla in diadectomorphs is very
some weathering. This is due in part to the premaxillary sculpturing similar to that in basal, pelycosauria-grade synapsids, differing mainly in
being associated with a series of foramina that probably allowed passage the latter having a dorsal contact with the nasal (Romer and Price, 1940;
of blood vessels and nerves to the skin of the snout. Three processes Wible et al., 1990).
unite the premaxilla with the maxilla, vomer, and nasal. Joined, narrow Nasal - The nasals (Figs. 3A, B, 6, 8) are extremely thick, and in
dorsal processes of the premaxillae extend posterodorsally along the overall length are the longest of the paired midline skull table elements,
midline between anterior processes of the nasals. Fracasso (1983) incor reflecting the great elongation of the snout. They not only form the
rectly described the dorsal processes as much shorter and wider and as dorsal surface of the snout, but also contribute significantly to the vent
enclosing a midline internasal bone. The maxillary process of the premax rolateral curvature of the cheek. Their external surface retains at least
191

FIGURE 4. Reconstruction of cranium of Limnoscelis paludis, holotype, YPM 811, in A, ventral (palatal), and B, occipital views.

remnants of the original ornamentation. Long anterior and posterior pro bital, and supratemporal, respectively. An unusual feature is an extremely
cesses of the nasals taper laterally as they extend far beyond their mid narrow, spike-like projection of the posterolateral corner of the parietal
line union. The anterior process contacts the entire lateral margin of the that partially separates the tabular and supratemporal. The parietal not
dorsal process of the premaxilla to form the dorsal margin of the naris and only has a strongly interdigitating suture with the frontal, which on the
end in a short, narrow, anteroventrally directed splint of bone. The pos ventral surface of the skull roof underlies extensively the frontal. A break
terior processes of the nasals, on the other hand, are deeply wedged through the left side of the skull table reveals that the parietal also
between the anterior portions of the frontal and prefrontal. narrowly underlies the supratemporal. The sculpturing of the parietals
Frontal - As in all diadectomorphs, the paired frontals (Figs. 3A, ends posteriorly in a very low occipital ridge narrowly anterior to and
5-6, 9) are relatively narrow and excluded widely from the orbital margin parallel with their posterior margin. The midline parietal suture is inter
by a wide contact between the postfrontal and prefrontal. In contrast to rupted at its midlength by a large pineal foramen, and, although it is
the condition in other diadectomorphs, however, where the frontals are relatively smaller than that in Tseajaia, Orobates, and Diadectes, it occu
narrowly subrectangular in outline, in Limnoscelis paludis the anterior pies 30% of the midline length of the parietal.
portions converge strongly on the midline, wedging deeply between Tseajaia, Orobates, and Diadectes also possess a lateral lappet of
posterior processes of the nasals, whereas posteriorly they expand gradu the parietal, but it differs from that in Limnoscelis paludis in having a
ally to their contact with the parietals. The anterior half of their midline relatively greater transverse width and projecting directly laterally. The
contact is essentially straight, whereas posteriorly it is strongly inter contacts of the lateral lappet to adjacent elements are identical in L.
digitating. paludis, Orobates, and Diadectes, but in Tseajaia the lateral margin also
Parietal - The large parietals (Figs. 3A, 5, 10) dominate the pos contacts the squamosal. The parietal of Tseajaia is also distinguishable
torbital skull table. Their great width is due in part to a well-defined, by its lack of the spike-like projection of its posterolateral corner that
subrectangular lateral lappet that extends a short distance anterolaterally partially separates the tabular and supratemporal.
and represents either a replacement or an incorporation of the Postparietal - The postparietal (Figs. 3A, 4B, 5, 12) is a single,
intertemporal bone (Berman et al., 1992, 1998). The anterior, lateral, and median element with a trapezoidal outline in which the transverse width
posterior margins of the lappet are bordered by the postfrontal, postor is about five times the midline length, and its smoothly finished surface
192
lacks the typical, midline ridge of cotylosaurs. The postparietal is also a short distance anterior to the orbits (Fig. 9) reveals the dorsal lamina as
single element in Orobates and Diadectes, but paired in Tseajaia. As in being slightly thinner and subequal in height to the alveolar shelf. At this
other diadectomorphs, the dorsal surface of the postparietal, although level the dorsal lamina has a strongly interdigitating suture with the
exhibiting a slight anterior bowing, occupies a plane intermediate be ventral margin of the anterior end of the suborbital process of the jugal, as
tween the horizontal skull table and the slightly anterodorsally inclined well as sheathing its lateral and medial surfaces. The entire dorsal margin
plane of the occiput. In non-cotylosaurs the postparietal occupies a of the maxilla appears to be narrowly overlapped laterally by the lacri
horizontal plane that is continuous with skull table, whereas in amniotes mal in a nearly straight suture.
it occupies a near vertical plane that is continuous with the occiput. The The maxillary teeth are set in very deep sockets. The implanta
occipital margin of the postparietal overlaps the dorsal edge of the su tion, however, cannot be considered strictly as thecodont, because the
praoccipital in a rounded, slightly overhanging projection in what ap ventral surface of the alveolar shelf slopes steeply ventrolingually, creat
pears to be a loosely sutured contact. ing a superficially pleurodont-like implantation (Figs. 9, 11). In addition,
Prefrontal - The large prefrontal (Figs. 3A, B, 5-8, 10), is almost reabsorption dental pits are clearly visible at the base of the lingual
entirely restricted to the dorsal skull table as essentially an margin of several teeth (Fig. 14). The tooth row, which includes 23 tooth
anteroposteriorly elongated rectangular, plate-like structure. It exhibits positions, extends nearly the entire length of the maxilla with only the
an otherwise strong ventral curvature into the vertical cheek surface of short, subnarial, premaxillary process being edentulous. The maxillary
the snout. A transverse break through the skull roof at a level just anterior teeth exhibit a gradual but steady increase in size serially from the first to
to the orbits (Fig. 9) reveals the prefrontal as forming a massive, out about the fifth tooth, then gradually decrease to the end of the series.
wardly directed, wedge-shaped, supraorbital shelf. The greater lower Although all the previously exposed teeth have lost their superficial
portion of the ventrolateral surface of the shelf has an extensive contact features, the fourth, left maxillary tooth (Fig. 14), which was not previ
with the dorsal half of the medial surface of the lacrimal. Posterior to its ously exposed, exhibits a sharp, very narrow, low, cutting ridge that
contact with the lacrimal the free, ventromedial, angular extension of the extends the entire fore and aft lengths of the crown and passes through
shelf is quickly reduced to a low, rounded ridge that is continued a short the crown tip. The distal third of the crown, as well as the three preced
distance along the internal surface of the prefrontal-frontal contact and ing teeth, is slightly recurved. Presumably these features were present in
then onto the postfrontal (Fig. 10A). Remnants of the surface sculptur all the maxillary except possibly those of the posterior end of the series.
ing of the prefrontal, as in the nasal, are present as shallow pits and faint In contrast to the condition in Limnoscelis paludis, the maxillary
ridges and grooves. dentitions in L. dynatis and Tseajaia include a pronounced caniniform
Postfrontal - The postfrontals (Figs. 3A, B, 5, 7, 10) are well tooth at the anterior end and the midlength of the series, respectively.
preserved and almost entirely restricted to the dorsal skull table, where The upper and lower marginal dentitions in Desmatodon, Orobates, and
they appear subrectangular in outline. They are plate-like except for a Diadectes are unique in that the teeth of the premaxilla and the anterior
slight ventral curvature along the orbital margin, especially at the ends of the maxilla and dentary are incisiform and much larger than the
posterodorsal margin of the orbit. The medial margin parallels the dorsal succeeding cheek teeth, which are molariform.
orbital margin and contacts mainly the frontal, but also occupies the Lacrimal - The entire length of the right lacrimal (Fig. 8) has been
right-angle notch between the anterior margin of the parietal and its severely crushed and fractured, whereas the left (not illustrated) is nearly
lateral lappet. The anterior and posterior contacts with the prefrontal intact, but both have been buckled medially (Fig. 9). Restored (Fig. 3A,
and postorbital, respectively, are oriented normal to the dorsal orbital B) the lacrimal appears as an elongate, rectangular, strap-like structure
rim. Near its posterior, orbital margin the ventral surface thickens greatly that extends between substantial contributions to the external naris and
and from it a transversely expanded, flange-like process that projects a orbit. For most of the lacrimal’s anterior length, however, it exhibits a
short distance ventrally along the posterior orbital wall. rather slight ventrolateral curvature in transverse section, so that in the
Postorbital - Exposure of the postorbital is divided nearly equally reconstructed skull nearly its upper half is visible in dorsal view. Anteri
between the skull table and cheek (Figs. 3A, B, 5, 7). On the skull table orly the lacrimal narrows slightly to match the height of the maxilla, as it
it contacts the entire lateral margin of the lateral lappet of the parietal forms the greater ventral portion of the posterior rim of the external
between contacts with the anterior postfrontal and posterior supratem naris. Here it joins with the maxilla in forming a small, medially directed,
poral. In lateral view the cheek portion of the postorbital forms the narial shelf at the posteroventral corner of the naris. Above the narial
greater portion of the posterior rim of the orbit from which it projects shelf the lacrimal duct (not visible in the illustrations) exits the narial
posteriorly in a wing-like outline. Its orbital margin is greatly thickened margin of the lacrimal. The ventral margin of the lacrimal contacts the
and ends ventrally in a laterally overlapping contact with a short, narrow, entire dorsal margin of the maxilla except for a narrow contact with the
dorsal orbital process of the jugal. Superficially this suture is directed suborbital process of the jugal at the anteroventral corner of the orbit. A
posterodorsally and continues in a broad, dorsally arching contact be shallow, broadly concave depression of the posterior third of the lacrimal
tween the ventrolateral margins of the postorbital and supratemporal deepens slightly posteriorly, where it is bordered by the thickened,
and the dorsal margin of the squamosal. This suture is unusual in that the rounded orbital rim. A small V-shaped notch on the lateral surface of the
postorbital and supratemporal project about 3 to 5 mm beyond the orbital rim marks the exit of a single, large foramen for the lacrimal duct.
lateral surface of the squamosal in what can be likened to the eave of a In overall proportions the lacrimals in Tseajaia, Orobates, and
house roof. Whereas the external surface of the eave is a continuation of Diadectes differs from that in Limnoscelis paludis in being relatively
the sculptured skull table, the internal surface is smoothly beveled to an much shorter and taller, with the height greatly exceeding that of the
acute-angled edge along its free margin. maxilla.
Maxilla - The long, low maxilla, which extends between the Jugal - Although the jugals have suffered severe breakage and
midlength levels of the external naris and orbit, reflects the skull’s profile some distortion (Figs. 13-14), their structure can be reconstructed with
(Figs. 3A, B, 4A, 8, 11, 13-14). In lateral view the ventral edge of the confidence (Fig. 3B). It is a relatively long, triradiate element, extending
maxilla is slightly convex for almost its entire length except at its anterior posteriorly from the level of the anterior orbital margin to shortly be
end, where it is redirected slightly anteroventrally beneath the external yond the midlength of the postorbital cheek. A broad, anterior suborbital
naris. Here it not only overlaps laterally the maxillary process of the process forms almost the entire ventral orbital rim, whereas a short,
premaxilla, but also contributes to the ventral shelf of the external naris narrow, dorsal process contributes to the posterior orbital margin. The
to support partially the base of the septomaxilla. The lateral surface of distal end of the dorsal process of the jugal, which is exposed in cross
the maxilla extends a short distance above a thick alveolar shelf as a low, section in the right (Fig. 7, 10) and medial view in the left (Fig. 10), is
vertical, dorsal lamina. A transverse break passing through the snout a laterally overlapped by the ventral, orbital process of the postorbital.
193

FIGURE 5. Limnoscelis paludis, holotype, YPM 811. Posterior portion of cranium in dorsal view. A, Illustrated as preserved, and B, emphasizing preserved
limits of elements.
194

FIGURE 6. Limnoscelis paludis, holotype, YPM 811. Anterior portion of cranium in dorsal view. A, Illustrated as preserved, and B, emphasizing preserved
limits of elements.

However, as seen on the left side of the posterior portion of the cranium squamosal from a narrow contact with the distal end of the dorsal margin
exposed in the ventral view (Fig. 10), dorsoventral crushing has tele of the posterior process of the jugal to the occipital margin of the skull
scoped the contact between the two processes, producing the false ap roof. The quadratojugal then thickens as it curves abruptly medially onto
pearance depicted in the reconstructions by Romer (1946) and Fracasso the occipital surface of the skull roof (Figs. 3B, 4, 10, 12, 14). Here its
(1980, 1983) that the jugal lacks a dorsal orbital process. Although the exposure is abruptly narrowed dorsally by a ventral, occipital expansion
dorsal process in the reconstruction presented here is short, a longer of the squamosal before contacting the lateral half of the posterior sur
process, as in other diadectomorphs, is likely. This is most strongly face of the articular condyle of the quadrate. The quadratojugal ends
supported by the dorsal process of the isolated jugal of Limnoscelis medially in a short, sheet-like expansion that with the quadrate form the
dynasties, which is much longer and formed nearly the ventral half of the lateral and medial margins of the quadrate foramen, respectively. The full
posterior orbital margin. In L. paludis, as well as probably in L. dynatis, dorsal extent of the quadratojugal-quadrate contact above the quadrate
the posterior process is extremely long, extending along the ventral mar foramen is covered by the lateral margin of the occipital exposure of the
gin of the skull roof to a short distance beyond the midlength level of the squamosal (Figs. 4, 12). However, within the subtemporal adductor
postorbital cheek (Fig. 13). The dorsal margin of the process quickly chamber, as in amniotes, the squamosal-quadrate suture is supported by
descends to form an acute-angled terminus below the anterior end of the a dorsal process of the quadratojugal that overlaps their contact (Fig. 10).
quadratojugal. As a result of its great length the posterior process of jugal Whereas the pattern of sculpturing on the lateral exposure of the
accounts for about 28% of the ventral margin of the skull roof. In marked quadratojugal is similar to that seen on the jugal and squamosal, it is
contrast, the posterior process in Tseajaia, Orobates, and Diadectes greatly subdued on its occipital surface.
ends well short of the midlength level of the postorbital cheek and is Squamosal - In lateral view the squamosal occupies most of the
directed away from the ventral margin of the skull roof as it ends wedged postorbital cheek region, forming a rather flat, disk-like structure whose
between the quadratojugal and squamosal. dorsal margin forms a broad, dorsally arching contact with the postor
Quadratojugal - From about midlength of the postorbital cheek bital and supratemporal (Figs. 3B, 5, 7, 13). Two unusual features are
the quadratojugal forms the posteroventral corner of the skull roof, with exhibited by the squamosal along this contact: 1) sculpturing covers all of
both lateral and occipital exposures. Its lateral exposure (Figs. 3B, 13) its surface except for a narrow band of smoothly finished bone along its
increases in height posterodorsally in a broadly concave contact with the dorsal margin; and 2) its dorsal sutural margin is expanded dorsomedially
195

FIGURE 7. Limnoscelis paludis, holotype, YPM 811. Posterior portion of cranium in right lateral view. A, Illustrated as preserved, and B, emphasizing
preserved limits of elements.

in the form of a well-developed, approximately 5 mm wide flange that evolved (Kemp, 1980), or simply a “loose junction” (Fracasso, 1983).
thins to an acute-angled edge. The flange contacts the ventral surfaces of Fundamental to these interpretations is the flawed observation that the
the postorbital and supratemporal 3 to 5 mm from their ventrolateral, contact between the cheek and skull table is a loose junction that was
eave-like margin. An identical, medially directed sutural flange has also inherited from osteolepiform antecedents, where it was integral to an
been described in an isolated right squamosal of Limnoscelis dynatis. intracranial kinetic complex. On the basis of the redescription given here,
The eave-like structure of the contact of the squamosal with the however, the so-called line of weakness in L. paludis must be reinter
postorbital and supratemporal in Limnoscelis paludis is unique among preted as representing a complex, sutural contact that was surely immo
Paleozoic tetrapods and has been variously explained as either a “line of bile. Ventral exposure of the skull roof (Fig. 10) further strengthens this
weakness” (Romer, 1946), a “kinetic junction” (Panchen, 1972), the conclusion in revealing that the dorsomedially directed sutural flange of
“temporal line” site where the upper temporal fenestra of synapsids the squamosal extends beyond the postorbital-supratemporal to contact
196
also the ventral surface of the posterolateral corner of the lateral lappet two components, but somewhat modified from that of Limnoscelis
of the parietal. In addition, the anterior corner of the squamosal medially paludis. Similarly, a proximal, horn-like component forms almost the
overlaps the sutural contact between the postorbital and jugal (Fig. 7), entire posterolateral corner of the skull table. The distal occipital flange
effectively uniting these elements in an immobile union. Although the component in both diadectids, however, differs in forming a greatly
unique structural relationship of the union between the postorbital cheek elongated, narrow, process-like extension that curves abruptly ventrally
and skull table can no longer be characterized as a line of weakness, a into the plane of the occiput. Whereas the ventral extension is separated
plausible, alternative functional explanation also cannot be offered. by a narrow gap from the occipital surface of the quadrate, rather than
The posterior margin of the cheek exposure of the squamosal the squamosal, the distal end narrowly contacts the distal end of the
(Figs. 3B, 4B, 5, 7, 12) curves abruptly medially onto the occipital paroccipital process. In Tseajaia the supratemporal appears to consist
surface of the skull roof in what can be referred to as the occipital flange only of the proximal, horn-like component, which also occupies nearly
of the squamosal. Ventrally the flange has a short, bluntly pointed pro the entire posterolateral extension of the skull table. Interestingly, it
cess that overlaps the quadratojugal. At the posterior margin of the appears to contact the lateral margin of the postparietal, rather than that
lateral exposure of the squamosal the occipital flange is demarcated by an of the tabular.
approximately 1 mm, step-like, depression of its surface that produces a Tabular - The tabular (Fig. 3A, 4B, 5, 7, 12) is a complex bone
ridge between the two exposures. The two exposures of the squamosal with large exposures on both the skull-roof table and occiput that can be
are also distinguished by a dramatic change in surface texture; a very fine described as divisible into three components (Berman, 2000). A proxi
sculpturing pattern of alternating grooves and ridges radiating from a mal, horn-like component joins with the much larger, complementary
central area of minute pits covers all of its lateral cheek exposure except component of the supratemporal in forming the posterolateral corner of
the narrow, smoothly finished band along the dorsal margin, whereas the the skull table. This component, considered part of the skull table, is
occipital flange is entirely smoothly finished. The occipital flange of the exposed almost entirely as a flat, smoothly finished surface that slopes
squamosal not only extends medially across the posterior surface of the slightly posteroventral at the same plane as the postparietal, intermedi
dorsal extension of the quadratojugal, but also a narrow margin of the ate between the skull table and the nearly vertical occiput. The distal end
dorsal extension of the quadrate above the articular condyle. Dorsally the of the horn-like component does exhibit, however, a distinct, low, dome
occipital flange extends about 2 mm beneath a freely projecting margin of like swelling of unknown function. A second much smaller component,
an occipital exposure of the supratemporal. referred to as the distal occipital flange of the tabular, is smoothly con
Supratemporal - The supratemporal is best described as con tinuous with ventral margin of the distal end of the proximal, horn-like
sisting of two components (Berman, 2000). One, referred to as the proxi component. It is poorly preserved on both sides of the skull roof, but is
mal, horn-like component, is triangular in outline, finely sculptured, and most complete on the left. It faces posteriorly and slightly medially and
occupies the greater portion of the posterolateral, horn-like extension of has the form of a smoothly surfaced, thin, flange-like structure with a
the skull table (Fig.3A, B, 4B, 5, 7, 12). Posteriorly it curves slightly broadly rounded, ventrolateral margin that overlaps the medial margin of
ventrally, where the greater distal portion of medial margin of the horn the distal occipital flange of the supratemporal. As restored, the tabular
like component is defined by an approximately 1 mm, step-like depres- supratemporal suture ends distally at the lateral margin of what has been
sion of surface, producing a ridge that lines up with the identical ridge identified (Romer, 1946; Berman, 2000) as the posttemporal fenestra.
along the lateral margin of the occipital flange of the squamosal. The ridge Romer (1946) and Berman (2000) suggested that the distal occipital
also demarks the second component of the supratemporal, a thin, flat, flange of the tabular may have continued farther medially, forming the
smoothly finished structure referred to as the distal occipital flange. It ventral margin of the posttemporal fenestra and ending in a narrow con
extends a short distance ventrally before curving abruptly medially around tact with the dorsolateral corner of the paroccipital process of the
the ventrolateral margin of a much smaller similar component of the opisthotic. This contact, as in the case of the complementary structure
tabular (Figs. 3A, B, 4B, 5, 12). An approximately 2 mm space separates of the supratemporal, was likely lost due to weathering and/or excessive
the flange from the underlying occipital flange of the squamosal. As preparation. This interpretation is supported by a small, triangular frag
restored (Fig. 4B), the distal occipital flange of the supratemporal forms ment of what appears to be part of the distal occipital flange of the right
the ventrolateral and most of the ventral margin of what has been identi tabular that retains its contact with the posterior surface of the dorsolat
fied (Romer 1946; Berman, 2000) as the posttemporal fenestra and ends eral corner of the paroccipital process. In addition, immediately above
in a bluntly rounded process just a few mm from the distal end of the the contact is a small, smoothly rounded notch in the lateral margin of the
paroccipital process of the opisthotic. However, in Romer’s (1946) de occiput (lost on the right side of the occiput) that undoubtedly repre
scription (p. 158) and reconstruction (fig. 8, p. 154) of the occiput of sents the medial border of the posttemporal fenestra. Fracasso (1983,
Limnoscelis paludis the distal, occipital flange of the supratemporal 1987) misinterpreted the occipital-flanges of the supratemporal and tabu
(misidentified as part of the tabular) ends in a broad contact with the lar and the squamosal exposure on the occipital surface of the cheek as
distal margin of the paroccipital process. The same relationship between parts of a long, ventral extension of the tabular that extended along the
these two structures was proposed by Berman (2000), which is sup occipital surface of the skull cheek to nearly the ventral margin of the
ported by three lines of evidence: 1) the distal end of the more complete skull.
left distal occipital flange of the supratemporal is at least marginally The remaining component of the tabular, the ventromedial occipi
incomplete; 2) the left paroccipital process of the opisthotic, as seen in tal flange (referred to as the tabular cone by Fracasso, 1983, 1987), is
occipital view (Fig. 12), is definitely incomplete distally, as it is notice restricted to the occiput and occupies the medial angle formed between
ably shorter than the right. Weathering and/or aggressive preparation the other two components of the tabular. It has the form of a thick, flat
could easily account for the approximately 3 mm gap separating the two plate that broadens slightly as it extends medially to contact the lateral
structures on the left side; and 3) the very narrow separation could also margin of the supraoccipital region of the occiput and the posterior
be accounted for by the dorsoventral crushing of the skull, which has surface of the dorsolateral corner of the paroccipital process of the
resulted in the cheeks being splayed ventrolaterally from their original, opisthotic. For most of its extent it exhibits the same roughened, surface
more vertical orientation (Fig. 12). All these altered features were either texture as the occiput. The smoothly finished, lip-like structure of the
overlooked or ignored by Kissel and Reisz (2004) and Reisz (2007), who occipital margin of the postparietal is continued laterally onto the tabular
stated that there is no direct evidence of a contact between the supratem in a broad, dorsal arch that defines the boundary between the proximal
poral and distal end of the paroccipital. horn-like component and the ventromedial occipital flange of the tabular.
The supratemporal in Orobates and Diadectes can also be de The lateral portion of the ventral margin of the ventromedial occipital
scribed (Berman, 2000; Berman et al., 2004) as divisible into the same flange of the tabular forms the dorsal margin of the posttemporal fenes
197

FIGURE 8. Limnoscelis paludis, holotype, YPM 811. Anterior portion of cranium in right lateral view. A, Illustrated as preserved, and B, emphasizing
preserved limits of elements.
198
tra. reveals a vertical thickening of the palatine medial to the channel on its
The same three components of the tabular in Limnoscelis paludis ventral surface. At its medial margin the palatine bifurcates into thinning,
can be recognized, though greatly modified, in Orobates and Diadectes shelf-like processes between which the lateral margin of the pterygoid is
(Berman, 2000; Berman et al., 2004). In both diadectids the posterolat clasped. The much thinner, dorsal shelf of the bifurcation extends the
eral corner of the sculptured, proximal, horn-like component of the tabu farthest medially and, although not visible in the cross-section, is in turn
lar curves abruptly ventrally from the skull table into the plane of the overlapped dorsally at its anterior and posterior ends by the medial
occiput in the form of a greatly elongated, narrow process. Distally it margins of the vomer and ectopterygoid, respectively. In the anterior,
ends in a narrow contact with the dorsal margin of a greatly shortened suborbital region of the skull roof the palatine also contacts the internal
paroccipital process. A narrow, smoothly finished, flange-like extension area of juncture of the maxilla, jugal, and lacrimal. A narrow field of
of the medial margin of the proximal, horn-like component, considered minute denticles extends along the posterior half of the medial margin of
equivalent to the combined distal medial and occipital flange components the ventral surface of the palatine.
of the tabular, contacts the lateral margin of the occiput. It encloses the Ectopterygoid - Although smallest of the palatal elements (Fig.
posttemporal fenestra laterally and extends a short distance farther ven 4A), the greatly thickened structure of the ectopterygoid unites firmly
trally to contact the dorsal margin of the paroccipital process. In Tseajaia the palate with the suborbital region of the skull roof. As in the palatine,
all that is visible of the tabular is a narrow, dorsally sculptured, thumb it thickens greatly laterally, forming not only a buttress-like contact with
like process, presumably the proximal, horn-like component, which the medial surface of the alveolar shelf of the maxilla, but also with the
projects a short distance beyond the distal end of its contact with the jugal. The posterolateral corner of the ectopterygoid has a short, verti
ventral surface of the same component of the supratemporal. Together cally rounded edge that narrowly contributes to the anterior margin of
they project posteroventrally over the dorsal margin of the occipital the narrow, notch-like anterior extension of the subtemporal adductor
surface of the quadrate. fossa. The posterior border of the ectopterygoid is thickened ventrally
toward its transverse contact with the transverse flange of the pterygoid,
Palate so as to be slightly convex ventrally in transverse section. The medial
margin of the convexity borders an anteroposteriorly elongated, narrow,
Vomer - The vomers are long in Limnoscelis paludis, equaling
oval basin-like depression that deepens posteriorly as it extends along
almost 30% of the total cranium length, and form the entire medial margin the greater posterior length of the ectopterygoid-pterygoid contact. The
of the internal nares (Figs. 4A, 11). The approximately anterior half of
the vomer, the premaxillary process, has the form of a vertical, depression does not correspond to the suborbital foramen seen in some
primitive amniotes, which is located typically at the juncture between
subrectangular sheet that closely approaches, possibly even contacts, the palatine, ectopterygoid, and pterygoid. However, at or near the junc
the ventral surface of the nasal, and is narrowly separated from its mate
along the mid-sagittal plane. Anteriorly they nearly overlap the ventral ture of these elements a very similar depression is also present in Orobates
margin of the vomerine process of the premaxilla and end anteriorly in a and Diadectes. The ectopterygoid is the only edentulous element of the
short, forked margin. Posteriorly the premaxillary processes of the vomers palate.
thicken along their ventral margin as they curve strongly medially to Pterygoid - The pterygoid is divisible into the three standard
components: anterior palatal process, transverse flange, and posterior
contact one another. Farther posteriorly the vertical, sheet-like vomer
quadrate process. Anteriorly the long palatal processes (Figs. 4A, 9, 11)
twists about its long axis as it diverges from the midline to form a poste
rolateral, wing-like extension with the lateral surface now facing ventro converge on the midline, wedging a short distance between the postero
lateral wings of the vomers. A field of fine denticles covers the palatal
laterally and slightly anteriorly. The posterolateral wing of the vomer
process except for a narrow margin bordering the interpterygoid vacuity.
narrowly overlaps the ventral, lateral margin of the wedge-shaped, united In Tseajaia, Orobates, Desmatodon, and Diadectes a single row of large
anterior ends of the pterygoids. Posteriorly the vomer ends in a V
shaped, marginal contact with the anterior margin of the palatine. A teeth, supported by a well-developed ridge, extends along the medial
short, narrow field of minute denticles, which closely parallels the pos margin of the palatal process. With the exception of the vomer not being
terior end of the midline union of the premaxillary processes, is contin known in Desmatodon, the pterygoid tooth row is smoothly continuous
with that of the premaxillary process of the vomer. A short, rectangular
ued posteriorly onto the pterygoid. In Tseajaia, Orobates, and Diadectes lateral expansion at the posterior end of the palatal process contacts the
a single row of well-developed teeth closely parallels the medial margin
of the premaxillary process. medial margin of the ectopterygoid. The oval depression described above
as extending along this contact is bordered medially by a well-developed,
Fracasso’s (1983, 1987) description of the vomer had several
buttress-like ridge of the pterygoid.
important errors. Because he was unable to detect the midline palatal
suture between the premaxilla and vomer, he misinterpreted the maxil Between the levels of the posterior margin of the transverse flange
lary process of the premaxilla as a lateral, wing-like extension of the and basicranial articulation the pterygoid is constricted to a short, neck
like strut with a dorsoventrally thickened, rounded medial margin (Figs.
vomer. Furthermore, he incorrectly indicated the vomer-palatine suture
as nearly straight and directed slightly anterolaterally, and positioned 4A, 10). On its lateral margin a web-like flange curves anterolaterally a
slightly anterior to the posterior margin of the internal naris, which short distance onto the posterior margin of the transverse flange. The
falsely portrays the palatine as having a short entrance into the medial posterior margin of the interpterygoid vacuity is closed on either side by
margin of the internal naris. a short, broad, posteromedially directed basal process of the pterygoid.
Palatine - In ventral view (Figs. 4A, 9, 11) the palatine has an The process has a triangular notch for the reception of the basipterygoid
anteroposterior elongate, subrectangular outline. A channel-like depres- process of the braincase, which is roofed dorsally by the base of the
sion extends along the lateral half of its ventral surface that deepens as it epipterygoid to form a freely mobile basicranial articulation. A similar
extends to the posterior margin of the internal naris. Laterally the channel basicranial articulation has been described in Diadectes and Orobates.
is bounded by a thin, ventrally reflected, lip-like flange that contacts the The transverse flange of the pterygoid (Figs. 4A, 10, 14) is dors
medial surface of the alveolar shelf of the maxilla. Anteriorly the lateral oventrally thickened into a massive structure, particularly toward its
flange of the palatine continues for a short distance as it tapers ventrally posterior, tooth-bearing margin. The bases of a single, transverse row of
to a sharply pointed process. A transverse section of the cranium at the eight or nine large teeth dominate the dentition along the posterior margin
midlength level of the palatines (Fig. 9) reveals a pronounced, dorsal of the transverse flange. Their basal diameters vary considerably, but in
thickening of their lateral margin to form a thick, buttress-like contact general exhibit a pronounced serial increase in size to a mid-serial posi
tion of three teeth of subequal size. The teeth are implanted in deep
with the medial surface of the alveolar shelf. The transverse section also
sockets and are bordered laterally and anteriorly by a narrow field of
199

FIGURE 9. Limnoscelis paludis, holotype, YPM 811. Posterior view of anterior portion of cranium at cross-sectional break through cranium at level of
anterior margins of orbits.

small denticles. As in amniotes, the transverse flange projects ventrolat essentially a transverse vertical plane that is inclined slightly
erally far below the ventral margin of the cranium (Fig. 3B). dorsomedially. Its surface is slightly convex, so as to be smoothly con
The quadrate ramus of the pterygoid (Figs. 4A, 10, 14) extends tinuous with the curvature of the posterior margin of the cheek region of
posterolaterally from the level of the basicranial articulation to the pos the skull roof, where it is overlapped narrowly by the squamosal. The
teromedial edge of the occipital exposure of the quadrate and can be occipital lamina in Tseajaia, Orobates, and Diadectes differs only in that
divided into two distinct components. One forms a thin, tall, the occipital surface is flat and inclined slightly anterodorsally. The me
dorsomedially inclined lamina-like structure that ends posteriorly con dial margin of the occipital lamina curves anteromedially, where it is
tacting the entire medial surface of the dorsal process of the quadrate just continuous with the second, vertical lamina of the dorsal process (Figs.
above its medial condyle. In ventral view all that is visible of this compo 4A, 7, 10, 12), referred to here as the pterygoid lamina. It extends ante
nent is a thickly rounded, keeled ventral edge, which medially borders the riorly and slightly medially and has the form of an expansive, subtriangular,
subtemporal adductor fossa. The second component of the quadrate sheet that forms approximately the posterior half of the medial wall of
ramus forms a prominent, narrowly arcuate flange that projects medially the subtemporal adductor chamber. Proximally the pterygoid lamina prob
and slightly dorsal to the level of the keeled ventral edge of the quadrate ably reached the height of the occipital lamina, but anteriorly its dorsal
ramus. The arcuate flange, present also in Limnoscelis dynatis, is appar margin tapers ventrally to the ventral margin of the subtemporal adduc
ently a unique feature of the limnoscelids among the diadectomorphs. tor chamber and is overlapped medially by approximately the posterior
half of the quadrate ramus of the pterygoid. In amniotes the dorsal
Palatoquadrate Complex process of the quadrate lacks an occipital lamina, consisting only of the
relatively flat, vertical pterygoid lamina and, therefore, does not contrib
Quadrate - The morphology of the quadrate probably represents
the primitive condition for diadectomorphs and possibly all cotylosaurs ute to the posterior wall of the adductor chamber.
in having a large ventral condyle and a dorsal process that consists of two Although the articular surfaces of both condyles are only partially
vertical sheets or laminae that arise from the condyle. One lamina, re exposed due to the jaw being tightly attached to the cranium, their ap
ferred to here as the occipital lamina (Figs. 4B, 10, 12), extends directly proximate shape can be discerned (Figs. 4, 10, 12, 14). As in amniotes,
above the posterior margin of the condyle as a transversely narrow, the condyle is greatly expanded mediolaterally and divided by a well
subrectangular plate that nearly reaches the entire dorsal height of the developed, anteroposteriorly oriented, saddle-like groove into two dis
tinct condylar facets. The axis of the slightly longer medial facet is ori
subtemporal adductor chamber. Within the chamber it forms the medial
half of the posterior wall. The greater portion of the occipital lamina is ented slightly anteromedially, whereas that of the lateral facet is oriented
also exposed on the occipital surface of the chamber, where it occupies strongly anteromedially.
200

FIGURE 10. Limnoscelis paludis, holotype, YPM 811. Posterior portion of cranium in ventral (palatal) view. A, Illustrated as preserved, and B,
emphasizing preserved limits of elements.
201
Just above the medial margin of the medial condyle the quadrate continuous with the smoothly finished lateral walls of the prootic region
and adjacent to the distal end of the quadrate ramus of the pterygoid of the braincase. Posteriorly the cristae end in a thin, ventrolaterally
there is a deep pocket of unfinished bone (Figs. 4B, 12) that presumably convex, sleeve-like structure that partially encloses the fenestra ovalis
received the distal end of the stapes, although this element is unknown in anteriorly. The thin sheet of bone spanning the depressed area between
the cristae extends onto the anterior margin of the ventral surface of the
Limnoscelis paludis.
Epipterygoid - Although rarely exposed in tetrapods, both basioccipital in a strongly interdigitating suture. A low, midventral ridge
epipterygoids are present in Limnoscelis paludis. As in amniotes, where between the cristae is continued posteriorly onto the basioccipital, where
this bone is known, the epipterygoid consists of a large basal plate (Figs. it merges with the raised, rounded, ventral lip of the occipital condyle.
4A, 7, 10) and a dorsal, narrow, vertical, rod–like process. Only the base Posterior and immediately dorsal to the level of the distal free margins of
of the right dorsal process is visible in lateral view in any of the figures the cristae ventrolaterales the opisthotic portion of the braincase is ex
(Fig.7). The epipterygoids are preserved in place and essentially com panded greatly laterally into two distinct structures. Suturally adjoining
plete except for the distal halves of both dorsal processes being sepa the lateral margin of the basioccipital and projecting posteroventrally
rated narrowly by a break and reoriented due to the dorsoventral crusting and slightly laterally from the base of the paroccipital process of the
of the skull so that the right and left distal portions are redirected poste opisthotic is a short, gutter-shaped structure originally termed the otic
riorly and anteriorly, respectively. Despite the jagged margins of the trough by Fracasso (1983, 1987). Because Fracasso (1987, p. 9) was
basal plates, due to small loss of bone, they have an expansive contact unable to detect the suture separating the opisthotic from the basioccipi
with the dorsolaterally facing surface of the quadrate ramus of the ptery tal and exoccipital, he mistakenly described the otic trough as appearing
goid from the level of the basicranial articulation to approximately the “to be continuous with the basioccipital, as there is no evidence of
midlength of the ramus. The basal plate is thickest along its ventrolateral sutures or other separation between the trough bases and the basioccipi
margin in the area of the basicranial articulation, where it roofs the V tal.” The otic trough, with its anterolaterally facing, gutter-like channel
forms essentially an extension of the posteromedial border of the fenes
shaped notch in the basal process of the pterygoid to provide a ventrally
tra ovalis. An otic trough has also been described in Diadectes and
facing articular facet for the basipterygoid process of the braincase (Figs.
4A, 10). The dorsal process of the epipterygoid arises from the medial Limnoscelis dynatis, as well as in some primitive basal synapsids (Berman,
margin of the dorsal surface of the basal plate as a conical structure that 2000). Although an otic trough was not described by Berman et al.
gradually narrows to a slender, rod-like structure. It reaches the ventral (2004) in Orobates, reexamination of one specimen (MNG 8760) has
surface of the parietal to provide a structural brace between skull roof revealed its presence. An otic trough was also tentatively identified in
and palate. A small recess on the ventral surface of the parietal at the level Tseajaia (Berman et al., 1992). In both species of Limnoscelis a distinct
of the posterior orbital margin and lateral to the sphenethmoid marks the cleft separates the otic trough from the base of the paroccipital process.
location of its contact with the skull roof (not visible in figures). In Occiput - The supraoccipital and opisthotic are indistinguishably
previous interpretations of L. paludis (Romer, 1946; Fracasso, 1983, fused into a single complex (Figs. 4B, 12). Romer (1946) and Fracasso
1987) the role of the epipterygoid in the basicranial articulation was not (1983, 1987) misinterpreted symmetrical fractures, most likely due to
the dorsoventral crushing of the skull, as the supraoccipital-opisthotic
recognized.
Braincase sutures extending directly dorsolaterally from the midheight of the lateral
margin of the exoccipital to the lateral margin of the occiput. The
General - The braincase was described and illustrated as pre supraoccipital-opisthotic suture was also misidentified by Berman (2000)
served and reconstructed by Fracasso (1983, 1987). In contrast to the and Berman et al. (1992) as extending dorsolaterally a short distance
present study, he removed portions of the dermal skull that obscured the from the middorsal margin of the foramen magnum that limited the su
braincase in dorsal and much of its lateral views. This was not attempted praoccipital exposure to rather narrow, dorsolaterally arching bands. The
here in order to avoid further damage to the cranium, and a detailed apparent absence of a supraoccipital-opisthotic suture in the holotype
description of the braincase in these views is deferred to Fracasso (1983, of Limnoscelis paludis is best attributed to late adult fusion. The suture
1987). The description of the braincase presented here, therefore, is is, however, easily traced in the isolated otico-occipital portion of the
limited to palatal and occipital views. braincase of L. dynatis, which lacks the exoccipitals and basioccipital.
Palatal view - The fused basisphenoid and parasphenoid, the Here the suture extends dorsolaterally from about midheight of the lat
basiparasphenoid complex, is extensively exposed in ventral view of the eral margin of the foramen magnum in a broadly sigmoid course that ends
braincase (Figs. 4A, 10). As exposed, the narrow cultriform process at the lateral margin of the occiput, where the occiput would have con
extends the entire length of the interpterygoid vacuity, reaching the level tacted the lower portion of the medial margin of the ventromedial occipi
of the posterior border of the internal naris. Its lance-like outline widens tal flange of the tabular. Using the occiput of L. dynatis as a guide, the
quickly but smoothly to a greatest width in the first third of its posterior opisthotic in L. paludis would approximate a shallow depression that
length and then gradually narrows to an acutely angled point. In cross encompasses the ventrolateral region of the occiput and includes the
section at the level of the anterior margins of the orbits it has the form of entire paroccipital process. Its dorsally convex margin ends with the
a narrow V-shaped trough with a well-developed ventral keel (Fig. 9). paroccipital process contacting about the ventral third of the medial
The trough of the posterior portion of the process supports the ventral margin of the ventromedial occipital flange of the tabular. On the basis of
margin of the sphenethmoid (Fig. 7). The stout basipterygoid process the right side of the occiput, the supraoccipital-opisthotic suture would
(Figs. 4A, 10) is triangular in outline with a transverse anterior margin have ended distally near or just above the medial margin of the small,
and is directed ventrolaterally into the notch of the same outline in the rounded notch in the lateral margin of the occiput that is identified here as
basal process of the pterygoid. Only the anterior surface of the the posttemporal fenestra. There is a further deepening of the depression
basipterygoid process exhibits a well-defined, roughened articular sur on the distal portion of the paroccipital process that begins in a step-like,
face, which is transversely oval in outline. medially convex arc and appears to extend dorsally beneath the lip-like,
Posterior to the basipterygoid processes the smoothly finished ventral margin of the ventromedial occipital flange of the tabular. The
basiparasphenoid expands greatly laterally, with the lateral margins be surface of the supraoccipital portion of the occiput is in contrast rela
ing dominated by narrow, well-developed, rounded ridges of the cristae tively flat.
ventrolaterales (Figs. 4, 10). The cristae increase in height as they merge In a reexamination of the otico-occipital portion of the braincase
anteromedially to form a deep, concave posteromedial border between in Limnoscelis dynasties Berman (2000) described a suture that divides
them. The lateral margins of the cristae extend dorsally for a short dis the supraoccipital into lateral halves. It extends from the middorsal mar
tance to a narrow zone of unfinished bone that separates them from being gin of the foramen magnum to the dorsal margin of the occiput and is
202

FIGURE 11. Limnoscelis paludis, holotype, YPM 811. Anterior portion of cranium in ventral (palatal) view. A, Illustrated as preserved, and B,
emphasizing preserved limits of elements.
203

FIGURE 12. Limnoscelis paludis, holotype, YPM 811. Cranium in occipital view. A, Illustrated as preserved, and B, emphasizing preserved limits of
elements.

traceable also on the interior, posterior wall of the cranial cavity. This the large jugular foramen, presumably for cranial nerves IX-XI and the
was compared to a similar division of the supraoccipital in two jugular vein, on the exoccipital-opisthotic suture at the ventral, lateral
anthracosaurian amphibians. There is no evidence in either species of margin of the dorsal, wing-like process of the exoccipital (Figs. 4, 10, 12).
Limnoscelis of a synotic bone or sutural evidence of its former presence,
which Fracasso (1987, p. 11) described in L. paludis by as “a small Mandible
median wedge-shaped element situated posterodorsally between the otic Dentary - The dentary occupies a large proportion of the lateral
capsules…tightly apposed against the overlying anteroventral surface of surface of the jaw, extending over more than two-thirds of its anterior
the supraoccipital.” length (Figs. 3C, 13). Most noticeable, at the symphysis the lateral
The basioccipital and exoccipitals (Figs. 4, 12) are indistinguish
surface of the dentary is swollen into a massive, hemispherical structure,
ably fused into a single complex in Limnoscelis paludis. The exoccipitals undoubtedly to accommodate the deep, large roots of the anterior three
undoubtedly did not consist, as Fracasso (1983, 1987) described, solely teeth. On the dorsal surface of the symphysis and medial to the three
of small, triangular, wing-like elements restricted to the lateral margins of anteriormost teeth the dentaries form a large, flat, rugose-surfaced plate
the foramen magnum. Rather, they almost certainly included, as is stan
(Figs. 15A, B), which may have provided additional reinforcement to the
dard in tetrapods and illustrated here, a short, ventral, column-like struc symphysis. Its midline suture is irregular, so that the left dentary makes
ture, which, along with the dorsal, wing-like process, contacted the pos a significantly greater contribution to the plate. In lateral view the dentary
terior surface of the supraoccipital-opisthotic in forming the lateral walls ends posteriorly in typical cotylosaurian pattern, in an oblique contact
of the foramen magnum. Furthermore, expanded bases of the exoccipitals with the angular and surangular as it narrows posterodorsally in a sharp,
contributed undoubtedly to the dorsolateral portions of the occipital
wedge-shaped process that rises very slightly as it extends onto the base
condyle and met or almost met on the floor of the foramen. Fracasso of the low coronoid eminence of the surangular. In medial view (Fig. 14)
(1983, 1987) was unable to identify the jugular foramen and mistook an the dentary is nearly entirely covered by other elements except for a
extremely small pocket of matrix at the ventrolateral margin of the dorsal, short anterior portion of the alveolar shelf.
wing-like process of the left exoccipital as the foramen for the hypoglo Transverse breaks of the right and left mandibles at the approxi
ssal nerve XII. Additional preparation has located both these foramina in mate levels of the sixth and tenth tooth positions (Fig. 15C, D), respec
their normal positions, with the hypoglossal foramen on the lateral sur
tively, reveal a marked reduction in the transverse thickness of the dentary
face of the neck of the occipital condyle (not visible in the figures) and posteriorly that is accompanied by a corresponding expansion the
204
205
206
Meckelian canal. Concomitantly, there is a ventral expansion of the the dorsal margin of the prearticular except anteriorly, where the contact
dentary from a level just below the Meckelian canal to the ventral midline is continued a short distance along the dorsal margin of the splenial. A
of the mandible that gradually thickens to equal that of its more anterior shagreen of denticles increases in density and width posteriorly as it
cross-section. The ventral expansion, however, is excluded from the extends across the posterior half of the anterior coronoid and the poste
lateral wall of the expanded Meckelian canal by a thin, sheet-like flange rior coronoid anterior to the adductor fossa. At the anterior margin of the
of the angular. As in the maxilla, the lateral margin of the alveolar shelf of adductor fossa a thin, wing-like extension of the posterior coronoid ex
the dentary is noticeably higher than the medial margin, creating a super tends along the dorsal margin of the medial surface of the surangular,
ficially pleurodont-like implantation of the teeth, particularly anteriorly nearly reaching the summit of a low coronoid eminence of the surangular.
in the series, and the teeth are set in very deep sockets (Fig. 15C, D). Contrary to Fracasso’s (1983, p. 194) description, there is no indication
There are places for 23 dentary teeth. The second and third teeth greatly of an expansion of the anterior end of the anterior coronoid to form a
dominate the series, with the second being slightly reduced. The first “lance-like configuration” that nearly reaches the mandibular symphy
tooth is reduced considerably compared to the second and third, but is sis. This was likely confused with the large, flat, rugose-surfaced plate of
still slightly larger, especially in its diameter, than the largest remaining bone formed by the dentaries on the dorsal surface of the symphysis.
teeth of the series, which decrease gradually in size posteriorly. The As in Limnoscelis paludis, Tseajaia possesses two coronoids,
teeth are identical in structure to those of the upper marginal series, with the posteriormost one contacting the dorsal margin of the medial
which is best exemplified by the five teeth posterior to the canine-like surface of the surangular in a low cornonoid eminence at the anterolateral
third and fourth of the left mandible (Fig. 13). Crenulations at the base of border of the adductor fossa. In contrast to L. paludis the diadectids
the roots are indicative of a labyrinthine enfolding (Figs. 13, 14-15). Desmatodon, Orobates, and Diadectes have a single coronoid, which
Splenial – In medial view of the mandible (Figs. 3C, 14) the forms a high, well-developed, prominent coronoid eminence at the ante
splenial extends posteriorly from its contact with its mate at the rolateral border of the adductor fossa that is visible in lateral and medial
posteroventral portion of the symphysis to about midlength of the man view.
dible. In this course the splenial is restricted to a very narrow, short In contrast to Fracasso’s (1983) description of the medial fenestra
exposure along the ventrolateral margin of the dentary posterior to its as an ovate opening that narrows anteriorly to an acute angle, here it is
symphyseal swelling (Figs. 3C, 13). A deep pocket in the posterior interpreted as having a narrow, near-elliptical outline that narrows only
surface of the symphysis is floored by the union of transversely narrow, slightly anteriorly. Tseajaia, Orobates, Desmatodon, and Diadectes also
shelf-like structures of the splenials (Fig. 15A, B). For a short distance possess a medial fenestra of the same size, general proportions, and
posterior to the symphyseal region on the medial surface of the mandible relationship to bordering elements as in Limnoscelis paludis. Otherwise,
the splenial expands quickly into a tall, flat, vertical plate that contacts in cotylosaurs a medial fenestra is absent with the possible exception of
dorsally the ventral half of the medial surface of the alveolar shelf of the the ophiacodonts, basal synapsids Ophiacodon and Varanosaurus, where
dentary (Fig. 15C). Farther posteriorly the splenial bifurcates around the a very small, possibly equivalent structure is present at the angular
anterior end of the medial fenestra (identification of this opening follows prearticular suture near its anterior juncture with the splenial.
Welles, 1941, rather than Fracasso’s (1983, 1987) identification of an Prearticular - The prearticular is visible only in medial view of
inframandibular fenestra) into two, narrow, strap-like posterior pro the mandible (Figs. 3C, 14). For most of its anterior length it has basi
cesses of very unequal lengths (Figs. 3C, 14). The much longer ventral cally the outline of a vertically narrow, anteroposteriorly elongated plate
process extends nearly the entire length of the medial fenestra, medially that extends from the posterior margin of the mandible to nearly the level
overlapping the angular beneath the fenestra, as it narrows slightly to end of the anterior margin of the medial fenestra. In this course the prearticular
in a bluntly pointed margin. The much shorter dorsal, posterior process can be divided into three regions of anteriorly decreasing vertical widths.
of the splenial extends between the ventral and dorsal margins of the The prearticular is most greatly expanded posterior to the adductor
anterior ends of the anterior coronoid and medial fenestra, respectively, fossa, where it contacts the medial surface of the articular a short dis
as it overlaps medially the anterior end of the prearticular. tance below the glenoid facet and ventrally overlaps the medial surface of
The cross-sectional breaks of the right and left mandibles at the the dorsal margin of the angular. At its posterior margin the prearticular
levels of the sixth and tenth tooth positions, respectively, reveal a dra joins with the articular in forming the ventral, triangular portion of a very
matic restructuring of the splenials (Fig. 15C, D). Anteriorly the cross large, vertical, anteriorly inflected, shelf-like structure of the pterygoideus
sectional area of that portion of the splenial exposed on the ventrolater or medial process for presumably the attachment of the pterygoideus
ally surface of the mandible has a massive, rectangular structure, which is musculature. The posterior margin of the process is angled anteroventrally,
excluded from the ventral margin of the Meckelian canal by the angular. so as to intersect the vertical, anterior margin ventrally in a blunt point.
In contrast, that portion of the splenial exposed on the medial surface of At its posterior margin the process is massive, but is quickly reduced to
the mandible appears as a vertical flange that bounds the Meckelian canal a thin edge along its anterior margin, where it is separated from the medial
as it extends dorsally to a narrow contact at the midheight level of the face of the mandible by about 3 cm, forming a deep, anteriorly facing,
medial surface of the alveolar shelf. Farther posteriorly, at the level of the smoothly finished, convexity. The free anterior edge of the process is
tenth tooth of the cross-sectional break of the left mandible the thick, roughly finished. A low ridge near the ventral margin of the medial surface
ventrolateral portion of the splenial is replaced by a ventral expansion of of the expanded, posterior, plate-like portion of the prearticular extends
the dentary, resulting in a greatly narrowed contact of the splenial with across its contribution to the inside curvature of the pterygoideus pro
the dentary at the ventral midline of the mandible. The vertical, medial cess to its ventral tip. In Tseajaia there is a low but prominent, medially
flange of the splenial, on the other hand, has become thickened, reori projecting ridge at the posteromedial corner of the prearticular that ex
ented dorsomedially, greatly shortened dorsally to the ventral margin of tends anteroventrally on the posterior portion of the prearticular’s con
the medial fenestra, and excluded from the Meckelian canal by the angu tact with the medial surface of the articular and adjacent to the posterior
lar. rim of the adductor fossa. It is apparently a greatly reduced equivalent of
Coronoids - Two coronoids are present, referred to here as the the pterygoideus process in Limnoscelis paludis, presumably serving
anterior (c1) and posterior (c2) coronoids, and exposed only in medial the same function, whereas Desmatodon, Orobates, and Diadectes lack
view of the mandible (Figs. 3C, 14). The anterior coronoid and the ante any indication of the process.
rior half of the posterior coronoid form a narrow, rectangular strip that The mid-region of the prearticular is marked by a posterior, step
widens slightly as it extends posteriorly in contact with the dorsal half of like narrowing of its dorsal and ventral margins that extends the length of
the medial surface of the alveolar shelf of the dentary (Fig. 15D). The the adductor fossa, as it forms the upper half of the lateral wall of the
ventral margin of the coronoids narrowly contact the medial surface of adductor chamber and ventrally overlaps the medial surface of the dorsal
207

FIGURE 15. Limnoscelis paludis, holotype, YPM 811. Symphyseal portion of lower jaw in dorsal view. A, Illustrated as preserved, and B, emphasizing
preserved limits of elements. Cross-sectional, posterior views of C, right mandible at level of sixth tooth, and D, left mandible at level of tenth tooth.
208
margin of the angular. A further narrowing of the prearticular anterior to of the adductor fossa. It is most extensively exposed in medial view of
the adductor fossa defines its anterior region. It borders the dorsal margin the mandible (Fig. 14), where its posterior margin forms the upper por
of the medial fenestra except for the short distance at its anterior end, tion of the large, anteriorly inflected, flange-like pterygoideus process.
where it is overlapped medially by the splenial. At the cross-sectional Projecting posteriorly from the upper margin of the posterior surface of
break of the left mandible at the level of the tenth dentary tooth position the pterygoideous process is a short, rounded, rugose knob that is prob
(Fig. 15D) the medially exposed prearticular contacts the ventral half of ably equivalent to a retroarticular process. An anterior process extends
the medial surface of the alveolar shelf of the dentary. However, at the from the anteromedial edge of the articular body. It has the form of a
cross-sectional break of the right mandible at the level of the sixth dentary vertical, transversely thin flange that is partially exposed above the ex
tooth position (Fig. 15C) the depth of the prearticular is greatly reduced. panded posterior portion of the prearticular that contacts its medial
Here it narrowly contacts on the ventral half of the medial margin of the surface. Farther anteriorly it is visible only in dorsal view of the man
alveolar shelf and is overlapped medially by the splenial. dible, as it sheaths the lateral surface of the prearticular to approximately
Angular - The angular is extensively exposed on the lateral and the midlength of the medial wall of the adductor fossa. In Desmatodon
medial surfaces of the posterior half of the mandible (Figs. 3C, 13-14). the anterior process of the articular extends along the lateral surface of
Here it forms the lateral and medial walls of the ventral half of the the prearticular to at least level of the posterior third of the medial
adductor chamber, which converge ventrally to the broadly convex, foramen, as indicated by its very narrow exposure along the posterior
posteroventral angle of the mandible in a sharply angled, V-shaped, keel dorsal border of the foramen, whereas in Diadectes the process extends
like structure. The keeled edge extends to within a short distance from to the level of the anterior end of the foramen.
the posterior end of the mandible. Here only the laterally exposed por Unfortunately, the articular glenoid surface of both articulars are
tion of the angular continues posteriorly, narrowing to an acute angle at mostly hidden from view by their tight articulation with the quadrate
the posterodorsal corner of the mandible as its posterior margin curves condyles (Fig. 14), but it is possible to discern that the glenoid is divided
medially to contact the lateral margin of the posterior surface of the into two, anteroposteriorly elongated concave surfaces that are sepa
pterygoideus process. Almost the entire dorsal margin of the laterally rated by a ridge of approximately the same orientation.
exposed portion of the angular contacts the surangular in a broadly con
vex suture. Anteriorly the angular has a sinuous contact with the dentary SUMMARY
that quickly angles anteroventrally to the ventral margin of the mandible, The above description, in addition to expanding on the most re
where it continues a short distance as a narrow, splint-like exposure
cent detailed descriptions of the holotypic skull of Limnoscelis paludis
between the ventral margins of the dentary and splenial. The lateral
presented by Fracasso (1983, 1987) since that of Romer’s (1946), also
surface of the angular bears a fairly rugose sculpturing that extends to the recognizes the following major errors:
ventral edge of the mandible. 1) the dorsal processes of the premaxillae are incorrectly described
In medial view of the mandible the angular extends anteriorly to
as short and wide and as enclosing an internasal bone, rather than as long,
the posteroventral margin of the medial fenestra. Here its medial surface
narrow, united processes that extend posterodorsally along the midline
is overlapped by the distal end of the ventral, posterior process of the in the absence of an internasal;
splenial, limiting the exposure of the angular to two, short, splint-like 2) failed to recognize presence of palatal vomerine processes of
processes: one extending along the ventral margin of the fenestra and the the premaxillae, which extend posteriorly as long, narrow processes on
other, as described above, along the ventral margin of the mandible.
either sided of the midline of the anterior portion of the palate;
Farther anteriorly the angular looses its superficial exposure as it contin 3) misinterpreted the jugal as lacking a the dorsal orbital process,
ues within the Meckelian canal to at least a short distance beyond the which is present as a long, narrow process that probably formed nearly
level of the anterior margin of medial fenestra. The cross-sectional break the ventral half of the posterior orbital margin;
of the left mandible at the level of the tenth marginal tooth (Fig. 15D)
4) eave-like structural contact of the squamosal with the postor
exposes an extensive, well-developed angular that is divisible into two bital and supratemporal was misinterpreted as a “loose junction” be
portions. A narrow, rectangular portion is slightly imbedded into the tween the cheek and skull table, rather than as a complex, sutural contact
internal surface of the splenial. It has a broad, midsagittal contact with that was surely immobile;
the dentary that is separated from the ventral margin of the mandible by
5) misinterpreted the occipital-flanges of the supratemporal and
a narrow dentary-splenial contact. The second portion of the angular tabular and the exposure of the squamosal on the occipital surface of the
extends dorsally from the ventromedial surface of the internal surface of cheek as parts of a long, ventral extension of the tabular occipital surface
the first portion as a thin, flange-like extension that sheaths the ventral
of the skull cheek that nearly reached its ventral margin;
half of the internal surface of the dentary. Thus, at this level of the
6) failed to recognize the midline palatal suture between the vomer
mandible the angular bounds the ventrolateral and the ventral, medial half and premaxilla and misinterpreted maxillary process of the premaxilla as
of the Meckelian canal. At the cross-sectional break of the right mandible
a lateral, wing-like extension of the vomer;
at the level of the sixth marginal tooth (Fig. 15C) all that remains of the
7) misinterpreted the vomer-palatine suture as nearly straight,
angular is a greatly reduced, rectangular portion, which floors the greatly directed slightly anterolaterally, and positioned slightly anterior to the
constricted Meckelian canal and retains much reduced contacts with the posterior margin of the internal naris and, therefore, falsely portrayed
splenial and dentary. the palatine as having a short entrance into the medial margin of the
Surangular - Exposures of the surangulars on the medial and internal naris. Rather, the posterior margin of the vomer has a V-shaped,
lateral surfaces (Figs. 3C, 13-14) of the mandibles are partially hidden by marginal contact with the anterior margin of the palatine at a level slightly
the close apposition of the mandibles to the cranium. The surangular
posterior to the posterior margin of the internal naris and, therefore, the
forms the lateral wall of the adductor fossa, thickening dorsally to a palatine does not enter the internal naris;
smoothly rounded, free dorsal margin that is elevated slightly into a 8) did not recognize the moderately deep, anteroposteriorly elon
broadly convex coronoid eminence. In lateral view the surangular has gate, narrow, basin-like depression on the ventral surface of the palate
roughly a diamond outline, with the anterior portion wedging between that straddles the contact of the medial margin of the ectopterygoid with
the dentary and angular, whereas posteriorly it narrows into a rather the pterygoid;
attenuated process that curves slightly medially as it contacts the articu 9) failed to recognize the participation of the epipterygoid in the
lar. basicranial articulation, which dorsally roofs a V-shaped notch in the
Articular - The articular (Figs. 3C, 14) forms the posterior corner
basal process of the pterygoid to provide a ventrally facing articular facet
of the mandible, and its concave anterior surface forms the posterior wall for the basipterygoid process of the braincase;
209
10) failed to identify the presence of the suture marking the con dorsal, wing-like portion of the exoccipital and the latter on the lateral
tact between the opisthotic and the basioccipital-exoccipital complex surface of the neck of the occipital condyle;
and mistakenly described the otic trough as a structure of the basioccipi 15) misidentified the presence of a “lance-like configuration” of
tal rather than the opisthotic; the anterior end of the anterior coronoid as nearly reaching the mandibu
11) misinterpreted paired symmetrical fractures on the posterior lar symphysis, which likely was confused with a single, large, flat, rug
surface of the occiput (most likely due to the dorsoventral crushing of ose-surfaced plate of bone formed by dentaries on the dorsal surface of
the skull) as the supraoccipital-opisthotic sutures, which are absent due the symphysis;
to probable late adult fusion; 16) misinterpreted the medial fenestra of the lower jaw as having
12) mistakenly identified the presence a synotic bone of the brain an ovate outline that narrows anteriorly to an acute angle, rather than a
case for which there is also no sutural evidence of its former presence; narrow, near-elliptical outline that narrows only slightly anteriorly.
13) misidentifed the exoccipitals as consisting solely of small,
triangular, wing-like element restricted to the lateral margin of the fora ACKNOWLEDGMENTS
men magnum. They almost certainly included also the presence of a The authors wish to express their thanks to Yale Peabody Mu
short, ventral, column-like structure that extended between the dorsal, seum for the extended loan of the holotype of Limnoscelis paludis. We
wing-like portion and an expanded base that contributed to the dorsolat are also greatly indebted to Mark A. Klingler of the Carnegie Museum of
eral portions of the occipital condyle, with the paired elements forming Natural History for his exacting, full skeletal reconstructions and
the lateral walls and at least the greater portion of the floor of the foramen
whole-body restoration of L. paludis and expertise in composing and
magnum; formatting the figures. Special gratitude is extended to the reviewers, S. S.
14) did not recognize the presence of the jugular foramen and
Sumida (California State University, San Bernardino) and Robert Sullivan
misidentified the location of the foramen for the hypoglossal nerve XII. (The State Museum of Pennsylvania, Harrisburg), whose careful, con
Both foramina are present in their standard locations, with the former on
structive comments and criticisms had considerable impact on the im
the exoccipital-opisthotic contact at the ventral, lateral margin of the
provement of this paper.

REFERENCES

Berman, D.S, 2000, Origin and early evolution of the amniote occiput: Laurin, M., and Reisz, R.R., 1995, A reevaluation of the amniote phylog
Journal of Paleontology, v. 74, p. 938-956. eny: Zoological Journal of the Linnean Society, v.113, p. 165-223.
Berman, D. S, and Sumida, S.S., 1990, A new species of Limnoscelis (Am Laurin, M., and Reisz, R.R., 1997, A new perspective on tetrapod phylog
phibia, Diadectomorpha) from the Late Pennsylvanian Sangre de Cristo eny, in Sumida, S.S., and Martin, K.L.M., eds., Amniote Origins, Com
Formation of central Colorado: Annals of Carnegie Museum, v. 59, p. pleting the Transition to Land: Academic Press, San Diego, p. 9-59.
303-341. Lee, M.Y.S., and Spencer, P.S., 1997, Crown-clades, key characters and
Berman, D.S, Sumida, S.S., and Lombard, R.E., 1992, Reinterpretation of taxonomic stability: when is an Amniote not an Amniote? in Sumida,
the temporal and occipital regions in Diadectes and the relationships of S.S., and Martin, K.L.M., eds., Amniote Origins, Completing the Tran
diadectomorphs: Journal of Paleontology, v. 66, p. 481-499. sition to Land: Academic Press, San Diego, p. 61-84,
Berman, D.S, and Sumida, S.S., 1995, New cranial material of the rare Moss, J.L., 1972, The morphology and phylogenetic relationships of the
diadectid Desmatodon hesperis (Diadectomorpha) from the Late Penn Lower Permian tetrapod Tseajaia campi Vaughn (Amphibia:
sylvanian of central Colorado: Annals of Carnegie Museum, v. 64, p. Seymouriamorpha). University of California Publications in Geological
315-336. Sciences, v. 98, p. 1-72.
Berman, D.S, Reisz, R.R., Bolt, J.R., and Scott, D., 1995, The cranial Olson, E.C., 1947, The family Diadectidae and its bearing on the classifica
anatomy and relationships of the synapsid Varanosaurus (Eupelycosauria: tion of the reptiles: Fieldiana, Geology, v. 7, p. 2-53.
Ophiacodontidae) from the Early Permian of Texas and Oklahoma: Panchen, A.L., 1972, The interrelationships of the earliest tetrapods, in
Annals of Carnegie Museum, v. 64, p. 99-133. Joysey, K.A., and Kemp, T.S., eds., Studies in Vertebrate Evolution:
(Diadectomorpha:
Berman, Diadectidae)
D.S, Sumida, from Martens,
S.S., and the Early Permian of central
T., 1998, Ger
Diadectes Winchester Press, New York, p. 65-87,
Reisz, R.R., 2007, The cranial anatomy of basal diadectomorphs and the
many, with description of a new species: Annals of Carnegie Museum, v. origin of amniotes, in Anderson, J.S., and Sues, H-D., eds., Major Tran
67, p. 53-93. sitions in Vertebrate Evolution: Indiana University Press, Bloomington
Berman, D.S, Henrici, A.C., Kissel, R.A., Sumida, S.S., and Martens, T., and Indianapolis, p. 228-252.
2004, A new diadectid (Diadectomorpha), Orobates pabsti, from the Romer, A.S., 1946, The primitive reptile Limnoscelis restudied: American
Early Permian of central Germany: Bulletin of Carnegie Museum of Journal of Science, v. 244, p. 149-188.
Natural History, v. 35, p. 1-36. Romer, A.S., and Price, L. I., 1940, Review of the Pelycosauria: Geological
Fracasso, M.A., 1980, Age of the Permo-Carboniferous Cutler Formation Society of America Special Paper 28, 538 p.
vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale Vaughn, P.P., 1964, Vertebrates from the Organ Rock Shale of the Cutler
ontology, v. 41, p. 1256-1261. Group, Permian of Monument Valley and vicinity, Utah and Arizona:
Fracasso, M.A., 1983, Cranial osteology, functional morphology, system Journal of Paleontology, v, 38, p. 567-583.
atics of Limnoscelis paludis Williston [Ph.D. dissertation]: Yale Uni Vaughn, P.P., 1969, Upper Pennsylvanian vertebrates from the Sangre de
versity, New Haven, Connecticut, 624 p. Cristo Formation of central Colorado: Contributions in Science, Los
Fracasso, M.A., 1987, Braincase of Limnoscelis paludis Williston: Postilla, Angeles County Museum of Natural History, v. 164, 1-28 pp.
no. 201, p. 1-22. Vaughn, P.P., 1972, More vertebrates, including a microsaur, from the
Kemp, T.S., 1982, Mammal-like reptiles and the origin of mammals: Aca Upper Pennsylvanian of central Colorado: Contributions in Science,
demic Press, London, 363 p. Los Angeles County Museum of Natural History, v. 223, p. 1-30.
Kissel, R.A., and Reisz, R.R., 2004, Ambedus pusillus, new genus, new Welles, S., 1941, The mandible of a diadectid cotylosaur: University of
species, a small diadectid (Tetrapoda: Diadectomorpha) from the Lower California Publications in Geological Sciences, v. 25, p. 423-432.
Permian of Ohio, with a consideration of diadectomorph phylogeny: Wideman, N.K., 2002, The postcranial skeleton of the late Paleozoic
Annals of Carnegie Museum, v. 73, p. 197-212. Family Limnoscelidae and its significance for diadectomorph taxonomy:
210
Journal of Paleontology, v. 22, p. 119A. monotremes and armadillos: Zoological Journal of the Linnean Society,
Wideman, N.K., Sumida, S.S., and O’Neil, M., 2005, A reassessment of the v. 98, p. 203-228.
taxonomic status of the materials assigned to the Early Permian tetra Williston, S.W., 1911a, A new family of reptiles from the Permian of New
pod genera Limnosceloides and Limnoscelops, in, Lucas, S.G. and Zeigler, Mexico: American Journal of Science, v. 31, p. 378-398.
K.E., eds., The Nonmarine Permian: New Mexico Museum of Natural Williston, S.W., 1911b, American Permian Vertebrates: Chicago University
History and Science Bulletin, 30, p. 358-362. Press, 145 p.
Wible, J.R., Miao, D., and Hopson, J.A., 1990, The septomaxillae of fossil Williston, S.W., 1912, Restoration of Limnoscelis, a cotylosaur reptile
and Recent synapsids and the problem of the septomaxillae of from New Mexico: American Journal of Science, v. 34, p. 457-468.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

211
REDESCRIPTION OF THE POSTCRANIAL SKELETON OF LIMNOSCELIS
PALUDIS WILLISTON (DIADECTOMORPHA: LIMNOSCELIDAE) FROM
THE UPPER PENNSYLVANIAN OFELCOBRE CANYON, NORTHERN NEWMEXICO

NATALIAK. KENNEDY
Department of Ecology and Evolutionary Biology, University of California Los Angeles, 621 East
Charles Young Drive, Los Angeles, California 90024, U.S.A.; natalia.k.kennedy@ucla.edu

Abstract—The Late Pennsylvanian diadectomorph tetrapod Limnoscelis is known from two species, the most
complete of which is L. paludis from El Cobre Canyon, Rio Arriba County, New Mexico. The most complete
specimen of L. paludis is Yale Peabody Museum specimen 811, which is redescribed here. Distinctive features of
the axial skeleton include an atlas-axis complex demonstrating the characteristic anteriorly-directed processes of
the atlantal and axial intercentra, and alternation of the neural spine height in certain regions of presacral vertebrae.
There is a single coracoid plate, and the scapula is robust; however, the earlier suggestion of the presence of a
suprascapular cartilage is questionable. The humerus is extremely robust with heavily constructed condyles and
muscular attachment processes. The radius and ulna are relatively short, measuring approximately 60 and 85% the
length of the humerus, respectively. The ilium demonstrates the well-developed lateral iliac shelf characteristic of
diadectomorphs, and the anteroventral margin of the acetabulum extends anteriorly in a narrow channel that
reaches toward the anterior border of the pubis. The femur is notable in its prominently developed rugosities for
the puboischiofemoralis internus and ischiotrochantericus muscles. The tibia and fibula range from 77 to 80% of
the length of the femur. Both manus and pes are incomplete in this specimen of Limnoscelis.

INTRODUCTION enemies (Williston, 1912). Romer (1946) redescribed Limnoscelis, con


The Late Pennsylvanian Limoscelidae is a pivotal taxon in under cluding that it was indeed a reptile, but that its anatomy was so general
standing the origin and early evolution of amniotes (e.g., Williston, 1911a,b, ized that it could be “regarded as representing the common stem of all
lines of reptilian descent.”
1912; Romer, 1966; Berman et al., 1992; Sumida et al., 1992; Sumida,
1997). It is part of the more inclusive tetrapod grouping, the Diadecto Limnoscelis dynatis was described from the Upper Pennsylva
morpha, which is generally considered to be the sister taxon to Amniota, nian of Fremont County in central Colorado (Berman and Sumida, 1990).
as traditionally defined (Heaton, 1980), but has also been suggested to be It was distinguished from L. paludis by a few differences in cranial
a basal member of Amniota (Berman et al., 1992; Lee and Spencer, 1997). elements, and Berman and Sumida (1990) suggested that L. paludis is
The Diadectomorpha is composed of three families: Diadectidae, more derived than L. dynatis. Although Berman and Sumida assigned L.
Tseajaiidae and Limnoscelidae. The latter two families have a strictly dynatis to Diadectomorpha, they placed it in the Class Amphibia.
North American record, whereas the family Diadectidae is found both in Here the holotypic postcranial skeleton of Limnoscelis paludis
North America and central Europe (Berman et al., 1997). Significantly, (YPM 811) is fully described for the first time, with dorsal and ventral
the limnoscelids have been suggested as basal members of the grouping, illustrations of the skeleton.
making them a reasonable model for the skeletal structure characteristic Institutional Abbreviations: CM, Carnegie Museum of Natural
of extremely primitive amniotes. History, Pittsburgh, Pennsylvania; FMNH, Field Museum of Natural
Although Limnoscelis is vital to the studies of early amniote ori History, Chicago, Illinois; MCZ, Museum of Comparative Zoology,
gins and evolution, the greater portion of the limnoscelid postcranial Harvard University, Cambridge, Massachusetts; YPM , Yale Peabody
skeleton has never been fully described or illustrated. Previous studies Museum, New Haven, Connecticut.
have focused on the skull (Fracasso, 1983, 1987; Berman et al., 1992; Anatomical Abbreviations: cor, coracoid; delt, deltoid process;
Berman, 2000), the atlas-axis complex (Sumida et al., 1992), and the ect, ectepicondyle; ent, entepicondyle; ent for, entepicondylar fora
vertebral column (Sumida, 1990). Because the Middle Pennsylvanian men; f, femur; fi, fibula; gl for, glenoid foramen; h, humerus; icl, inter
“limnoscelid” Limnostygis has been proposed as a combination of clavicle; int, intermedium; mc, medial centrale; ole, olecranon process;
ophiacodontid and captorhinid specimens (Wideman and Sumida, 2004), r,radius; rade, radiale; sup, supinator process; t, tibia; u, ulna; ule,
and Limnoscelis is the only well-defined member of the Limnoscelidae ulnare; I-V, digits 1-5.
(Wideman et al., 2005), L. paludis is included in this study, and com MATERIALS AND METHODS
pared to L. dynatis when differences are present.
Limnoscelis paludis was first described by Williston (1911a, b) The description of the postcranial skeleton of Limnoscelis is pri
based mainly on two specimens: the nearly complete, articulated skel marily based on the almost complete articulated holotype of L. paludis
eton of L. paludis (the holotype, YPM 811) and a nearly complete (YPM 811). Two other specimens of Limnoscelis exist (MCZ 1947 and
skeleton with some articulated portions (MCZ 1948, formerly YPM 1948, formerly YPM 819 and 809, respectively). The pes of MCZ 1948
809). Unfortunately, he illustrated only representative parts of the skel was used to supplement this description of Limnoscelis, because it is not
eton instead of the skeleton as a whole, and included only a brief descrip complete in the holotype. I could not study the remainder of MCZ 1948
tion. At the time, Williston classified Limnoscelidae in the Class Reptilia. or MCZ 1947, and would like to do so in the future. These materials
The following year, Williston (1912) expanded his study of L. paludis, were compared to the disarticulated postcranial materials of L. dynatis
reconstructing it as a slow, crawling reptile with a long body and long tail (CM 47653), and differences were noted.
that hunted slow-moving prey. He suggested that L. paludis probably Length, height and width measurements were made on all available
lived a semi-terrestrial lifestyle in marshes and hid in the water from its skeletal elements. The processes and depressions were measured, as far
212
as possible, at a right angle to the sagittal and dorsoventral axes. Speci
precluding ventral exposure of that element in the axial column. Most
mens were measured in as many ways as possible as determined by the notably, both the atlantal and axial intercentra bear ventromedial pro
element and its preservation. In addition, photographs were also taken. cesses that are directed anteriorly into furrows of the exoccipital and
The holotypic postcranial skeleton of Limnoscelis paludis is well atlantal intercentrum, respectively. There is little or no development of
preserved and nearly complete; however, it is partially encased in plaster an atlantal neural spine from the paired atlantal neural arches; however,
in the dorsoventral midsection, making access to some of the dorsal and the axial neural arch is robust and very well developed. A proatlas was
ventral aspects of the specimen difficult, and sometimes impossible present.
(Figs. 1,4,5). Much of the skeleton is visible in dorsal view, with the Dorsal Vertebrae and Ribs: Williston (1911a, b, 1912) first
exception of most of the ribs, portions of the pectoral and pelvic girdles described the vertebrae of Limnoscelis paludis (YPM 811) as being rather
and the vertebrae. The ventral anterior portion has been prepared, expos uniform, with swollen neural arches and neural spines all of about the
ing the interclavicle and the ventral surfaces of some of the vertebrae and
same length. Additional descriptions of L. paludis and the disarticulated
ribs and a portion of the ventral surface of the pelvis are visible. Unfor
vertebrae of L. dynatis (CM47653; Berman and Sumida, 1990; Sumida,
tunately, at some point in time, the plaster supporting the holotype of L. 1990; Sumida et al., 1992) showed that the size and shape of the verte
paludis was cracked, and this crack was repaired with plaster (Fig. 4), brae of limnoscelids vary significantly throughout the column.
obscuring some of the ventral skeleton from view. Furthermore, none of Limnoscelis paludis (YPM 811) has 26 presacral vertebrae. The
the ventral surfaces of the limb bones are visible in YPM 811. For the third to seventh dorsal vertebrae occupy the level between the scapular
sake of completeness, the only other valid limnoscelid, L. dynatis, was blades of the pectoral girdle (Fig. 2). The centra are amphicoelous and
used for comparison. notochordal. The centra of the most anterior vertebrae are small in diam
Specimen and locality data for the specimens examined include: eter compared to more posterior vertebrae and are approximately 25%
YPM 811, holotype of Limnoscelis paludis, complete skeleton; Upper
longer than they are wide (Sumida, 1990). The planes of the anterior and
Pennsylvanian El Cobre Canyon Formation, Cutler Group, El Cobre posterior zygapophyses slope anteromedially at about 30° and 25°,
Canyon, Rio Arriba County, northern New Mexico. Collected by David respectively. The neural spines vary in height between tall and short
Baldwin. CM 47653, holotype of L. dynatis, partial associated skeleton
spines in a random pattern. The neural spines in presacral vertebrae 3-7
of a single individual including premaxilla, maxilla, partial occiput, atlantal are the least expanded of the presacral column, but have a greater ante
neural arch, dorsal vertebrae, ribs, partial pectoral and pelvic limbs. Up rior-posterior measure. The third and fifth neural spines are well-devel
per Pennsylvanian Sangre de Cristo Formation, Fremont County, central oped with bases about 9 mm in transverse diameter. Usually, a fore and
Colorado. Collected by Peter P.Vaughn.
aft midline ridge is present on the ends of these tall spines. On the other
AXIAL SKELETON hand, the fourth neural spine is narrow and low with a longitudinal ridge.
Furrows on either side of the neural spine may have facilitated passage of
Atlas-axis Complex: The atlas-axis complex of Limnoscelis interspinus musculature (Sumida, 1990, 1997). Isolated anterior verte
paludis was first described by Williston (1911a, b). However, due to brae in L. dynatis confirm this pattern (Fig. 3). The neural spine on the
incomplete preparation of the specimen (YPM 811), only the dorsal sixth vertebra of the holotype is tall, measuring approximately 16.7 mm
aspect of the complex was described. Sumida (1990) redescribed the in height and 11 mm in transverse basal diameter. The seventh and tenth
atlas-axis complex in more detail; nonetheless, only the dorsal aspect neural spines are also tall. Even though the height alteration pattern does
was described. This complex was finally fully prepared and described in not have a precise arrangement and is not identical in all specimens, it is
complete detail by Sumida et al. (1992). Sumida et al. (1992) also de clear that some type of alteration of height and structure does occur in
scribed the atlas-axis complex of L. dynatis, which, due to its fragmen every specimen in the anterior dorsal vertebrae.
tary nature, provided additional insight only into the medial side of the Stout, spatulate ribs are visibly associated with dorsal vertebrae
left neural arch. three through seven (Fig. 2). The ribs on the right side of YPM 811 are
Re-examination of materials ascribed to Limnoscelis suggests no directed dorsolaterally, with some vertebrae abutting the scapula, whereas
significant emendations to the work of Sumida et al. (1992) are neces the ribs on the left side are directed ventrally. The anteriormost ribs are
sary, so only a brief summary is presented here. Both the atlas and axis approximately 17 cm long and increase serially in length posteriorly at
are multipartite elements in L. paludis (Sumida, 1992). The axial the level of the pectoral girdle. The ribs have a single head, but with
intercentrum is fused to the ventral aspect of the axial pleurocentrum, subequally developed areas for the capitulum and tubercle. Tubercular

FIGURE 1. Limnoscelis paludis, holotype, YPM 811, in dorsal view.


213
broadened proximal end with a long, narrow shaft.
The most posterior dorsal vertebrae exhibit serial changes in the
structure of the centra, pedicels, zygapophyses, transverse processes,
and neural spines. The centra continue to be longer than they are wide.
The posterior zygapophyses slope ventromedially 15 to 20 degrees,
whereas those of the anterior zygapophyses are horizontal. The trans
verse processes and neural arch pedicels are proportionately shorter
than those in the anterior portion of the column. Vertebrae 17 to 26 have
tall neural spines except the 22nd, which appears to have had a low,
ridge-like spine, which, though not preserved is evidenced by a long,
slender break scar. In some of the vertebrae coarse projections extend
from the lateral surface of the ventral portion of the neural spine. The tip
of the last neural spine is strongly bifid. All the presacral vertebrae up to
23 have been described as having accompanying ribs (Williston, 1912).
This description appears to be based on the holotype, but may have also
included supplementary information from MCZ 1947 and 1948. Cur
rently, only the left rib of vertebra 18 is present in the holotype, and the
ribs of CM 47653 are disarticulated, so they cannot be assigned to a
particular vertebra. The 18th left rib of L. paludis is shaped similarly to
the preceding ribs except for being about 50% shorter and less recurved.
Sacral Vertebrae and Ribs: Differing interpretations of the sac
ral region of Limnoscelis paludis have been presented. Williston (1911a,b,
1912) described a single sacral vertebra in the holotype with a following
FIGURE 2. Limnoscelis paludis, holotype, YPM 811, vertebrae 3-7 and large caudal vertebra, whereas Romer (1946) suggested that Limnoscelis
associated ribs. Cranial end of column is toward top of page. was a transitional model between “…a one-ribbed and two-ribbed condi
tion.” Sumida (1990) described two sacral vertebrae and ribs. Examina
and capitular facets are visible in anterodorsal and posteroventral views tion of YPM 811, the only specimen with an intact pelvic girdle, sug
of the rib, respectively. As preserved, all ribs in limnoscelids are single gests that L. paludis likely had two sacral vertebrae. Although the neural
headed. Romer (1946) hypothesized that the proximal ends of the ribs spines have been broken off and the vertebrae are not visible in ventral
had cartilaginous caps to allow the movement or exit of the vertebral view, the first sacral vertebra is obvious and well-developed. It is a large,
artery between the capitulum and tubercle. The shaft of the ribs is nar robust vertebra, approximately 5.3 cm wide across the anterior zygapo
row and roughly oval in cross-section. A low, triangular protrusion is physes, with swollen neural arches and widely-spaced anterior zygapo
visible on the posteroventral margin of the shaft just distal to its narrow physes. The first sacral ribs are robust and extend laterally for a short
est point. The distal ends are very thin dorsoventrally and spatulate in distance and then are abruptly angled posteriorly. The distal ends of the
shape. ribs are wide and point ventrally, where they contact the ilium almost
The more posterior vertebrae of the presacral series differ from vertically. The second vertebra is much smaller than the first and has
those at the anterior end of the series. The centra are significantly longer relatively much narrower neural arches and more closely-spaced anterior
than they are wide and they have beveled anterior and posterior edges zygapophyses. The neural spines are also missing on this vertebra. Con
ventrally, presumably for reception of the dorsal intercentra. A midventral fusion over the presence of one or two sacral vertebrae is understandable
depression is formed by the edges of these anterior and posterior ridges given the morphology of the associated ribs. The second sacral rib is
by the middle of the column. The zygapophyseal planes of the middor represented only by 1.8 cm, but retains its contact with the ilium. It is far
sal vertebrae slope medially about 15° without a posterior component as less robust than the first. Even though this rib does not directly contact
in the anterior vertebrae. The neural spines of vertebrae 12, 14, 16, 17, the second sacral vertebra, it is clear from its size and placement that it
and 19 are tall, in a manner similar to those described in the more anterior distally articulated with the posterior margin of the second sacral verte
vertebrae, whereas vertebrae 13 and 15 have virtually no neural spine, bra.
but only a low longitudinal ridge. Although the spine of vertebra 18 is Caudal Vertebrae and Ribs: In his restoration of Limnoscelis
tall, it is only about 5 mm in transverse width and much narrower than paludis, Williston (1912) based his description of it as having approxi
those of the anterior vertebrae. Clearly, the pattern of neural spine height mately 60 caudal vertebrae on the almost complete articulated tail of
alteration continues in the middorsal vertebrae. Isolated dorsal vertebrae MCZ 1947. He described the holotype as having 23 caudal vertebrae.
in Limnoscelis dynatis also exhibit the high- and low-spined morphology Although approximately 60 caudal vertebrae are currently found
(Fig. 3). in the holotype of L. paludis, this does not necessarily suggest that this
Most of the ribs in the middle and posterior of the column are not is the accurate tail count for this specimen. For accuracy, the following
visible in dorsal view in Limnoscelis paludis (YPM 811); only the left rib description is based only on the first 23 caudal vertebrae.
of vertebra 10, and partial right ribs of vertebrae 10, 11, 12 and 13, are The first few caudal vertebrae resemble the second sacral vertebra.
visible. They are much narrower and longer than the pectoral ribs and are Caudal centra are significantly longer than they are wide, but decrease
sharply recurved. In cross-section, the size and shape of the rib shaft serially in length posteriorly. Intercentra are present. The neural arches
remain constant throughout the rib’s length. In ventral view, the ribs of are not swollen, and the transverse processes project directly laterally.
vertebrae 9-16 are visible (Fig. 4). These ribs extend in natural position The neural spines are tall and blade-like and present only at the anterior
from the vertebrae to about their midpoint. At the midpoint, the ribs end of the series and decrease serially in height quickly posteriorly. No
collapse on themselves ventrally and their distal ends point alteration of spine height is present. Chevrons in the holotype are de
posteromedially, most likely indicating postmortem crushing of these scribed by Williston (1911a,b, 1912) as first appearing with the third
ribs. The ribs are slender in ventral view, but with their distal ends caudal, and ribs are fused to the first 10 or 11 caudals and distally curve
slightly dilated. A much shorter rib (5.5 cm) lies close to the ventral abruptly posteriorly (Sumida, 1990). However, in the holotype the chev
surface of vertebra 17. This rib ventrally overlies another short rib that is rons are not visible due to its encasement in plaster and only the first two
mostly hidden by matrix. This rib is visible in lateral view and has a left caudal ribs are visible.
214

FIGURE 3. Limnoscelis dynatis, holotype, CM 47653. Dorsal vertebrae demonstrating variable neural spine height. A, Tall-spined and B, low-spined
vertebra.

APPENDICULAR SKELETON girdle is partially crushed in this specimen, making the relative positions
of the scapular and coracoid regions even more difficult to characterize.
Pectoral Girdle and Forelimb: The pectoral girdle is completely The scapulocoracoid is composed of a dorsal scapula and a posterior
preserved only in YPM 811, but is not completely exposed. However, ventral coracoid. The scapular blade is relatively short dorsoventrally,
Limnoscelis dynatis (CM 47653) provides important information that is but expanded in its dorsal region anteroposteriorly. The dorsal margin of
not visible in the holotype. CM 47653 has a partial pectoral girdle the scapular blade in Limnoscelis paludis is thick and convex. The ante
preserved, including the right clavicle and both scapulocoracoids. rior border is almost vertically straight, whereas distally the posterior
The limnoscelid pectoral girdle consists of an interclavicle and border curves posteriorly.
paired cleithra, clavicles, and scapulocoracoids (Fig. 9A). The interclavicle Williston (1911a) and Romer (1946) both suggested that the dor
is known only in Limnoscelis paludis, but only recently has been ex sal portion of the scapular blade had a cartilaginous suprascapula at
posed in ventral view. However, the articulation of the interclavicle ob tached in Limnoscelis paludis. If there was a cartilaginous suprascapular
scures the ventral surface of the anterior portion of the expanded ante element, the base of attachment for this element would have been thick in
rior, diamond-shaped head of the interclavicle. What is visible is com L. paludis, but would probably have been too thin for an extensive
posed of a large, robust head with a long posterior process. The visible suprascapula in L. dynatis (Berman and Sumida, 1990). The
posterior portion of the head is slightly convex ventrally, and in poste scapulocoracoid is preserved in a single plane in L. dynatis, and Berman
rior view has the shape of a small arc (Fig. 5). The shaft is dilated and Sumida (1990) suggested that the coracoid plate had a significant
anteriorly where it contacts the head and slightly waisted posteriorly, ventromedial curvature. Unfortunately, the coracoid plate is not com
continuing to a rounded end. pletely visible in L. paludis (Figs. 4-5), so a ventromedial curvature of
The cleithra of limnoscelids are small, vestigial splints of bone the coracoid plate cannot be confirmed. A triangular, supraglenoid but
that contact the upper portion of the anterior margin of the scapular tress faces posterolaterally and is well developed at the base of the
blade (Williston, 1911a,b). The clavicles are larger bones located ventral scapula. A vertically expanded supraglenoid foramen is located near its
to the cleithra along the anterior margin of the scapular blade. They are dorsal apex. The existence of a suture between the anterior and posterior
partially visible in YPM 811. The clavicles consist of a dorsal stem and coracoids in L. paludis was described by Williston (1911a), but this
ventral plate that meet at an angle of approximately 110°, forming a half suture’s existence cannot be confirmed in YPM 811, as this portion of
sling-like shape. A posteriorly directed lamina on the lateral margin of the the girdle is not visible, and the suture is not present in L. dynatis.
dorsal stem contacts the anterior margin of the scapular blade. A groove However, an angular notch at approximately the same level as the cora
on the upper portion of the lateral surface of the clavicular stem likely coid suture in L. dynatis may indicate their separate ossifications. The
received the ventral end of the cleithrum. The ventral plate turns sharply coracoid plate in L. dynatis is smooth and thin, where it would have been
dorsally to form a high ridge, which is reduced as it continues as the covered ventrally by the clavicle and interclavicle. The glenoid fossa is
dorsal stem. This plate is subdivided into subequal anterior and posterior screw-shaped and faces slightly ventrally and posterolaterally at its
portions by a deep groove on the medial margin. The ventral plate is anterior end and dorsally at its posterior end. The glenoid fossa is sup
bowed ventrally and is slightly sculptured with dense, transverse striae ported by the supraglenoid buttress dorsally and a flange of thickened
(Berman and Sumida, 1990). In the fully articulated YPM 811, the ante bone anteroventrally. A coracoid foramen probably lies directly anterior
rior suture between the two clavicles is not apparent. to this flange in a deep fossa that undercuts the anterior portion of the
Characterization of the positioning of the scapulocoracoid in glenoid fossa in L. dynatis (Berman and Sumida, 1990).
limnoscelids is problematic because the scapulocoracoid is not fully Part of the left dorsal scapular blade is visible in YPM 811, as are
visible in the only articulated specimen of YPM 811. Also, the pectoral ventral parts of both coracoid plates. On the left side, the coracoid is
215

FIGURE 4. Limnoscelis paludis, holotype, YPM 811, in partial ventral view showing detail of anterior portion of postcranial skeleton.

crushed, and the glenoid fossa is present as a flat basin. A small foramen, extends from the dorsal surface of the ectepicondyle to the proximal
which may be the glenoid foramen, also known as the coracoid foramen, articular surface. Based on the descriptions of the humeri in L. paludis
extends dorsally from a small depression in the coracoid. The coracoid and L. dynatis, Sumida (1997) reconstructed the distal, ventral aspect of
extends posteriorly beyond the level of the distal end of the interclavicle the humerus in Limnoscelis (Fig. 9C).
stem. Repair of the posterior portion of the coracoid has resulted in its The radius is a short, stout element with flared dorsal and ventral
slight displacement posteriorly and medially as evidenced by the offset ends and a very short, narrow shaft connecting them. It is about 58% of
of a fracture on the medial side of the element. The right coracoid is the length of the humerus in Limnoscelis paludis. The dorsal surface is
slightly better preserved. The glenoid foramen is crushed closed, but the slightly convex with a small, blunt, midline ridge that runs the length of
glenoid fossa is well-preserved. The coracoid foramen and fossa are the shaft. The proximal end is only slightly expanded, and the proximal
located on the posterior end of the coracoid, directly lateral to the poste facet is slightly cupped for the articulation with the capitulum of the
rior end of the interclaviclar stem. Despite these preservational defects, humerus. The distal end of the radius is more expanded than the proximal
there appear to be no significant differences between the pectoral girdle end, with proximodistal parallel striae on its dorsal surface.
in YPM 811 and the better preserved examples in Limnoscelis dynatis. The ulna is approximately 84% of humeral length in Limnoscelis
Elements of both forelimbs are present in Limnoscelis paludis, paludis and more heavily built than the radius. A narrow shaft is deeply
but their exposure in YPM 811 is obscured partially by plaster, espe concave along its medial margin, but has a nearly straight lateral margin.
cially on the ventral surface (Figs. 4-5). However, due to postmortem The proximal end is flared medially to the semilunar or sigmoid notch,
anteroposterior crushing, the ventral surface of the left humerus of YPM which is matched on the medial margin on the distal end. The dorsal
811 is partially visible (Fig. 6). surface is slightly convex. The distal articular surface is clearly divided
The humerus of Limnoscelis paludis (Figs. 6,9) is an extremely into two facets for articulation with the carpals (Berman and Sumida,
stout bone, approximately 13.2 cm long. It resembles two tetrahedra set 1990). A slightly laterally facing facet articulates with the ulnare and
one on top of the other and offset approximately 90°. It has a large, pisiform, whereas the slightly medially facing surface articulates with
broad, quadrangular entepicondyle that is convex in outline. The the intermedium. Interestingly, Williston (1911a) described the ventral
ectepicondyle is robust, extends significantly anteriorly, and has a con surface of the radius and ulna in L. paludis as more flattened than the
cave outline. A large entepicondylar foramen appears to have been crushed convex dorsal surface. This may indicate slight postmortem compres-
closed in YPM 811. Stout supinator and deltoid processes project ante sion on the ventral surface. The ulna and radius have been reconstructed
riorly, and a pectoral process projects posteriorly. A large, hook-shaped in Figure 9E, based on the elements in L. paludis and L. dynatis.
notch separates the supinator and deltoid processes. A narrow ridge The manus is only preserved in the holotype of Limnoscelis
216

FIGURE 5. Limnoscelis paludis, holotype, YPM 811. Oblique ventral view


of pectoral girdle and associated structures.

paludis. The right manus has three proximal carpal bones, whereas the
left manus has four. Williston (1911a,b) described the right manus as also
having four proximal carpals, but only three are currently visible. The
four carpal bones are from lateral to medial: the pisiform, ulnare,
intermedium and radiale. The pisiform is semi-oval, tapered slightly on
its lateral margin, articulating with the ulna on its proximal margin and the
ulnare on its medial margin. The ulnare is the largest of the carpal bones, FIGURE 6. Limnoscelis paludis, holotype, YPM 811. Dorsal view of left
almost circular in shape, with a ridge running from the lateral margin forelimb and manus.
approximately three-fourths of the width of the bone, where it splits
into a small proximo-distal ridge. A small fossa is visible on either side of two carpals of the left manus is not sufficient for a confident identifica
the ridge. The ulnare articulates with the ulna proximally and intermedium tion. This indicates that the one remaining carpal in the right manus is
medially, but does not seem to have clearly defined facets for articula most likely correctly positioned, whereas the three distal carpals in the
tions with any other carpal elements (Williston, 1911a,b). The left manus may have shifted from their natural positions. These distal
intermedium is a small, dorsoventrally thick bone with two distinct carpals most likely represent the centrale and the third and fourth distal
ridges running proximo-distally on its lateral and medial sides. A fossa carpals (Williston, 1911a,b).
separates the two ridges in the left intermedium. In the right manus, the Williston (1911a) described digits II, III, and IV as being pre
intermedium articulates with the ulna, ulnare, and radius, whereas in the served complete, except for the distal, or ungual, phalanges of digits II
left manus, the intermedium has shifted slightly medially, giving the and IV. The ungual phalanx of digit I is also described as missing, as well
misleading impression of articulating only with the ulnare and radius. as the phalanges of digit V not being correctly articulated to the metacar
The radiale is the smallest of the four proximal carpals, almost oval in pal. However, those digits were preserved in the right manus, allowing
shape, flat on its dorsal surface and has a straight, flat radial border. It for a reconstruction of both hands (Williston, 1911a). The phalangeal
contacts the intermedium, but does not have a large articular surface for formula for the manus in limnoscelids appears to be 2-3-4-5-3. The
it. Williston (1911a,b) described the ventral surface of these proximal phalanges serially decrease in length distally. It is important to note that
carpal bones as flattened. the carpals and tarsals are possibly not fully ossified peripherally, and
Three distal carpal elements are preserved in YPM 811. Origi some of the more distal elements were probably not ossified at all, and
nally, Williston (1911a,b) described three distal carpals in both manus; this may affect the completeness of the specimen. The overall shape of
however, only one is visible in the right and three in the left. The distal the manus is rather broad, with the ungual phalanges shaped like small
carpals are all smaller than the proximal carpals and are all roughly circu hooves, with a thin rounded distal edge, which may have had a keratinous
lar in outline. Williston’s (1911a,b) illustration of the right manus indi covering in life (Williston, 1911a).
cates one directly proximal to metacarpal III, one lying between distal Pelvic Girdle and Hindlimb: The pelvic girdle is partially pre
carpal III and the intermedium, and one proximal to metacarpal IV. How served in Limnoscelis paludis. In YPM 811 the complete pelvis is pre
ever, the one distal carpal present in the right manus is located directly served in articulation; however, due to its having been embedded in
proximal to metacarpal III, whereas the carpals present in the left manus plaster only a portion of the dorsal and ventral surfaces are visible. The
are located proximal to metacarpal V, proximal to and directly in between pelvic girdle is reconstructed in Figure 9B based on specimens of L.
metacarpals III and IV, and proximal to metacarpal II. If distal carpal II is paludis and L. dynatis.
preserved in approximately correct position, then it may tentatively be The iliac blade is subrectangular in lateral outline except for the
identified as the medial centrale. The positional information for the other posterior process, which extends posteriorly. Romer (1946) suggested
217
that this posterior extension of the ilium was tipped with cartilage and
served as an attachment point for caudal tendons and ligaments. There is
no anterior expansion of the iliac blade. The ilium is proportionally short
and wide. A low sinuous or wavy ridge, the iliac shelf, extends
anteroposteriorly across the dorsal portion of the blade (Romer, 1946).
This shelf is a characteristic unique to diadectomorphs (Berman and
Sumida, 1990).A lateral iliac ridge and depression are present in Seymouria
and the pelycosaurian-grade synapsid Ophiacodon (Romer and Price,
1940; Berman and Sumida, 1990). However, the lateral iliac shelf is
developed to a much greater degree in Limnoscelis and in all other
diadectomorphs for which data on the ilium are available, thus making
this well-developed iliac shelf exclusively a diadectomorph characteris-
tic.
The pubis extends anteriorly beyond the level of the anterior
margin of the ilium and has a large anterior process, which has a rounded
convex anterior outline in Limnoscelis paludis. This is unlike the ilium,
which has a more quadrangular outline. An anteroposteriorly-directed
obturator foramen is partially crushed, yet visible in the right pubis of
YPM 811. Williston (1911a,b) described the pelvis of YPM 811 as
having a large, ventral midline keel formed by a ventral deepening of the
pubic and ischiadic symphyses. As preserved, the medial surface of the
ilium is convex and is touched by the two sacral ribs on the right side, and
by extrapolation, on the left side as well.
Currently, the ischium is not visible in Limnoscelis paludis, but
has been described in detail by Williston (1911a,b) and Romer (1946).
The ischium has a large posterior extension and is slightly concave in
posterior outline.
The acetabulum in the holotype of Limnoscelis paludis is only
partially visible in the right pelvis. What is visible suggests it is large and
oval with its long axis directed anteroposteriorly. The acetabulum is
supported dorsally by a small, ventrolaterally-expanded buttress in the
ilium. Notably, the anteroventral margin of the acetabulum has been
illustrated as reaching anteriorly, but not to the pubic margin, in YPM
811 (Romer, 1946).
Most elements of the hindlimb are present in the holotype of
Limnoscelis paludis (Fig. 7). However, the pes is only preserved com
pletely in articulation in MCZ 1948. The femur of L. paludis is very
robust with large, expanded ends connected by a relatively short, narrow FIGURE 7. Limnoscelis paludis, holotype, YPM 811. Dorsal aspect of
shaft, and has an average length of approximately 12.4 cm. It is visible in femur, tibia and fibula.
dorsal view in YPM 811.
The proximal head is angled posteriorly with an almost straight proximal articular surface is slightly sigmoid in outline with the anterior
anterior border and a concave posterior border. The dorsal surface of the end curving ventrally. The cnemial crest is well-developed originating on
proximal head is convex and smooth, except for a significantly rugose the dorsal surface of the proximal head and terminating proximally in a
area on its posterior margin where the puboischiofemoralis internus and small posterior knob. Posterior to the crest is a broad, shallow concavity.
ischiotrochantericus muscles probably inserted. The distal end of the The element narrows substantially to produce a well-defined shaft, then
femur is divided into two distinct condyles that articulate with the tibia expands again, although not as much as the proximal head. The dorsal
and fibula. The distal condyles of the femur are bulbous in shape and surface of the distal head is smooth and has a rounded convex distal
expanded dorsoventrally more than the proximal head. The anterior mar margin.
gin of the femur has a straight border, although it angles slightly anteri The fibula is longer than the tibia, being on average 80% of the
orly towards its distal end. The posterior margin of the femur expands length of the femur in the holotype of Limnoscelis paludis. As in the
posteriorly at both proximal and distal ends to accommodate the articu tibia, it is visible in only dorsal view in YPM 811. Its proximal head is
lar surfaces. The distal articular surface of the posterior condyle is shaped slightly flared with a proportionally longer shaft and a flared distal end.
like a parallelogram, with the dorsal edge slightly longer than the ventral The anterior margin is more convex, whereas the posterior margin is
margin. A small intercondylar fossa is present between the two condyles. almost straight. The articular surface of the proximal head is rugose, with
The surface of the distal end of the femur is smooth and convex, and the a slightly convex dorsal margin that is greatly expanded posteriorly. The
posterior condyle is slightly longer than the anterior condyle. The dorsal surface of the proximal head is flat, and appears much thicker
anteroproximal portion of the internal trochanter is visible in dorsal view dorsoventrally than the distal head. The dorsal surface of the distal head
in YPM 811. is slightly concave and has very fine parallel striae. The distal articular
The tibia is a robust bone with flared proximal and distal ends. surface appears very thin dorsoventrally and longer anteroposteriorly
Both tibiae are present in YPM 811, although only their dorsal surfaces than that of the proximal head.
are visible. It is on average 77% of the length of the femur in Limnoscelis Williston (1911a,b) described the hind foot of YPM 811 as only
paludis. The tibia of YPM 811 is thick dorsoventrally, especially at its being partially preserved in the left, and those bones being weathered and
proximal end. The tibia is anteroposteriorly broad, with an almost straight mostly disarticulated. However, this cannot be confirmed as the hind
anterior margin and a distinctly concave posterior margin, due primarily foot bones of YPM 811 have been reconstructed with plaster, and it is
to a more extensive posterior flaring of the proximal and distal ends. The impossible to determine which bones are the original elements Williston
218
“Diadectoides,” a taxon that Olson (1947) synonymized with Diadectes.
Olson’s synonymy reinforces the similarity of the pes in Limnoscelis
and diadectids, and lends some support to the potential application of
Berman and Henrici’s (2003) functional interpretation to Limnoscelis.
Their functional interpolation should be considered tentatively accepted
until the discovery of additional or more complete specimens of the pes
in Limnoscelis.
SUMMARY
The high degree of ossification of the postcranium and robust
development of axial and appendicular processes confirm the highly
terrestrial nature of Limnoscelis. Although this study is not intended as
the basis of a phylogenetic analysis of Limnoscelis or the
Diadectomorpha, its results do provide data relevant to future studies of
the group. Limnoscelis demonstrates variability in neural spine height, a
condition common in many basal amniotes and their close relatives. As in
basal amniotes, Limnoscelis has a phalangeal formula of 2-3-4-5-3 for the
manus and 2-3-4-5-4 for the pes.
This study reinforces the contention that the anterior processes
of the atlantal and axial intercentra, and the external iliac shelf are well
developed in Limnoscelis and likely defining features of the
Diadectomorpha, as recently suggested by Sumida et al. (1997) and
FIGURE 8. Limnoscelis paludis, MCZ 1948 (formerly YPM 809). Dorsal Sumida (1997).
aspect of the right pes. A variety of morphological differences are present between the
holotype of Limnoscelis paludis and L. dynatis. In the pectoral girdle and
described without preparation of the specimen. Examination of the cur hindlimb, differences can be seen in the scapulocoracoid, humerus, radius
rent mount of the left pes of MCZ 1948 (Fig. 8) indicates that the and ulna. First, the dorsal margin of the scapulocoracoid is thinner and
relationship of elements is not biologically reliable, and thus a complete less convex in L. dynatis than that in L. paludis. Also, the humerus in L.
paludis is 2.2 cm longer than that in L. dynatis. The radius is slightly
current description of the pes of MCZ 1948 is not possible. Because of
these limitations and to facilitate comparison with other taxa, Williston’s longer in L. dynatis than L. paludis, with the former being about 62% of
humeral length, while the latter is only 58% of humeral length. As in the
(1911 a,b) description of the pedes of MCZ 1948 is included here as the
humerus, the ulna (84% and 80% of humeral length, respectively) of L.
general model in Limnoscelis paludis.
In MCZ 1948 (formerly YPM 809) the pedes are much better paludis is longer than that in L. dynatis.
preserved, with two tarsals and partial digits present in the left pes and Similar differences are also present in the pelvic girdle and hindlimb.
two tarsals and complete digits present in the right pes (Williston, 1911a,b, The femora are approximately 1.3 cm in length on average shorter in
1912). In the pedes of MCZ 1948, Williston (1911a,b) identified the Limnoscelis dynatis than those in L. paludis. The femora in L. paludis
proximal tarsals as the fibulare and the fused tibiale and intermedium. He also appear more robust and stout in proportion than those in L. dynatis.
illustrated the fibulare as an almost circular element and the joined tibiale Both the tibia and fibula are longer in L. dynatis than they are in L.
intermedium as cuboidal, with articular surfaces for the fibula and tibia, paludis. In L. dynatis, the tibia and fibula are 80% and 88% of the femoral
separated by a slight notch. However, he admitted the possibility that length, whereas the tibia and fibula in L. paludis are 77% and 80% of
the tibiale was cartilaginous and the two proximal tarsals are the fibulare femoral length.
and intermedium. Romer (1946) agreed with this view, and I concur. If A synapomorphy of Limnoscelis is the presence of an anteriorly
the tibiale and intermedium were fused, the element would be much larger directed channel extending from the acetabulum to the anterior edge of
than what is present in the current mount of the right pes of MCZ 1948. the pubis. Whereas this channel is illustrated by Romer (1946), it is not
It is most likely that the two proximal tarsals in this specimen are the clearly visible in YPM 811. However, its presence is confirmed by its
fibulare and the intermedium, with the tibiale missing. Williston (1911 occurrence in L. dynatis. The anteroventral margin of the acetabulum in
a,b) also recovered some disarticulated phalanges. this specimen extends anteriorly as a narrow channel that reaches the
The phalanges of the right pes in MCZ 1948 were found in two anterior border of the pubis (Berman and Sumida, 1990). Pelletier et al.
separated blocks whose surfaces had been damaged during excavation, (2007) have described a similar, though shorter, channel in the Permian
thus distorting the articular surfaces of the phalanges and making their diadectomorph Tseajaia. The presence of a strongly developed adductor
ridge of the femur is not unique to Limnoscelis, however, Pelletier et al.
serial positions difficult to determine. However, Williston (1911a,b)
articulated them on the basis of their similar morphologies and similar (2007) have shown that the degree of angulation of this feature may be
matrix in a seamless, anatomical association. The phalanges of the pes are useful in distinguishing the genera of Diadectomorpha. Although this
similar to those of the manus except for being slightly broader. The feature is difficult to assess in YPM 811, it is visible in CM 47653. This
phalanges serially shorten distally. The ungual phalanges are hoof-like. If study confirms the analysis of Pelletier et al. (2007) that the angulation
the above interpretation is accepted, then the phalangeal count for the of the adductor crest relative to the long axis of the femur in Limnoscelis
right pes of MCZ 1948 is 2-3-4-5-4 (Fig. 8). Romer (1946) suggested dynatis is greater than that in either Tseajaia or Diadectes. The construc
that the incomplete preservation of the carpus and tarsus is due to tion of the pes in Limnoscelis presents intriguing potential for compari
son with the more thoroughly described condition in diadectid
imperfect ossification of the elements. Berman and Henrici (2003) have
have described in greater detail the pedes in the diadectid diadectomorphs diadectomorphs (Berman and Henrici, 2003); however, a functional in
terpretation is outside the scope of this review.
Diadectes and Orobates, and suggest it served as a mechanism to allow
anterior orientation of the manus and pes during locomotion. Of note, ACKNOWLEDGMENTS
Williston (1911 a, b; 1912) based further description of the hind foot of
L. paludis on FMNHUC 650. This specimen has since been identified as I thank Dr. Stuart Sumida for sharing his knowledge of limnoscelids
219

FIGURE 9. Reconstructions of various elements of Limnoscelis. A, Left lateral aspect of pectoral girdle, B, Left lateral view of pelvic girdle, C, Distal
ventral aspect of left humerus, D, Ventral view of left femur, E, Dorsal aspect of right radius and ulna, F, Dorsal aspect of left tibia and fibula. The
reconstruction is of slightly disarticulated elements that were laid flat for ease of viewing. Also, the preaxial curvature evident in the tibia may be slightly
exaggerated in the illustration due to the fact that elements available for examination were somewhat flattened. All scale bars equal 1 cm. Adapted from
Sumida (1997).
220
and for all his support throughout my time at California State University and borrowing specimens. Sincere thanks also to Jessica Cundiff of the
San Bernardino (CSUSB) and beyond, both with this project and others. Museum of Comparative Zoology for her help with the history and
This study fulfilled in part the requirements for the Masters of Science at study of the MCZ Limnoscelis specimens. In the CSUSB Vertebrate
California State University, San Bernardino. Sincere thanks are also due Paleontology laboratory, thanks are acknowledged to Gavan Albright
my thesis committee: Drs. Joan Fryxell, Anthony Metcalf, and Jim and Kim Scott for assistance with the illustrations, and to Ken Noriega
Ferrari. I am grateful to the staff of the Raymond M. Alf Museum of who provided support and assistance during this study. Thanks are also
Paleontology for all of their support and patience. My sincere apprecia due to Brett Austin Kennedy for his support and encouragement through
tion is also extended to Dr. David S Berman and Ms. Amy Henrici of the out the challenges of my graduate studies. Thanks to Dr. Dave S Berman,
Carnegie Museum of Natural History for all of their assistance in study Ms. Amy Henrici, Dr. Spencer G. Lucas, and Dr. Stuart S. Sumida for
ing the holotypic skeleton of Limnoscelis, as well as their generous thorough reviews of drafts of this manuscript; they helped make this a
hospitality. Thank you also to Chuck Schaff of the Museum of Com significantly clearer and more concise paper. All mistakes or misinterpre
parative Zoology for his above-and-beyond-the-call help with studying tations however, remain my own. This project was funded in part by the
California State University, San Bernardino ASI and IRP grants.

REFERENCES

Berman, D.S, and Henrici, A. C., 2003, Homology of the astragalus and structure of the rare Lower Permian diadectomorph Tseajaia campi
structure and function of the tarsus of Diadectidae: Journal of Paleontol Vaughn: Journal of Vertebrate Paleontology, v. 27, p. 128A.
ogy, v. 77, p. 172-188. Romer, A.S., 1946, The primitive reptile Limnoscelis restudied: American
Berman, D.S, and Sumida, S.S., 1990, A new species of Limnoscelis (Am Journal of Science, v. 244, p. 149-188.
phibia, Diadectomorpha) from the Late Pennsylvanian Sangre de Cristo Romer, A.S., 1966, Vertebrate paleontology, Third Edition. Chicago, The
Formation of central Colorado: Annals of Carnegie Museum, v. 59, p. University of Chicago Press, 468 p.
303-341. Sumida, S. S., 1990, Vertebral morphology, alternation of neural spine height,
Berman, D.S, Sumida, S.S, and Lombard, R.E., 1992, Reinterpretation of and structure in Permo-Carboniferous tetrapods, and a reappraisal of
the temporal and occipital regions in Diadectes and the relationships of primitive modes of terrestrial locomotion: University of California
the diadectomorphs: Journal of Paleontology, v. 66, p. 481-499. Publications in Zoology, v. 122, p. 1-133.
Berman, D.S, Sumida, S.S., and Lombard, R.E., 1997, Biogeography of Sumida, S.S., 1997, Locomotor features of taxa spanning the origin of
primitive amniotes; in Sumida, S.S. and Martin, K.L.M., eds., Amniote amniotes; in Sumida, S.S. and Martin, K.L.M., eds., Amniote origins:
origins: Completing the transition to land. San Diego, Academic Press, Completing the transition to land. San Diego, Academic Press, p. 353-
p. 85-139. 398.
Berman, D.S. 2000. Origin and early evolution of the amniote occiput: Sumida, S. S., Lombard, R. E. and Berman D. S, 1992, Morphology of the
Journal of Paleontology, v. 74, p. 938-956. atlas-axis complex of the late Palaeozoic tetrapod suborders
Fracasso, M.A., 1980, Age of the Permo-Carboniferous Cutler Formation Diadectomorpha and Seymouriamorpha: Philosophical Transactions of
vertebrate fauna from El Cobre Canyon, New Mexico: Journal of Pale the Royal Society of London, Series B, v. 336, p. 259-273.
ontology, v. 54, p. 1237-1244. Wideman, N. K., and Sumida, S. S., 2004, Taxonomic status of the tetrapod
Fracasso, M.A., 1983, Cranial osteology, functional morphology, system Limnostygis relictus and its bearing on the temporal distribution of basal
atics and paleoenvironment of Limnoscelis paludis Williston [Ph.D. amniotes: Journal of Vertebrate Paleontology , v. 24, p. 129A.
dissertation]: Yale University, New Haven, 624 p. Wideman, N. K., Sumida, S. S., and O’Neil, M., 2005, A reassessment of the
Fracasso, M.A., 1987, Braincase of Limnoscelis paludis Williston: Postilla, taxonomic status of the materials assigned to the Early Permian tetra
no. 201, p. 1-22. pod genera Limnosceloides and Limnoscelops: New Mexico Museum of
Lee, M.Y.S., and Spencer, P., 1997, Crown clades, key characters and taxo Natural History and Science, Bulletin, 30. p. 358-362.
nomic stability: when is an amniote not an amniote?; in Sumida, S.S. and Williston, S.W., 1911a, A new family of reptiles from the Permian of New
Martin, K.L.M., eds., Amniote origins: Completing the transition to Mexico: American Journal of Science, v. 31, p. 378-398.
land. San Diego, Academic Press, p. 61-84. Williston, S.W., 1911b, American Permian Vertebrates. Chicago, The Uni
Olson, E. C., 1947, The family Diadectidae and its bearing on the classifica versity of Chicago Press, 145 p.
tion of reptiles: Fieldiana: Geology, v. 11, p. 3-53. Williston, S.W., 1912, Restoration of Limnoscelis, a cotylosaur reptile
Pelletier, V., Sumida, S.S., Walliser, J., and Berman, D.S., 2007, Hindlimb from New Mexico: American Journal of Science, v. 34, p. 457-468.
Lucas et al., eds., 2010, Carb-Permian transition in Cañon del Cobre. New Mexico Museum of Natural History and Science Bulletin 49.

221
TYPOTHORAX COCCINARUM (ARCHOSAURIA: STAGONOLEPIDIDAE)
FROM THE UPPERTRIASSIC (REVUELTIAN)PETRIFIED FOREST FORMATION,
ELPUERTOCITO, CAÑON DELCOBRE, RIOARRIBA COUNTY, NEWMEXICO

JUSTINA. SPIELMANN AND SPENCER G. LUCAS


New Mexico Museum of Natural History and Science, 1801 Mountain Rd. NW, Albuquerque, NM 87104-1375

Abstract—Specimens of the Late Triassic aetosaur Typothorax coccinarum from localities near El Puertocito in
the Cañon del Cobre, Rio Arriba County, New Mexico, consists of dorsal osteoderms, vertebrae and other
postcrania. These fossils are from the Painted Desert Member of the Petrified Forest Formation, as are other
records of T. coccinarum from the Chama basin, and confirm the Revueltian age of these strata.

INTRODUCTION odont, Aetosaurid?” Both AMNH 7634 and 7635 have similar preserva
Aetosaurs (Archosauria: Stagonolepididae) are a suborder of tion, and it can be surmised that they were collected from the same
archosaurs whose fossils are known from the Upper Triassic strata of locality and cataloged separately based on elements, with the vertebral
North and South America, Greenland, Europe, India, North Africa and fragments given one number (AMNH 7634) and the osteoderms, limb
Madgascar (Heckert and Lucas, 2000). Aetosaur fossils are primarily bones and unidentified postcrania given another (AMNH 7635). The
osteoderms (bony armor) that formed a carapace around their neck, locality information listed for AMNH 7634 and 7635 is “Chinle, Trias
trunk and tail. Associated and articulated specimens are rare (Huntet al., sic, El Cobre Canyon, New Mexico,” El Cobre Canyon being one of the
1993; Heckert et al., 2010), but individual osteoderms have taxonomic other names used for the Cañon del Cobre. Given that Upper Triassic
utility, allowing identification to genus or species level (e.g., Long and strata are limited to the southern end of the canyon, around El Puertocito,
the AMNH material was likely collected not far from the NMMNH
Ballew, 1985; Long and Murry, 1995; Heckert and Lucas, 2000).
Numerous Late Triassic vertebrate localities in Rio Arriba County material.
yield aetosaur fossils (Hunt and Lucas, 1989, 1993; Lucas and Hunt,
NMMNH P-56493
1992; Long and Murry, 1995). These sites include the Snyder (Typothorax
coccinarum and Rioarribasuchus chamaensis) (Heckert et al., 2003), NMMNH P-56493 consists primarily of poorly preserved
Canjilon (Typothorax coccinarum) (Zeigler et al., 2002a; Nesbitt and osteoderm fragments with a handful of well preserved vertebral centra,
Stocker, 2008) and Hayden (Typothorax coccinarum) (Irmis et al., 2007; diagnostic osteoderm fragments and a left proximal humerus.
A. Downs, personal commun.) quarries, all within the Petrified Forest
Vertebral centra
Formation (Revueltian) and one locality from the Rock Point Formation
(Apachean) that yields a nearly complete skeleton of Aetosaurus that Five vertebral centra are present: one anterior dorsal (Fig. 2A-B),
awaits description. While vertebrate fossil collecting in Cañon del Cobre one mid dorsal (Fig. 2C-E), one posterior dorsal (Fig. 2F-H), one sacral
(= El Cobre Canyon) has been ongoing for over a century, this effort has (Fig. 2I-J) and one posterior caudal (Fig. 2K-M).
been primarily focused on Pennsylvanian and Permian strata. Few Trias The anterior dorsal consists of the body of the centrum, missing
sic specimens have been reported from the canyon. Here, we describe both ends, one side of the neural arch and the base of a transverse process
various postcranial fragments of three specimens of Typothorax (Fig. 2A-B). The centrum has a poorly developed ventral keel, and the
coccinarum (NMMNH P-56493 and AMNH 7634, 7635) and summa base of the transverse process is subtriangular in cross section. We iden
rize their significance for regional biostratigraphy. tify it as an anterior dorsal because of the position of the transverse
Institutional abbreviations: AMNH, American Museum of process high on the neural arch and its relatively large size; cervical
Natural History, New York and NMMNH, New Mexico Museum of vertebrae are extremely reduced in size compared to dorsal vertebrae in
Natural History and Science, Albuquerque. Typothorax coccinarum (Long and Murry, 1995, p. 104; Heckert et al.,
2010).
PROVENANCE The mid dorsal is more complete, missing all pre- and
NMMNH P-56493 was collected from NMMH locality 7141, a postzygapophyses, the neural spine and lateral ends of the transverse
pale green mudstone slope in an unsurveyed section of the southwest process (Fig. 2C-E). The anterior articular surface of the centrum is
corner of the Canjilon SE7.5’ quadrangle, immediately west-southwest circular (Fig. 2C), whereas the posterior surface is elliptical, and the
of El Puertocito (Fig. 1). Numerous outcrops of the Late Triassic Poleo, anterior surface is offset slightly higher than the posterior surface (Fig.
Petrified Forest and Rock Point formations bracket both east and west 2D). The centrum is arched in lateral view and is double keeled (Fig. 2E).
sides of El Puertocito, which forms the natural southern boundary of the The neural canal is filled with matrix and cannot be discerned. The trans
Cañon del Cobre (Fig. 1) (Kempter et al., 2007; Lucas et al., this volume). verse processes are rectangular in dorsal view and project posterolaterally.
Kempter et al. (2007) mapped these outcrops as Petrified Forest Forma The lack of a distinct capitular facet identifies the vertebra as a dorsal,
tion/Rock Point Formation undivided, however, based on the green, slope and the position of the transverse process on the neural arch at the level
forming mudstones and stratigraphic position, NMMNH locality 7141 of the dorsal margin of the centrum places it in the middle of the series.
is in the Painted Desert Member of the Petrified Forest Formation. The posterior dorsal is very incomplete, preserving only the body
The location that the AMNH material was collected from cannot of the centrum and its posterior articular surface and portions of the
be so easily identified. All the material was collected by S.J. Olsen and A. neural arch (Fig. 2F-H). The centrum has a single keel and an elliptical
Lewis in 1954 during an expedition for the Museum of Comparative posterior articular surface. The large size of the vertebra and total lack of
Zoology, Harvard University, and exchanged with the AMNH in 1956. a distinct capitular facet suggest it is a posterior dorsal, very close to the
The elements were originally identified as “Thecodont” and later “Thec sacrum.
222

FIGURE 1. Index map of the Cañon del Cobre. NMMNH locality 7141, where NMMNH P-56493 was collected, marked with a star.
223

FIGURE 2. Typothorax coccinarum, NMMNH P-56493. A-B, Anterior dorsal vertebra in A, dorsal and B, end views. C-E, Mid dorsal vertebra in C,
anterior, D, right lateral and E, ventral views. F-H, Posterior dorsal vertebra in F, anterodorsal, G, right lateral and H, posterior views. I-J, Sacral vertebra
in I, posterior and J, ventral views. K-M, Posterior caudal vertebra in K, left lateral, L, ventral and M, anterior views.
224
The sacral vetebra is only represented by an incomplete centrum Typothorax antiquum, which has coarser, less dense pitting on its para
and the bases of the transverse processes (Fig. 2I-J). The centrum is not median osteoderms (Lucas et al., 2002), and Rioarribasuchus chamaensis,
keeled, and both anterior and posterior articular surfaces are incomplete. which has a prominent boss with fine ridges radiating out from it (Zeigler
The bases of the transverse processes are very large, running et al., 2002b; Lucas et al., 2006).
anteroposteriorly nearly the entire length of the centrum, identifying the Left proximal humerus
element as a sacral centrum.
The posterior caudal vertebra is nearly complete, missing only The left proximal humerus is broken just distal to the end of the
one of its transverse processes and the dorsal tip of the neural spine (Fig. deltopectoral crest (Fig. 3A-B). The articular surface of the humerus is
2K-M). The double keeled centrum is elongate and cylindrical, with tetralobate and offset from the shaft. The deltopectoral crest, when in
circular articular ends. The transverse process is reduced in size and articulation, is directed ventrally. All these features are consistent with
splint-like. The neural spine, as preserved, is low, and no indication of the humeri of aetosaurs (Long and Murry, 1995).
pre- or postzygapophyses is present. The elongate centrum, splint-like
transverse process and lack of zygapophyses confirm the identification AMNH 7634 AND 7635
as a posterior caudal vertebra. None of the vertebral material in AMNH 7634 was complete
Osteoderm fragments enough to describe, though given its preservation it likely came from the
same locality as AMNH 7635, as noted above, and may represent parts
No complete osteoderms were collected, but all fragments pos of the same individual.
sess a consistent patterning offine, randomly arranged pits (Fig. 3C-H). The identifiable material from AMNH 7635 comprises a right iliac
All osteoderm fragments are flat or only slightly keeled; none are flexed, blade fragment (Fig. 4A-B), a proximal humerus (Fig. 4C-D), both proxi
as in lateral elements, and thus are paramedian osteoderms. The random mal ulnae (Fig. 4E-I), proximal and distal fragments of a femur (Fig. 5A
pattern of fine pitting is a diagnostic feature of Typothorax coccinarum F), a proximal tibia (Fig. 5G-I), a proximal ?fibula (Fig.5J-L), one com
(Heckert and Lucas, 2000; Heckert et al., 2010), contrasting with plete (Fig. 6A) and two fragmentary paramedian osteoderms (Fig. 6B-C)

FIGURE 3. Typothorax coccinarum, NMMNH P-56493. A-B, Proximal left humerus in A, anterior and B, posterior views. C-D, Osteoderm fragment in
C, dorsal, D, ventral and E, end view. F-H, Osteoderm fragments in dorsal view.
225

FIGURE 4. Typothorax coccinarum, AMNH 7635. A-B, Iliac blade fragment in A, medial and B, lateral views. C-D, Proximal right humerus in C, posterior
and D, anterior views. E-H, Proximal ulnae in E, G, posterior, F, I, anterior and H, medial views.
226
and two complete caudal ventral osteoderms (Fig. 6D-E). Osteoderms
Iliac blade fragment Five osteoderms, three paramedian and two caudal ventral, are
complete enough for description (Fig. 6). The three paramedians are
The posterior portion of a heavily concreted, right iliac blade is
complete (Fig. 6A), nearly complete (Fig. 6B) and fragmentary (Fig. 6C),
preserved (Fig. 4A-B). It is subrectangular in lateral outline and has a
respectively. Both the complete and nearly complete paramedians are
prominent ridge running posteriorly along its medial side. This medial right sacral or caudal osteoderms based on the presence of a small,
ridge is more pronounced in Typothorax coccinarum (compare Fig. 4A to subtriangular dorsal boss that projects beyond the posterior margin of
Long and Murry, 1995, fig. 106) than in other aetosaurs such as the rest of the osteoderm. This boss only occurs on sacral and caudal
Desmatosuchus haplocerus (Long and Murry, 1995, fig. 91) and
osteoderms, increasing in size down the tail (Heckert et al., 2010). Also,
Stagonolepis wellesi (Long and Murry, 1995, fig. 80). both share prominent anterior bars, ventral keels and the random, rela
tively coarse pitting characteristic of Typothorax coccinarum. The frag
Proximal humerus
mentary osteoderm preserves only the right half of a right paramedian
A right proximal humerus is preserved, though heavily concreted, osteoderm, but shares with the other osteoderms a pattern of fine pit
with almost no natural bone surface showing (Fig. 4C-D). It is identical ting.
in morphology to the proximal humerus described in NMMNHP-56493 The caudal ventral osteoderms are distinct from the paramedians;
(e.g., tetraradiate proximal articulation, proximal humerus offset from they are square to subrectangular in outline, and lack anterior bars or
shaft, etc.). prominent bosses (Fig. 6D-E). One is larger than the other and is inter
preted as anterior. Their overall morphology is consistent with caudal
Proximal ulnae
ventral osteoderms of Typothorax coccinarum described by Heckert et
The left and right proximal ulnae are identical, with an undevel al. (2010).
oped, blunt olecranon process, shallow articular surface for contacting
the distal humerus and anteroposterior expansion of the proximal ulna DISCUSSION
compared to the compressed shaft (Fig. 4E-I). Both ulnae are approxi The morphology of all elements from NMMNH P-56493 and
mately the same size, suggesting they come from the same individual. AMNH 7634 and 7635 is consistent with Typothorax coccinarum, most
The L-shaped olecranon process is identical to the ulnae of Typothorax notably the diagnostic patterning of the paramedian osteoderms (Long
coccinarum described by Heckert et al. (2010). and Murry, 1995; Heckert and Lucas, 2000; Heckert et al., 2010).
The preservation of all three specimens is consistent, with creamy
Proximal and distal femoral fragments
white to black bone, some elements partially concreted with yellow-gray
An isolated femoral head is complete with both the proximal ends matrix and all elements possessing dark staining or spots, possibly the
of the greater and fourth trochanters preserved (Fig. 5A-C). The distal result of manganese. In addition, no elements are duplicated between the
left femur fragment preserves both condyles, the lateral condyle pos specimens. While it cannot be definitively demonstrated that all three
sessing a large, subtriangular, posteriorly-directed process and the lateral specimens were collected from the same locality, especially given the
condyle possessing a lower, posterolaterally-directed process (Fig. 5D incomplete data for the AMNH specimens, it suggests that they all were
F). The larger lateral condyle has been noted in other specimens of recovered from the same stratigraphic interval, i.e., the Painted Desert
Typothorax coccinarum (Martz, 2002; Heckert et al., 2010). A fossa is Member of the Petrified Forest Formation.
developed between these processes on the posterior side of the element. Typothorax coccinarum is an index fossil of the Revueltian land
vertebrate faunachron of Lucas (1998). Its type material and other speci
Proximal tibia mens are known from localities in the Painted Desert Member of the
The left proximal tibia is expanded at its articular end, and the Petrified Forest Formation east of Cañon del Cobre in the Chama basin
(e.g., Lucas and Hunt, 1992; Heckert et al., 2005). The occurrence of T.
articular surface has an elliptical to kidney-shaped outline (Fig. 5G-I).
coccinarum in the Painted Desert Member of the Petrified Forest For
Proximal ?fibula mation near El Puertocito thus is consistent with its stratigraphic distri
bution elsewhere in the Chama basin.
The proximal ?fibula has two distinct articular surfaces on its While Late Triassic tetrapod fossils are rare locally in the Cañon
proximal end, one subcircular and slightly larger than the other, rectangu del Cobre, the recognition of Typothorax coccinarum within the Painted
lar surface (Fig. 5J-L). In mediolateral view, an elongate fossa extends
Desert Member of the Petrified Forest Formation further confirms its
distally down the element. We tentatively identify it as a proximal fibula
presence within stata of Revueltian and its utility as a biostratigraphic
based on its overall size; it is approximately the same size as the tibia and index taxon regionally (Lucas et al., 2007).
much larger than the ulnae, suggesting it cannot be a radius. However,
descriptions of Typothorax coccinarum fibulae do not describe two dis ACKNOWLEDGMENTS
tinct, proximal articulations (Long and Murry, 1995; Martz, 2002; Heckert
et al., 2010), thus making identification uncertain. Randy Pence and Susan Harris discovered and collected NMMNH
P-56493. Carl Mehling provided collection access at the AMNH. An
drew Heckert and Adrian Hunt provided helpful reviews that improved
the manuscript.
227

FIGURE 5. Typothorax coccinarum, AMNH 7635. A-C, Femoral head in A, posterior, B, anterior and C, articular views. D-F, Left distal femur in D,
anterior, E, posterior and F, distal views. G-I, Left proximal tibia in G, medial, H, lateral and I, articular views. J-L, Proximal fibula in J-K, mediolateral
views and L, proximal views.
228

FIGURE 6. Typothorax coccinarum, AMNH 7635. A, Complete right paramedian osteoderm in dorsal view. B, Nearly complete paramedian osteoderm in
dorsal view. C, Fragmentary paramedian osteoderm in dorsal view. D-E, Caudal ventral osteoderms in ventral view.
229
REFERENCES

Heckert, A.B. and Lucas, S.G., 2000, Taxonomy, phylogeny, biostratigra rial from the Chinle Formation of Petrified Forest National Park: Mu
phy, biochronology, paleobiogeography, and evolution of the Late Tri seum of Northern Arizona, Bulletin 47, p. 45-68.
assic Aetosauria (Archosauria: Crurotarsi): Zentralblatt für Geologie und Long, R.A. and Murry, P.A., 1995, Late Triassic (Carnian and Norian)
Paläontologie, Teil I 1998, Heft 11-12, p. 1539-1587. tetrapods from the southwestern United States: New Mexico Museum of
Heckert, A.B., Zeigler, K.E. and Lucas, S.G., 2003, Aetosaurs (Archosauria: Natural History and Science, Bulletin 4, 254 p.
Stagonolepididae) from the Upper Triassic (Revueltian) Snyder quarry, Lucas, S.G., 1998, Global Triassic tetrapod biostratigraphy and
New Mexico: New Mexico Museum of Natural History and Science, biochronology: Palaeogeography, Palaeoclimatology, Palaeoecology, v.
Bulletin 24, p. 115-126. 143, p. 347-384.
Heckert, A.B., Lucas, S.G., Sullivan, R.M., Hunt, A.P. and Spielmann, J.A., Lucas, S.G., Heckert, A.B. and Hunt, A.P., 2002, A new species of the
2005, The vertebrate fauna of the Upper Triassic (Revueltian: lower aetosaur Typothorax (Archosauria: Stagonolepididae) from the Upper
mid Norian) Painted Desert Member (Petrified Forest Formation: Chinle Triassic of east-central New Mexico: New Mexico Museum of Natural
Group) in the Chama basin, northern New Mexico: New Mexico Geo History and Science, Bulletin 21, p. 221-233.
logical Society, Guidebook 56, p. 302-318. Lucas, S.G. and Hunt, A.P., 1992, Triassic stratigraphy and paleontology,
Heckert, A.B., Lucas, S.G., Rinehart, L.F., Spielmann, J.A., Celeseky, M.D. Chama basin and adjacent areas, north-central New Mexico: New Mexico
and Hunt, A.P., 2010, Articulated skeletons of the aetosaur Typothorax Geological Society, Guidebook 43, p. 151-172.
coccinarum Cope (Archosauria: Stagonolepididae) from the Upper Tri Lucas, S.G., Hunt, A.P. and Spielmann, J.A., 2006, Rioarribasuchus, a new
assic Bull Canyon Formation (Revueltian: early-mid Norian), eastern name for an aetosaur from the Upper Triassic of north-central New
New Mexico, USA: Journal of Vertebrate Paleontology, in press. Mexico: New Mexico Museum of Natural History and Science, Bulletin
Hunt, A.P. and Lucas, S.G., 1989, Late Triassic vertebrate localities in New 37, p. 581-582.
Mexico; in Lucas, S.G. and Hunt, A.P., eds., Dawn of the age of dinosaurs Lucas, S.G., Hunt, A.P., Heckert, A.B. and Spielmann, J.A., 2007, Global
in the American Southwest: Albuquerque, New Mexico Museum of Natu Triassic tetrapod biostratigraphy and biochronology: 2007 status: New
ral History, p. 72-101. Mexico Museum of Natural History and Science, Bulletin 41, p. 229
Hunt, A.P. and Lucas, S.G., 1993, Stratigraphy and vertebrate paleontology 240.
of the Chinle Group (Upper Triassic), Chama basin, north-central New Lucas, S.G., et al., this volume [Geology]
Mexico: New Mexico Museum of Natural History and Science, Bulletin Martz, J.W., 2002, The morphology and ontogeny of Typothorax coccinarum
2, p. 61-69. (Archosauria: Stagonolepididae) from the Upper Triassic of the Ameri
Hunt, A.P., Lucas, S.G. and Reser, P.K., 1993, A complete skeleton of the can Southwest [M.S. thesis]: Texas Tech University, Lubbock, 279 p.
stagonolepipid Typothorax coccinarum from the Upper Triassic Bull Nesbitt, S.J. and Stocker, M.R., 2008, The vertebrate assemblage of the
Canyon Formation of east-central New Mexico, USA: New Mexico Late Triassic Canjilon quarry (northern New Mexico, USA) and the
Museum of Natural History and Science, Bulletin 3, p. 209-212. importance of apomorphy-based assemblage comparisons: Journal of
Irmis, R.B., Nesbitt, S.J., Padian, K., Smith, N.D., Turner, A.H., Woody, D. Vertebrate Paleontology, v. 28, p. 1063-1072.
and Downs, A., 2007, A Late Triassic dinosauromorph assemblage from Zeigler, K.E., Heckert, A.B. and Lucas, S.G., 2002a, A tale of two sites: A
New Mexico and the rise of dinosaurs: Science, v. 317, p. 358-361. taphonomic comparison of two Late Triassic (Chinle Group) vertebrate
Kempter, K., Zeigler, K., Koning, D. and Lucas, S., 2007, Preliminary fossil localities from New Mexico: New Mexico Museum of Natural
geologic map of the Canjilon SE quadrangle, Rio Arriba County, New History and Science, Bulletin 21, p. 285-290.
Mexico: New Mexico Bureau of Geology and Mineral Resources, Open Zeigler, K.E., Heckert, A.B. and Lucas, S.G., 2002b, A new species of
file Geological Map 150, 1:24,000. Desmatosuchus (Archosauria: Aetosauria) from the Upper Triassic of
Long, R.A. and Ballew, K.L., 1985, Aetosaur dermal armor from the Late the Chama basin, north-central New Mexico: New Mexico Museum of
Triassic of southwestern North America, with special reference to mate Natural History and Science, Bulletin 21, p. 215-219.

You might also like