Sie sind auf Seite 1von 138

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds.

, 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

11

ARCTOCYON (MAMMALIA, ARCTOCYONIDAE)


FROM THE PALEOCENE OF NORTH AMERICA
PETER E. KONDRASHOV1 AND SPENCER G. LUCAS2
1

Department of Biology, Northwest Missouri State University, 800 University Drive, Maryville, MO 64468
and Paleontological Institute of the Russian Academy of Sciences, Moscow, Russia;
2
New Mexico Museum of Natural History, 1801 Mountain Road, Albuquerque, NM 87104

AbstractThe cranial and dental morphology of two of the North American arctocyonids, Arctocyon ferox and
A. corugatus, is described, in part based on newly collected specimens from the San Juan Basin, housed in the
New Mexico Museum of Natural History. We revise the species-level taxonomy of the genus Arctocyon, which
served as a wastebasket taxon for large arctocyonids for more than 100 years, and we discuss the status of the
North American species previously assigned to the genus Claenodon. Claenodon is a junior subjective synonym
of the genus Arctocyon, originally described from the upper Paleocene of Western Europe. Three species of
North American Arctocyon are valid: A. ferox, A. corrugatus and A. acrogenius. Neoclaenodon montanensis is
a junior subjective synonym of A. corrugatus; Arctocyonides mumak is a junior subjective synonym of A.
acrogenius. Two species of Arctocyon are present in the Torrejonian of the San Juan Basin, New Mexico, A.
ferox and A. corrugatus.
Key words: Arctocyon, Claenodon, Paleocene, San Juan Basin, New Mexico

INTRODUCTION
The genus Arctocyon was originally described from the upper
Paleocene of Western Europe based on a single species, Arctocyon
primaevus (Blainville, 1841). This genus remained monotypic until
1964, when Russell synonymized it and the nearly contemporaneous
North American genus Claenodon. North American Claenodon had been
a waste-basket taxon for all large arctocyonids from the beginning of
the twentieth century. One by one, most of the species previously referred to Claenodon were assigned to several new genera: Neoclaenodon,
Colpoclaenus and Mentoclaenodon. Van Valen (1978) stated that
Claenodon is the synonym of Arctocyonides, rather than Arctocyon,
and referred all the North American species of Claenodon to
Arctocyonides. Williamson (1996) mentioned that Russells conclusions
about the synonymy of Arctocyon and Claenodon were based on a thorough morphological comparison and used Arctocyon for Claenodon
ferox. Gazin (1956) described a very large species of Claenodon, C.
acrogenius from the Tiffanian of Wyoming. Van Valen (1978), without
any discussion, placed this species in the European genus
Mentoclaenodon and described another large arctocyonid from the
Tiffanian, Arctocyonides mumak.
No recent revision of North American species of Arctocyon has
been published, although the genus urgently needs it. The morphology
of the North American species of Arctocyon is still poorly known, as
the last description of A. ferox was published by Matthew (1937). So,
here we describe the cranial and dental morphology of A. ferox and A.
corrugatus and revise the species-level taxonomy of the genus
Arctocyon.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; NMMNH = New Mexico Museum of Natural History, Albuquerque; USNM = National Museum of Natural History, Smithsonian Institution, Washington, D.C.; and YPM-PU =
Princeton University collection at Yale University, New Haven.
Other abbreviations: P upper premolars; M upper molars;
p lower premolars, m lower molars; L maximum tooth crown
length; W maximum tooth width or trigonid width (for m1-3); TW
talonid width.

Subfamily Arctocyoninae Giebel, 1855


Genus Arctocyon Blainville, 1841
Arctocyon Blainville, 1841, p. 73.
Paleocyon Blainville, 1841, p. 73, 112, 121.
Heteroborus Cope, 1880, p. 79.
Hyodectes Cope, 1880, p. 79.
Claenodon Scott, 1892, p. 298.
Arctotherium Lemoine, 1896, p. 342
Neoclaenodon Gidley, 1919, p. 547 (in part).
Claenodon: Matthew, 1937, p. 23 (in part).
Claenodon: Simpson, 1937, p. 174 (in part).
Arctocyon: Russell, 1964, p. 134.
Arctocyonides: Van Valen, 1978, p. 55 (in part).
Mentoclaenodon: Van Valen, 1978, p. 55 (in part).
Claenodon: Winterfield, 1982, p. 98.
Claenodon: Krause and Gingerich, 1983, p. 181.
Arctocyon: McKenna and Bell, 1997, p. 360.
Claenodon: McKenna and Bell, 1997, p. 360.
Mentoclaenodon: McKenna and Bell, 1997, p. 360 (in part).
Neoclaenodon: McKenna and Bell, 1997, p. 360.
Arctocyon: Archibald, 1998, p. 304.
Mentoclaenodon: Archibald, 1998, p. 303.
Type speciesA. primaevus Blainville, 1841.
Included speciesA. primaevus Blainville, 1841; A. ferox
(Cope, 1883); A. corrugatus (Cope, 1883) (= Neoclaenodon montanensis
Gidley, 1919; = Neoclaenodon latidens Gidley, 1919); A. acrogenius
(Gazin, 1956) (= Arctocyonides mumak Van Valen, 1978).
DistributionPaleocene of North America (TorrejonianTiffanian) and Western Europe (Cernaysian).
Revised diagnosisLarge arctocyonids that differ from
Arctocyonides in having less bunodont lower molars with better developed crests; differ from Anacodon in having large lower canines with
well developed shearing surfaces and in the shape of the talonid basin
of the lower molars, which is triangular, not rectangular. The genus
Arctocyon differs from Colpoclaenus and Mentoclaenodon in lacking
the crenulations of the enamel on the upper and lower molars.

SYSTEMATIC PALEONTOLOGY

A. ferox (Cope, 1883)

Family Arctocyonidae Giebel, 1855

Figs. 1-2, 3G-I, 4A-E, G, 7C-G, Tables 1, 2

12
Mioclaenus ferox Cope, 1883, p. 547.
Claenodon ferox: Scott, 1892, p. 298.
Claenodon ferox: Matthew, 1937, p. 23, pp. 376-377, pl. 2, fig. 1.
Claenodon ferox: Simpson, 1937, p. 180, fig. 36.
Arctocyon ferox: Russell, 1964, p. 137.
Arctocyon ferox: Rigby, 1980, p. 109.
Arctocyon ferox: Williamson and Lucas, 1993, p. 125.
Arctocyon ferox: Williamson, 1996, p. 44.
HolotypeAMNH-3268, teeth and fragments of the postcrania
(Figs. 3G-I, 4A-C). These teeth are not demonstrably from the same
individual, so we regard them as syntypes. We restrict the lectotype to
the right m2 (Fig. 4A).
Referred specimensFrom the Nacimiento Formation, San
Juan Basin, New Mexico:
From NMMNH locality 312: NMMNH 1030, right m3, tooth
fragment (isolated); 1116, right p2 partial-isolated; 1566, left P4-isolated; 1567, left M2 partial-isolated; 1644, right M1 partial-isolated;
1742, dP4 isolated; 1764, m? partial-isolated; 2104, left m? partialisolated; 2129, left M3-isolated; 2139, right M3-isolated; 2169, right
p3-isolated; 2179, left m3-isolated; 2184, right M2 paracone; 2186,
right M1 paracone; 2230, left P4 partial-isolated; 2622, left M3-isolated; 2684, left m2-isolated; 803, c partial-isolated; 820, right p? partial-isolated; 964, left m2-isolated.
From NMMNH locality 321: NMMNH 16248, skull fragments,
jaw fragments, tooth fragments.
From NMMNH locality 400: NMMNH 21025, right m1-m2.
From NMMNH locality 672: NMMNH 15368, left M2; 8627,
skull (Figs. 1-2).
From NMMNH locality 1080: NMMNH 15281, left m2 fragment-isolated.
From NMMNH locality 1083: NMMNH 16293, left M2 - isolated.
From NMMNH locality 1152: NMMNH 15785, left p4-m2.
From NMMNH locality 1244: NMMNH 15939, left p3?-isolated;
1244: 15943 right M1 - isolated.
From NMMNH locality 1398: NMMNH 16153, right M1 and
tooth fragment (isolated).
From NMMNH locality 1403: NMMNH 16169, right m2.
From NMMNH locality 1480: NMMNH 16352, right m2 partial-m3.
From NMMNH locality 2557: NMMNH 21589, left p3-m1 partial, right p4 partial-m1 partial, right P2-P3?.
From NMMNH locality 2601: NMMNH 27836, left m1, m2, pt
m3; phalangeal fragments.
From NMMNH locality 2662: NMMNH 21067, left m1 partial,
right m3 partial, premolar fragments.
From NMMNH locality 2690: NMMNH 22038, left p3-m3, right
p4-m3, canine, limbs, partial skeleton.
From NMMNH locality 2699: NMMNH 21577, left m1-m3.
From NMMNH locality 2707: NMMNH 21763, left p4-m3, right
m1-m2.
From NMMNH locality 3544: NMMNH 12281, right m3-isolated.
From Canyon Blanco, San Juan Basin, New Mexico, Baldwin,
1883: AMNH 772, right dentary with p2-m3; 3272, right M2 isolated.
From Chaco Canyon, San Juan Basin, New Mexico, collected by
Baldwin, 1882: AMNH 3269, isolated right P4, M2-3, left p4, m3,
right m3 (partial); 3266, left dentary with p2-3, m1; left dentary with
p4-m2.
From East Fork of Torrejon Arroyo, San Juan Basin, New Mexico:
AMNH 16542, limb and foot bones.
From red beds of Puerco, San Juan Basin, New Mexico, collected by Baldwin, 1882: AMNH 3269, isolated right M1, M3, left
M1, skeletal fragments.

FIGURE 1. Skull of Arctocyon ferox, NMMNH P-8627, dorsal stereophotograph


(A), right lateral (B) and left lateral (C) views.

From Rio Torrejon, San Juan Basin, New Mexico: AMNH 4405,
left dentary with m2-3; 3261, right dentary with p3-m3.
From San Juan Basin, New Mexico, collected by Baldwin: 3940,
isolated left p4, m1, m2 and foot bones; 3260, right dentary with m1-3;
3271, left dentary with p4-m1; 3270, right M2 isolated.
From San Juan Basin, New Mexico, collected by Baldwin, 1884:
AMNH 3262, right dentary with m1-2 (partial).
From Torrejon Arroyo, San Juan Basin, New Mexico: AMNH
16002, right dentary with c1, m1-2 and alveoli for p1-4; 16009, partial
right dentary with p4, m1-2 and m3 talonid; 16007, left maxillary fragment with M1-3; 16006, right maxillary fragment with M1-4, left maxillary fragment with P4-M2; 16004, right dentary with p3-m2 and partial m3; 16005, left dentary with p3-4; 16541, left dentary with p3-m2,
humerus, phalanges; 16003, isolated right P4, M3, left M3, right m3;
3862, right dentary with m2, partial; 3854, left m2 isolated; 3857a,
isolated right P4, m3; 16002, right dentary with p4, m1-3.
From Torrejon Wash, San Juan Basin, New Mexico, Brown, 1896:
AMNH 2457, left dentary with p2, p4-m3; 2466, left dentary with p4,
m3; left dentary with m1-3.
From Torrejon, San Juan Basin, New Mexico, expedition of 1896:
AMNH 2459, left dentary with p3-m3, right dentary with c1, p4-m3
(Matthew, 1937, pp. 376-377, pl. 2)
From between east-west forks of Arroyo Torrejon, San Juan Basin, New Mexico: USNM 15333, left M1-3, right M2; 15335, right m2;
15337, left dentary with m2; 15339, right m1.
From Rio Torrejon, San Juan Basin, New Mexico, collected by
B. Brown, 1898: USNM 5905, left dentary with p4-m3; 407535, right
dentary fragment with p4-m3 (Fig. 4D).

13

FIGURE 3. Selected upper cheek teeth of Arctocyon: A. corrugatus (A-F) and A.


ferox (G-I). A, YPM-PU 17730, right M1-3. B, YPM-PU 16593, right P4-M3. C,
YPM-PU 20394, right DP4-M2. D, AMNH 3258, right P4-M3 (holotype of A.
corrugatus). E, YPM-PU 13756e, left M1-3. F, YPM-PU 14857, left M1-3. G-I,
AMNH 3268, syntypes of Claenodon ferox, right M1 (G), left M1 (H) and left
M3 (I).

FIGURE 2. Skull of Arctocyon ferox, NMMNH P-8627, ventral stereophotograph


(A) and occipital (B) views.

From San Juan Basin, New Mexico, collected by Gazin, 1945:


USNM 407537, left M2.
From San Juan Basin, New Mexico, collected by Gazin, 1949:
USNM 15338, left dentary with p2-3, m3; 407538, right m3; 407539,
right M1-2; 407540, left dentary with p4-m3.
From San Juan Basin, New Mexico, collected by Gazin, 1964:
USNM 407533, left maxillary fragment with M1-3; 407536, maxillary
fragments with right P4-M3 and left P2-M3; 407541, right m2; 407543,
right dentary with m2-3; 407585, left dentary with p4-m3.
From the south rim of the Bison basin, sec. 29, T27W, R 95W,
Fremont Co, Wyoming: USNM 20576, left dentary with partial m2-3;
405047, left dentary with m2-3.
From the small saddle below south rim of the Bison basin, Fremont Co, Wyoming: USNM 21006, right dentary with p3, m3.
From the west end of the Bison basin, sec. 29, T27N, R 95W,
Fremont Co, Wyoming: USNM 20796, right M2-3.
From loc. 13, Crazy Mountain Field, Montana: YPM-PU 14194,
right maxillary fragment with M1-2.
From loc. 20, Crazy Mountain Field, Montana: YPM-PU 13756,
left maxillary fragment with M1-3 (Simpson, 1937, p. 180, fig. 36e);
13755b, right m1 (Simpson, 1937, p. 180, fig. 36b); 13755c, right dentary
with m1-3 (Simpson, 1937, p. 180, fig. 36c); 13755d, right maxillary
fragment with M1-3 (Simpson, 1937, p. 180, fig. 36d).
From loc. 91, Melville: YPM-PU 14622, isolated left M1, right
M2, M3.
From Polecat Bench Quarry: YPM-PU 14257, right dentary with
p1-4, 14258, right dentary with p3-m1 (Fig. 4E).
From Rock Bench Quarry: YPM-PU 17527, right dentary with
m3 (Fig. 4G); 18530, left M2.
Revised diagnosisA species of Arctocyon that differs from A.
corrugatus by its larger size (15-20 %) and from A. acrogenius by
smaller size (25 %); also differs from A. corrugatus by its more robust
lower jaw and shorter postorbital constriction.

DistributionLower Paleocene (Torrejonian) of New Mexico


(Nacimiento Formation) and Montana (Lebo Formation).
DescriptionThe skull (Figs. 1-2) is robust with wide zygomatic arches; its maximum width is at the anterior edge of the orbit, and
the minimum width is posterior to the postorbital constriction. The
snout is longer than the braincase.
The occiput is narrow and well defined by a strong occipital crest.
The condyles embrace the foramen magnum for more than a half of its
height. The condyles widen medially. There are several small, dorsallyfacing foramina in the supraoccipital. The basioccipital and basisphenoid
are flattened, and the tuberculum pharyngeum does not project ventrally.
The parietals form a strong sagittal crest that begins at the postorbital
constriction and fuses with the occipital crest posteriorly.
The frontals are flattened; their suture with the nasals is W-shaped.
There is no postorbital process. The crista frontalis externa begins at the
lacrimal and fuses with its counterpart far anterior to the postorbital
constriction.
The maxilla is excluded from the orbital ring by a well developed
lacrimal bone and the maxillary process of the jugal. The palatine process
is almost round in shape. The processus zygomaticus is weak and doesnt
form a surface for the attachment of the masseter muscle.
The premaxilla does not contact the frontal. The premaxillary
bones project anteriorly; they contact the nasals posteriorly and the
maxillaries laterally. The nasal bones are large, widening posteriorly.
The squamosal forms the lateral wall of the braincase. The zygomatic process is shifted laterally and forms more than one-third of the
zygomatic arch. The long axis of the glenoid fossa is not perpendicular to
the long axis of the skull. The postglenoid process is well developed, tall
and forms the posterior wall of the glenoid fossa, which is slightly concave and widens laterally.
The lacrimal bone is rather large and lies in the anterior corner of
the orbit. The tuberculum lacrimale is not developed. The foramen lacrimale
is situated in the orbital rim and faces posterolaterally.
The jugal bone is slightly reduced, and the zygomatic arch is much
weakened. The jugal bifurcates at the anterior end; it doesnt reach the
glenoid fossa posteriorly.
The pterygoid bones are small. The palatine forms part of the
floor of the orbit, where it contacts the frontal and lacrimal bones and
reaches the orbital fissure posteriorly. The palatine bones form most of

14
TABLE 1. Lower tooth measurements of A. ferox (mm).
Coll. #
AMNH-772
AMNH-2457
AMNH-2459
AMNH-2459
AMNH-2460
AMNH-2466
AMNH-2466
AMNH-3260
AMNH-3261
AMNH-3262
AMNH-3266
AMNH-3266
AMNH-3268
AMNH-3269
AMNH-3271
AMNH-3854
AMNH-3857
AMNH-3862
AMNH-3940
AMNH-4405
AMNH-16001
AMNH-16002
AMNH-16003
AMNH-16004
AMNH-16005
AMNH-16009
AMNH-16010
AMNH-16541
NMMNH-803
NMMNH-964
NMMNH-1030
NMMNH-1116
NMMNH-2169
NMMNH-2179
NMMNH-12281
NMMNH-15785
NMMNH-15939

cL

cW

10.5
10

5.8
6.8

p2 L p2 W
5.9
2.9
5.1
2.8

p3 L

p3 W
4.9

10.2

4.8

p4 L
10.9
11.1
12.9

p4 W
7.6
6.1
6.6
6.7

m1 L
11.6
10.6
10.8

m1 W m1 TW
8.3
7.8
7.9
8.6
7.5
8.6
8.7

m2 L m2 W
12.6
12.7
12.9

10.1
10.3
10.4

m2 TW
9.5
10
10.2
10.2

13.4
8.1

3.2

9.8

5.9

9.8

4.3

11

6.1

11.2
10.9

5.8
6.3

10.8
12.5
11.9
10.8
11.1

11.1

6.9
8.2
8
8.2
7.1

7.8

8.3
9.3
8.7
8.7

12.7
13

9.2
10.9

10.1
10.9

12.1
14.3

9.5
9.9

10.1
10.9

10.9

9.3

8.3

9.9
9.9
9.2
7.8

11.2
10.3
10.4
9.4
9.3

12.2

7.8

9.2

10.4

5.9

10.3

7.1

8.2

13.8
11
12.1

1.1
11.9
10.6

6.9
5.3
6
6.2
6.4

10.4

7.4

9.3

11.8

10.3

11.2

11.7
12.1
9.8

7.3
7.3
7

9.4
8.2
8.2

12

9.8
9.9
8.9

10.1
9.2

8.9
10.1
9.1

7.8

12

5.7
4.9

4.9
7.9

5.1

10.6

13.29 9,72
8.17

m3 W
8.9
8.7
9.2
9.2

m3 TW
8
7.2
8.9
8.7

12.9
13.7
12.8
12.4
14.4

9
9.1
8.6
9.2
8.9

8.3
8.7
7.9
8.5
7.8

14.5
12.1

8.9
8.2

8
7.6

13.7

8.6

7.9

14.6
10.6
12.3
12.1

9.8
8.7
8.5
8.4

7.9
7.9
7.5
7.2

12.3

7.2

9.2

13.6
6.6

m3 L
12
11.3
13.4
12.9

6.3

10.02
11.89

7.96

7.44

10.84
11.78
13.33

7.28
8.5

7.06
7.45
6.17

3.99
9.65

5.57

6.83
8.5

11.14

7.22

8.87

9.75

4.51

the hard palate; their length is about half of the length of P3-M3. The
choanal notch is rectangular, without a median process.
The infraorbital foramen is large, oval in shape and faces anteriorly. The infraorbital canal is rather short. The foramen maxillare is
larger than the foramen infraorbitale and faces posterolaterally, so the
canal itself is slightly curved. The foramen sphenopalatinum is united
with the foramen palatinum caudale in a single depression, and both
foramina face posterolaterally.
The canalis alaris is short; its anterior foramen is situated near
the posterior edge of the orbital plate of the palatine bone. The foramen
alaris anterius faces anterodorsally, and the foramen alaris posterius
faces posteroventrally, so the long axis of the canal is inclined at about
15o to the horizontal plane of the skull.
The fissura orbitalis is situated anteriorly and above the foramen alaris anterius; it faces anterodorsally. The lateral wall of the fissure is formed by the alisphenoid bone and the medial by the orbitosphenoid. The foramen opticum is situated anteriorly and above the
orbital fissure, in the orbitosphenoid bone. The foramen is rather large
and faces anteroventrally. The foramen ethmoideum is situated in the
frontal bone within the orbit, it is rather large and faces ventrally. The
foramen ovale is situated posterolaterally to the foramen alaris posterius
in a small depression at the base of the zygomatic process of the
squamosum; it faces dorsally. The postglenoid foramen is rather large
and is situated behind the medial wall of the foramen mandibularis; it

faces ventrally. The foramen lacerum anterior is at the anterior edge of the
petrosal, at the same level as the postglenoid foramen, but located medially.
The lower jaw is elongate and curved, with a procumbent rostrum due to the large canines. The vertical ramus is about one-quarter
of the length of the horizontal ramus. The angular process is well developed, round in shape and projects posteroventrally. There is a semilunar notch anterior to it. The articular process is robust, and the neck
is the same width as the condyle, slightly tapering anteriorly. The condyle
is cylindrical in shape and slightly constricted in the middle. The coronoid process is tall and wide. The incisura mandibulae is shallow because the articular process projects more posteriorly than dorsally. The
masseteric fossa is deep and well defined. The foramen mandibularis
is situated on the medial side of the lower jaw, under the coronoid
process, on the level of the tooth row. It faces posteriorly and slightly
ventrally. There are two pairs of mental foramina on each side of the
jaw. The anterior is under the p2, and the posterior is under the p3.
The upper canines are large and curved, with sharp, shearing
posterior edges. The upper premolars are simplethe P1 is singlerooted, the P2 is double rooted, and the P3 is three-rooted with a small
conule anterior to the paracone. The P4 is slightly molarized and has a
triangular shape, a well developed protocone, tiny parastyle and a metacone.
The upper molars are large with well defined hypocones. The M2

15
Table 1 continued
Coll. #
cL
NMMNH-16169
NMMNH-16248
NMMNH-16248 11.56
NMMNH-16352
NMMNH-21067
NMMNH-21258
NMMNH-21258
NMMNH-21577
NMMNH-21763
NMMNH-21763
NMMNH-22038
NMMNH-22038
NMMNH-27836
NMMNH-29412
NMMNH-30539
USNM 5905
USNM 15335
USNM 15337
USNM 15338
USNM 15339
USNM 20576
USNM 21006
USNM 407538
USNM 407540
USNM 407541
USNM 407543
USNM 407585
YPM-PU 13755b
YPM-PU 13755c
YPM-PU 14258
YPM-PU 17527
PU-1704
YPM-PU 13215
Mean
10.41
Number of spec. 5
Stand. Deviat.
1.69

cW

p2 L p2 W

p3 L

p3 W

p4 L

p4 W

m1 L

12.01 6.3
12.09 6.67

7.4

9.7

5.04

8.38

9.08

12.19 6.74
14.43 7.13

9.87
11.11
11.4
11.3
11.28

8.32
6.58
7.48
7.18
7.66
7.69

7.55
8.86
8.73
9.31
9.28

11.4

11.1

8.6

9.2

11.8

7.2

8.2

11.78 6.61

11.79 6.36

9.53

5.9

3.4

9.2

4.96

m1 W m1 TW m2 L m2 W
12.56 9.62
7.28 8.53
13.55 10.4

6.3

m2 TW m3 L
9.72
10.18
12.23

10.96
11.87 7.43

9.46

11.05
11.87
12.3
12.23
12.61
12.8
14.65
13.27
12.3
12.3
13.2

8.64
10.34
10.33
10.54
10.95
9.8
11
10.49
10.7
9.8
9.9

10.96
8.59
9.92
10.09
10.43
10.27
10.51
10.17
10.11
10.4
9.3
9.7

4.2

8.65

7.55

13.19
12.09
11.48

8.9
6.88
8.91

8.09
5.29
7.23

11.44
12.08

7.5
8.83

6.55
7.25

14.43
14.4
12.98

9.49
9.52
8.84

7.87
7.75
6.45

12.1

7.8

12.9

8.6

7.8

13.1
13.4
12.9

8.6
9
9.4
8.9

7.1
7.6
7.5
7.9

12.5
11.22

7.7
7.94

6.8
7.19

11.4

8.3

7.6

12

7.9

6.1

15.1
12.67
40
1.09

9.8
8.6
41
0.74

8.1
7.53
41
0.72

9.2
10.8

4.9
8.9

8.8

6.8
5
1.5

6.21
5
1.15

3.26
5
0.47

9.47
13
0.77

4.1

4.92
14
0.54

11.34 5.68

10.39
12.1

6.76
7.6

11.5

10.1

6.6

5.5

11.06 6.45
20
26
2.52 0.62

is the largest molar, and M3 is the smallest. The M1 is square; the


hypocone is nearly the size of the protocone, and the metaconule is
larger than the paraconule.
The M2 is rectangular, with a well-developed hypocone and a
metaconule equal in size to the paraconule. The hypocone of M3 is
vestigial or absent, the paraconule is distinct, and there is no metaconule.
The lower canines are large and oval in cross section. The first
lower premolar is simple and single-rooted; there is a short diastema
between the p1 and the p2. The p2 is double-rooted and single-cusped.
The p3 is simple and single-cusped, but the posterior cingulid forms a
tiny talonid; there is no precingulid. The p4 is slightly molarizedthe
protoconid is the largest cuspid, the precingulid is small (but forms a
vestigial anterior cuspid), and the talonid is small, with two tiny cuspids.
The lower molars are rectangular or square in shape. The m1 is
square, and the talonid is longer and wider than the trigonid, but lower.
The protoconid is equal to the metaconid in size, and the paraconid is
slightly reduced and shifted medially, smaller than the other two cuspids. There are two cuspids on the talonid: the larger hypoconid and the
smaller entoconid. The cingulid is developed on the anterior, posterior
and labial sides of the tooth. The talonid basin is closed. The m2 is
slightly larger than the m1, but doesnt differ significantly in morphology. The trigonid is equal in length to the talonid. The paraconid is more

11.11
27
0.70

7.5
29
0.55

7.89
9.3
9.7
7.8

8.72
30
0.55

12.8
13.3
13
11.22

9.8
9.7
9
8.84

10.4
9.9
9.1
8.9

12.2

10.5

9.9

16.5

14.8

13.6

12.67 9.8
37
38
1.11 1.10

10.1
38
0.89

m3 W m3 TW

reduced than on the m1. The m3 is elongate, oval in shape, and the
paraconid is less reduced than on the other molars. The talonid is triangular in shape, much longer, than the trigonid; it bears a small hypoconulid,
which lies close to the entoconid and is united with it by a strong cristid.
The talonid basin is closed.
Comparison to A. primaevusHaving skulls of both North
American and European representatives of the genus allows us to provide a comparison of cranial morphology of these two large species of
Arctocyon. The description of the skull of A. primaevus can be found in
Russell (1964).
The skull is more elongated in A. primaevus, with narrower
zygomatic arches and a taller sagittal crest. The structure of the zygomatic arch is very similar in both species: the zygomatic processes of
the squamosal are extremely well-developed and form most of the arch;
the jugal is slightly reduced and bifurcated anteriorly; it does not reach
the glenoid fossa posteriorly (apomorphic character).
The hamulus of the pterygoid bones is well-developed in A.
primaevus, but this part is absent in the A. ferox skull. The infraorbital
foramen is relatively smaller in A. primaevus, and the infraorbital
canal is longer than in A. ferox.
Important similarity is observed in the structure of the foramen
sphenopalatinum, which is united with the foramen palatinum caudale
in a single depression. All species of Arctocyon (the condition for A.
acrogenius is unknown) have a short alisphenoid canal (apomorphic

16
TABLE 2. Upper tooth measurements of A. ferox (mm).
Coll. number C L
AMNH-3259
AMNH-3268
AMNH-3268
AMNH-3269
AMNH-3270
AMNH-3272
AMNH-3852a
AMNH-16003
AMNH-16003
AMNH-16006
AMNH-16006 7.6
AMNH-16007
NMMNH-817
NMMNH-1566
NMMNH-2129
NMMNH-2230
NMMNH-2239
NMMNH-2622
NMMNH-8627
NMMNH-8627
NMMNH-15368
NMMNH-15943
NMMNH-16153
NMMNH-16248
NMMNH-16293
NMMNH-21589
NMMNH-30700
USNM-20796
USNM-407533
USNM-407533
USNM-407536
USNM-407536
USNM-407537
USNM-407539
YPM-PU 13755
YPM-PU 13755d
YPM-PU 14194
YPM-PU 18530
Mean
7.6
Number of spec. 1
Stand. Deviat. 0

CW

P2 L

P2 W

P3 L

P3 W

P4 L

P4 W

M1 L

M1 W

11
10.8

11.4
11.5

9.6

11.4
12.5
12.5

9.9
10.8

8
9.3

8.9

9.2

5.4
8.18

M2 L
10.8

10.05
10.1
10.2

11.2
11.1
12.7

10.3
10.3
11.1

M2 W
13.4

M3 L

M3 W

9.1

11.6

8.7

12.2

8.6
7.7

11.2
11.3

7
7.9

9.7
12.1

7.18

9.78

7.23
8.01
9
9.4

10.46
11.02
11.99
12.14

8.75

11.38

7.4
7.9

11.7
11.3

8.4
8.5

10.9
10.9

7.4
7.5

11.6
11.8

8.09
18
0.73

11.28
18
0.736

15.3
13.6

13.1
13.4
14.9

6.3
9.78

9.14
9.18

9.04
8.95

5.8

5.4
1
0

5.8
1
0

2.9

2.9
1
0

6.18
5.85

9.77
9.95

9.99

7.11

10.27

9.81

6.91

10.51

8.6

9.1
6
0.7

6.8

6.525
6
0.488

9.4
9.5

9.853
11
0.532

character), and the arrangement of fissura orbitalis and foramen opticum


is the same. The three species of Arctocyon (the condition for A.
acrogenius is unknown) share the following symplesiomophy: the
postglenoid foramen is large. In all these species the foramen ovale is
situated in a depression at the base of the zygomatic process. Position
and shape of the mental and mandibular foramina is similar in A. ferox,
A. primaevus and A. corrugatus.
Inner ear structure is similar in the two species. The
promontorium is slightly more bulbous in A. ferox, than in A. primaevus,
the sulcus of the promontorium is deeper in A. primaevus. The sulcus
for the facial nerve (VII) is deep in both species.
CommentsMost characters of the skull of Arctocyon are
plesiomorphic for eutherian mammals and not typical of later ungulates. But, among other arctocyonids, Arctocyon is a relatively advanced
genus. Unlike the other archaic ungulates, in Arctocyon the crista frontalis externa, beginning from the lacrimal bone, extends medially and
fuses with the same crest from the other side of the skull far ahead of
the postorbital constriction. This elongates the sagittal crest, which is
the place of origin of the temporalis muscle, which was dominant over

8.69
8.58

9.76

9.2
9.3

9.035
10
0.489

10.48
10.56

11.23
12.01

10.93
11.08
10.37

11.2
11.76
11.26

10.32
10.84

12.62

11.5

12.5

9.9
10.7

11.9
11.01

11.1
10.6
10.5

12.2
12.9
11.9
11.8

10.61
18
0.422

11.79
18
0.599

11.05
11.45
12.14

15.13
14.96
15.05

11.99
10.61
11.33

15.2
14.09

10.8
11.2
11.9
11
11.1
10.9
13.1
11.4
11.4
10.1
10.3
11.28
24
0.767

15
15
14
13.6
14.1
14.1
16
14.1
14.9
13.2
13.1
14.33
22
0.846

the masseter muscle in mastication in these animals.


Arctocyon corrugatus (Cope, 1883)
Figs. 3A-F, I-K, 4H-J, 6, 7A-B, Tables 3-4
Claenodon corrugatus Cope, 1883, p. 556.
Claenodon corrugatus: Cope, 1884, p. 349, fig. 16.
Neoclaenodon montanensis Gidley, 1919, p. 547.
?Neoclaenodon latidens Gidlley, 1919, p. 554.
Claenodon corrugatus: Matthew, 1937, p. 36, 378-385, pl. 3, figs. 1-3,
pl. 4, figs. 1-4, pl. 5, figs. 1, 2, pl. 6, figs. 1, 2.
Cllaenodon latidens: Simpson, 1937, p. 187, fig. 41.
Claenodon montanensis: Simpson, 1937, p. 181, figs. 37-39.
Arctocyonides montanensis: Van Valen, 1978, p. 55.
Claenodon cf. montanensis: Krause and Gingerich, 1983, p. 181, fig.
19 A, B.
HolotypeAMNH-3258, right maxillary fragment with P4-M3
(Fig. 3D).
Referred specimensFrom the south rim of the Bison Basin,

17

FIGURE 5. Bivariate plot of m2 length and width in North American Arctocyon


species.

FIGURE 4. Selected lower cheek teeth of Arctocyon: A. ferox (A-E, G) and A.


corrugatus (F, H-J). A-C, AMNH 3268, syntypes of Claenodon ferox, right m2
(A), left m3 (B) and right m3 (C). D, USNM 407535, right p4-m3. E, YPM-PU
14258, right p3-m1. F, YPM-PU 20394, right m2-3. G, YPM-PU 17527, right m3.
H, YPM-PU 18771, right m2-3. I, YPM-PU 18317, left m1-2. J, USNM 8388,
right m1-3 (holotype of Claenodon latidens).

sec. 29, T27W, R 95W, Fremont Co, Wyoming: USNM 20573, left
dentary with m1-2 ; 20574, left dentary with m2-3; 405046, left m2.
From the east side of the east pocket of Arroyo Torrejon, San
Juan Basin, New Mexico: USNM 15344, left dentary with p4-m3.
From Gidley quarry, Sweetgrass Co., Crazy Mountain Field,
Montana: USNM 8362, skull (Fig. 6), dentaries with most dentition
and postcranial fragments (Simpson, 1937, p. 183-184, figs. 37-39);
8386, right dentary with m2-3; 8388, right dentary with m2-3 (Fig. 4J;
holotype of Neoclaenodon latidens); 13781, left M2-3.
From Aurora locality, upper Polecat Bench Formation, NW 1/4,
sec. 26, T58 N, R101W, Park County, Wyoming: YPM-PU 18771, left
dentary with m1-3, right dentary with m2-3 (Fig. 4H).
From Big Sand Coulee, NW 1/4, sec. 33, T57 N, R100W, Park
County, Wyoming: YPM-PU 18440, right dentary with m1-2.
From east of Polecat Bench Formation, 1 mile SW of camp # 3,
SW 1/4, sec. 11, T56N, R99W, Park County, Wyoming: YPM-PU 13204,
left maxillary fragment with M1-3.
From East of Sand Covice NE 1/4, sec. 30, T57 N, R100 W, Park
County, Wyoming, Polecat Bench Formation: YPM-PU 18918, right
dP4-M1.
From loc. 3, Melville: YPM-PU 13756e, left maxillary fragment
with M1-3 (Fig. 3E).
From NE 1/4, sec. 28, T58N, R101W, Park County, Wyoming:
YPM-PU 20394, right dentary with m2-3 (Fig. 4F), right maxillary
fragment P4-M2 (Fig. 3C).
From Big Sand Coulee, NW 1/4, sec. 33, T57 N, R100W, Park
County, Wyoming: YPM-PU 20280, right dentary with m3.
From Polecat Bench Formation, NE 1/4, sec. 28, T58 N, R100W,
Park County, Wyoming: YPM-PU 18996, right maxillary fragment with
M1-2.
From Polecat Bench Formation, mile south of Canyon road,
Park County, Wyoming: YPM-PU 18317, left dentary fragment with
m1-2 (Fig. 4I).
From Polecat Bench Formation, NE of Big Sand Coulee, NE 1/
4, sec. 20, T57 N, R100W, Park County, Wyoming: YPM-PU 18925,
left dentary with m1-2, right M2, left M3.

From Polecat Bench Formation, SE 1/4, sec. 36, T55N, R96W,


Park County, Wyoming: YPM-PU 14857, left maxillary fragment with
M1-3 (Fig. 3F).
From Polecat Bench Quarry: YPM-PU 14020, right dentary with
p4-m1.
From Polecat Bench Quarry, sec. 36, T57N, R99W: YPM-PU
16593, right maxillary with P4-M3 (Fig. 3B).
From Rock Bench Quarry: YPM-PU 17730, right maxillary fragment with M1-3 (Fig. 3A).
Revised diagnosisA species of Arctocyon about 15-20 %
smaller than A. ferox. Differs from the Arctocyon acrogenius in being
40 % smaller.
DescriptionThe skull and lower jaw of Claenodon
montanensis (USNM 8362) have been illustrated by Simpson (1937),
accompanied by the skull description by Gidley. Here, we provide additional information on skull characters of this species.
The skull is similar to that of A. ferox, but is relatively more
elongate and not as robust. The anterior portion of the skull (USNM
8362) is heavily covered with plaster, which precludes accurate description. The premaxilla projects posteriorly between the canines on
the ventral side of the skull. The nasals are elongated, widening rostrally. The infraorbital foramen faces anteriorly and is situated above
the posterior root of the P3.
The sagittal crest is extremely tall, the occipital crest is relatively well developed, the occiput is narrow, triangular in shape. The
foramen magnum is rhomboid in shape. The condyles surround only
half of the foramen magnum. The basioccipital bears a weak ventral
crest that continues onto the basisphenoid. There are two well-expressed
tuberosities along the crest for the insertion of the musculus longus
capitis. The well-developed basioccipital crest is also the evidence of
the relatively strong development of this muscle (Anton and Galobart,
2000). The degree of development of the musculus longus capitis is
directly correlated with the strength of vertical head movement and
suggests A. corrugatus possessed a strong ventral bite, which correlates well with the size of the canines and tall sagittal crest. The tuberosities are also more shifted anteriorly in A. corrugatus, which may
suggest that this species was more carnivorous than the other, more
robust representatives of the genus.
Several foramina are preserved in the basicranium of USNM
8362. The fissura orbitalis is situated very low in the posterior part of
the orbit and faces anteriorly. A deep groove on the medial wall of the
orbital wall lies anterior to the orbital fissure. The foramen opticum is
situated directly above the orbital fissure and faces anterolaterally. The
foramen rotundum is situated posterior and lateral to the orbital fissure, it is perfectly round in shape and faces anteriorly. Its position
apart from the alisphenoid canal is a primitive feature for ungulates. The

18
The lower jaw is relatively tall. The coronoid process, judging
from its base, was rather high and oriented almost at 900 to the horizontal plane of the jaw. The anterior and dorsal edges of the coronoid
process are thickened and form a strong crest for temporal muscle attachment. This crest extends anteriorly on the lateral side of the jaw to
the posterior root of the m1. The masseteric fossa is very deep and
triangular in shape, its base thickened for the attachment of the masseteric muscle. The angular process is long, slightly curved dorsally and
medially as in other Arctocyon species. The foramen mandibulare is
situated relatively high, at the level of the cheek tooth row. It faces
posterodorsally. The posterior mental foramen is situated in a shallow
depression under the posterior root of the p3, and faces anterolaterally.
No anterior dentition is preserved. The upper canines are strong,
oval in cross-section and slightly curved posteriorly. The enamel is not
serrated, and there is a very weak posterior crest.
The P1 is situated close to the canine and has one root. The P2 is
relatively small, double-rooted and laterally compressed. It bears two
cusps large paracone and a small metacone. The cingulum is not
developed. The P3 is much larger, has three roots and is triangular in
shape. The paracone is the largest cusp; there is a small parastyle, vestigial protocone and larger metacone. The cingulum is complete. The
P4 is similar in structure to the P3, but considerably larger.
The M1 is subquadrate in shape. The paracone and metacone
are of equal size, the protocone is much larger. There is a small parastyle.
The cingulum is complete and forms a vestigial hypocone on the
posterolingual slope of the tooth. The M3 is much reduced, and almost
oval in shape. Its long axis is perpendicular to the long axis of the
skull. The protocone is the largest cusp; the metacone is much reduced
and shifted lingually. The cingulum is complete and there is no trace of
hypocone.
The lower canines are large. The p2 is double-rooted. The p3 is
double-rooted with a tall crown. There is a small talonid that bears a
pointed hypoconid. The cingulid is weak. The p4 differs from the p3 in
being larger with a strong cingulid, well developed paraconid and larger
talonid.
The m1 is the smallest of the molars. The trigonid is triangular
in shape with an unreduced paraconid. The talonid is slightly shorter
than the trigonid, basined; the hypoconulid is vestigial. The m2 is larger
than the m1, with a considerably reduced paraconid that is pressed to
the metaconid. The talonid is relatively and absolutely wider and deeper
than on m1. The hypoconulid is vestigial. The m3 is narrow and more
elongate than the m1-2. The paraconid is almost completely reduced.
The hypoconulid is large and posteriorly projecting. There is an entocristid

FIGURE 6. USNM 8362, holotype skull and lower jaw of Claenodon


montanensis, in lateral (above) and ventral (below) views (after Simpson, 1937).

alisphenoid canal is relatively short. Its anterior foramen is situated


posterior to the foramen rotundum. The posterior foramen is situated
very close to the foramen ovale, so this part of the basicranium is relatively short. The foramen ovale is large, its longs axis is at about 1350 to
the long axis of the skull. It is situated on the level of the glenoid fossa.
There is a strong crest that separates it from the posterior part of the
basicranium. The postglenoid foramen is large, situated posterior to the
glenoid fossa. It faces posteroventrally. The glenoid fossa is relatively
shallow and the postglenoid process is weak. The foramen lacerum anterior is shifted anteriorly, so it lies at the level of the posterior part of the
glenoid fossa. Only one more foramen can be observed on the basicranium
the hypoglossal foramen, which is situated at the anteromedial part of
the occipital condyle.

TABLE 3. Lower tooth measurements of A. corrugatus (mm).


Coll. number
NMMNH-21025
USNM-8362
USNM-8386
USNM-20573
USNM-20574
USNM-405046
YPM-PU 14020
YPM-PU 14257
YPM-PU 18440
YPM-PU 18771
YPM-PU 18925
YPM-PU 18996
YPM-PU 20280
YPM-PU 20394
Mean
Number of spec.
Stand. Deviat.

p2 L

5.2

5.2
1
0

p2 W

2.5

2.5
1
0

p3 L

p3 W

p4 L

p4 W

7.7

3.6

9.6

4.9

8.1

7.9
2
0.3

4.1

3.85
2
0.35

10.5
11.1

10.4
3
0.75

5
5.2

5.03
3
0.15

m1 L
9.17
9.1

m1 W
5.71
6

m1 TW
5.83

9.3

5.9

6.2

9.4

6.1

7.3

8.9
8.8
8.5

6.7
6.4
6.4

7.3
7.5
6.9

9.02
7
0.31

6.173
7
0.343

6.8383
6
0.6771

m2 L
9.97
8.6
10.15

m2 W
7.1
8.45

m2 TW
7.88
7.3
8.22

9.7
10

7.2
8.2

6
8

9.8
9.8
9.2
8.5

8.2
8.5
7.8
6.8

7.4
8
7.7
8

9.6
9.532
10
0.579

9.3
7.95
9
0.801

8.1
7.66
10
0.6546

m3 L

m3 W

9.4
9.91

6.7
7.59

m3 TW
7.04
5.7
6.2

9.4

5.2

9.3

7.4

6.3

10.1
9.8
9.65
6
0.33

6.5
7.1
6.882
6
0.596

5.4
6.1
5.9914
7
0.6206

19
The number of specimens now known, many collected since the
work of Matthew and Simpson, is enough for meaningful comparison.
They indicate that A. corrugatus is a separate species from A. ferox
that differs from the latter only in size and in having absolutely and
relatively less robust molars. On the other hand, A. corrugatus is indistinguishable from the similar arctocyonid from Montana, Claenodon
montanensis, described by Gidley (1919). As A. corrugatus has priority, A. montanensis is here considered a subjective junior synonym of
A. corrugatus. There are clearly two size groups of large Arctocyon
species from the Torrejonian - Tiffanian of New Mexico and Montana,
representing two species A. ferox (larger species) and A. corrugatus
(smaller species).
A. acrogenius (Gazin, 1956)
Claenodon acrogenius Gazin, 1956, p. 34, pl. 7, figs. 1, 6.
Arctocyonides mumak Van Valen, 1978, p. 55, pl. 3, fig. 1.

FIGURE 7. A-B, holotype of Claenodon latidens, USNM 8388, right dentary


fragment with m1-3, in occlusal (A) and labial (B) views. C-G, selected teeth of
Arctocyon ferox; C, USNM 6156, left m2; D, YPM-PU 13755, right m1; E, YPMPU13755, left m1-3; F, YPM-PU 13755, right M1-3; G, YPM-PU 13756, right
M1-3. (All after Simpson, 1937).

connecting the entoconid to the posterior wall of the metaconid that


closes the talonid basin. The cingulid is developed only on the labial wall
of the molars.
DistributionLower Paleocene (Torrejonian) of New Mexico
and Paleocene (Torrejonian - Tiffanian) of Montana (Gidley, 1919;
Simpson, 1937, Krause and Gingerich, 1983).
CommentsCope (1883) described Claenodon corrugatus as
a species differing from Claenodon ferox in its smaller size and less
robust teeth. Matthew (1937) treated C. corrugatus as a separate species, but he mentioned that it may be not valid due to its similar morphology and minor difference in size compared to C. ferox. Simpson
(1937) discussed the validity of this species and stated that it differs in
size too little to represent a species separate from A. ferox, although he
wrote that the number of A. ferox specimens available does provide a
meaningful comparison. So, in a comparative table, Simpson (1937, p.
177, table 37) included two specimens of A. ferox and four specimens
of A. corrugatus.

Holotype USNM 20634, right dentary with p1, m1-2.


Referred specimensFrom 2 miles SW of camp #3, SW 1/4,
sec. 11, T56N, R99W, Park County, Wyoming: YPM-PU 13215, right
maxillary fragment with P4-M1, M2, right m3; YPM-PU 17406, left
dentary with p4-m3 (Van Valen, 1978, p. 55, pl. 3, fig. 1).
Revised diagnosisDiffers from the A. ferox and A. corrugatus
in its larger size (25 and 40 % respectively). The crowns of the upper
teeth have a more square shape than in A. ferox. The cusps are low, and
the paracone and metacone are close to each other. The p2 is greatly
reduced or absent.
Measurements YPM-PU 13215: the m2 length 16.5 mm,
the m2 trigonid width 14.8 mm, the m2 talonid width 13.6 mm;
YPM-PU 1704: the m3 length 15.1 mm, the m3 trigonid width 9.8
mm, the m3 talonid width 8.1 mm; YPM-PU 13215: the P4 width
11.6 mm; the M1 length 10.6 mm, the M1 width 12.6 mm, the M2
length 13.7 mm, the M2 width 17.1 mm.
DistributionUpper Paleocene (Tiffanian) of Wyoming.
Comments Gazin (1956) assigned this species to Claenodon
and noted that the p2 was either very small or absent in this species a
tendency seen in most large arctocyonids. Van Valen (1978), without
any discussion, placed this species in the European genus
Mentoclaenodon. The lower molar morphology does not confirm this
allocation, because Claenodon arcogenius lacks the crenulations of
the enamel that are so typical for Mentoclaenodon. Instead, lower jaw
and molar morphology are very similar to Arctocyon ferox and A.
corrugatus. Van Valen (1978) described a large arctocyonid from the
Tiffanian of Wyoming, Arctocyonides mumak, based on a dentary with
p4-m3, and mentioned the similarity of this species to Arctocyon. Indeed, the morphology of the lower molars in the holotype dentary is
very close to A. ferox and A. primaevus. When treated as Arctocyon, this

TABLE 4. Upper tooth measurements of A. corrugatus (mm).


Coll. number
USNM-8362
USNM-8363
AMNH-3258
USNM-13781
YPM-PU 13204
YPM-PU 13756e
YPM-PU 14857
YPM-PU 16593
YPM-PU 17730
YPM-PU 18918
YPM-PU 20394
Mean
Number of spec.
Stand. Deviat.

P2 L
4.4

4.4
1
0

P2 W
2.1

2.1
1
0

P3 L
7.4
7.52

7.46
2
0.08

P3 W
5.3
5.12

5.21
2
0.127

P4 L
7.3
6.59
9

P4 W
7.7
7.82
8.1

7.7

8.6

8.2
7.76
5
0.91

8.1
8.064
5
0.347

M1 L
9.1
7.96
9.1

M1 W
10.3
9.4
11.1

9.5
9.2
8.3
8.7
9.6

11.4

8.6
8.9
9
0.55

9.1
9.7
10.5
10.6
10.7
10.31
9
0.77

M2 L

M2 W
11.81
11.6
12.6
12.2
12.5
12.1
12.7
13.3

M3 L
5.2
5.7
7.3
6.79
7.3
7.7
6.1
6.6
7.6

M3 W
8.1
8.99
10.5
10.17
10.2
10.6
8.5
10.4
12.1

8.23
9.9
8.9
9.7
9.7
8.6
9.9
10.4
9.1
9.38
9
0.71

12.9
12.41
9
0.538

6.7
9
0.88

9.951
9
1.23

20
species is indistinguishable in size or morphology from A. acrogenius.
Thus, the distribution of the genera Arctocyonides and Mentoclaenodon
is restricted to Western Europe.
We referred YPM-PU 13215, represented by a lower m3 and a
maxillary fragment with P4-M2, to A. acrogenius. The upper teeth of
this species were unknown before. The associated uppers and lower
m3 exceed in size any of the known specimens of A. ferox, so we assign
this specimen to A. acrogenius. The morphology of both upper and
lower teeth is very close in A. ferox and A. acrogenius, so these species
can be distinguished mostly by size.
DISCUSSION
Van Valen (1978) stated, without discussion, that Claenodon
Scott, 1892 is a junior subjective synonym of Arctocyonides Lemoine,
1885, rather than Arctocyon, as was concluded by Russell (1964) based
on a detailed description. Simpson (1937) noted that Neoclaenodon,
described by Gidley (1919), is a synonym of Claenodon and assigned
Neoclaenodon silberlingi Gidley, 1919 and Neoclaenodon vecordensis
Simpson, 1935 to Claenodon.
Thorough comparison of A. ferox with Arctocyon primaevus and
Arctocyonides species shows that the North American species is much
closer in dental morphology to A. primaevus. The major differences are
in the overall shape and crown structure. Arctocyonides has square and
rectangular upper and lower molars, whereas the molars in A. primaevus
and A. ferox are rounded. The conules on the upper molars of
Arctocyonides are much better developed than in Arctocyon species.
The lower tooth cuspids of Arctocyon are united by well developed
crests, and the teeth appear to be lophodont, even in a slightly worn
state. The lower teeth of Arctocyonides are much more bunodont and
do not lose the bunodont structure, even on much worn teeth. There are
other minor differences, such as stronger upper molar cingula and narrower lower premolars in Arctocyonides, that differentiate that genus
from all Arctocyon species. A. ferox and A. primaevus share most dental

features and differ only in slight details. So, considering the very similar
morphology of Claenodon ferox and Arctocyon primaevus and the
differences in morphology between them and the Arctocyonides species,
we conclude that Claenodon is a junior subjective synonym of Arctocyon,
rather than Arctocyonides.
The analysis of the skull and jaw morphology of Arctocyon species reveals that they have several features of carnivorous mammals.
They are: strong development of the canines with well-defined shearing surfaces, strong sagittal and occipital crests and a high coronoid
process of the lower jaw. On the other hand, Arctocyon species have
bunodont molars, which were not adapted for shearing meat. Of all the
species of Arctocyon, A. corrugatus appears to be more carnivorous
than the others. Several morphological features support this assumption: (1) A. corrugatus has taller sagittal and occipital crests than other
Arctocyon species; (2) the insertion facets for the longus capitis muscle
are much better developed in this species than in larger Arctocyon,
which suggests stronger ventral bite with the canines; (3) the insertion
facets for the longus capitis muscle are shifted anteriorly, which provides more strength during the ventral bite; (4) the premolars in A.
corrugatus are more laterally compressed and have more developed
shearing surfaces, than in other species of Arctocyon that have more
bunodont premolars. All this suggests that the smaller species, A.
corrugatus was more adapted for the carnivorous way of life, while the
larger species such as A. ferox, A.acrogenius and A. primaevus were
more omnivorous.
ACKNOWLEDGMENTS
R. Tedford made it possible to study specimens in the AMNH
and R. Emry provided access to USNM specimens. Funding from the
NMMNH Foundation made this study possible. The work of PK was
supported by the Russian Foundation for Basic Research, project 0204-48458, and the Grant Council of the President of Russian Federation, project NSh-1840.2003.4.

REFERENCES
Anton, M. and Galobart, A., 1999, Neck function and predatory behavior in the
scimitar toothed cat Homotherium latidens (Owen): Journal of Vertebrate Paleontology, v. 14, p. 771-784.
Archibald, J.D., 1998, Archaic ungulates Condylarthra; in Scott, K.M., Janis,
C.M. and Jacobs, L. L., eds., Tertiary Mammals of North America: Cambridge,
Cambridge University Press, p. 292-331.
Blainville, H. M., 1841, Osteographie et description icinographique des Mammiferes
recentes et fossiles (Carnivores). Vol. I, II. Paris.
Cope, E.D., 1880, On the genera of the Creodonta: Proceedings of the American
Philosophical Society, v. 19, p. 76-82.
Cope, E.D., 1883, First addition to the fauna of the Puerco Eocene: Proceedings of
the American Philosophical Society, v. 20, p. 545-563.
Cope, E.D., 1884, The Creodonta: American Naturalist, v. 18, p. 344-353.
Gazin, C.L., 1956, Paleocene mammalian faunas of the Bison Basin in southwestern Wyoming: Smithsonian Miscellaneous Coll., v. 131(6), p. 1-57.
Gidley, J.W., 1919, New species of claenodonts from the Fort Union (basal Eocene)
of Montana: Bulletin of the American Museum of Natural History, v. 41, p.
541-555.
Krause, D.W. and Gingerich, P. D., 1983, Mammalian fauna from Douglass Quarry,
earliest Tiffanian (Late Paleocene) of the Crazy Mountain Basin, Montana: Contributions Museum Paleontology, University of Michigan, v. 26(9), p. 157-196.
Lemoine, V., 1896, Etude sur les couches de lEocene inferieur remois qui continental la faune Cernaysienne, et sur deux types nouveaux de cette faune: Bulletin
Societe Geologique de France, v. 3(24), p. 333-344.
Matthew, W.D., 1937, Paleocene faunas of the San Juan Basin, New Mexico: Trans-

actions of the American Philosophical Society, v. 30, p. 1-510.


McKenna, M.C. and Bell, S.K., 1997, Classification of mammals above the species
level. N.Y.: Columbia Univ. Pr., 631 p.
Rigby, J. K., 1980, Swain quarry of the Fort Union Formation, middle Paleocene
(Torrejonian), Carbon County, Wyoming: geological setting and mammalian
fauna: Evolutionary Monographs, v. 3, p. 1-178.
Russell, D.E., 1964, Les Mammiferes paleocenes dEurope: Memoirs Museum
National dHistoire Naturelle, Ser. C., v. 13, p. 1-324.
Scott, W.D., 1892, A revision of North American Creodonta with notes on some
genera which have been referred to that group: Proceedings Academy Natural
Sciences, Philadelphia, v. 44, p. 291-323.
Simpson, G.G., 1937, The Fort Union Group of the Crazy Mountain Field and its
mammalian faunas: U.S. National Museum Bulletin, v. 169, p. 1-287.
Van Valen, L., 1978, The beginning of the age of mammals: Evolutionary Theory, v.
4, p. 45-80.
Williamson, T.E., 1996, The beginning of the age of mammals in the San Juan
Basin, New Mexico: Biostratigraphy and evolution of Paleocene mammals of
the Nacimiento formation: New Mexico Museum Natural History and Science,
Bulletin 8, 141 p.
Williamson, T.E. and Lucas, S.G., 1993, Paleocene vertebrate paleontology of the
San Juan Basin, New Mexico: New Mexico Museum of Natural History and
Science, Bulletin 2, p. 105-135.
Winterfield, G.F., 1982. Mammalian paleontology of the Fort Union Formation (Paleocene), eastern Rock Springs uplift, Sweetwater County, Wyoming: Contributions to Geology, University of Wyoming, v. 21, p. 73-112.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

21

OXYCLAENUS FROM THE EARLY PALEOCENE OF NEW MEXICO AND


THE STATUS OF THE OXYCLAENINAE (MAMMALIA, ARCTOCYONIDAE)
PETER E. KONDRASHOV1 AND SPENCER G. LUCAS2
1

Northwest Missouri State University, 800 University Drive, Maryville, MO 64468 and Paleontological Institute of RAS, Moscow, RUSSIA;
2
New Mexico Museum of Natural History, 1801 Mountain Road N. W., Albuquerque, NM 87104

AbstractNewly collected specimens of one of the most enigmatic Paleocene condylarth genera, Oxyclaenus,
are described. Three species are present in the Puercan of the San Juan Basin: O. cuspidatus, O. antiquus and O.
simplex, and differ from each other mostly in size. O. antiquus (Simpson, 1936) is the largest species of Oxyclaenus.
Oxyclaenus pearcei Gazin, 1941 is a junior subjective synonym of Prothryptacodon furens Simpson, 1937. A
new species of Oxyclaenus is described by Middleton and Dewar (2004) in this volume. It differs from the other
species in its small size and primitive morphology. Oxyclaenus pugnax does not belong to this genus, and it
is a junior subjective synonym of Loxolophus priscus. Carcinodon filholianus from the early Paleocene of New
Mexico, and O. corax from the early Paleocene of Alberta, are subjective junior synonyms of O. simplex. The
subfamily Oxyclaeninae has been a wastebasket taxon for all primitive arctocyonids for more than a century. We
place only two genera, Oxyclaenus and Baioconodon, in the tribe Oxyclaenini, which is included in Oxyclaeninae
together with the tribe Loxolophini, and conclude that the Chriacus group should be referred to Chriacinae,
as originally proposed by Osborn and Earle (1895) and Matthew (1937).
Key words: Oxyclaenus, Arctocyonidae, Oxyclaeninae, Paleocene, San Juan Basin, New Mexico

INTRODUCTION
Cope (1884) originally described Oxyclaenus as a subgenus of
Mioclaenus and differentiated it from the latter by its lack of the internal tubercle (protocone) on the third upper premolar. Besides the type
species, O. cuspidatus, he included O. corrugatus and O. ferox in the
same subgenus. Cope (1888), however, did not refer to Oxyclaenus in
the revision of the Puerco fauna. Scott (1892) resurrected Oxyclaenus
as a separate genus with a revised diagnosis.
Subsequently, several species of Oxyclaenus were described.
Simpson (1935) described Chriacus pugnax from the lower Paleocene
(Torrejonian) of Montana. This species was later assigned to Oxyclaenus
(Van Valen and Sloan, 1965; Robison, 1986). Oxyclaenus antiquus was
also originally referred to Chriacus (Simpson, 1936) and later to
Oxyclaenus (Van Valen and Sloan, 1965). Van Valen (1978) referred
this species to Baioconodon, but Middleton (1983) disagreed and assigned it to Oxyclaenus. Oxyclaenus pearcei was described by Gazin
(1941) from the Dragonian of Utah. Middleton (1983) doubted the
validity of this species and believed it may be conspecific with O. simplex. Another poorly known species of Oxyclaenus, O. corax, was described from the early Paleocene of Alberta (Johnston and Fox, 1984).
It is very close in size to O. simplex, and, considering the high degree of
variation in tooth morphology of early arctocyonids, and we consider it
conspecific with O. simplex, although the scarcity of material makes
detailed comparison difficult.
Middleton (1983), in his doctoral dissertation, mentioned a new
species of Oxyclaenus, which is formally described in this volume by
Middleton and Dewar (2004). Newly collected material of Oxyclaenus
cuspidatus and Oxyclaenus simplex allows us to describe the tooth morphology and lower jaw structure, assess the validity of the previously
described species of Oxyclaenus and to redefine the status and composition of the Oxyclaeninae.
Abbreviations: AMNH = American Museum of Natural History, New York, New York; KUVP = University of Kansas, Museum of
Paleontology, Lawrence, Kansas; NMMNH = New Mexico Museum of
Natural History and Science, Albuquerque, New Mexico; UA = University of Alberta, edmonton, Alberta, Canada; UCM = University of
Colorado, Museum of Paleontology, Boulder, Colorado; UCMP = University of California Museum of Paleontology, Berkeley, California;

USNM = National Museum of Natural History, Smithsonian Institution, Washington, D.C.


P = upper premolars; M = upper molars; p = lower premolars, m
= lower molars; L = maximum tooth crown length; W = maximum
tooth width or trigonid width (for m1-3); TW = talonid width.
SYSTEMATIC PALEONTOLOGY
Order Procreodi Matthew, 1909
Family Arctocyonidae Giebel, 1855
Subfamily Oxyclaeninae Scott, 1892
Tribe Oxyclaenini Scott, 1892
Oxyclaenidae Scott, 1892, p. 294 (in part).
Oxyclaeninae: Matthew, 1937, p. 38 (in part).
Oxyclaenidae: McKenna and Bell, 1997, p. 359 (in part).
Type genusOxyclaenus Cope, 1884 (=Carcinodon Matthew,
1937).
Included generaThe type genus and Baioconodon Gazin, 1941
(=Ragnarok Van Valen, 1978).
Revised diagnosisArchaic arctocyonids with primitive upper
molars, hypocone vestigial or absent, lower molar paraconid slightly
reduced, cusps bulbous; constriction between the trigonid and the talonid distinct. Differ from Loxolophini in the weak development of crests
on lower and upper molars and absence or weak development of the
upper molar hypocones and in having a constriction between the trigonid and the talonid on the lower molars.
DistributionLower Paleocene (Puercan) of North America.
Genus Oxyclaenus Cope, 1884
Oxyclaenus Cope, 1884, p. 312.
Mioclaenus: Cope, 1884, p. 312 (in part).
Mioclaenus: Cope, 1888, p. 303, 329 (in part).
Protochriacus: Osborn and Earle, 1895, p. 22 (in part).
Carcinodon: Scott, 1892, p. 323.
Oxyclaenus: Matthew, 1937, p. 38.
Carcinodon: Matthew, 1937, p. 57
Oxyclaenus: Van Valen, 1978, p. 54.
Oxyclaenus: Johnston and Fox, 1984, p. 192.

22
Oxyclaenus: Robison, 1986, p. 124.
Oxyclaenus: McKenna and Bell, 1997, p. 359.
Oxyclaenus: Archibald, 1998, p. 316.
Oxyclaenus: Eberle and Lillegraven, 1998, p. 59.
Type speciesOxyclaenus cuspidatus (Cope, 1884).
Included speciesThe type species, Oxyclaenus simplex (Cope,
1884) (=Mioclaenus filholianus Cope, 1888; =Oxyclaenus corax
Johnston and Fox, 1984), and Oxyclaenus antiquus (Simpson, 1935).
Revised diagnosisAn oxyclaenine that differs from
Baioconodon in the slight development of the constriction between the
trigonid and the talonid on the lower molars, in having less obtuse
cuspids on the lower molars and in the weak development of the hypocone.
DistributionLower Paleocene (Puercan) of North America.
Oxyclaenus cuspidatus (Cope, 1884)
Figs. 1B-C, 2B, G-H, 3A, C-D, G-H, Tables 1-2
Mioclaenus
Oxyclaenus
Oxyclaenus
Oxyclaenus
Oxyclaenus

(Oxyclaenus) cuspidatus Cope, 1884, p. 312.


cuspidatus: Matthew, 1937, p. 39, fig. 4.
cuspidatus: Van Valen, 1978, p. 54.
cuspidatus: Williamson, 1996, p. 42.
cuspidatus: Eberle and Lillegraven, 1998, p. 59, fig. 5C.

LectotypeAMNH-3252, left maxillary fragment with P4-M3


(Figs. 1B, 2G; Matthew, 1937, fig. 4).
Referred specimensFrom the Nacimiento Formation in the
San Juan Basin, New Mexico: From 2 miles above Ojo Alamo Arroyo,
San Juan Basin, New Mexico: AMNH 16344, left dentary with m1-3,
right dentary with p4-m3, left P4, right P4; 16345, left dentary with
m2-3, right dentary with p4-m3; 16346, left dentary with m2-3 (Fig.
3G), right dentary with p4-m3 (Fig. 3H); 16348, left dentary with p3m2; 16349, left m1, m2, right p4; 16350, left maxillary fragment with
M1-3; 16352, left dentary with m2-3; 16362, left dentary with m2-3,
right dentary with m1-3; 16365, left m1, right m2, m3; 23139, right
m1-2; 23141, left m1; 81887, right dentary with m2-3.
From Barrel Springs Arroyo, San Juan Basin, New Mexico:
AMNH 21138, left dentary with p2-4; 21145, right dentary with m1-2;
23142, right dentary with m2-3; 23144, left dentary with p4; 23146,
left m3; 23147, left m1, m2; 23163, right p4; 27680, right p2, p3, m23; 27687, left dentary with p4, m2; 27731, left M2-3; USNM 405675,
left dentary with p3-4; 405678, left dentary with m2-3; 405679, left
maxillary fragment with M1-2, left p4; 405680, right dentary with p34; 405681, fragments of left and right dentaries with p4-m1; 405682,
right dentary with m1; 405682, right dentary with m2; 405685, left m1
(isolated).
From Coal Creek Canyon, Polymastodon beds, San Juan Basin, New Mexico: AMNH 16352, left M2; 793f, right m2-3; 793g, left
p4; 793h, right m3; 793m, left M3 (Matthew, 1937, p. 40, fig. 4); 794,
left dentary with m1-2 (Figs. 1C, 3A), right dentary with m2-3 (Fig.
3D), right dentary with m1-3 (Fig. 3C); 799, right dentary with m2-3,
left maxillary fragment with M1-2.
From Coal Creek Canyon, San Juan Basin, New Mexico: AMNH
793, left dentary with m1-2; 797, left dentary with m1-3, right dentary
with m2-3; 798, left dentary with m1-3; 798a, left m3; 798b, right dentary
with m1-3; 801, right maxillary fragment with P3-M1 (Fig. 2B); 804,
right dentary with m2-3.
From divide between Ojo Alamo and Barrel Springs Arroyos,
San Juan Basin, New Mexico: USNM 405677, right dentary with m12.
From Kimbetoh Arroyo, San Juan Basin, New Mexico: AMNH
23134, right m1; 23136, left M1; 23137, left M2; 23148, left m1, m2;
23149, right m2-3; 2315, left p4; 23150, left m2; 23152, right m2;
23175, right M2; 23182, right M2-3; USNM 405672, left maxillary
fragment with M1.

FIGURE 1. Occlusal views of selected cheek teeth of Oxyclaenus simplex, O.


cuspidatus and O. antiquus. A, O. simplex, AMNH 3107, left M1-3 (holotype). B,
O. cuspidatus, AMNH 3252, left M1-3. C, O. cuspidatus, AMNH 794, left m1-3.
D, O. simplex, AMNH 3205, left p3-m3. E, O. antiquus, AMNH 27714, right P3
and M1-3 (image reversed) (holotype). A-D after Matthew (1937), E after Simpson
(1936).

From Ojo Alamo Arroyo, San Juan Basin, New Mexico: AMNH
23141, right dentary with m1-2.
From San Juan Basin, New Mexico, collected by Baldwin, 1885:
AMNH 3313, left dentary with p4, m2, right dentary with m1-3.
From San Juan Basin, New Mexico, collected by Baldwin:
AMNH 3056a, right m2; 3252, skull fragment with left P4-M3; 3607a,
left dentary with m2, right dentary with m3.
From San Juan Basin, New Mexico: AMNH 792, right dentary
with m1-3.
From Tsosie Arroyo, San Juan Basin, New Mexico: AMNH
27607, left M1-2; 58076, right dentary with m2-3; 58125, right M1;
58193, left m2; 58215, left m3; 58222, right M2, M3 (isolated); 58223,
left M2; 58240, right M1; 58570, left m3; 59835, right M2; 59997,
right m2.
From the University of Kansas locality 3, Ojo Alamo area, chiefly
Barrel Springs drainage (sec. 10 and W1/2 sect. 11, T24N, R11W),
Taeniolabis zone, San Juan Basin, New Mexico: KUVP 9435, right
maxillary fragment with M1-3 (Fig. 2H); 9425, left dentary with p4m2; 13366, left dentary with m2-3, right dentary with m2-3; 9437, right
dentary with p4-m1; 9439, left M3; 9440, left m1, m2; 9441, right
dentary with m2-3; 9462, left m1; 9470, right dentary with m1; 9472,
left p4, right M2; 9423, right dentary with m1-3; 12703, left dentary
with m2-3; 9424, right dentary with m2, left dentary with m2-3; 13367,
right dentary with m2; 12097, right M2; 8090, right dentary with m3;

23

FIGURE 2. Occlusal views of selected upper cheek teeth of Oxyclaenus. A, O. antiquus, B, G-H, O. cuspidatus. C-F, O. simplex. A, AMNH 27714, right P3, M1-3
(holotype of Chriacus antiquus). B, AMNH 801, right P3-M1. C, AMNH 3107, right M1-3 (lectotype of O. simplex). D, AMNH 16415, right M1-3. E, AMNH 3107, left
M1-3. F, NMMNH 8661, right M1-3 of O. simplex. G, AMNH 3252, left P4-M3 (lectotype of O. cuspidatus). H, KUVP 9435, right M1-3 of O. cuspidatus.

12117, right dentary with m2-3; 12121, right dentary with m2-3; 12913,
left M2; 12694, right M2; 12696, left m2; 12704, right dentary with
m3; 12795, left dentary with p4-m1; 12718, left dentary with m3.
From the Ferris Formation in the Hanna basin, Wyoming: UW
26231, left m3 (Eberle and Lillegraven, 1998, fig. 5C); UW 26142,
right m3.
Revised diagnosisA species of Oxyclaenus that differs from
O. simplex in the weak development of the hypocone on M1-2, lingually projecting parastyle on the M2 and larger size (15-20 %). Differs from O. antiquus in its much smaller (30 %) size.
DistributionLower Paleocene (Puercan) of New Mexico and
Wyoming.
CommentVan Valen (1978) mentioned that the type specimen is atypical, but it is within the range of intraspecific variability of
this species. We restrict the holotype of O. cuspidatus to a maxillary
fragment with P4-M3.
Description The p4 is very much compressed laterally. It has
a tall protoconid, somewhat inclined posteriorly. The talonid is also
compressed laterally and bears a single cuspid. It is connected to the
base of the main cuspid by a well-developed cristid.

The m1 is rather slender, smaller than the m2, but larger than the
m3. The trigonid is higher than the talonid, but narrower. The protoconid
is the largest trigonid cuspid, and the metaconid is only slightly smaller.
The paraconid is somewhat reduced and connected to the metaconid by
a well-pronounced, arched paracristid. The protoconid and the metaconid are connected by a metacristid. There is a weak cristid that runs
from the paraconid to the metaconid, closing the trigonid basin. There is
a slight constriction between the trigonid and the talonid. The cristid
obliqua begins at the base of the posterior wall of the trigonid between
the protoconid and the metaconid. The talonid basin is wide and deep.
The hypoconid is the largest talonid cuspid, and the entoconid is the
smallest. The hypoconulid is situated closer to the entoconid, but not
connected either to the hypoconid or to the entoconid. There is a weak
entocristid, but it does not reach the posterior wall of the metaconid,
leaving the talonid basin slightly open lingually. The cingulid is well
developed on the anterior, labial and posterolabial walls of the crown.
The postcingulid rises from the base of the crown up to the hypoconulid.
The m2 is larger than the m1, and the trigonid is equal in width to
the talonid. The cuspid structure is almost identical to that of the m1.
The hypoconid is relatively larger and twice the size of the hypoconulid.

24
TABLE 1. Lower tooth measurements (in mm) of O. cuspidatus.
Coll. number
AMNH 792
AMNH 793
AMNH 793
AMNH 793f
AMNH 793g
AMNH 793h
AMNH 794
AMNH 794
AMNH 794
AMNH 794
AMNH 799
AMNH 797
AMNH 797
AMNH 798
AMNH 798a
AMNH 798b
AMNH 799
AMNH 804
AMNH 2315
AMNH 3056a
AMNH 3218
AMNH 3313
AMNH 3313
AMNH 3607a
AMNH 16302
AMNH 16344
AMNH 16344
AMNH 16345
AMNH 16345
AMNH 16346
AMNH 16346
AMNH 16346
AMNH 16348
AMNH 16349
AMNH 16352
AMNH 16365
AMNH 21138
AMNH 21145
AMNH 23134
AMNH 23139
AMNH 23141
AMNH 23141
AMNH 23142
AMNH 23144
AMNH 23146
AMNH 23147

p2 L

p2 W

p3 L

p3 W

p4 L

6.8

p4 W

m1 L

m1 W

m1 TW

5.7

4.1

4.3

m2 L
6.1
6.9
6.7
7.4

m2 W
5
5.3
5.2
4.9

m2 TW
4.9
5.1
5.1
5

6.2
6.1
5.9
5.81
6.74
6.2
6.7

4.8
4.9
4.8
4.88

4.75
4.5
4.9
5.05
4.74
4.6
4.9
4.9

7
7.1
6.9

5.3
5.3
5.2

5.98
6.6
6.3

5.2
4.42
4.6
4.8

m3 L
6.4

m3 W
4.2

m3 TW
3.4

6.1

4.7
4.3

3.7

6.7

4.3

3.1
5.2

3.8

5.8
4.68

3.8
3.36

4
3.92

6.1

3.9

4.2

3.7

4.1

3.8

4.7

6.1

4.9
5.1

4.8
4.9
4.8

5.4
7

4.3
5.4
4.12
4.06
4.3
4.5
4.1
3.1
4.2
4.1
4.9

6.05

3.08

7
6.6
6.9
6.7
7.1

3.8
3.2
3.4
3.8
4
3.9
4.8

3.5
6

4.2
4.1
4.1
4.8
4.4
4.6
5.5
3.29
2.8
3.5

6.2
6

4.5
4

3.7
3.25

5.9

4.4

7.3

4.2

5
5.94
5.52
6.7
6.2
5.2

2.9
3.53
3.48
3.6
3.9
2.9
3.1
3.8

2.8
5.94

3.15

3.84

3.6

3.9

6.3

3.4
3.6
3.3

6.3
6.3
6.4

4
3.9
3.9

3.9
4.4
4.3
4.3

5.83
6.1

2.18
3.4

5.42
5.7

3.12
3.4
3.6

3.5

6.1
4.8
5.3

3.9

4.1

6.6
6.1
6.3
6.2
6

3.1
3.9
4
3.8
4.3

3.9
4.1
4.9
3.9
4

7.4

7.1

5.6

4.1

5.1

6.9
6.8

3.49
3.9
3.9
4.3

6.8
7.1
7
7.2
6.9
6.15
5.95
6.1
7.1
6
6.3

4.9
5.2
5.3
5.2
5.2
4.33
4.6
4.5
4.8
5.3
5.1
4.7

4.05
4.8
4.9
4.9
4.9
5.1
5.2
5
5.1
3.98
4.2
4.5
5.6
4.9
4.5

6.3

2.9

6.6
6.3
7.1
7

5
4.5

5
4.7
4.1

6.2

3
6.7

The hypoconulid is central in position, and its apex is often split into
two, which cannot be seen on the much worn specimens. The hypoconulid
and the entoconid are connected by a weak cristid.
The m3, although much variable in size and shape, is usually much
smaller than the other two molars. There is no cristid connecting the
paraconid and the metaconid, so the trigonid basin is open lingually. The
hypoconulid is the same size as the hypoconid, and the entoconid is
much smaller. The entocristid is better developed than on the other two
molars and almost closes the talonid basin.
The P3 is triangular in cross section with no protocone and a small
parastyle. Only weak anterior and posterior cingula are developed. The
P4 has a strong parastyle and a well-developed protocone. The labial
cingulum is rather strong.

4.6

3.8

4.9

The M1 is smaller than the M2 and more triangular in shape. The


protocone is almost the same size as the paracone and the metacone.
Both paraconule and metaconule are developed, and the metaconule is
slightly larger. There are strong anterior and posterior cingula; the posterior cingulum is wider, but does not form a distinct hypocone. The
parastyle is better developed than the metastyle. There is a weak
centrocrista running from the parastyle through the paracone and metacone to the metastyle. The labial cingulum is well developed.
The M2 is larger than the M1, and has a larger protocone and a
wider posterior cingulum, so it is more rectangular in shape than the M1.
The metaconule is slightly larger than the paraconule. Both conules are
situated close to the protocone. Preparaconule and postmetaconule wings
are rather weak. The latter is usually better developed and reaches the

25
Table 1 continued.
Coll. number
AMNH 23148
AMNH 23149
AMNH 23150
AMNH 23152
AMNH 23163
AMNH 27680
AMNH 27687
AMNH 58076
AMNH 58193
AMNH 58215
AMNH 59997
AMNH 81887
KUVP 8090
KUVP 9423
KUVP 9424
KUVP 9424
KUVP 9425
KUVP 9437
KUVP 9440
KUVP 9462
KUVP 9470
KUVP 9472
KUVP 12117
KUVP 12121
KUVP 12696
KUVP 12703
KUVP 12704
KUVP 12705
KUVP 12718
KUVP 13366
KUVP 13366
KUVP 13367
NMMNH-30502
UCM 39114
UCM 39114
UCM 39114
UCM 39114
USNM 15525
USNM 405675
USNM 405677
USNM 405678
USNM 405679
USNM 405680
USNM 405682
Mean
Number of spec.
Stand. Deviat.

p2 L

4.1

p2 W

2.3

p3 L

5.9

p3 W

p4 L

p4 W

7.2

6.3

7.2

2.9

m1 L

m1 W
4

m1 TW
4.4

m2 L
6.5
7
6.9
7.1

m2 W
5.5
4.9
4.9
4.3

m2 TW
4.9
4.8
4.8
5

m3 L

m3 W

m3 TW

3.2

6.8
6.8
5.9
5.9

5.2
4.9
4.4
4.6

4.8
4.8
4.5
4.4

6
6.6

4.5

4.5

7.16
7.05
6.89
6.61

4.89
5.18
4.9
4.88

5.2

3.2
3.8

3.2

6.75
6.03

2.42
3.58

6.5

3.56

3.28

5.84

3.66
4.08
3.43
3.44
3.71

3.91

6.74
6.38
6.26
6.68

3.62
4.04
4.12

2.7

6.8
7.1

6
4.1
2
0

2.3
1
0

5.7
5
0.4

5.9

2.85
2
0.21

3.18
3.28

6.31

4.18

3.68

6.62
6.01

4.41
4.28

3.85
3.45

6.39

4.12

3.27

6.24
6.56

4.24
6.64

3.41
3.42
4.96

4.18
3.24

3.55
2.75

4.41
4.2

3.4

2.62

3.12

5.96

3.85

4.92
4.96
4.96

4.61
4.63
4.37
4.54

3.95
4.57

5.1

3.96
4.26

4.58

6.73
6.31
7.06
7.57
6.49

4.77
5.34
5.03
4.71

6.2
6.75
6.5

5.13
5.44

3.47
3.55

4.02
3.91

4.9

3.5

3.9

6.1

3.9

4.4

7.06
5.98
5.8
6.03
6.21
6

4.82
4.64
4.62
4.55
5.1
4.9

4.76
4.4
4.76
4.56
4.65
5

6.9
7.3

5.2
5.1

5.1
5.2

4.3

3.5

7
6.59
58
0.48

5
4.868
57
0.281

5
4.7325
62
0.302

6.15
41
0.75

4.281
46
0.607

3.4733
39
0.4665

6.78
4.78

5.9
7.22

3.6

6.6188 3.03
19
18
0.4706 0.86

5.93
32
0.55

metastyle.
The M3 is much reduced. The metaconule is much smaller than
the paraconule. The conules are small, but well pronounced. The cingulum is very strong and almost circular.
Oxyclaenus simplex (Cope, 1884)
Figs. 1A, D, 2C-F, 3B, E-F, I, Tables 3-4
Chriacus simplex Cope, 1884, p. 314.
Mioclaenus filholians Cope, 1888, p. 303, 329.
Protochriacus simplex: Osborn and Earle, 1895, p. 23.
Oxyclaenus simplex: Simpson, 1936, p. 3, fig. 2.
Carcinodon filholianus: Matthew, 1937, p. 57, fig. 10.

3.65
34
0.326

3.915
35
0.3081

Oxyclaenus simplex: Matthew, 1937, p. 40, fig. 5.


Oxyclaenus simplex: Van Valen, 1978, p. 54.
Oxyclaenus corax Johnston and Fox, 1984, p. 192, pl. 7, figs. 1-7, table
11.
Oxyclaenus simplex: Williamson, 1996, p. 42.
LectotypeAMNH 3107, right maxillary fragment with M1-3
(Fig. 2C; Matthew, 1937, fig. 5) from Polymastodon beds, San Juan
Basin, New Mexico.
Referred specimensFrom the Nacimiento Formation in the
San Juan Basin, New Mexico: NMMNH locality 317: NMMNH P8661, right maxillary fragment with M1-3 (Fig. 2F); 8764, right p4 and
right dentary m2-3, bone fragments.

26
TABLE 2. Upper tooth measurements (in mm) of O. cuspidatus.
Coll. Number
AMNH 793m
AMNH 799
AMNH 801
AMNH 3252
AMNH 16344
AMNH 16350
AMNH 16352
AMNH 23136
AMNH 23137
AMNH 23175
AMNH 23182
AMNH 27607
AMNH 27731
AMNH 58125
AMNH 58222
AMNH 58223
AMNH 59835
KUVP 9435
KUVP 9439
KUVP 9472
KUVP 12097
KUVP 12694
KUVP 12913
USNM 405672
USNM 405679
Mean
Number of spec.
Stand. Deviat.

P2 L

4.2

P2 W

P3 L

P3 W

P4 L

P4 W

5.8

3.2

5.9
5.8
6.9

5.8
5.7
5.6

M1 L

M1 W

M2 L

M2 W

5.9
6.1
6.3

6.9
7.3
6.5

6.5

8.3

6.93

6.2

6.4

5.6

6.2

5.9

7.3

5.9

6.5

M3 W
7.5

8.4

4.6

6.2

6.6
6.6

8.6
8.6

4.8

6.9

6.5
6.8
6.8
6.2
6.1

7.9
7.8
8.6
8.3
8.3

3.9

7.2

4.3

4.5

7.9

5
4.56

7.24
6.49

4.5
8
0.34

7.054
8
0.54

2
5.99

4.2
1
0

M3 L
4.3

2
1
0

5.8
1
0

3.2
1
0

6.2
3
0.61

5.7
3
0.1

NMMNH locality 646: NMMNH P-15175, left dentary with


m2-3 and alveoli for p1-m1.
From 2 miles above Ojo Alamo Arroyo, San Juan Basin, New
Mexico: AMNH 16347, left dentary with m2-3 (Fig. 3B); 16375, right
dentary with m2-3.
From 3 miles east of Kimbetoh, San Juan Basin, New Mexico:
AMNH 15415, right maxillary fragment with M1-3; 16355, right dentary
with m1; 16367, left dentary with p3, m1-3 (Fig. 3I).
From Barrel Springs Arroyo, San Juan Basin, New Mexico:
AMNH 27703, left m3; USNM 15526, right dentary with m2-3; UCMP,
left maxillary fragment with P4-M3 (Simpson, 1936, fig. 2).
From Coal Creek Canyon, San Juan Basin, New Mexico: AMNH
808, right dentary with m1-2.
From Kimbetoh Arroyo, San Juan Basin, New Mexico: AMNH
23124, left dentary with p3-4; 23132, left m2; 23133, left m2; 23135,
left dentary with m2-3; 23154, left m1; 23162, right M2; USNM 405683,
left m2.
From Ojo Alamo Arroyo, San Juan Basin, New Mexico: AMNH
23129, right dentary with p3, m2; 23156, right dentary with dp3-m1;
23158, left dentary with p4-m2; 23159, right dentary with p3-4; 23160,
left dentary with p4-m3; USNM 15528, left dentary with m2-3.
From Polymastodon beds, San Juan Basin, New Mexico: AMNH
3107, left maxillary fragment with M1-3 (Figs. 1A, 2E), right maxillary fragment with M1-3 and left m2 (Fig. 3F).
From San Juan Basin, New Mexico, collected by Baldwin, 1885:
AMNH 3205, left dentary with p3-m3 (Fig. 1D; Matthew, 1937, fig.
10), right dentary with m1-3 (Fig. 3E, here designated the lectotype of
Carcinodon filholianus); 3206, left m2, m3.
From San Juan Basin, New Mexico, collected by Baldwin:
AMNH 3206B, right dentary with m2-3; 3887, left dentary with p4m3;16415, right M1-3 (Fig. 2D).

5.9
6.8
6.06
10
0.32

7.01

7.1
6.7
6.791
10
0.388

6.53

8.45

5.84
6.57
6.31
6.94

7.12
9.05
7.95
8.78

6.4
6.51
15
0.31

9.1
8.35
15
0.509

From Tsosie Arroyo, San Juan Basin, New Mexico: USNM


405686, left M1; AMNH 23153, right M2; 58051, right M2; 58056,
right M1; 58100, right m1; 58122, right m2; 58124, right m2; 58129,
right m2 (talonid), left m1; 58132, right m3; 58188, left m2-3; 58229,
left M1; 58237, right M1; 58277, right M1; 58278, left M1; 58357, left
m1; 58365, left M3; 58410, left m2; 58411, right M1; 58462, left M1;
58501, left M1; 58534, left m2; 58534, right M2; 58535, right m2;
58568, right M1; 58597, right M3; 59776, right m2; 59777, right M2;
59829, right m1; 59830, right m2; 59909, left m1; 59964, left m2; 59993,
left m2-3.
From University of Kansas locality 3, Ojo Alamo area, chiefly
Barrel Springs drainage (sec. 10 and W1/2 sect. 11, T24N, R11W),
Taeniolabis zone, San Juan Basin, New Mexico: KUVP 9469, left m23; 8092, right dentary with m1-2; 8089, left m3.
From the Ravenscrag Formation, Saskatchewan, Canada: UA
15217, right M2 (holotype of O. corax; Johnston and Fox, 1984, pl. 7,
fig. 2); 15218, M1 (Johnston and Fox, 1984, pl. 7, fig. 1); 15155, M1;
15190, M2; 15216, M3 (Johnston and Fox, 1984, pl. 7, fig. 3); 15156,
15208, 15209, M3s; 15199, m1s (Johnston and Fox, 1984, pl. 7, figs.
4-5); 15214, 15194, 15195, 15196, 15197, m2s; 15215, m2 (Johnston
and Fox, 1984, pl. 7, figs 6-7); 15106, 15191, m3s; 151298, 15200,
15201, 15202, 15203, molar fragments.
Revised diagnosisA small species of Oxyclaenus that differs
from O. cuspidatus in having a more developed cingulum on M1-2 and
smaller size (15-20 %). Differs from O. antiquus in being much smaller
(40-45 %).
DistributionLower Paleocene (Puercan) of North America.
DescriptionNMMNH P-15175 is an unusually complete
dentary of O. simplex, so we describe it here. The lower jaw is relatively
short and stout. The maximum height of the horizontal ramus is beneath
the m2. Anterior to the m1, the height of the horizontal ramus gradually

27

FIGURE 4. Bivariate plot of m2 size in Oxyclaenus species.

FIGURE 3. Occlusal views of selected lower cheek teeth of Oxyclaenus. A, C-D,


G-H, O. cuspidatus. B, E-F, I, O. simplex. J, O. antiquus. A, AMNH 794, left m12. B, AMNH 16347, left m2-3. C, AMNH 794, right m1-3. D, AMNH 794, left
m2-3. E, AMNH 3205, right m1-3 (holotype of Carcinodon filholianus). F, AMNH
3107, left m2. G, AMNH 16346, left m1-3. H, AMNH 16346, right p4-m3. I,
AMNH 16367, left p4-m3. J, KUVP 13364, left p4-m3.

diminishes. The vertical ramus is rather high; its height is about 40% of
the length of the horizontal ramus. The angular process is well developed
and projects posteroventrally and slightly laterally. There are two welldeveloped anteroposterior crests on its medial surface. The articular
process projects posteriorly, and the lower jaw condyle is rather dorsal
in position; it is at approximately 65 % of the height of the vertical ramus
from the base of it. The condyle is spindle-shaped, with a broader medial
portion, and the condyle neck is well developed. The incisura mandibulae
is rather deep, and its axis is oriented at approximately 70o to the long
axis of the lower jaw. The coronoid process is relatively short, rather
narrow and slightly inclined posteriorly. The masseteric fossa is deep
and well defined anteriorly, with a strong crest. The mandibular foramen
is situated on the medial side of the dentary, on the occlusal level halfway
between the anterior base of the coronoid process and the articular process. The anterior mental foramen is situated under the p1; it is large, oval
in shape and faces anterodorsally. The posterior mental foramen is situated under the anterior root of the p3; it is round in shape and faces
anterolaterally.
The p1 is single-rooted, and the p2-4 are double-rooted. Two
teeth are preserved in the dentary (NMMNH P-15175), the m2 and
m3. There are nine alveoli anterior to the m2, eight of them in pairs,
whereas the anteriormost is single. We identify the anterior eight as the
double-rooted alveoli of the m1, p4, p3 and p2 and the single alveolus
as that of the single-rooted p1. Anterior to the p1 alveolus, there is a
partially-preserved large alveolus for the canine. The p4-m1 alveoli are
situated close to each other, suggesting that the p2-m1 were closely
spaced. The p1 alveolus is slightly anterior to the p2 alveolus, suggesting there could be a very small diastema between these two teeth.
The lower molar cuspids are rounded. The m2 trigonid is slightly
wider than the talonid. The paraconid is much smaller than the protoconid and the metaconid, although well developed. It is slightly shifted
labially and lies anterolabial to the metaconid. The paracristid is well
developed. The precingulid is weak, and the labial cingulid is well
expressed between the protoconid and the hypoconid. The metaconid
is slightly larger than the protoconid. The postcingulid begins at the base
of the crown posterior to the hypoconid; as it extends lingually, it rises
up to the hypoconulid. The talonid basin is almost square. The cristid
obliqua begins at the posterior wall of the protoconid and extends posteriorly to the hypoconid, so it is parallel to the long axis of the tooth. The

hypoconid, hypoconulid and the entoconid are situated almost on the


same line, and the hypoconulid projects slightly posteriorly. All three
cuspids are connected by a weak cristid. The entocristid is well developed and closes the talonid basin, merging with the weak lingual cingulid.
The hypoconid is the largest cuspid, and the hypoconulid is the smallest.
The m3 is similar in structure to the m2, but the trigonid is significantly wider than the talonid. The hypoconulid is large and strongly
projects posteriorly. It is equal in size to the hypoconid. The cristid
connecting the hypoconid to the hypoconulid is rather weak. The
hypoconulid and the entoconid are united by a strong entocristid that
runs anteriorly down to the base of the metaconid. The entoconid is
weakly expressed on this cristid.
The upper molars are more rectangular than in O. cuspidatus. All
the molars have a well-developed labial cingulum. The trigon basin is
rather deep. The M1 is smaller than the M2, but larger than the M3. The
parastyle is rather small and lacks a well-pronounced cuspule. The anterior cingulum is rather weak. The posterior cingulum is much wider, with
a well-developed hypocone. The paraconule and metaconule are situated
closer to the paracone and metacone, respectively, than to the protocone.
The M2 is similar in structure to the M1. The parastyle is much
larger, and the paraconule and metaconule are equidistant form the protocone and paracone and metacone, respectively. The preparaconule and
postmetaconule crests do not reach the parastyle and the metastyle,
respectively. There is a weak centrocrista that begins from the parastyle
and runs through the paracone and metacone down to the metastyle.
The M3 is considerably reduced and triangular in shape. The
metacone is much smaller than the paracone and the protocone. The
parastyle is rather large. The postcingulum is narrow and does not bear
any cusps. The paraconule is much larger than the metaconule. All the
molars have deep trigon basins.
CommentVan Valen (1978) synonymized Carcinodon
filholianus Cope, 1884 with O. simplex; this synonymy was later questioned by Johnston and Fox (1984), but Williamson (1996) concluded
that it was justified, and we concur.
Oxyclaenus antiquus (Simpson, 1936)
Figs. 1E, 2A, 3J, Tables 5-6
Chriacus antiquus Simpson, 1936, p. 4, fig. 3.
Oxyclaenus antiquus: Van Valen and Sloan, 1965, p. 745.
Baioconodon antiquus: Van Valen, 1978, p. 57.
Oxyclaenus antiquus: Williamson, 1996, p. 42.
HolotypeAMNH-27714, maxillary fragment with P3-M3 (Figs.
1E, 2A; Simpson, 1936, fig. 3).
Referred specimensFrom the Nacimiento Formation in the
San Juan Basin, New Mexico: From Barrel Springs Arroyo, San Juan
Basin, New Mexico: AMNH 23181, right M1; 23183, left m1; 23184,

28
TABLE 3. Lower tooth measurements (in mm) of O. simplex.
Coll. number
AMNH 808
AMNH 3107
AMNH 3887
AMNH 16347
AMNH 16355
AMNH 16375
AMNH 23124
AMNH 23129
AMNH 23132
AMNH 23133
AMNH 23135
AMNH 23154
AMNH 23156
AMNH 23158
AMNH 23159
AMNH 23160
AMNH 27703
AMNH 58100
AMNH 58124
AMNH 58129
AMNH 58132
AMNH 58188
AMNH 58357
AMNH 58410
AMNH 58534
AMNH 58535
AMNH 59829
AMNH 59830
AMNH 59909
AMNH 59964
AMNH 59993
KUVP 8089
KUVP 8092
KUVP 9441
KUVP 9469
NMMNH-1817
NMMNH-2724
NMMNH-8764
NMMNH-15175
NMMNH-21866
NMMNH-21867
USNM 15526
USNM 15528
USNM 405683
USNM 405684
Mean
Number of spec.
Stand. Deviat.

p3 L

p3 W

p4 L

p4 W

m1 L
4.4

m1 W
2.8

m1 TW
3.1

2.8

4.8

3.1

5.1

3.2

m2 L
5.9
5.8

m2 W
4
4.1

5.2

4.1

m2 TW
4
3.9
4
4.2

2.1
2.3

2.2
4.9

m3 TW

5.6
4.7

3.7
3.3

3.1
3.1

4.6

5.6

3.5

2.9

4.8

2.9

2.5

5.7

3.2
3.5
4.2

2.8

3
5.2
5.8
5.7

3.6
4.2

m3 W

3.6
3.3

3.9
4.6

m3 L

2.1
2.8
3.6

5.4
5.5
4

3.6
3.9
2.8

3.8
4.1
2.7

3.6

4.8

3.1

2.8

4.6

3.2

3.4

4.2
4.2
4.1
4.1

4.5
4
3.9

3.1

3.3

3.4
3.8

3.4

5.6

4.1

5.9

4.1
4.2
4

4
3.9
4

3.8

4.2
3.4

3.9
3.2

2.9

5.3
4.8

3.2
3.8
5.7
4.5

4.97

3.7

4.23333 2.2
3
3
0.35119 0.1

4.2
5
0.7

2.41

2.93

2.01

2.72
6
0.59

4.72
10
0.59

3.246
11
0.653

right M2, left M2; 27689, left dentary with m2, right dentary with m23; 96406, left dentary with p4-m3, right dentary with p4-m2.
From Ojo Alamo Arroyo, San Juan Basin, New Mexico: AMNH
21714, right maxillary fragment with P3, M1-3.
From San Juan Basin, New Mexico, collected by Baldwin:
AMNH 3206, right maxillary fragment with M2-3.
From the University of Kansas locality 3, Ojo Alamo area, chiefly
Barrel Springs drainage (sec. 10 and W1/2 sect. 11, T24N, R11W),
Taeniolabis zone, San Juan Basin, New Mexico: KUVP 13364, left dentary
with p4-m3 (Fig. 3J), right maxillary fragment with M1-3.
Revised diagnosisThe largest species of Oxyclaenus; larger

3.3391
11
0.4359

4.83
5.65

3.42
3.82

5.73

3.94

3.48
3.81
3.77
3.9

4.7
5.04

3.61
3.79

3.65
3.59

4.84
5.2
5.1
5.3

3.32
4.2
3.9
4.1

3.26
4.1
3.6
4.5

5.22
22
0.68

3.87
27
0.335

3.8331
26
0.3325

4.47

3
3

3.29

5.52
4.77
4.72
4.71
4.88
4.96
5.64

3.44
3.34
3.04
3.21
3.24
3.57
2.94

3.1
2.94
2.45
2.94
3.02
3.13
2.83

4.9
4.9

3.2

2.1
2.1

5.6
5.06
17
0.43

3.4
3.299
19
0.317

2.6
2.7588
17
0.4001

than O. cuspidatus (30 %) and O. simplex (45 %).


DistributionLower Paleocene (Puercan) of the San Juan Basin, New Mexico.
DescriptionThe p4 is tall, trenchant; the cingulid is weak.
There is a small anterior cuspid, and the protoconid is tall and laterally
compressed. A sharp elevated cristid runs through the protoconid
anteroposteriorly and reaches the small talonid.
The trigonid cuspids of the m1 are close to each other, the paraconid
is reduced, but not shifted labially, and the paracristid is weak. The
trigonid basin is open lingually. The cingulid is weak and is formed on the
anterior, labial and posterior sides of the tooth. The postcingulid is

29
TABLE 4. Upper tooth measurements (in mm) of O. simplex.

TABLE 5. Lower tooth measurements (in mm) of O. antiquus.

Coll. Number
AMNH 3107
AMNH 3107
AMNH 23153
AMNH 23162
AMNH 58051
AMNH 58056
AMNH 58229
AMNH 58237
AMNH 58365
AMNH 58411
AMNH 58501
AMNH 58568
AMNH 58597
AMNH 59777
NMMNH 8661
USNM 405686
Mean
Number of spec.
Stand. Deviat.

Coll. Number

Tooth

Length

AMNH 27689

m2
m2
m3
m1
p4
m1
m2
m3

8.00
7.90
7.40
8.20
6.83
6.75
7.52
6.66

M1 L
4.9
4.8

M1 W M2 L
5.9
5.8
5.9
6.3
5.9
5.9
5.9
6.5

M2 W
7
7.1
7.1
7.2
6.8

M3 L
4

M3 W
6

AMNH 23183
KUVP 13364

4
4.4
3.2
5.2
4.8
4.3

5.8

Talonid
width
5.90
5.90
4.10
4.80
4.74
5.38
3.96

5
5.7

5
4.68
9
0.4

Trigonid
width
5.90
6.00
5.20
4.90
3.31
4.3
5.86
4.7

6.8
5.9
6
7
0.62

5.97

6.1
7.77

4.2

5.75

5.96
7
0.17

7.01
8
0.5

3.8
4
0.53

5.813
5
0.131

connected to the hypoconulid. The hypoconid is the largest talonid


cuspid, and the hypoconulid is the smallest; it is strictly median in
position and connected to the hypoconid and entoconid by low crests.
The entocristid is relatively well developed, but does not reach the posterior trigonid wall, leaving the talonid basin open. The m2 is more robust
than the m1, and the trigonid is wider and taller than the talonid. The
trigonid cuspids are more widely spaced than on the m1. The paraconid
is close to the metaconid and reduced, but not shifted medially. The
paracristid is arched. The talonid is identical in structure to the talonid of
m1. The m3 is slightly reduced, the trigonid is much wider than the
talonid, and the paraconid is much reduced. The hypoconulid is large; it
is close in size to the hypoconid. The entoconid is not distinct within the
entocristid. The cingulid is well developed, and the postcingulid reaches
the hypoconulid.
CommentsSimpson (1936) described this species as Chriacus
antiquus, and Van Valen and Sloan (1965) assigned it to Oxyclaenus.
Later, Van Valen (1978), following Mannings opinion (collection notes),
referred this species to the genus Baioconodon. Middleton (1983) demonstrated that this species is closer to O. cuspidatus and lacks characters typical of Baioconodon, such as the well-defined constriction between the talonid and the trigonid on the lower molars and the rectangular talonid basin (Fig. 3J). Williamson (1996) referred the species to
Oxyclaenus. This species is almost identical in tooth morphology to O.
cuspidatus and differs in its much larger size, so we agree with
Williamson and treat O. antiquus as the largest species of the genus
Oxyclaenus.
TAXONOMY
Among the seven species assigned to Oxyclaenus, four are valid:
O. cuspidatus, O. simplex, O.antiquus and O. sp. nov. from the Denver
Basin (Middleton and Dewar, this volume). Three species occur in the
Puercan of the San Juan Basin: O. cuspidatus, O. antiquus and O. simplex, and differ from each other mostly in size (Fig. 4).
Gazin (1941) described a new species of Oxyclaenus, O. pearcei,
from the early Paleocene of Utah, based on USNM 16186, left and right
dentary fragments with m2 and m3. In the diagnosis, Gazin mentioned
that O. pearcei is close in size to O. simplex, but differs in having a less
reduced m3 and more developed precingulid. Middleton (1983) noted
that the differences listed by Gazin may be insufficient to distinguish O.
pearcei from O. simplex, considering that the range of morphological
variation of the teeth of O. simplex was much widened by Van Valen

TABLE 6. Upper tooth measurements (in mm) of O. antiquus.


Coll. Number
AMNH 3206
AMNH 23184
AMNH 23181
AMNH 21714

Tooth
M2
M3
M2
M1
M2
P3
M1
M2
M3

Length
7.1
5.9
7.2
8.1
7.1
6
6.9
7.5
5.9

Width
9.3
7.4
8.9
9.2
4.6
7.9
9.1
8.2

(1978), who synonymized Carcinodon filholianus with O. simplex.


Williamson (1996) suggested that O. pearcei is a junior subjective synonym of Chriacus baldwini. Indeed, the lower teeth of O. pearcei do not
have the features diagnostic of Oxyclaenus, such as obtuse cusps, reduced m3 and a small paraconid. But, they have little in common with
Chriacus either. Study of the specimens from the USNM reveals that O.
pearcei is very close in size and tooth morphology to Prothryptacodon
furens from the Fort Union of Montana. So, we regard O. pearcei as a
junior subjective synonym of Prothryptacodon furens.
Simpson (1935) described Chriacus pugnax from the Fort Union
of Montana. Van Valen and Sloan (1965) referred this species to the
genus Oxyclaenus. Van Valen (1978) described a new arctocyonid genus and species, Thangorodrim thalion, and referred Chriacus pugnax
to the newly described genus. But, Robison (1986) described new material of O. pugnax from central Utah and argued that Thangorodrim is
a subjective junior synonym of Oxyclaenus, and that Thangorodrim
thalion is indistinguishable from Oxyclaenus pugnax. Oxyclaenus
pugnax does not possess the dental characters typical of Oxyclaenus
and has bulbous molars with rugose enamel very similar to the species
of Loxolophus. It has an entoconid that is much larger than the
hypoconulid, which is not typical of Oxyclaenus either. When treated
as Loxolophus, it is indistinguishable from Loxolophus priscus in size
and tooth proportions. So, we consider Chriacus pugnax and
Thangorodrim thalion junior subjective synonyms of Loxolophus priscus
and Thangorodrim a junior synonym of Loxolophus.
Oxyclaenus corax was described from the Puercan of Saskatchewan,
Canada (Johnston and Fox, 1984). This species falls well within the size
variation of O. simplex and does not differ significantly in morphology.
Considering the amount of variation in the other Oxyclaenus species,
such as O. cuspidatus, we conclude that O. corax is a junior subjective
synonym of O. simplex.
PHYLOGENY
The systematic position of Oxyclaenus has been the subject of
some discussion. Scott (1892) erected a new family, Oxyclaenidae, to

30
include archaic arctocyonids, which were thought to be creodonts. Osborn
and Earle (1895) referred Chriacus, Loxolophus and Tricentes to a new
family, Chriacidae, which they placed in Primates. Oxyclaenus, Carcinodon
and Mioclaenus were left as incertae sedis. Matthew (1897) noted differences between Oxyclaenus and the Chriacus group. He suggested that
Oxyclaenus is intermediate between the Chriacidae and Triisodontidae,
which is very close to recent suggestions about the taxonomic position of
Oxyclaenus (Prothero et al., 1988; Archibald, 1998).
Matthew (1937) modified Scotts Oxyclaenidae as a subfamily
of Arctocyonidae to include Oxyclaenus, Loxolophus, Protogonodon,
Carcinodon, Tricentes and Mixoclaenus. He also erected a new subfamily, Chriacinae, which was originally described as a separate family, Chriacidae (Osborn and Earle, 1895). Matthew (1937) included
Chriacus and Deltatherium in this subfamily. Simpson (1937) criticized Matthew for erecting a subfamily for Chriacus. Simpson believed
that this genus was rather close to other oxyclaenines and so included
all small arctocyonid genera in Oxyclaeninae. In his definition of
Chriacus, Simpson (1937) included such species as Ch. pugnax and
Ch. antiquus, which are now referred to Loxolophus and Oxyclaenus,
respectively. With the removal of these species, Chriacus becomes a
more distinct genus, clearly different from Oxyclaenus and its allies.
Van Valen (1978) erected a new subfamily Loxolophinae to include Loxolophus (=Protogonodon), Ragnarok, Baioconodon,
Mimotricentes, Desmatoclaenus and Deuterogonodon and placed
Oxyclaenus, Thangorodrim, Chriacus, Protungulatum, Oxyprimus,
Deltatherium, Prothryptacodon and Thryptacodon in Oxyclaeninae.
Later, several other genera were referred to Oxyclaeninae, such as
Princetonia (Gingerich, 1989). It is clear, though, that the Oxyclaeninae
in this sense is not a monophyletic group, because Deltatherium belongs close to Tillodontia and Protungulatum and Oxyprimus differ
much from other arctocyonids. On the other hand, Chriacus (including
Metachriacus, Protochriacus and Spanoxyodon), Prothryptacodon
(=Princetonia) and Thryptacodon are more closely related to the base
of the artiodactyl stock (Rose, 1996), whereas Oxyclaenus and
Baioconodon are considered to be closer to the base of the cetacean
stock. Prothero et al. (1988) and Archibald (1998) modified
Arctocyonidae and removed the genus Oxyclaenus to Cete. Williamson
(1996) mentioned that Oxyclaenus appears to be an arctocyonid.
We believe that Oxyclaenus is not specialized enough to be included in Cete, and its morphology well supports its referral to
Arctocyonidae. On the other hand, Oxyclaenus and Baioconodon do
differ from the Chriacinae. The dental morphology of Oxyclaenus and

Baioconodon is very close to that of the arctocyonids referred by Van


Valen (1978) to Loxolophinae, but does differ in details, such as the
weaker hypocones and better expressed constriction between the talonid and the trigonid on the lower molars. So, we suggest recognition of
two tribes in the subfamily Oxyclaeninae: Oxyclaenini and Loxolophini
(Fig. 5). We therefore suggest the following arrangement of Arctocyonidae:
Order Procreodi Matthew, 1909
Family Arctocyonidae Giebel, 1855
Subfamily Arctocyoninae Giebel, 1855
Genus Arctocyon Blainville, 1841
Genus Arctocyonides Lemoine, 1891
Genus Anacodon Cope, 1882
Genus Mentoclaenodon Weigelt, 1960
Genus Colpoclaenus Patterson and McGrew, 1962
Subfamily Chriacinae Osborn and Earle, 1895
Genus Chriacus Cope, 1883
Genus Prothryptacodon Simpson, 1935
Genus Thryprtacodon Matthew, 1915
Subfamily Oxyclaeninae Scott, 1892
Tribe Oxyclaenini Scott, 1892 stat. nov.
Genus Oxyclaenus Cope, 1884
Genus Baioconodon Gazin, 1941
Tribe Loxolophini Van Valen, 1978 stat. nov.
Genus Loxolophus Cope, 1885
Genus Mimotricentes Simpson, 1937
Genus Lambertocyon Gingerich, 1979
Genus Deuterogonodon Simpson, 1935
Genus Desmatoclaenus Gazin, 1941
Arctocyonidae inc. sedis.
Genus Protungulatum Sloan and Van Valen, 1965
Genus Oxyprimus Van Valen, 1978
ACKNOWLEDGMENTS
R. Tedford made it possible to study specimens in the AMNH, L.
Martin and D. Miao made it possible to study specimens in the KUVP,
and R. Emry made it possible to study specimens in the USNM. Funding from the NMMNH Foundation made this study possible. The work
of PK was supported by the Russian Foundation for Basic Research,
project 02-04-48458, and the Grant Council of the President of Russian
Federation, project NSh-1840.2003.4.

REFERENCES
Archibald, J. D., 1998, Archaic ungulates (Condylarthra); in Janis, C. N.,
Scott, K. M. and Jacobs, L. L., eds., Evolution of Tertiary mammals of
North America: Cambridge, Cambridge University Press, p. 292-331.
Cope, E. D., 1884, Second addition to the knowledge of the Puerco epoch: Proceedings of the American Philosophical Society, v. 21, p. 309-324.
Cope, E. D., 1888, Synopsis of the vertebrate fauna from the Puerco series: Transactions of the American Philosophical Society, new series, v. 16, p. 298-361.
Eberle, J. J. and Lillegraven, J. A., 1998, A new important record of earliest Cenozoic mammalian history: Eutheria and Paleogeographic/ biostratigraphic summaries: Rocky Mountain Geology, v. 33, p. 49-117.
Gazin, C. L., 1941, The mammalian fauna of the Paleocene of Central Utah, with
notes on geology: Proceedings of the United States National Museum, v. 91, p.
1-53.
Gingerich, P. D., 1989, New earliest Wasatchian mammalian fauna from the Eocene
of northwestern Wyoming: Composition and diversity in a rarely sampled highfloodplain assemblage: University of Michigan, Papers on Paleontology, no. 28, p. 1-97.
Johnston, P. A. and Fox, R. C., 1984, Paleocene and Late Cretaceous
mammals from Saskatchewan, Canada: Palaeontographica Abteilung A,
v. 186, p. 163-222.
Matthew, W. D., 1897, A revision of the Puercan fauna: Bulletin of the

American Museum of Natural History, v. 9, p. 259-323.


Matthew, W. D., 1937, Paleocene faunas of the San Juan Basin, New Mexico: Transactions of the American Philosophical Society, v. 30, p. 1-510.
McKenna, M. C. and Bell, S. K., 1997, Classification of mammals above the species level. New York: Columbia University Press, 631 p.
Middleton, M. D., 1983, Early Paleocene vertebrates of the Denver Basin, Colorado [Ph.D. dissertation]: Boulder, University of Colorado, 403 p.
Middleton, M. D. and Dewar, E. W., 2004, New mammals from the early Paleocene
Littleton Fauna (Denver Formation, Colorado): New Mexico Museum of Natural History and Science Bulletin 26, this volume.
Osborn, H. F. and Earle, C., 1895, Fossil mammals of the Puerco beds.
Collection of 1892: Bulletin of the American Museum of Natural History, v. 7, p. 1-70.
Prothero, D., Manning, E. and Fischer, M., 1988, The phylogeny of ungulates; in
Benton, M. J., ed., The phylogeny and classification of tetrapods. Volume 2: Mammals. Oxford, Clarendon Press, p. 201-234.
Robison, S. F., 1986, Paleocene (Puercan-Torrejonian) mammalian faunas of the
North Horn formation, central Utah: Brigham Young University Geology Studies, v. 1, p. 87-133.
Rose, K. D., 1996, On the origin of the order Artiodactyla: Proceedings of the National Academy of Sciences, v. 93, p. 1705-1709.

31
Scott, W. D., 1892, A revision of North American Creodonta with notes on
some genera which have been referred to that group: Proceedings of the
Academy of Natural Sciences of Philadelphia, v. 4, p. 291-323.
Simpson, G. G., 1935, New Paleocene mammals from the Fort Union of
Montana: Proceedings of the U. S. National Museum, v. 83, p. 221-224.
Simpson, G. G., 1936, Additions to the Puerco fauna, lower Paleocene:
American Museum Novitates, no. 849, p. 1-11.
Simpson, G. G., 1937, The Fort Union Group of the Crazy Mountain Field
and its mammalian faunas: U.S. National Museum Bulletin, v. 169, p. 1-

287.
Van Valen, L., 1978, The beginning of the age of mammals: Evolutionary
Theory, v. 4, p. 45-80.
Van Valen, L. and Sloan, R. E., 1965, The earliest primates: Science, v. 150, p. 743745.
Williamson, T. E., 1996, The beginning of the age of mammals in the San
Juan Basin, New Mexico: Biostratigraphy and evolution of Paleocene
mammals of the Nacimiento formation: New Mexico Museum Natural
History and Science, Bulletin 8, 141 p.

32

Dissacus navajovius, skull and lower jaw (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

33

TAXONOMIC REVISION OF THE GENUS ECTOCONUS (MAMMALIA, PERIPTYCHIDAE)


FROM THE PALEOCENE OF THE WESTERN UNITED STATES
SPENCER G. LUCAS1 AND PETER E. KONDRASHOV2
2

1
New Mexico Museum of Natural History, 1801 Mountain Road N. W., Albuquerque, NM 87104;
Department of Biology, Northwest Missouri State University, 800 University Drive, Maryville, MO 64468

AbstractWe revise the species-level taxonomy of one of the most common Puercan mammals from the San
Juan Basin, New Mexico, the periptychid condylarth Ectoconus, based primarily on the study of specimens
collected in the San Juan Basin, New Mexico. Among the four described species of Ectoconus, only two are
valid, E. ditrigonus and E. symbolus. Ectoconus majusculus, originally described from the San Juan Basin as a
more robust species than E. ditrigonus and as having a larger M3, is a junior subjective synonym of E. ditrigonus.
The smaller species of Ectoconus, E. symbolus from the early Paleocene of Utah, is valid, and differs from E.
ditrigonus mostly in size. Ectoconus cavigellii, recently described from the Puercan of Wyoming, is a junior
subjective synonym of E. symbolus. Both Ectoconus species thus have a stratigraphic range of middle to late
Puercan, which undermines the use of Ectoconus, or its species, as index fossils of subdivisions of the Puercan
land-mammal age.
Key words: Ectoconus, Periptychidae, Paleocene, San Juan Basin, New Mexico
INTRODUCTION
Ectoconus (Fig.1 ) is one of the most common mammals in the
early Paleocene (Puercan) strata of the San Juan Basin, New Mexico.
The genus remained monotypic until Matthew (1937) described a new
species, Ectoconus majusculus, that supposedly differed from the type
species in being slightly larger, having more robust teeth and a different third molar structure. Gazin (1941) reported Ectoconus from the
lower Paleocene of Utah and named a new species, E. symbolus (Fig.
2). Recently, Eberle and Lillegraven (1997) found Ectoconus in the
Puercan deposits of Wyoming and referred the material partly to E.
ditrigonus and partly to a new species, E. cavigellii. Study of the collections of the Puercan mammals at AMNH and USNM together with
the recently collected San Juan Basin specimens at the NMMNH provide an opportunity to study the variability of the tooth morphology of
Ectoconus and to assess the validity of the previously described species.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; BYU = Brigham Young University, Utah;
NMMNH = New Mexico Museum of Natural History, Albuquerque;
UCM = University of Colorado, Museum of Paleontology, Boulder,
Colorado; USNM = National Museum of Natural History, Smithsonian
Institution, Washington, D.C.; and UW = University of Wyoming,
Laramie, Wyoming.
SYSTEMATIC PALEONTOLOGY
Order Condylarthra Cope, 1881
Suborder Taligrada Cope, 1881
Family Periptychidae Cope, 1882
Subfamily Periptychinae Cope, 1882
Genus Ectoconus Cope, 1884
Ectoconus Cope, 1884, p. 795.
Ectoconus: Cope, 1888, p. 298.
Ectoconus: Osborn and Earle, 1895, p. 56.
Ectoconus: Matthew, 1937, p. 126.
Ectoconus: Middleton, 1983, p. 282.
Ectoconus: Robison, 1986, p. 104.
Ectoconus: McKenna and Bell, 1997, p. 364.
Ectoconus: Archibald, 1998, p. 312
Ectoconus: Eberle and Lillegraven, 1998, p. 92.

FIGURE 1. Lateral view of skull and lower jaw of Ectoconus ditrigonus,


based on AMNH 16500 (skull) and 16502 (lower jaw) (after Matthew,
1937).

Type species Ectoconus ditrigonus (Cope, 1882) (=E.


majusculus Matthew, 1937).
Included species Type species and Ectoconus symbolus Gazin,
1941 (=E. cavigellii Eberle and Lillegraven, 1998).
Revised diagnosis A large periptychine that differs from
Periptychus in having less folded enamel and a well-developed ectoflexus
on the upper molars. Upper and lower molars are more bunodont than in
Periptychus, with upper molars possessing up to nine well-developed
cusps.
Distribution Early Paleocene (Puercan) of western North
America (New Mexico, Utah, Wyoming, Colorado).
Comments Zhou et al. (1977) described ?Ectoconus from the
Shanghu Formation (early Paleocene) of southern China based on an
almost complete sacrum that generally resembled the sacrum of E.
majusculus, illustrated by Matthew (1937). However, Lucas and
Wiliamson (1995) pointed out that the described sacrum does not differ
from that of one of the most common mammals in the same formation,
the pantodont Bemalambda, so the putative occurrence of Ectoconus in
China can be discounted.

34

FIGURE 2. Occlusal views of cheek teeth of the two species of Ectoconus. A, E.


ditrigonus, left C, P2-M3 (AMNH 16500). B, E. symbolus, left P4-M2 (USNM
16189). C, E. symbolus, left P4-M1 (USNM 16188). D, E. symbolus, left m2-3
(USNM 16190). E, E. ditrigonus, left p4-m3 (AMNH 16502). A and E after
Matthew (1937); B-D after Gazin (1941).

Ectoconus ditrigonus (Cope, 1882)


Figs. 1, 2A, E, 3A-I, 4, 5, Tables 1-2
Periptychus ditrigonus Cope, 1882, p. 465.
Conoryctes ditrigonus: Cope, 1883, p. 968.
Ectoconus ditrigonus Cope, 1884, p. 794, figs. 4-5.
Periptychus ditrigonus: Cope, 1885, p. 40, pl. 23g, fig. 12.
Conoryctes ditrigonus: Cope, 1885, pl. 29d, figs. 2-6.
Ectoconus ditrigonus: Cope, 1888, p. 356.
Ectoconus ditrigonus: Osborn and Earle, 1895, p. 56.
Ectoconus ditrigonus: Matthew, 1937, p. 128.
Ectoconus majusculus Matthew, 1937, p. 129, figs. 27-28, pls. 23-32.
Ectoconus ditrigonus: Van Valen, 1978, p. 62.
Ectoconus ditrigonus: Middleton, 1983, p. 282, pl. 7, figs. 5, 8, pl. 8, fig.
1.
Ectoconus ditrigonus: Robison, 1986, p.104, pl. 2, figs. 5-7.
Ectoconus ditrigonus: Williamson and Lucas, 1992, fig. 12D-E.
Ectoconus ditrigonus: Williamson and Lucas, 1993, p. 125.
Ectoconus ditrigonus: Williamson, 1996, p. 45.
Ectoconus ditrigonus: Eberle and Lillegraven, 1998, p. 92, fig. 13, C.
HolotypeAMNH 3798, right dentary fragment with m2 (Fig.
4A).
Referred specimensFrom the Nacimiento Formation in the
San Juan Basin, New Mexico:
NMMNH locality 317: NMMNH 001, juvenile skull and lower
jaws (Fig. 5; Williamson and Lucas, 1992, fig. 12E); 8671, LM1, LM2;
8681, LM3; 15161, Lp1; 15166, RM1 partial-isolated; 15173, Lp3 (part)m3 (all isolated) (Williamson and Lucas, 1992, fig. 12D) and bone fragments.
From NMMNH locality 646: NMMNH 8842, LM partial-isolated; 8904, M3 partial-isolated; 15232, m1 partial; 15234, p3 partialisolated; 21812, Rp2-isolated.
From NMMNH locality 701: NMMNH 8972, Lc, tooth fragments.
From NMMNH locality 703: NMMNH 19453, dentary fragments with tooth fragments; 19464, Lp3-m3, Rm2-m3 (concreted).

From NMMNH locality 844: NMMNH 15112, Rp4-m3 (all


isolated).
From NMMNH locality 1119: NMMNH 15188, Lp1, Rm3 isolated, bone fragments.
From NMMNH locality 1121: NMMNH 19459, LM1-M3 and
upper tooth fragments (isolated).
From NMMNH locality 1288: NMMNH 16296, Rp4-m3.
From NMMNH locality 2550: NMMNH 19338, Lc1, p4-m3;
19340, RP2-M3, LP4, LM3, Lm3 (isolated), postcrania fragments.
From NMMNH locality 2551: NMMNH 19343, Lp4,m1,m2,m3
& tooth fragments (isolated); 19344, (8) miscellaneous isolated teeth,
bone fragments.
From NMMNH locality 2552: NMMNH 19345, LM3, misc.
upper tooth fragments; 19346, LM2, LM3 and upper tooth fragments
(isolated).
From NMMNH locality 3230: 2757, Lp2-isolated.
From NMMNH locality 3233: NMMNH 2932, Ldp3-m2, RdP4M2.
From Kimbetoh Arroyo, San Juan Basin, New Mexico: USNM
405022, right m2.
From San Juan Basin, New Mexico, collected by Baldwin, 1885:
USNM 5889, left dentary with p3-m1; 5890, maxillary fragment with
right P2-M2, left M1-3.
From San Juan Basin, New Mexico, collected by Gazin, 1936:
USNM 405076, right dentary with p3-m1; 405077, left dentary with
p4-m3, right dentary with m2-3, maxillary fragment with right P3-4,
left P3; 405078, left m2.
From San Juan Basin, New Mexico, collected by Gazin, 1949,
USNM 405011, isolated tooth fragments; 405012, right P4; 405013,
left P2, M2; 405015, left dentary with c1, p2-m1, right p4; 405016, left
p4; 405017, left maxillary fragment with M1-2.
From San Juan Basin, New Mexico, collected by Gazin, 1964:
USNM 405018, isolated tooth fragments; 405019, left dentary with
p2-m3, right dentary with p2-m2, maxillary fragment with RP4, RM3.;
405020, right m2; 405021, left dentary with p2-m3, right dentary with
p2-m3; 405080, left dentary with p2-m3, right dentary with p4-m3;
405081, left dentary with p3-m3, right dentary with p2-m3 (Fig. 4DE); 405090, broken upper lower teeth; 405090, isolated upper and lower
molars.
From 3 miles East of Kimbetoh Arroyo, San Juan Basin, New
Mexico: USNM 405008, right dentary with m1-3; 405009, left m2.
From Tsosie Arroyo, San Juan Basin, New Mexico, collected by
Gazin, 1936: USNM collection: 405079, right dentary with m1-3, left
M2; 15435, left dentary with p4-m3; 405023, right M1.
From the first arroyo west of Kimbetoh, San Juan Basin, New
Mexico: USNM 14549, portions of skull and lower jaws.
From Kimbetoh Arroyo, near the yellow hill, in the center of the
basin, San Juan Basin, New Mexico: USNM 15436, right dentary with
p3-m1; 15437, left dentary with p3-m3, right dentary with p3-m3 (Fig.
4I), maxillary fragment with right P4-M1, left P2-M3 (Fig. 3D); 15458,
maxillary fragments with RP3-M3, LP4-M1, LM3, dentary fragments
with Lp3-m1, rp4; 15459, left dentary with p3-m2, right dentary with
p4-m3; 405010, skull fragments with LP4-M3, RM2-3.
From 2 miles above Ojo Alamo Arroyo, San Juan Basin, New
Mexico: AMNH 10490, left maxillary fragment with P2-M3.
From 3 miles east of Kimbetoh, San Juan Basin, New Mexico:
AMNH 16495, left dentary with p2-m3, right dentary with p2-m3 (Matthew, 1937, pl. 23, figs. 4-5); 16496, right maxillary fragment with P4M3 (Fig. 3A), left maxillary fragment with P3-M3 (Fig. 3B), left dentary
with p2-m2 (Fig. 4B), right dentary with p2-4, m2-3 (Fig. 4F-G; Matthew, 1937, pl. 23, figs. 1); 16501, left maxillary fragment with P4M3, right dentary with p3-m2; 16505, left p1-3, m2, m3, M1-3, right
m1-3.
From Barrel Springs Arroyo, San Juan Basin, New Mexico:
AMNH 27670, isolated left P4-M2; 27721, right dentary with m2-3;

35

FIGURE 3. Occlusal views of upper cheek teeth of Ectoconus ditrigonus (A-I) and E. symbolus (J). A-B, AMNH 16496, right P4-M3 (A) and left P3-M3. C, AMNH
27603, right P3-M3. D, USNM 15437, left P4-M3. E, AMNH 27602, very worn right P2-M3. F, AMNH 3813, left P4-M3. G, AMNH 16494, right DP4-M1. H, AMNH
880, left P2-M3. I, Seven M3s catalogued as AMNH 58197 and 93219, showing wide range of variation. J, USNM 16189, left P4-M2 (holotype of E. symbolus).

27725, right dentary with m1; 27726, left dentary with p4-m2 (partial,
concreted).
From Chaco Canyon, San Juan Basin, New Mexico: AMNH 886,
right M1, tooth fragments; 888, humerus, left and right concreted
dentaries, right maxillary fragment with P3-M3.
From Coal Creek Canyon, San Juan Basin, New Mexico: AMNH

887, left P3.


From Eduardo Arroyo, San Juan Basin, New Mexico: AMNH
27602, maxillary fragments with right P2-M3 (Fig. 3E), left p3-M3;
27603, maxillary fragments with right P3-M3 (Fig. 3C) and left P3-M3
(isolated); 27605, right c, p3-4, m2 (isolated), left c, p4, m2, m3 ; 27632,
left m2; 27633, left P3-4 (isolated); 27634, right dentary with p4-m3,

36
TABLE 1. Measurements (in mm) of selected upper cheek teeth of Ectoconus.
P1 L
E. ditrigonus:
NMMNH 2932
NMMNH 19340
NMMNH 19340
NMMNH 19335
NMMNH 8671
NMMNH 8657
NMMNH 19346
NMMNH 8681
USNM 15437
USNM 15437
USNM 5890
USNM 5890
USNM 405013
USNM 405077
USNM 405077
USNM 405012
USNM 405079
USNM 15459
USNM 15459
USNM 405010
USNM 405010
USNM 405010
USNM 15458
USNM 15458
USNM 405019
AMNH 27603
AMNH 27603
AMNH 27661
AMNH 27656
AMNH 27670
AMNH 16494
AMNH 16494
AMNH 16143
AMNH 93219
AMNH 93219
AMNH 93219
AMNH 93219
AMNH 93219
AMNH 93219
AMNH 93219
AMNH 27633
AMNH 27602
AMNH 27602
AMNH 3821
AMNH 3830
AMNH 3827
AMNH 3823
AMNH 3823
AMNH 3803
AMNH 3803
AMNH 3800
AMNH 3800
AMNH 3812
AMNH 3812
AMNH 3813
AMNH 3801
AMNH 3828
AMNH 3802
AMNH 3802
AMNH 3822

P1 W P2 L

7.15

P2 W

P3 L

P3 W

P4 L

P4 W

9.28

8.61
8.86

11.91
11.90

8.40

13.31

8.19

12.30
6.80

9.15

6.80

9.26

7.70

7.95
8.06

7.00

6.70
6.50

9.60

8.50
9.00

11.49
11.59

7.69

11.14

7.90
7.70

12.00
11.90

8.40
8.20
8.10

7.10

11.05

10.90
12.10
11.80

8.43
8.12
8.16

M1 W

M2 L

M2 W

M3 L

M3 W

9.79

12.03

10.06

12.80

9.87
9.88
9.96
10.93

13.27
14.88
14.59
14.75

9.08
8.55

13.38
12.82

10.19
9.68

14.11
12.72
9.33

13.85

10.50
9.95
9.45
10.09
11.10

14.50
14.87
13.81
14.58
14.41

8.25
8.63
8.23

12.45
12.34
12.65

8.16

11.30

9.53
9.46
10.68
9.84
9.08
10.12
9.70

13.62
14.79
13.74
15.18
14.22
12.66
14.35

9.80
9.40
10.30

15.10
14.50
13.70

8.72
8.62
8.26
8.04
9.05
8.91
9.31
9.84
8.90
8.10

12.77
13.89
12.11
12.42
12.86
12.16
11.22
12.66
13.50
12.90

8.90

14.40
12.70

9.10
8.20
8.80
8.20
8.10
8.10
8.20

13.60
11.30
13.20
12.10
12.90
13.00
11.40

7.30
7.70
8.90
8.90
8.90
8.70
8.20
8.90
7.90

11.80
13.80
13.60
14.10
12.60
11.90
12.50
10.90
12.90

11.91

12.65
12.48
11.65

8.25

11.52

8.45

12.25

7.31

12.62

8.06
8.14
8.94
7.90
7.80

12.32
11.71
12.30
12.30
12.80

8.60
8.90

12.00
11.80

8.80
8.40

M1 L

13.30
12.20

10.15
10.35
9.69
9.92

15.10
14.73
12.65
13.02

9.94
10.09
9.79

14.86
14.86
15.16

9.49
9.54

14.31
13.52

9.10
9.50
9.20

13.80
13.40
13.90

9.40
8.80
10.00
9.30
10.70
9.90
9.90

12.10
12.90
12.90
14.00
12.20
14.50
13.60

8.30
10.00
10.00
9.80

12.20
13.80
14.40
14.00
15.00

7.10
7.90
7.90

9.90
11.30
11.10

8.60
8.00
8.10
7.90

11.40
12.70
12.10
11.60

10.00
9.70
10.00

13.80
14.30

8.90
8.70
9.90

13.70
15.20
16.10

7.70
7.20

11.80

6.90

12.10

8.90
9.10

14.90
14.80
13.10
14.80
14.60

9.30
9.00
9.10
9.60

13.40
16.30
11.90
13.80

12.90
13.80

10.00
9.90
9.80
9.90

13.90
13.20
16.00
14.20

8.90
8.50
7.60

12.10
12.70
12.70

13.20
11.40
12.60

9.60
10.10
10.30

12.10
13.90
14.70

9.40
8.90
8.60

12.80
12.90
13.40

7.10
7.40

10.30
10.60

7.30
7.30
7.90

11.40
11.20
6.80

9.90
10.20
11.20
9.80
9.90
10.00

10.80

8.80
8.20

12.10
11.80

9.00
9.50
9.30

37
Table 1 continued.
P1 L
AMNH 3822
AMNH 3824
AMNH 880
AMNH 880
AMNH 3804
AMNH 884
AMNH 58155
AMNH 58155
AMNH 59967
AMNH 58158
AMNH 58179
AMNH 58191
AMNH 58197
AMNH 58078
AMNH 58156
AMNH 58111
AMNH 58046
AMNH 58108
AMNH 58108
AMNH 16496
AMNH 16496
AMNH 888
AMNH 10490
AMNH 16501
AMNH 16505
KU 8128
Mean
Std. Dev.
N
E. symbolus:
USNM 16188
USNM 16189
USNM 16214
AMNH 36073
Mean
Std. Dev.
N

P1 W P2 L

P2 W

P3 L

P3 W

P4 L

P4 W

M1 L

M1 W

M2 L

M2 W

M3 L
8.70

6.80
7.10

8.50
8.20

11.60
11.90

8.20
7.90

7.40
11.70
12.60

11.30
10.10

13.60

10.10

15.90

9.10

13.40

8.00
8.20

8.20
10.60
10.50

9.90
13.50
9.90
9.90

14.40
8.20
13.60
13.10

8.20
11.30
8.10

11.60

8.20

13.10

10.10

14.60
9.20

13.10

9.30

14.00
8.40

14.10

9.60
9.80

7.20
5.10

9.30
9.50

4.70

12.00
8.40

7.10

8.70
8.20

8.60
11.60
10.90
11.30

10.00

13.30

M3 W
13.20

11.40

12.10
10.90

13.80

12.60

9.40

11.00
8.60
9.10

13.60
13.40
9.80
11.10
12.60

11.83
8.17
0.53
39

9.42
11.95
1.05
42

10.20
10.90
12.90
9.90
10.00
9.20
12.31
9.81
0.51
47

14.90
13.90
14.70
14.70
9.60
12.70
13.70
13.10
9.93
13.64
0.95
47

8.18
8.14

10.28
10.03

8.79

12.10

8.16
0.03
2

10.16
0.18
2

8.79
1

9.00
7.40

5.10

4.70

8.00

9.60
8.90
8.90

11.00
11.00
12.20

8.20

8.10
8.00

10.90
8.40

8.00
9.90

7.06
0.44
12

9.08
0.61
12

10.78
8.08
0.58
27

11.25
0.67
29

6.44

10.10

7.99

7.35

8.43

7.66

7.44
1.41
2

8.88
1.73
2

0.00

0.00

7.99

7.35

8.90

left m2 isolated; 27635, right dentary with p4-m1, m3.


From first wash of Kimbetoh Arroyo, San Juan Basin, New Mexico:
AMNH 93220, right dentary with p4-m3 (Matthew, 1937, figs. 27-28).
From first wash west of Kimbetoh Arroyo, San Juan Basin, New
Mexico: AMNH 16502, left dentary with p4-m3 (Figs. 1, 2E).
From Kimbetoh Arroyo, San Juan Basin, New Mexico: AMNH
16500, skull and nearly complete skeleton (holotype of E. majusculus)
(Figs. 1, 2A; Matthew, 1937, figs. 27-29, pl. 23, figs 2, 10, pls. 24-32);
27656, isolated left P4, m2; 27661, isolated left dP4, M1, M2; 27662,
right p2, p4-m3 (isolated); 27664, left and right dentaries in matrix;
93235, maxillary fragments with right M1-2 (concreted).
From San Juan Basin, New Mexico, collected by Baldwin, 1882:
AMNH 3799, left dentary with m2-3; 3800, right dentary with dp3-4
(Fig. 4G), m1, left dp3-4, left maxillary with dP4-M1, right maxillary
with dP2-M1 (Cope, 1885, pl. 29d, figs. 2-4; Matthew, 1937, pl. 23,
figs. 6-9).
From San Juan Basin, New Mexico, collected by Baldwin, 1885:
AMNH 3816, skeletal fragments, femur, pelvis, vertebrae, right m2-3;
3818, broken upper molars, postcrania fragments.
From San Juan Basin, New Mexico, collected by Baldwin, 1887:
AMNH 3476, bone and tooth fragments; 3723, left p4, m2, right m2
(isolated); 3739, left dentary with p3-m3; 3802, left maxillary with P3-

9.90
10.80
10.30
10.90
13.20
9.90
9.60
9.00
14.64
9.78
0.58
54

14.90
13.10
16.30
15.80
7.60
13.60
13.80
13.10
7.74
14.17
1.00
57

12.10
8.20
9.20
7.30
12.01
8.51
0.56
52

12.10

5.91
5.91

8.95
8.95

0.00

0.00

9.50

14.80
15.00
12.20
13.10

12.62
1.03
53

M3, right maxillary with P4-M3, right dentary with c1, p2-m3, left
dentary with p2-3, m3; 3804, right maxillary fragment with M2-3, right
dentary with m1-3; 3805, right dentary with p4-m3; 3806, left dentary
with p3-m3, right dentary with p1-m3 (concreted); 3813, left maxillary
with P4-M3 (Fig. 3F); 3820, isolated broken teeth; 3821, right M3, left
P4, M1 (partial), M2 (isolated); 3822, left m2, M3, right m3, M2, M3
(isolated); 3824, left maxillary fragment with P4 and broken molars;
3825, left m1; 3829, tooth fragments; 3830, left dentary with p3-4,
right dentary with rm2; right M3.
From San Juan Basin, New Mexico, collected by Baldwin:
AMNH 3780, left dentary with p3-m2; 3828, right M1, tooth fragments.
From San Juan Basin, New Mexico: AMNH 880, right maxillary fragment with P2-4, left maxillary fragment with P2-M3 (Fig. 3H),
left dentary with p4, m2-3 (Matthew, 1937, pl. 23, fig. 3); 884, right
dentary with p1-4, left dentary with m1-3, left maxillary fragment with
P4, M2-3; 3694, postcranial fragments; 3799, postcranial fragments;
3801, left maxillary with P4-M1 (broken); 3803, left dentary with p2m2 (Fig. 4C), maxillary fragment with right M1-3; isolated right P2,
left P2, P3, M2, M3; 3812, left maxillary with P3-M3, right maxillary
with P3-M3; 3817, left dentary with p2,m1; 3823, concreted maxillary
bones with left LP3-M3, right M1-3, right dentary with p2, p4-m3;
3827, left M3; 93219, isolated teeth (Figs. 3I, 4J); 93439, right p3-4, m3,

38
TABLE 2. Measurements (in mm) of selected lower cheek teeth of Ectoconus.
Collection Num.
E. ditrigonus:
NMMNH 2932
NMMNH 30506
NMMNH 19464
NMMNH 15188
NMMNH 8972
NMMNH 16296
NMMNH 19338
NMMNH 15173
NMMNH 19340
NMMNH 19344
NMMNH 19344
MMNH 19343
NMMNH 15112
NMMNH 15161
NMMNH 2757
USNM 405081
USNM 405081
USNM 15437
USNM 15437
USNM 5889
USNM 405016
USNM 405077
USNM 405077
USNM 405015
USNM 405015
USNM 405008
USNM 405077
USNM 405077
USNM 405009
USNM 15436
USNM 405080
USNM 405080
USNM 405078
SNM 405079
USNM 405020
USNM 15459
USNM 15459
USNM 15458
USNM 15458
USNM 405019
USNM 405019
USNM 405021
USNM 405021
USNM 405022
AMNH 16495
AMNH 16495
AMNH 27656
AMNH 27726
AMNH 27725
AMNH 93220
AMNH 27662
AMNH 93439
AMNH 16494
AMNH 27635
AMNH 16503
AMNH 93219
AMNH 93219
AMNH 93219
AMNH 3816
AMNH 27632

c L c W p1 L p1 W p2 L p2 w

p3 L p3 W p4 L p4 W m1 L m1 w m1 t.w. m2 L m2 w m2 t.w. m3 L m3 w m3 t.w.


10.10 9.10

4.07

8.27

10.16 8.49
11.39 9.26

8.38

3.96

13.76 9.99
12.47 8.91

8.45
7.79

11.97
11.61
13.14
13.04
12.44

7.21
9.18
8.07
8.61

5.46 6.72
6.60 6.61

6.45

4.45

9.24

6.17

7.96
8.32 7.85
10.29 8.91

9.91 9.20
11.42 9.92

8.50
9.23

11.13 10.45 9.26


10.28 9.06

10.39 9.29
10.62 9.31

9.08
9.09

9.14
8.06
8.61

10.40 9.89
10.15 8.73
9.25

8.28
8.49
7.82

9.84
9.86

7.91
5.21

9.77 8.86
10.04 9.93
0.00
10.76 9.89
10.97

9.30
10.01
10.02
10.49

11.27 10.79 9.83

13.61 10.37 8.20


8.66

11.24 10.17 9.38


10.40 10.01 9.81
10.42 10.09 10.22
9.33

12.47
12.82
12.20
12.40

9.54
9.20
9.56
9.39

7.65
8.03
8.09
7.32

9.10

9.37
8.01
9.71
8.90
8.93

9.40

7.69

10.93 10.06 9.35


12.40 11.30 9.48

11.20 9.18
12.26 9.36

8.17
7.99

9.74

9.09

9.02
9.18

8.36
8.74

10.38
11.08
10.55
10.57
10.59
10.86
10.84

8.13

5.09
6.75
5.95

6.16 5.92

10.91
10.63
11.46
0.00
11.17
11.14

6.60
9.79

6.34
6.30

6.01
7.69

8.90
8.59
8.42
8.60
8.18

7.72
7.15
7.44
8.26

8.26 7.17
10.25 8.89

9.05
8.70
7.85
9.01
8.76
9.38
8.27

8.36 8.24
9.26 8.77
10.01 9.58
9.06

6.27

8.84

8.61
8.56

7.69
7.92

7.99

10.08 7.99
8.39 8.89
8.14 8.96

8.83
9.82
9.68

10.21 9.88 8.92


10.02 10.54 9.15
10.02 9.85 9.41
10.78 9.59

8.47
8.14

7.96
7.02

7.45
8.47
6.89
6.74

7.08
6.51
5.60
5.51

10.95
9.79
6.68
8.44

8.10
8.29
7.09
7.18

7.80
7.90

6.40
5.70

8.40
8.90

7.90
7.40

8.00

6.2

7.2

8
8.80

8.24
8.98
8.57
8.49
10.54
10.13
9.14

9.26
9.17
7.88
8.08
9.04
9.30
8.37
8.20

11.09 10.11 9.59


10.60 10.25 9.63
8.49

8.60
8.70

8.30
7.30

10.80 8.20
10.00 8.20

7.10

8.60
9.20
8.30

6.90
8.20
8.20

7.3

7.9

8.3

5.60

7.30

8.76

8.60 8.20
8.80 8.20
8.60 8.40

11.51
11.56
11.50
10.06

10.54
10.61
9.40
9.27

10.21
10.07
8.36
8.58
8.00
8.00

10.80
10.10 8.20 7.90
10.30
10.10 9.20 8.20
10.8
9.6
11.10
10.60

9.6
9.7
9.40
9.90

7.9
8.9
9.00
9.00

10.09
10.29
10.25
9.96
10.84
10.60
10.01
9.41
11.12 9.45
10.21
10.66 10.99

9.49
9.45
9.55
8.89
10.25
10.25
8.91
9.07
9.19
9.91
10.21

11.76 9.79
12.59 9.47

12.00
12.00
11.02
11.12
10.72
11.10
10.90
11.20
11.10

11.57
11.51
9.46
9.50
9.57
9.40
9.70
8.40
9.20

14.09 10.52 9.26

10.70 9.50
10.50 9.90

9.00
9.70

12.10 9.90 9.20


12.90 9.10 7.90
12.90 10.20 8.00

11.1

10

8.9

12.10
10.90
11.60
11.30
10.10
11.60

11.00 10.70
10.70 9.60
11.20 10.00
10.60 10.10
9.10 8.60
11.20 9.90

11.84
11.76
10.16
9.99
9.99
10.00
9.60
9.60
10.10

13.36 9.71 7.98


11.68 9.58 8.05
11.51 10.91 8.56
13.48 9.25

8.03

12.50 9.15

7.79

11.40
11.60 9.20

7.60
8.00

10.30

7.40

10.8
12.90
12.20
12.10
12.50
11.90

9.1
10.30
9.80
9.00
10.00
8.60

7.8
8.80
7.80
7.90
8.40
7.80

39
Table 2 continued.
Collection Num.
AMNH 27632
AMNH 27605
AMNH 27605
AMNH 3825
AMNH 3830
AMNH 3817
AMNH 27634
AMNH 27634
AMNH 3823
AMNH 3803
AMNH 3800
AMNH 3805
AMNH 3805
AMNH 3806
AMNH 3802
AMNH 3802
AMNH 3822
AMNH 3820
AMNH 880
AMNH 3804
AMNH 884
AMNH 3739
AMNH 3723
AMNH 3723
AMNH 58155
AMNH 58155
AMNH 58198
AMNH 58161
AMNH 58218
AMNH 58231
AMNH 58150
AMNH 58082
AMNH 58074
AMNH 58011
AMNH 885
AMNH 885
AMNH 3780
AMNH 3799
AMNH 16496
AMNH 16496
AMNH 888
AMNH 888
AMNH 16502
AMNH 16501
AMNH 27721
AMNH 16505
AMNH 16505
Mean
Std. Dev.
N
E. symbolus:
USNM 16190
USNM 18518
BYU 3794
BYU 3788
BYU 3788
Mean
Std. Dev.
N

c L c W p1 L p1 W p2 L p2 w
7.30 6.50

p3 L p3 W p4 L
9.30
8.80 7.80 9.00
8.90
8.60

7.20

6.80
6.40

5.90 5.60
5.40 5.90

6.90
7.40

7.40

9.20

p4 W m1 L m1 w m1 t.w.
8.90 10.20 9.70 9.10
8.20
8.60
8.90
8.10
10.30 9.30 8.20
8.70 10.60 9.30 9.70

8.60
8.80

9.20
8.10

8.90
7.30

7.90

6.30

6.50
5.90

6.20
6.20

8.10

7.30
8.30
8.30

7.00

7.90
7.90
8.10

8.80

10.30 8.40
10.60 9.10
11.00 10.00
10.10

9.20
8.30
9.80
9.20

8.00
8.90

9.80 8.90
10.90 8.60

8.30
8.90

m2 L
10.20
11.10
10.90

m2 w
10.10
10.10
10.20

10.10 9.90

m2 t.w. m3 L m3 w m3 t.w.
9.30
13.20 10.10 8.20
9.10
10.10
11.30 9.80 8.10
9.20

11.10 10.50 10.20


11.30 10.60 9.90
11.00
8.90
10.30 9.90 8.90

12.90 9.90

11.50 11.30 11.10

11.80
11.80
11.10
11.90
12.30
13.20
14.20
12.10
11.10
11.10

9.90 9.70
11.10

8.90

11.10 10.50 9.50


7.80
10.30 8.60
3.90

3.10

7.70

7.90
7.60

5.10

7.90
7.00

6.50
7.10

6.80
6.90

8.10
8.10
9.10

7.80

9.20
9.30

7.80
8.90

9.00

8.60

11.20
11.00
10.10
10.00
10.60
10.60
11.70

11.70
9.90
8.50
9.50
9.90
9.70
8.80

10.30

8.00

9.10
7.90
8.50

9.90 8.70
10.00 8.80
9.90 9.20

8.20
8.10
6.90

10.30

9.80

10.20 9.90
10.00 9.90

9.10
8.90

9.10

8.30

11.20 9.40

9.00

8.90

10.10
10.10
12.10
12.10

9.80
9.40
10.70
10.50

8.20

8.80
9.10
8.90
8.80
7.90
8.90

8.50

9.70

9.90
9.90
9.10
9.20
9.10
10.00
10.20
9.60

8.20
7.80
8.10
7.80
7.20
7.00
8.10
8.10

8.20

6.90
6.90

12.10 8.60
12.80 9.20

7.80
7.90

11.10 8.40
12.10 8.30

7.90
7.00

12.10 9.90

8.20

10.40
9.70
9.60
8.80

9.10
8.40
8.40
7.80

10.10
9.20
9.10
9.56
0.59
56

8.20
7.30
7.00
8.00
0.54
57

7.10 6.60

7.80 7.50

6.70

5.80

6.90

6.30

7.60

6.30

8.40

7.60

9.10

7.90

9.20

8.40

10.50

9.70

8.10

9.80

9.00

11.20 10.30 9.70

9.30

8.10

6.90

6.10

8.10

7.60

6.47 6.42 4.32


0.88 0.60 0.60
8
8
4

4.04
0.82
4

7.20
0.83
26

6.29
0.85
26

8.58
0.90
35

7.61
0.60
35

8.10

7.10

4.00

4.29 4.24 4.8 5.01


4.29 4.24 4.80 5.01

9.10
8.40

7.00

4.10

4.93 4.63
6.69
5.81 4.63
1.24
2
1

8.60
8.90

7.44 6.30
6.09
7.44 6.20
0.15
1
2

9.20
9.10

8.85
0.65
54

7.41
7.43
7.42
0.01
2

8.10
7.90

8.41
0.53
55

7.09
6.92
7.01
0.12
2

10.80 9.40
10.40 8.50

9.80
10.33
0.66
54

8.55
7.82
8.19
0.52
2

9.00
9.38
0.66
52

7.52
8.03
7.78
0.36
2

8.90
8.30

7.90
8.80
0.66
57

7.31
7.10
7.21
0.15
2

11.00 10.20 9.80

10.00
10.00
11.90
11.30

10.70 9.90
9,29,1
11.20 10.90
10.90 10.00
10.90 8.20
10.80 10.07
1.43 1.44
69
69

9.20

13.20
12.70
12.50
12.10

10.60
8.90
8.70
9.55
0.71
67

13.20
11.90
12.50
12.24
0.95
59

10.15 9.13

8.18

10.65 8.14

9.07

7.97 7.01

10.30 7.39

6.82
6.76
6.27

9.56
9.59
0.54
3

8.13
8.41
0.63
3

10.48 7.77
0.25 0.53
2
2

6.62
0.30
3

7.37
7.52
0.60
3

40
M2 (isolated).
From the Head of Coal Creek Canyon, San Juan Basin, New
Mexico: AMNH 885, left dentary with m1-3, right dentary with m3.
From three miles above Kimbetoh Arroyo, San Juan Basin, New
Mexico: AMNH 10506, left p4-m3 (isolated).
From three miles east of Kimbetoh Arroyo, San Juan Basin, New
Mexico: AMNH 16503, right dentary with m1-3.
From Tsosie Arroyo, San Juan Basin, New Mexico: AMNH
58000, broken right M1; 58011, right p4; 58045, broken left M2; 58046,
right M3; 58052, left m2; 58074, right dentary with c1, p2-m1; 58078,
left P4; 58097, molar fragments; 58108, left P3, M1, right M2; 58111,
left M1, right m2; 58119, right m3; 58150, left p2, m2; 58155, isolated
teeth; 58156, right P4, M1; 58158, right P1, P4, M1, M2; 58161, right
p4; 58179, left M3; 58191, left M2, m1; 58197, left M3 (Fig. 3I); 58198,
right dentary w m2-3; 58218, right c; 58231, left p3, m2; 58966, broken M2; 59967, right M3, left P4, M1 (partial), M2 (isolated); 5816a,
molar fragments.
From two miles above Ojo Alamo Arroyo, San Juan Basin, New
Mexico: AMNH 16143, isolated right P3 and left M2-3; 16494, maxillary fragments with right P3-4, dP4, M1 and left M1-2, right dentary
with m1-3.
From Kansas University loc. 4, San Juan Basin, New Mexico:
KU 8128, left maxillary fragment with P3-M3.
From the North Horn Formation at Dairy Creek, Wagonroad,
Utah: BYU 3753, right dentary with m1-3; BYU 3778, Rm2 (Robison,
1986, pl. 2, fig. 6); BYU 3777, Lm3 (Robison, 1986, pl. 2, fig. 7); BYU
3845, LP4-M1 (Robison, 1986, pl. 2, fig. 5); BYU 3847, RM1, broken,
isolated.
From the Denver Formation in the Denver Basin, Colorado: UCM
39108, right dentary with p2-m3 (Middleton, 1983, pl. 7, fig. 5, pl. 8,
fig. 1); 40531, right maxillary with P3-M2 (Middleton, 1983, pl. 7, fig.
8).
From the upper Ferris Formation, western Hanna Basin, Wyoming: UW locality V-91022: UW 26267, RM3; UW locality V-91024:
UW 26493, Lp3-m1 (Eberle and Lillegraven, 1998, fig. 13C).
Revised diagnosisLarge species of Ectoconus that differs from
E. symbolus by its larger size (15-40 %) and paraconid on the p4 always present.
DistributionEarly Paleocene (Puercan) of New Mexico, Colorado, Utah and Wyoming.
DescriptionThe canine is relatively larger than in E. symbolus.
The P1 is rather simple and oval in cross section. The P2 and P3 are
similar in structure and differ only in sizethe P3 is much larger. The
paracone is much larger than the protocone. The cingulum is well developed, almost circular and terminates only lingual to the protocone.
The P4 is more molarized, and the protocone and paracone are almost
equal in size. There is no metacone. Both paraconule and metaconule
are well developed and lie close to the protocone. There is a small
parastyle and weaker metastyle. There is no hypocone or pericone, but
the anterior and posterior cingula are both well developed. The labial
cingulum is rather strong and connects the parastyle and the metastyle.
The M1 and the M2 are very similar in shape, size and morphology. The parastylar lobe of the M1 is smaller than the metastylar lobe,
whereas the M2 has a parastylar lobe that is larger than the metastylar
lobe. But, a well-developed cuspule is present only on the metastylar
lobe of both molars. All major cusps are close in size, the protocone
being slightly larger than the paracone and the metacone. The parastylar
lobe is large and bears a small cuspule; the metastylar lobe is more
inflated, but bears a large cuspule almost equal in size to the metaconule. Both the paraconule and the metaconule are well developed
and equal in size. They lie very close to the protocone, and the
preprotocrista and postprotocrista are rather weak. The mesostyle is
well expressed and has the shape of a small crest on M1 and the shape
of a round cuspule on M2. The hypocone is rather large and situated
posterolingual to the protocone. The pericone is slightly smaller than the

hypocone, but well expressed.


The M3 is rather variable in shape and structure. Usually it is
trapezoidal or triangular in shape, having a much reduced metaconule and
the paracone as the largest cusp. The parastyle is large, with one or two
accessory cusps. The paraconule is usually larger than the metaconule.
Sometimes there is an additional cusp labial to the metaconule. Both the
hypocone and the pericone are usually developed, and the hypocone is
always larger. The anterior and posterior cingula on all the molars are
rather weak, whereas the labial cingula are rather strong.
The lower teeth are much like those of E. symbolus but larger.
The p1 is simple and round in cross-section. The p2 has a very small
talonid, bearing one small cuspulid. The p3 is more molarized; it has a
small talonid with two or three cuspids: hypoconid, hypoconulid and
sometimes a very small entoconid. The trigonid consists of two cuspids, closely adjoined to each other, so on much worn teeth they cannot
be distinguished. The labial cingulid is well expressed on all the
premolars. The p4 has a talonid, consisting of three cuspids, and the
paraconid is always present. The protoconid and the metaconid are almost equal in size. The talonid is the same length as the trigonid, and it
consists of three cuspids: the largest cuspid is the hypoconid, the entoconid is the smallest.
The m1 is the smallest lower molar; the trigonid is slightly wider
than the talonid. The cristids are somewhat better developed than in E.
symbolus. The trigonid consists of three major cuspids. The paraconid
is a little smaller than the protoconid and the metaconid and is closely
attached to the metaconid, closing the trigonid basin. Both the paracristid
and the metacristid are well developed. The cristid obliqua begins at
the labial wall of the metaconid. The talonid basin is relatively wider
than in E. symbolus. The hypoconulid is the largest talonid cuspid, and
the hypoconulid and entoconid are usually equal in size. The entocristid
is low and wide and closes the talonid basin lingually.
The m2 is much larger than the m1, but very close in morphology. The hypoconid is situated more lingually, so the talonid basin is
relatively smaller than on the m1. There usually is a small cuspulid
lingual to the paraconid. The hypoconulid is somewhat larger than the
entoconid. The entocristid is well expressed.
The m3 is much variable in shape, from oval to triangular, and
the trigonid is always wider than the talonid. As on the m2, sometimes
there is a small cuspulid. The hypoconid is shifted lingually, and the
cristid obliqua begins between the protoconid and the metaconid. The
hypoconulid is usually slightly larger than the hypoconid, and connected
to the former by a weak cristid. The labial cingulid is very strong on all
the molars.
Juvenile skull, jaws and dentitionNMMNH P-001 is an incomplete juvenile skull and lower jaws of Ectoconus ditrigonus (Fig.
5) with left DP2-4, M1, right DP3-4, M1 and left dp2-4, m1, right dp24, m1. The skull is dorso-ventrally crushed, missing most of the premaxillae (no incisors are preserved) and most of the basicranium and
occiput. The dentaries are separated at the symphysis, and no incisors
or canines are present.
In dorsal view, the skull has a relatively short rostrum (rostrum
length is about 50% of total skull length) and is widest across the anterior roots of the zygomatic arches. The nasals are long but broaden
posteriorly to their digitate suture with the frontals posteriorly. The
skull gently narrows posterior to that suture and then widens again at
the posterior roots of the zygomatic arches. The frontals cover this postorbital constriction but are not raised to a sharp sagittal crest, as in the
adult skull of E. ditrigonus illustrated by Matthew (1937, pl. 24). Total
length of the juvenile skull is estimated at ~ 90 mm, and width across
the anterior end of the orbits is ~ 52 mm.
In left lateral view, the large infraorbital foramen above the DP3
opens anteriorly. The fact that the sagittal crest is not as well developed as in the adult skull is especially evident. Nevertheless, a detached piece of the posterior portion of the skull roofs shows that there
was some development of a sagittal crest posterior to the orbits. The

41

FIGURE 4. Occlusal views of lower cheek teeth of Ectoconus ditrigonus. A, AMNH 3798, left m2 (holotype of E. ditrigonus). B, AMNH 16496, left p2-m2. C, AMNH
3803, left p2-m2. D, USNM 405081, right p2-m3. E, USNM 405081, right p3-m3. F, AMNH 16496, right p2-4. G, AMNH 3800, left dp3-4. H, AMNH 16496, right m23. I, USNM 15437, right m2-3. J, Five m3s catalogued as AMNH 93219 showing wide range of variation.

anterior edge of the orbit is above the anterior edge of the M1.
Ventrally, the palate is broad and narrows relatively little from
posterior to anterior. The internal nares (choanae) open immediately
behind a transverse line through the M1s.
A relatively large canine alveolus is anterior to the DP2 and is at
least as large as the combined DP2 alveoli. The DP2 has three roots
and has a small, low crown mostly composed of a single cusp, the paracone. A low, slightly cuspidate cingulum extends around the anterior
edge of the crown base, and a more pronounced cingulum surrounds
the posterior crown base. A posterior crest on the paracone slope divides this posterior cingulum. A weakly developed ridge (rib) is also
present on the labial face of the paracone.
The DP3 is a triangular tooth in occlusal view. It has a large

paracone and a slightly smaller and more labially positioned metacone.


There is a prominent parastyle anteriorly and a thin, cuspidate cingulum
labially. The protocone is larger than the paracone and connected to the
parastyle by a low preprotocrista. The postprotocrista runs posterolabially to a small but distinct metastyle. A small cingulum is present
posterior to the protocone.
The DP4 has a nearly trapezoidal outline. It has a distinct parastyle
and metastyle on the prominent labial cingulum and a shallow
ectoflexus. The paracone is slightly larger than and more labial than
the metacone. The protocone is about as large as the metacone and
located just lingual to a small paraconule and metaconule. The paraconule
is on the preprotocrista, whereas the metaconule is on the postprotocrista;
clefts in the cristae separate these cuspules from the protocone. The

42

FIGURE 5. Juvenile skull, lower jaws and cheek teeth of E. ditrigonus, NMMNH P-001. A, Dorsal view of skull. B, Ventral view of skull. C, Lateral view of skull. D,
Labial view of left dentary. E, Lingual view of left dentary. F, Occlusal view of left DP2-M1. G, Occlusal view of left dp2-m1.

preprotocrista is confluent with the parastyle, and there is an anterolingual cingulum between the paraconule and the protocone. There is
another cingulum on the postero-lingual edge of the tooth between the
metaconule and the protocone. There is a long protocone slope. Measurements (in mm) of the upper deciduous teeth of NMMNH P-001 are:
DP2 length = 5.8, width = 3.2, DP3 length = 6.7, width = 7.4, DP4 length
= 6.9, width = 9.5.
The dentary of NMMNH P-001 has a horizontal ramus of nearly
equal depth from the symphysis to under the m1. There is a single
mental foramen under the dp3. The ascending ramus is incomplete, but
its preserved base suggests it was inclined slightly posteriorly. The
symphysis is not fused and is an inclined, rugose surface that extends
posteriorly to under the anterior edge of the dp3.
The canine alveolus is nearly round and larger than the combined alveoli of the dp2. There is no postcanine diastema. The dp2 is a
trenchant tooth dominated by a tall protoconid with some vertical striae
(ribs) on its labial surface. A smaller and lower metaconid is slightly
postero-lingual to the protoconid. A low talonid heel is located posterior to the paraconid.
The dp3 is an elongate, submolariform tooth. The trigonid is dominated by a large protoconid and a smaller paraconid and metaconid. The
talonid basin is dominated by a large hypoconid, and there are smaller
entoconid and hypoconulid lingually. A cristid obliqua runs from the

hypoconid towards the metaconid. A low, finely cuspidate cingulid runs


from the anterior edge of the crown labial to the paraconid, all along the
labial edge of the tooth to the posterior face of the hypoconid.
The dp4 is a molariform tooth similar to the dp3 but differing in
being relatively wider and having a much better developed talonid basin
with relatively larger entoconid and hypoconulid. Measurements (in
mm) of the lower deciduous teeth of NMMNH P-001 are: dp2 length =
5.4, width = 2.9, dp3 length = 7.5, trigonid width = 4.5, talonid width =
4.3, dp4 length = 8.2, trigonid width = 5.9, talonid width = 6.3.
Ectoconus symbolus Gazin, 1941
Figs. 2B-D, 3J, Tables 1-2
Ectoconus symbolus Gazin, 1941, p. 39, fig. 21.
Ectoconus symbolus: Robison, 1986, p.106, pl. 2, figs. 8-9.
Ectoconus cavigellii Eberle and Lillegraven, 1998, p. 94, fig. 13D-F.
Ectoconus sp.: Eberle and Lillegraven, 1998, p. 96, fig. 13G.
HolotypeUSNM 16189, left maxillary fragment with partial
P4 and M1-2 (Figs. 2B, 3J; Gazin, 1941, fig. 21b).
Referred specimensFrom the North Horn Formation, Emery
County, Utah:
From w1/2 sec. 7, T. 19 S, R. 6 E, Emery Co., Utah: USNM
16188, left maxillary fragment with P4-M1 (Fig. 2C; Gazin, 1941, fig.

21c); 16190, left dentary with m2-3 (Fig. 2D; Gazin, 1941, fig. 21a);
16213, right dentary fragment with p2; 16214, left dentary with p3-4.
From the west side of Dragon Canyon, Emery Co., Utah: USNM
18517, isolated m1 fragment; 18518, isolated Lm3 talonid.
From Blue Lake, Wagonroad, North Horn Formation, Utah: BYU
3794, right dentary with m2-3 (Robison, 1986, pl. 2, fig. 9); 3788, right
dentary with c, p1, p3-m2, left dentary with p3-m1, Lp2 and Rp2 isolated (Robison, 1986, pl. 2, fig. 8).
From Wagonroad Ridge, Wagonroad, North Horn Formation,
Utah: AMNH 36073, isolated LM3.
From the upper Ferris Formation, western Hanna Basin, Wyoming:
UW locality V-91005: UW 26237, incomplete RM3 (holotype
of E. cavigellii: Eberle and Lillegraven, 1998, fig. 13E); UW 26240,
LP1 (Eberle and Lillegraven, 1998, fig. 13F); UW 26539, RP1; UW
RM1 fragment; UW 26541, right premolar fragment (Eberle and
Lillegraven, 1998, fig. 13D).
From UW locality V-91031: UW 26183, RP2; UW 26272, right
molar fragment (Eberle and Lillegraven, 1998, fig. 13G).
From UW locality 26539: RP1.
Revised diagnosisSmall Ectoconus, differs from the only other
known species, E. ditrigonus, by its smaller size (15-40 %) and lack of
a paraconid on the p4.
DistributionEarly Paleocene (Puercan) of Utah and Wyoming.
DescriptionThe P4 is oval in shape, and the paracone is much
larger than the protocone. Both anterior and posterior cingula are well
developed.
The M1 is almost round in shape. The protocone is the largest
cusp, and the paracone and metacone are slightly smaller and equal in
size. The metastyle is better developed than the parastyle and forms a
small cuspule. The mesostyle is small, but well expressed. The
paraconule and metaconule are almost equally developed and lie close
to the protocone; the preparaconule and postmetaconule wings are well
expressed. The hypocone is slightly larger than the pericone. There are
weak anterior and posterior cingula joined to the pericone and hypocone, respectively. The M2 is slightly larger than the M1, but almost
identical in shape and structure. The M2 has a parastylar lobe that is
larger than the metastylar lobe, whereas the metastylar lobe of the M1
is larger.
The M3 is much reduced. There is a well-developed anterior
cingulum. The paracone is the largest cusp, whereas the metacone is
much reduced and shifted lingually. The paraconule is slightly larger
than the metaconule. There is no postprotocrista and postmetaconule
wing, as the metacone is situated very close to the metaconule and
protocone. The hypocone is very slightly expressed, and the pericone is
absent. Both labial and posterior cingula are well developed. The
parastyle is small but forms a separate cuspule anteriolabial to the
paraconule. There is no metastyle or mesostyle.
The lower canine is small and round in cross section, with rugose enamel. The p2 has a small postcingulid. There is a well-expressed
cuspulid at the base of the crown, posterior to the main cuspid. The p3
is much more differentiated and has a well developed trigonid and talonid. The trigonid is formed by two cuspids: large protoconid and small
metaconid. There is a well-expressed paracristid, but no paraconid anterior to the protoconid. The talonid is formed by a large hypoconulid,
connected to the protoconid by a rather strong cristid. The postcingulid
is rather strong and forms part of the talonid as well. The p4 is much
like p3: the trigonid is formed by two main cuspids, but the talonid has
two small cuspids, a hypoconulid and an entoconid in addition to the
large hypoconid. The postcingulid is well expressed labial to the talonid.
The m1 is the smallest molar. There is a very strong lingual cingulid
that embraces the tooth labially. The trigonid is almost equal in size to
the talonid. The trigonid is formed by three well-developed cuspids:
slightly larger protoconid and equal-sized paraconid and metaconid. The
cristids connecting the major cuspids are weak. The hypoconid is the

43
largest talonid cuspid; the hypoconulid is the smallest, median in position and does not merge either with the hypoconid or the entoconid. The
hypoconid is expanded lingually, so the talonid basin is very small. The
cristid obliqua is weak, but well expressed. There is a weak crest connecting the hypoconid and the hypoconulid.
The m2 is the largest molar, and the trigonid is wider than the
talonid. The trigonid on the m2, as on the m1, is formed by a large
protoconid and smaller paraconid and metaconid. The cristid obliqua
is much weaker than on m1. The hypoconid is shifted lingually and is
situated posterior to and between the protoconid and metaconid. The
hypoconulid lies almost posterior to the hypoconid and is connected to
it by a broad cristid. The entoconid and the hypoconulid are almost
equal in size. The entoconid is connected by a very weak entocristid to
the posterior wall of the metaconid, closing the very small and shallow
talonid basin.
The m3 is much narrower than the m2, but can exceed the latter in
length. The trigonid is more compressed than on the m2. The hypoconid
is the largest talonid cusp, is median in position and is connected to the
protoconid and the metaconid by a very weak cristid obliqua. The large
hypoconulid lies posterolingual to the hypoconid. The small entoconid
is merged with the very strong lingual cingulid. The labial cingulid is very
well developed and embraces the tooth labially, beginning from the base
of the protoconid and reaching the hypoconulid posteriorly.
DISCUSSION
The range of variation of the size and morphology of the teeth of
Ectoconus suggests that only two species, E. ditrigonus and E. symbolus,
are valid, differing mostly in size. The other named species, E.
majusculus and E. cavigellii, are the subjective junior synonyms of E.
ditrigonus and E. symbolus, respectively. Van Valen (1978) stated,
without discussion, that E. majusculus is a subjective junior synonym
of E. ditrigonus, which was later supported by Williamson (1996) and
by Eberle and Lillegraven (1998).
Gazin (1941) described a new species of Ectoconus, E. symbolus,
from the early Paleocene of Emery County, Utah. According to Gazin,
the Utah species differs from E. ditrigonus in its smaller size, relatively larger premolars, weaker parastyle on the upper molars, and weak
or absent parastylid on the lower molars. The dental characters listed
by Gazin are very variable in Ectoconus, but the difference in size well
supports the referral of the Utah specimens to a species separate from
E. ditrigonus.
Eberle and Lillegraven (1998) described a new species of
Ectoconus, E. cavigellii, from the middle Puercan of the Hanna Basin,
Wyoming, based on an isolated, broken M3. The major difference of
the new species from E. ditrigonus was in size. On the other hand, the
new species appeared to be very close in size to E. symbolus and even
some features listed in the diagnosis for E. cavigellii well fit Gazins
diagnosis of E. symbolus, such as the lack of a large parastylar lobe.
We should restrict the discussion of the validity of E. cavigellii to the
holotype, because the other four teeth, referred to the new species, are
not associated with the type. The third molars are the most variable
teeth in all archaic ungulates and especially in large periptychids such
as Ectoconus (Figs. 3I, 4J), and most of the characters listed by Eberle
and Lillegraven (1998), such as weakly rugose enamel; cusps and stylar cingula less inflated; upper molars less transverse are variable in
E. ditrigonus and presumably in E. symbolus. So, based on the same size
and close dental morphology, we treat E. cavigellii as a subjective junior
synonym of E. symbolus.
Ectoconus ditrigonus is found in the middle and late Puercan of
the San Juan Basin, the middle Puercan of Colorado and the late Puercan
of Utah and Wyoming. E. symbolus is known from the middle Puercan of
Wyoming and the late Puercan of Utah. So, both species are found in the
middle and late Puercan and thus cannot be used to identify subdivisions
of Puercan time.

44
ACKNOWLEDGMENTS
R. Tedford made it possible to study specimens in the AMNH, R.
Emry made specimens in the USNM available and L. Martin made it

possible to study the KUVP collection. Funding from the NMMNH


Foundation made this study possible. The study was partially supported by the Russian Foundation for Basic Research, projects nos. 9804-49089 and 99-04-48636.

REFERENCES
Archibald, J.D., 1998, Archaic ungulates Condylarthra; in Scott, K.M., Janis,
C.M. and Jacobs, L.L., eds., Tertiary mammals of North America: Cambridge,
Cambridge University Press, p. 292-331.
Cope, E.D., 1882, Synopsis of the vertebrate fauna of the Puerco epoch: Proceedings of the American Philosophical Society, v. 20, p. 461-471.
Cope, E.D., 1883, Some new Mammalia from the Puerco Formation: American
Naturalist, v. 17, p. 968.
Cope, E.D., 1884, The Condylarthra: American Naturalist, v. 18, p. 790-805.
Cope, E. D., 1885, The Vertebrata of the Tertiary formations of the West. Book 1:
Report of the U. S. Geological Survey of the Territories [Hayden], v. 3, 1009 p.
Cope, E.D., 1888, Synopsis of the vertebrate fauna of the Puerco Series. Transactions of the American Philosophical Society, v. 16, p. 298-361.
Eberle, J. J. and Lillegraven, J. A., 1998, A new important record of earliest Cenozoic mammalian history: Eutheria and Paleogeographic/ biostratigraphic summaries: Rocky Mountain Geology, v. 33, p. 49-117.
Gazin, C.L., 1941, The mammalian fauna of the Paleocene of central Utah, with
notes on geology: Proceedings of the United States National Museum, v. 91, p.
1-53.
Lucas, S.G. and Williamson, T.E., 1995, Systematic position and biochronological
significance of Yuodon and Palasiodon, supposed Paleocene condylarths from
China: Neues Jahrbuch fr Geologie und Palontologie, v. 196, p. 93-107.
Matthew, W. D., 1937, Paleocene faunas of the San Juan Basin, New Mexico: Transactions of the American Philosophical Society, v. 30, p. 1-510.

McKenna, M. C. and Bell, S. K., 1997, Classification of mammals above the species level. New York, Columbia University Press, 631 p.
Middleton, M. D., 1983, Early Paleocene vertebrates of the Denver Basin, Colorado [Ph.D. dissertation]: Boulder, University of Colorado, 403 p.
Osborn, H. F. and Earle, C., 1895, Fossil mammals of the Puerco beds. Collection of
1892: Bulletin of the American Museum of Natural History, v. 7, p. 1-70.
Robison, S. F., 1986, Paleocene (Puercan-Torrejonian) mammalian faunas of the
North Horn Formation, central Utah: Brigham Young University Geology Studies, v. 33, p. 87-133.
Van Valen, L., 1978, The beginning of the age of mammals: Evolutionary Theory, v.
4, p. 45-80.
Williamson, T. E., 1996, The beginning of the age of mammals in the San Juan
Basin, New Mexico: Biostratigraphy and evolution of Paleocene mammals of
the Nacimiento Formation: New Mexico Museum Natural History and Science,
Bulletin 8, 141 p.
Williamson, T. E. and Lucas, S. G., 1992, Stratigraphy and mammalian biostratigraphy of the Paleocene Nacimiento Formation, southern San Juan Basin, New
Mexico: New Mexico Geological Society, Guidebook 43, p. 265-296.
Williamson, T. E. and Lucas, S. G., 1993, Paleocene vertebrate paleontology of the
San Juan Basin, New Mexico: New Mexico Museum of Natural History and
Science, Bulletin 2, p. 105-135.
Zhou, M., Zhang, Y., Wang, B. and Ding, S., 1977, Mammalian fauna from the
Paleocene of Nanxiong Basin, Guandong: Palaeontologia Sinica, v. 153, p. 1100.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

45

A NEW SPECIES OF DELTATHERIUM (MAMMALIA, TILLODONTIA)


FROM THE PALEOCENE OF NEW MEXICO
SPENCER G. LUCAS1 AND PETER E. KONDRASHOV2
2

1
New Mexico Museum of Natural History and Science, 1801 Mountain Road, Albuquerque, NM 87104
Paleontological Institute of RAS, Profsoyuznaya 123, Moscow 117868 RUSSIA and Northwest Missouri State University, Maryville, Missouri, USA 64468

AbstractDeltatherium dandreai is a new species from the Torrejonian Paleocene of the San Juan Basin, New
Mexico. D. dandreai can be distinguished from the type and only other species of Deltatherium, D. fundaminis,
by the new species larger size, sharper molar cusps, larger upper molar conules, larger parastyles, weaker
metastyles, larger hypoconal flares and better developed lower molar entoconids. Deltatherium is best assigned
to a separate family Deltatheriidae within the order Tillodontia.
Key words: Deltatherium, Paleocene, San Juan Basin, New Mexico, Tillodontia

INTRODUCTION
Deltatherium (Fig. 1) is one of the most poorly understood genera of Paleocene primitive ungulates. The phylogenetic relationships
of Deltatherium to other groups of eutherians have been long discussed
without consensus. Deltatherium fossils are found only in the Paleocene (Torrejonian) deposits of the San Juan Basin, New Mexico. They
have long been assigned to a single species, but a careful restudy of the
entire sample of Deltatherium reveals two distinct morphotypes. For
this reason, we name here a new species of Deltatherium.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; KUVP = University of Kansas, Museum of
Natural History, Lawrence; NMMNH = New Mexico Museum of Natural History, Albuquerque; USNM = National Museum of Natural History, Smithsonian Institution, Washington, D. C.
PREVIOUS STUDIES
Cope (1881a) first described Deltatherium and gave a brief description and named the new species, D. fundaminis. He based the new
genus on upper teeth, and soon thereafter (Cope, 1881b) named
Lipodectes penetrans for what he later recognized (Cope, 1884, 1885)
is the lower dentition of D. fundaminis. Cope (1881a, 1884, 1885, 1888)
considered Deltatherium to be a leptictid, a family he included in the
Creodonta. Subsequent authors (Schlosser, 1887; Scott, 1892; Osborn
and Earle, 1895; Matthew, 1897, 1899, 1937) continued to include
Deltatherium in the Creodonta, though its family-level assignment varied from Leptictidae to Oxyclaenidae to Arctocyonidae.
Matthew (1937) provided a detailed description of the skull (Fig.
1) and teeth of D. fundaminis and assigned the genus to Arctocyonidae,
allying it closely to Chriacus. Matthew considered arctocyonids to be
primitive carnivorans (Creodonta), but Arctocyonidae are now assigned
to archaic ungulates (Prothero et al., 1988).
Van Valen (1978, 1988) thought Deltatherium to be the ancestor
of Pantodonta, thus deriving pantodonts from a condylarthran stock.
McKenna (1975) assigned it to Pantodonta without discussion, a point
of view later supported by Sloan (1987). Cifelli (1983) suggested that
a Deltatherium-like mammal could be the ancestor of both pantodonts
and tillodonts. This idea of the close relationships between Deltatherium
and Pantodonta was rejected by some authors (Zhou et al., 1977; Muizon
and Marshall, 1992), but the latter suggested that tillodonts are the
sister group of pantodonts. Lucas (1993) stressed the close resemblance
between Deltatherium and primitive tillodonts (Esthonyx) and identified Deltatherium as the sister taxon of tillodonts. However, he argued
that tillodonts are not true ungulates because they are not condylarthran
descendants. Williamson and Lucas (1993) and Williamson (1996) repeated this opinion, and they included Deltatherium in Tillodontia.

FIGURE 1. Skull and lower jaws of Deltatherium fundaminis, AMNH 16610.


Skull length ~ 140 mm (after Matthew, 1937).

SYSTEMATIC PALEONTOLOGY
Order Tillodontia Marsh, 1875
Family Deltatheriidae Van Valen, 1988
Deltatheriinae Van Valen, 1988, p. 39.
Deltatheriidae: Williamson and Lucas, 1993, p. 128.
Deltatheriinae: McKenna and Bell, 1997, p. 215.
Type genusDeltatherium Cope, 1881a (= Lipodectes Cope,
1881b).

46
Included generaKnown only from the type genus.
Revised diagnosisPrimitive tillodonts with simple upper molars; with a vestigial hypocone and poorly defined conules, but large
parastyles and tall lower molar trigonids. Deltatheriidae differ from
the other tillodonts in having low-crowned and less lophodont lower
teeth and in having widely open talonid basin on the lower molars.
Unlike the other tillodonts, the P4/p4 are very slightly molarized. Also,
unlike other tillodonts, deltatheriids lack an elongate rostrum, have a
narrow skull in the area of the postorbital constriction, a low sagittal
crest, and relatively short basicranium; lack a deep mandibular symphysis, have small P2-3/p2-3, lack molariform P4/p4 and dilambdodont
molars, lack labially placed upper molar protocones, lack distinct
metastylids, lower molars and lack labially extended enamel (crown
hypsodonty).
Distribution Paleocene (Torrejonian) of North America (New
Mexico).
Genus Deltatherium Cope, 1881a
Deltatherium Cope, 1881a, p. 337.
Lipodectes Cope, 1881b, p. 1019.
Deltatherium: Cope, 1884, p. 352.
Deltatherium (in part): Cope, 1885, p. 277.
Deltatherium: Osborn and earle, 1895, p. 39.
Deltatherium: Matthew, 1937, p. 67.
non Deltatherium: Van Valen, 1978, p. 53.
Deltatherium: Williamson and Lucas, 1992, p. 285.
Deltatherium: Williamson and Lucas, 1993, p. 128.
Deltatherium: Williamson, 1996, p. 48.
Deltatherium: McKenna and Bell, 1997, p. 215.
Type species: D. fundaminis Cope, 1881a (= Lipodectes
penetrans Cope, 1881b).
Included species: The type species and D. dandreai sp. nov.
Diagnosis: As for the family.
Stratigraphic and geographic distribution: As for the family.
Discussion: Deltatherium remained monotypic for nearly a century until Van Valen (1978) described the new species D. durini based
on an isolated incomplete M2 (AMNH 102161: Fig. 2E), which supposedly differed from D. fundaminis in being twice as small and in
having a strong metastyle. But, regardless of some superficial similarity, this tooth (Fig. 2E; Van Valen, 1978, pl. 1, fig. 6) is characterized
by features that distinguish it from Deltatherium. For example, it has a
strongly developed metastyle, which is typical of small hyaenodontids.
For that reason, Williamson (1996) suggested D. durini is a senior synonym of Prolimnocyon macfaddeni Rigby, 1980, which is accepted here.
Deltatherium fundaminis Cope, 1881
Figs. 1, 2A-B, D, 3A-B, 4B-C, 5B-C, Tables 1-2
Deltatherium fundaminis Cope, 1881a, p. 337.
Lipodectes penetrans Cope, 1881b, p. 1019.
Deltatherium fundaminis: Cope, 1884, p. 352, fig. 20.
Deltatherium fundaminis: Cope, 1885, p. 278, pl. 23e, figs. 8-11, pl.
25a, fig. 10, pl. 25d, fig. 3.
Deltatherium fundaminis: Cope, 1888, p. 304.
Deltatherium fundaminis: Osborn and Earle, 1895, p. 40, figs. 10-11.
Deltatherium fundaminis: Matthew, 1937, p. 70, figs. 11-14.
Deltatherium fundaminis: Williamson and Lucas, 1992, fig. 15K-M.
Deltatherium fundaminis: Wiliamson, 1996, p. 48.
Deltatherium fundaminis (in part): Williamson and Lucas, 1997, p. 263.
Holotype AMNH-3315, palate with left P4-M3 and right P2M2 (Fig. 2A-B; Cope, 1884, pl. 25a, fig. 10; Cope, 1885, pl. 23e, figs.
8-11; Williamson and Lucas, 1992, fig. 15K).
Horizon and locality of holotype Torrejonian interval of the

Nacimiento Formation, San Juan Basin, New Mexico, mouth of Caon


Largo.
Revised diagnosisA species of Deltatherium that differs from
D. dandreai in having less sharp upper molar cusps, smaller upper
molar conules, smaller parastyles, stronger metastyles, smaller
hypoconal flares and a prominent hypoconulid and no entoconid on the
lower molars.
Referred specimensFrom the Torrejonian interval of the
Nacimiento Formation, San Juan Basin, New Mexico, specific locality
unknown: AMNH 3130, left dentary fragment with part of m2-3 (referred to Pentacodon? by Matthew, 1937, p. 226); 3316, skull and lower
jaws with left P3-M1, right I1-3, C and M2-3, doubtfully associated
ulna (Figs. 3B, 4B, 5B; Cope, 1884, pl. 23e); 3318, right dentary with
badly damaged p2-m2; 3326, right dentary with p3, m2-3 (collected by
Baldwin 8-28-?); 3327, left dentary fragment with p4; 3329, right dentary
with m1-2 (bag 2, collected by Baldwin 3-25-1882); 3333, left P4-M2
and left m1-2; 3335, jaw with left p3, m1-2 (Figs. 3A, 4C, 5C; holotype
of Lipodectes penetrans: Cope, 1885, pl. 25d, fig 3; Williamson and
Lucas, 1992, fig. 15L-M); 3338, very concreted upper and lower molars, right M2 and m1 visible; 3343, left M1?; 3349, badly damaged
left maxillary fragment with M1-3; 3350, skull, jaw and some postcrania
(collected by Baldwin 11-26-1883); 3382, left dentary fragment with
m3 (collected by Baldwin, 1882); 4054, right M1 (collected by Baldwin,
1883); 16610, skull, lower jaws and humerus (Fig. 1; Matthew, 1937,
figs. 11-14, pls. 9-10).
From the head of Coal Creek Canyon, collected by Wortman
and Peterson, 1892: AMNH 763, right maxilla with P3-M3 (Fig. 2D).
From Escavada Canyon, 1892: AMNH 783, left dentary with
p3-4, m3 and some associated postcrania; 780, lower jaw with left p4,
right p3, m1-2 (Osborn and Earle, 1895, fig. 10); USNM 5871 (originally AMNH 781), left dentary fragment with p2-4, m3.
From sec. 26, T27N, R11W, Coots Caon, 1916: AMNH 17084,
left dentary fragment with m2; 17086, right dentary fragment with p4m1.
From 1 mi west of Angel Peak, Coots Canyon, 1916: AMNH
17090, dentary fragments with left m1-2, right m3.
From the east fork of Arroyo Torreon, 1913, lower level of
Torrejon: AMNH 16742, right M1, M3.
From Kimbeto Arroyo (1913 coll.): AMNH 93141, left m3 talonid.
From Kimbeto Arroyo: USNM 407548, left dentary fragment
with m2-3.
From University of Kansas locality 7, Kimbeto area, exposures
along highway one and a half mile SE of J & R Trading Post on both
sides of the road (SE 1/4 sec. 1, T23N, R9W): KUVP isolated left m1,
m2, m3, right M1.
From University of Kansas locality 8, Kimbeto area, W1/2 sec.
1, T23N, R9W: KUVP 14044, left dentary with p3-4, right dentary
with c, p3, m2.
From University of Kansas locality 9, Kimbeto area, NW 1/4
sec. 2, T23N, R9W: KUVP 10623, right dentary with m2, concreted;
11945, left M1; 7924, right dentary with p4-m1, concreted; 11944, left
M2; 13807, right maxillary fragment with P3-M2.
From west side, west fork of Kimbetoh: USNM 15341, left m2
and right m3.
From head of east branch Kimeto Arroyo: USNM 407549, left
maxillary fragment with P4-M3, left dentary fragment with p3-p4, m3,
right p2, p4, m2-3.
From University of Kansas locality 13, Angel Peak area, NW 1/
4 sec. 14, T27N, R11W), 50 m below western rim of Angel Peak badlands in first main red zone below rim: KUVP 9597, left maxillary
fragment with M1-3; 10593, right m1; 7795, right dentary with p2-3,
m1-2; 7796, left maxillary fragment with P4-M3; 9596, right maxillary
fragment with M1-2, M3 (partial).
From NMMNH locality 625: NMMNH P-21390, left dentary

47

FIGURE 2. Selected upper cheek teeth of Deltatherium in occlusal view. A-B, AMNH 3315, holotype of D. fundaminis, right P4-M2 (A) and left P3-M3 (B: photograph
reversed). C, AMNH 16611, holotype of D. dandreai, right P4-M3. D, AMNH 763, D. fundaminis, right P3-M3. E, AMNH 10261, right M2, holotype of Deltatherium
durini.

48

FIGURE 3. Selected lower cheek teeth of Deltatherium in occlusal view. A, AMNH


3335, left p3 roots, p4, m2-3 of D. fundaminis. B, AMNH 3316, right p4-m3 of D.
fundaminis. C, AMNH 3966, right p4-m3 of D. dandreai.

fragment with p3, m1-3.


From NMMNH locality 1258: NMMNH P-15974, left M3.
From NMMNH locality 1260: NMMNH P-15987, right m1.
From NMMNH locality 2548: NMMNH P-25035, right p3-4.
From NMMNH locality 2569: NMMNH P-19431, palate with
left P2-M3 and right P4-M3.
From NMMNH locality 2590: NMMNH P-18817, right dentary
fragment with m1-3; P-18907, right dentary fragment with p2-3; P19772, right maxilla fragment with P4.
From NMMNH locality 2630: NMMNH P-21548, left maxilla
with incomplete M1-3.
From NMMNH locality 2654: NMMNH P-20973, left maxilla
with M1.
From NMMNH locality 2722: NMMNH P-21733, left dentary
fragment with m1-3.
From NMMNH locality 3249: NMMNH P-2993, left dentary
fragment with m3; P-12268, left dentary fragment with m3 talonid.
DescriptionThe skull and dental morphology of Deltatherium
were described by Matthew (1937) in his monograph on the Paleocene
mammals of San Juan Basin. But, since that time numerous specimens
have been collected, which necessitate a redescription.
The skull (Fig. 1) is elongate with a massive snout and narrow,

laterally compressed braincase. The sagittal crest is very strong, and the
zygomatic arches are very broad. The anterior edge of the orbit is under
the posterior root of P4. The palate is relatively broad, and the tooth
rows converge anteriorly. There are two or three pairs of palatine foramina, and the largest pair is at the level of the posterior edge of
M1. The choanal notch is rectangular in shape, with a small median
process, and the choanal edge is a little thickened. The foramen
infraorbitale is situated above the anterior root of M3. The infraorbital
canal is relatively short. The long axis of the glenoid fossa is perpendicular to the long axis of the skull. The postglenoid process is short,
round and relatively high. The postgleniod foramen is quite large, opens
at the posteromedial edge of the postglenoid process and faces ventrally. The occiput is narrow and oval in shape. The occipital crest is
very strong. The occipital condyles are relatively small and embrace
the foramen magnum for about half of its height. The nasals are narrow
and long; posteriorly, they are close to the crista frontalis externa. The
jugal bifurcates anteriorly; its squamosal process does not reach the
glenoid fossa. The frontals are large and form most of the orbit. The
postorbital process is small, but massive; it forms a tubercle on the
upper edge of the orbit. The crista frontalis externa begins at the postorbital process, extends towards the midline of the skull, then turns
posteriorly and forms a strong sagittal crest. The basioccipital bears
the median ventral crest and two weaker crests situated lateral to the
central crest. The foramen sphenopalatinum is situated above the M3.
The promontorium of the petrosal projects ventrally.
There are three small, subequal-sized incisors and a large trenchant canine. The P2 is a small, simple tooth. The P3 is wider than the
P2. The protocone is small and situated lingually. There are two crests
on the paracone: one on the anterior and one on the posterior slopes of
the tooth. There is a weak labial cingulum. The P4 has a well-developed protocone, which is slightly smaller than the paracone. The
parastyle is very well developed. The metacone is absent. The cingulum is circular but weak.
The M1 is triangular in shape; the protocone is slightly larger
than the paracone and metacone, which are equal in size. The paraconule
and metaconule are absent. The cingulum is circular and better expressed than on the P4. The hypocone is small and circular; it lies
posterior to and a little labial to the protocone. The parastyle is less
developed than on the P4 and the M2. There is a notch on the labial
side of the tooth between the paracone and the metacone, which is well
expressed on all molars. This gives the tooth a heart-shape, which
resembles very much the shape of the molars in Esthonyx. The M2 is a
little larger than the M1; the parastyle is slightly larger and there is
vestigial paraconule and metaconule, which have the shape of weak
crests. The hypocone is more lingual, than on the M1, but never lies
directly lingual to the protocone, as in Deltatherium dandreai.
The M3 has a very remarkable shape. The protocone and paracone are well developed and equal in size, whereas the metacone is
reduced and much smaller. The parastyle is much stronger than on the
other molars and is larger than the metacone. Because of this, the tooth
acquires the shape of a triangle with the metacone at the base of the
right angle. The paraconule and metaconule are absent, and the hypocone is weak and situated posterolingual to the protocone.
The lower jaw (Figs. 1, 4-5) is elongate and slender with a tall
coronoid process. The minimum depth of the horizontal ramus is near
the anterior edge of the coronoid process; the maximum depth is near
the p2-3. The masseteric fossa is deep and well defined anteriorly, posteriorly and ventrally. The coronoid process is tall, broad and inclined
posteriorly. The incisura mandibulae is not very deep, and it is round in
shape. The articular process is shifted caudally. The condyle lacks the
neck; its long axis is perpendicular to the long axis of the skull. The
angular process is well developed. There is a notch between the articular and angular processes. The processus angularis is directed ventromedially. The symphyseal region is massive due to the strong canines.
The foramen mandibularis is under the posterior edge of the coronoid

49

FIGURE 4. Dentary fragments of Deltatherium in lingual view. A, AMNH 3966, D. dandreai, right dentary with p2-m3. B, AMNH 3316, right dentary fragment with m13 of D. fundaminis. C, AMNH 3335, left dentary fragment with c, p4 and m3 of D. fundaminis.

process, below the tooth row. There are two mental foramina on each
side of the jaw. Both are situated directly under or a little posterior to
the canine.
There are three small, subequal-sized lower incisors and a large
trenchant canine. The p2 is simple with a vestigial talonid. The p3 is
simple, too, but the talonid is a little larger and has one small cuspid.
The cingulum is circular and is better expressed than on m1. The p4 is
slightly molarized, and the trigonid has three major cuspids: protoconid,
paraconid and metaconid. The protoconid is larger than the other cuspids, and the paraconid is the smallest. The talonid length is not less
than one-half of the length of the trigonid. There is only one small
cuspulid on the center of the length of the talonid. There is a longitudinal cristid on the talonid, which connects the posterior wall of the trigonid with the cuspid on the talonid. The trigonid is much higher than the
talonid.
The m1 and m2 are very much alike. The trigonid is equal in length
to the talonid, but much higher. The paraconid is the smallest of the
cuspids. The protoconid and metaconid are equal in size. The cuspids are
sharp, and the cristid obliqua begins from the middle of the posterior wall
of the trigonid. The hypoconid is the largest talonid cuspid. The entoconid occupies a central position, and the hypoconulid is smaller, but
well expressed. The talonid basin is widely open lingually. The m3 is
much more elongate than m1-2, but more narrow and lower in height.
There is a cristid on the trigonid uniting the paraconid and the metaconid

and closing the trigonid basin lingually. The cingulid is weak, but almost
circular and absent only on the lingual side of the tooth. The entoconid is
large, almost equal in size to the hypoconid. There is a small entoconulid
situated anterolingual to the entoconid. The entocristid is absent, and the
talonid basin is widely open lingually.
Deltatherium dandreai, new species
Figs. 2C, 3C, 4A, 5A, Tables 1-2
Deltatherium fundaminis (in part): Matthew, 1937, p. 70.
Deltatherium fundaminis (in part): Williamson and Lucas, 1997, p. 263,
fig. 2B.
HolotypeAMNH 16611, very concreted skull and jaw fragments with right P4-M3 visible (Fig. 2C).
Horizon and locality of holotypeFrom the Torrejonian interval of the Nacimiento Formation, San Juan Basin, New Mexico, specific
locality unknown.
ParatypeNMMNH P-30675, left dentary, with part p1, p3-4,
m2-3.
Horizon and locality of paratypeFrom the Torrejonian interval of the Nacimiento Formation, San Juan Basin, New Mexico,
NMMNH locality 4343.
DiagnosisA species of Deltatherium that differs from D.

50
TABLE 1. Measurements (in mm) of selected upper cheek teeth of Deltatherium. Asterisks (*) indicate type specimens.
CL

CW

P2 L

P2W

P3 L

P3 W

P4 L

P4 W

M1 L

M1W

M2 L

2.71

3.85

5.37

5.29

5.99

6.15

7.28

8.78

7.49

NMMNH 19431
AMNH 3333

5.90

6.71

6.22

7.81

6.75
7.40

8.48
9.11

7.61
7.30

AMNH 3316
AMNH 3350

6.11

6.10

6.40

8.11

7.20
6.20

8.40
7.41

6.60

6.50
5.60

6.50

7.44
7.55

6.60
6.61

9.10
7.40

6.60
7.20

7.80
7.50

M2W

M3 L

M3W

6.50

10.52

9.88
10.40

7.55

9.34

7.11
6.55

9.50
10.20

6.91

9.66

7.30
8.21

9.88
10.50

7.12

7.55

6.60

8.70

7.05

10.5

D. fundaminis:
USNM 407549
NMMNH 15974

AMNH 763
AMNH 3315*

3.61

3.22

AMNH 4054
AMNH 16742
KU 13807
KU 11945

5.03

4.96

KU 9597
KU 7796

6.94

6.91

6.76
6.2

8.37
8.2

6.64

8.03

7.53

9.24
9.95

8.23

10.9
11.4

7.67
8.08

10.6
11.8

KU 9596
Mean

3.16

3.54

5.80

5.86

6.45

7.43

7.63
6.92

8.19
8.42

7.6
7.45

9.99
10.32

6.76
7.15

10.3
9.80

Std. Dev.
N

0.64
2

0.45
2

0.62
5

0.692
6

0.33
6

0.694
7

0.48
13

0.76
14

0.51
10

0.56
10

0.56
8

1.29
8

4.50

5.80

6.00

6.30

6.88
6.50

8.61
7.80

8.16
7.40

10.13
10.10

5.54
7.10

8.94
8.00

5.73

6.13

6.51
6.81

6.19
7.60

8.75
7.91

6.93

8.56

5.82

8.06

7.82

6.33

10.01

7.05

8.13

7.21

10.08
6.85

10.54

6.40
6.20

7.80
6.90

D. dandreai:
USNM 15340
USNM 15343
USNM 15355
NMMNH 18874

8.81

6.52

5.20

5.41

NMMNH 19749
NMMNH 19756
NMMNH 20718
NMMNH 25027
AMNH 3319
AMNH 3340

4.90

AMNH 3342
AMNH 16611*

5.3

AMNH 16612
KU 7793

7.30

7.10
7.10

9.60
8.71

6.41

9.60

5.80
5.5

8.90
8.9

6.3

6.9

7.6

7.4

9.4

7.50
9.1

6.80

7.80
9.61

4.1

3.23

5.61

5.66

6.12

7.61

7.20
7.23

8.10

Mean
Std. Dev.

8.81

6.52

5.20
0.78

5.41
1.54

5.12
0.68

5.97
0.241

5.90
0.72

6.93
0.723

6.81
0.47

8.11
0.88

7.27
0.45

9.19
0.81

6.17
0.6

9.01
0.9

11

11

fundaminis in having sharper upper molar cusps, larger upper molar


conules, larger parastyles, weaker metastyles, larger hypoconal flares
and a prominent entoconid and a weaker hypoconulid on the lower
molars.
Referred specimens: From the Torrejonian interval of the
Nacimiento Formation, San Juan Basin, New Mexico, specific locality
unknown: AMNH 3319, right maxilla with P4-M2; 3340, concreted
upper and lower teeth with right P3-M3 and right p4-m3 visible; 3342,
left M3 (collected by Baldwin, 1885); 3966, right dentary with p3-m3
(Figs. 3C, 4A, 5A); 4034a, left p4; 16612, left maxilla with P4-M3 and
left m2 92 mi above Chico Spring, 1913); USNM 5870 (formerly AMNH
3337), left dentary fragment with m1-2 (collected by Baldwin, 1882).
From Gallego Caon, collected by Baldwin, 1883: AMNH 3332,
dentary fragments with right m1-3 and left m2-3.
From 2 mi SE of Aztec, Animas Valley, 1916: AMNH 17037,
left dentary with p4-m3.
From divide between Kimbeto Arroyo and west Kimbeto Arroyo: USNM 15343, right C, M1, M3; 15355, skull with left C, P2-M3,
right P1-M3, lower jaw with left c, p2-m3 and right c, p2-4.
From east side of Kimbeto Arroyo: USNM 15340, right maxillary
fragment with M1-2, left M3.
From University of Kansas locality 9, Kimbeto area, NW 1/4 sec.

2, T23N, R9W: KUVP 7923, left dentary with m1-2.


University of Kansas locality 13, Angel Peak area, NW 1/4 sec.
14, T27N, R11W, 50 m below western rim of Angel Peak badlands in
first main red zone below rim: KUVP 7793, left maxillary fragment
with C, P2-M2.
From NMMNH locality 400: NMMNH P-21042, left dentary
fragment with m1-2.
From NMMNH locality 624: NMMNH P-15327, right dentary
fragments with m1-3.
From NMMNH locality 625: NMMNH P-21391, dentary fragment with incomplete m1; P-21392, left dentary fragment with p4-m1;
P-21393, left dentary fragment with p4-m1; P-25011, dentary fragments
with left p4 and right m1; P-25015, left dentary fragment with m2.
From NMMNH locality1264: NMMNH P-16110, right dentary
fragment with part p3, p4, left dentary fragment with p4-m1.
From NMMNH locality 1397: NMMNH P-16147, right dentary
fragment with p3.
From NMMNH locality 2543: NMMNH P-19264, left dentary
with m2-3 (Williamson and Lucas, 1997, fig. 2B); P-19265, left dentary
fragment with part m2-3, right dentary fragment with part m1-2.
From NMMNH locality 2548: NMMNH P-25027, right maxilla
fragment with M3.

51
TABLE 2. Measurements (in mm) of selected lower cheek teeth of Deltatherium. Asterisks (*) indicate type specimens.
Lower
D. fundaminis:
USNM5871

p2L

p2 W

p3 L

p3 W

p4 L

p4 W

3.2

5.7

3.7

6.3

4.1

m1 L

m1tri

m1tal

USNM15341
USNM407548
USNM407549
NMMNH21733

4.9

2.9

6.4

6.2

5.7

2.8

5.9

m2tri

m2tal

8.1
7.5

3.4
3.4

4.2
3.8

6.8

4.2
3.4

4.2
4

3.7

3.8

3.7

NMMNH18817
NMMNH15987
NMMNH18907
NMMNH2993

m2 L

3.1

3.5

3.1

3.8

m3 L

m3 tri

m3tal

7.9

4.5

3.8

8.2

4.5

3.3

8.4

4.6
3.9

3.7
4.2

7.7

4.1

3.8

7.5

3.5
3.5

8.1

4.2
4.2

2.8

NMMNH12268
NMMNH 21390

6.5

AMNH3327
AMNH3382

5.6

AMNH 3335
AMNH780

5.9
5.6

3.3
3.5

AMNH783
AMNH3326

6.1
6.2

3.5
4.1

3.5

3.5

6.1

3.7

6.3

3.6

3.9
3.6

8.1
7.3
7.6

AMNH93141
AMNH17084

7.6

3.9

4.5

8.5
8.7

3.7
4.3

8.7

4.9

5.3
4.5

6.8

AMNH 3329
AMNH 3316

6.6

KU 10623
KU 7924

6.2
4.1

3.7
7.6

7.1
5.1

4.4

4.3
3.4

KU 14044
KU 10593

3.5

6.3

3.9

7.4
4.3

7.1

KU 7795
Mean

4.9
5.125

2.9
2.95

5.3
5.81

2.6
3.38

6.17

3.7

6.52

3.5

4.4167

7.11

4.16

4.0833

8.13

4.32

3.77

Std. Dev.
N

0.38622
4

0.17
4

0.31
7

0.48
8

0.29
6

0.24
7

0.37
5

0.42
5

1.4878
6

1.04
11

0.73
9

0.2714
6

0.44
10

0.32
9

0.3093
10

4.5
5.6

6.9

2.9

7.7

3.7

4.3

8.9

3.9

4.1
3.6

8.4

4.8

4.4

3.9

4.3

8.5

3.9

4.1

D. dandreai:
USNM15355
NMMNH16147

5.3

NMMNH15327
NMMNH16110

6.3

3.8

7.5
7.5

3.5
3.4

4.1
3.9

NMMNH19893
NMMNH21042

5.8

3.5

7.3

3.5

4.1

3.5

7.5

3.7

7.9

NMMNH21393
NMMNH19730
NMMNH19265
NMMNH25011

6.9
7.8

NMMNH25015
NMMNH30675

6.2
5.9

3.5

NMMNH19264
AMNH4034

3.5

4.9

3.8

4
4.4

7.9
7.9

3.4

7.2
6.9

AMNH3332
AMNH 3966
AMNH17037
Mean

5.3

5.6

Std. Dev.
N

0.8
4

3.2

4.9

9.6

4.4

9.8

4.4

3.4
3.9

6.3

4.4

3.6

15.1

7.1
6.4

4.1
3.79

7.9
7.4

23
4

0.5
6

0.35
7

0.3
8

From NMMNH locality 2577: NMMNH P-20718, right maxillary fragment with M1.
From NMMNH locality2589: NMMNH P-19893, right dentary
fragment with part of m1 and m2.
From NMMNH locality 2590: NMMNH P-18874, right maxilla
fragment with P-M1; P-19749, left maxilla fragment with M3; P-19756,
left M1; P-19730, right dentary fragment with part m2 and m3.

3.7

7.1
7.1

3.5

3.5
3.883

6.5
7.4

0.3
7

0.223
6

0.5
9

4.5
4.6

8.7

4.5

3.8

4.2
4.563

8.1
8.9

4.6

4.1
4.217

0.5
5

0.32
8

0.6
7

0.5
4

0.214
6

From NMMNH locality 3249: NMMNH P-2852, left dentary


fragment with m3; P-2917, right dentary fragment with incomplete m2;
P-12251, left dentary fragment with part p4-m1.
From NMMNH locality 4343: NMMNH P-30627, right dentary
fragment with inomplete m3.
EtymologyThe species is named after N. V. (Dan) DAndrea
to honor his many contributions to paleontology in New Mexico.

52

FIGURE 5. Dentary fragments of Deltatherium in labial view. A, AMNH 3966, right dentary with p2-m3 of D. dandreai. B, AMNH 3316, right dentary fragment with
m1-3 of D. fundaminis. C, AMNH 3335, left dentary fragment with c, p4 and m3 of D. fundaminis.

Deltatherium sp.
Referred specimensNMMNH P-2672, canine (NMMNH locality 1870); P-2673, canine (locality 1870); P-2828 , edentulous left
dentary fragment (locality 3249); P-2918, dentary and molar fragments
(locality 3249); P-12154, left dentary fragment with part p2 (locality
3249); P-15909, dentary and maxilla fragments (locality 1230); P-20637,
tooth fragments (locality 2588); P-20961, concreted maxilla and dentary
fragments (locality 2649); P-20969, concreted dentary fragments (locality 2652); P-21041, edentulous dentary fragment (locality 400); P-21066,
upper molar fragment (locality 2662); P-21379, dentary and tooth fragments (locality 2629); P-21572, canine (locality 400); P-27807, edentulous maxillary fragments (locality 2569); P-30766, dentary and tooth
fragments (locality 2662).
DiscussionThe specimens listed here clearly pertain to
Deltatherium but are too fragmentary and/or poorly preserved to be
assigned to a species.
PHYLOGENETIC RELATIONSHIPS
The skull and dental morphology of Deltatherium are much more
similar to those of a primitive tillodont than to any known condylarth.

Thorough comparison of Deltatherium with the primitive tillodont


Esthonyx reveals several specific synapomorphic features of these two
taxa: (1) well developed ectoflexus on the upper molars; (2) strong
parastyle on the upper molars; (3) transverse M3; (4) shape of M3
the protocone and paracone are well developed and of equal size,
whereas the metacone is reduced and much smaller; the parastyle is
much stronger than on the other molars and is larger than the metacone; because of this the tooth acquires the shape of a triangle with the
metacone at the base of the right angle; (5) stylar shelves well developed; and (6) entoconid much reduced (secondary), central in position
and talonid basin widely open (secondary). These features are not found
in arctocyonids or in other true condylarths. They are strong evidence that Deltatherium and early tillodonts are closely related.
ACKNOWLEDGMENTS
R. Tedford made it possible to study specimens in the AMNH.
The generosity of Dan DAndrea made this study possible. The work of
PK was supported by the Russian Foundation for basic research project
02-04-48458 and the Grant Council for the President of Russian Foundation project Nsh-1840.2003.4.

53
REFERENCES
Cifelli, R.L., 1983, The origin and affinities of the South American Condylarthra
and early Tertiary Litopterna (Mammalia): American Museum Novitates, no.
2772, p. 1-49.
Cope, E.D., 1881a, Mammalia of the lower Eocene beds: American Naturalist, v.
15, p. 337-338.
Cope, E. D., 1881b, Notes on Creodonta: American Naturalist, v. 15, p. 1018-1020.
Cope, E.D., 1884, The Creodonta: American Naturalist, v. 18, p. 255-267, 344353, 478-485.
Cope, E. D., 1885, The Vertebrata of the Tertiary formations of the West. Book 1:
Report of the U. S. Geological Survey of the Territories [Hayden], v. 3, 1009 p.
Cope, E. D., 1888, Synopsis of the vertebrate fauna of the Puerco Series: Transactions of the American Philosophical Society, new series, v. 16, p. 298-361.
Lucas, S.G., 1993, Pantodonts, tillodonts, uintatheres and pyrotheres are not ungulates; in Szalay, F.S., Novacek, M.J. and McKenna, M.C., eds., Mammal phylogeny: Placentals: New York, Springer-Verlag, p. 182-194.
Matthew, W. D., 1897, A revision of the Puerco fauna: Bulletin of the American
Museum of Natural History, v. 9, p. 259-323.
Matthew, W. D., 1899, A provisional classification of the fresh-water Tertiary of the
West: Bulletin of the American Museum of Natural History, v. 12, p. 19-75.
Matthew, W.D., 1937, Paleocene faunas of the San Juan Basin, New Mexico: Transactions of the American Philosophical Society, new series, v. 30, p. 1-510.
McKenna, M.C., 1975, Toward a phylogenetic classification of the Mammalia; in
Luckett, W.P. and Szalay, F.S., eds., Phylogeny of the primates: New York,
Plenium Press, p. 21-46.
McKenna, M. C. and Bell, S. K., 1997, Classification of mammals above the species level. New York, Columbia University Press, 631 p.
Muizon, C. de and Marshall, L.G., 1992, Alcidedorbignia inopinata (Mammalia,
Pantodonta) from the early Paleocene of Bolivia: Phylogenetic and
paleobiogeographic implications: Journal of Paleontology, v. 66, p. 499-520.
Osborn, H. F. and Earle, C., 1895, Fossil mammals of the Puerco beds. Collection of
1892: Bulletin of the American Museum of Natural History, v. 7, p. 1-70.
Prothero, D., Manning E. and Fischer M., 1988, The phylogeny of ungulates; in

Benton, M.J., ed., The phylogeny and classification of tetrapods. Vol. 2: Mammals: Oxford, Clarendon Press, p. 201-234.
Rigby, J. K., Jr., 1980, Swain quarry of the Fort Union Formation, middle Paleocene
(Torrejonian), Carbon County, Wyoming: Geologic setting and mammalian
fauna: Evolutionary Monographs, v. 3, 179 p.
Schlosser, M., 1887, Die Affen, Lemuren, Chiropteren u.s.w. des europ. Tertiars:
Beitrage Palontologie Oest.-Ungarn, v. 6.
Scott, W. B., 1892, A revision of the North American Creodonta: Proceedings of the
Academy of Natural Sciences of Philadelphia, v. 1892, p. 291-323.
Sloan, R.E., 1987, Paleocene and latest Cretaceous mammal ages, biozones,
magnetozones, rates of sedimentation, and evolution: Geological Society of
America, Special Paper 209, p 165-200.
Van Valen, L., 1978, The beginning of the age of mammals: Evolutionary Theory, v.
4, p. 45-80.
Van Valen, L., 1988, Paleocene dinosaurs or Cretaceous ungulates in South America:
Evolutionary Monographs 10, p. 1-79.
Williamson, T.E., 1996, The beginning of the age of mammals in the San Juan
Basin, New Mexico: Biostratigraphy and evolution of Paleocene mammals of
the Nacimiento Formation: New Mexico Museum Natural History and Science,
Bulletin 8, 141 p.
Williamson, T. E. and Lucas, S. G., 1992, Stratigraphy and mammalian biostratigraphy of the Paleocene Nacimiento Formation, southern San Juan Basin, New
Mexico: New Mexico Geological Society, Guidebook 43, p. 265-296.
Williamson, T. E. and Lucas, S. G., 1993, Paleocene vertebrate paleontology of the
San Juan Basin, New Mexico: New Mexico Museum of Natural History and
Science, Bulletin 2, p. 105-135.
Williamson ,T.E. and Lucas, S.G., 1997, The Chico Springs locality, Nacimiento
Formation, San Juan Basin, New Mexico: New Mexico Geological Society,
Guidebook 48, p. 259-265.
Zhou, Chow M., Zhang, Y., Wang, B. and Ting, S., 1977, Mammalian fauna from
the Paleocene of Nanxiong Basin, Guandong: Paleontologia Sinica Series C., v.
20, p. 1-100.

54

Ectoconus majusculus, type skull and paratype mandible (from Matthew, 1937)

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

55

MAGNETIC STRATIGRAPHY OF THE MIDDLE EOCENE (DUCHESNEAN)


BACA FORMATION, WEST-CENTRAL NEW MEXICO
DONALD R. PROTHERO1, JOSHUA A. LUDTKE1 AND SPENCER G. LUCAS2
1
Department of Geology, Occidental College, Los Angeles, CA 90041;
New Mexico Museum of Natural History, 1801 Mountain Road N.W., Albuquerque, NM 87104

AbstractThe Baca Formation in west-central New Mexico consists of about 180 m of siliciclastic red beds
deposited by fluvial and lacustrine systems. It yields a fragmentary but important Duchesnean (late middle
Eocene) mammalian fauna, including the primitive entelodont Brachyhyops wyomingensis, the creodont
Hyaenodon sp. (both of which first occur in the Duchesnean), plus the primitive artiodactyls Protoreodon
pumilus and the brontothere Diplacodon sp. (both Uintan holdovers), and fragmentary agriochoerid oreodonts,
camelids, and protoceratids. The section at Mariano Mesa (north of the town of Quemado in Catron County)
yields Duchesnean age mammals and was sampled for magnetostratigraphy. The samples were then subjected
to alternating field demagnetization at 25, 50, and 100 Gauss, followed by thermal demagnetization at 50C
steps from 200 to 630C. These samples yielded a stable single component of remanence that passed a reversal
test, and was held largely in magnetite with minor goethite overprints. The entire sampled section is of reversed
polarity. Based on radioisotopic dates of ~ 38.0 Ma from the overlying Spears Formation volcanics (lower part
of the Datil Group), we correlate the Duchesnean interval of the Baca Formation with Chron C17r (38.2-38.3
Ma), which is middle Duchesnean. This correlates to the Duchesnean interval of the Galisteo Formation in
north-central New Mexico, suggesting a synchronous transition from Laramide tectonism to intermediate volcanism in north-central and west-central New Mexico that began during the middle Eocene.
Keywords: Eocene, New Mexico, Baca Formation, Duchesnean.

INTRODUCTION
The Baca Formation of Wilpolt and Wanek (1946) crops out in a
discontinuous belt for about 250 km in west-central (Catron and Socorro
Counties) New Mexico, from near Magdalena, New Mexico, westward
across the border into Springerville, Arizona (Fig. 1). Most outcrops
average about 180 m in thickness, with a maximum of 370 m (Snyder,
1971; Cather, 1980, 1982). The Baca Formation consists mostly of red
mudstones, sandstones, and lesser amounts of claystone and conglomerate. This unit is thought to represent a braided fluvial-alluvial fan
system to the west, and a lacustrine system to the east (e.g., Cather,
1982). The Baca Formation disconformably overlies the Upper Cretaceous Mesaverde Group (mostly the Crevasse Canyon Formation). It is
overlain by the Spears Formation (lower Datil Group) volcanics; the
contact between these units is conformable and gradational (Cather,
1982).
Fossil mammals have been found in the braided-alluvial plain
facies of the Baca Formation. Snyder (1970, 1971) first reported specimens of the artiodactyl Protoreodon pumilus, which suggested a Uintan
age for the Baca Formation. Schrodt (1980) and Schiebout and Schrodt
(1981) discovered and reported additional mammal fossil including
specimens they assigned to brontotheres (which they called Menodus
sp.), the creodont Hyaenodon sp., the entelodont Brachyhyops
wyomingensis, the oromercyid Eotylopus sp., and the ruminant
Leptomeryx sp. They assigned these mammals a Chadronian age, based
largely on comparison with the Porvenir local fauna of Trans-Pecos
Texas, which they considered Chadronian, in agreement with Wilson
(1978).
However, Lucas (1983) re-examined these fossils, and presented
different identifications. He referred the brontothere to Diplacodon sp.,
the Leptomeryx to a protoceratid such as Leptoreodon, and the
Eotylopus to an indeterminate camelid. He also reported the presence of a fragmentary agriochoerid as well. This revised interpretation
of these fossils no longer supports a Chadronian age for the Baca Formation, but instead is consistent with a middle or late Duchesnean age.

FIGURE 1. Location map of the Baca Formation (after Lucas, 1983).

Diplacodon, Leptoreodon, and Protoreodon are Uintan holdovers that


are also known from the Duchesnean (but not the Chadronian), while
Hyaenodon and Brachyhyops first appear in the Duchesnean and continue into the early Chadronian. Thus, the Duchesnean age of the fauna
(as currently known) is well established (Lucas, 1992; Lucas and
Williamson, 1993). In addition, there was some confusion because earlier authors considered the Uintan to be late Eocene, the Duchesnean
late Eocene or early Oligocene, and the Chadronian early Oligocene.
Recent radioisotopic dating and paleomagnetic analysis has shown that
the Uintan is late middle Eocene, the Duchesnean latest middle Eocene,
and the Chadronian is late Eocene in age (Prothero and Swisher, 1992;
Prothero and Emry, 1996). Lucas (1990) reported a brontothere metacarpal from stratigraphically low in the Baca Formation, which suggests
that the base of the formation is Bridgerian or Uintan in age.
Further confirmation of this interpretation comes from radioisotopic dates of the overlying Spears Formation (lowermost unit of the
Datil Group). A spectrum of K-Ar and fission-track ages from these
volcanics yield ages that cluster around 38 Ma for the base of the Datil
volcanics (Cather et al., 1987), with a 40Ar/39Ar date as young as 37.02
0.15 Ma, and one K/Ar date as old as 41.8 3.0 Ma (this last date

56

FIGURE 2. Orthogonal demagnetization (Zijderveld) plots of representative


samples. Solid squares indicate declination (horizontal component); open squares
indicate inclination (vertical component). First step is NRM, followed by AF steps
of 25, 50, and 100 Gauss, then thermal steps from 200 to 630C in 50C increments.
Each division equals 10-5 emu.

was considered questionable by Cather et al., 1987). Thus, the Baca


Formation must be older than 38 Ma, which rules out a Chadronian age
for the fauna, since the Chadronian ranges from 34 to 37 Ma (Prothero
and Swisher, 1992). Instead, this is consistent with a Duchesnean interpretation of the fossils, which ranges in age from 37 to 40 Ma
(Prothero and Emry, 1996). However, the lower limit of the age of the
Baca Formation cannot be determined from radioisotopic dates, and
can only be constrained by the fauna, and by magnetic stratigraphy.
METHODS
In the spring of 2003, sampling was conducted on the thickest,
most fossiliferous section of the Baca Formation on the southwestern
face of Mariano Mesa from the base of Sonoreno Draw to the tuff bed
at the top of the mesa (NW NW SE , sec. 21, T3N R16W,
Mariano Springs 7.5-minute quadrangle, Catron County, New Mexico).
Sixteen paleomagnetic sites (3 samples per site) were taken from regularly-spaced intervals covering 80 m of section, with samples of both
resistant sandstone ledges and also mudstones. Samples were taken as
oriented blocks of rock with simple hand tools, and then wrapped and
carried back to the laboratory. There, they were subsampled into cores
using a drill press, or if the sample was too crumbly, casts into disks of
Zircar aluminum ceramic. The samples were then analyzed on a 2G
cryogenic magnetometer with an automatic sample changer at the California Institute of Technology. After measurement of NRM (natural
remanent magnetization), they were demagnetized in alternating fields
(AF) of 25, 50, and 100 Gauss to prevent the remanence of multi-domain grains from being baked in, and to examine the coercivity behavior of each specimen. AF demagnetization was followed by thermal
demagnetization of every sample in 50C steps from 200 to 630C to
get rid of high-coercivity chemical overprints due to iron hydroxides
such as goethite, and to determine how much remanence was left after
the Curie temperature of magnetite (580C) was exceeded.
Results were plotted on orthogonal demagnetization
(Zijderveld) plots, and average directions of each sample were determined by the least-squares method of Kirschvink (1980). Mean directions for each sample were then analyzed using Fisher (1953) statistics, and classified according to the scheme of Opdyke et al. (1977).

FIGURE 3. Stereonet of mean of reversed sites. Open circle and dashed line indicate
mean of reversed samples (upper hemisphere projection). Solid square and solid
line indicates projection of reversed mean and error estimates to the lower hemisphere
of the stereonet. Solid square indicates the expected Eocene pole position, which is
antipodal to the reversed mean for the Baca Formation.

RESULTS
Orthogonal demagnetization (Zijderveld) plots of representative samples are shown in Figure 2. In nearly every sample, there was
a single component of remanence that was pointed south and down at
NRM, indicating that the sample is reversed in polarity. There was a
slight high-coercivity component (shown by the minimal drop of intensity in the first three demagnetization steps in Figure 2), probably due
to some chemical remanence due to iron hydroxides, such as goethite.
This is not surprising, given the red color of the rocks, but apparently
the overprint was not significant, because it was quickly removed by
thermal demagnetization at 200C (above the temperature at which
iron hydroxides are dehydrated to hematite). The rest of the remanence
appears to be held in magnetite, because it declined in intensity rapidly
through thermal demagnetization, vanished above the Curie temperature of magnetite and no remanence was left at 600 or 630C (Fig. 2).
All 16 sites at Mariano Mesa were reversed in polarity (Fig. 3).
Site statistics are given in Table 1. All but one site was statistically
TABLE 1. Paleomagnetic data from the Baca Formation
SITE
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

DEC
163.7
184.6
140.3
171.2
180.1
187.2
194.5
199.3
143.6
161.7
205.9
163.6
179.5
192.0
173.6
182.7

INC
-74.1
-57.8
-62.9
-50.5
-75.5
-49.9
-56.9
-63.1
-66.1
-68.4
-42.0
-59.5
-43.6
-54.2
-62.9
-39.2

K
28.4
11.1
59.6
5.8
32.0
6.6
11.0
9.8
52.1
48.0
14.1
9.8
10.8
17.9
9.3
13.4

95
23.6
38.9
16.1
56.4
22.2
52.6
39.1
41.8
17.2
36.9
34.2
41.6
39.6
30.1
42.8
35.0

57

FIGURE 4. Lithostratigraphy and magnetic stratigraphy of the Baca Formation at


Mariano Mesa. Declination and inclination of magnetic sites are shown. Solid circles
are sites that are statistically removed from a random distribution at the 95%
confidence level (Class I sites of Opdyke et al., 1977); open circle represents a site
with only two samples, so site statistics could not be calculated (Class II site of
Opdyke et al., 1977).

significant, i.e., separated from a random distribution at the 95% confidence level (Class I sites of Opdyke et al., 1977); the exception (site
10) failed the significance test because one sample crumbled, so only
two samples remained to be measured (Class II site of Opdyke et al.,
1977).
Because only reversed polarity sites and no normal sites were
obtained, it is impossible to conduct a conventional reversal test for
stability. However, the reversed mean (D = 179.1, I = -59.1, k = 32.1,95
= 6.6, n = 16) is antipodal to the normal magnetic direction expected
for this latitude in the Eocene (Fig. 4), so it is clearly a primary or
characteristic remanence, and overprinting has been removed.
DISCUSSION
By itself, the 80 m of reversed polarity in the Mariano Mesa
section of the Baca Formation is not age diagnostic. However, the presence of a Duchesnean mammal fauna in the interval of reversed polarity, and the overlying date of ~ 38.0 Ma places age constraints that give
one unambiguous correlation of this reversed polarity magnetozone (Fig.
5). According to Prothero and Emry (1996), the early Duchesnean is
entirely normal in polarity (Chron C18n), and the only reversed magnetic interval in the time scale of Berggren et al. (1995) that is older
than 38.0 Ma is Chron C17r (38.0-38.3 Ma). Thus, our best interpretation is that the Mariano Mesa section of the Baca Formation correlates
with Chron C17r.
This correlation is consistent with the interpretation of the
Galisteo Formation in north-central New Mexico, which yields middlelate Duchesnean mammals and correlates with Chron C17n-C18n
(Prothero and Lucas, 1996). It is also consistent with the earlier interpretation that the Baca mammals are comparable to those of the Porvenir

FIGURE 5. Correlation of the Baca Formation paleomagnetic section, based on the


dates and age constraints discussed in the text. Magnetic stratigraphy of the Galisteo
Formation after Prothero and Lucas (1996), and of the Texas section after Prothero
(1996). Time scale after Berggren et al. (1995) and Prothero and Emry (1996).

local fauna of Trans-Pecos Texas, which correlates with Chron C17nC17r, and is underlain by a 40Ar/39Ar date of 37.8 0.15 Ma on the
Buckshot Ignimbrite (Prothero, 1996). The middle Eocene age of the
upper parts of both the Baca and Galisteo formations in north-central
and west-central New Mexico indicates that the transition from
Laramide deformation to intermediate volcanism began in this part of
New Mexico in the late middle Eocene (around 38 Ma).
CONCLUSION
Magnetostratigraphic analysis of the Mariano Mesa section shows
that the upper part of the Baca Formation correlates with Chron C17r
(38.0-38.3 Ma), i.e., is late Duchesnean in age. Older parts of the formation may be as old as Bridgerian or Uintan. This correlation of the
upper part of the Baca Formation is comparable to that of the Duchesnean
Galisteo Formation of north-central New Mexico, and suggests that the
transition from Laramide tectonism to intermediate volcanism in northcentral and west-central New Mexico began in the middle Eocene.
ACKNOWLEDGMENTS
We thank Paula Dold, Matthew Liter, Jingmai OConnor, and
Francisco Sanchez for help with sampling, and Francisco Sanchez for
help with the lab work. We thank Dr. Joseph Kirschvink for access to
the Caltech paleomagnetics lab. We thank Justin Spielman for help
with the stratigraphy and geology. We thank Steve Cather and Kate
Zeigler for helpful reviews of this paper. This research was supported
by grants to Prothero by the Donors of the Petroleum Research Fund of
the American Chemical Society, and by NSF grant 03-09538.

REFERENCES
Berggren, W.A., Kent, D. V., Swisher, C. C., III, and Aubry, M.-P., 1995, A revised
Cenozoic geochronology and chronostratigraphy: SEPM Special Publication, v.
54, p. 129-212.
Cather, S.M. 1980. Petrology, diagenesis, and genetic stratigraphy of the Eocene
Baca Formation, Alamo Navajo Reservation and vicinity, Socorro County, New
Mexico [M.A. thesis]: Austin, University of Texas, 243 p.
Cather, S.M. 1982. Lacustrine sediments of Baca Formation (Eocene), western
Socorro County, New Mexico: New Mexico Geology, v. 4, p. 1-6.
Cather, S. M., McIntosh, W. C. and. Chapin, C.E., 1987, Stratigraphy, age, and
rates of deposition of the Datil Group (upper Eocene-lower Oligocene), westcentral New Mexico: New Mexico Geology, v. 9, p. 50-54.
Fisher, R. A., 1953, Dispersion on a sphere: Proceedings of the Royal Society, v.

A217, p. 295-305.
Kirschvink, J. L., 1980, The least-squares line and plane and the analysis of paleomagnetic data: examples from Siberia and Morocco: Geophysical Journal of the
Royal Astronomical Society, v. 62, p. 699-718.
Lucas, S. G., 1983, The Baca Formation and the Eocene-Oligocene boundary in
New Mexico: New Mexico Geological Society, Guidebook 34, p. 187-192.
Lucas, S. G., 1990, Middle Eocene mammal from the base of the Baca Formation,
west-central New Mexico: New Mexico Journal of Science, v. 30, p. 35-39.
Lucas, S. G., 1992, Redefinition of the Duchesnean land mammal age, late Eocene
of western North America; in Prothero, D. R. and Berggren, W. A., eds., EoceneOligocene climatic and biotic evolution: Princeton, Princeton University Press,
p. 88-105.

58
Lucas, S. G. and Williamson, T. E., 1993, Eocene vertebrates and Laramide stratigraphy of New Mexico: New Mexico Museum of Natural History and Science,
Bulletin 2, p. 145-158.
Opdyke, N. D., Lindsay, E. H., Johnson, N. M., and Downs, T., 1977, The paleomagnetism and magnetic polarity stratigraphy of the mammal-bearing section
of Anza-Borrego State Park, California: Quaternary Research, v. 7, p. 316-329.
Prothero, D.R., 1996, Magnetostratigraphy of Eocene-Oligocene transition in TransPecos Texas; in Prothero, D. R. and Emry, R. J., eds., The terrestrial EoceneOligocene transition in North America: Cambridge, Cambridge University Press,
p. 174-183.
Prothero, D. R., and Emry, R. J. 1996, Summary; in Prothero, D. R. and Emry, R. J.,
eds., The terrestrial Eocene-Oligocene transition in North America: Cambridge,
Cambridge University Press, p. 646-664.
Prothero, D. R., and Lucas, S. G., 1996, Magnetostratigraphy of the Duchesnean
part of the Galisteo Formation, New Mexico; in Prothero, D. R. and Emry, R. J.,
eds., The terrestrial Eocene-Oligocene transition in North America: Cambridge,
Cambridge University Press, p. 184-190.
Prothero, D.R. and Swisher, C. C., III, 1992, Magnetostratigraphy and geochronology of the terrestrial Eocene-Oligocene transition in North America; in Prothero,

D. R. and Berggren, W.A., eds., Eocene-Oligocene climatic and biotic evolution: Princeton, Princeton University Press, p. 46-74.
Schiebout, J.A., and Schrodt, A.K. 1981. Vertebrate paleontology of the lower Tertiary Baca Formation of western New Mexico: Geological Society of America
Bulletin Part I, v. 92, p. 976-979.
Schrodt, A. K., 1980, Depositional environments, provenance, and vertebrate paleontology of the Eocene-Oligocene Baca Formation, Catron County, New Mexico
[M.S. thesis]: Baton Rouge, Louisiana State University, 174 p.
Snyder, D. O., 1970, Fossil evidence of the Eocene age of the Baca Formation: New
Mexico Geological Society, Guidebook 21, p. 65-67.
Snyder, D.O., 1971, Stratigraphic analysis of the Baca Formation, west-central New
Mexico [Ph.D. dissertation]: Albuquerque, University of New Mexico, 160 p.
Wilpolt, R. H. and Wanek, A.A., 1946, Geology of the region from Socorro and San
Antonio east of Chupadera Mesa, Socorro County, New Mexico: U.S. Geological Survey, Oil and Gas Investigations Map OM-121.
Wilson, J. A., 1978, Stratigraphic occurrence and correlation of early Tertiary vertebrate faunas, Trans-Pecos Texas. Part I: Vieja area. Texas Memorial Museum,
Bulletin 25, p. 1-42.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

81

THE SKULL OF THE EOCENE PERISSODACTYL LAMBDOTHERIUM


AND ITS PHYLOGENETIC SIGNIFICANCE
SPENCER G. LUCAS1 AND LUKE T. HOLBROOK2
1

New Mexico Museum of Natural History, 1801 Mountain Rd. NW, Albuquerque, NM 87104
2
Department of Biological Sciences, Rowan University, Glassboro, NJ 08028

AbstractWe describe the most complete skull known of Lambdotherium popoagicum, which is from the
Eocene Wasatch Formation near Big Piney in the northern part of the Bridger basin of southwestern Wyoming.
A cladistic analysis of the relationships of Lambdotherium and other perissodactyls places it in a clade with the
palaeotheriids Palaeotherium and Plagiolophus and not with the brontotheriid Eotitanops. Thus, the analysis
does not support a close relationship between Lambdotherium and brontotheriids, but instead supports the
notion that Lambdotherium is most closely related to palaeotheres.
Key words: Lambdotherium, Eocene, Wyoming, Brontotheriidae, Perissodactyla

INTRODUCTION
Cope (1880) coined the name Lambdotherium popoagicum for a
left dentary fragment with p2-m3 collected from lower Eocene strata in
the Wind River basin of Wyoming. The perissodactyl affinities of
Lambdotherium were immediately evident, and Cope (1880, 1884) allied it with the brontotheres. Osborn (1929) reviewed subsequent discoveries in Wyoming and Colorado and recognized five species of
Lambdotherium, which he considered the oldest and most primitive
brontothere. Osborn (1929, fig. 233) also presented a reconstruction of
the skull of Lambdotherium based on fragmentary material (Fig. 1).
Bonillas (1936) revised the species-level taxonomy of
Lambdotherium, recognizing only one valid, dentally variable species,
L. popoagicum. This view has been followed by subsequent revisers,
and Lambdotherium fossils are now known from lower Eocene strata
in Colorado, Wyoming and Ellesmere Island, Canada (Wallace, 1980;
Mader, 1998). And, Lambdotherium is generally regarded as the index
taxon of late Wasatchian (Lostcabinian) time (the Lambdotherium
biochron of Lucas, 1998).
Gazin (1952, p. 67-68, pl. 10) briefly described and illustrated a
skull of Lambdotherium popoagicum from the Wasatch Formation near
Big Piney in the Bridger basin of western Wyoming. This is the most
complete skull known of Lambdotherium, and it merits more extensive
description, illustration and an assessment of its phylogenetic significance to the relationships of Lambdotherium and brontotheres.
Institutional abbreviations: USNM = National Museum of
Natural History, Smithsonian Institution, Washington, D. C.
DESCRIPTION
USNM 19761 is a nearly complete skull (Figs. 2-3; Table 1). Its
greatest deficiencies are the incomplete zygomatic arches and the virtual lack of a basicranium. Close examination also suggests that the
specimen was found in essentially three piecespalate, skull roof and
braincasethat were fused together with plaster. This plaster occupies
much of the medial walls of the orbits, the lateral aspect of the rostrum
and parts of the lateral edges of the braincase.
The skull is relatively long and dolichocephalic. It has a tubular
rostrum anterior to the p2, then posteriorly widens rapidly to its broadest point, which is across the zygomatic arches. A supraorbital notch is
visible on the anterolateral aspect of the left supraorbital process. The
braincase is long and broadest just behind the postorbital constriction,
and peaks dorsally at a long, relatively tall sagittal crest. The sagittal
crest meets a well-developed lambdoidal crest posteriorly that projects
further posteriorly to overhang the occiput. In dorsal view, only the
anterior portions of the nasal sutures are visiblethe suture that sepa-

FIGURE 1. Reconstruction of the skull of Lambdotherium popoagicum,


from Osborn (1929, fig. 233).

rates the two nasals, and their lateral sutures to the maxillae.
In left lateral view, the nasal incision is retracted to above the
canine. The premaxilla-maxilla suture is evident between the I3 alveolus and the canine. The maxilla excludes the premaxilla from contact
with the nasal dorsally. The rostrum (lateral aspect of the maxilla) between the nasal incision and anterior root of the zygomatic arch is evenly
convex; there is no preorbital fossa in the maxilla. A small infraorbital
foramen in the maxilla is above P3 and opens anteriorly. The anterior
root of the zygomatic arch is thick and begins above the M1. The jugal
maxilla suture is a curved line just anterior to the orbit, so the jugal
makes up the antero-ventral border of the orbit.
The anterior edge of the orbit begins over the M1/M2 juncture.
A distinct postorbital ridge overhangs the postero-dorsal edge of the
orbit. There is a pronounced postorbital constriction that separates the
orbit from the braincase. The glenoid fossa is short, wide, trapezoidal
in outline and is a flat to slightly concave surface that extends onto the
root of the zygomatic arch. The postglenoid process is peg-like and
anteriorly oriented, as in Palaeotherium (Holbrook, 2001, fig. 7A). There
is no postglenoid foramen, and the postglenoid process is followed by a
narrow external auditory meatus. The mastoid-paroccipital process is
also small and connected to the lambdoidal crest by a ridge. A distinct
exposure of the mastoid is visible in posterior view between the squamosal and exoccipital, and there is a mastoid foramen.

82

FIGURE 2. Lambdotherium popoagicum, USNM 19761, skull in right lateral (A), left lateral (B), dorsal (C) and ventral (D) views.

In ventral view, the left premaxilla is a thin arc of bone with parts
of the roots of two incisors. Apparently there was a single, large incisive
foramen. The canine is relatively large, with an antero-posteriorly oval
cross section. A moderately long diastema follows, and there is a trenchant, two-rooted P1. A short diastema separates the P1 from the P2.
The P2-M3 form an unbroken tooth row, and the cheek tooth rows
diverge posteriorly only slightly. They thus border a relatively long and
narrow palate. The internal nares are partly covered by the palatines,
which have separated and been overlapped along the midline, and they
are relatively long, narrow and extend anteriorly to a line through the
posterior portions of the M2s. Only a fragment of the basisphenoid is
preserved in matrix behind the palatines, and the configuration of the
foramen ovale cannot be determined.
The occiput is tall and narrow, with a relatively narrow and dorsally flat lambdoidal crest. The posterior aspect of the occiput is distinctly concave dorsally, and convex immediately dorsal to the foramen
magnum.
The cheek teeth correspond well to already described and illustrated upper teeth of Lambdotherium popoagicum (Osborn, 1929;
Bonillas, 1936). Dental measurements of USNM 19761 (Table 1) are
within the range of or very close to the range of measurements of the

teeth of L. popoagicum reported by Bonillas (1936) and Wallace (1980).


PHYLOGENETIC SIGNIFICANCE
Lambdotherium was long considered the basal brontothere (e.g.,
Osborn, 1929; Simpson, 1945) until Wallace (1980) suggested, based
on dental evidence, that Lambdotherium is more closely related to
palaeotheres (also see Mader, 1989, 1998). Cladistic analysis by
Froehlich (1999), however, placed brontotheres within the palaeothere
clade, so Froehlich concluded that Lambdotherium may, indeed, represent the first brontothere, and he suggested a derivation of North American brontotheres from a European, palaeothere ancestry.
The general appearance of USNM 19761 is that of a primitive
perissodactyl, particularly the long rostrum with shallow maxillary fossa
and a shallow narial incision. The anterior orientation of the postglenoid process is shared with several perissodactyl taxa, including
brontotheriids, chalicotheres, palaeotheres and equids. There are some
features that distinguish this skull from those of brontotheriids. A supraorbital notch or foramen is absent from brontotheriid skulls, although
it is found in chalicotheres, palaeotheres, and at least some equids.
Also, the dolichocephalic skull proportions of USNM 19761 differ from
the relatively short rostral regions that are a derived feature of

83

FIGURE 3. Lambdotherium popoagicum, USNM 19761, occipital view of skull (A), and occlusal views of left cheek teeth (B) and right cheek teeth (C).

brontotheriids (Mader, 1989).


In order to test whether these differences constitute evidence
against a close relationship between Lambdotherium and brontotheriids,
we undertook a phylogenetic analysis of 53 characters among 25 selected perissodactyl taxa and one outgroup, using data primarily from
Holbrook (1999, 2001) and USNM 19761. (A more robust analysis
should, nevertheless, be performed that includes several brontothere
taxa, not just Eotitanops.) Characters include features of the skull, dentition, and postcranial skeleton. A branch-and-bound search using
PAUP* 4.0 produced 220 shortest trees of 106 steps with a consistency
index of 0.6509. The strict consensus of these shortest trees is given in

Figure 4. Lambdotherium is placed in a clade with the palaeotheriids


Palaeotherium and Plagiolophus and not with the brontotheriid Eotitanops.
Thus, the data do not support a close relationship between
Lambdotherium and brontotheriids, but instead support Wallaces (1980)
notion that Lambdotherium is related to palaeotheres.
ACKNOWLEDGMENTS
Robert Emry and Robert Purdy made it possible to study specimens in the USNM collection. Comments on the manuscript by R. Emry,
P. Kondrashov and M. Mihlbachler improved its content.

84
TABLE 1. Measurements (in mm) of USNM 19761, Lambdotherium
popoagicum.

FIGURE 4. Strict consensus of 96 shortest trees resulting from a branch-and-bound


analysis of 26 taxa and 62 characters.

Dorsal length of skull


(tip of nasals to tip of lambdoidal crest)
Length of skull anterior to orbits
Length of skull posterior to orbits
Width of nasals
Width across zygomatic arches
Width between orbits
Width across postorbital constriction
Width of lambdoidal crest
Maximum diameter of orbit
Length of palate
(anterior edge palatines to anterior edge internal nares)
Length of glenoid fossa
Width of glenoid fossa
Length C-P1 diastema
Length P1-P2 diastema
Length P1-M3
P1 length
P1 width
P2 length
P2 width
P3 length
P3 width
P4 length
P4 width
M1 length
M1 width
M2 length
M2 width
M3 length
M3 width

182
85
80
13
98
67
33
36
26
87
16
21
12.3
5.5
75
5.4
3.1
7.3
6.2
8.2
10.1
8.6
11.4
11.4
14.6
12.5
16.1
12.1
17.3

REFERENCES
Bonillas, Y., 1936, The dentition of Lambdotherium: Journal of Mammalogy, v.
17, p. 139-142.
Cope, E. D., 1880, The badlands of the Wind River and their fauna: The American
Naturalist, v. 14, p. 745-748.
Cope, E. D., 1884, The Vertebrata of the Tertiary formations of the West, Book I: U.
S. Geological Survey of the Territories Report, v. 3, 1009 p.
Froehlich, D. J., 1999, Phylogenetic systematics of basal perissodactyls: Journal of
Vertebrate Paleontology, v. 19, p. 140-159.
Froehlich, D. J., 2002, Quo vadis eohippus? The systematics and taxonomy of early
Eocene equids (Perissodactyla): Zoological Journal of the Linnean Society, v.
134, p. 141-256.
Gazin, C. L., 1952, The lower Eocene Knight Formation of western Wyoming and
its mammalian fauna: Smithsonian Miscellaneous Collections, v. 117, no. 18,
82 p.
Holbrook, L. T., 1999, The phylogeny and classification of tapiromorph perissodactyls (Mammalia): Cladistics, v. 15, p. 331-350.
Holbrook, L. T., 2001, Comparative osteology of early Tertiary tapiromorphs (Mammalia, Perissodactyla): Zoological Journal of the Linnaean Society, v. 132, p. 154.
Hooker J. J. 1989. Character polarities in early perissodactyls and their significance
for Hyracotherium and infraordinal relationships; in Prothero, D. R. and Schoch,
R. M., eds., The evolution of perissodactyls: New York, Oxford University Press,
p. 79-101.
Hooker, J.J., 1994, The beginning of the equoid radiation: Zoological Journal of the
Linnaean Society of London, v. 112, p. 29-63.

Lucas, S. G., 1998, Fossil mammals and the Paleocene/Eocene Series boundary in
Europe, North America, and Asia; in Aubry, M-P., Lucas, S. G. and Berggren,
W. A., eds., Late Paleocene-early Eocene climatic and biotic events in the marine and terrestrial records: New York, Columbia University Press, p. 451-500.
Mader, B. J., 1989, The Brontotheriidae: A systematic revision and preliminary
phylogeny of North American genera; in Prothero, D. R. and Schoch, R. M.,
eds., The evolution of perissodactyls: Oxford, Oxford University Press, p. 458484.
Mader, B. J., 1998, Brontotheriidae; in Janis, C. M., Scott, K. M. and Jacobs, L. L.,
eds., Evolution of Tertiary mammals of North America. Volume 1: Terrestrial
carnivores, ungulates, and ungulate like mammals: Cambridge, Cambridge
University Press, p. 525-536
Osborn, H. F., 1929, The titanotheres of ancient Wyoming, Dakota, and Nebraska:
U. S. Geological Survey, Monograph 55, 953 p.
Radinsky L. B., 1966, The families of the Rhinocerotoidea (Mammalia,
Perissodactyla): Journal of Mammalogy, v. 47, p. 631-639.
Simpson, G.G., 1945, The principles of classification and a classification of mammals: Bulletin of the American Museum of Natural History, v. 85, p. 1-350.
Thewissen, J. G. M. and Domning, D. P., 1992, The role of phenacodontids in the
origin of the modern orders of ungulate mammals: Journal of Vertebrate Paleontology, v. 12, p. 494-504.
Wallace, S. M., 1980, A revision of North American early Eocene Brontotheriidae
(Mammalia, Perissodactyla) [M.S. thesis]: Boulder, University of Colorado, 157
p.

85
APPENDIX 1
Character matrix used for parsimony analysis.
Homogalax

Cardiolophus

?
?

Isectolophus

Eomoropus

Moropus

Lophiodon

Heptodon

Helaletes

Colodon

Protapirus

Tapirus

Hyrachyus

Rostriamynodon

Amynodon

Triplopus

Hyracodon

Teletaceras

Trigonias

Subhyracodon

Uintaceras

Hyracotherium

Eotitanops

Lambdotherium

Phenacodus

Palaeotherium

Plagiolophus

0/1

Homogalax

Cardiolophus

Isectolophus

Eomoropus

Moropus

Lophiodon

Heptodon

Helaletes

Colodon

Protapirus

Tapirus

Hyrachyus

Rostriamynodon

Amynodon

Triplopus

Hyracodon

Teletaceras

Trigonias

Subhyracodon

Uintaceras

Hyracotherium

Eotitanops

Lambdotherium

Phenacodus

Palaeotherium

Plagiolophus

86
Homogalax

Cardiolophus

Isectolophus

Eomoropus

Moropus

Lophiodon

Heptodon

Helaletes

Colodon

Protapirus

Tapirus

Hyrachyus

Rostriamynodon

Amynodon

Triplopus

Hyracodon

Teletaceras

Trigonias

Subhyracodon

Uintaceras

Hyracotherium

Eotitanops

Lambdotherium

Phenacodus

Palaeotherium

Plagiolophus

APPENDIX 2
Description of characters used in this analysis. More detailed descriptions of these characters and their scoring can be found in Holbrook
(1999, 2001).
1. Nasals short (1). The nasals of perissodactyls primitively extend to the point above the anterior tip of the premaxilla or further, as
in Hyracotherium. Possession of significantly shorter nasals is a derived condition.
2. Nasals posteriorly broad (1). The nasals of most mammals are
splint- or diamond-shaped, and the posterior portion of the nasals intrudes between the frontals. This is probably the primitive condition
for eutherians. The nasals of perissodactyls are unique in having a triangular shape, where the base of the triangle is a suture with the frontal that does not intrude but, instead, runs transversely.
3. Lacrimal small, not contacting nasal (1); or large and not contacting nasal (2). Nasolacrimal contact is primitive for perissodactyls,
as demonstrated by the presence of this contact in many tapiromorph
and non-tapiromorph perissodactyls. This contact is the consequence
of the broad posterior portion of the nasals and the prominent facial
exposure of the lacrimal. Reduction of the facial exposure of the lacrimal results in loss of nasolacrimal contact, a derived state. The absence of nasolacrimal contact in Phenacodus is a fundamentally different condition from that seen in some tapiromorphs, such as Heptodon.
Phenacodus possesses a prominent facial exposure of the lacrimal, but
its nasals are very narrow and do not spread posteriorly (as in perissodactyls) to reach the lacrimals. To reflect this fundamental difference,
Phenacodus is scored as 2 for this character. Although the condition
in Phenacodus may not actually be derived, this score reflects the fact
that the condition in this genus is different from that in the other taxa.
4. Premaxilla robust, not contacting nasal (1); or small and not
contacting nasal (2). The primitive condition of the premaxilla, seen in
many perissodactyls and non-perissodactyl outgroups, is a relatively
small bone with a prominent ascending process that contacts the nasals. In some perissodactyls, such as tapirids, the premaxilla is robust
and does not contact the nasals. Despite the fact that its only contact is
with the maxilla, the premaxilla is still a well-developed element in

condition 1. In rhinocerotids, the ascending process of the premaxilla is reduced and does not contact the nasals; the premaxillae of
rhinocerotids are generally not well-developed.
5. Incisive foramen single and median (1). The contact between
the premaxilla and maxilla on the palate is pierced by one or two incisive foramina. Primitively, a bilateral pair of foramina is present, as in
Hyracotherium. In some perissodactyls, a single, median foramen is
present.
6. Maxillary (preorbital) fossa well-developed pocket (1); or
vertical groove (2). The facial portion of the maxilla primitively possesses a shallow or no fossa of any note, as in Hyracotherium. In some
rhinocerotoids, a prominent fossa is present. Some tapiroids possess
another type of fossa, a vertical groove anterior to the orbit.
7. Infraorbital foramen positioned over molars (1). The infraorbital foramen of most mammals is positioned over the upper premolars,
usually P2 or P3, and this is the position is in most perissodactyls. The
infraorbital foramen of amynodontids is positioned more posteriorly,
over the molars and often within the posterior part of the maxillary
fossa.
8. Posterior edge of narial incision retracted to a point over P4 or
molars (1); or retracted as in state 1 and posteroventrally excavated
(2). The narial incision primitively has its posterior border over the
first premolar or more anterior. Several tapiromorphs have a border
positioned much more posteriorly. In Helaletes and Colodon, the ventral aspect of the posterior border is deep and rounded, giving the impression of a keyhole shape. This character was treated as ordered.
9. Supraorbital foramen present (1). The postorbital process of
the frontal is primitively not notched or pierced, as in Hyracotherium
and Phenacodus. In a number of perissodactyl taxa, a foramen pierces
this process, or else there is a distinct notch. Ontogenetically, the foramen forms by the closing of the notch; so, the presence of a notch or a
foramen is scored the same.

10. Foramen ovale confluent with medial lacerate foramen (1).


In most perissodactyls and Phenacodus, the foramen ovale and medial
lacerate foramen are separate.
11. Postglenoid foramen absent (1). This foramen is present in
Phenacodus and a several perissodactyl taxa, and its presence is considered to be primitive for perissodactyls.
12. Postglenoid process facing anterolateral (1). The postglenoid process of the squamosal of Hyracotherium and Phenacodus is
small, peg-like, and faces anteriorly; this is the primitive condition. In
several tapiromorphs, this process is large, flattened, and obliquely oriented.
13. Anterior face of postglenoid process convex with median
ridge (1). The anterior face of the postglenoid process is primitively
flat or concave. In some rhinocerotoids, this face has become convex
and divided into medial and lateral portions by a ridge.
14. Posttympanic process short (1). In many perissodactyls, the
posttympanic process of the squamosal is about as long as the postglenoid process. The posttympanic process is significantly shorter in
Eomoropus and Moropus.
15. Bulla fused to skull (1). In most perissodactyls, the bulla is
not fused to the skull and is rarely preserved.
16. Mastoid exposure lateral (1) or amastoid (2). In most eutherians, the mastoid portion of the periotic is exposed posteriorly.
17. Postcotyloid process of dentary present (1). Present in
rhinocerotids and Uintaceras, this process is a buttress on the posterior
edge of the ascending ramus of the mandible, just below the mandibular condyle. This process is absent in other perissodactyls and nonperissodactyls.
18. Acromion process of scapula absent (1). A distinct acromion
process is present on the scapula of Hyracotherium and many nonperissodactyls, including Phenacodus, and its presence is considered
to be primitive.
19. Deltopectoral crest of humerus prominent and hooking laterally (1). The deltopectoral crest of non-tapiromorph perissodactyls and
non-perissodactyls is either absent or a low ridge. Both of these conditions are scored as primitive here.
20. Entepicondylar foramen of humerus absent (1). The presence of this foramen is considered to be primitive for eutherians
(Thewissen and Domning, 1992), and it is present in Phenacodus.
21. Capitulum of humerus keeled (1). The capitulum is primitively a rounded surface articulating with the radius, as in Phenacodus.
In perissodactyls, the capitulum is trochleate due to a median ridge or
keel.
22. Lateral process of proximal radius weak or absent (1). The
radius of Hyracotherium shows the primitive condition, where the lateral articular facet for the humerus extends beyond the shaft laterally
on a prominent process or tuberosity.
23. Radial facet of scaphoid with emarginated lunar contact (1)
or long, straight lunar contact (2). The border of the radial facet of the
scaphoid is relatively short in Hyracotherium and brontotheres, and
this condition is considered to be primitive.
24. Anterior iliac crest concave (1). The anterior crest of the
ilium is primitively convex, as in Phenacodus and many other eutherians.
25. Fovea capitis of femur centrally placed (1). In most perissodactyls and Phenacodus, the fovea capitis interrupts the margin of the
articular facet of the femoral head.
26. Lesser trochanter of femur weak or absent (1). The lesser
trochanter of Phenacodus and Hyracotherium is present as a prominent flange.
27. Medial trochlear ridge of femur expanded into tuberosity
(1). The medial and lateral trochlear ridges are primitively about equal
in size, as in Phenacodus and Hyracotherium.
28. Gastrocnemius (supracondylar) fossa of femur present (1).
A distinct fossa for attachment of the gastrocnemius is not present above

87
the lateral condyle on the posterior side of the femur in Phenacodus,
Hyracotherium, and brontotheriids.
29. Adductor tubercle of femur pronounced (1). In most perissodactyls and Phenacodus, the adductor tubercle is small or absent.
30. Patella broad and flattened (1) or broad with a distince medial process (2). Primitively, the patella is teardrop-shaped and
anteroposteriorly thick, as in Hyracotherium and Phenacodus.
31. Trochlea of astragalus laterally offset from neck (1) or overlapping onto short neck (2). The trochlea of equids and primitive
brontotheriids such as Eotitanops lies more or less directly above a
distinct neck of the astragalus.
32. Sustentacular and distal calcaneal facets of astragalus
confluent (1), confluent with a ridge formed at their junction (2), or
joined by a third facet (3). These facets are separate in many perissodactyls. Since state 2 is simply a special case of state 1, this character
was treated as ordered.
33. Process on posterodistal aspect of ectocuneiform present (1).
The distal facet of the ectocuneiform in non-tapiromorph perissodactyls is flat with no processes.
34. Navicular facet of astragalus saddle-shaped (1). The astragalar
head (navicular facet) of non-perissodactyls is not saddle-shaped. In
Phenacodus, this facet is rounded for a ball-and-socket joint.
35. Pes tridactyl (1). Five digits on the pes is primitive for eutherians and is seen in Phenacodus.
36. Upper incisors conical (1), buccolingually compressed (2),
or absent (3). Spatulate upper and lower incisors are found in all outgroup
taxa and many tapiromorphs, and this condition is scored as (0).
37 and 38. I1 chisel-shaped (1) (char. 37); i2 lanceolate (1) or
absent (2) (char. 38). State 1 for both of these characters is characteristic of rhinocerotids (Radinsky, 1966). The primitive condition is both
incisors spatulate or conical.
39 and 40. Number of upper (char. 39) and lower (char. 40) incisors two (1), one (2), or none (3). The primitive condition for perissodactyls is three upper and lower incisors.
41 and 42. Upper (char. 41) and lower (char. 42) canines absent
(1). Canines are primitively present for perissodactyls.
43. Postcanine diastema long (1). Primitively, the diastema between the canine and first premolar is short, as in Phenacodus and
Hyracotherium, or absent, as in Eotitanops and Isectolophus.
44 and 45. P1 (char. 44) and p1 (char. 45) abutting P2/p2 (1) or
absent (2). In Phenacodus and Hyracotherium, there is a short diastema
between the first and second premolars. In many tapiromorphs, this
diastema has become closed, so that there are no spaces between the
seven cheek teeth, or, in others, one or both of the first premolars may
be lost.
46. Upper molar parastyles small and narrow (1). Primitively,
the parastyle is large and shaped like a lozenge.
47 and 48. Upper molar paraconules (char. 47) and metaconules
(char. 48) indistinct or absent (1). These cusps are present as bumps or
swellings on the protoloph and metaloph of less lophodont (and therefore primitive) forms.
49. M3 metastyle labially deflected (1), lingually deflected (2),
or absent (3).
50. Lower molar cristid obliqua (metalophid) labially positioned
(1). In all non-perissodactyl outgroups, as well as all brontotheres and
equoids, the metalophid is an oblique, mesiolingually directed crest.
51. Lower molar metastylid absent (1). A lower molar metastylid
(or twinned metaconid of Hooker [1994]) is found in non-perissodactyl outgroups, as well as many non-tapiromorph perissodactyls; thus,
the presence of this cusp is considered to be primitive.
52. m3 hypoconulid small and narrow (1) or absent (2). A prominent hypoconulid is present on the m3 of Hyracotherium and other
non-tapiromorphs.
53. Upper molar premetaconule crista/lower molar lingual
postcristid absent (1). This character comes from Hooker (1989).

88

Periptychus rhabdodon, right manus and left pes (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

89

ACHAENODON (MAMMALIA, ARTIODACTYLA) FROM THE EOCENE CLARNO


FORMATION, OREGON, AND THE AGE OF THE HANCOCK QUARRY LOCAL FAUNA
SPENCER G. LUCAS1, SCOTT E. FOSS2 AND MATTHEW C. MIHLBACHLER3
1

New Mexico Museum of Natural History, 1801 Mountain Rd. NW, Albuquerque, NM 87104; 2 John Day Fossil Beds National Monument, 32651 Hwy.
19, Kimberly, OR 97848; 3Division of Paleontology, American Museum of Natural History, Central Park West at 79th St., New York, NY 10024

Abstract The artiodactyl family Achaenodontidae, represented by the new species Achaenodon fremdi, can be
added to the fossil mammal assemblage of the Eocene Hancock mammal quarry local fauna of the Clarno Formation
in north-central Oregon. Radioisotopic dates from the base of the John Day Formation directly above the Hancock
quarry indicate it is older than 39.5-40 Ma. A review of the fossil mammal assemblage from the Hancock quarry
indicates it is essentially a mixture of North American endemic taxa (Protapirus, Epihippus, Diplobunops,
Achaenodon) of Uintan age, and Asian immigrant taxa (Hemipsalodon, Teletaceras, Heptacodon) that do not
appear elsewhere in North America until the Duchesnean. A late Uintan age for the Hancock quarry local fauna
best fits the radioisotopic and mammalian biochronologic data.
Keywords: Achaenodon, Eocene, Clarno Formation, Oregon, Uintan

INTRODUCTION
The Clarno Formation of north-central Oregon (Fig. 1) yields an
Eocene mammal assemblage that has long been assigned either a
Duchesnean or a Chadronian age (e.g., Emry et al., 1987; Krishtalka et al.,
1987; Lucas, 1992; Hanson, 1996). Hanson (1996) reviewed this mammal assemblage, which he termed the Hancock [mammal] quarry local
fauna, and identified the following taxa: Hemipsalodon grandis,
Nimravidae, Haplohippus texanus, Epihippus gracilis, Protapirus
hancocki, Procadurcodon n. sp., Teletaceras radinskyi, Protitanops n.
sp., Heptacodon sp., Diplobunops n. sp. and Rodentia. Here, we document an addition to the Hancock quarry local fauna, a new species of the
artiodactyl Achaenodon. We then review the radioisotopic and mammalian biochronologic data to argue that the Hancock quarry local fauna is
late Uintan in age.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; CM = Carnegie Museum of Natural History, Pittsburgh; FMNH.PM = Field Museum of Natural History, fossil
mammal collection, Chicago; OMSI = Oregon Museum of Science and
Industry collection (now at the Condon Museum of Geology, Eugene);
PU = Princeton University collection (now at the Yale Peabody Museum, New Haven); SDSNH = San Diego Natural History Museum, San
Diego.
SYSTEMATIC PALEONTOLOGY
Order Artiodactyla
Genus Achaenodon
Achaenodon fremdi, new species
Fig. 2, Table 1
HolotypeOMSI 590, a right maxillary fragment with P4-M2
(Fig. 2).
Horizon and localityHancock mammal quarry, Clarno Formation, Oregon (Fig. 1).
Diagnosis Differs from other species of Achaenodon in the
more rounded shape of the P4 and the nearly equal size of the para- and
protocones; and in having a M1 with six (not just four) equally prominent cusps (protocone, paracone, metacone, paraconule, metaconule,
hypocone)
Description and comparisonThe P4 is dominated by a tall
paracone with a strong distal ridge that terminates with a small conule
(metacone?) near a well-developed ectocingulum. A strong protocone is
nearly the size of the paracone that also has a well-developed distal ridge

FIGURE 1. Index map showing location of Hancock quarry and generalized


stratigraphic section of the Clarno Formation and base of the overlying
John Day Formation (after Bestland and Retallack, 1994).

that meets a similarly strong posterior cingulum; this contains a small


cuspule (metaconule?). OMSI 590 differs from other specimens of
Achaenodon (including A. robustus and A. uintense) in the more rounded
shape of the tooth and the nearly equal size of the para- and protocones.
The M1 is four-rooted, square, and six-cusped, with strong preand post-cingula. This differs from nearly all other known specimens of
Achaenodon, in which the M1 and M2 are dominated by four cusps. In

90
TABLE 1. Dental measurements (in mm) of Achaenodon fremdi sp.
nov. from the Hancock quarry local fauna (holotype: OMSI 590) and
selected other specimens of Achaenodon.
Tooth position
P4L
P4W
M1L
M1W
M2L
M2W
a
b

OMSI 590
23.2
25.6
29.1
28.5
~29
~32

A. robustusa
23.0
27.5
21.8
25.8
24.5
31.8

A. uintensisb
27.5
35.5
25.0
31.0
27.0
37.0

PU 10033, from Gazin (1955, p. 40)


AMNH 2047, from Gazin (1955, p. 41)

tions to paleontology in Oregon.


AGE OF THE HANCOCK QUARRY LOCAL FAUNA
Two lines of data bear on the age of the Hancock quarry local
fauna, radioisotopic ages from the overlying welded tuff at the base of
the John Day Formation (Fig. 1) and the biochronological distribution of
the mammal taxa from the Hancock quarry itself. Here, we review these
data to argue that the Hancock quarry is of late Uintan age.
Radioisotopic Ages
FIGURE 2. OMSI 590, right maxillary fragment with P4-M2, holotype of
Achaenodon fremdi sp. nov., from the Clarno Formation, in labial (A) and
occlusal (B) views. Scale in mm.

other specimens of Achaenodon (including A. robustus and A. uintense),


accessory cusps may be present (especially the paraconule), but they
are never prominent, and in most specimens the condition is not observable due to wear. OMSI 590 also has a small hypocone; this is also not
observable on other specimens of Achaenodon.
As in the other species of Achaenodon, the M2 is significantly
larger than M1. The paraconule is more developed and the ectoflexus
more prominent, but the tooth is otherwise morphologically similar to
M1.
DiscussionThe size and morphology of OMSI 590 is diagnostic of Achaenodon. Indeed, it resembles other well-preserved upper dentitions of Achaenodon, including AMNH 2047 and MCZ 7749. Particularly significant is the postero-lingual cusp, which is an enlarged metaconule, a feature diagnostic of Achaenodon (Stucky, 1998).
OMSI 590 is difficult to compare to the type species of
Achaenodon, A. insolens, which is only known from the lower jaws,
AMNH 5143 (Cope, 1885, pls. 57-57a). However, the teeth of A. insolens
are not readily discernable from those of A. robustus, with which it cooccurs, and both may be the same species. Measurements of OMSI 590
(Table 1) are similar to those of both A. robustus and A. uintense. Gazin
(1955) noted that A. robustus and A. uintense may be the same species,
however, features of both the jaw ramus and skull roof support each as
a unique species.
OMSI 590 is not an entelodont, but is convergently similar in
much of its derived cusp pattern. Its relatively large size readily distinguishes it from the oldest North American entelodont, DuchesneanChadronian Brachyhyops (Colbert, 1938; Wilson, 1971; Russell, 1980;
Effinger, 1998). The size of OMSI 590 is within the size range of certain
Archaeotherium species. However, the earliest entelodont to display
even similar morphology is Archaeotherium crassum, which has its earliest record in the late Chadronian (Foss, 2001), which is 5.5 million
years later than the latest possible date for the Hancock quarry local
fauna.
EtymologyNamed for Ted Fremd, to honor his many contribu-

The basal welded tuff of the John Day Formation, which overlies
the Clarno Formation and is thus above the Hancock quarry (Fig. 1), has
been radioisotopically dated at least four times since the 1960s. Early K/
Ar analysis yielded an age of 32.8 Ma (Evernden and James, 1964,
corrected to 35.7 Ma using the revised decay constants of Dalrymple,
1979). Fission tracks ages of 37-38 Ma (Swanson and Robinson, 1968)
and 36.8 and 37.4 Ma (Vance, 1988) have also been published.
The most recent dates are those published by Bestland and
Retallack (1994), Bestland et al. (1994a, b) and Retallack et al. (1996).
These are Ar/Ar dates calculated by C. C. Swisher III at the Berkeley
Geochronological Center. They are 39.2 + 0.2 Ma and 39.22 + 0.03 Ma
(at Clarno) and 39.7 + 0.03 Ma and 39.72 + 0.03 (at Painted Hills). Using
the Fish Canyon tuff recalibration developed by the U. S. Geological
Survey for Ar/Ar dates run at the Berkeley Geochronology Center prior
to 1997, the 39.22 Ma date changes to 39.49 Ma, and the 39.72 Ma date
changes to 39.99 Ma.
Hanson (1996) also summarized these ages, and they provide
compelling evidence that the base of the John Day Formation is about 39
to 40 Ma. This means the Hancock quarry is certainly older than 39 Ma
and probably older than 40 Ma. Given that the beginning of Duchesnean
time is about 40 Ma (Prothero and Emry, 1996), the Hancock quarry
must be much older than Chadronian, and older than all but perhaps the
very earliest Duchesnean. With an age older than 39-40 Ma, the radioisotopic evidence suggests Hancock is as old as late Uintan.
Mammalian Biochronology
Careful review of the biochronological distribution of mammal
taxa (except for Rodentia, which has little biochronological significance) from the Hancock quarry local fauna supports assignment of a
late Uintan age (Fig. 3).
Hemipsalodon grandis
Outside of the Hancock quarry, the hyaenodontine creodont
Hemipsalodon grandis is known from the early Chadronian Yoder fauna
of Wyoming (Kihm, 1987) and the early Chadronian Cypress Hills fauna
of Saskatchewan (Russell, 1938). Hemipsalodon sp. is present at the
base of the Chadronian section at Flagstaff Rim, Wyoming (Emry, 1992).
Duchesnean records of Hemipsalodon are from the Porvenir local fauna

91

FIGURE 3. Temporal distribution of mammal taxa of the Hancock quarry local fauna at other North American localities.

of Trans-Pecos Texas (a different species, H. viejensis: Gustafson, 1986),


and from the Duchesnean interval of the Galisteo Formation (H. sp.) in
New Mexico (Lucas, 1982). Hemipsalodon is almost certainly a Eurasian immigrant (together with its close relative Hyaenodon), and outside
of Hancock has a Duchesnean-early Chadronian age range (Mellett, 1969,
1977; Gunnell, 1998).
Nimravidae
A single, incomplete, serrated canine is evidence of the Nimravidae
at the Hancock quarry (Hanson, 1996, p. 224, fig. 9). In North America,
nimravids range in age from Chadronian to Hemphillian and are regarded
as immigrants from Eurasia (e.g., Bryant, 1991; Martin, 1998). The
oldest Asian nimravid is Nimravus mongoliensis from the Ergilian of
Mongolia (Gromova, 1959; Russell and Zhai, 1987)
Haplohippus texanus
Outside of the Hancock quarry, Haplohippus is known only from
the Duchesnean Porvenir local fauna in Trans-Pecos, Texas (McGrew,
1953). MacFadden (1998, p. 543) noted that Haplohippus is unusual
because it is at the approximate same grade of premolarization of
Orohippus, but occurs 10 Ma later in time, with no known morphological intermediates. Actually, by MacFaddens (1998, fig. 37-7) own
diagram, the Orohippus-Haplohippus temporal gap is only about 2 million years, the duration of the late Uintan. This gap shortens if Clarno
Haplohippus is assigned a late Uintan age. Hanson (1996), however,
suggested that Haplohippus may be derived from an Old World equoid
clade, but this has not been documented. Indeed, given the lack of similar
middle-late Eocene equoids in Asia, a North American origin of
Haplohippus seems most likely.

Epihippus gracilis
In North America, Epihippus is a late Bridgerian-Duchesnean taxon
(MacFadden, 1998). E. gracilis, the species from Hancock, is known
from the early-late Uintan of Texas, Utah, Wyoming and Saskatchewan
(e.g., MacFadden, 1980; Eaton, 1985). As Hanson (1996) stressed, the
E. gracilis specimens from Hancock are particularly similar to those
from the latest Uintan Candelaria local fauna in Trans-Pecos, Texas.
Protapirus hancocki
The tapir Protapirus hancocki is one of the most abundant taxa in
the Hancock quarry (Hanson, 1996). It is known elsewhere only from
the late Uintan Candelaria local fauna of Trans-Pecos, Texas (Radinsky,
1963; Wilson and Schiebout, 1984). The genus Protapirus continues on
through much of the Oligocene (Colbert and Schoch, 1998).
Protitanops curryi
A large, horned brontothere is known from the Hancock quarry
from two nearly complete skulls, a complete mandible and various other
maxillary and dental elements, although it has never been fully described
or even clearly taxonomically identified. A preliminary conclusion is
offered here, although a fuller description is warranted.
The Clarno brontothere is a eubrontothere (sensu Schoch and
Lucas, 1985) based on its small, rounded, spherical incisors and
semimolarized upper premolars (i.e., two lingual cusps on P2P4). The
presence of three lower incisors and long upper and lower postcanine
diastemata in the Clarno brontothere rules out typical North American
eubrontotheres from classic Chadronian deposits (e.g., Megacerops
coloradensis sensu Mihlbachler et al., 2004). Likewise, the upper canine

92
diastema, the wider mandibular symphysis, and the absence of the large
parietal dome in the Clarno brontothere rules out the typical Duchesnean
eubrontothere, Duchesneodus.
Other North American eubrontotheres are Protitanops curryi Stock,
1936, from the early Chadronian or late Duchesnean Titus Canyon Formation of California, and Notiotitanops mississippiensis Gazin, 1942
from the Archusa Marl Member of the Cook Mountain Formation of
Mississippi. Notiotitanops mississippiensis has been considered Uintan
or Duchesnean in age due to its similarity with brontothere species from
those land-mammal ages (Krishtalka et al., 1987), but marine mollusc
data suggest it is Bartonian (~ 41 Ma), which means it is late Uintan (e.g.,
Dockery, 1998).
Both of these species share with the Clarno brontothere an upper
postcanine diastema, a plesiomorphic trait that distinguishes them from
more derived eubrontotheres such as Megacerops and Duchesneodus.
However, the Clarno brontothere differs from Notiotitanops
mississippiensis in having a more posteriorly positioned orbit, a much
deeper nasal incision, and more anteriorly positioned posterior nares. To
the extent that we have examined the material, the Clarno brontothere
differs from Protitanops curryi in the size of the horns (not known in
Notiotitanops) and the position of the posterior nares. However, these
characters are variable within some eubrontothere species, and, perhaps
do not warrant separate taxonomic status for the Clarno brontothere
(Mihlbachler et al., 2004). Unfortunately, Notiotitanops mississippiensis
and Protitanops curryi are known only from their holotypes, and nothing is known of the intraspecific variability of these species. However ,
the Clarno brontothere is most similar to Protitanops curryi and could
belong to that species. Hanson (1996) considered the Clarno brontothere
to represent a new species of Protitanops, but did not provide a diagnosis. We are not presently convinced that the differences between the
Clarno brontothere and Protitanops curryi warrant a new species.
The Duchesnean and Chadronian occurrences of eubrontotheres
seem to conflict with our interpretation of the Hancock quarry as latest
Uintan. However, the Clarno brontothere seems plesiomorphic in comparison to typical eubrontotheres, as indicated by the retention of three
lower incisors, an upper and lower postcanine diastema, and relatively
large canines. Although a complete phylogeny of brontotheres is not yet
established, it seems likely that eubrontotheres stemmed from an earlier,
more primitive horned brontothere such as the Uintan Protitanotherium,
or the Asian Irdin Manhan (~Uintan equivalent) genus Protitan. It is,
therefore, not unrealistic to expect early eubrontotheres to occur in latest
Uintan faunas, such as the Hancock quarry assemblage.
Teletaceras radinskyi
The genus Teletaceras is known at the Hancock quarry, as well as
in the late Duchesnean or early Chadronian Titus Canyon local fauna of
California (T. mortivallis of Stock, 1949) and in the Duchesnean Porvenir
local fauna of Trans-Pecos, Texas (T. mortivallis: Wilson and Schiebout,
1984; Prothero, 2004). The species T. radinskyi is endemic to the Hancock
quarry (Hanson, 1989). Teletaceras borissiaki is known in Asia from the
Uglov svita at Artyom (Artm) near Vladivostok in eastern Russia (see
below) (Belyayeva, 1959; Belyayeva et al., 1974; Hanson, 1989).
Procadurcodon sp.
Procadurcodon has no North American records other than the
Hancock quarry. Its only other record is at Artyom in Russia (Gromova,
1960). One of us (SGL) has studied the Hancock quarry Procadurcodon
specimens, and the following conclusions can be offered: (1) the specimens appear to represent the same genus-level taxon described by Gromova
(1960) as Procadurcodon, though this taxon may be a nomen dubium
(see below); (2) the Clarno amynodontid is at the same evolutionary
grade as Zaisanamynodon, an Asian Ergilian taxon, but is distinct as
indicated by Hanson (1996); (3) the Clarno amynodont definitely represents a new species, perhaps a new genus. Thus, the biochronological
significance of the Clarno amynodont is somewhat uncertain.

Heptacodon sp.
Outside of the Hancock quarry, most records of the anthracothere
Heptacodon are of Chadronian-Orellan age (Kron and Manning, 1998).
The oldest record is from the Duchesnean Lac Pelletier fauna of
Saskatchewan (Storer, 1983). Anthracotheres are generally thought to
have immigrated from Eurasia to North America near the end of the
Eocene (Lucas, 1992; Kron and Manning, 1998). Coombs and Coombs
(1977), however, suggested that Apriculus from the late Uintan of Wyoming is an anthracothere, though this has been questioned by Stucky
(1998).
Achaenodon fremdi
Outside of the Hancock quarry, all specimens of Achaenodon are
known from Eocene localities in California, Utah and Wyoming. These
span from the Bridgerian achaenodont, Parahyus vagus Marsh; to the
Uintan achaenodont localities that include the Friars Formation of California (SDSNH 47730; Walsh, 2000), the Washakie Formation, earliest
Uintan Adobetown Member (FMNH.PM 60693; Foss et al., 2001), and
the Uinta B, Wagonhound Member (CM 3182, 3183, 3196) and Uinta C,
Myton Member (CM 3954). Achaenodon is only known from North
America and has not been reported from beds later than Uintan (Gazin,
1955; Walsh, 2000; Foss et al., 2001).
Diplobunops sp.
Outside of the Hancock quarry, the agriochoerid artiodactyl
Diplobunops (Lander, 1998, considers the genus a synonym of
Agriochoerus) is a well known early-late Uintan taxon in Utah, Wyoming and Arkansas (e.g., Peterson, 1919, 1931; Gazin, 1955; Westgate
and Emry, 1985; Lander, 1998). Its youngest record is late Uintan, from
the Brennan Basin Member of the Duchesne River Formation in Utah
(Scott, 1945).
AGE OF THE ARTYOM LOCALITY, EASTERN RUSSIA
The Artyom mammal locality is located in a coal mine north of
Vladivostok in eastern Russia in strata termed the Uglov svita (literally, coal formation) (Russell and Zhai, 1987). Fossil mammals from
this locality are undescribed carnivores (Trofimov, 1953), the brontothere
Rhinotitan orientalis (Yanovskaya, 1957), the rhinocerotoid Teletaceras
borissiaki (Belyayeva, 1959; Hanson, 1989) and the amynodontid
Procadurcodon orientalis (Gromova, 1960). Russell and Zhai (1987, p.
233) regarded the Artyom locality as Sharamurunian in age, based largely
on the stage of evolution of the few mammal taxa. We are less certain of
this, and believe the locality might just as easily be assigned an Ergilian
age.
Rhinotitan orientalis Yanovskaya, 1957, from Artyom is not
Rhinotitan. Indeed, Yanovskaya (1957) explains how the Artyom specimen differs from Rhinotitan but never justifies its assignment to Rhinotitan.
She assumed that it is Rhinotitan because she thought Rhinotitan was the
ancestor of Embolotherium, and the material has some Embolotheriumlike traits. The lower premolars somewhat resemble those of Rhinotitan,
but isolated Rhinotitan premolars cannot easily be differentiated from
several other taxa (Protitanotherium, Diplacodon, Parabrontops etc.).
However, the incisors are small, which rules out Rhinotitan and
Embolotherium grangeri. It could be Parabrontops gobiensis,
Embolotherium andrewsi or Metatitan, so R. orientalis is a nomen dubium. Judging by the fossils, it could be of Irdinmanhan, Sharamurunian
or Ergilian age.
Teletaceras has no other Asian occurrences, but the Artyom specimen (holotype of Eotrigonias borissiaki) has also been assigned to
Forstercooperia or Juxia, two hyracodontid taxa of SharamurunianErgilian age (Belyayeva, 1959; Belyayeva et al., 1974; Lucas et al., 1981;
Lucas and Sobus, 1989). Therefore, a grade level of evolution suggests
either a Sharamurunian or Ergilian age for Teletaceras borissiaki.
Procadurcodon is very similar to and may be a synonym of

Zaisanamynodon (= Gigantamynodon). The holotype material from


Artyom described by Gromova (1960) simply is not diagnostic, and
Procadurcodon thus is a likely nomen dubium (Wall, 1989).
Zaisanamynodon is an Ergilian taxon (Lucas et al., 1996).
Thus, a Sharamurunian or Ergilian age for the Artyom locality
seems certain, but deciding between the two ages is impossible based on
current data. Most workers correlate the Sharamurunian to parts of the
North American Uintan-Duchesnean and the Ergilian to parts of the
North American Duchesnean-Chadronian (e.g., Li and Ting, 1983;
Holroyd and Ciochon, 1994; Tong et al., 1995). A late Uintan age for the
Hancock quarry local fauna makes a Sharamurunian age for the Artyom
locality seem most probable.
DISCUSSION
The Hancock quarry mammal assemblage has been assigned either
a Duchesnean or early Chadronian age (e.g., Emry et al., 1987; Krishtalka
et al., 1987; Lucas, 1992; Hanson, 1996). Consensus on the age assignment was not reached because several of the taxa from the Hancock
quarry are found elsewhere in Uintan, Duchesnean or Chadronian strata
(Fig. 3), making the Hancock quarry assemblage difficult to correlate
biostratigraphically. Furthermore, some of the Hancock quarry taxa
(Procadurcodon, Teletaceras) are known from the Russian Artyom

93
locality of Sharamurunian or Ergilian age.
An important key to evaluating the age of the Hancock quarry
local fauna is to divide its mammal taxa into two groups, those of obvious
North American origin (Protapirus, Epihippus, Haplohippus,
Diplobunops and Achaenodon) and those that are immigrants from Eurasia
that first appear in North American assemblages of Duchesnean age
(Hemipsalodon, Teletaceras, Heptacodon). Except for Haplohippus, the
North American endemics are restricted to Uintan faunas. If a late Uintan
age is assigned to Hancock, then the temporal gap between Orohippus
and Haplohippus shortens.
A late Uintan age for the Hancock quarry extends back the first
appearances in North America of Hemipsalodon, Teletaceras, Heptacodon
and Nimravidae. This indicates some dispersal lag between Hancock and
the other North American localities, or simply a lack of discovery of
these taxa in Uintan faunas outside of Oregon. We conclude that a late
Uintan age for the Hancock quarry mammals best fits all data, radioisotopic and biochronologic.
ACKNOWLEDGMENTS
Bill Orr made it possible to study the fossil documented here.
Comments on the manuscript by Lia Vella, Robert Emry, Peter
Kondrashov and Donald Prothero improved its content.

REFERENCES
Belyayeva, E. I., 1959, Sur la dcouverte de rhinocros tertiaires anciens
dans la Province Maritime de lURSS: Vertebrata PalAsiatica, v. 3, p. 8192.
Belyayeva, E. I., Trofimov, B. A. and Reshetov, V. Yu., 1974, Osnovnyye
etapy evolyutsii mlekopitayushchikh v pozdnem Mezozoye-Paleogene
Tsentralnoy Azii [Basic stages in the evolution of mammals in the late
Mesozoic-Paleogene of Central Asia]: Trudy Sovmestnaya-Mongolskaya
Paleontologicheskaya Ekspeditsiya, v. 1, p. 19-45.
Bestland, E. A. and Retallack, G. J., 1994, Geology and paleoenvironments
of the Painted Hills unit, John Day Fossil Beds National Monument,
Oregon: Final report, NPS Contract CX-9000-1-10009, 203 p.
Bestland, E. A., Retallack, G. J., Fremd, T. and Swisher, C. C. III, 1994a,
Geology and age assessment of late Eocene fossil localities in the Clarno
unit, John Day Fossil Beds National Monument, central Oregon: Geological Society of America, Abstracts with Programs, v. 26, no. 6, p. 4.
Bestland, E. A., Retallack, G. J. and Fremd, T., 1994b, Sequence stratigraphy of the Eocene-Oligocene transition: Examples from the non-marine volcanically influenced John Day basin; in Swanson, D. A. and
Haugerud, R. A., eds., Geologic field trips in the Pacific Northwest:
Seattle, University of Washington, p. 11-19.
Bryant, H. N., 1991, Phylogenetic relationships and systematics of the
Nimravidae (Carnivora): Journal of Mammalogy, v. 72, p. 56-78.
Colbert, E. H., 1938, Brachyhyops, a new bunodont artiodactyl from Beaver Divide, Wyoming: Annals of the Carnegie Museum, v. 27, p. 87108.
Colbert, M. W. and Schoch, R. M., 1998, Tapiroidea and other
moropomorphs; in Janis, C. N., Scott, K. M. and Jacobs, L. L., eds.,
Evolution of Tertiary mammals of North America: Cambridge, Cambridge University Press, p. 569-582.
Coombs, W. P., Jr. and Coombs, M. C., 1977, The origin of the anthracotheres:
Neues Jahrbuch fur Geologie und Palaontologie, Monatshefte 1977: 8489.
Cope, E. D., 1885, The Vertebrata of the Tertiary formations of the West.
Book 1: Report of the U. S. Geological Survey of the Territories [Hayden],
v. 3, 1009 p.
Dalrymple, G. B., 1979, Critical tables for conversion of K-Ar ages from
old to new constants: Geology, v. 7, p. 558-560.
Dockery, D. T. III, 1998, Molluscan faunas across the Paleocene/Eocene
Series boundary in the North American Gulf Coastal plain; in Aubry, MP., Lucas, S. G. and Berggren, W. A., eds., Late Paleocene-early Eocene

climatic and biotic events in the marine and terrestrial records: New
York, Columbia University Press, p. 296-322.
Eaton, J. G., 1985, Paleontology and correlation of the Eocene Tepee Trail
and Wiggins formations in the North Fork of Owl Creek area, southeastern Absaroka Range, Hot Springs County, Wyoming: Journal of Vertebrate Paleontology, v. 5, p. 345-370.
Effinger, J. A., 1998, Entelodontidae; in Janis, C. M., Scott, K. M. and
Jacobs, L. L., eds., Evolution of Tertiary mammals of North America.
Volume 1: Terrestrial carnivores, ungulates and ungulatelike mammals:
Cambridge, Cambridge University Press, p. 375-398.
Emry, R. J., 1992, Mammalian range zones in the Chadronian White River
Formation at Flagstaff Rim, Wyoming; in Prothero, D. R. and Berggren,
W. A., eds., Eocene-Oligocene climatic and biotic evolution: Princeton,
Princeton University Press, p. 106-115.
Emry, R. J., Bjork, P. R. and Russell, L. S., 1987, The Chadronian, Orellan,
and Whitneyan land mammal ages; in Woodburne, M. O., ed., Cenozoic
mammals of North America, geochronology and biostratigraphy: Berkeley, University of California Press, p. 118-152.
Evernden, J. F. and James, G. T., 1964, Potassium-argon dates and the
Tertiary floras of North America: American Journal of Science, v. 262,
p. 945-974.
Foss, S. E., 2001, Systematics and paleobiology of the Entelodontidae
(Mammalia, Artiodactyla) [Ph.D. dissertation]: Dekalb, Northern Illinois University, 222 p.
Foss, S. E., Turnbull, W. D. and Barber, L., 2001, Observations on a new
specimen of Achaenodon (Mammlia, Artiodactyla) from the Eocene
Washakie Formation of southern Wyoming: Journal of Vertebrate Paleontology, v. 21, supplement to no. 3, p. 51A.
Gazin, C. L., 1942, A new titanothere from the Eocene of Mississippi with
notes on the correlation between the marine Eocene of the Gulf Coastal
Plain and continental Eocene of the Rocky Mountain region: Smithsonian
Miscellaneous Collections, v. 101, no. 13, 13 p.
Gazin, C. L., 1955, A review of the upper Eocene Artiodactyla of North
America: Smithsonian Miscellaneous Collections, v. 128, no. 8, 96 p.
Gromova, V, 1959, Premire dcouverte dun chat primitif au Palogene
dAsie Centrale: Vertebrata PalAsiatica, v. 3, p. 59-72.
Gromova, V., 1960, Pervaya nakhodka v Sovetskom Soyuze amynodonta
(novyy rod Procadurcodon) [The first discovery of amynodonts in the
Soviet Union (new genus Procadurcodon)]: Akadaemiya Nauk SSSR,
Trudy Paleontologicheskovo Instituta, v. 77, p. 128-151.

94
Gunnell, G. F., 1998, Creodonta; in Janis, C. N., Scott, K. M. and Jacobs, L.
L., eds., Evolution of Tertiary mammals of North America: Cambridge,
Cambridge University Press, p. 91-105.
Gustafson, E., 1986, Carnivorous mammals the late Eocene and early Oligocene of Trans-Pecos Texas: Texas Memorial Museum, Bulletin 33, 66
p.
Hanson, C. B., 1989, Teletaceras radinskyi, a new primitive rhinocerotid
from the late Eocene Clarno Formation, Oregon; in Prothero, D. R. and
Schoch, R. M., eds., The Evolution of Perissodactyls; Oxford University Press, New York, p. 379-398.
Hanson, C. B., 1996, Stratigraphy and vertebrate faunas of the BridgerianDuchesnean Clarno Formation, north-central Oregon; in Prothero, D.
R. and Emry, R. J., eds., The terrestrial Eocene-Oligocene transition in
North America: Cambridge, Cambridge University Press, p. 206-239.
Holroyd, P. A. and Ciochon, R. L., 1994, Relative ages of Eocene primatebearing deposits of Asia; in Fleagle, J. G. and Kay, R. F., eds., Anthropoid
origins: New York, Plenum Press, p. 123-141.
Kihm, A. J., 1987, Mammalian paleontology and geology of the Yoder
Member, Chadron Formation, east-central Wyoming: Dakoterra, v. 3,
p. 28-45.
Krishtalka, L., Stucky, R. K., West, R. M., McKenna, M. C., Black, C. C.,
Bown, T. M., Dawson, M. R., Golz, D. J., Flynn, J. A., Lilleghraven, J. A.
and Turnbull, W. D., 1987, Eocene (Wasatchian through Duchesnean)
biochronology of North America; in Woodburne, M. O., ed., Cenozoic
mammals of North America, geochronology and biostratigraphy: Berkeley, University of California Press, p. 77-117.
Kron, D. G. and Manning, E., 1998, Anthracotheriidae; in Janis, C. N.,
Scott, K. M. and Jacobs, L. L., eds., Evolution of Tertiary mammals of
North America: Cambridge, Cambridge University Press, p. 381-388.
Lander, B., 1998, Oreodontoidea; in Janis, C. N., Scott, K. M. and Jacobs,
L. L., eds., Evolution of Tertiary mammals of North America: Cambridge, Cambridge University Press, p. 402-420.
Li, C. and Ting, S., 1983, The Paleogene mammals of China: Bulletin of the
Carnegie Museum of Natural History, v. 21, p. 1-93.
Lucas, S. G., 1982, Vertebrate paleontology, stratigraphy, and biostratigraphy of Eocene Galisteo Formation, north-central New Mexico: New
Mexico Bureau of Mines and Mineral Resources Circular 186, 34 p.
Lucas, S. G. 1992. Redefinition of the Duchesnean land mammal age, late
Eocene of western North America; in Prothero, D. R. and Berggren, W.
A., eds., Eocene-Oligocene climatic and biotic evolution: Princeton,
Princeton University Press, p. 88-105
Lucas, S. G. and Sobus, J. C., 1989, The systematics of indricotheres; in
Prothero, D. R. and Schoch, R. M., eds., The Evolution of Perissodactyls; Oxford University Press, New York, p. 358-378.
Lucas, S. G., Emry, R. J. and Bayshashov, B. A., 1996, Zaisanamynodon, a
late Eocene amynodontid (Mammalia, Perissodactyla) from Kazakhstan
and China: Tertiary Research, v. 17, p. 51-58
Lucas, S. G., Schoch, R. M. and Manning, E., 1981, The systematics of
Forstercooperia, a middle to late Eocene hyracodontid (Perissodactyla:
Rhinocerotoidea) from Asia and western North America: Journal of
Paleontology, v. 55, p. 826-841.
MacFadden, B. J., 1980, Eocene perissodactyls from the type section of the
Tepee Trail Formation of northwestern Wyoming: Contributions to
Geology, University of Wyoming, v. 18, p. 135-143.
MacFadden, B. J., 1998, Equidae; in Janis, C. N., Scott, K. M. and Jacobs, L.
L., eds., Evolution of Tertiary mammals of North America: Cambridge,
Cambridge University Press, p. 537-559.
Martin, L. D., 1998, Nimravidae; in Janis, C. N., Scott, K. M. and Jacobs,
L. L., eds., Evolution of Tertiary mammals of North America: Cambridge, Cambridge University Press, p. 228-235.
McGrew, P. O., 1953, A new and primitive early Oligocene horse from
Trans-Pecos, Texas: Fieldiana Geology, v. 10, p. 161-171.
Mellett, J. S., 1969, A skull of Hemipsalodon (Mammalia, Deltatheridia)
from the Clarno Formation of Oregon: American Museum Novitates,
no. 2387, 19 p.
Mellett, J. S., 1977, Paleobiology of North American Hyaenodon (Mammalia, Creodonta): Contributions to Vertebrate Evolution 1, 134 p.
Mihlbachler, M. C., Lucas, S. G. and Emry, R. J., 2004, The holotype of

Menodus giganteus, and the insoluble problem of Chadronian


brontothere taxonomy: New Mexico Museum of Natural History and
Science, Bulletin 26.
Peterson, O. A., 1919, Report upon the material discovered in the upper
Eocene of Uinta basin by Earl Douglass in the years 1908-1909, and by
O. A. Peterson in 1912: Annals of the Carnegie Museum, v. 12, p. 40168.
Peterson, O. A., 1931, New species from the Oligocene of the Uinta:
Annals of the Carnegie Museum, v. 21, p. 51-78.
Prothero, D. R., 2004, The evolution of North American rhinoceroses:
Cambridge, Cambridge University Press, in press.
Prothero, D. R. and Emry, R. J., 1996, Summary; in Prothero, D. R. and
Emry, R. J., eds., The terrestrial Eocene-Oligocene transition in North
America: Cambridge, Cambridge University Press, p. 666-683.
Radinsky, L. B., 1963, Origin and early evolution of North American
Tapiroidea: Peabody Museum of Natural History, Yale University, Bulletin 17, 106 p.
Retallack, G., Bestland, E. A. and Fremd, T., 1996, Reconstructions of
Eocene and oligocene plants and animals of central Oregon: Oregon
Geology, v. 58, p. 51-69.
Russell, D. E. and Zhai, R., 1987, The Paleogene of Asia: Mammals and
stratigraphy: Mmoires du Musum National dHistoire Naturelle Srie
C, v. 52, 488 p.
Russell, L. S., 1938, The skull of Hemipsalopdon grandis, a giant Oligocene
creodont: Transactions of the Royal Society of Canada, Series 3, v. 32,
p. 61-66.
Russell, L. S., 1980. A new species of Brachyhyops? (Mammalia, Artiodactyla) from the Oligocene Cypress Hills Formation of Saskatehewan:
ROM Life Sciences Occasional Paper No. 33, 5 p.
Schoch, R.M. and S. G. Lucas 1985, The Brontotheriidae, a group of Eocene
and Oligocene perissodactyls from North America, Asia and Eastern
Europe: Abstracts of Papers and Posters, Fourth International
Theriological Congress, Abstract 0558.
Scott, W. B., 1945, The Mammalia of the Duchesne River Oligocene:
Transactions of the American Philosophical Society, v. 34, p. 209-253.
Stock, C., 1936, Titanotheres from the Titus Canyon Formation, California: Proceedings of the National Academy of Sciences: 22: 656661.
Stock, C., 1949, The mammalian fauna from the Titus Canyon Formation,
California: Publications Carnegie Institution of Washington, no. 584, p.
229-244.
Storer, J. E., 1983, A new species of the artiodactyl Heptacodon from the
Cypress Hills Formation, Lac Pelletier, Saskatchewan: Canadian Journal
of Earth Sciences, v. 20, p. 1344-1347.
Stucky, R. K., 1998, Eocene bunodont and bunoselenodont Artiodactyla
(dichobunids); in Janis, C. N., Scott, K. M. and Jacobs, L. L., eds.,
Evolution of Tertiary mammals of North America: Cambridge, Cambridge University Press, p. 358-374.
Swanson, D. R. and Robinson, P. T., 1968, Base of the John Day Formation
in and near Horse Heaven mining district, north-central Oregon: U. S.
Geological Survey, Professional Paper 600-D, p. 154-161.
Tong, Y., Zheng, S. and Qiu, Z., 1995, Cenozoic mammal ages of China:
Vertebrata PalAsiatica, v. 33, p. 290-314.
Trofimov, B. A., 1953, Drevnetretichnye mlekopitayushchiye na Dalnem
Vostoke SSSR [Ancient Tertiary mammals in the far east of the USSR]:
Priroda, v. 12, p. 111-112.
Vance, J. A., 1988, New fission track and K-Ar ages from the Clarno
Formation, Challis age volcanic rocks in north central Oregon: Geological Society of America, Abstracts with Programs, v. 20, no. 6, p. 473.
Wall, W. P., 1989, The phylogenetic history and adaptive radiation of the
Amynodontidae; in Prothero, D. R. and Schoch, R. M., eds., The evolution of Perissodactyls; Oxford University Press, New York, p. 340-354.
Walsh, S. L., 2000, Bunodont artiodactyls (Mammalia) from the Uintan
(middle Eocene) of San Diego County, California: Proceedings of the
San Diego Society of Natural History, no. 37, 27 p.
Westgate, J. W. and Emry, R. J., 1985, Land mammals of the Crow Creek
local fauna, late Eocene, Jackson Group, St. Francis County, Arkansas:
Journal of Paleontology, v. 59, p. 242-248.
Wilson, J. A., 1971, Early Tertiary Vertebrate faunas, Vieja Group, Trans-

95
Pecos Texas: Entelodontidae: The Pearce-Sellards Series, 17, 17 p.
Wilson, J. A. and Schiebout, J. A., 1984, Early Tertiary vertebrate faunas,
Trans-Pecos Texas: Ceratomorpha less Amynodontidae: The Pearce
Sellards Series, no. 39, 47 p.

Yanovskaya, N. M., 1957, Brontoterii Mongolii [Brontotheres of Mongolia]:


Trudy Sovmestnaya-Mongolskaya Paleontologicheskaya Ekspeditsiya,
v. 12, 219 p.

96

Haploconus angustus, front of skull and jaw (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

97

THE ENTELODONT BRACHYHYOPS (MAMMALIA, ARTIODACTYLA)


FROM THE UPPER EOCENE OF FLAGSTAFF RIM, WYOMING
SPENCER G. LUCAS1 AND ROBERT J. EMRY2
1

New Mexico Museum of Natural History, 1801 Mountain Rd. NW, Albuquerque, NM 87104; 2Department of Paleobiology, National Museum of Natural History,
Smithsonian Institution, Washington, D. C. 20560

AbstractSpecimens of the entelodont Brachyhyops from the early Chadronian of Flagstaff Rim, Wyoming,
document new morphological characteristics of the upper dentition and cranial anatomy of the genus. Brachyhyops
fossils are present in Saskatchewan, Montana(?), Wyoming, Utah, New Mexico and Texas in strata of late
Duchesnean-early Chadronian age. The genus first appeared in North America as an immigrant from Asia,
where it occurs in Irdinmanhan-Ergilian (middle-late Eocene) strata in China, Kazakstan and Mongolia.
Keywords: Brachyhyops, Artiodactyla, Entelodontidae, Wyoming, Flagstaff Rim, Eocene
INTRODUCTION
Strata of the White River Formation exposed at Flagstaff Rim in
central Wyoming (Fig. 1) yield an extensive record of Chadronian (late
Eocene) fossil mammals directly tied to magnetostratigraphy and a series of radioisotopic ages between ~35 and 32 Ma (Emry, 1973, 1992;
Emry et al., 1987; Prothero, 1996a). This section (Fig. 1) thus is critical to understanding Chadronian mammal biostratigraphy, biochronology
and correlation. Emry (1992) plotted the stratigraphic ranges of the
mammalian taxa in the Flagstaff Rim section, and indicated the presence of two entelodont taxa. Brachyhyops records are low in the section, 20-60 ft above its base, whereas Archaeotherium occurs higher in
the section, 220-480 ft above its base (Fig. 1). Here, we document the
Brachyhyops specimens from Flagstaff Rim and summarize the geographic and stratigraphic distribution of the genus.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; CM = Carnegie Museum of Natural History, Pittsburgh; USNM = National Museum of Natural History,
Smithsonian Institution, Washington, D. C.
DESCRIPTION AND MORPHOLOGICAL SIGNIFICANCE
Two specimens of Brachyhyops are known from the Flagstaff
Rim section, and they provide new morphological data on this rare
taxon. USNM 521245 is a left maxillary fragment with the posterior
end of the canine alveolus, P1 alveolus, P2-3 and DP4 (Fig 2D-E). The
specimen has no teeth in common, and thus little basis for direct comparison, with the holotype of B. wyomingensis Colbert, 1938, which is
most of a skull (CM 12048) that includes P1, but which has no other
premolars preserved except for part of one broken and very worn P4.
However, the measurement from the posterior edge of the canine alveolus to the posterior edge of DP4 is 78 mm in USNM 521245, whereas
the same measurement in B. wyomingensis (but to P4 rather than DP4)
is 62 mm. Thus, based on this one comparable measurement, USNM
521245 is approximately 25% larger than the B. wyomingensis type.
Further, USNM 521245 is of the appropriate size to occlude almost
perfectly with a cast of the holotype lower jaw of B. viensis Russell,
1980, the species from the Cypress Hills Formation of Saskatchewan.
Russell (1980a, p. 3) calculated that B. viensis is about 30% larger
than B. wyomingensis, based on comparison of the type of B. viensis to
a mandible from Texas referred by Wilson (1971) to B. wyomingensis.
We conclude that USNM 521245 is the right size to be B. viensis, and
we refer it to the species on that basis.
The maxilla is concave laterally and has an antero-posteriorly
oriented cavity (sinus) medially that extends back to over the P3. There
is a large infraorbital foramen above the anterior edge of P4. The C
alveolus is larger than the double alveolus of P1, which is at an angle to
the long axis of the tooth row. A short alveolus follows P1, and the

FIGURE 1. Location map and stratigraphic section at Flagstaff Rim, Wyoming,


showing stratigraphic distribution of entelodont taxa (after Emry, 1992).

trenchant P2 follows. This tooth has not previously been described in


Brachyhyops. It is antero-posteriorly long, transversely narrow, and two
rooted. The single cusp is a tall cone with a distinct posterior ridge that
runs onto a very low, narrow talon heel (with a posterior cingulum).
The enamel is slightly rugose and lineated. Measurements (in mm)
are: length = 15.2, width = 5.7
The P3 is not complete but has two large, pointed labial cusps,

98

FIGURE 2. Brachyhyops from Flagstaff Rim, Wyoming. A, USNM 521246, incomplete skull roof of B. viensis in lateral (A), dorsal (B) and ventral (C) views. D-E,
USNM 521245, left maxillary fragment with C and P1 alveoli and P2-3, DP4, lateral (E) and occlusal (F) views.

the paracone and metacone, and only a small, low, very narrow labial
cingulum, better developed labial to the metacone than the paracone. A
parastyle is present but no better developed than the cingulum. Low
ridges run between the paracone and metacone, creating an incomplete
ectoloph-like crest. The crown lingually is only a small lobe between
the paracone-metacome upon which presumably a low, bulbous protocone was located. Lingual to the metacone, the posterior cingulum is
taller and better developed than the labial cingulum. Length (in mm) =
18.6. Note that Colbert (1938, fig. 3) reconstructed P3 incorrectly as a
single-coned, trenchant tooth.
The DP4 is a bunodont tooth with a nearly triangular outline. It
has a large paracone and metacone labially and prominent cingula anteriorly and posteriorly that essentially surround the crown. The protocone is the largest lingual cusp, and is followed by a somewhat smaller
hypocone on the postero-lingual cingulum. The metaconule and
paraconule are slightly larger than the hypocone. Measurements (in
mm) are: length = 18.8, width = 18.2.
USNM 521246 (Fig. 2A-C) is a skull roof from the anterodorsal
margin of the orbits (left side) to the occipital (nuchal) crest, the posterior parts of both zygomatic arches (not connected to the brain case)

with glenoid fossae, and other fragments of basicranial elements. It


represents an individual substantially larger than CM 12048, the holotype of B. wyomingensis: for example, the width across the frontals
above the dorsal edges of the orbits (79 mm in CM 12048; 101 mm in
USNM 521246) is about 28% larger, and the distance from the posterior edge of the orbit to the posterior end of the nuchal crest (108 mm in
CM 12049; 152 mm in USNM 521246) is about 41% larger (the greater
degree of difference in this latter measurement is due in part to the
exaggerated posterolateral wings of the nuchal crest in USNM 521246).
The size of USNM 521246 is in the range expected for B. viensis,
based on the size of the lower jaw, and we refer the specimen to that
species.
In dorsal view, the roof is broadest across the orbits and tapers
slowly back to the postorbital constriction, flares again slightly to taper
more sharply back to the posterior tip of the braincase, then flares out
again slightly with parasagittal crests that are sinusoidal (cf. Colbert,
1938, fig. 2). The top of the skull has three sets of long ridges: (1) a
medial ridge that runs with slight irregularity from frontal to occiput;
(2) two ridges laterally that are lower, broader and converge to meet
over the braincase constriction; and (3) the parasagittal ridges that de-

99
fine the outer margin of the dorsal surface of the skull (except for the
orbit, which projects laterally from them). Between the ridges the bone
is textured into irregular rugose lines and pits, similar to the texture of
crocodilian dermal bone. Parts of the glenoid portion of the zygomata
are preserved on each side. These have a wide, shallowly concave glenoid fossa with a short, blunt post-glenoid process medially. The anterior wall of the glenoid roof is a broad plate that is slightly concave
anteriorly. There is a prominent process on the dorso-lateral margin.
Ventrally, the frontal is invested with irregular, concave pockets
and grooves (sinuses). The orbital wall is smooth, with a complete
postorbital bar; the braincase dorsal roof is a broad, roughened cup for
the cerebral hemispheres with a distinct, fairly digitate frontal-parietal
suture that cuts across it transversely; also, there is an irregular pattern
of foramina on this surface. A convex ridge cuts off the anterior from
the posterior concavity for the occipital lobe, which is antero-posteriorly short and bulbous.
Laterally, the skull roof rises behind the orbits, and there is a
concave depression between the braincase wall and overhanging
parasagittal crests. The bone is generally smooth on the lateral aspect
of the skull. The frontal-parietal suture continues down the lateral side
of the skull. In occipital view, the occiput is high and narrow, and deeply
concave between backward projecting nuchal crests.
Archaeotherium coarctatum has a broad skull roof across the
orbits that tapers rapidly posterior to the orbits to form very narrowly
spaced parasagittal crests above the braincase (e.g., Peterson, 1909).
Brachyhyops, in contrast, has much broader parasagittal crests above
the braincase. The textured, lineated braincase roof of Brachyhyops is
most similar to the condition in Archaeotherium coarctatum, and it
may be a derived feature uniting Brachyhyops to an Archaeotherium
coarctatum clade. This pattern here differs from the holotype skull of
B. wyomingensis, so it may be unique to B. viensis or it may just be
variable among Brachyhyops.
DISTRIBUTION OF BRACHYHYOPS
Brachyhyops, the oldest North American entelodont, is an immigrant from Asia (e.g., Brunet, 1979; Lucas and Emry, 1996; Foss,
2001). In Asia, Brachyhyops (= Eoentelodon) has a distribution in China,
Kazakstan and Mongolia in strata of Irdinmanhan-Ergilian (middlelate Eocene) age (Russell and Zhai, 1987; Wang and Qiu, 2002).
Brachyhyops fossils have a broad distribution in western North America,
from Sakatchewan to the Big Bend of Texas in strata of late Duchesneanearly Chadronian age (Fig. 3). These records are:
1. Southfork local fauna in the Cypress Hills Plateau of
Saskatchewan (Russell, 1980, a, b; Storer, 1984, 1996). This is the
type locality of Brachyhyops viensis Russell. Storer (1984, 1996) regarded the Southfork local fauna as late Duchesnean, but the taxa present
include species of Hyaenodon, Mesohippus, Hyracodon and
Leptomeryx that suggest an early Chadronian age.
2. Ostrander (1985, p. 213) lists the genus Brachyhyops without
documentation from the early Chadronian Prickly Pear Creek local fauna
in Montana. A. Tabrum (written commun., 2004) informs us that Prickly
Pear Creek is a drainage east of Helena and west of Canyon Ferry
where the White River deposits yielded a small collection of mammal fossils made by Earl Douglass in 1902 and housed at the Carnegie
Museum of Natural History. Tabrum says the Carnegie collection is
mostly fragmentary but includes material of Colodon, Menodus and
Oreonetes, but no specimens of Brachyhyops. The basis for Ostranders
(1985) report of Brachyhyops may be in another collection and thus
remains undocumented.
3. The Flagstaff Rim records documented here, which are early
Chadronian (Fig. 1; Emry, 1992).
4. The type locality of Brachyhyops wyomingensis, the early
Chadronian (see Lucas et al., this volume) Big Sand Draw Sandstone
Lentil of the White River Formation at Beaver Divide, Wyoming
(Colbert, 1938; Emry, 1975).

FIGURE 3. Geographic and temporal distribution of Brachyhyops in North America.


Localities are: B = Baca Formation, west-central New Mexico; BD = Beaver Divide,
Wyoming; CC = Canyon Creek, Wyoming; F = Flagstaff Rim, Wyoming; G =
Galisteo Formation, north-central New Mexico; L = Lapoint fauna, Utah; P = Porvenir
local fauna, Trans-Pecos, Texas; Pr = Prickly Pear Creek local fauna, Montana; S =
Southfork local fauna, Saskatchewan.

5. Lower Chadronian strata of the White River Formation at Canyon Creek, Wyoming, about 50 km east of Beaver Divide. These are skull
fragments in the USNM collection (including USNM 455679) that match
well the cranial material of B. viensis from Flagstaff Rim described here.
6. The Lapoint Member of the Duchesne River Formation, in
northeastern Utah, which is the type fauna of the Duchesnean landmammal age (Lucas, 1992). Brachyhyops wyomingensis from Lapoint
(Gazin, 1956; Wilson, 1971; Emry, 1981) is stratigraphically above an
ash with an Ar/Ar age of 39.74 + 0.07 Ma and probably correlates to
part of Chron 18n (Prothero and Emry, 1996). The Lapoint mammals
suggest, however, a younger, late Duchesnean age, perhaps equivalent
to Chrons 17r or 17n (Prothero and Emry, 1996).
7. Duchesnean interval of the Galisteo Formation in north-central New Mexico (Lucas and Estep, 1999). The Brachyhyops specimen, a m1 assigned to B. viensis, occurs at the Duchesneodus quarry
(Stearns quarry) near the top of the Galisteo Formation. This stratigraphic level is in an interval of reversed magnetic polarity that Prothero
and Lucas (1996) correlated to Chron 17r. The base of the Espinaso
Formation, about 10 m above the Duchesneodus quarry, is in a normal
polarity interval correlated to Chron C17n3 and yields radioisotopic
ages of ~38 Ma. Thus, the Brachyhyops specimen from the Galisteo
Formation is of late Duchesnean age.
8. Duchesnean interval of the Baca Formation at Mariano Mesa
in west-central New Mexico. Originally reported by Schrodt (1980)
and Schiebout and Schrodt (1981), the specimens are an edentulous
mandible fragment (Louisiana State University Museum of Geosciences
[LSUMG] 2313-1) and a right p4-m3 (LSUMG 2313-2) of Brachyhyops
wyomingensis from the upper part of the Baca Formation. The latter,
illustrated by Schrodt (1980, apppendix 3, fig. 1) is the correct morphology and size (measurements in mm are: m1l = 16.2, w = 11.6, m2l
= 17.7, w = 13.9, m3l = 18.6, w = 12.0) to be assigned to B.
wyomingensis. Recent paleomagnetic stratigraphy indicates that the
Brachyhyops locality is in an interval of reversed polarity that is immediately below the base of the Spears Formation, which yields radioisotopic ages of ~ 38 Ma. Thus, the Brachyhyops locality is best correlated to Chron 17r (Ludtke et al., 2003), and it is approximately coeval
with the Galisteo Formation record of Brachyhyops.

100
9. Late Duchesnean Porvenir local fauna in Trans-Pecos, Texas,
where Brachyhyops wyomingensis is well documented (Wilson, 1971).
The Porvenir local fauna is late Duchesnean; it is above an ash with an
Ar/Ar age of 37.8 + 0.15 Ma and in an interval of normal polarity
equivalent to Chron 17n2 (Prothero, 1996b; Prothero and Emry, 1996).

ACKNOWLEDGMENTS
J. Schiebout and M. Dawson provided access to specimens in
their care. A. Tabrum provided helpful information on the Prickly Pear
Creek local fauna We thank P. Kondrashov and S. Foss for comments
on the manuscript.

REFERENCES
Brunet, M., 1979, Les grands mammifres chefs de file de limmigration Oligocne
et le problme de la limite ocne-Oligocne en Europe. Paris, ditions de la
Fondation Singer-Polignac, 281 p.
Colbert, E. H., 1938, Brachyhyops, a new bunodont artiodactyl from Beaver Divide, Wyoming: Annals of the Carnegie Museum, v. 27, p. 87-108.
Emry, R. J., 1973, Stratigraphy and preliminary biostratigraphy of the Flagstaff
Rim area, Natrona County, Wyoming: Smithsonian Contributions to
Paleobiology, no. 18, 42 p.
Emry, R. J., 1975, Revised Tertiary stratigraphy and paleontology of the western
Beaver Divide, Fremont County, Wyoming: Smithsonian Contributions to
Paleobiology, no. 25, 20 p.
Emry, R. J., 1981, Additions to the mammalian fauna of the type Duchesnean, with
comments on the status of the Duchesnean age: Journal of Paleontology, v.
55, p. 563-570.
Emry, R. J., 1992, Mammalian range zones in the Chadronian White River Formation at Flagstaff Rim, Wyoming; in Prothero, D. R. and Berggren, W. A., eds.,
Eocene-Oligocene climatic and biotic evolution: Princeton, Princeton University Press, p. 106-115.
Emry, R. J., Russell, L. S. and Bjork, P. R., 1987, Chadronian, Orellan, and
Whitneyan North American land mammal ages; in Woodburne, M. O., ed., Cenozoic mammals of North America. Geochronology and biostratigraphy: Berkeley, University of California Press, p. 118-152.
Foss, S. E., 2001, Systematics and paleobiology of the Entelodontidae (Mammalia,
Artiodactyla) [Ph.D. dissertation]: Dekalb, Northern Illinois University, 222 p.
Gazin, C. L., 1956, The geology and vertebrate paleontology of upper Eocene strata
in the northeastern part of the Wind River basin, Wyoming, Part 2. The mammalian fauna of the Badwater area: Smithsonian Miscellaneous Collections 131,
35 p.
Lucas, S. G. 1992. Redefinition of the Duchesnean land mammal age, late Eocene
of western North America; in Prothero, D. R. and Berggren, W. A., eds., EoceneOligocene climatic and biotic evolution: Princeton, Princeton University Press,
p. 88-105.
Lucas, S. G. and Emry, R. J., 1996, Late Eocene entelodonts (Mammalia: Artiodactyla) from Inner Mongolia, China: Proceedings of the Biological Society of
Washington, v. 109, p. 397-405.
Lucas, S. G. and Estep, J. W., 1999, Brachyhyops (Mammalia, Artiodactyla) from
the Galisteo Formation and its biochronological significance: New Mexico Geological Society, Guidebook 50, p. 25-26.
Ludtke, J., Prothero, D. R. and Lucas, S. G., 2003, Magnetic stratigraphy of the
upper middle Eocene Baca Formation, west-central New Mexico: Geological
Society of America, Abstracts with Programs, v. 35, no. 6, p. 427.
Ostrander, G. E., 1985, Correlation of the early Oligocene (Chadronian) in north-

western Nebraska: Dakoterra, v. 2, p. 205-231.


Peterson, O. A., 1909, A revision of the Entelodontidae: Memoirs of the Carnegie
Museum 4: 41-158.
Prothero, D. R., 1996a, Magnetostratigraphy of the Eocene-Oligocene transition in
Texas; in Prothero, D. R. and Emry, R. J., eds., The terrestrial Eocene-Oligocene
transition in North America: Cambridge, Cambridge University Press, p. 189198.
Prothero, D. R., 1996b, Magnetic stratigraphy of the White River Group in the
High Plains; in Prothero, D. R. and Emry, R. J., eds., The terrestrial EoceneOligocene transition in North America: Cambridge, Cambridge University Press,
p.262-277.
Prothero, D. R. and Emry, R. J., 1996, Summary; in Prothero, D. R. and Emry, R. J.,
eds., The terrestrial Eocene-Oligocene transition in North America: Cambridge,
Cambridge University Press, p. 664-683.
Prothero, D. R. and Lucas, S. G., 1996, Magnetic stratigraphy of the Duchesnean
part of the Galisteo Formation, New Mexico; in Prothero, D. R. and Emry, R. J.,
eds., The terrestrial Eocene-Oligocene transition in North America: Cambridge,
Cambridge University Press, p. 199-205.
Russell, D.E. and Zhai, R.J., 1987, The Paleogene of Asia: Mammals and stratigraphy: Memoires Museum National dHistoire Naturelle, v. 52, 488 p.
Russell, L. S., 1980a, Tertiary mammals of Saskatchewan Part V: The Oligocene
entelodonts: ROM Life Sciences Contributions 122, 42 p.
Russell, L. S., 1980b, A new species of Brachyhyops? (Mammalia, Artiodactyla)
from the Oligocene Cypress Hills Formation of Saskatehewan: ROM Life Sciences Occasional Paper 33, 5 p.
Schiebout, J. A. and Schrodt, A. K., 1981, Vertebrate paleontology of the lower
Tertiary Baca Formation of western New Mexico: Geological Society of America
Bulletin, v. 92, p. 976-979.
Schrodt, A. K., 1980, Depositional environments, provenance, and vertebrate paleontology of the Eocene-Oligocene Baca Formation, Catron County, New Mexico
[M. S. thesis]: Baton Rouge, Louisiana State University, 174 p.
Storer, J. E., 1984, Fossil mammals of the Southfork local fauna (early Chadronian)
of Saskatchewan: Canadian Journal of earth Sciences, v. 21, p. 1400-1405.
Storer, J. E., 1996, Eocene-Oligocene faunas of the Cypress Hills Formation,
Saskatchewan; in Prothero, D. R. and Emry, R. J., eds., The terrestrial EoceneOligocene transition in North America: Cambridge, Cambridge University Press,
p. 240-261.
Wang, B. and Qiu, Z., 2002, A new species of Entelodontidae (Artiodactyla, Mammalia) from late Eocene of Nei Mongol, China: Vertebrata PalAsiatica, v. 40, p.
194-202.
Wilson, J. A., 1971, Early Tertiary vertebrate faunas, Vieja Group, Trans-Pecos
Texas: Entelodontidae: The Pearce-Sellards Series, 17, 17 pp.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

113

LATE EOCENE BRONTOTHERES (MAMMALIA, PERISSODACTYLA) FROM BEAVER


DIVIDE, WYOMING, AND THEIR BIOCHRONOLOGICAL SIGNIFICANCE
SPENCER G. LUCAS1, ROBERT J. EMRY2 AND MATTHEW C. MIHLBACHLER3
1
New Mexico Museum of Natural History, 1801 Mountain Rd. NW, Albuquerque, NM 87104;
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, D. C. 20560;
3
Division of Paleontology, American Museum of Natural History, Central Park West at 79th St., New York, NY 10024

AbstractWe document lower jaws of brontotheres from the Big Sand Draw Sandstone Lentil of the White
River Formation at Beaver Divide in Fremont County, Wyoming and redescribe a brontothere skull collected
much earlier from the same locality. The jaws do not belong to typical Chadronian brontotheres such as Menops,
Brontops, or Megacerops (sensu Mader, 1989 = Megacerops coloradensis sensu Mihlbachler et al., this volume) but could belong to Duchesneodus primitivum (Lambe) or Protitanops curryi (Stock). The skull most
likely belongs to Menodus heloceras (=Megacerops coloradensis in part sensu Mihlbachler et al., this volume)
but we cannot completely rule out the possibility that the skull belongs to the same species to which the mandibles belong. Although the specific taxonomic identity of these specimens is not certain, the brontothere material and the other mammals from the Big Sand Draw Sandstone Lentil suggest an early Chadronian age.
Keywords: Eocene, brontothere, Beaver Divide, Wyoming, Duchesneodus
INTRODUCTION
Granger (1910) and Sinclair and Granger (1911) first reported
fossil mammals of Eocene-Oligocene age from the Beaver Divide area
of central Wyoming (Fig. 1). Since then, some uncertainty developed
regarding the precise ages of some of the Paleogene mammals from the
section at Beaver Divide, most of which was resolved by Van Houten
(1964) and Emry (1975). Here, we describe lower jaws and redescribe
a skull of brontotheres from Beaver Divide and discuss their taxonomic
identity and biochronological significance.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; CM = Carnegie Museum of Natural History, Pittsburgh; LACM = Los Angeles County Museum, Los Angeles;
NMC = National Museum of Canada, Ottawa; TMM = Texas Memorial Museum, Austin; USNM = National Museum of Natural History,
Smithsonian Institution, Washington, D. C.; YPM = Yale Peabody
Museum of Natural History.
PROVENANCE
The three brontothere mandibles described here were collected
in 1971 and 1972 by Emry from the type section of the Big Sand Draw
Sandstone Lentil of the White River Formation, on the slope immediately north of Wagon Bed Spring. This is essentially locality 9 of Van
Houten (1964, plate 2) and is in the NE 1/4 SW 1/4 sec. 34, T32N,
R95W, Fremont County. The stratigraphic level is the Big Sand Draw
Sandstone Lentil, which is a valley fill representing the earliest phase
of White River deposition locally. The three brontothere mandibles
occurred from 3 to 6 meters below the base of the crystalline facies
of the Beaver Divide Conglomerate (Fig. 1), which directly overlies
the Big Sand Draw Sandstone Lentil. The locality is also represented
by Plate XXII of Granger (1910), the caption of which says the skull
of Titanotherium was from the draw in the distance, and by Plate XI,
B of Osborn (1929), the caption of which mentions that the skull of
Menodus heloceras came from the draw just to the right of this view.
These photographs and their captions permit us to determine conclusively that the three mandibles reported here are from in the same draw
identified in these two plates, so were undoubtedly found within a few
meters of the spot where Granger collected the Menodus heloceras
skull (AMNH 14576) in 1909.
DESCRIPTION
USNM 437625 (Fig. 2B, D-E) is a mandible with left i2-3, c

FIGURE 1. Index map and stratigraphic section of Paleogene, mammal-bearing


rocks at Beaver Divide, Wyoming (after Emry, 1975).

(damaged), p1-4, m1-3 and right c, a supernumerary tooth, p1-4 and


m1-3. The two incisors are small, round (globular) and closely appressed.
The i2 is 12 mm long and 8 mm wide, whereas the i3 is 14 mm long
and 10 mm wide. The canines are relatively small (length = 25 mm,
width = 22 mm), somewhat procumbent and have a round cross section. A distinct (~ 40 mm long) postcanine diastema separates the canines from the premolars.
The right side of the jaw has a supernumerary tooth located anteromedial to the p1. This tooth has a single root, whereas the normal p1
has two divided and partly fused roots. The crown of the supernumerary tooth is conical with a round cross section, a cingulid around nearly
the entire circumference and has smooth enamel. This small tooth has
a diameter of 11 mm.
The p1 is oval in occlusal outline with a tall and blunt paraconid

114

FIGURE 2. Brontothere and rhinocerotoid specimens from Beaver Divide, Wyoming. A, C, USNM 510081, lower jaw of Duchesneodus primitivum in right lateral (A)
and occlusal (C) views. B, D-E, USNM 437625, lower jaw of Duchesneodus primitivum in right lateral (B) and occlusal (D) views and occlusal close-up of anterior
dentition (E). F, USNM 521243, cf. Hyracodon sp., occlusal view of right P2-3. G-H, USNM 521244, cf. Hyracodon sp., lateral and occlusal view of left dentary
fragment.

115
TABLE 1. Measurements (in mm) of cheek teeth of brontothere lower jaws from Beaver Divide compared to those of selected Duchesneodus
(from Lucas and Schoch, 1989) and Notiotitanops mississippiensis.
USNM
USNM
USNM Duchesneodus Duchesneodus
437625
510081
521242 primitivuma
uintensisb
p2l
26
25
27
26
22
p2w
15
16
16
18
14
p3l
29
29
30
32
26
p3w
19
21
21
23
19
p4l
34
32
35
35
30
p4w
23
25
26
27
26
m1l
49
50
48
42
m1w
30
31
29
26
m2l
63
65
58
53
m2w
36
38
33
31
m3l
90
93
82
77
m3w
36
39
33
31
a
NMC 6421
b
Mean values for the species from Lucas and Schoch (1989, table 27.1)
c
LACM/CIT 1854
d
USNM 16146

forming the trigonid, followed by a much lower, posteriorly-sloping


talonid heel. Length = 15 mm, width = 10 mm. The p2 is a much larger,
molariform tooth. It has a trigonid that is slightly wider than the talonid. The crown is low, with a broad and blunt paraconid and small
paracristid. There is a shallow talonid basin (hypoflexid) bordered by a
low, blunt hypoconid connected to the metastylid ridge by a cristid
obliqua. The p3 is similar to the p2 but larger, with a wider talonid
than trigonid, and more prominent paracristid, metaconid/metastylid
and talonid basin (hypoflexid). The p4 is nearly molariform with prominent trigonid and talonid basins and a very tall and peaked metaconid.
The m1 and the m2 are similar to the p4, but have labial cingulids
and tall entoconids, much taller than the metaconids. The m3 trigonid
and talonid are much like those of the m1-2, but the labial cingulids are
more prominent, and the hypoconulid lobe is strong, with a large lingual cingulid.
All the cheek teeth have lineated and rugose enamel. The mandible has a long symphysis that extends back to the p4 trigonid and has
a deep, concave dorsal aspect. The right side has two mental foramina,
one under the anterior part of p2 and the other under the anterior edge
of the p3. The left side has a single foramen below the juncture of p23 (this is the condition on both sides of USNM 51008 and 521242).
The horizontal ramus increases in height posteriorly to a plate-like ascending ramus with a tall coronoid process and condyle well above the
cheek-tooth row. The mandibular angle is rugose but not greatly thickened.
USNM 510081 (Fig. 2A, C) is a lower jaw very similar to USNM
437625 in size and morphology. It, however, does not have the anterior
tip of the symphyseal region preserved, so the nature of its incisors
cannot be determined. However, particularly significant features of
USNM 510081 include its relatively small and slightly procumbent
canines, distinct postcanine diastema, and four relatively molariform
premolars.
The third specimen, USNM 521242, is the anterior part of a
mandible with the symphysis. It contains right i2-3, c (damaged), p2-4,
fractured parts of m1, and left i2, c (damaged), p2-4, and fractured
parts of m1-2. The specimen represents an older individual, with molars well worn and with the incisors having the crowns nearly worn
away. This specimen adds little new information, but is consistent with
the other specimens in having a diastema at least 30 mm long between
the canines and the p1 alveoli, and also shows that at least two incisors
were not vestigial and lost early, but were functional and retained even

Protitanops
curryic

45
57
89
38

Notiotitanops
mississippiensisd
26
19
29
23
38
26
47
30
69
28

in older individuals.
IDENTIFICATION
Mandibles
Duchesneodus mandibles typically have three incisors (Lucas
and Schoch, 1989), whereas brontotheres that are unambiguously
Chadronian in age (Megacerops, Brontops, Menops sensu Mader, 1989;
Megacerops sensu Mihlbachler et al., this volume) have no more than
two (Osborn, 1929; Lucas and Schoch, 1982). However, the vestigial
incisors of Chadronian brontotheres tend to fall out during life, with
the alveoli being reabsorbed. Therefore, the presence of two lower incisors in USNM 437625 and 521242 is not necessarily of taxonomic
significance.
However, the distinct postcanine diastema in USNM 437625,
510081 and 521242 narrows the taxonomic diagnosis of these specimens considerably. All typical Chadronian brontotheres (Menops,
Brontops, Megacerops sensu Mader [1989]; Megacerops coloradensis
sensu Mihlbachler et al., this volume) most frequently lack postcanine
diastemata. A very short upper and lower postcanine diastema is occasionally present in typical Chadronian brontotheres, but it is always a
very minor gap in comparison to the comparatively long postcanine
diastemata of USNM 437625, 510081 and 521242 . Duchesneodus
uintensis mandibles have a very short postcanine diastema or lack one
altogether (Lucas and Schoch, 1989).
At least three taxa should be more closely considered:
Duchesneodus primitivum (Lambe, 1908), Protitanops curryi Stock,
1936, and Notiotitanops mississippiensis Gazin and Sullivan, 1942.
Unfortunately, each of these taxa are presently known only from their
holotypes, so we have no knowledge of their intraspecific variability.
For this reason, all we can do is compare the differences and similarities of these holotype specimens with the Beaver Divide mandibles. In
size (Table 1), the Beaver Divide brontothere jaws resembles the holotype (a mandible) of Duchesneodus primitivum. However, the postcanine
diastema of D. primitivum is shorter and the anterior end of the symphysis is much narrower. Consequently, the transverse width across the
canines of D. primitivum is much shorter than the transverse width
across the p2s. In these respects, D. primitivum more closely resembles
D. uintensis rather than the Beaver Divide mandibles, where the transverse width across the canines is more similar to the transverse width
across the p2s.

116
Notiotitanops mississippiensis and Protitanops curryi are each
known from an associated skull and mandible. However, the anterior
portions of each mandible are not preserved. Nonetheless, the skulls
indicate that these taxa possessed vestigial (globular) incisors and distinct upper postcanine diastemata that are comparable to the incisors
and postcanine diastemata of the Beaver Divide mandibles. It is reasonable to infer that Notiotitanops and Protitanops had similar mandibular incisors and diastemata. This introduces the possibility that the
Beaver Divide mandibles belong to either of these taxa.
The holotype of Notiotitanops mississippiensis is smaller than
the Beaver Divide mandibles (Table 1). The available lower cheek tooth
dimension of Protitanops curryi are much closer to those of the Beaver
Divide mandibles, and generally fall below the size range by no more
than a few millimeters. Although the actual size ranges of these species are unknown, Protitanops curryi and Duchesneodus primitivum
fall nearest to the size range of the Beaver Divide mandibles. However,
this information alone is not enough to make a specific identification.
Nevertheless, the two most likely possibilities seem to be: (1)
the Beaver Divide mandibles are either Duchesneodus primitivum or
Protitanops curryi; or (2) D. primitivum and Protitanops curryi are the
same species, with D. primitivum (Lambe, 1908) serving as the senior
synonym and the proper reference species of the Beaver Divide
brontothere mandibles. A third possibility is that the Beaver Divide
mandibles represent a new taxon, although sufficient evidence to diagnose a new species is not present in these specimens. An associated
skull and mandible is ultimately needed to solve this problem.
Skull
In addition to the brontothere mandibles described above, a single
partial cranium (AMNH 14576) of a brontothere was reported from the
Big Sand Draw Sandstone Lentil at Beaver Divide by Osborn (1929).
The specimen (Figs. 3-4) lacks the right side of the anterior portion of
the skull and lacks all teeth except the left molars.
In general, the skull of AMNH 14576 can be described as rather
small and gracile compared to most typical Chadronian brontotheres,
but falling within the range of variation for brontotheres from that landmammal age (skull length ~61 cm, M1-M3 length = 18 cm). This
specimen was referred by Osborn (1929) to Menodus heloceras (Cope,
1873), and it compares favorably to the holotype of that species (AMNH
6360 from the Titanotherium zone, Chadron A at Horsetail Creek, NE
Colorado). (Note that the apparent differences in shape between AMNH
14576 and AMNH 6360, such as the apparent width of the premaxilla,
are artificial and result from extensive plaster reconstruction of the
latter specimen.) AMNH 6360 and 14576 also resemble, in general
size and proportions, at least three other Chadronian brontothere species according to Osborns (1929) taxonomy: Brontops brachycephalus,
Brontotherium leidyi, and Allops walcotti. Other specimens including
USNM 20971 (from Cameron Springs, Fremont County, WY) and a
specimen referred by Wilson (1977) to Menodus bakeri Stovall, 1948,
TMM 40932-1 (from the early Chadronian Chisos Formation, Big Bend
National Park, Texas) fit the following description as well.
The short horns of AMNH 14576 and the above mentioned specimens are trihedral at their bases, but are rounded at their peaks. They
are elevated on short but distinct superorbital pillars that ascend nearly
vertically from the dorsal margin of the orbit. These horns are not connected by a large crest of bone, and the nasal bone extends a considerable distance anteriorly from the base of the horns. This is because the
horns are comparatively small, unlike many of the larger, Chadronian
specimens with massive horns that are connected by a large crest that
essentially absorbs the nasal process (e.g., YPM 12048, holotype of
Brontops robustus). The dorsal surface of the skull is concave, from
the horns to the nuchal crest, but the concavity is not as deeply saddleshaped as in many Chadronian specimens. The zygomatic arches are
relatively slender and are not laterally expanded. The parasagittal ridges
are widely separated over the parietal region, the nuchal crest is deeply

FIGURE 3. Osborns (1929, fig. 436) illustration of AMNH 14576, a brontothere


skull from Beaver Divide he identified as Menodus heloceras (here assigned to
Megacerops coloradensis) in left lateral (A), dorsal (B) and ventral (C) views.

notched at the midline and the occiput is angled strongly backwards.


The taxonomy of Chadronian brontotheres continues to be poorly
understood, although it is generally recognized that historically they
have been severely oversplit (Mihlbachler et al., this volume). Due to
their similarity, the specimens matching the above description are possibly the same species, of which Menodus heloceras would serve as the
senior name. The horns of these specimens are shorter and less massive
than those of the holotypes of Megacerops coloradensis Leidy 1870 and
Megacerops acer (Cope, 1873), which are the only two species of
Chadronian brontotheres named from good cranial material prior to the
naming of Menodus heloceras (Cope, 1873). However, Mihlbachler et
al. (this volume) argue that all Chadronian brontotheres with unbifurcated
horns previously recognized as belonging to the genera Brontops, Menops,
and Megacerops (sensu Mader, 1989) could be a single, sexually dimorphic species, Megacerops coloradensis. They caution that a more extensive analysis of Chadronian brontothere morphology and biostratigraphy may reveal multiple diagnosable species, but if their suggestion is
further corroborated, the vast majority of known Chadronian brontothere
specimens should be referred to Megacerops coloradensis.
The skull from Beaver Divide (AMNH 14576) is most consistent
with the concept of Menodus heloceras (which is possibly synonomous
with Megacerops coloradensis) described above but it may not actually
belong to this group of specimens. Typical Chadronian brontotheres
including those specimens that possibly belong to Menodus heloceras
either have a short postcanine diastema as in CM 93 (referred to
Brontotherium leidyi by Osborn, 1929) or there is no postcanine diastema, as is often the case with older (geriatric) individuals (e.g., USNM
20971). However, these specimens can be distinguished from Protitanops
curryi in which a relatively long postcanine diastema is present. Unfortunately, the premaxilla and the anterior dentition of the skull from Beaver Divide (AMNH 14576) are not preserved. We are therefore unable to
determine the nature of the postcanine diastema in that particular specimen. This introduces the possibility that AMNH 14576 is actually

117

FIGURE 4. AMNH 14576, skull of Megacerops coloradensis from Beaver Divide, left lateral (A), ventral (B) and dorsal (C) views.

Protitanops curryi. Therefore, AMNH 14567 may actually belong to the


same species represented by the brontothere mandibles from Beaver
Divide, which, in contrast to typical Chadronian brontotheres, are characterized by a long postcanine diastema.
However, AMNH 14576 differs in other ways from the holotype
(and only known specimen) of Protitanops curryi: the minimum width
of the dorsal parietal roof is wider, the orbit is somewhat more anterior,
the zygomatic arches are less laterally expanded, and the horns are less
bulbous and less eliptical in cross section. It therefore seems less likely
that AMNH 14576 is Protitanops curryi. Ultimately, we cannot rule out
the possibility that AMNH 14576 is Protitanops curryi despite possible differneces, because the amount of intraspecific variation in traits
like horn size and shape and zygomatic arch thickness are likely to vary
greatly within species due to sexual dimorphism (Mihlbachler et al., this
volume). For this reason, the Beaver Divide skull and jaws may actually
belong to the same species (possibly Protitanops curryi).
OTHER TAXA
Van Houten (1964, p. 67) listed the following mammals from the
Big Sand Draw Sandstone Lentil: Brachyhyops wyomingensis Colbert,
Caenopus yoderensis Schlaikjer, Metamynodon cf. M. planifrons Scott &
Osborn, Hyracodon cf. H. priscus [should be priscidens] Lambe, Menodus
heloceras (Cope), cf. Epihippus intermedius Peterson or a small species
of Mesohippus, cf. Leptomeryx and cf. Teleodus [= Duchesneodus].
Only the Brachyhyops and Megacerops specimens have previously been documented (Osborn, 1929; Colbert, 1938). Emry et al. (1987,

p. 134) discussed the equid fossil, an isolated m3 (USNM 19108), and


assigned it to Mesohippus, noting that it is slightly larger than the
Duchesnean-early Chadronian species M. texanus. The other specimens
upon which Van Houten (1964) based his identifications need to be
located and documented.
Rhinocerotoid fossils, also collected in 1971, from the Big Sand
Draw Sandstone Lentil at Beaver Divide, are USNM 521243, a right
maxillary fragment with P2-3 (Fig. 2F) and 521244, a left dentary fragment with well worn p2-m3 (Fig. 2G-H). Because these specimens
lack the anterior teeth (incisors, canines) and because the lower dentition is very worn, their identification is somewhat uncertain. The P2-3
have postprotocristae that wrap labially around the lingual edges of the
teeth and are not connected to the metalophs, so that the trigon basins
open posteriorly, not lingually. In size (P2L = 15.5 mm, w = 18.2 mm,
P3L = 17.8 mm, w = 21.9 mm) and morphology they closely match
relatively primitive species of Hyracodon (Russell, 1982; Prothero,
1996), though it is possible that they could represent Teletaceras
radinskyi (cf. Hanson, 1989, fig. 20.4, table 20.2).
The lower jaw, USNM 521244, also well matches either primitive Hyracodon or Teletaceras radinskyi in size and morphology. Note
especially the relatively nonmolariform p2-3 (though, note that this is a
judgment based on overall shape of very worn teeth), the well developed
p4 hypolophid that is not as wide as the metalophid and the low connection of the metalophid and the protolophid on m3. Thus, we identify
both USNM 521243 and USNM 521244 as cf. Hyracodon sp. Measurements (in mm) of USNM 521244 are: p2 l = 13.8, w = 8.5, p3 l = 16.6,

118
trigonid w = 12.1, talonid w = 14.5; p4 l = 17.5, trigonid w = 13.9, talonid
w = 14.1; m1 l = 20.9, trigonid w = 15.4, talonid w = 16.3; m2 l = 22.5,
trigonid w = 17.2, talonid w = 16.6; m3 l = 27.3, trigonid w = 18.2,
talonid w = 16.4
BIOCHRONOLOGICAL SIGNIFICANCE
Emry (1975) established the lithostratigraphic succession at Beaver Divide used here (Fig. 1). Particularly significant to this establishment was recognition that the Big Sand Draw Sandstone Lentil of the
White River Formation fills a large paleovalley that was cut into sediments of the Wagon Bed Formation, and that it also was cut through the
volcanic facies of the Beaver Divide Conglomerate. The volcanic
facies and the crystalline facies are not facies of one unit.
Fossil mammals from the Wagon Bed Formation indicate it is of
Uintan (Uinta B) age (Emry, 1975). Strata of the volcanic conglomerate member of the Wiggins Formation (Van Houtens volcanic facies of the Beaver Divide Conglomerate) also yield Uintan (Uinta
C) age mammals (Emry, 1975), and these mammal localities were
previously (and incorrectly) assigned to the Beaver Divide Conglomerate Member of the White River Formation (Van Houten, 1964). Some
50 km east of Beaver Divide, the White River Formation yields middle
Chadronian age mammals (the Cameron Spring local fauna) (Van
Houten, 1964; Emry et al., 1987). The Big Sand Draw Sandstone Lentil is thus between Uintan and Chadronian strata (Fig. 1), and it has
previously been assigned an early Chadronian age (e.g., Van Houten,

1964; Emry, 1975; Emry et al., 1987; Lucas, 1992).


Several of the taxa reported from the Big Sand Draw Lentil
strongly indicate a Chadronian age: Leptomeryx, Mesohippus (larger
than Duchesnean species), Metamynodon, Hyracodon and Menodus
(e.g., Lucas, 1992; Prothero and Emry, 1996). The brontothere jaws and
skull are consistent with this age even though they could belong to one or
two of several species including Duchesneodus primitivum, Protitanops
curryi, and Menodus heloceras. Although Duchesneodus has been considered an index taxon of the Duchesnean land-mammal age, the type
specimen of D. primitivum comes from the Cypress Hills Formation,
strata which have not yielded other unambiguously Duchesnean-age
mammals, but instead have long been assigned a Chadronian age (e.g.,
Lucas, 1992). Protitanops curryi is clearly Chadronian, as is Menodus
heloceras and other specimens from Texas, Colorado and Wyoming that
likely belong to this taxon. Therefore, we advocate the following:
1. The mammal assemblage from the Big Sand Draw Sandstone
Lentil is of early Chadronian age.
2. Duchesneodus primitivum is an early Chadronian taxa.
3. The species Duchesneodus uintensis (not the genus
Duchesneodus) is an index taxon of the Duchesnean LMA.
ACKNOWLEDGMENTS
We thank Peter Kondrashov and Donald Prothero for comments
on the manuscript.

REFERENCES
Colbert, E. H., 1938, Brachyhyops, a new bunodont artiodactyl from Beaver Divide, Wyoming: Annals of the Carnegie Museum, v. 27, p. 87-108.
Cope, E.D. 1873, Second notice of extinct Vertebrata from the Tertiary of the plains:
Palaeontological Bulletin, no. 15, p. 1-6.
Emry, R. J., 1975, Revised Tertiary stratigraphy and paleontology of the western
Beaver Divide, Fremont County, Wyoming: Smithsonian Contributions to
Paleobiology, no. 25, 20 p.
Emry, R. J., Russell, L. S. and Bjork, P. R., 1987, Chadronian, Orellan, and
Whitneyan North American land mammal ages; in Woodburne, M. O., ed., Cenozoic mammals of North America. Geochronology and biostratigraphy: Berkeley, University of California Press, p. 118-152.
Gazin, C.L. and Sullivan, J.M. 1942, A new titanothere from the Eocene of Mississippi, with notes on the correlation between the marine Eocene of the Gulf Coastal
Plain Rocky Mountain Region: Smithsonian Miscellaneous Collections, v. 13,
p. 1-13.
Granger, W., 1910, Tertiary faunal horizons in the Wind River Basin, Wyoming,
with descriptions of new Eocene mammals: Bulletin of the American Museum
of Natural History, v. 28, p. 235-251.
Hanson, C. B., 1989, Teletaceras radinskyi, a new primitive rhinocerotid from the
late Eocene Clarno Formation of Oregon; in Prothero, D. R. and Schoch, R. M.,
eds., The evolution of perissodactyls: New York, Oxford University Press, p.
379-398.
Lambe, L. M., 1908, The Vertebrata of the Oligocene of the Cypress Hills,
Saskatchewan: Contributions to Canadian Paleontology, v. 3, p. 1-64.
Leidy, J. 1870, Remarks on Megacerops coloradensis: Proceedings of the Academy of Natural Sciences of Philadelphia, v. 22, p. 1-2.
Lucas, S. G., 1992, Redefinition of the Duchesnean land mammal age, late Eocene
of western North America; in Prothero, D. R. and Berggren, W. A., eds., EoceneOligocene climatic and biotic evolution: Princeton, Princeton University Press,
p. 88-105.
Lucas, S. G. and Schoch, R. M., 1982, Duchesneodus, a new name for some
titanotheres (Perissodactyla, Brontotheriidae) from the late Eocene of western

North America: Journal of Paleontology, v. 56, p. 1018-1023.


Lucas, S. G. and Schoch, R. M., 1989, Taxonomy of Duchesneodus
(Brontotheriidae) from the late Eocene of North America; in Prothero, D. R.
and Schoch, R. M., eds., The evolution of perissodactyls: New York, Oxford
University Press, p. 490-503.
Mader, B. J. 1989, The Brontotheriidae: A systematic revision and preliminary phylogeny of North American genera; in Prothero, D. R. and Schoch, R. M., eds.,
The evolution of perissodactyls: New York, Oxford University Press, p. 458584.
Osborn, H. F., 1929, The titanotheres of ancient Wyoming, Dakota, and Nebraska:
U. S. Geological Survey, Monograph 55, 953 p.
Prothero, D. R., 1996, Hyracodontidae; in Prothero, D. R. and Emry, R. J., eds., The
terrestrial Eocene-Oligocene transition in North America: Cambridge, Cambridge
University Press, p. 664-683
Prothero, D. R. and Emry, R. J., 1996, Summary; in Prothero, D. R. and Emry, R. J.,
eds., The terrestrial Eocene-Oligocene transition in North America: Cambridge,
Cambridge University Press, p. 664-683.
Russell, L. S., 1982, Tertiary mammals of Saskatchewan Part VI: The Oligocene
rhinoceroses: Life Science Contributions Royal Ontario Museum, no. 133, 58
p.
Sinclair, W. J. and Granger, W., 1911, Eocene and Oligocene of the Wind River and
Bighorn basins: Bulletin of the American Museum of Natural History, v. 30, p.
83-117.
Stock, C. 1936, Titanotheres from the Titus Canyon Formation, California: Proceedings of the National Academy of Sciences, v. 22, p. 656-661.
Stovall, J.W. 1948, Chadron vertebrate fossils from below the Rim Rock of Presidio
County, Texas: American Journal of Science, v. 246, p. 78-95.
Van Houten, F. B., 1964, Tertiary geology of the Beaver Rim area, Fremont and
Natrona Counties, Wyoming: U. S. Geological Survey, Bulletin 1164, 99 p.
Wilson, J.A. 1977, Early Tertiary vertebrate faunas, Big Bend area Trans-Pecos
Texas: Brontotheriidae: The Pearce Sellards Series, no. 25, p. 1-17.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

119

O. C. MARSH AND THE EOCENE BRONTOTHERE TELEODUS:


A PALEONTOLOGICAL HOAX
SPENCER G. LUCAS
New Mexico Museum of Natural History, 1801 Mountain Road N. W., Albuquerque, NM 87104

AbstractIn 1890, O. C. Marsh purchased from L. W. Stilwell for $100 the lower jaw of a brontothere that he
made the holotype of Teleodus avus. The distinctive characteristic of this jaw was its apparent possession of six
lower incisors instead of the four or fewer incisors characteristic of other, coeval brontotheres. However, the two
extra incisors of the Teleodus avus jaw were added by Stilwell to a jaw with four incisors in order to induce
Marsh to pay more for the fossil. Marsh did not spot the fake, nor did other paleontologists for nearly a century.
The Teleodus avus hoax is yet another example of the authenticity problems inherent to the commercial purchase of fossils as well as the great capacity all paleontologists have for seeing what they want to see in a fossil,
not what actually is there.
Key words: O. C. Marsh, Brontotheriidae, Teleodus, L. W. Stillwell, hoax
INTRODUCTION
The Brontotheriidae are a family of perissodactyls with a fossil
record from Eocene strata in Europe, Asia, and North America (e.g.,
Osborn, 1929; Yanovskaya, 1980; Lucas and Schoch, 1989a; Mader,
1989). Also called titanotheres, the initial discoveries of these rhinoceros-like mammals (Fig. 1) were made in the American West during
the 1840s (Osborn, 1929). No man was more central to the early discoveries and scientific understanding of the brontotheres than the Yale
Peabody Museums paleontologist, Othniel Charles Marsh (1831-1899),
who coined the family name Brontotheriidae (thunder beasts) in 1873.
Marsh collected some brontothere fossils himself, but he mostly
obtained brontothere fossils from his own paid collectors and others
out West from whom he purchased fossils. One of these purchased
fossils was a lower jaw, for which Marsh (1890) coined the name
Teleodus avus. Marsh purchased this lower jaw (Yale Peabody Museum [YPM] 10321) in 1890 for $100 from L. W. Stilwell of Deadwood, South Dakota. Already, in 1890, Marsh knew that most of the
brontotheres he had studied, which he considered to be of Miocene age
(they actually are of latest Eocene [Chadronian] age) had none, or only
one or two lower incisors on each side of the lower jaw. However, the
lower jaw he bought from Stilwell (it arrived at YPM on 15 March
1890) had three lower incisors on each side, and Marsh felt that this
feature justified creating a new genus and species for it.
Nearly a century later, Lucas and Schoch (1982) revealed that
the lateral incisors of YPM 10321 were added to the fossil, and are not
part of the original jaw, which contained only four lower incisors, like
most other Chadronian brontothere jaws. They concluded that either
Stilwell or an assistant may have added the two i3s to a specimen of
Brontops or Brontotherium with only four lower incisors (Lucas and
Schoch, 1982, p. 1021).
The question of who added the extra incisors to the holotype of
Teleodus avus was not fully resolved by Lucas and Schoch (1982). Here,
I review the correspondence of Stilwell and Marsh regarding purchase
of the fossil to conclude that Stilwell added the incisors to the fossil in
order to induce Marsh to purchase it for a high price.
MARSH AND THE BRONTOTHERIIDAE
Osborn (1929, p. 141-149) reviewed in detail nineteenth century discoveries and early research on brontotheres. The first discoveries in the 1840s came from the White River badlands of South Dakota
and northwestern Nebraska. Early descriptions of these fossils, which
were only teeth and jaw fragments, especially by American paleontologist Joseph Leidy (1823-1891), failed to recognize the distinctive nature of the brontotheres. Thus, although Leidy (e.g., 1854) and others

FIGURE 1. Skeleton and restoration of the brontothere Brontops, original skeleton


about 3.6 meters long (from Osborn, 1929).

early referred them to Perissodactyla, even using the generic name


Palaeotherium (a kind of extinct horse) for some specimens, he later
(e.g., Leidy, 1869) assigned them to the Artiodactyla.
Marsh entered the study of brontotheres in 1870 when he led a
Yale College expedition to northern Colorado, where they collected
many brontothere fossils, including skulls, jaws and postcrania. Based
on this new, more complete material, Marsh (1873, p. 486) recognized
that his genus Brontotherium is a true perissodactyl with limb bones
resembling those of Rhinoceros. The genus is related to Titanotherium,
and the two appear to form a distinct family, which may be called
Brontotheridae (correctly spelled Brontotheriidae). Marsh (1874, 1876)
followed with important articles that delineated the characteristics of

120

FIGURE 2. Reconstruction of the skeleton of Brontops, original skeleton about 3.6


meters long (from Marsh, 1889).

the Brontotheriidae. After that, he published several articles describing new brontothere taxa (see Osborn, 1929 for a review) culminated
by an elegant restoration of the skeleton of Brontops (Marsh, 1889)
(Fig. 2).
In his many articles on Miocene brontotheres, Marsh repeatedly stressed that they had two or fewer lower incisors on each side of
the lower jaw (four or fewer incisors total). Thus, in his 1876 paper,
Marsh listed the four genera of Miocene brontotheres then known,
giving their lower incisor count as two (Menodus, Brontotherium), one
(Diconodon) or zero (Megacerops). So, when Marsh acquired a
brontothere lower jaw with six incisors (Fig. 3), he was quick to make
it the type of a new genus and species, Teleodus avus, stating that it
resembles Brontotherium but differs from it in having six lower incisors instead of four (Marsh, 1890, p. 524). Marsh based the name
Teleodus on Greek teleos (finished, complete) and odon (tooth), and
the name avus is Latin for grandfather. It probably reflected his conviction that Teleodus avus was more primitive than and perhaps ancestral
to the other Miocene brontotheres.
TELEODUS
After Marsh named the genus in 1890, Teleodus remained a
widely used and well accepted generic name until 1982. Some early
workers (Hatcher, 1893; Osborn, 1896, 1902) considered it to be a
synonym of either Titanotherium Leidy or Megacerops Leidy. However, Osborn (1929), in his massive monographic revision of the
brontotheres, identified Teleodus as a distinct genus. He even assigned
to it a lower jaw from Saskatchewan that Lambe (1908) had named
Megacerops primitivus, as did Russell (1934). Peterson (1931) further
extended use of the name Teleodus by coining a new species, T. uintensis,
for brontotheres with six lower incisors from the upper Eocene of Utah.
Additional species were named, and additional specimens of Teleodus
were described through the early 1980s (e.g., Stock, 1935, 1938; Scott,
1940, 1945; Bjork, 1967; Clark et al., 1967; Lucas and Kues, 1979;
Nelson et al., 1980; Lucas, 1982, 1983). Teleodus was also considered
to be of biostratigraphic value as a characteristic fossil of the late Eocene
Duchesnean land-mammal age (e.g., Wood et al., 1941; Clark et al.,
1967; Wilson, 1978; Emry, 1981).
However, in 1982 Lucas and Schoch re-evaluated the holotype
of Teleodus avus (Fig. 3) to conclude that it originally had four lower
incisors, and two had been added to the specimen, presumably by L. W.
Stilwell. They coined the new generic name Duchesneodus for the species Peterson (1931) had named Teleodus uintensis, which actually is a
brontothere species with six lower incisors, as are the other species
that had been referred to Teleodus (Lucas and Schoch, 1989b; Lucas,

FIGURE 3. The holotype of Teleodus avus, YPM 10321. A-B, Left lateral (A) and
occlusal (B) views of lower jaw. C-D, Occlusal (C) and anterior (D) views of anterior
dentition. Arrows indicate the lateral incisors.

1992). Lucas and Schoch (1982) concluded that the holotype lower jaw
of Teleodus avus belongs to Brontops or Brontotherium (=Megacerops:
Mader, 1989). The name Teleodus thus became a synonym of one of
those genera.
Lucas and Schoch (1982) presented the following observations
to support their contention that YPM 10321 (Fig. 3) only had four actual incisors and that somebody had added two (the lateral incisors) to
the specimen:
1. The six incisors are not fully imbedded in bone. Both lateral
incisors are totally imbedded in plaster. Plaster also partially holds the
remaining four incisors in place.
2. The right lateral incisor is only a small fragment of a root,
missing the crown at a fresh break, suggesting relatively recent damage.
3. The left lateral incisor is a more complete root fragment with
only the base of the crown preserved.
4. Radiographs demonstrate that the four central incisors have
roots that extend into the mandible.
5. However, on the right lateral incisor its lower tip does not
even approach the edge of the mandibular bone, so the tooth is not
rooted in bone.
6. Also, the root tip of the left lateral incisor terminates above
the edge of the mandibular bone, so it too is not rooted in bone.
7. The six incisors are tightly crowded in a straight transverse
line anterior to the canines. This differs from other brontotheres in
which the incisors are either between the canines or are rooted in a
bony arc anterior to the canines (Osborn, 1929).
8. Except for the two lateral incisors, all features of the teeth
and dentary of YPM 10321 are identical to Brontops or Brontotherium,
Chadronian brontotheres that have four or fewer incisors.

121
9. No Chadronian brontothere found since 1890 has six lower
incisors.
THE MARSH-STILLWELL CORRESPONDENCE
L. W. (Lucien Warren) Stilwell was born in Manlius, New York
on 24 March 1844 and died in Deadwood South Dakota in 1932. Stilwell
grew up and attended school near Fond du Lac, Wisconsin and studied
literature and mathematics at Ripon College in Wisconsin. As a young
man, Stilwell worked in or owned several businesses, including a market and produce business and a wholesale commission grain business.
He arrived in Deadwood, South Dakota, in 1879 and took a position
with a bank. By 1880, Stilwell had begun selling natural history objects from a building behind his house. This enterprise became so successful that in 1890 he resigned his position at the bank to devote himself fulltime to his natural history sales. Helbock (2000) also identifies
Stilwell as an Indian trader (dealer in crafts and artifacts) in the years
1907 and 1914.
Stilwells letterhead to Marsh lists him as a wholesale and retail dealer in natural history specimens in Deadwood, South Dakota,
and correspondence between Marsh and Stilwell is preserved at the
Beinecke Library of Yale University, catalogued as items 16515 through
16655. These letters make reference to the fact that Marsh and Stilwell
had met, presumably in Deadwood, and that Stilwell, indeed, had
worked in a bank prior to becoming a dealer in natural history specimens. Marshs correspondence with Stilwell began in 1889, when he
received the following letter:
Apr. 23 89
Prof. Marsh

Yours truly,
L. W. Stilwell
Marshs reply is not available, but, in July of that year, Marsh
purchased a skull of Elotherium (a kind of artiodactyl) from Stilwell
for $150. By the Fall of 1889, Marsh had also sent Stilwell some of his
publications:
Nov 6 89
Prof. O. C. Marsh
Yale College New Haven, Conn.
Dear Sir:
Yesterday I received the two vols you kindly sent me
and for which I thank you very much. The one on the
Dinocerata I prize very highly as it will come in good use.
The other will be of interest to me if not of practical use.
I infer you have recd the Elotherium & C. I hope they
arrived in nice order.
Truly yours,
L. W. Stilwell
What followed from Stilwell, early in 1890, was a letter to Marsh
announcing new fossils, obviously for sale:
Jan 22 90

New Haven, Conn.


Prof O. C. Marsh
Dear Sir:
New Haven, Conn.
Have you ever obtained a complete head of either the
Brontortherium [sic] or the titanotherium [sic]? I have
handled a great many of these fossils & have yet to know
that I have seen the lower jaw of the Bronto or the upper
jaw of the titano unless the one is the complement of the
other. I know you have given this subject considerable
study and have had colectors [sic] in the Bad Lands and
I desire to ask if you ever thought the Bronto and Titano,
so called, were not one and the same animal as we find
the teeth? I know of a Bronto and a Titao [sic] the upper
jaws of one (skull) and the lower jaws of the other in
place together as one head. The teeth clashed together as
one mouth, the Bronto being the skull and the pair of Titano
jaws at the bottom, teeth up, all as one head. Found in
these positions in the Bad Lands. I am to get out a new
catalog at once and you would gratify me very much with
a little instruction. Have these names Bronto & Tiatano
[sic] been reclassified or changed in any way, or do they
remain the same, as names of the two genera? What do
you call the deposit in which they are found? I have a
newly found ammonite as pictured at the head of this sheet.
Can you tell me if it is a new specie [sic] or one that has
been seen and named before in some other locality. It is a
curled, corrugated animal imbedded in a sandstone nodule. From 10 to 20 Lbs. i.e. the nodule weighs that. It is a
slate colored concretion. Pllease [sic] give me an idea of
its name or if you think it an undetermined specie [sic].
Your early reply will greatly oblige.

Dear Sir:
I think I have something new now. Plate (?) No 3 Enclosed. The 4 teeth there represent very imperfectly, excepting as to measurement, are perfect, handsome teeth
in a solid piece of the upper jaw. Nos 3 & 4 appear in
form of the Brontotherium, but No 2 is entirely unlike any
Bronto I ever saw. It has the appearance of being half longer
anterior-post. than transversely, but measurement proves
the greatest length to be 1 & 5-8 and greatest diameter
near the roots to be 1 & 1-4 inches. No 1 is drawn proportionately too large. The space occupied by the 4 teeth is
just 6 inches. I have another specimen of the very same
type with 3 tee[th] like 2,3,&4 & 4 2,3,&4 only they are
1-4 smaller. They seem to be very similar to the Paleosyops,
only that tooth No 2, are they any thing new and do you
wish them? My plate No 1 is a Bronto. lower jaw, part of
one side gone as you see. It is a fine specimen and the 6
incisors make it valuable I should suppose. No 2 is very
nearly the same size specimen but in front of jaw between
the canines is the deep notch, like a sharp valley. One
incisor hugging each canine. Are these specimens desirable to you? If so can we make a bargain? I wish you would
reply at once if you do not wish them. And if you do not
wish them will you kindly return the drawings with such
word that I may use them to send out again. In any event
will you kindly telll [sic] me if 6 incisors in the jaw of

122

FIGURE 4. Stilwells drawing of what became the holotype lower jaw of Teleodus avus. Original on left, redrawn for clarity on right.

Bronto is often found. And what you think is name of


small teeth on sheet No 3. I would klike [sic] to know
what I must get to measure teeth with by lines and if I
should measure the greatest length and diameter.
You will observe that the canines of N10 [sic] I standup
nearly perpendicular and those of No 2 at an angle of 45
degrees.
I have a fine Placenticeras placenta that measures 24
inches across, and weighs 125 lbs. and also a turtle 2 feet
long.
Will you please tell me what name you found those
teeth to be that looked like the Oreodon major only the 4
lobes of each tooth were long and sharp, inner ones nearnly
[sic] as prominent as the exteriors. Have you a work illustrated on the Bronto. I am often puzzled at the different
sizes of series of 6 teeth. I find some more than half larger
than others in the same position. And I had a lower jaw
the other day that went to Europe that seemed about 2-3
size of the ordinary find. And a very short small row of 6
upper Bronto. teeth. These went to Europe.

I await your reply duly, and remain


yours truly
L. W. Stilwell
The sketch that accompanied the letter, drawn by Stilwell with
his annotations, clearly showed the jaw with six incisors (Fig. 4). Marsh
wrote a note on this last letter, Diplacodon? (a middle Eocene
brontothere with six lower incisors) and on 28 January 1890 fired off a
telegram:
To L. W. Stilwell
Deadwood, Dakota
Letter recd. Please send me on approval by freight all the
jaws and teeth mentioned and figured. Answer, when
shipped.

O. C. Marsh

123
What he received in reply was a letter to negotiate the price:

Dear Sir:

Feb 1 1890

Upon recpt of your letter of Feb 7th I concluded to send


you the 4 specimens of Bronto & C outlined for you previously. I so telegraphed you.

Prof. O. C. Marsh
New Haven, Conn
Dear Sir:
Yr. telegram red duly. I desire you to have the teeth I
wrote you about & hesitate about sending on approval
because if you should not give me what I can get from
others it would make more handling & risk of breakage &
would be a delicate matter to ship from yr. hands to others
in competition with you. I prefer something more definite
stipulated before they leave my hands. I think the Bronto
with the 6 incisors intact is a valuable specimen. They are
all just as represented nothing less than $250.00 for the
lot on the 3 sheets (4 spec) would at all suit me. I may be
able to get a higher bid. I want from you something definite as to price. I would wrap them carefully in much cotton.
It will be useless for you to offer me less than 250 for the
lot until at least I had given others of this fraternity a chance
to secure them. I will not promise them to anyone else
than you until I have your reply.

Very truly yours


L. W. Stilwell
Almost simultaneously, Marsh received a telegram from F.
R. Carpenter, one of his paid collectors, indicating the brontothere jaw
with six incisors might go out to Marshs rivals for bid:
Jan 29 1890
To Prof O. C. Marsh,
Stilwell has broncho [sic] with six incisors intact, and proposes to send drawings to Cape [sic, Cope?] Scott. Others
hoping for big price have asked him to wait until I hear
from you.

F. R. Carpenter
Deadwood, S. Da.
Both Stilwells letter and Carpenters telegram obviously
piqued Marshs interest, and he wrote Stilwell (letter not preserved)
and received a short telegram on 11 February 1890 from Stilwell (letter recd will ship at once and write fully) followed by a long letter:

I shall trust you not to use them for measurement or


description until sold to you. In case others wanted them
at a better price than you are willing to give I would not
consider it fair to the purchasing scientist for you to possess measurements and descriptions. On the score of your
paying what we may agree on, if you buy, I had no hesitancy to sending them on approval. I did express myself
willing to send on approval but when I did so at time you
were here I contemplated having some stipulated price
and then sending on approval.
But I can readily see the difficulty in my putting a price
on anything I know nothing about. I do know some thing
about the desirability of the Menodus. If an entirely new
animal to me I would know nothing of its value an[d] would
hvave [sic] to trust you to give me what it might be wort[h]
I got an impression from the manner of your trading
when here that your views of values and mine were too
far apart. I think you got the large specimens you bought
cheap enough. And I believe that there were some small
teeth in that 3.00 lot you bought that were worth their
weight in gold. Commercially, I havent [sic] the lest [sic]
hesitancy in trusting you with any amount on approval.
It was only on the ground that I want only the purchaser to have the advantage of the benefit of these fine
and rare finds and that we might be too far apart in prices.
I think the four pieces are worth $250. Easily and I have
packed them finely in cotton and excelsior in two boxes
and sent them to you p[e]r freight, upon the proposition
that I have a right, (as they are on approval,) in the mean
time, to try for higher figures than you may be willing to
give, but reserving the right first to you to take them if
you allow as much as any one else. I would like to have
you possess them because they are your specialty and because you are in a position to make them well known to
science. I think you may never have seen the 6 perfect
incisors in place as finely as these.
While you have described an animal with 6 incisors
was it not done by seeing part of the alveoli rather than all
6 teeth in one specimen.
Please unpack them very carefully. Three premolars
belonging to the Bronto. with two incisors are packed separately in cotton and put in same box, the smaller box contains only the one specimen with 6 incisors. The larger
box contains all the rest.
Truly yours

Feb 13 90
Prof. O. C. Marsh

L. W. Stilwell

New Haven, Conn.

The reply, more than a month later, is preserved as a typed copy

124
in Yale archives. In it, Marsh devalues the specimens Stilwell sent,
indicating that none of the fossils represent new taxa:

Very truly yours,


O. C. Marsh

New Haven, Conn., March 26, 1890


L. W. Stilwell, Esq.

Dear Sir:
The two boxes arrived in apparent good order a few
days ago, and I have now examined with some care the
specimens they contained, a description of which you had
previously sent me.
The lower jaw with six incisors came in good order, as
did also the two small specimens of upper jaws. The other
large lower with two incisors was rather badly broken,
although the box was complete. The jaw was broken near
the middle, where I think there was an old break, and also
split through between the incisors. These breaks, however, can be repaired without great trouble, but the job
must be done carefully.
All four specimens belong to the Brontotheridae, but
whether to the genus Brontotherium, I cannot say without
a more careful examination. The jaw with the six incisors
is not my genus Diplacodon, as I thought possible, from
your description, but it may be a Brontotherium with two
of the milk incisors still remaining in position. See my
Dinocerata, Page 38, figures 39 and 40, and you will understand what I mean. I have several Brontotherium jaws
which retain one or more of the deciduous incisors, and I
have one which has three incisors in place on one side.
There is no reason why this should be rare, as the incisors
vary in number very much.

P.S. The specimen I bought of you at the Bank was not


new as you suggest. It was the Hyopotamus of Leidy described long ago. I thought I wrote you the name soon
after my return. I shall always be glad to find you any
information I can.
In particular, Marsh indicates the jaw with six incisors is not
Diplacodon, as he previously thought, but may be a Brontotherium
with two of the milk incisors still remaining in position. Marsh even
states that he already has a specimen with three incisors on one side
(but, I have not been able to locate such a specimen in the YPM collection). Furthermore, a note written by Marsh to the PM Coll (Peabody
Museum collection) is kept with the letter in Yale archives:
Note for PM coll on receiving specimens.
2072 Brontos
(1) Lower jaw (6 pm & m) & 6 incisors medium size (spec
in good order) one ramus complete except angle and
condyle cor. process other ramus gone behind last PM
Of 6 incisors, middle one largest on each side, outer one
smallest & close to canine. PM & M all well worn, hence
adult fully
This seems to suggest that Marsh did not regard the third incisors of the lower jaw as the retained deciduous teeth, and his publication of the jaw soon thereafter (Marsh, 1890) confirms the point. Marsh
thus misled Stilwell in his 26 March letter in order to purchase the
specimen at a lower price.
Marshs ploy worked, for Stilwells subsequent letter lowered
prices, though for the jaw with six incisors he still insisted on $100:
April 1 90

The other larger lower jaw has nothing unusual about


it, except, perhaps, the notch between the incisors, but
this is a character of little importance. I have a half dozen
jaws which show this peculiarity.
The two portions of upper jaws belong to young individuals, probably of Brontotherium. I have quite a number of about this size, and many others larger and smaller.
I have thus stated frankly and fully to you what the
specimens are you have sent me. They are all of interest,
and I should like to keep them all for my collection at a
fair price. Had they represented new genera, or forms that
I have not seen, they would have been much more valuable than they are. They are worth more to me than to
anybody else, just now while I am preparing my book,
and I should think one hundred dollars would be a good
price for the lot. If we cannot agree on the price for all
four specimens, please write me. Frankly, as I have to
you, what value you place on each one of the different
specimens, if I do not take them all. I am much obliged to
you for sending them to me, and I should like to purchase
them and all others of interest you may send me.
Hoping soon to hear from you.

Prof O. C. Marsh
New Haven, Conn.
Dear Sir:
Your favor of the 20th MCH [sic] to hand. I am sorry to
hear of the cheaper of the two jaws being broken.
In reply will say that if none of the specimens are new
that it of course makes some difference in their value.
Whether or not the jaws with the six incisors is a new
thing, the fact that the six teeth are all there in place is a
rare find. 3 teeth on one side & 3 alveoli on the other side
are not like 6 teeth in place.
My prices were
Best spec

150.

Next (2 incisors)

65.

The two young Bronto. both

35.

125
As the one with two incisors was broken I will call
that specimen $40.00
2 young series both

25.00

The 6 incisors

100,

This is the lowest I can stand. If you can possibly take


the broken specimen at $40. please do so and not move it.
The young animals I am not particular whether I sell or
not as they are the only fine young Bronto. I ever got in 5
years and I would like to have them sent back by express
if not wanted for $25. The 6 incisor specimen, if you cannot use at $100 (one hundred) I wish you would very carefully pack it by itself alone in a box, with plenty of cotton
around the front and firmly placed in dry excelsior and
the box banded with hoop iron. and send it to L.W. Morris
& Son, 18 Broadway, New York, N.Y.
I have two parties who will take it for $100. and glad
to get it. But I gave you the first chance. If you decide not
to keep it please be kind enough to do this at once. Very
truley [sic] yours
L. W. Stilwell
Marsh accepted these lowered prices, as his telegram indicated:
Apr. 5 90
Letter recd. Will take all these specimens at price named
if you will give me refusal of any future discoveries [of
interest].

O. C. Marsh
Stilwells reply made it clear he wanted to continue selling fossils to Marsh:

And without the delay and risk incident to shipt [sic]


to you and reshipment.
But, Prof. I shall be glad to help you and the interests
of science all I can afford to. And am glad to let you see
and buy whatever new I get if there is no loss in price or
inconvenience to other good customers. I am very likely
to send you descriptions and send on approval, if you request it, any new objects you may think desirable, and
give you the first chance to buy. When I did not knotw
[sic] the value of a thing and dared nor set a price on it,
and you should appreciate it as I could not, I should want
you to offer me what it might be worth. If I have an idea of
the value of a thing I can put a price on it. And it would be
at your option to take it or not.
I shall soon collect in Wyoming among the Dinocerata
and shall find other animals that I very likely cannot determine. Some may be so valuable that I could not afford
to give them away to have them named for me; and some
sample teeth I would be willinhg [sic] to contribute to
have a scientist tell me the names.
If I favor you with new finds on approval and sometimes make a gift of some teeth, If I send to you for determination would you like to name fossils for me I may get
in Wyoming?
I shall be in that countr[y] and will want to know the
names of finds in order to catalogue them and label corectly
[sic] in shipping to Europ[e]
I promise to favor you all I can. Would it be any advantage to me and my business, If you thought it best to mention my name as collector of this 6 incisor Bronto., if you
are to mention the specimen particularly? in yr. book. I
have two collecors [sic] out roaming the country far and
wide and shall undoubtedly find ma[ny] fossils which I
have not seen before.

April 5 90
Prof. O. C. Marsh
New Haven, Conn.

Truly yours
L. W. Stilwell

Dear Sir:

Locality data followed after Stilwell received payment:

Your telegram recd. In reply will say that I am known


as one who will try to live up to promises and do the fair
thing. I do not wish to make any cast iron promises not
knowing what future circumstances might arise. I have
several strings to my bow and do not need to bind myself
to anyone regarding furture [sic] finds. Indeed it might be
a loss to me sometimes to make any fixed obligation. I
have several good customers to please and I am in the
business for a living and all there is in it. From the experience that we have had together, while pleasant to me, I
am of the opinion that I can often sell for more money
than you are willing to give. I do not want any unreasonable price, but what you may think too much I know of
others that would willingly give my prices, especially
in Europe, as well as America.

Deadwood, So. Dak May 24 90


Prof O. C. Marsh
New Haven, Conn.
Dear Sir!:
You remittance of May 6th 165.00 to hand duly. I waited
to thank you until now to get from my collectors the formation you asked for.
The 4 specimens were all found 5 miles west of the
Cheyenne River midway between French and Battle Creek.
In a stratum lowest of three stratums. Probably Miocene.

126
In a grayish white marl and chalcedony gravel.
Obviously, both men were satisfied with the transaction. Marsh
subsequently purchased large numbers of fossils from Stilwell, many
from the badlands of South Dakota, a relationship that continued until
Marshs death in 1899.
THE TELEODUS HOAX
Marshs published work on brontotheres and the Marsh-Stilwell
correspondence presented above indicate the following:
1. By 1890, Marsh had described several taxa of Miocene
brontotheres, and all had four or fewer lower incisors.
2. Marsh had shared at least some of his published work with
Stilwell. So, Stilwell apparently knew that a Miocene brontothere
lower jaw with six lower incisors was an important find.
3. In his letters to Marsh, Stilwell laid repeated emphasis on the
jaw with six lower incisors. His letters of 22 January, 1 February, 13
February and the Carpenter-Marsh telegram not only did this, but also
made it clear that others might buy the fossil, and that Stilwell was
willing to sell it to them. Indeed, even after Stilwell lowered the price
of the jaw to $100 he still indicated that if Marsh would not pay that
sum (letter of 1 April), then two buyers in New York would.
4. Stilwell provided Marsh with a labeled drawing (Fig. 4) of
the lower jaw with six lower incisors that became YPM 10321, the
holotype of Teleodus avus.
5. The specimen had six incisors when Marsh acquired it; Stilwell
had evidently added the two extra lateral incisors, using plaster to hold
them in place.
6. Marsh failed to perceive that the fossil had been doctored,
although he had managed to drive the price of the jaw down by misleading Stilwell regarding its uniqueness.
7. Marsh thus published the fossil in good faith, believing that
in naming Teleodus avus he was basing it on a jaw that actually contained six incisors.
DISCUSSION
The story of the holotype of Teleodus avus is a cautionary tale
for two reasons. First, it reveals the dangers inherent when buying fossils. Second, it underscores the lack of critical attention on the part of
Marsh and a whole series of subsequent researchers, none of whom

questioned the authenticity of the six incisors in the holotype lower jaw
of Teleodus avus.
The Marsh-Stilwell correspondence reveals two men driving a
hard bargain. Both was willing to practice deception to reach the desired price. Stilwell doctored the lower jaw to create a new, unique
kind of brontothere that he believed would merit a high price. Marsh
misled Stilwell about its uniqueness, in order to reduce the sale price.
Clearly, the pressure of the sale can distort honesty and judgment of
both seller and buyer, and such distortion is a danger inherent to the
buying and selling of fossils.
Marsh uncritically published the holotype of Teleodus avus as
having six lower incisors, and subsequent workers took him at his word
for nearly a century. In large part, this was because there are brontotheres
with six lower incisors. Indeed, all the fossils referred to Teleodus except the holotype belong to a taxon with six lower incisors,
Duchesneodus (Lucas and Schoch, 1982, 1989b). However, a critical
re-examination of the holotype of Teleodus avus that revealed it had
been doctored was not undertaken until 90 years after Marsh had first
published the specimen.
The lessons of the story of Teleodus will continue to be relearned
by paleontologists as long as fossils are bought and sold. A very recent
relearning is the case of Archaeoraptor, a supposedly remarkable
bird-dinosaur link from China (Simons, 2000). This fossil was doctored, sold and initially accepted as authentic by several paleontologists. Unlike Teleodus, however, the fakery of Archaeoraptor was revealed in less than one year. The desire to sell a fossil at a high price
may induce the seller to doctor it. The purchaser, enamored of the highpriced fossil, may not spot the fake. This is what happened with the
holotype of Teleodus avus, and it will surely happen again. Caveat
emptor!
ACKNOWLEDGMENTS
Access to archives in the Yale Peabody Museum and the Beinecke
Library of Yale University made this article possible. Robert Schoch
aided me in reviewing the archived correspondence of Marsh and
Stilwell. Arlette Hensen, Curator of the Adams Museum and House in
Deadwood, South Dakota, provided biographical information on L. W.
Stilwell in the form of an account by Elizabeth Berry Hagmann in Some
history of Lawrence County (1981, p. 442-445). Reviews by Robert
Emry, Vincent Morgan and Robert Sullivan improved the manuscript.

REFERENCES
Bjork, P. R., 1967, Latest Eocene vertebrates from northwestern South Dakota: Journal of Paleontology, v. 41, p. 227-236.
Clark, J., Beerbower, J. R. and Kietzke, K. K., 1967, Oligocene sedimentation, stratigraphy, paleoecology and paleoclimatology in the Big Badlands of South Dakota: Fieldiana: Geology Memoirs 5, 158 p.
Emry, R. J., 1981, Additions to the mammalian fauna of the type Duchesnean, with
comments on the status of the Duchesnean age: Journal of Paleontology, v.
55, p. 563-570.
Hatcher, J. B., 1893, The Titanotherium beds: American Naturalist, v. 27, p. 204221.
Helbock, R. W., 2000, Indian trader advertising covers: La Posta, November 2000,
p. 58-63.
Lambe, L. M., 1908, The Vertebrata of the Oligocene of the Cypress Hills,
Saskatchewan: Contributions to Canadian Paleontology, v. 3, p. 1-64.
Leidy, J., 1854, The ancient fauna of Nebraska, or a description of remains of extinct Mammalia and Chelonia from the mauvaises terres of Nebraska:
Smithsonian Contributions to Knowledge, v. 6 (7), 126 p.
Leidy, J., 1869, The extinct mammalian fauna of Dakota and Nebraska, including
an account of some allied forms from other localities, together with a synopsis of
the mammalian remians of North America: Academy of Natural Science Philadelphia Journal, Second Series, v. 7, 472 p.
Lucas, S. G., 1982, Vertebrate paleontology, stratigraphy, and biostratigraphy of
Eocene Galisteo Formation, north-central New Mexico: New Mexico Bureau of

Mines and Mineral Resources, Circular 186, 34 p.


Lucas, S. G., 1983, The Baca Formation and the Eocene-Oligocene boundary in
New Mexico: New Mexico Geological Society, Guidebook 34, p. 187-192.
Lucas, S. G., 1992, Redefinition of the Duchesnean land mammal age, late Eocene
of western North America; in Prothero, D. R. and Berggren, W. A., eds., EoceneOligocene climatic and biotic evolution: Princeton, Princeton University Press,
p. 88-105.
Lucas, S. G. and Kues, B. S., 1979, Vertebrate biostratigraphy of the Eocene Galisteo
Formation, north-central New Mexico: New Mexico Geological Society, Guidebook 30, p. 225-229.
Lucas, S. G. and Schoch, R. M., 1982, Duchesneodus, a new name for some
titanotheres (Perissodactyla, Brontotheriidae) from the late Eocene of western
North America: Journal of Paleontology, v. 56, p. 1018-1023.
Lucas, S. G. and Schoch, R. M., 1989a, European brontotheres; in Prothero, D. R.
and Schoch, R. M., eds., The evolution of perissodactyls: New York, Oxford
University Press, p. 485-489.
Lucas, S. G. and Schoch, R. M., 1989b, Taxonomy of Duchesneodus
(Brontotheriidae) from the late Eocene of North America; in Prothero, D. R.
and Schoch, R. M., eds., The evolution of perissodactyls: New York, Oxford
University Press, p. 490-503.
Mader, B. J., 1989, The Brontotheriidae: A systematic revision and preliminary
phylogeny of North American genera; in Prothero, D. R. and Schoch, R. M.,
eds., The evolution of perissodactyls: New York, Oxford University Press, p.

127
458-484.
Marsh, O. C., 1873, Notice of new Tertiary mammals: American Journal of Science, v. 5, p. 407-410, 485-488.
Marsh, O. C., 1874, On the structure and affinities of the Brontotheridae: American
Journal of Science, v. 7, p. 1-8.
Marsh, O. C., 1876, Principal characters of the Brontotheridae: American Journal
of Science, v. 11, p. 335-340.
Marsh, O. C., 1889, Restoration of Brontops robustus, from the Miocene of America:
American Journal of Science, v. 37, p. 163-165.
Marsh, O. C., 1890, Notice of new Tertiary mammals: American Journal of Science, v. 39, p. 523-525.
Nelson, M. E., Madsen, J. H., Jr. and Stokes, W. L., 1980, A titanothere from the
Green River Formation, central Utah: Teleodus uintensis (Perissodactyla:
Brontotheriidae): University of Wyoming, Contributions to Geology, v. 18, p.
127-134.
Osborn, H. F., 1896, The cranial evolution of Titanotherium: Bulletin of the American Museum of Natural History, v. 8, p. 157-197.
Osborn, H. F., 1902, The four phyla of Oligocene titanotheres: Bulletin of the American Museum of Natural History, v. 16, p. 91-109.
Osborn, H. F., 1929, The titanotheres of ancient Wyoming, Dakota, and Nebraska:
U. S. Geological Survey, Monograph 55, 953 p.
Peterson, O. A., 1931, New species of the genus Teleodus from the upper Uinta of
northeastern Utah: Annals of the Carnegie Museum, v. 20, p. 307-312.

Russell, L. S., 1934, Revision of the lower Oligocene vertebrate fauna of the Cypress Hills, Saskatchewan: Royal Canadian Institute Transactions, v. 20, p. 4967.
Scott, W. B., 1940, The mammalian fauna of the White River Oligocene. Part V
Perissodactyla: Transactions of the American Philosophical Society, v. 28, p.
747-980.
Scott, W. B., 1945, The Mammalia of the Duchesne River Oligocene: Transactions
of the American Philosophical Society, v. 34, p. 209-253.
Simons, L., 2000, Archaeoraptor fossil trail: National Geographic, October 2000,
p. 128-132.
Stock, C., 1935, Titanothere remains from the Sespe of California: Proceedings of
the National Academy of Science, v. 21, p. 456-462.
Stock, C., 1938, A titanothere from the type Sespe of California: Proceedings of the
National Academy of Science, v. 24, p. 507-512.
Wilson, J. A., 1978, Stratigraphic occurrence and correlation of early Tertiary vertebrate faunas, Trans-Pecos, Texas, Part 1: Vieja area: Texas Memorial Museum Bulletin 25, 42 p.
Wood, H. E., II et al., 1941, Nomenclature and correlation of the North American
continental Tertiary: Bulletin of the Geological Society of America, v. 52, p. 148.
Yanovskaya, N. M., 1980, Brontoterii Mongolii [The brontotheres of Mongolia]:
Trudy Sovmestnoi Sovyetskoi-Mongolskoi Paleontologicheskoi Ekspeditsii, v.
12, 219 p.

128

Ectoconus majusculus, new species, right manus and pes of type (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

129

THE HOLOTYPE SPECIMEN OF MENODUS GIGANTEUS, AND THE


INSOLUBLE PROBLEM OF CHADRONIAN BRONTOTHERE TAXONOMY
MATTHEW C. MIHLBACHLER1, SPENCER G. LUCAS2 AND ROBERT J. EMRY3
1

Division of Paleontology, American Museum of Natural History, Central Park West at 79th St., New York, NY 10024;
2
New Mexico Museum of Natural History, 1801 Mountain Rd. NW, Albuquerque, NM 87104;
3
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, D. C. 20560

AbstractThe holotype specimen of the brontothere Menodus giganteus Pomel, 1849, long presumed lost, is
in the Smithsonian collection. The validity of Menodus giganteus depends on an ultimate understanding of
Chadronian brontothere species-level diversity. Although Chadronian brontotheres have historically been split
into a multitude of species (N=47, by 1929), based primarily on horn variation, a preliminary analysis of
Chadronian brontothere morphometric data reveals that most of the variation in their horns can be explained by
sexual dimorphism. We are not able to objectively group Chadronian brontothere specimens into discrete morphological units with our morphometric data. Variation is apparently continuous among those variables that
were commonly used to diagnose Chadronian brontothere species. Chadronian brontotheres can be unambiguously grouped into no more than two discrete morphological units, a group with unbifurcated horns that includes
the vast majority of specimens, and a much smaller group with bifurcated horns. If these groups represent valid
taxa, Menodus giganteus is a nomen dubium. Megacerops coloradensis Leidy, 1870 is the earliest Chadronian
brontothere name whose holotype includes a pair of unbifurcated horns, and is therefore the valid name for the
more common species. The rare species with bifurcated horns is Megacerops kuwagatarhinus Mader and
Alexander, 1995.
Key words: brontothere, Menodus, Megacerops, holotype, neotype, horn, sexual dimorphism, Chadronian
INTRODUCTION
One of the first fossil mammals collected in the American West
and brought to scientific attention was a jaw fragment of a brontothere
named Menodus giganteus (Fig. 1). This specimen is the holotype of
the oldest available taxonomic name for a species of brontothere. Long
considered lost (supposedly destroyed in a fire in St. Louis in 1849),
the disappearance of the holotype prompted designation of a neotype
(Osborn, 1929; Mader, 1989). Nevertheless, the holotype of Menodus
giganteus does exist, and it has long been in the collection of the National Museum of Natural History, Smithsonian Institution, Washington, D. C. (USNM). Here, we briefly document this specimen and discuss its taxonomic significance.
HISTORY
Osborn (1929) provided a detailed (but not entirely correct) review of the history of the holotype of Menodus giganteus, briefly recounted and elaborated on here. Prout (1846, 1847) first reported on
and illustrated (Fig. 1) a jaw fragment he identified as a gigantic
Palaeotherium from what are now considered upper Eocene
(Chadronian) strata in the White River badlands of South Dakota. Pomel
(1849) realized that assignment of the fossil to the genus Palaeotherium
was unwarranted, so he coined the name Menodus giganteus for the
jaw fragment illustrated by Prout. In so doing, Pomel proposed the
first, available Linnaean binomen for a brontothere fossil. Prouts specimen, for which Pomel coined the name Menodus giganteus, is now
USNM 21820.
Owen et al. (1850) next mentioned brontothere specimens in a
preliminary report on material collected by Evans (who was employed
by Owen). In this short note, they mentioned specimens of brontotheres
thought to pertain to the same species, to which they applied the name
Paleotherium? proutii. They noted that these remarkable remains are
thus named in compliment to Dr. Prout of St. Louis, who first noticed
them in the American Journal of Science and Arts. The generic characters, however, are not yet satisfactorily decided (Owen et al., 1850, p.
66). However, the material in their possession when they coined this
name was the Owen/Evans collection, so it is not clear whether they

FIGURE 1. Prouts (1846, 1847) original illustrations of the holotype of


Menodus giganteus (from Osborn, 1929). A, Occlusal view of left m3. B,
Labial view of left dentary fragment with part m1-2 and complete m3. C,
Labial view of left m3. For scale, m3 length = 112 mm.

130
based the name Palaeotherium? proutii on this collection, or on Prouts
original specimen.
The Owen/Evans collection was subsequently turned over to
Joseph Leidy, who reported it in more detail (Leidy, 1852). Leidy noted
that in Dr. Owens collection is a fragment of the left side of a lower
jaw, apparently of the same animal that Leidy (1852, pl. 9, fig. 3)
illustrated in occlusal view. By same animal Leidy obviously meant
same kind of animal and not same individual, because he indicates
the specimen is of a smaller individual. Leidy (1852, p. 552) goes on to
state that all of the preceding specimens, except probably the latter
two [upper tooth fragments], I suspect belong to a different genus from
either Palaeotherium or Anchitherium, and should the suspicion prove
correct, Titanotherium would be a good name for this animal, as expressive of its very great size.
In 1853, Leidy again discussed this material. He thus mentioned
that Prout had described and illustrated the original jaw fragment, and
then says the specimen, which was the first fossil from the Eocene
cemetary of Nebraska, presented to the notice of the world, with another
corresponding to the opposite side, apparently of the same individual,
were kindly loaned to me by Dr. Prout for examination. (Pl. XVI, Fig. 1)
(Leidy, 1853, p. 72). However, this illustration reference is somewhat
ambiguousdid Leidy mean that figure 1 shows Prouts original specimen, or the jaw corresponding to the opposite side? Because the figure
is of a right jaw, it seems evident that Leidy meant the latter. Leidy (1853,
p. 73) again mentioned the left lower jaw in the collection of Dr. Owen
and illustrated it as Pl. XVI, Figs. 2, 3.
This makes it clear that Leidy confused specimens. Thus, the
right jaw in his plate XVI, figure 1 is very likely the opposite side of the
same individual as the Owen/Evans specimen he illustrated just below it
as figure 2 (and had also illustrated the year before), and not the opposite
side of the jaw of Prouts original specimen. Note that in 1852 Leidy had
mentioned that this left jaw was of a smaller individual than Prouts
specimen, and the size of the illustrated right jaw matches the Owen/
Evans left jaw, not Prouts left jaw.
The right jaw Leidy (1853, pl. XVI, fig. 1) illustrated is now
ANSP (Academy of Natural Sciences, Philadelphia) 12669. The left jaw
(Leidy, 1853, pl. XVI, fig. 2) is now USNM 113. Comparison of the
actual specimens by one of us (RJE) indicates that they are of identical
size, preservation and state of dental wear, and thus represent the same
individual. These were thus the two jaws Leidy (1853) was looking at
when he wrote apparently of the same individual. However, neither is
Prouts original specimen. Indeed, USNM 113 is the type specimen of
Leidys taxon Titanotherium proutii.
Osborn (1929, p. 204, fig. 159) later reproduced Leidys figure of
the right jaw (which Leidy had mistakenly referred to as the corresponding jaw of the opposite side of Prouts specimen), and Osborns
caption says Fig. 159.Type of Menodus giganteus. Prouts original
specimen. After Leidy, 1854. One-third natural size. However, inexplicably, on the previous page (Osborn, 1929, p. 203, fig. 158), Prouts
original illustration (a left jaw) is reproduced. Osborn (1929, p. 205, fig.
160) also reproduced Leidys (1852) illustration of the Owen/Evans left
jaw in occlusal view, the same specimen as illustrated by Leidy (1853, pl.
XIV, figs. 2-3).
By 1853, Leidy had received and examined fossil collections
from the White River from several sources, the major collections being
from Alexander Culbertson, from Spencer Fullerton Baird at the
Smithsonian (collected by Thaddeus Culbertson), from D. D. Owen
(collected by J. Evans), and a collection from Captain Stewart van
Vliet (sent by him to Baird, who forwarded it to Leidy), with smaller
collections from several sources, including Hiram Prout of St. Louis.
And, by this time, Leidy had confused the sources of some of this material. For example, in his 1853 monograph, Leidy cited the Van Vliet
collection as the source of the holotype of Hyracodon nebraskensis
(then Rhinoceros nebraskensis). However, Emry and Purdy (1984) cite
correspondence between Baird and Leidy that shows this specimen ac-

tually came from the Thaddeus Culbertson collection. In fact, Leidy had
already described and characterized Rhinoceros nebraskensis, and presented his paper to the Academy of Natural Sciences, about nine months
before the Van Vliet collection was forwarded to him. Emry and Purdy
(1984) also note other instances in which Leidy confused sources of
specimens in his 1853 monograph.
Marsh (1873) argued that Menodus Pomel, 1849 was preoccupied by Menodon Meyer, which is obviously incorrect. Nevertheless,
the name Menodus was then not generally used (e.g., Osborn, 1902)
until Osborn (1929) resurrected it.
Osborn (1929 p. 141-144, 202-205) reviewed the history of the
type specimen of Menodus giganteus, and was convinced the holotype
specimen had disappeared, perhaps destroyed in the great fire of St.
Louis of 1849. Therefore, Osborn (1929, p. 205) designated a neotype
for M. giganteus, a skull in the collection of the American Museum of
Natural History, AMNH 505 (Osborn, 1929, figs. 389, 391, 393, 394,
406, 444, pls. 133, 141). He then proceeded to review the morphology
of M. giganteus as a valid brontothere taxon (Osborn, 1929, p. 530537).
Osborns (1929) designation of a neotype apparently validated
Menodus giganteus as the oldest taxonomic name of a Chadronian
brontothere. Thus, those who followed, especially those who viewed
Chadronian brontothere taxonomy as greatly oversplit, considered
Menodus a valid name for many, or almost all of the Chadronian
brontotheres (e.g., Scott, 1941; Clark et al., 1967; Wilson, 1977).
Mader (1989), however, reached a different conclusion in his revision of the genus-level taxonomy of North American brontotheres. Mader
(1989) accepted Osborns conclusion that the holotype lower jaw fragment of M. giganteus was lost, but he considered Menodus sensu Osborn,
1929 to be a synonym of Menops Marsh, 1887.
HOLOTYPE
Accession record 214350 in the USNM archives includes a letter
written 1 April 1957 by Courtney Werner, Associate Professor of Geology at Washington University in St. Louis. The letter, addressed to
Remington Kellogg of the USNM staff, and an attached accession form,
indicate that the holotype of Menodus giganteus was donated by Washington University to the USNM and arrived there on 9 April 1957. The
specimen thus was not destroyed in the great fire of St. Louis, as
suggested by Osborn (1929). In addition, one of us (SGL) discovered a
painted plaster cast of the holotype, inscribed as a gift to Benjamin
Silliman from Hiram Prout, in the Yale Peabody Museum (YPM) collection. Indeed, Osborn (1929, p. 202) mentions this cast by quoting
Dana and Sillimans introduction to Prout (1846): we are also indebted
to him [Prout] for a cast of the jaw [holotype of M. giganteus], a view of
the posterior tooth of which is represented below. This cast, YPM
47285, thus could have served as a plastotype, and was available to
Osborn (1929).
The holotype of M. giganteus is USNM 21820, a left dentary
fragment with damaged m1-2 and a complete m3 (Fig. 2). The lower
molars are of typical Chadronian brontothere morphologyrectangular,
lophodont and brachydont, and the m3 has a labial cingulid. Measurements (in mm) are: m1 length ~ 59, trigonid width ~ 40, talonid width ~
47; m2 length ~ 81, trigonid width ~ 45, talonid width ~ 51; m3 length =
112, trigonid width = 50, talonid width = 47.
TAXONOMIC IMPLICATIONS
Osborns (1929) proposed designation of AMNH 505 as the
neotype of Menodus giganteus Pomel has no official taxonomic status,
as the proposal of a neotype was never acted upon by the International
Commission of Zoological Nomenclature. The existence of the holotype of M. giganteus documented here undermines the need for a neotype. Furthermore, Osborns (1929) choice of a neotype embodied a
curious error.
Thus, Osborn (1929, fig. 159) repeated an illustration from Leidy

131
TABLE 1. Statistical summary of Chadronian brontothere data.
(Abbreviations: N-number of measurements, SD-standard deviation,
CV-coefficient of variation, Min-minimum value, max-maximum
value.

N
Mean
SD
CV
Min
Max

FIGURE 2. USNM 21820, the holotype of Menodus giganteus. Occlusal (A),


labial (B) and lingual (C) views of left dentary fragment with part m1-2 and complete
m3.

(1853, pl. 16, fig. 1), claiming this illustrated the specimen Prout described, and hence the holotype of M. giganteus. He then stated that a
carefully made model, based on Leidys figures and measurements of the
lower jaw, was compared with various specimens of Menodus until an
upper dentition was found (in a skull, Am. Mus. 505) which appears to
fit very well the lower teeth of the type. Hence the skull (Am. Mus. 505)
has been selected as a neotype of Menodus giganteus.
In reality, the specimen illustrated by Leidy is not the holotype of
Menodus giganteusit is a right dentary fragment with part of m1 and
complete m2-3, not the left dentary fragment with part of m2 and all of
m3 illustrated by Prout. Leidy (1853, p. 72) referred to the fossil as
another corresponding of the opposite side, apparently from the same
individual [as the jaw fragment originally described by Prout]. However, it clearly is not the same specimen Prout illustrated. Thus, the
neotype skull designated by Osborn was not actually matched to the
holotype of M. giganteus, but to another fossil.
TOWARDS A RATIONAL TAXONOMY OF CHADRONIAN
BRONTOTHERES
The validity of Menodus giganteus as the first available name for
a Chadronian brontothere species ultimately rests upon a species-level
revision of Chadronian brontotheres, a group that, today, is almost universally recognized as severely oversplit and in need of revision, perhaps
more so than any other group of North American fossil mammals (e.g.,
Prothero, 1994; Alroy 2003). Leidy, Cope, and Marsh named Chadronian
brontothere species for nearly every new specimen that came into their

horn
length
83
162
101.2
62.4
10
380

horn
circumference
84
174
104.8
60.1
34
385

zygomatic
thickness
76
218
24.8
32.1
40
139

M1-M3
length
76
218
24.8
11.4
170
290

m1-m3
length
41
236
31.5
13.4
178
332

possession. Osborn (1896), who was the primary student of the


Brontotheriidae, believed at one point that Chadronian brontotheres belonged to a single genus, Titanotherium. Later, he considered valid an
overwhelming 37 out of 47 species that had had been named up to the
time of his monographic treatment of the Brontotheriidae (Osborn, 1929).
Chadronian brontotheres are known from numerous skeletons
and, literally, hundreds of complete skulls housed in numerous North
American museums. They vary conspicuously in the size, orientation
and shape of the frontonasal protuberances (horns), nasal bones, and
zygomatic arches, while other aspects of the skeleton are far less variable (Fig. 3). Osborn (1929) simultaneously attributed cranial variation in Chadronian brontotheres to multiple factors, including ontogeny, sexual dimorphism, and species-level differences. He did not, however, convincingly demonstrate how he differentiated, for instance, horn
variation due to sexual dimorphism versus horn variation due to speciation.
Osborns obviously dubious taxonomy was rejected, even by his
peers. For instance, in his review of Chadronian brontotheres, Scott
(1941, p. 908) stated ...it is hardly worth while even to list the species; that 37 species could not have co-existed within the relatively
brief space of the Chadron, is obvious; how many did exist and what
names should be given to them, are insoluble problems. Scott (1941,
p. 873) also concluded the probably great effects of sex and age and
fluctuating variability have not been sufficiently evaluated. This statement is as true today as it was in 1941.
Despite the obvious need for a revision of Chadronian brontothere
taxonomy, no such overhaul has taken place, although some have voiced
opinions. Clark et al. (1967) suggested that all Chadronian brontotheres
belong to a single, highly variable species, for which Menodus giganteus
could serve as the valid name. This conclusion was based on four horn
cores of variable morphology that were collected from a single graveyard in a layer 1 ft. thick and within 30 ft. of each other horizontally
(p. 50). Mader (1989), on the other hand, concluded that three genera
are validMenops, Brontops, and Megacerops. Mader (1989) based
these genus-level distinctions primarily on the cross-sectional shape of
the horns, which he regarded as an apomorphic condition that is diagnostic at the genus level (p. 477). In a further revision of just one of
these genera, Megacerops, Mader and Alexander (1995) accepted four
species and stated that specimens (attributed to Megacerops) constituted a monophyletic taxon that is characterized by dorsoventrally compressed wing-shaped zygomatic arches and by long laterally or anteriorly directed horns that are round to elliptical in cross section (p. 585).
If Maders (1989) and Maders and Alexanders (1995) revisions are accepted, Menodus giganteus would clearly be a dubious name,
given that the holotype (a mandible fragment) retains no diagnostic
morphology that would validate a genus (horn shape). However, we do
not accept these revisions because the application of horn shape, or
zygomatic shape, as genus-level traits is entirely conjectural at this point.
These authors undertook no morphometric analysis to demonstrate that

132

FIGURE 4. Bivariate plots of Chadronian brontothere cranial data.

FIGURE 3. A series of Chadronian brontothere skulls, from Osborn (1896). These


specimens demonstrate the wide range of cranial variation, which is most conspicuous
in horn size and shape, and thickness of the zygomatic arches.

these variables could be used to group specimens into discretely diagnosable taxa. Moreover, there has been no phylogenetic analysis that would

support their contention that these characters are apomorphies of higher


level, monophyletic groups (genera).
Nevertheless, we do agree with Mader (1989) that the degree of
variation among Chadronian brontothere fossils seems greater than that
which is generally acceptable for a single species, but it is almost impossible to objectively group the vast majority of the specimens into
discrete morphological units (taxa). One of us (MCM) has measured a
large number of Chadronian brontothere skulls and mandibles from
most of the major North American collections. These data are ultimately intended for a more sophisticated morphometric analysis, but a
simplified analysis is presented here. Table 1 contains the statistical
results for selected Chadronian brontothere variables. Horn length, horn
circumference, and zygomatic thickness are extremely variable, and
yield coefficients of variation ranging between 30 and 63. These range far
above the normal expectation of mammal intraspecific variability (cv <
10). However, the upper and lower molar row lengths fall much closer to
the expected degree of intraspecific variation (cv = 11.4 and 13.4, respec-

133
TABLE 2. Statistical summary of Giraffe (Giraffa camelopardalis)
cranial morphometric data (in mm). Abbreviations are as in Table 1.

N
Mean
SD
CV
Min
Max

FIGURE 5. Bivariate plot of adult giraffe (Giraffa camelopardalis) horn length


versus tooth row length.

tively).
The pattern of variability among Chadronian brontotheres resembles those of sexually dimorphic mammal species. For comparison, Menoceras arikarense, a small Miocene North American rhino,
shows bimodal sexual dimorphism in the nasal horn bosses and nasal
bones (Mihlbachler, unpublished data). The coefficients of variation
for horn and nasal characters in that species is as high as 51.3. However, in the same species, non-sexually dimorphic variables (e.g., toothrow length) always yield coefficients of variation below 10.
The disparity in the magnitude of variation in horn, zygomatic,
and dental variables suggests that much of the variation among
Chadronian brontotheres can be explained by sexual dimorphism. Additionally, it is also plausible that some of the variation in horn length
and zygomatic thickness could be ontogenetic. To test for an ontogenetic effect on horn length, we plotted that variable against a series of
dental wear stages illustrated by Osborn (1929, fig. 387). (The stages
used here, 10-15, are all adult stages; juvenile brontothere specimens
are very rare). The plot (Fig. 4A) does not suggest a significant ageeffect on horn length among adults, nor was a least squares regression
of these variables significant (R squared = 0.17, P = 0.23).
To examine the possibility of sexual dimorphism, horn length
was plotted against molar row length (Fig. 4B). The least squares regression of horn length and toothrow length is marginally significant
(R squared = 0.84, P = 0.04). Among significantly sexually dimorphic
mammals, including sexually dimorphic perissodactyls such as the white
rhino (Ceratotherium simum), body size and secondary sex traits (e.g.,
horn size) are larger in males (Owen-Smith, 1988; Rachlow and Berger,
1997). The pattern among Chadronian brontotheres is consistent with
the generalized ungulate pattern in which larger individuals (represented by longer tooth-rows) have larger horns. The third plot, of two
highly variable characters, horn length and zygomatic thickness (Fig.
4C), shows a highly significant correlation (least squares regression: R
squared = 0.66, P = 0.00). Although the high degree of variability in
horn size and zygomatic thickness is consistent with sexual dimorphism,
there are no obvious bimodal patterns that one might interpret as male
and female size clusters in the Chadronian brontothere data.
Ideally, sexual dimorphism should yield bimodal size clusters,
but the absence of bimodality, as is the case here, does not eliminate
the possibility of sexual dimorphism. For instance, supposing that the
horns and zygomatics are dimorphic, the lack of apparent bimodality
could result from two or more differently sized taxa mixed in the data,
thus smearing out any detectible intraspecific bimodality. However, this
is not necessarily the case. There are instances among extant mammals
where a sexually dimorphic character lacks clear bimodality in a single

ossicone
length
29
144
39.6
27.6
80
224

ossicone
circumference
29
134
45.5
34.0
63
270

toothrow
length
29
152
9.5
6.2
127
168

species. Among these are the ossicones of the giraffe (Giraffa


camelopardalis).
Perhaps, the best analogies for brontothere horns are giraffe
ossicones (Fig. 6). Chadronian brontothere horns and modern giraffe
ossicones share a combination of features not seen in other horned
mammals. Both have horns that are composed entirely of
nondeciduous bone and that are morphologically unelaborated. More
important for the argument presented below, they tend to develop conspicuous roughened surfaces on the distal ends. Frequently, in
Chadronian brontotheres, the rugosities extend down the shaft and continue around the nasal process and to the orbits. Likewise, the rugosities of giraffe ossicones often spread over the entire calvarium. Occasionally, there are small bony spurs on the medial surface of the horn
shaft in giraffes and Chadronian brontotheres (e.g., the right giraffe
skull in fig. 6). Giraffe ossicones are covered with skin, a condition
likely shared by all horned brontotheres.
A statistical summary of selected variables for adult giraffe skulls
is given in Table 2. Adults were defined as having a fully erupted M3,
showing some amount of wear. The pattern of variability is similar to
that of Chadronian brontotheres. The ossicones are extremely variable,
whereas tooth-row length is less variable. The plot of horn length and
tooth-row length (Fig. 5) shows a significant correlation (least squares
regression: R squared = 0.46, P = 0.01). Like Chadronian brontotheres,
there is no apparent bimodality in the data. Nonetheless, males giraffes
tend to have larger, more rugose horns than those of females (Dagg and
Foster, 1976).
Giraffe ossicones, particularly those of males, are subject to extensive secondary bony growth due to combat behaviors involving sideto-side head clashes in the regions of the orbits, occiputs, and ossicones
(Solounias and Tang, 1990). The force of impact damages bone cells
and stimulates the perisoteum to form secondary growth (Solounias
and Tang, 1990). Because this process is largely epigenetic, we postulate that the absence of clear bimodality in this sexually dimorphic
structure is related to the extreme epigenetic variation, particularly
among males, caused by secondary bone growth, obscuring any potential gene-based bimodality. The variably roughened surfaces of
Chadronian brontothere horns most likely indicate secondary bone
growth, due to head clashes of some sort. The extremely thickened
and laterally expanded zygomatic arches are consistent with some form
of head clashing as well. Clearly, a histological analysis of brontothere
horn tissue would further address the possibility of secondary bone
growth in brontothere horns. Nonetheless, the example of the giraffe
provides a strong case for concluding that phenotypic sexual dimorphism can exist without clear bimodal clustering. It is therefore plausible that the entire sample of Chadronian brontothere fossils may represent a single, sexually dimorphic species.
Given that the Chadronian brontothere cheek-tooth data (cv ~1113) are somewhat more variable than is typical for a non-dimorphic trait
(cv <10), it is alternatively possible that the sample may contain multiple species of overlapping body size, of which one, or all, are sexually

134

FIGURE 6. Selected adult giraffe (Giraffa camelopardalis) skulls in the AMNH collection. Scale bar = 5 cm.

dimorphic. However, the data presented here are continuous, and do not
provide an objective means for dividing specimens into discrete units.
Therefore, one could only conclude from this analysis that there is, at
most, one unambiguously diagnosable Chadronian brontothere taxon.
If there were only a single species, Menodus giganteus would
clearly be the valid name. However, recent discoveries have introduced
additional complexity to the problem of Chadronian brontothere taxonomy. A small number of Chadronian brontothere skulls with bifurcating horns (N=2 and some other horn fragments) were given a new
species name, Megacerops kuwagatarhinus Mader and Alexander, 1995.
An additional skull with bifurcating horns was reported by Ervin (2001).
These unusual specimens are clearly separable from the vast pool of
available Chadronian brontothere skulls whose horns are not bifurcated.
Additionally, the only complete molar known from this species, an M3
belonging with the holotype skull in the Frick Collection of the Am.
Mus. (F:AM128600), exhibits several peculiar traits not seen in typical Chadronian brontotheres, such as a small paraconule and a large
metaconule (Mader and Alexander, 1995). In addition, the ectoloph
seems unusually brachydont when compared to typical Chadronian
brontothere molars.
Given the rarity of bifurcating horns among Chadronian
brontothere skulls, these specimens may only represent atypical individuals of Menodus giganteus rather than a separate species. Several
brontothere skulls have small bony spurs or hornlets on the medial side
of the body of the horn (similar to that seen on the right giraffe skull of
fig. 6). The more obvious bifurcations of the rare Megacerops
kuwagatarhinus skulls may represent the extreme form of this condition in Menodus giganteus. However, the unusual third molar of the
holotype of M. kuwagatarhinus more compellingly suggests a distinct
species, but it has not been corroborated in other M. kuwagatarhinus
specimens, in which no teeth are preserved.
The validity of Megacerops kuwagatarhinus is ultimately contingent upon which species concept is applied. The application of a
biological species concept (BSC) (e.g., Mayr and Ashlock, 1991), where
emphasis is on population thinking rather than typological thinking, allows one to conjecture that those specimens with bifurcating
horns are rare variants of Menodus giganteus. Ultimately the validity
of this species under a BSC is a matter of opinion, since one could also
conjecture the opposite, that M. kuwagatarhinus represents an isolated
breeding population. In contrast, the application of species concepts
more amenable to cladistic analysis, such as the phylogenetic species
concept (e.g., Cracraft, 1989), where, in operation, fossils are partitioned
into the smallest diagnosable units (Nixon and Wheeler, 1992), there are
two discretely diagnosable species: a common species without bifurcat-

ing horns, and a rare species with bifurcating horns.


If multiple species are accepted (as we recommend), Menodus
giganteus, whose holotype is a mandible fragment, is presently a nomen
dubium, because the lower dentition M. kuwagatarhinus is unknown.
The morphology of the Menodus giganteus holotype is typical for
Chadronian brontotheres, however, if future discoveries demonstrate
that the dentition of M. kuwagatarhinus is differentiated, as the single
upper M3 of its holotype suggests as a possibility, Menodus giganteus
would again, become valid.
Nonetheless, if two species are accepted, Menodus giganteus, is
presently dubious. All other early Chadronian brontothere names, mostly
named by Leidy between 1850 and 1854, are based on premolars and
molar fragments that are as equally dubious as Menodus giganteus (see
Osborn, 1929 for summary). Megacerops coloradensis Leidy 1870, is
the first Chadronian brontothere species whose holotype, at the Academy of Natural Sciences in Philadelphia (ANSP 13362), includes a
pair of unbifurcated horns. Megacerops coloradensis, unambiguously
represents the more common species with unbifurcating horns.
Megacerops kuwagatarhinus remains the valid name for the species
with bifurcated horns.
One final precautionary comment should be added regarding
Megacerops coloradensiswe cannot rule out the possibility that one
or more brontothere species experienced anagenic evolution during the
Chadronian land-mammal age. Unfortunately, many of the best specimens, particularly those from older collections, have poor stratigraphic
data. The sample used here, which includes only specimens with
unbifurcated horns, spans the better part of the Chadronian. An
anagenically evolving species would account for some of the variability in this time-averaged sample. We can also postulate a scenario where
two or more anagenically evolving species, that at any instant were
morphologically distinct, may not be recognizable in a time-averaged
assemblage. We are not currently prepared to fully address the additional possibility of anagenesis, although a large collection of Chadronian
brontothere material with excellent stratigraphic data in the Frick Collection of the American Museum of Natural History, which still resides
in plaster jackets, may provide significant inroads into the problem of
Chadronian brontothere diversity and evolution. However, when one
considers the facts that Chadronian brontotheres are presently known
from hundreds of skulls that are each distorted to varying degrees (thus
resulting in possible taphotaxa sensu Lucas [2001]), and that the morphometric variation among them is apparently continuous, we may ultimately have to accept Scotts (1941) contention that the true number of
Chadronian brontothere species is one of paleontologys insoluble problems.

ACKNOWLEDGMENTS
Bob Purdy assisted us with Smithsonian archives. Jin Meng and
Chris Norris (American Museum of Natural History, Paleontology),
Jeanne Spence (American Museum of Natural History, Mammalogy),
Logan Ivy (Denver Museum of Nature and Science), Pat Holroyd (University of California Museum of Paleontology), Alan Tabrum and Chris

135
Beard (Carnegie Museum of Natural History), Jaelyn Eberle (University
of Colorado Museum), and Dan Brinkman (Yale Peabody Museum)
allowed MCM to examine Chadronian brontothere specimens in their
care. Carl Mehling and Robert Evander assisted with moving and repairing several large American Museum brontothere specimens that made
this study possible. P. Kondrashov provided helpful comments on the
manuscript.

REFERENCES
Alroy, J., 2003, Taxonomic inflation and body mass distributions in North American fossil mammals: Journal of Mammalogy, v. 81, p. 431-443.
Clark, J., Beerbower, J. R. and Kietzke, K. K., 1967, Oligocene sedimentation, stratigraphy, paleoecology and paleoclimatology in the Big Badlands of South Dakota: Fieldiana: Geology Memoirs 5, 158 p.
Cracraft, J., 1989, Species and it ontology: the empirical consequences of alternative species concepts for understanding patterns and processes of differentiation;
in Otte, D., and Endler, J.A., eds., Speciation and its consequences: Sunderland,
MA, Sinauer Associates, p. 28-59.
Dagg, A. I. and Foster, J. B., 1976, The giraffe: its biology, behavior, ecology. New
York: Van Nostrand Reinhold.
Emry, R. J., and Purdy, R. W., 1984, The holotype and would-be holotypes of
Hyracodon nebraskensis (Leidy, 1850): Notulae Naturae of The Academy of
Natural Sciences of Philadelphia, no. 460, p. 1-18
Ervin, P., 2001 Megacerops kuwagatarhinus skull, from Chadronian land mammal age, doubles this late Eocene species geographic range. Journal of Vertebrate Paleontology, suppliment to volume 21, p. 48A
Leidy, J., 1852, Description of the remains of extinct Mammalia and Chelonia from
Nebraska Territory, collected during the geological survey under the direction
of Dr. D. D. Owen; in Owen, D. D., Report of a geological survey of Wisconsin,
Iowa, and Minnesota and incidentally a portion of Nebraska Territory: Philadelphia, p. 533-572.
Leidy, J., 1854, The ancient fauna of Nebraska, or a description of remains of extinct Mammalia and Chelonia from the mauvaises terres of Nebraska:
Smithsonian Contributions to Knowledge, v. 6 (7), 126 p.
Lucas, S. G., 2001, Taphotaxon: Lethaia, v. 34, p. 30.
Mader, B. J., 1989, The Brontotheriidae: A systematic revision and preliminary
phylogeny of North American genera; in Prothero, D. R. and Schoch, R. M.,
eds., The evolution of perissodactyls: Oxford, Oxford University Press, p. 458484.
Mader, B. J., and J. P. Alexander, 1995, Megacerops kuwagatarhinus n. sp., an
unusual brontothere (Mammalia, Perissodactyla) with distally forked horns:
Journal of Paleontology, v. 69: 581-587.
Marsh, O. C., 1873, Notice of new Tertiary mammals: American Journal of Science, v. 5, p. 407-410, 485-488.
Marsh, O. C., 1887, Notice of new fossil mammals: American Journal of Science,
third series, v. 34, p. 323-331.

Mayr, E., and P. D. Ashlock, 1991, Principles of Systematic Zoology: New York,
McGraw-Hill, Inc.
Nixon, K. C., and Q. D. Wheeler, 1992, Extinction and the origin of species: in
Wheeler, M. J. and Wheeler Q. D., eds., Extinction and phylogeny: New York,
Columbia University Press, p. 119-143.
Osborn, H. F., 1896. The cranial evolution of Titanotherium: Bulletin of the American Museum of Natural History, v. 8: 157-197.
Osborn, H. F., 1902, The four phyla of Oligocene titanotheres: Bulletin of the American Museum of Natural History, v. 16, p. 91-109.
Osborn, H. F., 1929, The titanotheres of ancient Wyoming, Dakota, and Nebraska:
U. S. Geological Survey, Monograph 55, 953 p.
Owen, D. D., Norwood, J. G. and Evans, J., 1850, Notice of fossil remains brought
by Mr. J. Evans from the Mauvais Terres, or bad lands of White River, 150
miles west of the Missouri: Proceedings of the Academy of Natural Sciences, v.
5, p. 66-67.
Owen-Smith, R. N., 1988, Megaherbivores: the influence of very large body size on
ecology. Cambridge University Press, Cambridge.
Pomel, A., 1849, Description dun os maxillaire fossile de Palaeotherium par Hiram
Prout, Am. Jour. Sci. and arts by Sillimans and J. Dana, 2e srie, vol. 3, no. 8,
p. 248: Bibliothque Universelle de Geneve, Archives des Sciences Physiques
et Naturelles, v. 10, p. 73-75.
Prothero, D. R., 1994, The Eocene-Oligocene transition: paradise lost, New York,
Columbia University Press.
Prout, H. A., 1846, Gigantic Palaeotherium: American Journal of Science, second
series, v. 2, p. 288-289.
Prout, H. A., 1847, Description of a fossil maxillary bone of a Palaeotherium from
near White River: American Journal of Science, second series, v. 3, p. 240-250.
Rachlow, J. L., and Berger, J., 1997, Conservation implications of patterns of horn
regeneration in dehorned white rhinos: Conservation Biology, v. 11, p. 84-91.
Scott, W. B., 1941, The mammalian fauna of the White River Oligocene. Part V
Perissodactyla: Transactions of the American Philosophical Society, v. 28, p.
747-980.
Solounias, N. and Tang, N., 1990, The two types of cranial appendages of Giraffa
camelopardalis (Mammalia, Artiodactyla): Journal of Zoology, London, v. 222,
p. 293-302.
Wilson, J. A., 1977, Early Tertiary vertebrate faunas Big Bend area Trans-Pecos
Texas: Brontotheriidae: The Pearce-Sellards Series, no. 25, 17 p.

136

Loxolophus hyattianus, skull and lower jaw (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

185

LATE PALEOCENE (GASHATAN) NYCTITHERIIDAE


(MAMMALIA, LIPOTYPHLA) FROM MONGOLIA
PETER E. KONDRASHOV1,2, ALEXEY V. LOPATIN2 AND SPENCER G. LUCAS3
1

Department of Biology, Northwest Missouri State University, 800 University Drive, Maryville, MO 64468;
2
Paleontological Institute of RAS, Profsoyuznaya, 123, Moscow, Russia, 117868;
3
New Mexico Museum of Natural History, 1801 Mountain Road NW, Albuquerque, NM 87104

AbstractNew specimens of Nyctitheriidae from the Mongolian localities Tsagan-Khushu and Naran-Bulak
are described. The material comes from the Zhigden Member of the Naran-Bulak Formation, which is thought
to be latest Paleocene (Gashatan) in age. It includes several dentulous upper and lower jaws. Three taxa, all
extremely rare, are identified: Praolestes nanus, Praolestes maximus sp. nov. and Jarveia erronea sp. nov.
Previously, Praolestes nanus was known only by the type specimen with p3-m1 and was thus one of the most
problematic late Paleocene Asian mammals. The new material documents the dental morphology of the genus,
thus helping to define its systematic position as a nyctitheriid insectivore and to allow comparison with other
nyctitheriids. The new species of Praolestes resembles the type species, but differs in larger size and in the m3
structure. Both species are closer to the genus Bumbanius, known from the lower Eocene of Tsagan-Khushu,
than to other nyctitheriids. The new species of Jarveia is represented by both upper and lower teeth. The
peculiar dental morphology of this small species is most similar to Jarveia minuscula from the latest Paleocene
of Kazakstan. Praolestes nanus is an index fossil of the Gashatan land-mammal age, so the time range of
Nyctitheriidae in Mongolia extends back to the late Paleocene.
Key words: Nyctitheriidae, Praolestes, Jarveia, Mongolia, Paleocene, Gashatan

INTRODUCTION
For many years, nyctitheriid insectivores were known only from
the Upper Cretaceous and Paleogene of North America (Simpson, 1928;
McKenna, 1968; Fox, 1970; Krishtalka, 1976) and the Paleogene of
Western Europe (Sig, 1976). The first Asian nyctitheriids were described by Russell and Dashzeveg (1986), who documented a rich fauna
of insectivores from the lower Eocene deposits of Tsagan-Khushu,
Mongolia. Among four newly described genera, two appeared to be
nyctitheriids: Bumbanius and Oedolius. Later, Averianov (1995) redescribed Nessovs proteutherian Jarveia from the late Paleocene of
Kazakstan as a nyctitheriid insectivore. Meng et al. (1998) described a
late Paleocene nyctithere, Bayanulanius, from the Bayan Ulan locality
of Inner Mongolia, China, based on two maxillary fragments and one
dentary fragment.
Praolestes nanus, based on a dentary fragment, was described
from the upper Paleocene Gashato Formation (Mongolia) as ?Leptictidae
(Matthew et al., 1929). In a brief discussion, Matthew et al. said that
Praolestes appeared to be a leptictid, although they mentioned that the
m1 of Praolestes is more similar to the p4 of Leptictidae. Van Valen
(1967) placed Praolestes in Geolabidinae without discussion. Szalay and
McKenna (1971) redescribed Praolestes and assigned it to
Zalambdalestidae, as a possible descendant of Zalambdalestes. Russell
and Dashzeveg (1986) noted the similarities in dental morphology between Praolestes and Bumbanius, an early Eocene nyctitheriid, but did
not place the former directly in the Nyctitheriidae. Carroll (1988) regarded Praolestes as Anagalida fam. indet. McKenna and Bell (1997)
placed Praolestes in Leptictida inc. sedis. Ting (1998) included Praolestes
in Zalambdalestidae, whereas Meng et al. (1998) listed it as Insectivora.
So, Praolestes has been assigned to several mammalian groups, and its
systematic position has been uncertain, largely because of the scarcity of
material. Here, we describe new specimens of Praolestes nanus and two
new species of Nyctitheriidae: P. maximus sp. nov. and Jarveia erronea
sp. nov.
PROVENANCE
The new specimens described here come from the lower red hori-

zon of the Zhigden Member of the Naran-Bulak Formation at the TsaganKhushu and Naran-Bulak localities. The Tsagan-Khushu and NaranBulak localities are situated in the Nemegt basin of southern Mongolia.
Badamgarav and Reshetov (1985) described the section of Tsagan-Khushu
and Naran-Bulak, listed the known mammalian remains and regarded the
Zhigden Member as late Paleocene. We identified Pseudictops lophiodon,
Palaeostylops iturus and Hyracolestes ermineus in our collection from
the Zhigden Member, which are index fossils of the Gashatan landmammal age and correspond to the late Paleocene (Ting, 1998). The
specimens described here were collected by V.Yu. Reshetovs field party
of the Joint Soviet-Mongolian expedition of 1980, 1985 and 1987.
Institutional abbreviations: AMNH = American Museum of
Natural History, New York; PIN = Paleontological Institute of the Russian Academy of Sciences, Moscow.
SYSTEMATIC PALEONTOLOGY
Order Lipotyphla Haeckel, 1866
Suborder Soricomorpha Gregory, 1910
Family Nyctitheriidae Simpson, 1928
Subfamily Nyctitheriinae Simpson, 1928
Genus Praolestes Matthew, Granger and Simpson, 1929
Praolestes Matthew, Granger and Simpson, 1929, p. 3.
Praolestes: Saban, 1958, p. 875.
Praolestes: Szalay and McKenna, 1971, p. 301.
Praolestes: McKenna and Bell, 1997, p. 103.
Type speciesPraolestes nanus Matthew, Granger and Simpson,
1929.
Included speciesThe type species and P. maximus sp. nov.
Revised diagnosisSmall nyctitheriid, most similar to
Bumbanius Russell and Dashzeveg, 1986, from the lower Eocene of
Mongolia, but differs in the p4 structure, which is more elongated in
Bumbanius with a large, anteriorly shifted paraconid and also in having a more prominent, posteriorly projecting m3 hypoconulid. The upper teeth of Praolestes are wider and shorter than in Bumbanius and
have a weaker hypocone shelf. Differs from Jarveia Nessov, 1987 in

186
being larger, and having more anteroposteriorly compressed lower molar
trigonids with well expressed cristid obliqua, para- and metacristids.
The upper molars of Praolestes differ from those of Jarveia in having
well-developed conules, a larger parastyle, wider stylar shelf and weaker
hypocone shelf. Differs from Oedolius Russell and Dashzeveg, 1986 in
having more anteroposteriorly compressed trigonids, a wider and less
open talonid basin and a less expressed labial notch between the trigonid and the talonid. Lower molars of Praolestes differ from those of
Bayanulanius Meng et. al., 1998 in having considerably higher trigonids
and rectangular talonid basins. Upper molars differ in their less reduced M3 without the hypocone and more separated paracone and
metacone on M1-3. Differs from European Saturninia Stehilin, 1940
in having an unreduced p3, shorter lower molar trigonids and in upper
molar structure. Differs from North American Nyctitheriinae in its more
anteroposteriorly compressed, columnar lower molar trigonids, wider
and shorter upper molars, and less developed hypocones and stylar
shelves.
DistributionUpper Paleocene (Gashatan) of Mongolia.
Praolestes nanus Matthew, Granger and Simpson, 1929
Figs. 1-4
Praolestes nanus Matthew, Granger and Simpson, 1929, p. 3, fig. 1
Praolestes nanus: Szalay and McKenna, 1971, p. 301, figs. 19-23.
HolotypeAMNH 21718, incomplete left dentary with p3-m1
and alveolus for p2, talonid of m1 damaged (Fig. 1a, b).
Type localityGashato, southern Mongolia, latest Paleocene.
Diagnosis30 % smaller than P. maximus sp. nov., with no
postcingulid on the m3.
DistributionUpper Paleocene of the Tsagan-Khushu and
Gashato localities, Mongolia.
Referred specimensFrom the Zhigden Member of the TsaganKhushu locality: PIN 3104-360, fragment of the right dentary with m12 and alveoli for i1-3, c and p1-4 (Fig., 3a-c); 3104-361, fragment of
the right dentary with m1-3 (Fig. 1c, d); 3104-362, fragment of the
right dentary with m2-3 and coronoid process (Fig. 3d); 3104-427, fragment of the right dentary with i3-m2 (Fig. 1e, f, Fig. 2); 3104-433, left
maxillary fragment with M1-3 (Fig. 1g, Fig., 3e).
Description The mandible is low, not curved ventrally. The
posterior mental foramen is situated under the posterior root of the p3
at the midheight of the mandible. An interesting structure of the mental foramina is observed on PIN 3104-360 (Fig. 3a). On this specimen,
the posterior foramen is fissure-like, elongate and smile-shaped. There
is no anterior foramen on this specimen, but there is a shallow, round
fossa in its place (under the middle of the p2) that is connected to the
posterior foramen by a shallow groove. So, probably on this specimen
the blood vessel emerged from the posterior mental foramen, went anteriorly on the surface of the bone along the groove and only then went
to the soft tissues of the lower jaw. Such a structure of the mental
foramen is often observed in nyctitheriids (Sig, 1976; Nessov, 1987;
Averianov, 1995). However, the shape and placement of the mental
foramen in the holotype of Praolestes and in PIN 3104-427 (Figs. 1a, e,
2a) are normal for eutherians: the posterior mental foramen is round
in shape, and there is a well-developed anterior foramen. In the dentary,
PIN 3104-427, the anterior mental foramen is situated under the p1,
posterior to the anterior root of the p4. So, the observed aberrance of
the structure of the mental foramina in PIN 3104-360 may be typical,
but not obligatory for nyctitheriids.
The symphysis is short, and it reaches the posterior edge of the
alveolus of the canine. The masseteric ridge is strong, and the masseteric fossa is rather deep. The angle between the horizontal ramus and
the coronoid process is ~ 125o. Judging from the alveoli, all the teeth
were closely spaced without any diastemata.
The lower tooth formula is i3 c1 p4 m3. The first four alveoli are
inclined forward at 45o. The first three are interpreted as incisor alveoli,

FIGURE 1. Praolestes nanus, a, b, holotype, AMNH 21718, left dentary with p3m1: a, labial view, b, occlusal view; c, d , PIN 3104-361, right dentary with m1-3:
c, labial view, d, occlusal view; e, f , PIN 3104-427, right dentary with p1-m2: e,
labial view, f, occlusal view; g, 3104-433, left maxillary fragment with M1-3, occlusal
view.

and the fourth, which is rather large, has a protruding labial wall, is round
in shape and should be for the canine. The alveolus of the first incisor is
larger than the other two, so, the first incisor probably was the largest
and somewhat equal in size to the canine. The p1 alveolus is small, round
in shape, single and closely appressed to the alveolus of the anterior root
of the p2. The p2, p3 and p4 alveoli are double. They are elongate,
slightly widen anteroposteriorly and are oval in shape.
All the premolars and first two molars are preserved in the dentary
PIN 3104-427 (Figs. 1 e, f, 2). Two incisor-like teeth were found next to
this dentary, so we assign them to P. nanus. One of the teeth is much
larger than the other, and based on the comparison with the dentary of
Praolestes maximus sp. nov., PIN 3102-75, and according to the alveoli
size in PIN 3104-360 and 3104-427, we consider the larger tooth to be
the canine and the smaller the i3. The height of the canine, including the
root, is about 30 % more than that of the i3. The i3 crown is simple,
spatulate and tapers slightly apically. The wear facet is elongate and oval
in shape. The canine crown is inclined anteriorly. The root is slightly
curved anteriorly. Laterally, the crown is almost rectangular in shape.
There is a small posterior tubercle at the base of the crown. An oval
apical wear facet extends anteroposteriorly.
The p1 is single-rooted, with a single tall cuspid and a very weak
posterior cingulid. The p2 is larger in size, and the posterior cingulid

187

FIGURE 2. Praolestes nanus, PIN 3104-427, right dentary with i3-m2 (isolated i3
and c): a, labial view, b, occlusal view, c, lingual view.

forms a tiny talonid. The p3 has a vestigial paraconid closely appressed


to the base of the protoconid. The metaconid is present as a thickening
on the posterolingual wall of the protoconid. The talonid is small, but
well expressed. It bears a medial anteroposterior cristid. The p4 is rather
molarized. The trigonid is formed by a large protoconid, small paraconid
and larger metaconid. The paraconid is not appressed to the protoconid,
but situated anterolingual to the latter. The paracristid is rather weak.
The metaconid is situated directly lingual to the protoconid, and the
metacristid is well expressed. The wear facets on all the premolars
descend along the posterior slope of the protoconid (or the trigonid, if
present).
The lower molars are inclined slightly lingually. The m1 and m2
are almost equal in shape, and the talonid is wider than the trigonid
because of the labially protruding hypoconid. The lingual wall of the
lower molars is almost straight, whereas the labial is much curved,
with a well defined notch between the trigonid and the talonid. The
lower molar trigonid is twice as high as the talonid.
The m1 is much narrower than the m2. The paraconid is distinct, separate, projects anteriorly and is situated more medially than
on the m2. It is fused with the metaconid on the worn teeth. The
paraconid is connected to the protoconid by a weak paracristid. The
precingulid is well developed and runs from the middle width of the
crown to its anterolabial edge; it is separated from the paraconid by a
well-developed anterior groove. The protoconid and the metaconid are
united at the base up to three quarters of the height of the metaconid
and form a column. The metacristid is weak. The metaconid is slightly
taller than the protoconid, but it is reversed on the worn teeth. The
trigonid basin is open lingually, well defined, but shallow on the unworn teeth, and disappears even on slightly worn teeth. The talonid is
a little compressed anteroposteriorly. The cristid obliqua begins from
the middle of the posterior wall of the trigonid. The talonid basin is
wide and almost closed lingually by a weak entocristid extending from
the entoconid, but not reaching the metaconid. The entoconid is much
smaller than the hypoconid and slightly shifted posteriorly. The
hypoconulid is weak and is not observed on the worn teeth; the hypo-

FIGURE 3. Praolestes nanus, a-c, PIN 3104-360, right dentary with m1-2 and
alveoli for i1-3, c and p1-4: a, labial view, b, occlusal view, c, lingual view; d, PIN
3104-362, right dentary with m2-3 and coronoid process, labial view; e, PIN 3104433, left maxillary fragment with M1-3, labial view.

conid and the entoconid are united by a narrow and low, slightly curved
postcristid. The postcingulid is rather weak.
The m2 is similar to m1 in morphology, but differs in details: both
the trigonid and talonid are slightly wider, the paraconid is not so shifted
lingually and is situated more labially than on the m1. The talonid basin
is almost closed lingually by a low and thin entocristid, which is weaker
than that on the m1. The entoconid is larger and not shifted posteriorly,
so the postcristid is better developed than on the m1 and more curved
posteriorly. The hypoconulid is better expressed than on the m1.
The m3 has a trigonid very similar to the m2, with a slightly more
reduced paraconid. The talonid is much narrower than the trigonid, and
the hypoconulid is well defined, projecting posteriorly.
Based on size and occlusion pattern, we refer a left maxillary
fragment with M1-3, PIN 3104-433 (Figs. 1g, 3e, 4a), to Praolestes
nanus. The upper molars are low-crowned and transversely elongate.
The M1 and M2 are slightly curved anteriorly. The paracone is larger
and taller than the metacone on the M1. These two cusps are connected
at the base by a weak centrocrista. There is a weak postmetacrista on
the posterior slope of the metacone that runs to the moderately developed metastyle. The labial cingulum is weak and does not form a distinct stylar shelf between the parastylar and metastylar lobes. There is
no mesostyle. The parastyle is larger than the metastyle, projects anteriorly, and the preparacrista is absent. The ectoflexus is slightly expressed. The protocone is sharp; transversely, its apex is situated at the
level of the posterior slope of the paracone. The preprotocrista and
postprotocrista are sharp. They connect the protocone with the

188
TABLE 1. Mandible depth of Praolestes nanus (in mm).
Coll. Number
PIN 3104-360
PIN 3104-427

p2
2.4
2.3

p3
2.6
2.7

p4
2.7
2.75

m1-2
2.95
2.6

TABLE 2. Measurements of the lower teeth of Praolestes nanus (in


mm).
Coll. Number

Tooth

PIN 3104-360

m1
m2
m1
m2
m3
m2
m3
p1
p2
p3
p4
m1
m2

PIN 3104-361

PIN 3104-362
PIN 3104-427

Length Trigonid Talonid


width
width
1.8
1.3
1.3
1.8
1.45
1.4
1.7
1.3
1.13
1.65
1.3
1.17
1.8
1.3
2.0
1.5
1.3
2.25
1.35
1.1
0.75
0.55
0.95
0.7
1.55
0.75
1.75
0.9
1.85
1.15
1.1
2.0
1.4
1.2

TABLE 3. Measurements of the upper teeth of Praolestes nanus (in


mm).
FIGURE 4. Praolestes nanus, a, PIN 3104-433, left M1-3, occlusal view; b, PIN
3104-362, m2-3, occlusal view (reversed).

paraconule and the metaconule and form a relatively elongate and deep
trigon basin. The conules are distinct and almost equal in size, or the
paraconule is slightly larger. Each conule has two labial crests. The
preparaconule crest ends at the anterolingual base of the paracone. The
postparaconule crest is indistinct on the M1 due to wear. The
premetaconule crest extends to the anterolingual base of the metacone,
whereas the postmetaconule crest reaches the posterolingual base of
the metastyle. The precingulum is weak, short and situated at the anterior base of the protocone. The hypocone shelf is much stronger than
the precingulum and does not form a distinct hypocone. The hypocone
shelf extends labially to the labialmost part of the protocone.
The M2 is slightly wider than the M1, but similar in tooth structure. The metacone is slightly inclined posteriorly and shifted lingually.
The ectoflexus is deeper than on the M1; the parastyle projects
anteriolabially, and it is much stronger than on the M1. Transversely,
the protocone apex is situated at the same level as the paracone apex.
The paraconule is larger than the metaconule. The preparaconule crest
ends at the anterior base of the paracone. The postparaconule crest
reaches the lingual base of the paracone. The hypocone shelf is slightly
narrower than on the M1, but bears a small hypocone. The hypocone
shelf reaches the posterolingual base of the protocone.
The M3 is considerably more compressed anteroposteriorly than
the M1-2 and reduced. The large parastyle projects almost labially.
There is a weak, but distinct preparacrista that runs from the paracone
apex to the anterolabial base of this cusp. The metastylar lobe is absent. The labial cingulum is absent. The metacone is much reduced
(much smaller than the paracone), shifted lingually and posteriorly inclined. The precingulum and hypocone shelf are absent.

Coll. Number
PIN 3104-433

Tooth
M1
M2
M3

Length
2.2
2.0
1.45

Width
3.4
3.75
3.55

Measurements (in mm)The i3 (PIN 3104-427) height (including the root) 2.7; i3 crown height 0.7; i3 length 0.55; i3 width
0.5. The c height (including the root) 3.75; c crown height 1.2; c
length 0.8; c width 0.55. The M1-3 (PIN 3104-433) length - 5.45.
The other measurements are in Tables 1-3.
Praolestes maximus sp. nov.
Figs. 5-6
Etymologymaximus [Lat.] maximal, the species name is based
on the relatively large size, in contrast to the other known species nanus
[Lat.]tiny.
HolotypePIN 3104-363, right dentary with m1-m3 (Fig. 5a).
Type localityTsagan-Khushu, Zhigden Member of the NaranBulak Formation, latest Paleocene.
Referred specimensFrom the Zhigden Member of the TsaganKhushu locality: PIN 3104-364, fragment of right dentary with m3;
3104-365, isolated left m1 (Fig. 5b); 3104-366, isolated right m2 (Fig.
5c); 3104-367, isolated left m2 (Fig. 5d); 3104-368, left mandible fragment with alveoli for p4-m3; 3104-430, right dentary with m3; 3104429, right dentary fragment with the m1.
From the Zhigden Member of the Naran-Bulak locality: PIN 310275, right dentary with i3, p3-m3 and roots of i1-2, c and p1-2 (Fig. 6).
DiagnosisDiffers from the type species, P. nanus, in its larger
size (30 % larger) and in having a postcingulid on the m3.
DistributionUpper Paleocene of the Tsagan-Khushu and

189

FIGURE 5. Praolestes maximus sp. nov. a, holotype, PIN 3104-363, right dentary
with m1-m3: a, occlusal view; b, PIN 3104-365, left m1, occlusal view; c, PIN
3104-366, right m2, occlusal view; d, PIN 3104-367, left m2, occlusal view.

Naran-Bulak localities (Mongolia).


DescriptionThe mandible is robust, deep and slightly curved
ventrally (Fig. 6a, d). The symphysis is short and extends back to the
level of the p1. The posterior mental foramen is situated under the
posterior root of the p3. The ascending ramus is perpendicular to the
horizontal ramus. The height of the ascending ramus is one half of the
length of the horizontal ramus. The coronoid process is relatively tall,
vertical and narrow. Its apex is flattened and inclined posteriorly. The
masseteric ridge is rather strong, and the masseteric fossa is relatively
deep. The mandibular foramen is situated below the posterior slope of
the coronoid process on the level of the tooth row. It is round in shape
and faces posteriorly. The base of the articular process is situated relatively high and is twice as high as the height of the horizontal ramus.
The angular process is thin at the base. It is long and thin, as in other
nyctitheres (Sig, 1976, p. 81-82, fig. 95). There is a well-defined depression on the lingual side of the ascending ramus anterior and below
the mandibular foramen.
The dental formula is i3 c1 p4 m3. The lower incisors are inclined anteriorly. Judging from the preserved roots of the i1-2 and the
crown bases of i3, the i2 is the largest, and the i1 is the smallest, but
the difference in size is slight. The i3 is spatulate; its shearing surface is
not parallel to the long axis of the dentary. Judging from the base, the

FIGURE 6. Praolestes maximus sp. nov., PIN 3102-75, right dentary with i3, p3m3 and roots of i1-2, c and p1-2: a, labial view, b, occlusal view, c, roots of i1-2, i3,
roots of c-p1, d, lingual view.

canine was twice as large as the i3, whereas the p1 was almost equal in
size to the latter (Fig. 6c). P. maximus is similar to P. nanus in the
overall structure of the p4-m3. The p4 of P. maximus differs in having
a relatively small, anteriorly projecting paraconid and relatively large
metaconid, well separated from the protoconid. The talonid is relatively longer, with a well-expressed transverse cristid.
The lower molars lean slightly lingually in relation to the dentary.
The trigonids are twice as high as the talonids. As in P. nanus, the
lingual wall of the lower molars is almost straight, whereas the labial
is much curved due to the deep notch between the trigonid and talonid.
The m1-m2 are approximately the same length, or the m1 is slightly
shorter, and the m3 is much longer. The cuspids are rather massive.
The paraconid is relatively large and low, projects anteriorly and is not
fused with the metaconid. The metaconid is slightly higher than the
protoconid. The cristid obliqua begins at the posterior wall of the protoconid, and it is rather weak. The talonid is tricuspid, and the talonid
basin is open lingually. The hypoconid is the largest talonid cusp; it is
connected to the well defined hypoconulid by a strong cristid, which
extends far lingually and joins the entoconid. The entoconid is small;
the hypoconulid size is intermediate between the hypoconid and the
entoconid. The m3 talonid is narrower than on the m1-2, and the m3
hypoconulid is well-developed. There is a distinct postcingulid situated
labial to the hypoconulid.

190
Jarveia erronea sp. nov.
TABLE 4. Mandible depth of Praolestes maximus sp. nov. (in mm).
Coll. Number
PIN 3102-75

c
2.0

p1
2.2

p3
3.2

p4-m2
3.5

m3
3.9

TABLE 5. Measurements of the lower teeth of Praolestes maximus


sp. nov. (in mm).
Coll. Number

Tooth

PIN 3104-363

m1
m2
m3
m3
m1
m2
m2
m2
m3
i3
c
p4
m1
m2
m3

PIN 3104-364
PIN 3104-365
PIN 3104-366
PIN 3104-367
PIN 3104-429
PIN 3104-430
PIN 3102-75

Length Trigonid
width
2.4
1.75
2.4
1.95
2.95
1.8
2.85
1.8
2.25
1.5
2.1
1.65
2.1
1.56
2.3
1.6
2.7
1.8
0.75
0.7
1.1
1.2
2.05
1.4
2.5 (est.) 1.8
2.45
2.0
2.75
1.85

Talonid
width
1.55
1.65
1.45
1.5
1.5
1.56
1.44
1.45
1.5

1.75
1.35

Measurements (in mm)m1-m3 length (holotype) - 7.1; the


depth of the mandible under the m1 - 4.4, under the m3 - 4.1. The
height of the ascending ramus (PIN 3102-75) 12.0. The i1-m3 length
(PIN 3102-75) 17.4 (est.). The other measurements are in Tables 4-5.
Genus Jarveia Nessov, 1987
Jarveia Nessov, 1987, p. 209.
Jarveia: Averianov, 1995, p. 216.
Jarveia: McKenna and Bell, 1997, p. 285.
Type speciesJarveia minuscula Nessov, 1987.
Included speciesThe type species and J. erronea sp. nov.
Revised diagnosisSmall nyctitheriine, differs from
Bumbanius, Saturninia and North American genera in transversely wider
and anteroposteriorly shorter upper molars with vestigial hypocone.
The upper molars differ from those of Praolestes and Bayanulanius in
lacking the stylar shelf, and in having reduced conules and long lingual
slopes on the labial cusps. Jarveia also differs from Bayanulanius in
the wider M3 without a hypocone; in the lower molar structure in having a more labially placed cristid obliqua and rectangular rather than
triangular lower molar talonid basins. Differs from both Praolestes and
Bumbanius in having narrower, laterally compressed lower molars and
a reduced m3 talonid. Differs from Oedolius in having a twinned entoconid and hypoconulid on the lower molars and a wide labial notch
between the trigonid and the talonid.
DistributionLate Paleocene (Gashatan) of Mongolia and
Kazakstan.
CommentsAverianov (1995) redescribed Jarveia minuscula
Nessov, 1987, based on the holotype, represented by a dentary with
m1-3. He also described a maxillary fragment with M1-2 from the same
locality as Jarveia sp. He doubted reference of this specimen to J.
minuscula because of the slightly larger size. However, the same correlation of the size of the upper and lower molars is observed in Praolestes
and Bayanulanius. So, we refer the upper molars of Averianovs Jarveia
sp. to J. minuscula.

Figs. 7-9
Etymology erroneus [Lat.]wanderer, traveler. The species name
is due to the considerable distance between the geographic occurrence of
the new species and the type species.
HolotypePIN 3104-434, left maxillary fragment with P4-M3
(Figs. 7, 9).
Type localityTsagan-Khushu, Zhigden Member of the NaranBulak Formation, latest Paleocene.
Referred specimenFrom the Zhigden Member of the TsaganKhushu locality: PIN 3104-428, left dentary with incomplete m2 and
complete m3 (Fig. 8).
DiagnosisThe new species is slightly larger than the type species. It differs in the lower molar proportionsthe m3 is shorter or
equal in length to the m2whereas it is obviously longer in J.
minuscula. The entoconid of m3 is much weaker than in the type species. There is a distinct postcingulid on the m2. The upper molars differ in being less curved anteriorly with weaker metastyles and shorter
postmetacristae. The new species differs in having a distinct metaconule
on M2.
DistributionLate Paleocene (Gashatan) of the Tsagan-Khushu
locality.
DescriptionThe base of the vertical ramus is preserved in PIN
3104-428 (Fig. 8). The angle between the horizontal ramus and the
coronoid process is ~ 130o. The dentary is lower than in Praolestes and
slightly curved lingually. The masseteric fossa is rather deep, and the
anterior masseteric ridge is strong. There is a well-expressed tubercle
at the base of the anterior masseteric ridge at the midheight of the
dentary. The mandibular foramen is large, suboval in shape and faces
posteriorly. It is situated a little lower than the level of the tooth row,
under the middle of the coronoid process. There is a rounded, dorsallypointed tubercle directly posterior to the m3, at the base of the vertical
ramus. Such tubercles occur in other nyctitheriids, such as Saturninia
(Sig, 1976, p. 13, figs. 2 B-C).
The m2 is partly damaged. The trigonid is less compressed than
in Praolestes and Bumbanius. The posterior wall of the protoconid is
vertical. The paraconid is high and situated anterior to both the protoconid and metaconid and slightly shifted medially. The precingulid is
short but relatively strong. There is a weak postcingulid labial to the
hypoconulid that runs dorsolingually from the posterior base of the hypoconid to the posterolabial base of the hypoconulid. The cristid obliqua
begins at the base of the posterior wall of the trigonid. The talonid is
wide, with a large hypoconid. The hypoconulid is separated from the
hypoconid, it is relatively large and posteriorly projecting. The entoconid is broken, but judging from the base, it was situated close to the
hypoconulid.
Crowns of both the m2 and the m3 are curved labially, with a
deep notch between the trigonid and the talonid. The m3 trigonid is
wider than the talonid. The trigonid is slightly compressed. The
paraconid is not reduced and not shifted medially. It is situated directly
anterior to the metaconid. The posterior wall of the protoconid is slightly
concave. The cristid obliqua begins from the posterior base of the trigonid, closer to the protoconid, so the talonid basin is almost rectangular
in shape. The hypoconid is large and protrudes labially. The hypoconulid
is the same size as the hypoconid and projects posteriorly. The entoconid is close to the hypoconulid.
Part of the zygomatic process of the maxilla is preserved in the
holotype (Figs. 7, 9). The process begins at the middle of the M2. It is
closely appressed to the maxillary bone, and the angle between them is
about 150. There is a distinct lateral depression on the anterior part of
the zygomatic process.
The labial part of the P4-M1 and complete M2-3 are preserved in
the holotype. The P4 is shifted lingually in relation to the molars. The
paracone is large and situated in the center of the labial part of the crown.

191

FIGURE 7. Jarveia erronea sp, nov., holotype, PIN 3104-434, left maxillary
fragment with P4-M3, occlusal view.

There is no metacone. The stylar shelf is absent, and the parastyle is


well-developed, projecting anteriorly. The metastyle is relatively large
and connected to the paracone by a crest. The protocone is broken, but
judging from its lingual base, this cusp was rather large. There are no
conules.
The paracone is larger than the metacone on the M1. The
postmetacrista is short. The stylar shelf is absent, and the parastyle and
metastyle are the same shape as on the P4. The anterior wall of the
crown is almost straight. There is a narrow crest running along the
posterior wall of the metacone to the metastyle base.
The M2 is transversely elongate and slightly curved anteriorly.
The paracone and metacone are relatively smaller than on the M1 and
widely spaced. The lingual slopes of these cusps are long, extending
far into the trigon basin. The postmetacrista is very short. The parastyle
projects anterolabially, and the metastyle projects posterolbially, so the
labial wall of the crown is slightly curved lingually. The protocone is
massive, transversely, its apex is situated on the same level as the paracone apex. The preprotocrista reaches the anterior base of the paracone. There is no paraconule. The postprotocrista runs to the weak
metaconule. The metaconule is low without any wings. The thin crest
connects the metaconule with the metastyle. The precingulum is short
and weak. The hypocone shelf is inflated with a distinct low crest and
a small lingual cuspule.
The M3 is smaller than the M2. The metacone is relatively reduced and slightly shifted lingually. The parastylar lobe projects
anterolabially. The preparacrista is strong, but does not reach the
parastyle. The metastylar lobe is absent. The protocone is pointed. The
preprotocrista reaches the anterior base of the paracone. The
postprotocrista runs from the protocone to the small metaconule. There
is a short and weak crest between the metaconule and the posterior
base of the metacone, but no precingulum or hypocone shelf.
Measurements (in mm)Depth of the dentary under the m2
1.7; depth of the dentary under the m3 1.6; m2 length 1.45 (est.);
m3 length 1.35; m3 trigonid width 0.95 (est.); m3 talonid width
0.7 (est.); P4 length 1.4; M1 length 1.6; M2 length 1.5; M2 width
2.5; M3 length 1.1; M3 width 2.3.
DISCUSSION
The systematic position of Praolestes was long a subject of discussion. The holotype of Praolestes is represented by three cheek teeth,
initially identified as p3, p4 and m1, and two alveoli anterior to them
(Matthew et al., 1929). Szalay and McKenna (1971) listed the teeth
also as p3, p4, m1 and two alveoli for a double-rooted p2. Similarity in
dental morphology with Late Cretaceous Zalambdalestidae led them to
assign Praolestes to this family.
Saban (1958) referred Praolestes to the Leptictidae. He suggested
Praolestes was very similar to Diacodon and Prodiacodon in the p4
morphology (p5 in recent leptictid dental terminology). He wrote (p.

FIGURE 8. Jarveia erronea sp, nov., PIN 3104-428, left dentary with partial m2
and complete m3: a, labial view, b, occlusal view, c, lingual view.

875) that this tooth [p4] is more similar to the p3 [of Leptictidae]
because of the weak, one-cusped talonid and two small cusps: anterior
[paraconid] and posterior [metaconid] to the main cusp. The next tooth,
m1 is very similar to the p4 [actually p5] of Leptictidae, as mentioned
in the description. It possesses more developed trigonid than the previous tooth with better expressed paraconid. Its talonid is not concave
with a single cusp and a crest on the place of the second. So, Saban
suggested that the p4 of the holotype is actually the p3, and the m1 is
the p4. Probably based on this assumption, Praolestes was referred to
Leptictida (McKenna and Bell, 1997).
But, the specimens at hand allow us to conclude the following.
In the mandibles with m1-3 and m1-2 (PIN 3104-361, 360 and 427),
the first molar, m1, is identical to the posterior tooth in the holotype, so
it is a true m1 and not p4 or p5 as suggested by Saban. This is confirmed by the almost complete dentary, PIN 3104-427, with p1-m2, in
which the p3-4 and m1 are identical to the p3-4 and m1 of the holotype.
This dentary has two mental foramina, as in the holotype. In this dentary,
the alveolus situated anterior to the p1 is large and rounded and should
be for the canine, so there is no doubt as to the premolar count in
Praolestes. In the dentary, PIN 3104-360, all the alveoli anterior to the
m1 are preserved. Together with the upper teeth, this allows us to reconstruct the dental formula of Praolestes as I?/3 C?/1 P4/4 M3/3, which
does not correspond to Leptictida. So, the dental formula and the tooth
structure, such as a non-molariform posterior premolar, suggest that
Praolestes doesnt belong to Leptictida.
Szalay and McKenna (1971) assigned Praolestes to
Zalambdalestidae, but Kielan-Jaworowska (1984) argued that the posterior premolar of Praolestes is small and the m1 trigonid is not compressed, as in the other Zalambdalestidae. Wang (1995) also didnt in-

192

FIGURE 9. Jarveia erronea sp, nov., holotype, PIN 3104-434, left maxillary
fragment with P4-M3: a, labial view, b, occlusal view.

clude Praolestes in Zalambdalestidae, although he mentioned that the


scarcity of material did not allow for a thorough comparison. Praolestes
does differ significantly from Zalambdalestidae, mostly in the incisor
structure (the first incisor is not procumbent), larger canine, which is
weak in Zalambdalestidae, the closely spaced teeth with no diastemata
present, the less compressed molar trigonids with better developed
paraconids and the non-reduced m3 with posteriorly projecting
hypoconulid. But, Praolestes nanus and P. maximus sp. nov. possess
features that are diagnostic of the lipotyphlan family Nyctitheriidae: (1)
the lower molar trigonids are columnar, the constriction between the
trigonid and the talonid is well expressed; (2) the hypoconid is relatively
high; (3) the sagittal plane of the lower tooth row is somewhat inclined
lingually from the lower jaw sagittal plane; (4) the trigonid exceeds the
talonid in height no more than twice; (5) the hypoconulid is small, but
present; (6) the dentary of Praolestes maximus possesses a structure of
the vertical ramus identical to that of North American and European
nyctitheres (McKenna, 1968; Sig, 1976); and (7) the dentary of Praolestes
nanus, PIN 3104-360, possesses a large mental foramen under the p3,
which is typical of nyctitheriids (McKenna, 1968; Sig, 1976; Russell
and Dashzeveg, 1986; Nessov, 1987; Averianov, 1995). The upper molars referred to Praolestes in the present paper well support its assignment to Nyctitheriidae.
Jarveia was described as a Proterutheria by Nessov (1987).
Averianov (1995) redescribed it as a nyctitheriid insectivore. McKenna

and Bell (1997) referred this genus to Micropternodontidae, with


Hyracolestes and Sarcodon. Jarveia does not share any synapomorphies
with other micropternodontids and does not have any morphological
features typical of that family, such as extremely reduced or absent m3
and very tall lower molar trigonids. On the other hand, all the characters of both the upper and the lower teeth are typical of nyctitheriids
(see above). So, we follow Averianov (1995) and refer Jarveia to
Nyctitheriidae.
We identified three nyctitheriid species of different size in the
Zhigden Member of the Tsagan-Khushu locality: large P. maximus,
medium-sized P. nanus and small J. erronea. Available upper teeth
that can be referred to Nyctitheriidae belong to two size groups. The
smallest obviously correspond to Jarveia in molar structure (Averianov,
1995, p. 217-218, figs. 2-3). Based on size and occlusion pattern, we
referred the other to P. nanus.
The upper molars of both Praolestes and Jarveia are rather primitive, closer to Leptacodon than to Saturninia (Krishtalka, 1976; Sig,
1976). The structure of the labial part of the P4 is unknown in Praolestes,
but the P4 of Jarveia erronea differs from that of other nyctitheriids in
lacking the distinct metacone, which was either absent or vestigial and
disappeared with wear. The contemporaneous Asian nyctitheriid
Bayanulanius closely resembles Praolestes and Jarveia in upper molar
shape. The peculiar shape of the upper molars of the Paleocene Asian
nyctitheres, Praolestes, Bayanulanius and Jarveia, which are relatively
short and wide, with a weak hypocone shelf, differs significantly from
the later Asian, North American and European, and contemporaneous
North American Nyctitheriidae.
Most of the characters of the upper cheek teeth of the Paleocene
nyctitheres are plesiomorphic. So, similarity in upper molar structure
in Praolestes, Jarveia and Bayanulanius reflects the primitive stage of
the evolution of Asian Nyctitheriidae, and is not due to close phylogenetic relationship. Moreover, judging from both upper and lower tooth
morphology we can identify three phylogenetic lineages among Asian
nyctitheres: Praolestes-Bumbanius, Bayanulanius-Oedolius and Jarveia.
Praolestes nanus is an index fossil of the Gashatan land-mammal age. Considering the close relationship between Jarveia
minuscula and J. erronea, we suggest a Gashatan age for the Zhylga
fauna from southern Kazakstan (Nessov, 1987; Averianov, 1995). The
Zhylga mammalian fauna, as with the Gashatan fauna of Mongolia,
also includes arctostylopids and mixodonts.
Detailed comparison of the morphology of the Gashatan and
Bumbanian Nyctitheriidae reveals that the former represent a distinct
stage of faunal evolution of Asian mammals. That corresponds well
with the data on other mammalian groups (Ting, 1998), confirming the
validity of the Gashatan land-mammal age.
ACKNOWLEDGMENTS
R. Tedford made it possible to study specimens in the AMNH.
Funding from the NMMNH Foundation made this study possible. The
study was supported by the Russian Foundation for Basic Research,
projects 02-04-48458 and 04-05-64805, and the Grant Council of the
President of Russian Federation, projects MK-726.2004.4 and Nsh1840.2003.4, and the grant of the program Support of young scientists of the Presidium of the Russian Academy of Sciences.

REFERENCES
Averianov A., 1995, Nyctitheriid insectivores from the upper Paleocene of
southern Kazakhstan (Mammalia, Lipotyphla): Senckenbergiana
Lethaea, v. 75, p. 215-219.
Badamgarav D. and Reshetov, V. Yu., 1985, Paleontologiya i stratigrafiya paleogena
Zaaltayskoy Gobi. [Paleontology and stratigraphy of the Paleogene of Zaaltayski
Gobi]: Trudy Sovmestnoy Sovetsko-Mongolskoy Paleontologicheskii Expeditsii
[Proceedings of the Joint Soviet-Mongolian Paleontological Expedition], v. 25, 104 p.

Carroll, R.L., 1988, Vertebrate Paleontology and Evoluation. New York:


W.H. Freeman and Co., 698 p.
Fox, R.C., 1970, Eutherian mammal from the early Campanian (Late Cretaceous)
of Alberta, Canada: Nature, v. 227, p. 630-631.
Kielan-Jaworowska, Z., 1984, Evolution of the therian mammals in the
Late Cretaceous of Asia. Part V. Skull structure in Zalambdalestidae:
Palaeontologia Polonica, v. 46, p. 106-117.
Krishtalka, L., 1976, North American Nyctitheriidae (Mammalia,

193
Insectivora): Annals of the Carnegie Museum, v. 46, p. 7-28.
Matthew, W.D., Granger, W. and Simpson, G.G., 1929, Additions to the fauna of
the Gashato Formation of Mongolia: American Museum Novitates, no. 376, 12
p.
McKenna, M.C., 1968, Leptacodon, an American Paleocene nyctithere (Mammalia, Insectivora): American Museum Novitates, no. 2314, 12 p.
McKenna, M.C. and Bell, S.K., 1997, Classification of mammals above the species
level. New York, Columbia University Press, 631 p.
Meng, J., Zhai, R. and Wyss, A.R., 1998, The late Paleocene Bayan Ulan fauna of
Inner Mongolia, China: Bulletin of the Carnegie Museum Natural History, v.
34, p. 148-185.
Nessov L.A., 1987, Results of searches and investigations in the mammal-bearing
Cretaceous and early Paleogene on the territory of the USSR: Annual of AllUnion Paleontological Society, v. 30, p. 199-218.
Russell, D.E. and Dashzeveg D., 1986, Early Eocene insectivores (Mammalia) from the Peoples Republic of Mongolia: Palaeontology, v. 29, p.
269-291.

Saban R., 1958, Insectivora; in Piveteau J., ed. Traite de Paleontologie, v.


6(2), p. 822-909.
Sig, B., 1976, Insectivores primitifs de lEocene superieur et Oligocene inferieur
dEurope occidentale. Nyctitheriides: Memoires Museum National dHistoire
Naturelle, v. C34, p. 1-140.
Simpson, G.G., 1928, A new mammalian fauna from the Fort Union of southern
Montana: American Museum Novitates, no. 297, 15 p.
Szalay, F.S. and McKenna, M.C., 1971, Beginning of the age of mammals in Asia:
The late Paleocene Gashato fauna, Mongolia: Bulletin of the American Museum of Natural History, v. 144, p. 269-318.
Ting, S., 1998, Paleocene and early Eocene land mammal ages of Asia: Bulletin of
the Carnegie Museum Natural History, v. 34, p. 124-147.
Van Valen, L., 1967, New Paleocene insectivores and insectivore classification:
Bulletin of the American Museum of Natural History, v. 135, p. 217-284.
Wang, Y., 1995, A new zhelestid (Mixotheridia, Mammalia) from the Paleocene of
Quinshan, Anhui: Vertebrata Palasiatica, v. 33, p. 114-137.

194

Periptychus coarctatus, skull and jaw (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

195

PALAEOSTYLOPS ITURUS FROM THE UPPER PALEOCENE OF MONGOLIA AND THE


STATUS OF ARCTOSTYLOPIDA (MAMMALIA, EUTHERIA)
PETER E. KONDRASHOV1 AND SPENCER G. LUCAS2
1

Department of Biology, Northwest Missouri State University, 800 University Drive, Maryville, MO 64468
and Paleontological Institute of the RAS, Moscow, RUSSIA;
2
New Mexico Museum of Natural History, 1801 Mountain Road, Albuquerque, NM 87104

AbstractNew specimens, including complete lower and upper dentitions, of the rare Paleocene mammal
Palaeostylops iturus are described. The material comes from the upper Paleocene (Gashatan) deposits of the
Zhigden Member of the Naran-Bulak Formation at Tsagan-Khushu and the Gashato Formation at Shabarakh
Usu, both localities in Mongolia. Based on comparison of Palaeostylops iturus with Gashatostylops macrodon,
we conclude that Gashatostylops is a junior subjective synonym of Palaeostylops. The dental morphology of
Palaeostylops and arctostylopids is remarkably similar to that of true notoungulates, which is unlikely to be due
to evolutionary convergence. Indeed, arguments previously presented to distinguish an order Arctostylopida
separate from Notoungulata are largely based on unsupportable assumptions and an incorrect interpretation of
dental homologies, and thus are insufficient to demonstrate separate ordinal rank for Arctostylopida. So, we
abandon the term Arctostylopida and treat Arctostylopidae as a family within Notoungulata.
Key words: Palaeostylops, Paleocene, Mongolia, Gashatan, Arctostylopida, Notoungulata

INTRODUCTION

SYSTEMATIC PALEONTOLOGY

The first arctostylopid discovered, Arctostylops steini, was described by Matthew (1915) from the upper Paleocene (Clarkforkian) of
Wyoming, who referred it without question to the Notoungulata. The
new genus Palaeostylops and two new species of Arctostylopidae were
subsequently described from the late Paleocene Gashato fauna of
Mongolia (Matthew and Granger, 1925; Matthew et al., 1929). Later,
one of the species was assigned to a new genus, Gashatostylops (Cifelli
et al., 1989). Several new genera and species of arctostylopids were
subsequently discovered in Asia (Tang and Yan, 1976; Zhai, 1978;
Zheng, 1979; Zheng and Huang, 1986; Nessov, 1987; Huang et al.,
2001; Huang and Zheng, 2003). Arctostylops steini long remained a
poorly known species until new material was described from the upper
Paleocene of North America (Gingerich and Rose, 1977; Rose, 1981;
Cifelli et al., 1989; Bloch, 1999).
Traditionally, arctostylopids were assigned to Notoungulata based
on a set of characters held in common with South American
notoungulates. Matthew (1915) placed them in ?Isotemnidae, which
was later followed by Patterson (1934). Simpson (1934, 1945, 1948)
erected the notoungulate suborder Notioprogonia, in which he included
Arctostylopidae, Notostylopidae and Henricosborniidae, uniting them
mostly on the basis of shared primitive similarities. Cifelli et al. (1989)
rejected the idea of a close relationship between the true notoungulates
and arctostylopids and erected a new order Arctostylopida to include a
single family, Arctostylopidae.
Here, we describe newly collected, extensive material of
Palaeostylops iturus from the upper Paleocene (Gashatan) of Mongolia
(Kondrashov and Lucas, 2002). On the basis of comparison of
Palaeostylops iturus from Tsagan-Khushu and Gashatostylops
macrodon from Naran-Bulak, we consider Gashatostylops to be a junior subjective synonym of Palaeostylops. We also critically review the
evidence marshalled by Cifelli et al. (1989) to distinguish arctostylopids
from true notoungulates and conclude that the order Arctostylopida
should be abandoned and Arctostylopidae should remain a family within
the order Notoungulata.
Institutional Abbreviations: AMNH = American Museum of
Natural History, New York; IVPP = institute of Vertebrate Paleontology and Paleoanthropology, Beijing; PIN = Paleontological Institute,
Moscow.

Order Notoungulata Roth, 1903


Family Arctostylopidae Schlosser, 1923
Arctostylopidae Schlosser, 1923, p. 614
Arctostylopinae: Zheng, 1979, p. 391.
Arctostylopida Cifelli, Schaff and McKenna, 1989, p. 5.
Arctostylopidae: Cifelli, Schaff and McKenna, 1989, p. 5.
Arctostylopiformes Kinman, 1994, p. 38
Arctostylopida: McKenna and Bell, 1997, p. 365.
Arctostylopidae: McKenna and Bell, 1997, p. 365.
Arctostylopida: Meng et al., 1998, p. 169.
Arctostylopidae: Meng et al., 1998, p. 169.
Type genusArctostylops Matthew, 1915.
Included generaThe type genus; Anatolostylops Zhai, 1978;
Asiostylops Zheng, 1979; Allostylops Zheng, 1979; Palaeostylops
Matthew and Granger, 1925; Sinostylops Tang and Yan, 1976;
Bothriostylops Zheng and Huang, 1986; Kazachostylops Nessov, 1987;
Stenostylops Huang, Zheng and Ding, 2001.
Revised diagnosisDiffers from other notoungulates in having
more simple molar crown morphology and in having triserial, trenchant lower premolars.
DistributionPaleocene and Eocene of Asia, late Paleocene of
North America.
Genus Palaeostylops Matthew and Granger, 1925
Palaeostylops Matthew and Granger, 1925, p. 2.
Gashatostylops Cifelli, Schaff and McKenna, 1989, p. 15.
Palaeostylops: McKenna and Bell, 1997, p. 365.
Gashatostylops: McKenna and Bell, 1997, p. 365.
Type speciesPalaeostylops iturus Matthew and Granger, 1925.
Included speciesType species and Palaeostylops macrodon
Matthew, Granger and Simpson, 1925.
DistributionUpper Paleocene (Gashatan) of Asia.
Revised diagnosisMedium-sized arctostylopids that differ
from the most similar genus Arctostylops in having a simple, tricuspid
p4, without a well-defined trigonid and talonid. Differs from Sinostylops
in lacking the paracristid on the lower molars. Differs from

196

FIGURE 1. Original illustrations of Palaeostylops (from Matthew and Granger,


1925 and Matthew et al., 1929). A-B, P. iturus, AMNH 20415, left maxillary
fragment with P1-M3 in labial (A) and occlusal (B) views. C, P. macrodon, AMNH
22144, left maxillary fragment with P3-M2 (paratype), occlusal view. D, P. sp.,
AMNH 20426, p1? E-F, P. iturus, AMNH 20428, left p1-3, labial (E) and lingual
(F) views. G, J, P. macrodon, AMNH 21725, left dentary fragment with p3-m2
(holotype), occlusal (G) and lingual (J) views. H-I, K, P. iturus, AMNH 20414,
right dentary fragment with i3-m3 (holotype), occlusal (H), labial (I) and lingual
(K) views.

Anatolostylops and Stenostylops in having lower-crowned teeth. Differs


from Asiostylops in having a well-developed hypocone. Differs from
Kazachostylops and Bothiostylops in having a more reduced m3.
Palaeostylops iturus Matthew and Granger, 1925
Figs. 1A-B, E-F, H-I, K, 2-3, 6A, C, Tables 1-2
Palaeostylops iturus Matthew and Granger, 1925, p. 2, figs. 1-4.
Palaeostylops iturus: Chow and Qi, 1978, p. 83, pl. 2, figs. 5-7.
Arctostylops iturus: Dashzeveg and Russell, 1988, p. 131.
Palaeostylops iturus: Cifelli, Schaff and McKenna, 1989, p. 15, figs. 8B,
9B.
Palaeostylops iturus: Meng et al., 1998, p. 169.
HolotypeAMNH 20414, right dentary with i3, c, p1-4, m1-3
(Figs. 1H-I, K; Matthew and Granger, 1925, fig. 1).
Referred specimensFrom the Zhigden Member (upper Paleocene) of the Naran Bulak Formation, Tsagan Khushu locality, lower red
horizon: PIN 3104-316, right P2-M2; 3104-317, left M1-3 (Fig. 2B);
3104-318, right P3-M3; 3104-319, right P4-M3 (Fig. 2H); 3104-320,
left P4-M3; 3104-321, left p3-m3 (Fig. 3A); 3104-322, right p1-m3;
3104-323, left p2-m3; 3104-324, left p3-m3 (Fig. 3C); 3104-325, right
P4-M2; 3104-326, left P1-4; 3104-327, right p1-3; 3104-328, right m23; 3104-329, right m1-2; 3104-330, left M3; 3104-331, right m3; 3104332, left P2-4; 3104-333, right p4-m3; 3104-334, right m2-3; 3104-335,
right p4-m3; 3104-336, right M1-2 (Fig. 2E); 3104-337, left p2-4; 3104338, right P4-M1; 3104-339, left P1-M3; 3104-340, left M1-3; 3104341, right M1-2; 3104-342, right p4-m1; 3104-343, right M1-2; 3104344, left P3; 3104-345, left P4-M1 (Fig. 2C); 3104-346, right m2-3;
3104-347, left M2; 3104-348, left c1-p2; 3104-349, right P2-M2; 3104350, left c1-p4, right i2-p3; 3104-351, right P2-3; 3104-352, right P2-4;
3104-353, left P3-M3; 3104-354, right p4; 3104-355, left p2-3, m1;
3104-356, left P2-M2; 3104-357, left M3; 3104-358, left m1; 3104-374,
left P4-M3; 3104-375, right P2-M3; 3104-376, left P2-4; 3104-377, left
P2-M2; 3104-378, left P3-M1; 3104-379, left M1-2; 3104-380, right cm3; 3104-381, left m1-3; 3104-382, right P4-M3; 3104-383, right m1-3;
3104-384, right p1-m3; 3104-385, right i3-m2; 3104-386, right p1-m3;

FIGURE 2. Occlusal views of selected upper cheek teeth of Palaeostylops iturus.


A, PIN 3104-421, left P4-M2. B, PIN 3104-317, left M1-3. C, PIN 3104-345, left
P4-M1. D, PIN 3104-459, right M2-3. E, PIN 3104-336, right M1-2. F, AMNH
20415, left P1-M3 (paratype of P. iturus). G, PIN 3104-443, right P3-M1. H, 3104319, right P4-M3.

3104-387, right m1-2; 3104-388, left m2-3; 3104-389, right m1-3; 3104390, left p2-4; 3104-391, left m1-3; 3104-392, right p2-m2; 3104-393,
right m2-3; 3104-394, left p2-m2; 3104-395, left p3-4, m2-3; 3104-396,
left p1-2, p4-m1; 3104-397, left m2-3; 3104-398, left p2-m3; 3104-399,
left p2-m1; 3104-400, left p4-m3; 3104-401, right p4-m1; 3104-402,
right m2-3; 3104-403, left p1-3; 3104-404, left p4-m1; 3104-405, left
m2-3; 3104-406, left p3-m3; 3104-407, right m1-3; 3104-408, left m1-3;
3104-409, left p3-m2; 3104-410, left i1-p4; 3104-411, left m2-3; 3104412, right p3-m1; 3104-413, right p3-m3; 3104-414, right m2-3; 3104415, left p2-m3; 3104-416, right p2-m3 (Fig. 3B); 3104-417, right p4m3; 3104-418, right p1-m3; 3104-419, left p4-m1; 3104-420, left p1m3; 3104-421, left P4-M2 (Fig. 2A); 3104-422, right M1-2; 3104-411,
right p3-m2; 3104-412, right p3-m1; 3104-413, right p3-m3; 3104-414,
right m2-3; 3104-415, left p2-m3; 3104-416, right p2-m3; 3104-417,
right p4-m3; 3104-418, right p1-m3; 3104-419, left p4-m1; 3104-420,
left ?-m3, p1-m3; 3104-421, left p4-m2; 3104-422, right m1-2; 3104423, right p3-m2; 3104-424, right p4-m3; 3104-425, right m1-3; 3104426, left p2-m3; 3104-442, right p3-m2; 3104-443, right P3-M1 (Fig.
2G); 3104-444, left p2-m3; 3104-445, right i3-m3; 3104-446, right m13; 3104-447, right m1-2; 3104-448, left m3; 3104-449, left m1; 3104450, right M1-3; 3104-451, left M2-3; 3104-452, right m1; 3104-453,
right M1; 3104-454, right c-p2; 3104-455, left M2-3; 3104-456, left
M2; 3104-457, left p2-m2; 3104-458, right m2-3; 3104-459, right M2-3
(Fig. 2D); 3104-460, right p2-3; 3104-461, right p3-m1; 3104-462, right
p4-m2; 3104-463, right M3; 3104-464, left m2-3; 3104-465, right p4m2; 3104-466, left p4-m2; 3104-467, right P3-M3; 3104-468, left M1;
3104-469, left M1-2; 3104-470, left P4; 3104-471, left M3; 3104-472,
right p2-3; 3104-473, right p4; 3104-474, right p4-m1; 3104-475, left
m2; 3104-476, left m1-2; 3104-477, right M2-3; 3104-478, right P2-4.
From the Gashato Formation, Shabarakh Usu, Mongolia: AMNH
20415, left maxillary fragment with P1-M3 (Figs. 1A-B, 2F; Matthew
and Granger, 1925, fig. 2); 20417, left and right M3; 20419, left dentary
with p3-4, m1-3; 20426, anterior tooth (Fig. 1D; Matthew and Granger,
1925, fig. 4); 20428, left dentary fragment with p1-3 (Fig. 1E-F; Matthew and Granger, 1925, fig. 3); 21740, left dentary with p3-4, left P3M3; 21742, right dentary with i3-m2; 21743, left dentary with p3-4;

197
TABLE 1. Measurements (in mm) of upper cheek teeth of Palaeostylops iturus.
Coll. Number
P1 L
PIN 3104-316
PIN 3104-317
PIN 3104-318
PIN 3104-319
PIN 3104-320
PIN 3104-325
PIN 3104-326
1.35
PIN 3104-330
PIN 3104-332
PIN 3104-336
PIN 3104-338
PIN 3104-339
1.65
PIN 3104-340
PIN 3104-341
PIN 3104-343
PIN 3104-344
PIN 3104-345
PIN 3104-347
PIN 3104-349
PIN 3104-351
PIN 3104-352
PIN 3104-353
PIN 3104-356
PIN 3104-357
AMNH 20417
AMNH 20417
AMNH 21726-8
AMNH 21726-16 1.63
AMNH 21726-17
AMNH 21726-18
AMNH 21740
AMNH 101965
AMNH 101975
IVPP w/n
1.8
Mean
1.61
Number of spec. 4
Stand. Deviat.
0.19

P1 W

0.87

0.93

P2 L
1.7

P2 W
1.55

P3 L
1.85

P3 W
1.79

P4 L
1.7

P4 W
2.2

2.13

1.8

2.2
2.25
2.6
2.3
2

1.77

2.1

1.55

2.1
2.15
2.11
2
2.2

1.6

1.98

1.8

2.11

1.51

1.95
2.05

2.2
2.25

2.1

2.1

1.5

2.01

1.04

1.41
1.063
4
0.242

1.78
1.75

1
1.3

2.07

1.35

1.99
2.02
1.64

2.18
1.91
11
0.18

1.41
1.54
1.11

1
1.305
11
0.239

2.15
2.12
2.03
2.3
2.21

M1 L
2.85
2.62
2.84
3.12
2.9
2.75

M1 W
3.1
2.97
2.88
3.3
3.18
2.8

2.9
2.46
2.75
3
2.2
3

3.2
2.9
2.95
3.1
2.9
2.9

3.07

2.37

1.6

2.3

2.45

2.1
2.25
2.2

2.18
2.6
2.21

2.8
2.95

M2 L
3.95
3.8
3.82
4.05
4.1
3.45

M2 W
3.8
3.72
3.5
3.92
4.1
3.55

3.6

3.7

3.38
3.9
3.5
3.65

3.43
3.65
3.4
3.55

3.7
3.6

3.6
3.5

3.15
3.12

3.75
3.72

3.61
3.54

M3 L
1.87
2.5
2.16
2

M3 W
3.2
3.38
2.95
3.2
3.25

2.25

3.5

2.12
2.1

2.71
3

2.3

3.1

2.15
2.09
2.06

2.87
2.62
2.51

1.8
1.43
1.75
2.1
1.9

2.14
2.16
1.72

1.99
1.81
1.54

2.19

2.05

1.93
2.14
2.07
16
0.15

1.62
2.03
1.748
16
0.277

2.49

1.98

2.24
1.91
2.13
1.87
2.07
19
0.2

2.61
1.83
1.96
2.48
2.193
19
0.278

22413, left maxillary fragment with I1-M3 (Matthew et al., 1929, fig.
13); 101965, left maxillary fragment with P4-M2; 101975, skull fragment with right P3-M2; 101983, left dentary with p3-m3; 101985, right
dentary with p2-3, m1-2; 21726-1, right dentary with p1-4; 21726-2,
right m1; 21726-3, left m3; 21726-4, left dentary with p3-4; 21726-5,
left m1; 21726-6, left p1; 21726-7, right dentary with p1-2; 21726-8,
right P2-3; 21726-9, right m2; 21726-10, right m2-3; 21726-11, right
m1-2; 21726-12, right dentary with p4-m1; 21726-13, left dentary with
m2-3; 21726-14, right dentary with p2-4; 21726-15, right dentary with
p4-m1; 21726-16, right maxillary fragment with P1-M3; 21726-17, right
P2-4; 21726-18, left M1-3; 21726-19, right m1.
From the Nomogen Formation at Bayan Ula, Nei Monggol, China:
IVPP 5489-5530, jaw fragments (Chow and Qi, 1978, pl. 2, figs. 5-7).
DistributionUpper Paleocene (Gashatan) of Mongolia.
Revised diagnosisA species of Palaeostylops that differs from
P. macrodon by its smaller size and relatively smaller M2/m2.
DescriptionIn the description below we use the cusp nomenclature for arctostylopid molars proposed by Cifelli et al. (1989, fig. 1),
except we term the anterolabial cristid on the lower molar trigonid a
parastylid. We do so because the paraconid should lie lingual to the
protoconid, and this cristid extends anteroexternally, so it is an elongated
parastylid. Thus, the cristid, which begins at the protoconid and extends

2.81

2.95

3.46

3.65

1.91

2.3

2.68
3.3
2.33
2.52
2.78
2.78
22
0.27

2.66
3.61
2.46
2.34
3.11
2.952
22
0.299

3.55
5.26
3.49
3.05
3.87
3.75
21
0.42

3.24
4.24
2.88

2.46

2.86
3.67

3.62
3.61
20
0.288

2.4
2.17
14
0.19

3.51
3.039
16
0.385

anterolingually, should be called a paracristid. The cristid that connects


the protoconid and the metaconid should not be called the protocristid,
but the metacristid. The external cristid that embraces the lower molars
from the labial side is not actually the cristid obliqua, but the united
cristid obliqua and the hypolophid, so it is best called an ectolophid.
In the one skull fragment, the infraorbital foramen is small and
situated above the anterior root of the P4. The infraorbital canal is
rather long. The upper canine and the P1 are simple, single-rooted and
laterally compressed, with a well developed lingual cingulum. The P1
ectoloph does not bear any cusps.
The P2-4 have three roots. The P2 is the same length as the P3,
but a little narrower. The protocone on this tooth is shifted posteriorly
and is smaller in size than on the P3. The P3 is slightly longer than the
P4, but a little narrower, so it is triangular in shape. The protocone is
well developed, but smaller than on the P4. The parastyle is well developed. There is a small anterior cingulum. The P4 is more compressed
anteroposteriorly, the protocone is much larger than on P3; the anterior
and posterior cingula are well developed, and the ectoloph is strong.
There is a continuous lingual cingulum that begins at the parastyle and
reaches the metastyle without interruption on the lingual side of the
protocone.
The M1 is trapezoidal in shape. The ectoloph is strong, the fosette

198
TABLE 2. Measurements (in mm) of lower cheek teeth of Palaeostylops iturus.
Coll. Number

i3 L

i3 W

cL

cW

p1 L

p1 W

p2 L

p2 W

PIN 3104-321
PIN 3104-322

1.45

0.75

PIN 3104-323

1.22

0.75

p3 W

p4 L

p4 W

m1 L

m1 W

m1 TW m2 L

m2 W

m2 TW m3 L

m3 W m3 TW

0.93

2.35

1.12

2.53

1.35

1.23

3.22

2.05

1.95

2.6

1.45

1.2

1.87

0.8

2.12

0.92

2.35

1.18

2.73

1.4

1.52

3.42

1.75

1.75

2.7

1.5

1.2

1.92

0.8

2.19

0.95

2.25

2.48

1.45

1.38

3.25

1.6

1.45

2.45

1.3

1.1

2.3

1.1

2.25

1.2

2.9

1.45

1.35

3.32

1.95

1.85

2.3

1.4

1.25

1.75

0.9

2.27

1.05
3.65

1.9

1.7

3.31

1.4

1.2

2.64

1.2

1.4

3.3

1.7

1.65

1.7

1.7

PIN 3104-324
PIN 3104-327

p3 L
1.9

PIN 3104-328
PIN 3104-329
PIN 3104-331
PIN 3104-333

2.85

1.4

1.2

2.85

1.3

1.05

2.2

1.05

2.75

1.4

1.2

3.15
3.3

1.6

1.5

2.5

1.5

1.1

2.5

1.1

2.9

1.4

1.55

3.5

1.5

1.6

2.5

1.2

1.05

2.4

1.1

PIN 3104-342

1.7

1.4

2.95

1.3

1.1

PIN 3104-346

3.6

1.8

1.5

2.9

1.25

1.1

PIN 3104-334
PIN 3104-335
PIN 3104-337

1.4

PIN 3104-348

1.5

0.85

0.87

0.9

2.2

1.03

1.73

PIN 3104-350
PIN 3104-350
PIN 3104-354
PIN 3104-355

1.92

1.48

2.4

1.2

2.63

2.75

PIN 3104-358

2.35

1.2

1.25

2.81

1.3

1.2

PIN 3104-372
AMNH 20414

1.31

0.95

1.48

1.09

1.57

1.11

1.82

1.12

2.05

1.28

2.22

1.28

AMNH 20419

3.06

AMNH 20419
AMNH 21726-1

1.2

0.95

1.9

2.1

0.98

2.12

2.07

1.07

2.48

1.27

AMNH 21726-2

2.88

1.46

3.5

3.7

1.5

3.38

1.64

1.44

2.82

1.48

1.23

3.49

1.72

1.52

3.22

1.42

1.22

2.47

1.38

1.27

2.96

1.57

1.21

1.34

AMNH 21726-3
AMNH 21726-4

2.15

1.21

2.31

1.23

AMNH 21726-5

2.68

AMNH 21726-6

1.54

1.41

1.38

0.99

AMNH 21726-7

1.67

0.9

2.15

0.82

AMNH 21726-9

3.38

1.83

AMNH 21726-10

3.61

1.86

1.66

3.49

1.72

1.44

3.49

1.78

1.52

2.6

1.2

1.04

2.53

1.15

1.01

2.53

1.29

1.27

1.1556

AMNH 21726-11
AMNH 21726-12

2.65

1.27

2.28

1.13

2.15

0.96

2.88

1.27

1.26

3.16

1.36

1.45

AMNH 21726-13
AMNH 21726-14

1.79

1.91

2.04

1.08

AMNH 21726-15

2.82

1.32

1.37

AMNH 21726-19

3.04

1.39

1.14

AMNH 21740

2.83

1.2

AMNH 21740
AMNH 21742

1.21

0.76

1.35

0.94

1.44

1.06

1.79

1.02

AMNH 21743

2.14

1.17

2.06

AMNH 101983
AMNH 101985

1.79

IVPP w/n

0.77

1.94

1.64

1.27

3.45

1.5

1.48

2.31

1.21

2.64

1.26

1.04

3.31

1.47

1.21

2.47

1.14

2.82

1.33

1.21

3.62

1.87

1.57

2.65
1.18

2.28

1.08

2.75

1.42

1.52

3.42

1.82

1.84

0.98

2.12

2.55

1.03

1.15

3.15

1.41

1.26

1.62

1.11

1.96

1.42

2.31

1.99

2.19

0.86

1.44

0.96

1.51

0.95

1.81

1.04

2.1

1.13

2.34

1.21

2.77

1.33

1.3105

3.39

1.807

1.5709

2.72

1.361

Number of spec. 2

13

13

17

17

22

20

21

20

20

23

23

23

18

18

18

Stand. Deviat.

0.13

0.08

0.12

0.25

0.15

0.18

0.31

0.12

0.27

0.16

0.37

0.2

0.11

0.1386

0.17

0.442

0.18

0.28

0.118

0.0857

Mean

1.26

0.07

is usually well expressed and the parastyle is well developed. The labial
cingulum is small and is expressed at the mid-length of the ectoloph, but
usually does not form a mesostyle. The hypocone is usually much smaller
than the protocone, placed more lingually, and there is no sulcus separating the two cusps. The degree of development of the lingual cingulum is
very variable. On some specimens it is continuous, beginning at the
parastyle, ending at the metacone and uninterrupted on the lingual side of
the protocone. Then it forms two or more smaller cuspules lingual to the
protocone. On some specimens, the lingual cingulum forms two large
cuspules, situated directly lingual to the protocone and hypocone, respectively. Then, the cingulum is interrupted between these two cusps,

which are widely separated on some specimens (Fig. 2G).


The M2 is the largest molar. It is close in structure to M1, but
differs in details. The mesostyle is constantly present, together with a
short labial cingulum. The parastyle is well expressed. The hypocone
is very well developed, sometimes larger than the protocone and separated from the latter by a deep sulcus. The fosette is usually not expressed. As on M1 there are two kinds of lingual cingula. The first kind
forms a continuous lingual cingulum with two to five small cusps on its
lingual side. The other forms two large cusps directly lingual to the
protocone and hypocone, respectively, with the cingulum weak or absent between them.

199

FIGURE 3. Occlusal views of selected lower cheek teeth of Palaeostylops iturus.


A, PIN 3104-321, left p3-m3. B, PIN 3104-416, right p2-m3. C, PIN 3104-324,
left p3-m3.

The M3 is much reduced, the ectoloph tapering posteriorly and


the paracone large and situated more lingually than on the other molars. The ectocingulum is weak and short and does not form a mesostyle.
The postcingulum is lacking, and the precingulum is small, but constantly present.
The lower jaw is rather slender and elongate. The symphysis is
short; it ends under the anterior root of the p2. The mental foramina are
small. There are usually two to three pairs of mental foramina: one or
two anterior and one posterior pair. The anterior foramina are situated
under the c-p1, the posterior under the p3-4. The coronoid process is
rather tall. There is a small pit behind the m3 between the crest that
forms the anterior edge of the masseteric fossa and the crest extending
posteriorly from the m3.
The lower teeth are somewhat compressed laterally. The incisors, the canine and the p1 are single-rooted, and all the other teeth are
double-rooted. The incisors and the canine are procumbent. The incisors are simple and spatulate. The i1 crown is narrow, the i2 crown is
simple, wider than that of the i1, and the i3 crown is bicuspid, with the
anterior cuspid much larger. All the premolars and the canine are tricuspid. The p1 is suboval in shape, and the three cuspids are united by
the ectolophid. The middle cuspid (protoconid) is the largest; the anterior cuspid (paraconid) is equal in size to the posterior cuspid (metaconid). The p2 is larger than p1 and more elongate, with the tapering
posterior portion. The p3 is very similar to p2, slightly larger in size and
has a vestigial heel posterior to the metaconid. The p4 is slightly molarized,
with the paraconid well separated from the protoconid, which is the
largest cuspid, and the metaconid is situated close to the protoconid and
shifted laterally. The talonid is much larger than on the p3; it is about
one-third of the length of the trigonid. There are no cuspids expressed on

FIGURE 4. Bivariate plot of M2 measurements of Palaeostylops.

the talonid, but the cristid obliqua extends as far as the posterior edge of
the talonid.
Both the parastylid and paracristid are well developed on the
m1. The metaconid is also well developed. The talonid is much longer
than the trigonid. The talonid notch is very deep and wide, almost equal
in size to the talonid basin. The entolophid is strong, and the entoconid
is well expressed. The hypoconulid is generally more weakly developed than the entoconid and looks like a thickening on the posterior
edge of the ectolophid.
The m2 is much larger than the m1, and the paracristid is much
less developed than on the m1, whereas the parastylid is much stronger
and sometimes forms a separate cuspid. The metaconid is well expressed and much larger than the protoconid; the metacristid is short
and thick. The metaconid is almost posterior to and just a little lingual
to the protoconid. The entolophid is almost perpendicular to the
ectolophid, so the trigonid notch is close to rectangular in shape. The
talonid basin is rather narrow, tapering anteroexternally. The
hypoconulid is not expressed on the posterior edge of the ectolophid.
The m3 is much reduced and narrower than the other molars. The
parastylid is absent, the metaconid is well developed, but the metacristid
is much weaker than on the m2. The entolophid is curved, so it is perpendicular to the ectoloph at its labial part, but then it turns posteriorly, so
the talonid notch is widened lingually and the talonid basin deeper and
narrower than on the other molars. The hypoconulid is well expressed as
a large thickening on the posterior edge of the ectolophid.

TABLE 3. Measurements (in mm) of upper cheek teeth of Palaeostylops macrodon.


Coll. Number
AMNH 21144

I2 L

I2 W

I3 L

I3 W

CL

C W P1 L

P2 L P2 W P3 L
2.36

P3 W
1.87

AMNH 21726a-16
AMNH 21726a-17
AMNH 21740
AMNH 22142

2.19

AMNH 101962
AMNH 101963

1.87

1.37

2.15

2.05

1.88

P4 L
2.24

P4 W M1 L
2.48
3.38

M1 W
3.38

M2 L
5.27

M2 W
3.55

M3 L M3 W

2.20

2.05

3.04
3.23

2.96
3.25

4.71
4.62

3.76
3.60

2.56

3.10

2.24

2.61

3.30

3.61

5.26
4.45

4.24
3.82

2.64

3.67
3.06

3.61
3.87

2.27
2.55

2.82
3.08

2.18

2.34

2.06

2.60
3.14

2.76
3.02

4.57
4.51

2.88
2.78

3.21

4.36
4.43

AMNH 101964
AMNH 101966

1.90

3.63

2.18

2.91

AMNH 101979
AMNH 101986

2.02

1.77

1.01

1.77

1.73

1.71

1.98

2.37
2.08

3.00
3.05

3.11
3.16

4.78
4.51

3.99
4.54

2.81

0.79

2.02
2.09

2.20

1.72

2.00
2.18

Mean
Number of spec.

1.72
1.00

0.79
1.00

1.77
1.00

1.01
1.00

1.77
1.00

1.73
1.00

1.71
1.00

1.93
2.00

1.37
1.00

2.13
6.00

1.96
4.00

2.15
7.00

2.32
6.00

3.04
10.00

3.16
9.00

4.68
11.00

3.86
10.00

2.40
6.00

3.06
7.00

Stand. Deviat.

0.00

0.00

0.00

0.00

0.00

0.00

0.00

0.08

0.00

0.16

0.09

0.09

0.22

0.24

0.25

0.31

0.32

0.21

0.29

200
TABLE 4. Measurements (in mm) of lower cheek teeth of Palaeostylops macrodon.
Coll. Number i2 L
AMNH 101986

i3 L

cL

p1 L

p1 W

p2 L

p2 W

p3 L

p3 W

p4 L p4 W

m1 L
3.21

m1 W

m1 TW m2 L
4.05

m2 W
1.68

m2 TW m3 L
1.48

5.17
4.61

2.15
2.01

1.83
1.72

5.00
4.82

2.04
1.86

2.14
1.99

4.57
4.83

2.11
1.79

2.30
1.46

4.53
4.74

1.95
2.02

1.74
1.91

4.05

1.71

1.69

4.35

1.84

1.67

4.54

2.07

1.73

AMNH 21726a-8
AMNH 21726a-9

5.08
4.74

1.80
9.14

1.78
1.81

AMNH 21726a-10
AMNH 21726a-11

4.65
5.02

1.90
2.24

1.65
2.00

3.40

AMNH 21726a-12
AMNH 21726a-13

4.69
4.61

2.07
1.88

1.66
1.79

AMNH 21726a-14
AMNH 21726a-15

4.88
4.79

1.99
1.88

1.75
2.02

AMNH 2276
AMNH 2276
AMNH 2276
AMNH 22141

1.39

1.45

1.32

2.10
1.70

1.42

1.16

AMNH 21725
AMNH 101981
AMNH 101984
AMNH 20416

1.54

0.84

AMNH 21742
AMNH 21723

1.56
1.83

0.97
1.20

3.02
2.00

1.52

2.20
2.19

1.18

2.13
2.38

m3 W

m3 TW

2.84

1.18

1.46

2.46

1.22

1.09

2.98

1.56

1.33

2.89

1.57

1.25

2.99

1.62

1.44
1.30

2.63
2.65
2.77
2.92

1.12

1.69

3.33

1.48

3.84
3.34

AMNH 21726a-1
AMNH 21726a-2

2.19

1.06

2.41

1.22

AMNH 21726a-4
AMNH 21726a-5

2.28

1.32

2.62

1.44

2.71

1.21

AMNH 21726a-6
AMNH 21726a-7

3.44
3.42

1.43
1.23

1.56
1.75

3.77

1.84

3.33

1.45

1.54
1.34

1.45
1.68

1.00

Mean
1.39
Number of spec. 1.00

1.45
1.00

1.32
1.00

1.48
2.00

0.84
1.00

1.80
4.00

1.11
3.00

2.30
8.00

1.30
6.00

2.67
7.00

1.36
6.00

3.46
8.00

1.54
6.00

1.40
5.00

4.69
20.00

2.31
20.00

1.80
21.00

2.93
6.00

1.38
4.00

1.28
4.00

Stand. Deviat.

0.00

0.00

0.08

0.00

0.23

0.12

0.31

0.17

0.16

0.21

0.22

0.22

0.26

0.30

1.62

0.21

0.30

0.21

0.15

0.00

DISCUSSION
Matthew and Granger (1925) described Palaeostylops iturus based
on a lower jaw with all the cheek teeth (Fig. 1H-I, K). Later, Matthew et
al. (1929) described the second species of Palaeostylops, P. macrodon,
based on somewhat larger teeth (Fig. 1G, J). Cifelli et al. (1989), based on
the type specimen (AMNH 20414) and some dentulous dentaries and
maxillae of Palaeostylops iturus and on the holotype (AMNH 21725, a
left dentary), paratype (AMNH cast 109521) and several dentulous
dentaries and maxillae of Palaeostylops macrodon, stated that
Palaeostylops macrodon belongs to a new genus, Gashatostylops. The
following differences from Palaeostylops were listed in the diagnosis: (1)
relatively enlarged second upper and lower molars; (2) presence of
cuspules on the lingual cingula of upper molars; (3) weakness or absence
of the sulcus separating the lingual cusps of M1; (4) presence of two
rather than three upper incisors; and (5) laterally constricted snout with
dental arcade multiply curved.
As demonstrated by the extensive material from Tsagan-Khushu,
the teeth of arctostylopids are very variable in shape and structure (Figs.
2-3). Almost all the features listed by Cifelli et al. (1989) as the major
differences between Palaeostylops and Gashatostylops are found in different specimens of Palaeostylops iturus. Thus: (1) Palaeostylops iturus
has similarly enlarged upper and lower second molars, the slight difference can not be used as a generic difference; (2) there usually are several
small cuspules on the lingual cingula of the upper molars on some specimens of P. iturus; their number and size are very variable in both P. iturus
and P. macrodon; (3) the sulcus separating the lingual cusps of M1 is
absent only on the paratype specimen of P. iturus and is always present
on the specimens from Tsagan-Khushu; (4) the number of incisors in P.
macrodon cannot be established with certainty, because the specimen
upon which the conclusion about two rather than three incisors was
made is a badly damaged snout, and no incisors are preserved (Fig. 5); (5)

P. iturus has a very similarly constricted snout.


So, of the five characters listed as generic differences not one can
be used with certainty to distinguish these two genera. The only character mentioned by Cifelli et al. (1989) that could support the referral of P.
macrodon to a separate genus is the supposed reduced number of upper
incisors of P. macrodon. The specimen on which the conclusion about
the number of incisors in Palaeostylops macrodon was based on an
uncatalogued IVPP specimen from Bayan Ulan (Cifelli et al., 1989),
which is on loan at AMNH and is represented by a broken and badly
damaged skull (Fig. 5). Referred to P. macrodon by Cifelli et al. (1989), it
has no incisors, but fragments of their roots, and the cheek teeth are
present. Detailed study of this specimen reveals that it is within the size
range of P. iturus (Tables 1-4), but differs in dental morphology from
both described species of Palaeostylops and cannot be referred to this
genus with certainty, because of the narrower, more rounded molars. The
possibility that the IVPP specimen belongs to a third species was also
considered by Meng et al. (1998). The latter authors also mentioned that
the differences listed by Cifelli et al. (1989) were not observed among the
Bayan Ulan specimens of P. macrodon and could be insufficient to separate Gashatstylops as a different genus. Thus, the major difference between the two species of Palaeostylops is in size: P. iturus is smaller
(Tables 1-4). So, considering the lack of evidence that Gashatostylops
macrodon belongs to a separate genus we treat it as a species of
Palaeostylops and conclude that Gashatostylops is a junior subjective
synonym of Palaeostylops.
ARCTOSTYLOPIDA AND NOTOUNGULATA
Cifelli et al. (1989) argued that evidence from the dentition and
ankle region identifies a monophyletic Notoungulata that does not include the arctostylopids. We review here in detail their arguments to
support this assertion but conclude the opposite, that arctostylopids are

201

FIGURE 6. Upper (A) and lower (C) cheek teeth of Paleostylops iturus and upper
(B) and lower (D) cheek teeth of the notoungulate Notostylops murinus. A, C after
Matthew and Granger (1925); B, D after Simpson (1948).

FIGURE 5. Uncatalogued IVPP specimen, incomplete skull of ?Palaeostylops,


dorsal (A), left lateral (B) and ventral (C) views.

notoungulates.
Dentition
Arctostylopids and notoungulates have strikingly similar dentitions (Fig. 6), and this was long been the basis for inclusion of the
Arctostylopidae in the Notoungulata. However, Cifelli et al. (1989, p.
35) stated that derivation of the Arctostylopidae from within the order
[Notoungulata]would require many simplifications (reversals) in the
dentition. They thus noted that notoungulates have secondary molar complications such as crochets in the upper molars and M1-2
hypocones, features not seen in arctostylopids. Of course, the required
reversal is only a reversal if we assume that arctostylopids are derived from notoungulates, but if the reverse were the case, no evolutionary reversal took place. Furthermore, if the logic of Cifelli et al.
were correct, then no evolutionary simplification of crown pattern could
have occurred throughout mammalian evolution without a reversal.
Thus, the evolution of derived, bilophodont ceratomorph perissodactyl
molars from a more primitive and complex Hyracotherium-grade molarwhich all agree took placewould have occurred through evolutionary reversals according to Cifelli et al. (1989). Indeed, simplification of cheek-tooth crown morphology, from a multicuspate bunodont
form to lophodont or selenodont forms has happened many times in
eutherian evolution, especially in the Perissodactyla, Artiodactyla and
Proboscidea. Furthermore, Cifelli (1993, see especially fig. 15.6), in
his analysis of notoungulate phylogeny, argued for several simplifications of molar structure in notoungulate evolution, especially in the
derivation of some typotheres (intatheriids, mesotheriids and

hegetotheriids) and in the origin of derived toxodontids such as


Mixotoxodon. Simplification of crown morphology thus cannot be assumed to be an evolutionary reversal.
To further separate arctostylopids from notoungulates, Cifelli et
al. (1989) assessed the development of the upper molar hypocone in
notoungulates in a novel (different from the work of Patterson, 1934 and
Simpson, 1948, among others) and questionable fashion. The
posterolingual upper molar cusp of arctostylopids is not a hypocone,
but instead a pseudohypocone that is either a metaconule or a cusp
formed by a lingual extension of the ectoloph in the region of the metaconule (Cifelli et al., 1989, fig. 1). Cifelli et al. (1989, p. 36, fig. 12) argued
that M3 variation in the notoungulate Henricosbornia illustrates a plausible character state series for the addition of the posterolingual cusp on
primitive notoungulate anterior upper molars. They concluded that this
cusp is a derivative of the cingulum and therefore a true hypocone.
However, Simpsons (1945, fig. 48) illustration of M3 variants in
Henricosbornia lophodonta indicates otherwise (Fig. 7). Thus, the
postero-lingual cusp (crest) is readily seen to be a lingual extension of the
ectoloph, if not a true metaconule. It is thus not a hypocone, but a
pseudohypocone, as in arctostylopids. Therefore, Cifelli et al.s (1989,
p. 36) conclusion that the poterolingual upper molar cusp of southern
notoungulates and arctostylopids appears to have been acquired independently and in a nonhomologous fashion is in error.
Cifelli et al. (1989) also argued that the most primitive notoungulates
have submolariform posterior lower premolars in which p4 has a complete, curved talonid crescent. They conceded that the serially
multicuspate, blade-like p4s of Palaeostylops and some other
arctostylopids are a specialization (in their words, a secondary simplification), but noted that Asiostylops lacks the degree of molarization
seen even in henricosborniids. Of course, this does not preclude a
notoungulate-arctostylopid relationship if we see secondary simplification as a vector in notoungulate evolution. Also, note that among the
arctostylopids, both Bothriostylops and Peripanostylops have relatively
molariform p4s.
Cifelli et al. (1989) drew attention to Marshall et al. (1983), who
identified Perutherium as an arctostylopid and thereby homologized the
accessory trigonid cristids of notoungulate lower molars with the pre-

202

FIGURE 7. Occlusal views of M3s of the notoungulate Henricosbornia lophodonta showing variation (after Simpson 1948, fig. 48). Note that the postero-lingual cusp
on some specimens develops not from the postero-lingual cingulum, so it is not a hypocone. Instead it develops in the location of the metaconule or as a lingual extension
from the ectoloph (metaloph), so it is a pseudohypocone.

and postmetastylids. Most arctostylopids lack these accessory cristids,


so Cifelli et al. again saw this as a secondary loss. Ironically, and not
mentioned by Cifelli et al. (1989), Marshall et al. (1983) proposed derivation of arctostylopids from South American notoungulates via a simplification of lower molar trigonid struture. Again, Cifelli et al. s (1989)
argument assumes a notoungulate ancestry of arctostylopids and denies
the possiblity of simplification in dental evolution.
Indeed, their cladogram of arctostylopid relationships (Cifelli et
al., 1989, fig. 11; also see Cifelli and Schaff, 1998, fig. 21.3) assumes that
primitive arctostylopids have more simple molar crowns (e.g.,
Asiostylops), and more derived arctostylopids have more complex molar
crowns (e.g., Anatolostylops). The reverse is equally plausible. Further
arguments presented by Cifelli et al. (1989) against a notoungulatearctosylopid evolutionary vector, based on comparing henricosborniid
premolar and molar structure to that of arctostylopids, are also based on
unwarranted assumptions about possible evolutionary vectors and reversals.
The diagnosis of Arctostylopida presented by Cifelli et al. (1989,
p. 5; essentially repeated in Cifelli and Schaff, 1998) lists only dental
characteristics, most of which are found in notoungulates:
1. Upper and lower dentitions forming an evenly graded series and canines poorly or not differentiated and without diastemata
separating them from adjacent teethfeatures present in some
notoungulates (e.g., Colbertia, Homaladotherium), though most
notoungulates have diastemata and sharp differentiation of the incisors, canines and premolars.
2. Posterior upper premolars somewhat molarized except in
Asiostylops; P4, at least, with a metaconemolarization of upper
premolars and the presence of P3 metacones are virtually ubiquitous in
notoungulates.
3. Upper molars with well-developed centrocrista, becoming a
salient, straight ectoloph in advanced generathe prominent upper
molar ectolophs are a striking similarity shared by arctostylopids and
notoungulates.
4. Parastyle usually prominentAgain, seen in virtually all
notoungulates.
5. Pre- and postprotocristae of upper molars strong, conules
lackinga characteristic evident on many notoungulate upper molars.
6. Upper molars primitively triangular but M1-2 becoming quadrate in advanced forms by the addition of a posterolingual cusp
(pseudohypocone)as discussed above, this assumes a vector (triangular evolves to quadrate) in arctostylopid evolution, but the range of
shapes of the upper molars of notoungulates, and the manner in which a
posterolingual cusp (pseudohypocone) is formed (see above) is the same
as in arctostylopids.
7. Anterior lower premolars serially tricuspid, with strong shearing surfacesthis feature does distinguish arctostylopids from

notoungulates, but as Cifelli et al. (1989) noted, it is a specialization of


some arctostylopids, and as we note, both Arctostylops and
Borthriostylops show a more molariform p4 not that different from
notoungulate p4s.
8. Lower molars primitively biselenodont, with paracristid lost
and various secondary trigonid structures acquired in advanced taxa
the biselenodont lower molars of arctostylopids are very similar to those
of many notoungulates.
9. Lower molar hypoconid indistinct; entoconid transversely
expanded and, in advanced forms, developed into an antero-bucally
oriented entolophidagain, features well displayed by many
notoungulates.
Ankle
Cifelli et al. (1989, p. 36; also see Cifelli, 1983) posited the
following as synapomorphies of the notoungulate ankle: a long, constricted astragalar neck with an oblique dorsal crest; astragalar body
with a median (tibial) protuberance; astragalar foramen with posterolateral sulcus interrupting continuity of the tibial trochlea and flexor
tendon groove; and well-developed sustentacular-navicular facet contact on the astragalus. They went on to identify isolated calcanea and
astragali from Mongolia, which lack an unambiguous dental association, as those of Palaeostylops and Gashatostylops (Cifelli et al.,
1989, fig. 7). These bones differ from those of notoungulates in several
striking features, which Cifelli et al. (1989, p. 36-37) identified as
arctostylopid specializations: astragalus with a cylindrical, verticallywalled body, the tibial trochlea extensively developed anteroposteriorly;
lack of a fibular shelf; navicular facet developed so that the axis of
movement along the midtarsal joint would have been roughly parallel
(rather than oblique) to that of the proximal ankle joint; astragalar cuboid
facet lost(?); ectal facet steeply inclined with respect to inferior surface
of astragalus; calcaneal fibular facet strongly devloped into a semicylindrical surface; and sustentaculum of calcaneus distally located, at
or near distal (cuboid) end of the bone. Indeed, these features do differentiate the astragalus-calcaneum Cifelli et al. (1989) assign to
arctostylopids from notoungulate ankles. But, there are two reasons
why these distinctions are meaningless to assessing arctostylopidnotoungulate relationships:
1. The astragalus-calcaneum morphology that Cifelli et al.
attributed to Mongolian arctostylopids is virtually identical to that of
the anagalid Pseudictops, a Gashatan taxon that co-occurs with those
ankle bones in Mongolia (compare Cifelli et al., 1989, fig. 7 to Sulimski,
1968, pl. 13, figs. 3-4, pl. 14, figs. 2-3). Indeed, the specializations of
arctostylopid ankles listed by Cifelli et al. (1989), and quoted above,
are a good description of the salient features of the ankle of Pseudictops.
Therefore, it seems more likely that the ankle bones Cifelli et al. identified as arctostylopid are actually those of Pseudictops (or a closely
realted anagalid) which renders their anatomy irrelevant to an assess-

203
ment of arctostylopid-notoungulate relationships.
2. Bloch (1999) presented initial data on a partial skeleton of
Arctostylops from the Paleocene of Wyoming, which includes an unambiguously associated (found semi-articulated) ankle and dentition.
This ankle does not resemble the ankles from Mongolia assigned to
arctostylopids by Cifelli et al. (1989), and that increases our confidence in the conclusion that the Mongolian ankles are not arctostylopid,
and should be reassigned to Pseudictops. Bloch (1999, p. 32A) stated
that the ankle from Wyoming resembles that of notoungulates in possessing a long, constricted astragalar neck with an oblique dorsal crest;
the sustentaculum of the calcaneum proximally positioned; and a tibial
protuberance on the astragalus. He also noted (p. 32A) that the Wyoming ankle differs from that of notoungulates in lacking a fibular shelf
and having a steeply inclined ectal facet. These similarities and differences thus do not preclude a close arctostylopid-notoungulate relationship. Pending a full description and analysis of the postcranial skeleton of Arctostylops from Wyoming, we conclude that there is no postcranial evidence that excludes arctostylopids from the Notoungulata.

CONCLUSION
Thus, there seems to be little basis for separating arctostylopids
from notoungulates. They share many striking dental similarities that are
not the result of convergence (as Cifelli et al., 1989 would have us believe) but that are, as long concluded, evidence of a close relationship. We
thus return the arctostylopids to the Notoungulata as a family of
notoungulates primarily distingished by their simplifications of tooth
crown morphology and their serially tricuspid lower premolars (of most
genera).
ACKNOWLEDGMENTS
R. Tedford made it possible to study specimens in the AMNH.
Funding from NMMNH foundation made this study possible. The work
of PK was supported by the Russian Foundation for Basic Research,
project 02-04-48458, and the Grant Council of the President of Russian Federation, project NSh-1840.2003.4.

REFERENCES
Bloch, J. I., 1999, Partial skeleton of Arctostylops from the Paleocene of Wyoming:
Arctostylopid-notoungulate relationship revisited: Journal of Vertebrate Paleontology, v. 19, supplement to no. 3, p. 32A.
Chow, M. and Qi, T., 1978, Paleocene mammalian fossils from Nomogen Formation of Inner Mongolia: Vertebrata Palasiatica, v. 16, p. 77-85.
Cifelli, R. L., 1983, Eutherian tarsals from the late Paleocene of Brazil: American
Museum Novitates, no. 2761, 31 p.
Cifelli, R. L., 1993, The phylogeny of the native South American ungulates; in
Szalay, F. S., Novacek, M. J. and McKenna, M. C., eds., Mammal phylogeny:
Placentals: New York, Springer-Verlag, p. 195-216.
Cifelli, R. L. and Schaff, C. R., 1998, Arctostylopida; in Janis, C. N., Scott, K. M.
and Jacobs, L. L., eds., Evolution of Tertiary mammals of North America: Cambridge, Cambridge University Press, p. 332-336
Cifelli, R. L., Schaff, C. R. and McKenna, M. C., 1989, The relationships of
Arctostylopidae (Mammalia): New data and interpretation: Bulletin of the Museum of Comparative Zoology, v. 152, p. 1-44.
Dashzeveg, D. and Russell, D.E., 1988, Paleocene and Eocene Mixodontia (Mammalia, Glires) of Mongolia and China: Palaeontology, v. 31, p. 129-164.
Gingerich, P. D. and Rose, K. D., 1977, Preliminary report on the American Clark
Fork mammalian fauna, and its correlation with similar faunas in Europe and
Asia: Gobios Mmoire Special 1, p. 39-45.
Huang, X. and Zheng, J., 2003, Note on two new mammalian species from the late
Paleocene of Nanxiong, Guangdong: Vertebrata Palasiatica, v. 41, p. 271-277.
Huang, X., Zheng, J. and Ding, S., 2001, Arctostylopid fossil (Mammalia) of
Changtao basin, Hunan and comments on related stratigraphy: Vertebrata
PalAsiatica, v. 39, p. 14-23.
Kinman, K.E., 1994, The Kinman System: Toward A Stable Cladisto-Eclectic Classification of Organisms (Living and Extinct: 48 Phyla, 269 Classes, 1,719 Orders. Hays, Kan., K.E. Kinman, 81 pp.
Kondrashov, P. and Lucas, S. G., 2002, Late Paleocene Arctostylopidae (Mammalia, Notoungulata) from Mongolia: Journal of Vertebrate Paleontology, v. 22,
supplement to no. 3, p. 75A.
Marshall, L. G., Muizon, C. de and Sig, B., 1983, Perutherium altiplanense, un
notongul du Crtac Suprieur du Prou: Palaeovertebrata, v. 13, p. 145-155.
Matthew, W.D., 1915, A revision of the lower Eocene Wasatch and the Wind River
faunas. Part IV. Entelonychia, Primates, Insectivora (part): Bulletin of the American Museum of Natural History, v. 34, p. 429-483.
Matthew, W.D. and Granger, W., 1925, Fauna and correlation of the Gashato For-

mation of Mongolia: American Museum Novitates, no. 189, p. 1-12.


Matthew, W.D., Granger, W. and Simpson, G.G., 1929, Additions to the fauna of
the Gashato Formation of Mongolia: American Museum Novitates, no. 376, p.
1-12.
McKenna, M. C. and Bell, S. K., 1997, Classification of mammals above the species level. New York, Columbia University Press, 631 p.
Meng, J., Zhai, R. and Wyss, A.R., 1998, The late Paleocene Bayan Ulan fauna of
Inner Mongolia, China: Bulletin of the Carnegie Museum of Natural History, v.
34, p. 148-185.
Nessov, L.A., 1987, Results of searches and investigations in the mammal-bearing
Cretaceous and early Paleogene in the territory of the USSR: Annual of Allunion Paleontological Society, v. 30, p. 199-218. [in Russian]
Patterson, B., 1934, Upper molar structure in the Notoungulata, with notes on taxonomy: Geological Series Field Museum of Natural History, v. 6(6), p. 91-111.
Rose, K.D., 1981, The Clarkforkian land-mammal age and mammalian composition across the Paleocene-Eocene boundary: University Michigan, Papers on
Paleontology, v. 26, p. 1-189.
Schlosser, M., 1923, Saugetiere; in Broili, F. and Schlosser, M., eds., K.A. von Zittel,
Gundzge der Palaeontologie, Neubeart: Munich, M.R. Oldenbourg, p. 402689.
Simpson, G. G., 1934, Provisional classification of extinct South American hoofed
mammals: American Museum Novitates, no. 750, p.1-21.
Simpson, G. G., 1945, The principles of classification and a classification of mammals: Bulletin of the American Museum of Natural History, v. 85, p. 1-350.
Simpson, G. G., 1948, The beginning of the age of mammals in South America.
Part 1: Bulletin of the American Museum of Natural History, v. 91, p. 1-232.
Sulimski, A., 1968, Paleocene genus Pseudictops Matthew, Granger & Simpson,
1929 (Mammalia) and its revision: Palaeontologia Polonica, v. 19, p. 101-129.
Tang, Y. and Yan, D., 1976, Notes on some mammalian fossils from the Paleocene
of Qianshan and Xuancheng, Anhui: Vertebrata PalAsiatica, v. 14, p. 91-99.
Zhai, R., 1978, More fossil evidences favoring an early Eocene connection between
Asia and Neoarctic: Memoirs of the Institute of Vertebrate Paleontology and
Paleoanthropology, Academia Sinica, no. 13, p. 107-115.
Zheng, J., 1979, The Paleocene notoungulates of Jiang-xi; in Symposium on Cretaceous and early Tertiary red beds of South China: Beijing, Science Press, p.
387-394.
Zheng, J. and Huang, X., 1986, New arctostylopids (Notoungulata, Mammalia)
from the late Paleocene of Jiangxi: Vertebrata PalAsistica, v. 24, p. 121-128.

204

Dissacus navajovius, skull and lower jaw (from Matthew, 1937).

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

205

THE OLDEST KNOWN ASIAN ARTIODACTYL (MAMMALIA)


PETER E. KONDRASHOV1,2, ALEXEY V. LOPATIN2 AND SPENCER G. LUCAS3
1
2

Department of Biology, Northwest Missouri State University, 800 University Drive, Maryville, MO 64468;
Paleontological Institute of the Russian Academy of Sciences, Profsoyuznaya 123, Moscow 117868, Russia;
3
New Mexico Museum of Natural History, 1801 Mountain Road N. W., Albuquerque, NM 87104

AbstractA new artiodactyl genus and species, Tsaganohyus pecus, from the earliest Eocene of Mongolia is
described, based on a dentary with dp4-m1. The new genus shows close resemblance to North American
Homacodontinae. The most similar Asian genus, Siamotherium, belongs to the family Anthracotheriidae and is
known from Myanmar. The new form is too advanced to belong to Diacodexeinae because of the more bunodont
teeth, but less advanced than Anthracotheriidae, to which it probably has distant affinities. The closest group is
the subfamily Homacodontinae from the early-middle Eocene of North America, which probably migrated into
Asia via a boreal land bridge together with other North American genera, such as Hyopsodus, Hyracotherium
and Diacodexis.
Keywords: Artiodactyl, Eocene, Mongolia, Tsaganohyus, Homacodontinae
INTRODUCTION
Artiodactyls appeared almost simultaneously in North America,
Europe and Asia during the early Eocene. Some have argued that the
earliest artiodactyls evolved from an arctocyonid condylarth ancestor
(Rose, 1996), although artiodactyl origins remain unclear (Gentry and
Hooker, 1988; Stucky, 1998). The Paleocene Chriacus looks like a good
model for the ancestor of Diacodexis. If so, artiodactyls originated in
North America, and the diversity of early European and Asian artiodactyls is due to faunal interchange. This correlates well with the early
Eocene occurrence of the typical North American condylarth genus
Hyopsodus both in Europe (Hooker, 1979) and Asia (Dashzeveg, 1977;
Kondrashov and Agadjanian, 1999), and the genus Phenacodus in Europe (Crusafont, 1956).
Thewissen et al. (1983) described a new species of Diacodexis
from the lower Eocene of Pakistan and, based on its small size and
primitive morphology, argued that artiodactyls originated in Asia, although they excluded Diacodexis pakistanensis from the direct ancestry of the entire order. The age of this faunule was later considered to
be earliest Eocene (Leinders et al., 1999), although no new data were
presented to support this conclusion. Recently, several authors doubted
the referral of Diacodexis pakistanensis to the genus Diacodexis
(Gentry and Hooker, 1988; Stucky, 1998) and referred to it as an outgroup
to the primitive artiodactyls.
A wide diversity of artiodactyls is known from the early-middle
Eocene of Pakistan, Myanmar and Thailand. They belong to at least
three groups: diacodexeids (Gingerich et al., 1979; Thewissen et al.,
1983, 1987), helohyids (Holroyd and Ciochon, 1995; Ducrocq et al.,
1997) and anthracotheriids (Suteethorn et al., 1988; Ducrocq, 1999;
Ducrocq et al., 2000; Tsubamoto et al., 2002).
Compared to southern Asia, artiodactyls are much rarer in the
early Eocene of northern Asia. They are not even included in the
Bumbanian (early Eocene) faunal list (Ting, 1998). Dashzeveg (1982)
mentioned Artiodactyla indeterminate from the Bumban Member of
the Naran Bulak Formation (Mongolia). Gabunia (1971, 1973) described
two new genera of artiodactyls, Aksyiria and Paraphenacodus, from
the early middle Eocene of Kazakstan. Averianov (1996) described a
new diacodexeid genus and species, Eolantianius russelli, from the
lower Eocene of Kyrgyzstan. All these taxa, with some certainty, can
be referred to Dichobunidae or Diacodexidae. Tong and Wang (1998)
assigned a lower jaw fragment from the early Eocene Wutu fauna in
China to Entelodontidae as gen. et sp. nov.
The systematics of early artiodactyls is complicated and poorly
understood. The two latest revisions introduced a great deal of confusion by claiming that the traditional systematic groups of artiodactyls
are paraphyletic, and many new taxa were created (Gentry and Hooker,

1988; Stucky, 1998). In both listed publications, it is stated that


Dichobunidae, Homacodontidae and Diacodexis are not monophyletic.
Gentry and Hooker proposed a new family of Dichobunoidea, the
Bunomerycidae, to include all homacodontines except Homacodon.
Averianov (1996) reassessed the value and arrangement of the characters for the cladogram proposed by Gentry and Hooker (1988) and stated
that the conclusion about the paraphyletic nature of Diacodexis is rather
doubtful or, at least, premature. Here, we accept the arrangement of
early artiodactyls proposed by McKenna and Bell (1997).
The new artiodactyl described here does not fall in the range of
the early artiodactyl subfamilies known from Asia, so we assign it to
the North American dichobunid subfamily Homacodontinae based on
overall similarity and specific synapomorphies. Homacodontinae are
known from the early-middle Eocene of North America (West, 1984;
McKenna and Bell, 1997; Stucky, 1998), and do not include the
homacodontines described from western Europe (Erfurt, 1993; Sudre
and Marandat, 1993).
PROVENANCE
The Tsagan-Khushu locality, from which the artiodactyl specimen described here was collected, is situated in the Nemegt depression in southwestern Mongolia. Vertebrate remains come from the facies transitional from fluviatile to lacustrine sediments at the base of
the Bumban Member, which is early Eocene in age.
The specimen was obtained by screenwashing well-rounded
gravel with a silty sand matrix from small lenses 1.5 x 0.5 m in size,
situated in sandy variegated clays, by V.Yu. Reshetovs field party of
the Joint Soviet-Mongolian Paleontological Expedition in 1985. The
specimen is housed in the Paleontological Institute of the Russian Academy of Sciences, Moscow, Russia (PIN). The specimen was mentioned
as Artiodactyla gen. et sp. nov. by Kondrashov et al. (2001)
SYSTEMATIC PALEONTOLOGY
Order Artiodactyla
Superfamily Dichobunoidea Turner, 1879
Family Dichobunidae Turner, 1879
Subfamily Homacodontinae Marsh, 1894
Tribe Homacodontini Marsh, 1894
Genus Tsaganohyus, new genus
Type speciesTsaganohyus pecus, new species.
Included species Known only from the type species.
Diagnosis Small artiodactyl with bunodont lower molars, al-

206

FIGURE 1. Tsaganohyus pecus, PIN 3104-479, holotype, occlusal (A) and lingual
(B) views of left dp4-m1. Scale bar = 2 mm.

most lacking distinct cristids. On m1, the paraconid is considerably


reduced, shifted medially and is connected by cristids to both the protoconid and metaconid. The hypoconid and entoconid are of equal size; the
hypoconulid is much smaller and is situated posterior to the former
cuspids in a median position on m1. The posterocingulid forms the
flanks of the hypoconulid. There is a well-expressed entoconulid on the
anterior slope of the entoconid. The dentary is high, and the symphysis
is relatively long, reaching the p3.
Distribution Lower Eocene of Mongolia.
Etymology Tsagan - white [Mongolian], for the type locality,
Tsagan-Khushu, and hyus [Greek] - pig.
Tsaganohyus pecus, new species
Figs. 1-2
Holotype PIN 3104-479, left dentary fragment with dp4-m1
and alveoli of c-p3.
Diagnosis Same as for the genus.
Type locality Base of the Bumban Member of the Naran Bulak
Formation at the Tsagan-Khushu locality.
Etymology Pecus [Latin] - small cattle.
Description The dentary is high. The symphysis is longit
reaches the anterior root of the p3. The angle between the symphyseal
suture and the dentary is about 45 degrees. The anterior mental foramen
is situated under the p1. The posterior mental foramen is situated under
the anterior root of the p4. Two teeth are preserved in the holotype, the
dp4 and m1. Eight alveoli are preserved in the dentary anterior to the
dp4. The five alveoli anterior to the dp4 are equal in size, and we interpret them as the two pairs of alveoli for the double-rooted p2 and p3 and
an alveolus for a single-rooted p1. The alveolus anterior to the p1 is larger
than the others, so it evidently was for the canine. There are two alveoli
for the single-rooted i2 and i3 anterior to the canine.
The dp4 is elongate, with a sharp, anteriorly-projecting paraconid
lobe that is separated from the rest of the trigonid by the relatively deep
trigonid basin posteriorly and by the weak lingual flexus. The lobe is
inclined 45 degrees to the tooth row. The paraconid itself is situated on
the anteriormost part of the lobe. It is small and not pointed. The paraconid
is connected by the protolophid with the protoconid and by a weaker
cristid with the metaconid. The metaconid is slightly larger than the

FIGURE 2. Tsaganohyus pecus, PIN 3104-479, holotype, labial (A), occlusal (B)
and lingual (C) views of left dentary fragment with dp4-m1.

protoconid, both cuspids being more crescentic than on the m1. The
protoconid is connected to the metaconid by the wide and low metalophid.
Although the metaconid part of the tooth is heavily worn, there is a
metastylid, which is of relatively large size. The talonid is slightly wider
than the trigonid. The cristid obliqua begins at the posterolabial wall of
the metaconid. It is low, wide and diminishes posteriorly, reaching the
hypoconid as an indistinct wrinkle. The entoconid is larger than the
hypoconid. The hypoconulid is heavily worn to the base, but judging
from the facet it was situated in the midline of the tooth directly posterior to the hypoconid and entoconid. None of the talonid cuspids are
connected by cristids. Judging from the shape of the wear facet of the
entoconid and compared with the same structure on the unworn m1, we
conclude that there was a well-developed entoconulid anterior to the
entoconid. There are no cingulids on the dp4.
The m1 is almost rectangular in shape, and the talonid is slightly
wider than the talonid. The cuspids are more bulbous and less crescentic than on the dp4. The paraconid is central in position, anteriorly
projecting, but not forming such a strong paraconid lobe as on the dp4.
Although much smaller than the two other trigonid cuspids, the
paraconid is relativelly well developed and closely appressed to the
hypoconulid of the dp4. The protolophid is well developed; the cristid
between the paraconid and metaconid is much weaker but present. The
metaconid is larger than the protoconid, and these cuspids are not connected by cristids. The cristid obliqua is very weak; it runs from the top
of the metaconid to the base of the hypoconid. There is a small wear
facet closer to the base of the metaconid on the oblique cristid. There is
another cristid on the posterior slope of the metaconid, situated more
lingually than the cristid obliqua, that runs to the base of the entoconulid.
The hypoconid is slightly larger than the entoconid. The hypoconulid is
well-developed, pointed, and it is situated at the midline of the tooth,
posterior to the hypoconid and the entoconid. The posterocingulid forms
the lateral wings of the hypoconulid. There is a well-developed cus-

pid, the entoconulid, on the anterior slope of the entoconid.


Measurements (in mm):
Tooth
Length Trigonid width
Talonid width
dp4
2.9
1.5
1.6
m1
3.1
1.9
2.15
DISCUSSION
The new genus is the oldest known Asian artiodactyl as it comes
from the base of the Bumban Member of the Naran Bulak Formation,
so its age is close to the beginning of the early Eocene. The new genus
certainly belongs to Artiodactyla. The most similar group includes
phenacodontid condylarths, which are also known from the early Eocene
of Asia (Wang and Tong, 1997). The major differences include more
bunodont teeth in Tsaganohyus, reduction of crests, which are well
developed in the phenacodont Lophocion, and very high dentary with a
long symphysis. The hypoconulid structure of the m1 of Tsaganohyus
is not found in any known condylarth, including possible artiodactyl
ancestors among the Arctocyonidae.
On the other hand, the new genus differs considerably from all the
other Asian artiodactyls and cannot be referred to the artiodactyl subfamilies already known from the Asian subcontinent. Tsaganohyus is
not a diacodexine, as it has different morphology, which is more derived
on one hand and less on the other. The teeth of the new genus are more
bunodont than in any known dichobunid; the paraconid on m1 is much
reduced, the hypoconid and entoconid are almost equal in size, and the

207
hypoconulid is very different from that of any Diacodexeinae, because it
is separated from both the hypoconid and entoconid and lies on the flank
of the posterocingulid. All the crests are very weak. These characters
also differentiate Tsaganohyus from Helohyidae.
Of all the described Asian early artiodactyl genera, the closest in
morphology is Siamotherium krabiense from the upper Eocene of
Myanmar (Suteethorn et al., 1988) that was assigned to the
anthracotheriids. Described later, Siamotherium pondaungensis from the
middle Eocene of Thailand (Ducrocq et al., 2000), was synonymized
with Pakkokuhyus lahirii (Helohyidae) (Tsubamoto et al., 2002). The
morphology of the new genus resembles much the morphology of
Siamotherium, especially in the median position of the m1 paraconid and
the structure of the crests, connecting the latter with the protoconid and
metaconid. Among the North American and European forms, the new
genus closely resembles Homacodon from the early-middle Eocene of
North America. The difference is mostly in the more bunodont teeth of
the new genus, lacking distinct crests.
Nevertheless, the following synapomorphies can be identified
in Tsaganohyus and typical homacodontines: (1) the paraconid on m1
is reduced in a similar way; (2) the hypoconulid is situated on the
posterocingulid, posterior to the hypoconid and entoconid, and is not
connected by cristids to either; and (3) there is a small but well expressed entoconulid. These characters allow us to refer the new genus
to the tribe Homacodontini in the subfamily Homacodontinae, which
we include in Dichobunidae.

REFERENCES
Averianov, A.O., 1996, Artiodactyla from the early Eocene of Kyrgyzstan:
Palaeovertebrata, v. 25, p. 359-369.
Crusafont, P.M., 1956, Otro nuevo condilartro del Luteciense Pirenaico: Estratto
dal Bulletino della Societa Geologica Italiana, v. 75, p. 1-8.
Dashzeveg D., 1977, O pervoi nakhodke Hyopsodus Leidy, 1870 (Mammalia,
Condylarthra) v Mongolskoy Narodnoy respublike [About the first occurrence
of Hyopsodus Leidy, 1870 (Mammalia, Condylarthra) in the Mongolian Peoples
Republic]: Trudi Sovmestnoy Sovetsko-Mongolskoy Expeditsii [Proc. of the
Joint Soviet-Mongolian Expedition], v. 4, p. 7-13.
Dashzeveg, D., 1982, La faune de Mammiferes du Paleogene inferieur de Naran
Bulak (Asie Centrale) et ses correlations avec lEurope et lAmerique du
Nord.:Bulletin Societe Geologique de France, v. 7(24-2), p. 275-281.
Ducrocq, S., 1999, The late Eocene Antracotheriidae (Mammalia, Artiodactyla)
from Thailand: Paleontographica Abt. A, v. 252, p. 93-140.
Ducrocq, S., Aung, N.S., Aung, A.K., Benammi, M., B,o B., Chaimanee, Y., Tun,
T., Thein, T., and Jaeger, J., 2000, A new anthracotheriid artiodactyl from
Myanmar, and the relative ages of the Eocene anthropoid primate-bearing localities of Thailand (Krabi) and Myanmar (Pondaung): Journal Vertebrate Paleontology, v. 20, p. 755-760.
Ducrocq, S., Chaimanee, Y., Suteethorn, V. and Jaeger, J., 1997, First discovery of
Helohyidae (Artiodactyla, Mammalia) in the late Eocene of Thailand: A possible transitional form for Anthracotheriidae: Compte Rendus Academie Science Paris, v. 325, p. 367-372.
Erfurt, J., 1993, Interim report on taxonomy and biostratigraphy of Artiodactyla
from the Middle Eocene locality Geiseltal near Halle (Germany): Kaupia, v. 3,
p. 131-136.
Gabunia, L. K., 1971, O novom predstavitele condilartr (Condylarthra) iz eotsena
Zaisanskoi Kottovini [On a new representative of condylarths (Condylarthra)
from the Eocene of Zaisan depression]: Soobshcheniya Gruzinskoy Akademii
Nauk [Proceedings of the Georgian Academy of Sciences], v. 61, p. 234-235.
Gabunia, L. K., 1973, O prisutstvii diakodexein (Diacodexinae) v Eotsene Azii [On
the existence of diacodexines (Diacodexinae) in the Eocene of Asia]:
Soobshcheniya Gruzinskoy Akademii Nauk [Proceedings of the Georgian Academy of Sciences], v. 71, p. 741-744.
Gentry, A. W. and Hooker, J. J., 1988, The phylogeny of the Artiodactyla; in Benton,
M. J., ed., The phylogeny and classification of the tetrapods. Volume 2. Mam-

mals: Oxford, Clarendon Press, p. 235-272.


Gingerich, P. D., Russell, D. E., Sigogneau-Russell, D. and Hartenberger, J.-L., 1979,
Chorlakkia hassanni a new Middle Eocene dichobunid (Mammalia, Artiodactyla) from the early-middle Eocene of Kashmir: Contributions Museum Paleontology, University of Michigan, v. 25, p. 71-77.
Holroyd, P. A. and Ciochon, R. L., 1995, A new artiodactyl (Mammalia) from the
Eocene Ponduang Sandstones, Burma: Annals of the Carnegie Museum, v. 64,
p. 177-183.
Hooker, J. J., 1979, Two new condylarths (Mammalia) from the early Eocene of
southern England: Bulletin of the British Museum of Natural History, v. 32, p.
43-56.
Kondrashov, P.E. and Agadjanian, A. K., 1999, New material on the genus
Hyopsodus (Mammalia, Condylarthra) from the Eocene of Mongolia: Paleontological Journal, v. 33, p. 667-676.
Kondrashov, P.E., Lopatin, A.V. and Lucas, S.G., 2001, Early Eocene (Bumbanian)
mammal fauna from Tsagan Khushu locality (Mongolia): J. Vertebrate Paleontology, v. 21. suppl. no 3., p. 69 A.
Leinders, J .J. M., Arif, M., Bruijn, H., Hussain, S. T. and Wessels, W., 1999, Tertiary continental deposits of northwestern Pakistan and remarks on the collision
between the Indian and Asian plates; in Reumer, J.W.F. and De Vos, J., eds.,
Elephants have a snorkel! Papers in Honour of Paul Y. Sondaar: Deinsea, v. 7,
p. 199-213
McKenna, M.C. and Bell, S.K., 1997, Classification of mammals above the species
level: New York, Columbia University Press, 631 p.
Rose, K. D., 1996, On the origin of the order Artiodactyla: Proceedings of the National Academy of Sciences USA, v. 93, p. 1705-1709.
Stucky, R. K., 1998, Eocene bunodont and bunoselenodont Artiodactyla
(dichobunids); in Scott, K. M., Janis, C. M. and Jacobs, L. L., eds., Tertiary
mammals of North America: Cambridge, Cambridge University Press, p. 358374
Sudre, J. and Marandat, B., 1993, First discovery of an Homacodontinae (Artiodactyla, Dichobunidae) in the middle Eocene of Western Europe: Eygalayodon
montenati new genus, new species. Considerations on the evolution of primitive
artiodactyls: Kaupia, v. 3, p. 157-164.
Suteethorn, V., Buffetaut, E., Helmcke-Ingavat, R., Jaeger, J. J. and
Jongkanjanasoontorn, Y., 1988, Oldest known Tertiary mammals from south-

208
east Asia: Middle Eocene primate and anthracotheres from Thailand: Neues
Jahrbuch fr Geologie und Palontologie Monatshefte, v. 9, p. 563-570.
Thewissen, J. G. M., Gingerich, P. D. and Russell, D. E., 1987, Artiodactyla and
Perissodactyla (Mammalia) from the early-middle Eocene Kuldana Formation
of Kohat (Pakistan): Contributions Museum Paleontology, University of Michigan, v. 27, p. 247-274.
Thewissen, J. G. M., Russell, D. E., Gingerich, P. D. and Hussain, S. T., 1983, A
new dichobunid artiodactyl (Mammalia) from the Eocene of northwest Pakistan: Proceedings of the Koniklijke Nederlandse Academie van Wetenschappen,
Series B, v. 86, p. 153-180.
Ting, S., 1998, Paleocene and early Eocene land mammal ages of Asia: Bulletin of
the Carnegie Museum of Natural History, v. 34, p.124-147.
Tong, Y. and Wang, J., 1998, A preliminary report on the early Eocene mammals of
the Wutu fauna, Shandong Province, China: Bulletin of the Carnegie Museum

of Natural History 34, p. 186-193.


Tsubamoto, T., Takai, M., Egi, N., Shigehara, N., Tun, S.T., Aung, A.K., Aung,
N.S. and Thein, T., 2002, The Anthracotheriidae (Mammalia; Artiodactyla)
from the Eocene Pondaung Formation (Myanmar) and comments on some other
anthracotheres from the Eocene of Asia: Paleontological Research, v. 6, p. 363384.
Wang, J. and Tong, Y., 1997, A new phenacodontid condylarth (Mammalia) from
the Early Eocene of the Wutu basin, Shandong: Vertebrata Palasiatica, v. 35, p.
283-289.
West, R. M., 1984, Paleontology and geology of the Bridger Formation, southern
Green River Basin, southwestern Wyoming. Part 7. Survey of Bridgerian Artiodactyla, including description of skull and partial skeleton of Antiacodon
pygmaeus: Contributions in Biology and Geology, Milwaukee Public Museum,
no. 56, p. 1-47.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

209

REVISED DISTRIBUTION OF CONDYLARTHS (MAMMALIA, EUTHERIA) IN ASIA


PETER E. KONDRASHOV1 AND SPENCER G. LUCAS2
1
Northwest Missouri State University, 800 University Dr., Maryville, MO 64468
and Paleontological Institute of the Russian Academy of Sciences, Moscow, Russia;
2
New Mexico Museum of Natural History, 1801 Mountain Road NW, Albuquerque, NM 87104

AbstractRepresentatives of six condylarth families have been described from Asia, but a detailed review
demonstrates that only three families actually occur in Asia: Hyopsodontidae, Phenacodontidae and
Quettacyonidae. Valid Asian condylarth species are Hyopsodus orientalis and H. huashigouensis from the lower
Eocene of Mongolia and China, Midiagnus gracilis from the lower Eocene of Mongolia, Quettacyon parachai,
Q. aeruginus and Q. abbasi from the lower Eocene of Pakistan and Lophocion asiaticus from the lower Eocene
of China. All the valid species of Asian condylarths are from the Eocene, whereas the supposed Paleocene
condylarths are either anagalids or other typical Paleocene Asian mammals. This suggests that condylarths did
not reach Asia before the early Eocene, where they became relatively common during the Bumbanian. Phylogenetic analysis of condylarth relationships demonstrates that primitive sister taxa, ergo likely ancestors, of the
few Asian condylarths are present, abundant and diverse in Paleocene and Eocene strata of western North
America. The appearance in Asia of hyopsodontids (Hyopsodus, Midiagnus), phenacodontids (Lophocion) and
quettacyonids (Quettacyon, arguably arctocyonid derivatives) at approximately the beginning of the Eocene
thus was the result primarily of dispersal from North America, which clearly was the ancestral area for condylarths
during the Paleocene.
Key words: Condylarthra, Asia, Paleocene, Eocene, immigration

INTRODUCTION
Condylarths are abundant in North American Paleocene and Eocene
mammal faunas but rare in Asian Paleogene strata. Several taxa have been
described from the Paleocene and Eocene of Asia, but the taxonomic
status and validity of some of the described species are doubtful. Here,
we review all condylarth taxa that were described from Asia to establish
their distribution, taxonomic validity and affinities.
Institutional abbreviations: GSP-UM = Geological Survey of
Pakistan, Islamabad, Pakistan and University of Michigan, Museum of
Paleontology, Ann Arbor, Michigan; IVPP = Institute of Vertebrate Paleontology and Paleoanthropology, Beijing, China; PIN = Paleontological Institute of the Russian Academy of Sciences, Moscow, Russia.
SYSTEMATIC PALEONTOLOGY
Order Condylarthra Cope, 1881
Superfamily Hyopsodontoidea Trouessart, 1879, new rank
Family Hyopsodontidae Trouessart, 1879
Subfamily Hyopsodontinae Trouessart, 1879
Hyopsodus Leidy, 1870
H. orientalis Dashzeveg, 1977
Fig. 1
Hyopsodus orientalis Dashzeveg, 1977, p. 8, figs. 1-3.
Hyopsodus sp.: Zhai, 1978, p. 107, fig. 1.
Hyopsodus turpanensis Zhai, 1978 in Huang, 1995 p. 41, 46.
Hyopsodus orientalis: Kondrashov and Agadjanian, 1999, p. 670, pl.
6, figs. 1-7, pl. 7, figs. 1-6.
HolotypeGeological Institute of the Mongolian Academy of Sciences 21-35, right maxillary fragment with M1-3.
DiagnosisDiffers from most of the species of Hyopsodus by
the well-developed pericone. Similar to European H. wardi, but differs
from the latter by a well-developed parastyle on the upper molars and
in more elongate lower molars. Differs from H. huashigouensis, the
other Asian Hyopsodus, in being smaller (30 %) and in having less
developed crests on lower molars.

Referred specimensBase of the Bumban Member of the Naran


Bulak Formation at the Tsagan-Khushu locality. The Tsagan-Khushu
locality is situated in the Nemegt depression in southwestern Mongolia,
and its geology was described by Badamgarav and Reshetov (1985).
Vertebrate remains come from the facies transitional from fluviatile to
lacustrine sediments at the base of the Bumban Member: PIN 3104293, right M2-3; 3104-294, right M1-2; 3104-295, left M1-2; 3104296, right ?4-M1; 3104-297, right M1-3; 3104-298, left ?4-M1; 3104299, left M2; 3104-300, left M1; 3104-301 right M2; 3104-302, left
M2; 3104-303, left ?3; 3104-304, right m1; 3104-305, left m1-2; 3104306, right p3-m1; 3104-307, right m2-3; 3104-308 right m1-2; 3104309, left m1-2; 3104-310, right m3; 3104-311, right m2.
CommentsZhai (1978) described a dentary with m1-m3 and
alveoli for p2-4 as Hyopsodus sp. from the Shinsanjiangfang Formation in the Turpan Basin, Xinjiang, China (Fig. 1). Huang (1995) created a nomen nudum for this specimen without designating a holotype
or providing a diagnosis of this species by terming it Hyopsodus
turpanensis Zhai, 1978, although Zhai (1978) never mentioned this
name. The size of the lower molars of this species is very close to that
of H. orientalis. The m1 length is 3.1 average for H. orientalis and 3.2
for Hyopsodus sp. from the Turpan Basin. The lower molar morphology of the latter also closely resembles H. orientalis in the degree of
paraconid reduction, in its slightly posteriorly displaced metaconid on
m1 and m2 and in the talonid structure of the m3. So, the Turpan Basin
specimen belongs to H. orientalis.
DistributionLower Eocene (Bumbanian) of Mongolia and
Xinjiang, China.
H. huashigouensis Tong, 1989
Hyopsodus huashigouensis Tong, 1989, p. 182, fig. 1.
Hyopsodus fangxianensis: Huang, 1995, p. 33, fig. 1.
HolotypeIVPP-7921, right dentary with m2.
DiagnosisDiffers from other Hyopsodus in having more lophodont lower molars. Differs from H. orientalis in being considerably
larger (30%).
CommentsHuang (1995) described H. fangxianensis with the
following diagnosis: The m2 length is 4.5 mm. The m1 and m2 are of the

210

FIGURE 1. Hyopsodus orientalis, IVPP 66019, left dentary fragment with p4-m3 from the Turpan basin, China, occlusal (A) and left lateral (B) views.

same length and width. Both m1 and m2 are rather massive while the m3
is considerably narrower. Paraconid and metastylid is absent on lower
molars. Hypoconulid is better developed and is central in position on
m1-2. Both H. huashigouensis and H. fangxianensis have similarly
reduced paraconids on the lower molars, the metastylid is absent, and the
entoconulid is well developed. The size differences are minimal. Obviously, these two Hyopsodus are conspecific. H. huashigouensis Tong,
1989 has priority, so H. fangxianensis Huang, 1995 is the junior subjective synonym.
Distribution: Lower Eocene (Bumbanian) of China.
Midiagnus Russell in Godinot et al., 1987
Midiagnus gracilis Kondrashov, 2004
Midiagnus gracilis Kondrashov, 2004, fig. 1.
HolotypePIN 3104-312, left dentary with p4-m3.
Referred specimensFrom the type locality: PIN 3104-439,
right dentary with m1-2, alveoli for p2-p4, coronoid and articular processes.
Diagnosis: Differs from the type species in the position of the
mental foramina, which are located more posteriorly in M. gracilis sp.
nov. (anterior under the p3, posterior under the posterior root of the p4);
also differs in the relative size of the molars: the m1 is smaller than the
m3. There are differences in the p4 structure: the new species does not
have a metaconid on the p4 and has 2 cuspids on the p4 talonid. Lower
molars of the new species differ in having a vestigial hypoconulid and in
lacking the small cuspid at the base of the protoconid, described by
Russell for the type species (Godinot et al., 1987). The m3 is not as
reduced in the new species and has a very strong entocristid that bears a
small cuspid.
Type localityBase of the Bumban Member of the Naran Bulak
Formation at the Tsagan-Khushu locality, southwestern Mongolia.
DistributionThe Bumban Member of the Naran Bulak Formation is early Eocene in age. The species was found only at the type
locality.

Hyopsodontidae gen. et sp.


Tong and Wang (1998) listed Hyopsodontidae gen. et sp. nov.
from the Wutu fauna, Shandong Province, China. In their description,
they mention that the Wutu taxon is generically different from Hyopsodus.
Fam. Hyopsodontidae
Decoredon anhuiensis (Xu, 1976)
Xu (1977) described the species Decoredon elongatus from Paleocene strata in the Chiansian basin of China as a hyopsodontid condylarth.
Together with this species, she described two anagalid species: Diacronus
wanghuensis and Diacronus anhuiensis (Xu, 1976). Van Valen (1978)
suggested, without discussion, that Decoredon elongatus is a plesiadapid
primate.
Szalay and Li (1986) re-examined the holotypes of Diacronus
anhuiensis and Decoredon elongatus and stated that these two represent
a single taxon, Decoredon anhuiensis. The latter authors referred this
taxon to primates, particularly to the family Omomyidae within a new
subfamily, Decoredontinae, that supposedly represents the earliest primate radiation in Asia. Gingerich et al. (1991) mentioned that Decoredon
elongatus is so different from any known primate, fossil or living,
that they are unlikely to represent this order and suggested that
Condylarthra was a possible allocation for this species.
Decoredon does not exhibit features typical either of archaic ungulates or of omomyid primates. The upper molars are very wide transversely and anteroposteriorly compressed, closely resembling the condition seen in many anagalids, but not in primates and condylarths, except
periptychids, which have a long, lingual protocone slope not seen in
Decoredon. The lower molars of Decoreodon have a tall, prismatic trigonid, which is also more typical of either anagalids or insectivores, but not
of ungulates and primates. The m3 talonid is extremely elongated, which
immediately excludes the possibility of any condylarthran affinities and
also makes primate affinities very unlikely. The p4 structure is trenchant; the tooth is laterally compressed and has a well-developed shearing surface. The p4 in Omomyidae is short, the protoconid is the central

211
HolotypeGSP-UM 4000, right dentary fragment with p4-m3
(Fig. 2b).
CommentsGingerich et al. (1998) described the new genus and
species Sororocyon usmanii (Fig. 2a) from the Ghazij Formation of
Pakistan. The generic differences they listed between Sororocyon and
Quettacyon were mostly premolar characters, including spacing between
the p3 and p4, though the p3 and its position in relation to the p4 is
unknown for Quettacyon. On the other hand, the p4 and lower molar
structure is strikingly similar in these two genera. The m1 appears wider
in Sororocyon, but this tooth is damaged in Quettacyon and cannot be
used for comparison. The m2 and m3 are similar to those of Quettacyon.
Indeed, when treated as Quettacyon, this species is undistinguishable
from Quettacyon parachai both in size and morphology. These are the
measurements (in mm) of the m2: length = 11.1 in Quettacyon parachai
and 11.0 in Sororocyon usmanii; trigonid width = 9.9 in Quettacyon
parachai and 10.5 in Sororocyon usmanii; and talonid width = 9.7 in
Quettacyon parachai and 10.3 in Sororocyon usmanii. Neither morphology nor size support the generic and specific differences between the
two species, so Sororocyon usmanii is a junior synonym of Quettacyon
parachai.
DistributionPakistan, Ghazij Formation, lower Eocene.
Quettacyon aeruginis (Gingerich, Arif, Hussain and Abbas, 1998)
Fig. 2d
Obashtalkia aeruginis Gingerich, Arif, Hussain and Abbas, 1998, p. 7,
fig. 8.

FIGURE 2. Occlusal views of lower cheek teeth of selected species of Quettacyon


(from Gingerich et al., 1997, 1998, 1999, to same scale). A, Sororocyon
usmanii, holotype, GSP-UM 4021; B, Quettacyon parachai, holotype,
GSP-UM 4000; C, Q. parachai, GSP-UM 4021; D, Quettacyon
(=Obashtalkia) aeruginis, holotype, GSP-UM 4006; E, Quettacyon

and dominating cusp, and the tooth is rounded, not elongated. All the
above listed features are characters typical of anagalids, so we accept the
synonymy of Diacronus anhuiensis and Decoredon elongatus and regard Decoredon anhuiensis as an anagalid.
Hyopsodontidae gen. et sp. indet.
Zhang et al. (1979) referred a mandibular fragment with a m3 from
the Paleocene of Jiangxi Province of China to Hyopsodontidae gen. et sp.
indet. However, the illustrated specimen does not possess features characteristic of the hyopsodontid m3. Thus, the oblique cristid is extremely
strong and begins at the lingual side of the metaconid. The talonid basin
is narrow, triangular in shape and widely open lingually. These characters
are not typical of hyopsodontids or other condylarths. Due to the incompleteness of the material it is difficult to establish the affinities of the
fossil described by Zhang et al. (1979), so it is best treated as Eutheria
indet. until more specimens are discovered.
Family Quettacyonidae Gingerich, Abbas and Arif, 1997
Quettacyon Gingerich, Abbas and Arif, 1997
Quettacyon Gingerich, Abbas and Arif, 1997, p. 631.
Sororocyon Gingerich, Arif, Hussain and Abbas, 1998, p. 6.
Obashtalkia Gingerich, Arif, Hussain and Abbas, 1998, p. 7.
Machocyon Gingerich, Arif, Khan, Clyde and Bloch, 1999, p. 238.
Quettacyon parachai Gingerich, Abbas and Arif, 1997
Fig. 2a-c
Quettacyon parachai Gingerich, Abbas and Arif, 1997, p. 631, fig. 2.
Sororocyon usmanii Gingerich, Arif, Hussain and Abbas, 1998, p. 6, fig.
7.

HolotypeGGS-UM 4006, isolated right p4 (Fig. 2d).


Revised diagnosisA species of Quettacyon that differs from the
type species in being slightly larger.
CommentsGingerich et al. (1998) described the new genus and
species Obashtalkia aeruginis from the Ghazij Formation of Pakistan
based on a single lower fourth premolar (Fig. 2d). The generic differences
of Obashtalkia from Quettacyon were listed as: having a substantially larger p4 with a crown that narrows anteriorly, contacts p3 and has
a larger hypoconulid. Nevertheless, the distinctiveness of these characters is questionable (compare Figures 2b and 2d), so we judge them
insufficient to separate the two genera, though they may be enough to
separate two species. Thus, we treat Obashtalkia as a junior synonym
of Quettacyon and tentatively refer Obashtalkia aeruginis to Quettacyon
aeruginis, although more material is needed to demonstrate its specific
distinctiveness.
DistributionPakistan, Ghazij Formation, lower Eocene.
Quettacyon abbasi (Gingerich, Arif, Khan, Clyde and Bloch,
1999)
Fig. 2e
Machocyon abbasi Gingerich, Arif, Khan, Clyde and Bloch, 1999, p.
238, fig. 7.
HolotypeGSP-UM 4208, right dentary fragment with p2-m3,
associated C1 (Fig. 2e).
Revised diagnosisThe largest representative of the genus.
CommentsGingerich et al. (1999) described the new genus and
species Macocyon abbasi from the Ghazij Formation of Pakistan based
on a dentary with p2-m3 (Fig. 2e). The newly described quettacyonid
appeared to be the largest known representative of the family. However, the characters provided by Gingerich et al., (1999) are insufficient to substantiate generic separation of this species from Quettacyon.
Indeed, the characters listed by Gingerich et al. (1999) include
more massive premolars and no diastema between the p3 and the p4.
The molar structure is strikingly similar in Machocyon and Quettacyon,
including the arrangement of cusps and tooth proportions. Another character used by Gingerich et al. (1999) to distinguish Quettacyon from
Machocyon is the relative proportions of the m2, which has the

212

FIGURE 3. A- B, IVPP V 4235, holotype of Palasiodon siurensis, right dentary fragment with damaged p2-m3, occlusal (A) and labial (B) views. C-D,
IVPP 4236, holotype of Yuodon protoselenoides, right dentary fragment with p2-m3, occlusal (C) and labial (D) views.

talonid narrower than the trigonid.


However, if we turn to the measurements of Q. parachai, the
difference is minute. The m2 trigonid/talonid widths (in mm) of Quettacyon
parachai are 9.9/9.7 and 11.1/11.0 in GSP-UM 4000 and GSP-UM
4021, respectively. The trigonid/talonid width in Machocyon abbasi
is 12.5/12.7, based on a single specimen. The degree of wear is different
on the holotype of Machocyon abbasi and both the holotype and referred specimen of Quettacyon parachai, but the arrangement of cusps
remains the same. We thus consider the differences listed by Gingerich et
al. (1999) insufficient to establish the generic rank of Machocyon
abbasi, so we treat Machocyon as a junior synonym of Quettacyon and
refer Machocyon abbasi to the genus Quettacyon.
DistributionPakistan, Ghazij Formation, lower Eocene.
Superfamily Phenacodontoidea McKenna, 1975
Family Phenacodontidae Cope, 1881
Subfamily Meniscotheriinae Cope, 1882

Lophocion Wang and Tong, 1997


Lophocion asiaticus Wang and Tong, 1997
Lophocion asiaticus Wang and Tong, 1997, p. 288, fig. 1.
Lophocion asiaticus Tong and Wang, 1998, p. 191, fig. 3d.
HolotypeIVPP-V10707, maxillary fragment with left M1-3.
DiagnosisDiffers from all Meniscotherium species in having
lophodont rather than buno-selenodont teeth. Differs from Ectocion species in having more lophodont upper molars with more developed
protoloph and metaloph.
Distribution: Lower Eocene (Bumbanian) of China.
Phenacodontidae
Paraphenacodus solivagus Gabunia, 1971
Gabuniya (1971) described P. solivagus from the middle Eocene
of the Zaysan basin in eastern Kazakstan as a phenacodontid, but it was
later shown that this taxon belongs to Dichobunidae (Artiodactyla)

213
(Gabunia, 1984; Russell and Zhai, 1987).
Phenacodontidae
Eodesmatodon spanios Zheng and Chi, 1978
Zheng and Cui (1978) named the phenacodontid
Eodesmatodon spanios for a dentary fragment with p4 and m2 from the
Eocene of Guangxi, China. The specimen is obviously a representative
of Carnivora, and Russell and Zhai (1987, p. 239) assigned it to
Procyonoidea.
Phenacodontidae
Zhou et al. (1977) described and illustrated IVPP V 4237, a
right dentary fragment with p3 and roots of the cheek teeth from the
Paleocene Shanghu Formation of southern China as Phenacodontidae
gen. et sp. nov., but did not name it. Though the crowns of the cheek
teeth were broken, the authors mentioned that the enamel extended
down all the way into the alveoli, a feature that is unknown in
phenacodontid condylarths (Thewissen, 1990; Williamson and Lucas,
1992). Based on that feature and on comparison with other anagalids,
Lucas and Williamson (1995) referred this specimen to Anagalida, although they mentioned that the specimen, which is apparently lost, is
too incomplete for certain identification.
Family Mioclaenidae
Yuodon protoselenoides Chow, Chang, Wang and Ting, 1973
Fig. 3C-D
Chow et al. (1973) described Yuodon protoselenoides from the
early Paleocene of Nonhsiung Basin, China, as a mioclaenid condylarth
and placed it close to Protoselene opisthacus. However, the structure
of the lower teeth of Yuodon differs considerably from the condition
seen in Mioclaenidae. The most significant differences are in the structure of the lower premolars, which are bulbous and rounded in
mioclaenids with evident crushing adaptations, whereas in Yuodon the
lower premolars are laterally compressed and adapted for shearing.
Molar structure is also very different. Lucas and Williamson (1995) reevaluated the holotype and concluded that Yuodon protoselenoides is a
member of Anagalida, not a condylarth.
Palasiodon siurensis (Zhou, Chang, Wang and Ting, 1977)
Fig. 3A-B
Palasiodon siurensis, based on a badly damaged dentary with
p4-m3 from the Shanghu Formation, was described by Chow et al. (1973)
as a mioclaenid condylarth and assigned to the genus Promioclaenus.
Later the same authors (Zhou [Chow] et al., 1977) erected a new genus
Palasiodon to include Promioclaenus siurensis. The dentition of the
holotype appears to be broader than in Yuodon protoselenoides, which
is due to the type of preservation, as the teeth were deformed plastically by crushing. Lucas and Williamson (1995) noted the similarities
in overall morphology between Palasiodon siurensis and Yuodon
protoselenoides, such as lack of cingulids, small paraconids, strong
oblique crests and the extension of the enamel all the way down into
the alveoli. All these characters suggest that Palasiodon siurensis is
not a hyopsodontid condylarth, but an anagalid, probably congeneric
with Yuodon.
Family Periptychidae
Pseudanisonchus antelios Zhang, Zheng and Ting, 1979
Fig. 4
Zhang et al. (1979) described Pseudanisonchus antelios from the
Paleocene of Jiangxi Province, China, based on an isolated ?M2 (IVPP
5041), as a periptychid condylarth and placed it in the subfamily
Anisonchinae. The structure of the tooth does not show any periptychid
or condylarth features. Such features as a large protocone with extremely
well-developed pre- and postmetacristae, no para- and metaconules and

FIGURE 4. Cast of holotype of Pseudanisonchus antelios, occlusal view of an


isolated ?M2 (IVPP 5041).

almost undifferentiated hypocone exclude this taxon from the archaic


ungulates and suggest it has affinities with pseudictopid anagalids.
?Ectoconus sp.
Zhou et al. (1977) reported ?Ectoconus from the Paleocene
Shanghu Formation of China. The identification was based on an almost complete sacrum, which was similar to the sacrum of Ectoconus
majusculus (=E. ditrigonus), described by Matthew (1937). Lucas and
Williamson (1995) showed that this sacrum does not differ from the
sacrum of the pantodont Bemalambda, which is rather common in the
same strata.
DISCUSSION
Of six condylarth families and 18 species reported from Asia,
only three families and seven species do belong to the order
Condylarthra. These families are: Hyopsodontidae, Phenacodontidae
and Quettacyonidae. Valid Asian condylarth species are Hyopsodus
orientalis and H. huashigouensis from the lower Eocene of Mongolia
and China, Quettacyon parachai, Q. aeruginus and Q. abbasi from the
lower Eocene of Pakistan, Midiagnus gracilis from the lower Eocene
of Mongolia and Lophocion asiaticus from the lower Eocene of China.
It is interesting that all the species of Asian condylarths that are valid
were reported from the Eocene (early and middle), whereas the supposed Paleocene condylarths are either anagalids or other typical Paleocene Asian mammals. This suggests that condylarths did not reach
Asia before the early Eocene, where they became relatively common
during the Bumbanian.
Beard (1998, 2002; Beard and Dawson, 1999) argued that Asia
was an important ancestral area for many Paleogene placental mammal
clades that originated there, including Perissodactyla, Artiodactyla,
Cetacea, Dinocerata, Tillodontia, Pantodonta, Rodentia, Lagomorpha
and Primates. They called this concept east of Eden, and concluded
that at about the beginning of the Eocene these mammal groups spread
from Asia to North America and/or Europe via boreal land bridges.
Condylarths run counter to this purported vector of immigration.
Thus, there are no bona fide Paleocene condylarths known from Asia.
(Years ago, a Chinese colleague showed us artocyonid teeth supposedly
from the lower Paleocene of southern China, but these teeth have never
been published, and we are skeptical of their provenance.) Phylogenetic
analysis of condylarth relationships (Archibald, 1998) demonstrates that
primitive sister taxa, ergo likely ancestors, of the few Asian condylarths

214
are present, abundant and diverse in Paleocene strata of western North
America. The appearance in Asia of hyopsodontids (Hyopsodus,
Midiagnus), phenacodontids (Lophocion) and quettacyonids (Quettacyon,
arguably arctocyonid derivatives) at approximately the beginning of the
Eocene thus was the result of dispersal from North America, which
clearly was the ancestral area for condylarths during the Paleocene. Therefore, the condylarths were west of Eden during the Paleogene.

ACKNOWLEDGMENTS
We would like to thank Alexey Lopatin for comments on the
manuscript. Funding from the NMMNH Foundation made this study
possible. The study was partially supported by the Russian Foundation
for Basic Research, project 02-04-48458 and the Grant Council of the
President of the Russain Federation project Nsh-1840.2003.04.

REFERENCES
Archibald, J. D., 1998, Archaic ungulates (Condylarthra); in Janis, C. M.,
Scott, K. M. and Jacobs, L. L., eds., Evolution of Tertiary mammals of
North America. Volume 1: Terrestrial carnivores, ungulates, and ungulate like mammals: Cambridge, Cambridge University Press, p. 292-331.
Badamgarav, D. and Reshetov, V. Yu., 1985, Paleontologiya i stratigrafiya paleogena
Zaaltayskoy Gobi [Paleontology and stratigraphy of the Paleogene of
Zaaltayskaya Gobi]: Trudi Sovmestnoy Sovetsko-Mongolskoy
Paleontologicheskoy Expeditsii [Proceedings of the Joint Soviet-Mongolian
Paleontological Expedition], v. 25, p. 1-104.
Beard, K. C., 1998, East of Eden: Asia as an important center of taxonomic origination in mammalian evolution: Bulletin of the Carnegie Museum of Natural History, v. 34, p. 5-39.
Beard, K. C., 2002, East of Eden at the Paleocene/Eocene boundary: Science, v.
295, p. 2028-2029.
Beard, K. C. and Dawson, M. R., 1999, Intercontinental dispersal of Holarctic land
mammals near the Paleocene/Eocene boundary: Paleogeographic, paleoclimatic
and biostratigraphic implications: Bulletin Societe Gologique de France, v. 170,
p. 697-706.
Chow, M., Chang, Y., Wang, B. and Ting, S., 1973, New mammalian genera and
species from the Paleocene of Nanhsiung, N. Kwangtung: Vertebrata Palasiatica,
v. 11, p. 31-35.
Chow, M., Chang, Y., Wang, B. and Ting, S., 1977, Mammalian fauna from the
Paleocene of Nanxiong basin Guangdong: Vertebrata Palasiatica, v. 20, p. 1100.
Dashzeveg, D., 1977, O pervoi nakhodke Hyopsodus Leidy, 1870 (Mammalia,
Condylarthra) v Mongolskoy Narodnoy Respublike [On the first occurrence of
Hyopsodus Leidy, 1870 (Mammalia, Condylarthra) in the Mongolian Peoples
Republic]: Trudi Sovmestnoy Sovetsko-Mongolskoy Paleontologicheskoy
Expeditsii [Proceedings of the Joint Soviet-Mongolian Paleontological Expedition], v. 4, p. 7-13.
Gabunia, L. K., 1971, O novom predstavitele condylartr (Condylarthra) iz eotsena
Zaisanskoi kotlovini [About the new representative of condylarths (Condylarthra)
from the Eocene of Zaissan depression]: Soobsheniya Akademii Nauk Gruzinskoi
SSR [Reports of the Georgian Academy of Sciences], v. 61, p. 233-235.
Gabunia, L. K., 1984, Kratki obzor faun Paleogenovih mlekopitayushih Zaisanskoi
vpadini [Short review of the Paleogene mammalian fauna of the Zaysan Depression]; in Gabuniya, L.K., ed., Flora i fauna Zaysansskoi vpadini [Flora and
fauna of the Zaissan depression]: Tbilisi, Georgian Academy of Sciences, p.
142-141.
Gingerich, P. D., Abbas, S.G. and Arif, M., 1997, Early Eocene Quettacyon parachai
(Condylarthra) from the Ghazij Formation of Baluchistan (Pakistan); oldest
Cenozoic land mammal from south Asia: Journal of Vertebrate Paleontology, v.
17, p. 629-637.
Gingerich, P. D., Arif, M., Hussain, I. and Abbas, S.G., 1998, First early Eocene
land mammals from the upper Ghazij Formation of the Sor Range, Baluchistan;
in Ghazhnavi, M.I., Raza, S.M. and Hasan, M. T., eds., Siwaliks of south Asia,
Proceedings of the Third Geosas Workshop held at Islamabad, Pakistan, p. 117.
Gingerich, P.D., Arif, M., Khan, I.H., Clyde, W.C. and Bloch, J., 1999, Machocyon
abbasi, a new early Eocene quettacyonid (Mammalia, Condylarthra) from the
middle Ghazij Formation of Mach and Dakhari coal fields, Baluchistan
(Pakistan): Contributions Museum Paleontology, University of Michigan, v. 30, p. 233-250.
Gingerich, P.D., Dashzeveg, D. and Russell, D.E., 1991, Dentition and
systematic relationships of Altanius orlovi (Mammalia, Primates) from
the early Eocene of Mongolia: Geobios, v. 24, p. 637-646.

Godinot, M., Crochet, J.-Y., Hartenberger, J.-L., Lange-Badre, B., Russell,


D.E., Sige, B., 1987, Nouvelles donnees sur les mammiferes de Palette
(Eocene inferieur, Provence): Munchner Geowiss. Abh., v. 10, p. 273288.
Huang, X., 1995, A new Hyopsodus from the Early Eocene of Fangxian, Hubei:
Vertebrata PalAsistica, v. 33, p. 39-46.
Kondrashov, P. E., 2004, A new hyopsodontid (Mammalia, Condylarthra) from the
early Eocene of Mongolia: New Mexico Museum of Natural History and Science, Bulletin 26.
Kondrashov, P.E. and Agadjanian, A.K., 1999, New material on the genus Hyopsodus
(Mammalia, Condylarthra) from the Eocene of Mongolia: Paleontological Journal, v. 33, p. 667-676.
Lucas, S.G. and Williamson, T. E., 1995, Systematic position and biochronological
significance of Yuodon and Palasiodon, supposed Paleocene condylarths from
China: Neues Jahrbuch fr Geologie und Palontologie Abhandlungen, v. 196,
p. 93-107.
Matthew, W.D., 1937, Paleocene faunas of the San Juan Basin, New Mexico: Transactions of the American Philosophical Society, new series, v. 30, 510 p.
Russell, D.E. and Zhai, R.J., 1987, The Paleogene of Asia: Mammals and
stratighaphy: Memoires Museum National dHistoire Naturelle, v. 52, 488 p.
Szalay, F. S. and Li, C., 1986, Middle Paleocene euprimate from southern China
and the distribution of primates in the Paleogene: Journal of Human Evolution,
v. 15, p. 387-397.
Thewissen, J. G. M., 1990, Evolution of Paleocene and Eocene Phenacodontidae
(Mammalia, Condylarthra): Papers on Paleontology University of Michigan, v.
29, p. 1-107.
Tong, Y., 1989, Some Eocene mammals from the Uqbulak area of the Junggar basin, Xinjiang: Vertebrata Palasiatica, v. 27, p. 182-196.
Tong, Y. and Wang, J., 1998, A preliminary report on the early Eocene mammals of
the Wutu fauna, Shandong Province, China: Bulletin of the Carnegie Museum
of Natural History 34, p. 186-193.
Van Valen, L., 1978, The beginning of the age of mammals: Evolutionary Theory, v.
4, p. 45-80.
Wang, J. and Tong, Y., 1997, A new phenacodontid condylarth (Mammalia) from
the Early Eocene of the Wutu basin, Shandong: Vertebrata Palasiatica, v. 35, p.
283-289.
Williamson, T.E. and Lucas, S.G., 1992, Meniscotherium (Mammalia,
Condylarthra) from the Paleocene-Eocene of western North America: New
Mexico Museum of Natural History and Science, Bulletin 1, 75 p.
Xu, Q., 1976, New materials of Anagalidae from the Paleocene of Anhui: Verterbrata
Palasiatica, v. 14, p. 242-251.
Xu, Q., 1977, Two new genera of old Ungulata from the Paleocene of Qianshan
Basin, Anhui: Vertebrata Palasiatica, v. 15, p. 119-125.
Zhai, R., 1978, More evidence favouring on early Eocene connection between Asia
and Neoarctic: Memoirs Institute of Vertebrate Paleontology and
Paleoanthropology, v. 13, p. 107-116.
Zhang, Y., Zheng, J. and Ting, S., 1979, Description of some condylarths
from the Paleocene of Jiangxi; in Mesozoic and Cenozoic red beds of
south China: Academia Sinica Institute of Vertebrate Paleontology and
Paleoanthropology, Beijing: Science Press, p. 382-386.
Zheng, J. and Cui, H., 1978, Some of the latest Eocene Condylarthra mammals from Guangsi, south China: Vertebrata Palasiatica, v. 16, p. 97102.
Zhou [Chow], M., Zhang, Y., Wang, B. and Ting, S., 1977, Mammalian fauna from
the Paleocene of Nanxiong basin, Guandong: Paleontologia Sinica Series C, v.
20, 100 p.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

215

EARLY EOCENE (BUMBANIAN) PERISSODACTYLS FROM MONGOLIA


AND THEIR BIOCHRONOLOGICAL SIGNIFICANCE
SPENCER G. LUCAS1 AND PETER E. KONDRASHOV2
2

1
New Mexico Museum of Natural History, 1801 Mountain Road, Albuquerque, NM 87104;
Department of Biology, Northwest Missouri State University, 800 University Drive, Maryville, MO 64468
and Paleontological Institute of RAS, Moscow, RUSSIA 117868

AbstractPerissodactyls are relatively common in Wasatchian deposits of North America, but their occurrences in the lower Eocene (Bumbanian land-mammal age [lma]) of Asia are very rare, and only a few species
are known from that time interval. We describe new specimens of two species of basal perissodactyls, Homogalax
namadicus and Hyracotherium gabuniai, in the early Eocene collection from the Bumban Member of the NaranBulak Formation at the Tsagan-Khushu locality in Mongolia. Cladotaxonomy is the recognition of cladotaxa,
which are low-level taxa (genera and species) that correspond to clades in a cladistic analysis. We reject the
current cladotaxonomy of basal perissodactyls as typological, oversplit, devoid of biological significance and
premature. Orientolophus Ting is a junior subjective synonym of Hyracotherium and O. hendongensis a junior
subjective synonym of H. gabuniai. Hyracotherium and Homogalax characterize the Bumbanian lma but are
too poorly known in Asia to provide a basis for subdivision of that lma.
Key words: Perissodactyla, Hyracotherium, Homogalax, Mongolia, Eocene, Bumbanian, cladotaxon

INTRODUCTION
The systematics of basal perissodactyls, which first appear in
Asia, North America and Europe at about the Paleocene-Eocene boundary, has been the focus of extensive study. Here, we add to the meager
database of Asian basal perissodactyls the description and illustration
of new specimens from Tsagan-Khushu in Mongolia. We assign these
specimens to the perissodactyl taxa already named from Tsagan-Khushu,
Hyracotherium gabuniai and Homogalax namadicus (Dashzeveg,
1979a, b). We also discuss and reject the current cladotaxonomy of
basal perissodactyls, and demonstrate that Orientolophus hengdongensis
is a junior subjective synonym of Hyracotherium gabuniai. Finally, we
evaluate the significance of basal perissodactyls to Asian land-mammal biochronology.
Institutional abbreviations: IVPP Institute of Vertebrate Paleontology and Paleoanthropology, Beijing, Peoples Republic of China;
PIN Paleontological Institute of the Russian Academy of Sciences,
Moscow, Russia; PST Paleontology and Stratigraphy Department of
the Geological Institute of the Mongolian Academy of Sciences, Ulan
Bator, Mongolia.
SYSTEMATIC PALEONTOLOGY
Order Perissodactyla Owen, 1848
Family Isectolophidae Peterson, 1919
Genus Homogalax Hay, 1899
Homogalax namadicus Dashzeveg, 1979
Figs. 1E-G, 2A-B, E-F, Tables 1-2
Homogalax namadicus Dashzeveg, 1979a, p. 108, figs. 1, 2.
HolotypePST 20-10, right M1 (Fig. 1E; Dashzeveg, 1979a,
fig. 1).
Referred specimens From the Bumban Member (basal
Eocene) of the Naran-Bulak Formation, Tsagan-Khushu locality, upper
white horizon: PIN 3104-323, right maxillary fragment with P4-M3
(Fig.2A-B); 3104-480, isolated left m1 (Fig. 2E); 3104-481, isolated
right m3 (Fig. 2F). Also see the PST specimen (Fig. 1F-G) referred by
Dashzeveg (1979a).
Revised diagnosisMedium-sized Homogalax, differs from
most of the representatives of the genus, including H. wutuensis from

China, in lacking the distinct conules on the upper molars and in being
smaller.
DistributionLower Eocene (Bumbanian) of the TsaganKhushu locality, southwestern Mongolia, Nemegt depression.
DescriptionPIN 3104-323 is a maxillary fragment with the
right P4-M3 (Fig. 2A-B). The P4 is a triangular tooth with a prominent
parastyle connected to a straight ectoloph that joins the relatively large,
distinct paracone and metacone and terminates posteriorly with a small,
crest-like metastyle. The two ectoloph cusps, though, are large enough
to give the labial aspect of the ectoloph a ribbed (corrugated) appearance. The protocone is positioned lingually and is larger than the paracone and metacone. A strong preprotocrista well connected to the protocone runs antero-labially to connect to the ectoloph low on its lingual
face and just antero-lingual to the paracone. The postprotocrista is much
lower than the preprotocrista and is not as strongly connected to the
protocone as is the preprotocrista. The postprotocrista is confluent with
the metacone. Low anterior and posterior cingula nearly meet lingual
to the protocone. And, the parastyle postero-labially becomes a low
cingulum interrupted by the paracone rib on the ectoloph. A low cingulum is labial to the metacone rib on the ectoloph.
The M1 is a trapezoidal tooth, mostly because of the very large,
antero-labially projecting parastyle. This parastyle is a blunt flange
connected to the paracone by a low crest, so that there is a distinct
notch (cleft) between the parastyle and paracone. The paracone is distinctly larger than and taller than the metacone, and a low crest connects the cusps, forming an ectoloph with prominent paracone and
metacone ribs. The protocone is large, antero-lingual and has a prominent well-worn paraloph that runs antero-labially to the cleft between
the parastyle and paracone. The hypocone is somewhat smaller than
and postero-labial to the protocone. A deep groove separates these lingual cusps, but a low crista on the lingual aspect of the hypocone connects it to the lingual cingulum. A low, thin metaloph connects the
hypocone to the metacone. The metacone and paracone form a nearly
straight, ribbed ectoloph. The metastyle is a distinct blade-like ridge
that projects almost directly posterior to the metacone. A cingulum is
present along the posterior edge of the tooth, running from nearly the
protocone and connected to the metastyle. Both cross lophs are well
worn, making M1 the most worn of the four teeth.
The M2 is similar in most features to the M1. Key differences
are a relatively larger parastyle and a more trapezoidal crown shape.

216

FIGURE 1. Original illustrations of Hyracotherium gabuniai and Homogalax namadicus, from Dashzeveg (1979a, b). A-B, Holotype of Hyracotherium gabuniai, PST
20-55, right M2 in occlusal (A) and labial (B) views. C, Left M2, PST 20-56, of H. gabuniai, occlusal view. D, Left m1, PST 20-57, of H. gabuniai, occlusal view. E,
Holotype of Homogalax namadicus, PST 20-10, right M1 in occlusal view. F-G, Right dentary fragment with m1-3, PST 20-9, of H. namadicus, in occlusal (F) and
lingual (G) views.

The M3 is larger than the M2 but differs from the preceding


molars in the more reduced, laterally compressed, blade-like metacone,
which is close to the hypocone. The M3 parastyle is very large and
projects labially, so that the ectoflexus is more concave labially and the
metacone much more lingually positioned. The M3 also has a more
distinct metastyle and a more curved preprotocrista (convex anteriorly)
than do M1 or M2. Like M2, the M3 has an essentially complete lingual cingulum. The cingulum is circular and does not terminate at the
base of the hypocone. There is a small metastyle on the posterior slope
of the metacone on the cingulum. The ectoloph is higher than on the
other molars.
The m1 (Fig. 2E) is rectangular in shape. The talonid is slightly
wider than the trigonid. The paraconid is greately reduced and lies at the

base of the anterior wall of the metaconid. The protolophid is rather


strong and curved. It runs labially to the base of the protoconid and
than turns back to the protoconid apex. There is another wide and low
cristid on the anterolingual wall of the protoconid. There is also a blunt
cristid on the anterolabial surface of the metaconid. The two cuspids
are connected by a strong metalophid. There is a distinct metastylid
posterior to the metaconid with a short cristid running down the posterior wall of the trigonid, that begins from the metastylid. The cristid
obliqua is strong; it begins from the base of the posterior trigonid wall,
right in its middle. The hypoconid and entoconid are almost equal in
size, connected by a short and low hypolophid with a saddle in the
middle of it. There is a small, cingular, but distinct hypoconulid with
strong wings, formed by the posterocingulid. The cingulid is well devel-

217

FIGURE 2. New specimens of basal perissodactyls from Tsagan-Khushu, Mongolia. A-B, Homogalax namadicus, PIN 3104-323, right maxillary fragment with P4-M3
in labial (A) and occlusal (B) views. C, Hyracotherium gabuniai, PIN 3104-482, right M1, occlusal view. D, Hyracotherium gabuniai, PIN 3104-483, right P4, occlusal
view. E, Homogalax namadicus, PIN 3104-480, left m1, occlusal view. F, Homogalax namadicus, PIN 3104-481, right m3, occlusal view.

oped on the anterior, posterior and labial sides of the tooth.


The m3 (Fig. 2F) is elongated with a high and strong metalophid.
There is a vestigial metastylid on the posterior slope of the metaconid.
There is a weak cristid running from the metastylid posteriorly to the
base of the entoconid. The cristid obliqua is very weak and begins from
the middle of the posterior trigonid wall. The hypoconid is slightly larger
than the entoconid. The two cuspids are connected by a rather strong
hypolophid with a saddle in the middle. The hypoconulid is large, posteriorly projecting, and joined with the posterocingulid. There is a low,
wide cristid that runs from the base of the posterior slope of the
hypolophid, in its middle, posteriorly to the hypoconulid. The cingulid

is well-developed on the labial and posterior sides of the tooth.


DiscussionPIN 3104-323 is substantially larger than the holotype of Hyracotherium gabuniai and differs from it in several features, including the better developed ectolophs and cross lophs of PIN
3104-323. Size and morphology of the M1 match well the original holotype M1 of Homogalax namadicus, so we refer PIN 3104-323 to that
taxon. Whether or not H. namadicus is a species distinct from North
American Homogalax species remains to be determined.
Family Hyracotheriidae Cope, 1881
Subfamily Hyracotheriinae Cope, 1881

218
TABLE 1. Measurements (in mm) of the upper teeth of Homogalax
namadicus (PIN 3104-323).

TABLE 3. Measurements (in mm) of the upper teeth of


Hyracotherium gabuniai.

Tooth
P4
M1
M2
M3

Tooth
P4 (PIN 3104-483)
M1 (PIN 3104-482)

Length
6.69
8.25
9.48
10.11

Width
8.61
9.77
11.51
12.08

TABLE 2. Measurements (in mm) of the lower teeth of Homogalax


namadicus.
Tooth

Length

m1 (PIN 3104-480)
m3 (PIN 3104-481)

Trigonid
width
8.86
5.85
12.3 (est.) 6.48

Talonid
width
6.16
5.78

Genus Hyracotherium Owen, 1840


Hyracotherium gabuniai Dashzeveg, 1979
Figs. 1A-D, 2C-D, Table 3
Hyracotherium gabuniai Dashzeveg, 1979b, p. 109, fig. 1.
Orientolophus hendongensis Ting, 1993, p. 202, fig. 1.
HolotypePST 20-55, right M2 (Fig. 1A-B; Dashzeveg, 1979b,
fig. 1a-b).
Referred specimensFrom the Bumban Member (basal
Eocene) of the Naran-Bulak Formation, Tsagan-Khushu locality, upper
white horizon. PIN 3104-482, right M1 (Fig. 2C), 3104-483, right P4
(Fig. 2D). Also see the PST specimens (Fig. 1C-D) referred by
Dashzeveg (1979b).
Revised diagnosisSmall Hyracotherium that differs from the
European species in the square shape of the upper molars and more
lophodont teeth. Differs from H. angustidens and H. vasacciense from
the Wasatchian of North America in being smaller.
Distribution Early Eocene (Bumbanian) of the Tsagan-Khushu
locality, southwestern Mongolia, Nemegt depression and Hunan Province, China.
DescriptionThe P4 (Fig. 2D) is typical of hyracotheres. It is
triangular in shape. The parastyle is very large, anterolabially projecting. It is heavily worn as is the anterior slope of the paracone. The
paracone and the metacone are equal in size, close to each other and
connected by a low and sharp ectoloph. The protocone is considerably
shifted posteriorly and lies lingual to the metacone. The protoloph is
strong. It runs from the parastyle posteriorly to the protocone through
the well-developed paraconule and turns posteriorly after joining the
latter cusp. The metaloph is short, but wide with a larrge metaconule in
the middle. There is a crest on the posterior slope of the metacone that
joins the posterocingulum. The cingulum is strong and circular. The
trigon is deep and basined.
The M1 (Fig. 2C) is square in shape. The parastyle is large and
anteriorly projecting. It is connected to the paracone by a short, curved
crest. The paracone, metacone, protocone and hypocone are almost equal
in size. The paracone is slightly taller than the protocone and connected
to the latter by a low protoloph, that begins from the paracone apex and
is slightly curved anteriorly. The paracone and metacone are close to
each other, so the ectoloph is short and low. The metacone is connected
to the hypocone by a low metaloph, which is curved anteriorly. The
paraconule is vestigial and can be traced only on the worn teeth. The
cingulum is relatively weak, but almost circular and terminates only at

Length
7.10
6.78

Width
5.85
6.61

the base of the hypocone. There is no metastyle, but there is a crest on


the posterior wall of the metacone that runs to the base of the crown
and joins the posterocingulum.
DiscussionTing (1993) described a new isectolophid? genus and species, Orientolophus hendongensis, from the early Eocene
(Bumbanian) of Hendong County, Hunan Province, China. Lucas (1998)
discussed the status of Orientolophus and concluded that the characters listed by Ting (1993) are insufficient to distinguish it from
Hyracotherium and considered Orientolophus to be a junior subjective
synonym of Hyracotherium. In the description, Ting (1993) compared
the new genus and species with several Asian forms, but the tooth
structure of Orientolophus hendongensis is absolutely identical to
that of Hyracotherium gabuniai, and the measurements are very close
as well.
Ting (1993, p. 205) distinguished the Chinese species from H.
gabuniai by the lack of a lingual cingulum, labial cingulum interrupted
at the paracone and more labially positioned lower molar metalophids
(metaconids) of the Chinese species. However, the upper molars of
Orientolophus hendongensis do have weak lingual cingula (Ting, 1993,
figs. 1A-B) as in H. gabuniai (Daszheveg, 1979a, fig. 1a-c), and a similar
labial cingulum as well. Differences in the lower molar metalophids
claimed by Ting cannot be replicated by us.
Ting (1993) also suggested that Hyracotherium gabuniai probably belongs to the newly described genus and thus is an isectolophid.
The newly discovered P4 of Hyracotherium gabuniai supports its referral to Hyracotherium as it has a very different structure from that of
isectolophids. The tooth is triangular in outline, the paraconule is very
large and the protocone is shifted posteriorly (see McFadden, 1998;
Lucas et al., 2003). So, we conclude that Orientolophus is a junior
subjective synonym of Hyracotherium as proposed by Lucas (1998),
and Orientolophus hendongensis is a junior subjective synonym of
Hyracotherium gabuniai.
Distinction of H. gabuniai from American species of
Hyracotherium remains to be determined. Indeed, we expect a sound,
variation-based restudy of Hyracotherium will indicate that H. gabuniai
is a synonym of another, already named Hyracotherium species.
CLADOTAXONOMY OF BASAL PERISSODACTYLS
Hyracotherium (long ago regarded as having priority over the
more colorful name Eohippus) was long regarded as the only generic
name for the oldest and most primitive perissodactyl. Regarded by some
as the earliest equid, it was more widely considered a taxon that encompassed the ancestry of both equids and tapiroids (see discussion by
MacFadden, 1976). Hyracotherium as thus construed was well documented from the lower Eocene of North America (e.g., Granger, 1908;
Kitts, 1956; Gingerich, 1981, 1991; MacFadden, 1998), western Europe (e.g., Cooper, 1932; Simpson, 1952; Savage et al., 1965; Hooker,
1980; Godinot, 1981) and Asia (Dashzeveg, 1979b).
Sample sizes of Hyracotherium from Europe and Asia are small,
but in North America Hyracotherium is one of the most common
Wasatchian fossil mammals. Large samples from single localities and
horizons, as well as at least one mass-death assemblage, have allowed
some assessment of the wide range of variation in Hyracotherium cheektooth morphology in terms of intraspecific variation. Such assessments,

beginning with Granger (1908), continuing through Kitts (1956) and


Gingerich (1981, 1991), have identified several polymorphic species
of Hyracotherium. Nevertheless, the amount of polymorphism has never
been fully documented and analyzed, so that the species-level taxonomy
of Hyracotherium has remained somewhat open to discussion.
However, the work of Hooker (1984, 1989, 1994) has taken a
totally different approach to the alpha taxonomy of basal perissodactyls
that has been culminated by the recent work of Froehlich (1999, 2002).
This approach has split Hyracotherium into several genera that correspond to branches on a cladogram. We refer to this approach as
cladotaxonomy, and define a cladotaxon as a low-level taxon (genus or
species) that corresponds to a clade in a cladistic analysis.
Froehlichs (2002) most recent cladistic analysis of basal perissodactyls identifies a series of sister taxa within what was traditionally
termed Hyracotherium. He assigns these sister taxa to the genera
Hyracotherium (restricted to its type species H. leporinum), Hallensia,
Sifrihippus, Minippus, Arenahippus, Xenicohippus, Eohippus,
Pliolophus and Protorohippus. This is supposedly justified by recognizing that what was termed Hyracotherium (traditional usage) is revealed to be paraphyletic by the cladistic analysis, so every strictly
monophyletic taxon the analysis identifies merits a separate genus name.
We reject this cladotaxonomic approach to basal perissodactyls
for four reasons:
1. No attempt at gauging the amount of and the significance of
variation in Hyracotherium is incorporated into the cladistic analysis.
Instead, the variation, such as it is determined, is assumed to be of
phylogenetic significance only. Thus, the possibility of populational
variation in characters deemed to be of phylogenetic significance has not
been addressed.
2. The cladotaxonomy names nearly every branch on the cladogram, and thus results in taxonomic hypersplitting. Thus, for example,
examination of Winans (1989) five groups of North American Equus
indicates that the cladotaxonomic approach would split Equus into at
least five genera. Indeed, any genus currently considered speciose will be
split into multiple genera by the cladotaxonomic approach.
3. Cladotaxa only convey the topology of a cladogram and thus
are taxa devoid of other biological significance. The cladogram is based on
a character atomization that takes discrete dental characteristics of biological import, such as cheek-tooth lophodonty, and divides them into
many smaller characters, all deemed to be of equal phylogenetic significance. Thus, any biological significance of the characters is removed from
the analysis at the outset.
4. Finally, there is what we call the cladogram du jour factor.
Cladistic analysis has proven highly useful in constructing phylogenetic hypotheses that can be subjected to rigorous evaluation. But, when
such hypotheses are instantly turned into new taxonomic names, the
taxonomic nomenclature becomes burdened with numerous names based
on little-tested hypotheses. In effect, the cladogram of the moment,
even if it is the only published cladogram (the cladogram du jour),
becomes the basis for new taxonomy. Thus, for example, McKennas
(1975) phylogenetic analysis of mammals was valuable, but unfortunately he used it to create several new taxonomic names that further
phylogenetic analysis demonstrated were ill conceived. Froehlichs
(2002) basal perissodactyl cladogram was used to introduce new generic names and reinstate old generic names, and has yet to be seriously

219
evaluated by another investigator. New alpha taxonomy based on such a
cladogram is premature.
The cladotaxonomy of basal perissoadctyls can thus be seen as
basically typological, oversplit, of little biological significance and premature. With regard to alpha taxonomy, Lucas (1998, p. 452) stated,
taxonomic identity should be demonstrated by morphological similarity analyzed within the context of population variation. This approach,
when applied to basal perissodactyls, long identified a single genus
Hyracotherium composed of several species. We conclude it is preferable to the typology inherent to cladotaxonomy, which identifies several
genera of basal perissodactyls based only on their perceived cladistic
relationships.
MAMMALIAN BIOCHRONOLOGY
The oldest Eocene land-mammal age (lma) in Asia is the
Bumbanian lma of Russell and Zhai (1987). (We follow Lucas (1998)
and consider the term Bumbanian preferable to the term Lingchan introduced by Li and Ting (1983) and used by Tong et al. (1995), simply
because the mammalian assemblage of the Bumban Member is much
more diverse than the mammalian assemblage of the Lingcha Formation). In Asia, only Bumbanian assemblages include Hyracotherium
and Homogalax.
Nevertheless, both taxa are rare. Besides their co-occurrence at
Tsagan-Khushu in Mongolia, Hyracotherium is present in the Lingcha
Formation of Henan, China, and Hyracotherium and Homogalax cooccur in the Wutu Formation of Shandong (Chow and Li, 1963; Ting,
1993; Tong and Wang, 1995). Other basal perissodactyl fossils of
Bumbanian age are assigned to Heptodon and are from the Niushan
Formation in Shandong, the Shisanjiangfang Formation in Xinjiang
and possibly the lower member of the Xinyu Formation in Jiangxi (Chow,
1959; Chow and Li, 1965; Zhai, 1978). We are not able to verify the
supposed Heptodon record (Xu et al., 1979) from the Yuhuangding
Formation in Henan. In total, all basal perissodactyl taxa from
Bumbanian strata in Asia are now known from fewer than 20 specimens.
This, however, has not deterred Tong et al. (1995) from suggesting
that the Lingchan (= Bumbanian) LMA can be divided into three phases:
(1) an older Cocomys-Orientolophus phase (Lingcha assemblage); (2) a
middle Changlelestes-Arcitoparamys-Homogalax phase (Wutu assemblage); and (3) a younger Rhombomylus-Heptodon phase (Dajian assemblage). Ting (1998) also suggested the same subdivisions, terming them
(in ascending order) the Orientolophus, Homogalax and Heptodon interval zones. However, no stratigraphic superposition documents these
Bumbanian subdivisions, and the evolutionary trajectories that supposedly support it are not documented. Clearly, subdivision of the Bumbanian
lma should be possible, but the data provided by basal perissodactyls are
as yet insufficient to do so.
ACKNOWLEDGMENTS
Suyin Ting generously provided casts of the type and referred
specimens of Orientolophus hengdongensis. Donald Prothero and Robert
Sullivan provided helpful comments on the manuscript. The work of
PK was supported by the Russian Foundation for Basic Research, project
02-04-48458 and the Grant Council of the President of Russian Federation project Nsh-1840.2003.4.

REFERENCES
Chow, M., 1959, Eocene vertebrates from Sinyu, Kiangsi Province: Vertebrata
PalAsiatica, v. 3, p. 106.
Chow, M. and Li, C., 1963, A fossil of Homogalax from the Eocene of Shantung:
Scientia Sinica, v. 12, p. 1411-1412.
Chow, M. and Li, C., 1965, Homogalax and Heptodon of Shantung: Vertebrata
PalAsiatica, v. 9, p. 15-21.

Cooper, C. F., 1932, The genus Hyracotherium. A revision and description of new
specimens found in England: Philosophical Transactions Royal Society of London B, v. 221, p. 431-448.
Dashzeveg, D., 1979a, Nakhodka Homogalax (Perissodactyla, Tapiroidea) v
Mongolii i evo stratigraficheskoye znacheniye [Discovery of Homogalax
(Perissodactyla, Tapiroidea) in Mongolia and its stratigraphic significance]:

220
Byulletin Moskovskovo Obshchestvo Ispytately Prirody Otdel
Geologicheskiy, v. 54, p. 105-111.
Dashzeveg, D., 1979b, Nakhodka girakoteriya v Mongolii [The find of hyracothere
in Mongolia]: Paleontologicheskiy Zhurnal [Paleontological Journal], v. 3, p.
108-113.
Froehlich, D. J., 1999, Phylogenetic systematics of basal perissodactyls: Journal of
Vertebrate Paleontology, v. 19, p. 140-159.
Froehlich, D. J., 2002, Quo vadis eohippus? The systematics and taxonomy of the
early Eocene equids (Perissodactyla): Zoological Journal of the Linnean Society, v. 134, p. 141-256.
Gingerich, P. D., 1981, Systematics and evolution of early Eocene Perissodactyla
(Mammalia) in the Clarks Fork Basin, Wyoming: University of Michigan Contributions from the Museum of Paleontology, v. 28, p. 181-213.
Gingerich, P. D., 1991, Systematics and evolution of early Eocene Perissodactyla
(Mammalia) in the Clarks Fork Basin, Wyoming: Contributions from the Museum of Paleontology The University of Michigan, v. 28, p. 181-213.
Godinot, M., 1981, Les mammifres de Rians (ocne infrieur, Provence):
Palaeovertebrata, v. 10, p. 43-126.
Granger, W., 1908, A revision of the American Eocene horses: Bulletin of the American Museum of Natural History, v. 24, p. 221-264.
Hooker, J. J., 1980, The succession of Hyracotherium (Perissodactyla, Mammalia)
in the English early Eocene: Bulletin of the British Museum of Natural History
(Geology), v. 33, p. 101-114.
Hooker, J. J., 1984, A primitive ceratomorph (Perissodactyla, Mammalia) from the
early Tertiary of Europe: Zoological Journal of the Linnaean Society, v. 82, p.
229-244.
Hooker, J. J., 1989, Character polarities in early perissodactyls and their significance for Hyracotherium and infraordinal relationships; in Prothero, D. R. and
R. M. Schoch, eds., The evolution of the Perissodactyla: New York, Oxford
University Press, p. 79-101.
Hooker, J. J., 1994, The beginning of the equoid radiation: Zoological Journal of
the Linnaean Society of London, v. 112, p. 29-63.
Kitts, D. B., 1956, American Hyracotherium (Perissodactyla, Equidae): Bulletin
of the American Museum of Natural History, v. 110, p. 1-60.
Li, C. and Ting, D., 1983, The Paleogene mammals of China: Bulletin of the Carnegie
Museum of Natural History, v. 21, p. 1-93.
Lucas S.G., 1998, Fossil mammals and the Paleocene/Eocene series boundary in
Europe, North America, Asia; in Aubry, M.-P., Lucas, S.G., Berggren, W.A.,
eds., Late Paleocene-early Eocene cliamtic and biotic events in the
marine and terrestrial records: New York, Columbia University Press, p.

451-500.
Lucas, S. G., Holbrook, L. T. and Emry, R. J., 2003, Isectolophus (Mammalia,
Perissodactyla) from the Eocene of the Zaysan Basin, Kazakstan and its
biochronological significance: Journal of Vertebrate Paleontology, v. 23, p. 238243.
MacFadden, B. J., 1976, Cladistic analysis of primitive equids, with notes on other
perissodactyls: Systematic Zoology, v. 25, p. 1-14.
MacFadden, B. J., 1998, Equidae; in Scott, K. M., Janis, C. M. and Jacobs, L. L.,
eds., Tertiary mammals of North America: Cambridge, Cambridge University
Press, p. 537-559.
McKenna, M. C., 1975, Toward a phylogenetic classification of the Mammalia; in
Luckett, W. P. and Szalay, F. S., eds., Phylogeny of the Primates: New York,
Plenum Press, p. 21-46.
Russell, D. E. and Zhai, R., 1987, The Paleogene of Asia: Mammals and stratigraphy: Mmoire du Musum National dHistoire Naturelle Serie C Sciences de la
Terre, v. 52, 488 p.
Savage, D. E., Russell, D. E. and Louis, P., 1965, European Eocene Equidae: University of California Publications in Geological Sciences, v. 56, 94 p.
Simpson, G. G., 1952, Notes on British hyracotheres: Journal of the Linnaean Society of London, v. 42, p. 195-206.
Ting, S., 1993, A preliminary report on an early Eocene mammalian fauna from
Hengdong, Hunan Province, China: Kaupia [Darmstdter Beitrge zur
Naturgeshichte], v. 3, p. 201-207.
Ting, S., 1998. Paleocene and early Eocene land mammal ages of Asia: Bulletin of
the Carnegie Museum of Natural History, v. 34, p 124-147.
Tong, Y. and Wang, J., 1995, Mammals of the early Eocene Wutu fauna, Shandong
Province, China: Journal of Vertebrate Paleontology, 15, supplement to no.3, p.
57A.
Tong, Y., Zheng, S. and Qiu, Z., 1995, Cenozoic mammal ages of China: Vertebrata PalAsiatica, v. 33, p. 290-314.
Winans, M. C., 1989, A quantitative study of North American fossil specimens of
the genus Equus; in Prothero, D. R. and R. M. Schoch, eds., The evolution of
the Perissodactyla: New York, Oxford University Press, p. 262-297.
Xu, Y., Yan, D., Zhou, S., Han, S. and Zhong, Y., 1979, On stratigraphic subdivision of the red beds and the mammal fossils in the Liguanqiao Basin, Henan; in
The Mesozoic-Cenozoic red beds of southern China: Beijing, Science Press, p.
416-432.
Zhai, R.,1978, More fossil evidence favoring an early Eocene connection between
Asia and the Nearctic: Memoirs of the Institute of Vertebrate Paleontology and
Paleonanthropology, v. 13, p. 68-81.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

221

EOCENE MAMMALS FROM THE LIGUANQIAO BASIN, HENAN, CHINA, AND THE
BOUNDARY BETWEEN THE BUMBANIAN AND ARSHANTAN LAND-MAMMAL AGES
SPENCER G. LUCAS
New Mexico Museum of Natural History, 1801 Mountain Road N. W., Albuquerque, New Mexico 87104

AbstractIn the Liguanqiao basin of Henan, China, Eocene mammals from the uppermost Yuhuangding Formation and base of the overlying Dacangfang Formation are: Asiocoryphodon conicus (= A. progressivus),
Heterocoryphodon flerowi, Trogosus shantunensis (= Kuanchuanius danjiangensis), Gobiatherium minutum,
Yimengia yani, Eomoropus quadridentatus (= E. zhanggouensis), Schlosseria magister (= Lophialetes primus)
and Hyracodontidae. These mammals are of Arshantan age, and are stratigraphically above Bumbanian-age
mammals from the middle member of the Yuhuangding Formation. Rhombomylus is the best index fossil of the
Bumbanian. The mammal assemblages of the Yuhuangding Formation are a relatively rare record of
stratigraphically-superposed mammal assemblages of Bumbanian and Arshantan age.
Key words: China, Henan, Liguanqiao basin, Yuhuangding Formation, Bumbanian, Arshantan

INTRODUCTION
The Liguanqiao Basin, located on the border of Hubei and Henan
Provinces in China, is a Late Cretaceous-Cenozoic intermontane halfgraben developed in the eastern Qinling Mountains (Fig. 1). Paleogene
deposits in the basin are mostly mudstone, sandstone, conglomerate
and marlstone that represent alluvial and lacustrine lithofacies. Among
them, the Eocene deposits are the thickest and have, since the 1920s,
yielded a large number of mammalian fossils, thus providing a basis
for biostratigraphic subdivision of the strata (e.g., Li and Ting, 1983;
Russell and Zhai, 1987).
Particularly significant are the Eocene fossil mammal assemblages
from the Yuhuangding Formation. Stratigraphic organization of the mammal localities indicates that there are three, superposed mammal assemblages in the Yuhuangding Formation, two of Bumbanian age and one of
Arshantan age. Here, I revise the alpha taxonomy of the mammals from
the Arshantan assemblage, and discuss the significance of the fossil mam-

mals from the Yuhuangding Formation for Asian land-mammal


biochronology.
Institutional abbreviations: GV = Beijing Geological University; IVPP = Institute of Vertebrate Paleontology and Paleoanthropology,
Beijing.
STRATIGRAPHIC CONTEXT
The Eocene strata in the Liguanqiao basin have been subdivided
into three formations (from lower to upper): Yuhuangding, Dacangfang
and Hetaoyuan formations (Fig. 1). Disconformably overlying the Upper Cretaceous Hugang Formation and conformably underlying the
middle Eocene Dacangfang Formation (Fig. 1), the Yuhuangding Formation is well-exposed in an east-west trending outcrop belt, and is
subdivided into three members (lower, middle, and upper). There are
nine fossil localities in the formation distributed in the three members,
and they represent three fossiliferous intervals (Ma and Cheng, 1991).

FIGURE 1. Geologic map of the Liguanqiao basin in Henan, China (after Ma and Cheng, 1991) showing distribution of fossil mammal localities of the Yuhuangding and
Dacangfang formations.

222

FIGURE 2. Coryphodontid pantodonts from the Liguanqiao basin. A-B, Asiocoryphodon conicus (holotype of Asiocoryphodon progressivus), GV 86024, left dentary
fragment with p2-m3, occlusal (A) and labial (B) views. C, A. conicus, GV 85001, occlusal view of left P4. D, Heterocoryphodon flerowi, GV 86030, occlusal view of
right M3. E, H. flerowi, GV 85003, occlusal view of right m1. Scale bars are 1 cm long.

Cheng and Ma (1990) documented fossil mammals from two localities near Zhanggou village. The stratigraphically lower site, 84003,
is in the upper member of the Yuhuangding Formation and yielded the
specimens of Asiocoryphodon, Heterocoryphodon, Gobiatherium,
Schlosseria, Yimengia, Eomoropus and Hyracodontidae discussed below. The stratigraphically higher locality, 86001, is at the base of the
Dacangfang Formation and yielded the specimens assigned to Trogosus
here.
PALEONTOLOGY
Here, I revise the taxonomy of the fossil mammals from the
Liguanqiao basin documented by Cheng and Ma (1990).
Pantodonta
Asiocoryphodon and Heterocoryphodon
The most common taxa in the collections described by Cheng
and Ma (1990) are coryphodontid pantodonts. They created a new spe-

cies, Asiocoryphodon progressivus, for a left dentary fragment with p2m3 (GV 86024: Fig. 2A-B). They diagnosed this species as being of
medium size with weak pre- and postprotocristae on the upper premolars,
M2-3 hypocones directed at the metacones, upper molar paracones with
antero-posteriorly directed crista and m3 with a weak hypoconulid crista.
Only the last feature can be evaluated on the holotype, and it does not
differ significantly from the holotype of A. conicus (Xu, 1976; Lucas
and Tong, 1987). The size of the holotype of A. progressivus also is
slightly larger than that of the holotype of A. conicus (Table 1).
One problem with A. progressivus is that some of the teeth referred to the taxon by Cheng and Ma (1990) actually belong to
Heterocoryphodon. The holotype lower jaw fragment belongs to
Asiocoryphodonit is relatively small, and the molars lack the massive metaconids characteristic of Heterocoryphodon. However, a lower
molar Cheng and Ma (1990) assigned to A. progressivus as an m1 actually is a m1 of Heterocoryphodonnote especially its massive metaconid (Fig. 2E). In contrast, an upper premolar (Fig. 2C) has a short,
deep ectoflexus that precludes assignment to Heterocoryphodon. It is

223
Perissodactyla
TABLE 1. Measurements (in mm) of the holotype of Asiocoryphodon
progressivus (GV 86024) and the holotype of A. conicus (IVPP V
5141). Measurements are: l = length, tal = talonid, tri =trigonid, w =
width.
measurement:
p2 l
p2w
p3l
p3w
p4l
p4w
m1l
m1 trig w
m1 tal w
m2l
m2 trig w
m2 tal w
m3l
m3 trig w

GV 86024
19.4
18.9
21.4
22.6
22.1
25.1
32.5
24.1
25.1
37.6
28.5
28.7
37.3
28.4

IVPP V 5141
18.8
15.6
19.5
18.4
20.2
19.7
26.9
19.9
21.4
32.6
25.1
24.8
36.0
25.2

indistinguishable from the P4 of A. conicus. A right M3 (Fig. 2D), however, has a very long metaloph, a characteristic of Heterocoryphodon
that distinguishes it from Asiocoryphodon.
So, if I restrict A. progressivus to its holotype, the specimen is
slightly larger than but otherwise identical to A. conicus (Xu, 1976;
Lucas, 1984; Lucas and Tong, 1987). Therefore, I consider A.
progressivus to be a junior subjective synonym of A. conicus. The other
specimens assigned to A. progressivus by Cheng and Ma (1990) are
either A. conicus or Heterocoryphodon flerowi (Fig. 2).
Tillodontia
Trogosus
Cheng and Ma (1990) named a new species of tillodont,
Kuanchuanius danjiangensis, for GV 86033, a left P4 (Fig. 3C-D),
misidentified by them as a M1. Kuanchuanius was named by Chow
(1963) for an incomplete lower dentition, so assigning an upper tooth
to the genus was provisional, as Cheng and Ma (1990) conceded.
GV 86033 clearly belongs to a trogosine tillodont, and it is very
similar to P4s of Trogosus species from North America (see especially
Gazin, 1953, pls. 3, 7). Lucas (1993) considered Kuanchuanius to be a
synonym of Trogosus, and GV 86033 is consistent with this conclusion
(but see Miyata and Tomida [1998a, b] for another view of the taxonomic status of Kuanchuanius). The so-called diagnostic features of K.
danjiangensis are largely based on misidentification of the tooth as a
M1. The m1-3 length of Trogosus shantunensis is 64 mm (Chow, 1963,
p. 98), intermediate between the size of Trogosus hillsi, T. hyracoides
(larger) and T. grangeri (smaller). The size of the holotype P4 of K.
danjiangensis (length = 13.3 mm, width =20.9 mm) is also intermediate between the P4 sizes of these North American species. Therefore, I
consider K. danjiangensis a junior subjective synonym of T. shantunensis
and assign GV 86033 to T. shantunensis.
Dinocerata
Gobiatherium
Cheng and Ma (1990) named the new species Gobiatherium
minutum for GV 85011, a left dentary with dp2-m1 (Fig. 3A-B). Lucas
(2001) reviewed this taxon, concluding that it is a valid species of
Gobiatherium distinguished by its small size. G. minutum is only known
from its type locality in the Liguanqiao basin.

Yimengia
Cheng and Ma (1990) described GV 85015, a left dentary with
dp3-4 and m1-2 (Fig. 3E-F) as Yimengia sp. Measurements of the m1
(length = 7.7 mm, trigonid width = 4.9 mm, talonid width = 5.2 mm)
are very close to those of Yimengia yani (Wang, 1988). However, Cheng
and Ma (1990, p. 242) refrained from assigning GV 85015 to this species because of the supposed distinct lower molar conids and relatively
weak protolophids. Nevertheless, I am unable to replicate these differences, so I assign GV 85015 to Y. yani.
Eomoropus
Cheng and Ma (1990, p. 236) named the new species Eomoropus?
zhanggouensis for GV 86083, a left m1 (Fig. 3J). This tooth is relatively light colored and lacks roots, so it was probably never erupted. It
closely resembles the m1 of Eomoropus quadridentatus, so I am confident of its assignment to the genus (Cheng and Ma also suggested it
might belong to Propalaeotherium).
Cheng and Ma (1990) diagnosed E. zhanggouensis by its smaller
size, distinct main cusps and massive paralophid. However, the morphology of this tooth and its size (Fig. 3G; length = 13.1 mm, trigonid
width = 7.5 mm, talonid width = 8.8 mm) are indistinguishable from E.
quadridentatus (see Lucas and Schoch, 1989) and larger than E. minimus
(see Huang, 2002). A referred m3 (length = 16.5 + mm, trigonid width
= 9.1 mm, talonid width = 8.8 mm) is also a good match for E.
quadridentatus. Therefore, I consider E. zhanggouensis to be a junior
subjective synonym of E. quadridentatus. GV 85018 (Cheng and Ma,
1990, pl. 2, fig. 14) is a left dentary with a damaged dp3-4 that may
pertain to Eomoropus.
Schlosseria
Cheng and Ma (1990) named the new species Lophialetes?
primus for GV 86064, a right dentary fragment with part of m2 and a
complete m3 (Fig. 3I). Assignment to Lophialetes seems reasonable
(note especially the prominent and high m3 metalophid and the looplike hypoconulid), but the size of GV 86064 (m2 talonid width = 6.2
mm, m3 length = 15.7 mm, trigonid width = 6.4 mm, talonid width =
6.5 mm) is small for Lophialetes expeditus, though within the size range
of Schlosseria.
This specimen reopens the unresolved question of how to identify advanced specimens of Schlosseria and primitive specimens of
Lophialetes, which overlap in size and morphology. GV 86064 seems
to best match specimens Radinsky (1965, table 6) assigned to
Schlosseria cf. S. magister, so I synonymize L. primus with S. magister. However, I also note that GV 86064 is very similar to AMNH 81777,
a specimen intermediate in size and morphology between S. magister
and L. expeditus and that Radinsky (1965, p. 198) stated cannot with
certainty be assigned to one or the other genus.
Dashzeveg and Hooker (1997, p. 107) also discussed the problem and assigned material from Mongolia to Lophialetes sp. that may
belong to the same taxon as GV 86064. It may be that L. primus should
be a valid species of Lophialetes, used to identify the morphology intermediate between Schlosseria and Lophialetes.
Hyracodontidae
GV 86081 (Fig. 3H) is a badly damaged and distorted m1 of a
hyracodontid rhinoceros. Length = 17.6 mm, trigonid width = 12.8 mm,
talonid width = 12.8 mm. Cheng and Ma (1990) assigned this tooth to
Forstercooperia?, but I believe it cannot be identified more precisely
than Hyracodontidae
BUMBANIAN-ARSHANTAN BIOCHRONOLOGY
In the Liguanqiao basin, the Yuhuangding Formation yields fos-

224

FIGURE 3. Uintathere, tillodont and perissodactyls from the Liguanqiao basin. A-B, Gobiatherium minutum, GV 85011, holotype, left dentary fragment with dp2-m1. CD, Trogosus shantunensis (holotype of Kuanchuanius danjiangensis), GV 86033, left P4. E-F, Yimengia yani, GV 85015, left dp3-m2. G, Eomoropus quadridentatus,
GV 86084, left m3. H, Hyracodontidae, GV 86081, left m1?. I, Schlosseria magister (holotype of Lophialetes primus), GV 86064, right m2-3. J, Eomoropus quadridentatus
(holotype of Eomoropus zhanggouensis), GV 86083, left m1. Scale bars are 1 cm long.

225
sil mammals from three stratigraphic intervals (Gao, 1976; Xu et al.,
1979; Ma and Cheng, 1991; Du et al., 1991, 1992; Ma and Lucas, 1993)
(Fig. 4). The lowest stratigraphic interval is in the lower part of the
middle member of the Yuhuangding Formation and only yields the
coryphodontid pantodonts Asiocoryphodon conicus and
Heterocoryphodon flerowi (Chow, 1957; Xu, 1976; Lucas and Tong,
1987; Ma and Cheng, 1991).
The overlying interval (Fig. 4), higher in the middle member of
the Yuhuangding Formation, yielded the extensive assemblage of
Rhombomylus turpanensis recently documented by Meng et al. (2003),
as well as Asiocoryphodon conicus, Advenimus hubeiensis, Hohomys
lii, ?Hunanictis and cf. Heptodon sp. (Xu, 1976; Xu et al., 1979;
Dawson et al., 1984; Lucas and Tong, 1987).
Stratigraphically higher, the uppermost Yuhuangding Formation
and basal Dacangfang Formation yield the assemblage of fossils revised here (Fig. 4): Asiocoryphodon conicus (= A. progressivus),
Heterocoryphodon flerowi, Trogosus shantunensis (= Kuanchuanius
danjiangensis), Gobiatherium minutum, Yimengia yani, Eomoropus
quadridentatus (= E. zhanggouensis), Schlosseria magister (=
Lophialetes primus) and Hyracodontidae.
The lower two assemblages from the Yuhuangding Formation
can be assigned a Bumbanian age. The middle assemblage includes
abundant Rhombomylus, which is restricted to the Bumbanian (Ting,
1998). Indeed, Rhombomylus is the best Bumbanian index fossil because it has a broad distribution in Bumbanian localities in China and
Mongolia, being known from the Shisanjiangfang Formation in the
Turpan basin of Xinjiang (Zhai, 1978), the Zhangshanji Formation at
Laian in Anhui (Zhai et al., 1976), the Youping Formation at Fangxian,
Hubei (Lei, 1984; Huang, 1995), the Yuhuangding Formation (Meng et
al., 2003) and the Bumban Member of the Naran Bulak Formation at
Tsagan-Khushu, Mongolia (Dashzeveg et al., 1987; Dashzeveg and
Russell, 1988).
The occurrence of Schlosseria and Gobiatherium in the upper
assemblage of the Yuhuangding Formation indicates it is of Arshantan
age (Lucas, 2001). Neither the Bumbanian nor the Arshantan mammal
assemblages in the Yuhuangding Formation are diverse or extensive,
but they are stratigraphically superposed. This means the BumbanianArshantan boundary is in the Yuhuangding Formation, and the assemblages thus are a rare example of the direct superposition of mammal
assemblages of Bumbanian and Arshantan age.
ACKNOWLEDGMENTS
I am grateful to Ancheng Ma, who made it possible for me to
study the fossil mammals illustrated in this article, translated Chinese
text for me and provided other information. Peter Kondrashov provided helpful comments on the manuscript.

FIGURE 4. Summary of mammal distribution across the Bumbanian-Arshantan


boundary in the Liguanqiao basin.

REFERENCES
Cheng, J. and Ma, A., 1990. The new mammalian materials from the Eocene of the
Liguanqiao Basin in Qinling Mts.: Vertebrata PalAsiatica, v. 28, p. 228-243.
Chow, M., 1957, A new Coryphodon from Sintai, Shantung: Vertebrata PalAsiatica,
v. 1, p. 301-304.
Chow, M., 1963, Tillodont materials from Eocene of Shantung and Honan: Vertebrata PalAsiatica, v. 7, p. 98-104.
Dashzeveg, D. and Hooker, J. J., 1997, New ceratomorph perissodactyls (Mammalia) from the middle and late Eocene of Mongolia: Their implications for phylogeny and dating: Zoological Journal of the Linnaean Society, v. 1997, p. 105138.
Dashzeveg, D. and Russell, D. E., 1988, Paleocene and Eocene Mixodontia (Mammalia, Glires) of Mongolia and China: Palaeontology, v. 31, p. 129-164.
Dashzeveg, D, Russell, D. E. and Flynn, L. J., 1987, New Glires (Mammalia) from
the early Eocene of the Peoples Republic of Mongolia. 1. Systematics and description: Proceedongs of the Koninklijke Nederlandse Akademie van
Wetenschappen Series B Physical Sciences, v. 90, p. 133-154.

Dawson, M. R., Li, C. and Qi, T., 1984, Eocene ctenodactyloid rodents (Mammalia) of eastern central Asia: Carnegie Museum of natural History, Special Publication 9, p. 138-150.
Du, H., Cheng, J., Wu, W. and Ma, A., 1991, A discussion on Paleogene mammalian fossil horizons and concerned problems of the Liguanqiao and Yuanqu Basins: Earth Science Journal of China University of Geosciences, v. 16, p. 113126.
Du, H., Ma, A., Cheng, J. and Wu, W., 1992, Paleogene biostratigraphic characteristics of the red basins in western Henan and its adjacent areas, with an outline of
Paleogene paleobiogeographic provinces of China: Acta Geologia Sinica, v. 5,
p. 101-118.
Gao, Y., 1976, Mammalian fossil localities and horizons in the Wucheng and Xichuan
Basins, Henan: Vertebrata PalAsiatica, v. 14, p. 26-34.
Gazin, C. L., 1953, The Tillodontia: An early Tertiary order of mammals:
Smithsonian Miscellaneous Collections, v. 121(10), 110 p.
Huang, X., 1995, A new Hyopsodus from the early Eocene of Fangxian, Hubei:

226
Vertebrata PalAsiatica, v. 33, p. 39-46.
Huang, X., 2002, New eomoropoid (Mammalia, Perissodactyla) remains from the
middle Eocene of Yuanqu basin: Vertebrata PalAsiatica, v. 40, p. 286-290.
Lei, Y., 1984, The analysis of the mammal fauna of the Paleogene of central-southern China: Journal of the Yichang Institute of Geology and Minerals, Chinese
Academy of Geology, v. 8, p. 41-50.
Li, C. and Ting, S., 1983, The Paleogene mammals of China: Bulletin Carnegie
Museum of Natural History, v. 21, p. 1-98.
Lucas, S. G., 1984, Systematics, biostratigraphy and evolution of early Cenozoic
Coryphodon (Mammalia, Pantodonta) [PhD. dissertation]: New Haven, Yale
University, 649 p.
Lucas, S. G., 1993, Pantodonts, tillodonts, uintatheres and pyrotheres are not ungulates; in Szalay, F. S., Novacek, M. J. and McKenna, M. C., eds., Mammal phylogeny: Placentals: New York, Springer-Verlag, p. 182-194.
Lucas, S. G., 2001, Gobiatherium (Mammalia: Dinocerata) from the middle Eocene
of Asia: Taxonomy and biochronological significance: Palaontologische
Zeitschrift, v. 74, p. 591-600.
Lucas, S. G. and Schoch, R. M., 1989, Taxonomy and biochronology of Eomoropus
and Grangeria, Eocene chalicotheres from the western United States and Asia;
in Prothero, D. R. and Schoch, R. M., eds., The evolution of perissodactyls:
New York, Oxford University Press, p. 422-437.
Lucas, S. G. and Tong, Y., 1987, A new coryphodontid (Mammalia, Pantodonta)
from the Eocene of China: Journal of Vertebrate Paleontology, v. 7, p. 362-372.
Ma, A. and Cheng, J., 1991, On biostratigraphics in the sequences, newly collecting
of mammalian taxa from old localities, additional measuring of four stratigraphic
sections, and most important, the new discoveries of several Eocene vertebrate
localities in different formations: Scientia Geologia Sinica, v. 1, p. 21-29.
Ma, A. and Lucas, S. G., 1993, Stratigraphy and mammalian biochronology of the
Eocene Yuhuangding Formation, Liguanqiao Basin, Henan, China: Journal of
Vertebrate Paleontology, v. 13, supplement to no. 3, p. 47A.

Meng, J., Hu, Y. and Li, C., 2003, The osteology of Rhombomylus (Mammalia:
Glires): Implications for phylogeny and evolution of Glires: Bulletin of the
American Museum of Natural History, no. 275, 247 p.
Miyata, K. and Tomida, Y., 1998a, A new tillodont from the early middle Eocene of
Japan and its implication to the sufamily Trogosinae (Tillodontia: Mammalia):
Paleontological Research, v. 2, p. 53-66.
Miyata, K. and Tomida, Y., 1998b, Trogosus-like tillodont (Tillodontia, Mammalia) from the early middle Eocene of Japan: Paleontological Research, v. 2, p.
193-198.
Radinsky, L. B., 1965, Early Tertiary Tapiroidea of Asia: Bulletin of the American
Museum of Natural History, v. 129, p. 181-264.
Russell, D. E. and Zhai, R., 1987, The Paleogene of Asia: Mammals and stratigraphy: Mmoires du Museum National dHistoire Naturelle Srie C, Sciences de
la Terre, v. 52, 488 p.
Ting, S., 1998, Paleocene and early Eocene land mammal ages of Asia: Bulletin of
the Carnegie Museum of Natural History, v. 34, p. 124-147.
Wang, J., 1988, A new genus of ceratomorphs (Mammalia) from the middle Eocene
of China: Vertebrata PalAsiatica, v. 26, p. 20-34.
Xu, Y., 1976, Early Eocene coryphodontids from Xichuan, Henan: Vertebrata
PalAsiatica, v. 14, p. 185-193.
Xu, Y., Yan, D., Zhou, S., Han, S. and Zhong, Y., 1979, On stratigraphic subdivision of the red beds and the mammal fossils in the Liguanqiao Basin, Henan; in
The Mesozoic-Cenozoic red beds of southern China: Beijing, Science Press, p.
416-432.
Zhai, R., 1978, More fossil evidence favoring an early Eocene connection between
Asia and the Nearctic: Institute of Vertebrate Paleontology and Paleoanthropology
Beijing, Memoir 13, p. 107-115.
Zhai, R., Bi, Z. and Yu, Z., 1976, Stratigraphy of Eocene Zhanshanji Formation
with note on a new species of eurymylid mammal: Vertebrata PalAsiatica, v. 14,
p. 100-103.

Lucas, S.G., Zeigler, K.E. and Kondrashov, P.E., eds., 2004, Paleogene Mammals, New Mexico Museum of Natural History and Science Bulletin No. 26.

227

EOCENE PANTOLESTA FROM THE ZAYSAN BASIN, KAZAKSTAN


SPENCER G. LUCAS1 AND ROBERT J. EMRY2
2

1
New Mexico Museum of Natural History, 1801 Mountain Rd. NW, Albuquerque, NM 87104;
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, D. C. 20560

AbstractTwo taxa of pantolestan eutherians are known from upper Eocene (Ergilian-age) strata in the Zaysan
basin, Kazakstan from their holotype specimens. Kiinkerishella zaisanica Gabuniya & Biryukov is known from
a dentary fragment with m3 from the upper part of the Aksyir svita and appears to be a dyspternine. Oboia
argillaceous Gabuniya from the Kusto svita is known from a dentary fragment with p4-m1 and appears to be a
pantolestine. Kiinkerishella and Oboia thus extend the modest diversification of pantolestans during the late
Eocene (Ergilian) from Western Europe into Central Asia and are evidence of the paleobiogeographic continuity
of land-mammal faunas across Eurasia during the late Eocene.
Key words: Pantolesta, Eocene, Kiinkerishella, Oboia, Kazakstan
INTRODUCTION
Pantolesta is an order of eutherian mammals with a fossil record
from Paleogene strata in North America, Europe, Africa and Asia (e.g.,
Van Valen, 1965; Cray, 1973; Heissig, 1977; Koenigswald, 1983; Bown
and Simons, 1987; Russell and Godinot, 1988; McKenna and Bell,
1997; Pfretzschner, 1999; Smith, 2001). Rarely common in fossil mammal assemblages, Pantolesta are particularly rare in Asia, with only a
handful of published occurrences. Indeterminate Asian pantolestan
records are from the Bumbanian (early Eocene) Tsagan Khushu locality, Mongolia and the Irdinmanhan (middle Eocene) Irdin Manha locality in Nei Monggol, China (Russell and Dashzeveg, 1986; Russell and
Zhai, 1987). The oldest named Asian pantolestan is Bogdia orientalis
from the Irdinmanhan of the Valley of Lakes Depression, Mongolia
(Dashzeveg and Russell, 1985; Russell and Zhai, 1987). Two records
from the Ergilian (late Eocene) of the Zaysan basin in eastern Kazakstan
(Fig. 1) are reviewed here. The fourth record is of Gobiopithecus khan
from the Ergilian of Khoer Dzan, Mongolia (Dashzeveg and Russell,
1992). The Kazak records demonstrate that two of the principal clades
of Pantolesta were present in Asia during the late Eocene.
Institutional abbreviations: IPGAN = Institute of Paleobiology,
Georgian Academy of Sciences, Tbilisi, Georgia; KAN = Kazak Academy of Sciences, Institute of Zoology, Almaty, Kazakstan.
KIINKERISHELLA
The holotype and only known specimen of Kinkerishella
zaisanica Gabuniya & Biryukov, KAN 3/7/289/1406, is a right dentary
fragment with the roots/alveoli of m1 and m2, part of the talonid of m1
or m2 and a complete m3 (Fig. 2). The fossil is from sopka K (hill
K) near Kin Kerish Mountain on the northern shore of Lake Zaysan
(UTM zone 45, 312894E, 5333986N, WGS 84). This locality is in the
upper part of the Aksyir svita, strata of Ergilian (late Eocene) age (Emry
et al., 1998). Russell and Zhai (1987, p. 230-231) placed the stratigraphic level as lower part of the Aksyir svita, but this is in error.
Gabuniya and Biryukov (1978) described Kiinkerishella
zaisanica as a condylarth; the holotype lower jaw had earlier been
misidentified as Plesiadapis (Borisov, 1963; Bazhanov and Kostenko,
1964; Biryukov and Kostenko, 1965). Russell and Godinot (1988) illustrated the m3 of the type and referred Kiinkerishella to the
Paroxyclaenidae. Dashzeveg and Russell (1992) allied it with
Gobiopithecus and Dyspterna in the Dyspterninae, but McKenna and
Bell (1997, p. 219) listed Kiinkerishella as a paroxyclaenid outside of
the subfamilies Paroxyclaeninae and Merialinae. Lucas and Emry (2000)
also allied Kiinkerishella to Paroxyclaenus.
The m1 evidently was a double-rooted tooth that was relatively
short antero-posteriorly. The m2 appears to have been similar in overall proportions to the m1. The specimen now has a talonid in place as

FIGURE 1. Location map of Zaysan basin in eastern Kazakstan.

part of the m2, but Gabuniya and Biryukov (1978, fig. 2) illustrated
this as the talonid of the m1. It has a well developed basin with a
distinct hypoconid, hypoconulid and entoconid connected to each other
by low cristids. The m3 is a nearly round, peg-like tooth with a wide
anterior cingulid, large protoconid, essentially no paraconid and small
metaconid. The length = 4.8 mm and width = 3.5 mm. The talonid is
loop-like with a distinct hypoconid and cingulid-like talonid basin. The
horizontal ramus is relatively low, and its ventral border projects ventrally below the m3. The masseteric fossa is not divided.
OBOIA
The holotype and only known specimen of Oboia argillaceous is
IPGAN 221, a right dentary fragment with the canine and p4-m1 (Fig.
3). The fossil is from the pantserny sloy (carapace layer) locality.
This locality is in the middle part of the Kusto svita on the northern
shore of Lake Zaysan near Kiin Kerish Mountain (UTM zone 45,
31667E, 5334660N, WGS 84).The Kusto svita is of Ergilian age (Emry
et al., 1998).
Gabuniya (1989) originally assigned Oboia to the Pantolestinae,

228

FIGURE 3. IPGAN 211, holotype of Oboia argillaceous Gabuniya, right dentary


fragment with canine, p4-m1 in occlusal stereophotograph (A), lingual (B) and labial
(C) views.

FIGURE 2. KAN 3/7/289/1406, holotype of Kiinkerishella zaisanica Gabuniya &


Biryukov, right dentary fragment with m3 in occlusal stereophotograph (A), labial
(B) and lingual (C) views.

an assignment reiterated by Dashzeveg and Russell (1992) and McKenna


and Bell (1997). Lucas and Emry (2000) posited a close relationship
between Oboia and Euhookeria, a merialine paroxyclaenid.
The canine of Oboia is large, grooved, recurved and has a trihedral cross section. It is followed by three closely spaced alveoli, two
teeth and four alveoli for the last two molars. The alveoli suggest the
presence of a double-rooted p3 similar in shape to the p4. However,
anterior to the p3 alveoli is only one alveolus, and a portion of the
dentary immediately posterior to the canine is missing. This makes it
impossible to be certain of the premolar configuration anterior to the
p3, other than to say that a p2 was present.
The alveoli behind the m1 indicate a double-rooted, anteriorposteriorly rectangular m2 about the same size as the m1. The m3 alveoli, however, are about 25% smaller than those of the m2, and they

indicate an elongate tooth with a trigonid wider than the talonid.


The p4 is a trenchant tooth with two large, circular roots. The
crown is dominated by a large, pointed trigonid cuspid that has a small
antero-lingual cingulid connected to it by a median ridge. The p4 talonid is a low cingulid bounded lingually and posteriorly by a ridge on
the posterior slope of the main cuspid. The m1 is rectangular with equalsized metaconid and protoconid and a somewhat smaller paraconid located median to the protoconid-metaconid cristid and connected to the
protoconid by a distinct paracristid. The trigonid is higher than the
talonid, but the talonid is wider than the trigonid. The talonid has a
broad basin bounded labially by a large hypoconid connected by a distinct cristid obliqua to the posterior wall of the trigonid. A small, crestlike hypoconulid is present, as is a somewhat larger entoconid. In mm,
the p4 length = 5.5, width = 3.2; m1 length = 6.1 and width = 4.2.
The horizontal ramus is long and low with a mental foramen
between p4 and m1. The ascending ramus is low, and the condyle is
just above the occlusal plane of the cheek teeth. There is a ventral
extension of the horizontal ramus below m3, as in Kinkerishella, but it
is not as pronounced. The masseteric fossa is divided into two parts by
a horizontal ridge.
DISCUSSION
Pantolestans are extremely rare as fossils in Asia. Kiinkerishella
and Oboia thus extend the modest diversification of pantolestans during the late Eocene (Ergilian) from Western Europe into Central Asia
and suggest some continuity of land-mammal faunas across Eurasia
during the late Eocene.
A detailed reappraisal of the phylogenetic relationships of
Pantolesta is needed but beyond the scope of this article. We thus rely
primarily on the work of Russell and Godinot (1988) and of McKenna

and Bell (1997) to assess the relationships of Kiinkerishella and Oboia


to other pantolestans in a preliminary way. As Dashzeveg and Russell
(1992) argued, Kiinkerishella is very incompletely known, but displays
a few key features (note especially the prominent anterior cingulid on
m3, antero-posteriorly compressed m3 trigonid, and absence of a m3
paraconid and entoconid) that suggest it may be closely related to
Gobiopithecus and Dyspterna and thus belongs in the subfamily
Dyspterninae. In contrast, Oboia is more completely known and displays key features (distinct m1 entoconid and a well developed, median

229
paraconid) that may ally it with the pantolestines. The two Kazak
pantolestan taxa thus indicate the presence of two distinct lineages of
pantolestansdyspternines and pantolestinesin Central Asia during
the late Eocene.
ACKNOWLEDGMENTS
The National Geographic Society and the Walcott Fund of the
Smithsonian Institution supported this research. L. Gabuniya and V.
Chikhvadze made possible access to specimens in Tbilisi.

REFERENCES
Bazhanov, V. S. and Kostenko, N. N., 1964, Korrelyatsiya otlozheniy kaynozoya
Kazakhstana i Indii po faune mlekopitayushchikh [Correlation of Cenozoic deposits of Kazakstan and India based on mammal faunas]: Materialy k XXII
Sessii Mezhdunarodnovo Geologicheskovo Kongressa, p. 82-95.
Biryukov, M. D. and Kostenko, N. N., 1965, Otnositelno obaylinskoy fauny
mlekopitayushchikh Zaysanskoy kotloviny [On the obaylian mammal fauna
of the Zaysan depression]: Vestnik Akademiia Nauk Kazakhskoy SSR, v. 12, p.
75-77.
Borisov, B. A., 1963, Stratigrafiya verkhnevo Mela i Paleogena-Neogena
Zaysanskoy vpadiny [Stratigraphy of the Upper Cretaceous and Paleogene-Neogene of the Zaysan basin]: Trudy VSEGEI Gosudarstveniy Geologickeskovo
Komineta, new series, v. 94, p. 11-75.
Bown, T. E. and Simons, E. L., 1987, New Oligocene Ptolemaiidae (Mammalia:
?Pantolesta) from the Jebel Quatrani Formation, Fayum depression, Egypt: Journal of Vertebrate Paleontology, v. 7, p. 311-324.
Cray, P. E., 1973, Marsupialia, Insectivora, Primates, Creodonta and Carnivora from
the Headon Beds (upper Eocene) of southern England: Bulletin of the British
Museum (Natural History) Geology, v. 23, p. 1-102.
Dashzeveg, D. and Russell, D. E., 1985, A new middle Eocene insectivore from the
Mongolian Peoples Republic: Geobios, v. 16, p. 871-875.
Dashzeveg, D. and Russell, D. E., 1992, Extension of dyspternine Pantolestidae
(Mammalia, Cimolesta) in the early Oligocene of Mongolia: Geobios, v. 25, p.
647-650.
Emry, R. J., Lucas, S. G., Tyutkova, L and Wang, B., 1998, The Ergilian-Shandgolian
(Eocene-Oligocene) transition in the Zaysan basin, Kazakstan: Bulletin of
Carnegie Museum of Natural History, v. 34, p. 298-312.
Gabuniya, L. K., 1989, O pervoy nakhodke pantolestid (Pantolestidae, Insectivora)
v Paleogene SSSR [On the first discovery of a pantolestid (Pantolestidae,
Insectivora) in the Paleogene of the USSR]: Soobshcheniya Akademiia Nauk
Gruzinskoi SSR, v. 136, p. 177-180.

Gabuniya, L. K. and Biryukov, N. D., 1978, O prisutstvii svoyebraznovo


predstavitelya arktotsionoidey (Arctocyonoidea) v Paleogene Azii [On the presence of a peculiar representative of the arctocyonoids in the Paleogene of Asia]:
Soobshcheniya Akademiia Nauk Gruzinskoi SSR, v. 92, p. 489-492.
Heissig, K., 1977, Neues Material von Cryptopithecus (Mammalia, Pantolestidae)
aus dem Mitteloligozn von Mhren 13 in Mittelfranken: Mitteilungen der
Bayerische Staatssammlungen fr Palontologie und Historische Geologie, v.
17, p. 213-225.
Koenigswald, W. v., 1983, Skelettfunde von Kopidodon (Condylarthra, Mammalia) aus dem mitteleoznen lschiefer von Messel bei Darmstadt: Neues Jahrbuch
fr Geologie und Palontologie Abhandlungen, v. 167, p. 1-39.
Lucas, S. G. and Emry, R. J., 2000, Eocene pantolestids from the Zaysan basin,
Kazakstan: Journal of Vertebrate Paleontology, v. 20, supplement to no. 3, p.
54A-55A.
McKenna, M. C. and Bell, S. K., 1997, Classification of mammals above the species level. New York, Columbia University Press, 631 p.
Pfretzschner, H-U., 1999, Buxolestes minor n. sp. ein neuer Pantolestide (Mammalia, Proteutheria) aus der eoznen Messel-Formation: Courier
Forschungsinstitut Senckenberg, v. 216, p. 1-18.
Russell, D. E. and Dashzeveg, D., 1986, Early Eocene insectivores (Mammalia)
from the Peoples Republic of Mongolia: Palaeontology, v. 29, p. 269-291.
Russell, D. E. and Godinot, M., 1988, The Paroxyclaenidae (Mammalia) and a new
form from the early Eocene of Palette, France: Palontologische Zeitschrift, v.
62, p. 319-331.
Russell, D. E. and Zhai, R., 1987, The Paleogene of Asia: Mammals and stratigraphy: Mmoires du Museum National dHistoire Naturelle Srie C, Sciences de
la Terre, v. 52, 488 p.
Smith, R., 2001, Les pantolestides (Mammalia, Pantolesta) de lEocene inferieur de
Premontre (Aisne, France): Palaeovertebrata, v. 30, p. 11-35.
Van Valen, L., 1965, Paroxyclaenidae, an extinct family of Eurasian mammals: Journal of Mammalogy, v. 46, p. 388-397.

230

Deltatherium fundaminis, skull and lower jaw (from Matthew, 1937).

Das könnte Ihnen auch gefallen