You are on page 1of 21

Journal of Vertebrate Paleontology

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/ujvp20

A sauropodomorph (Dinosauria, Saurischia)


specimen from the Upper Triassic of southern
Brazil and the early increase in size in
Sauropodomorpha

Rodrigo T. Müller & Maurício S. Garcia

To cite this article: Rodrigo T. Müller & Maurício S. Garcia (2022): A sauropodomorph (Dinosauria,
Saurischia) specimen from the Upper Triassic of southern Brazil and the early increase in size in
Sauropodomorpha, Journal of Vertebrate Paleontology, DOI: 10.1080/02724634.2021.2002879

To link to this article: https://doi.org/10.1080/02724634.2021.2002879

View supplementary material

Published online: 11 Feb 2022.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ujvp20
Journal of Vertebrate Paleontology e2002879 (20 pages)
© by the Society of Vertebrate Paleontology
DOI: 10.1080/02724634.2021.2002879

ARTICLE

A SAUROPODOMORPH (DINOSAURIA, SAURISCHIA) SPECIMEN FROM THE UPPER


TRIASSIC OF SOUTHERN BRAZIL AND THE EARLY INCREASE IN SIZE IN
SAUROPODOMORPHA

RODRIGO T. MÜLLER,1,2* and MAURÍCIO S. GARCIA2


1
Centro de Apoio à Pesquisa Paleontológica da Quarta Colônia, Universidade Federal de Santa Maria, Rua Maximiliano Vizzotto,
598, 97230-000, São João do Polêsine, Rio Grande do Sul, Brazil, rodrigotmuller@hotmail.com;
2
Programa de Pós-Graduação em Biodiversidade Animal, Universidade Federal de Santa Maria, 97105-120, Santa Maria, Rio
Grande do Sul, Brazil, mauriciossauro@gmail.com

ABSTRACT—Whereas sauropod dinosaurs from the Jurassic and Cretaceous Periods were the largest land animals that ever
lived, some of their early relatives evolved relatively large bodies during the Triassic Period. The evolutionary pathways
followed by the earliest sauropodomorphs towards the acquisition of massive bodies are poorly understood. However,
new finds from South America and Africa are reshaping our knowledge of this issue. Here, we describe a new early and
relatively large sauropodomorph represented by a partial postcranial skeleton excavated from Carnian-aged beds (Upper
Triassic) of southern Brazil. The new specimen is recovered as a sauropodomorph more closely related to bagualosaurians
than saturnaliids or other early-diverging forms in two phylogenetic analyses. The new specimen is generically
indeterminate but provides important evidence of an early increase in body size in Sauropodomorpha, being significantly
larger than that of coeval or older forms (except Bagualosaurus agudoensis). Furthermore, the specimen is about 3.2 times
heavier than Buriolestes schultzi, the earliest-branching sauropodomorph. The slender hind limbs and typical cursorial
proportions present in the earliest sauropodomorphs are mostly maintained in the new specimen despite its larger body size.

SUPPLEMENTAL DATA—Supplemental materials are available for this article for free at www.tandfonline.com/UJVP

Citation for this article: Müller, R. T. and M. S. Garcia. 2022. A sauropodomorph (Dinosauria, Saurischia) specimen from the
Upper Triassic of southern Brazil and the early increase in size in Sauropodomorpha. Journal of Vertebrate Paleontology.
DOI: 10.1080/02724634.2021.2002879

INTRODUCTION interval, Santa Maria Formation) yield the earliest members of


Dinosauria worldwide (233.23 ± 0.73 Ma; mid-Carnian; Langer
From birds to sauropods, dinosaurs achieved an extreme range
et al., 2018), including sauropodomorphs such as Saturnalia tupi-
of body sizes during their evolutionary history (Benson et al.,
niquim (Langer et al., 1999) and Buriolestes schultzi (Cabreira
2014), with a striking span during the Jurassic and Cretaceous
et al., 2016). These animals were small (femur ∼15 cm long),
Periods (Lacovara et al., 2014; Riga et al., 2016; Carballido
bipedal, and faunivorous (Cabreira et al., 2016; Bronzati et al.,
et al., 2017; Otero et al., 2021). Some non-sauropod sauropodo-
2017, 2019; Müller et al., 2021). The next-youngest stratigraphic
morphs evolved massive and large bodies during the Triassic
horizons (middle, Santa Maria Formation, precise age
Period (Pol and Powell, 2007; Apaldetti et al., 2018; Peyre de
unknown) contain sauropodomorphs displaying dental changes
Fabrègues and Allain, 2019). Towards the end of the Triassic
related to an omnivorous or mostly herbivorous diet, such as
(mid-Norian), quadrupedal posture was adopted for the first
Pampadromaeus barberenai (Cabreira et al., 2011) and Bagual-
time, which evolved rapidly alongside the evolution of increased
osaurus agudoensis (Pretto et al., 2018). The latter species
body mass (Langer et al., 1999; Martinez et al., 2011; Sereno
reveals a slight increase in body size, with femoral length over
et al., 2013; Cabreira et al., 2016; McPhee et al., 2018; Müller
20 cm (Pretto et al., 2018). Finally, the upper stratigraphic interval
et al., 2018a; Novas et al., 2021).
(Caturrita Formation, 225.42 ± 0.37 Ma; early Norian; Langer
Until recently, the evolution of body size in the earliest sauro-
et al., 2018) preserves sauropodomorphs exhibiting the typical
podomorphs has remained obscure. However, increased field-
bauplan of the so-called ‘prosauropods’ (sensu Sereno, 2007),
work has yielded several important finds (e.g., Cabreira et al.,
such as Unaysaurus tolentinoi (Leal et al., 2004) and Macrocollum
2011, 2016; Apaldetti et al., 2018; Müller et al., 2018b; Pretto
itaquii (Müller et al., 2018b). These animals were herbivorous,
et al., 2018) that have reshaped our knowledge of their early evol-
longer-necked, and larger, with femoral lengths over 30 cm
ution (Novas et al., 2021; Pol et al., 2021). Among these deposits,
(Leal et al., 2004; Müller et al., 2018b). Both Unaysaurus tolenti-
those from southern Brazil provide clues to the first steps of the
noi and Macrocollum itaquii lack any evidence of a quadrupedal
sauropodomorph radiation via three distinct horizons (Müller
stance, retaining the plesiomorphic bipedal condition of earliest
and Garcia, 2020) that encompass a ∼8 Ma time interval
sauropodomorphs (McPhee et al., 2018).
(Langer et al., 2018). The oldest horizons (i.e., lower stratigraphic
The fossil record from the middle stratigraphic interval is
scarce in comparison to that of the lower and upper intervals,
and its limited nature hampers the assessment of the macroevo-
*Corresponding author. lutionary trends that took place during this crucial period. For
Color versions of one or more of the figures in the article can be found instance, the anatomical innovations that accompanied the
online at www.tandfonline.com/ujvp.

Published online 11 Feb 2022


Müller and Garcia—An early sauropodomorph from Brazil (e2002879-2)

body size increase during the early evolution of sauropodo- 2020a). This fauna is intriguing because it serves as a bridge
morphs are poorly understood. Here, we explore this issue with between the land ecosystems with few dinosaurs and the first
the description of a new specimen from this transitional stage. environments dominated by dinosaurs (Martinez et al., 2013;
It was excavated from the ‘Várzea do Agudo’ site (southern Pol et al., 2021). Regarding cynodonts, the most abundant
Brazil) and is notable for its large size in comparison to the taxon is the traversodontid Exaeretodon riograndensis (Oliveira
oldest members of Sauropodomorpha. et al., 2007; Liparini et al., 2013; Müller et al., 2020a), whereas
Institutional Abbreviations—BP, Evolutionary Studies Insti- specimens of the ecteniniid Trucidocynodon riograndensis were
tute, Johannesburg, South Africa (previously the Bernard Price rare (Oliveira et al., 2010; Stefanello et al., 2018). On the other
Institute); BRSUG, University of Bristol, Geology Department, hand, archosauromorphs are represented by dinosaurs (e.g.,
Bristol, United Kingdom; CAPPA/UFSM, Centro de Apoio à Pampadromaeus barberenai and Bagualosaurus agudoensis;
Pesquisa Paleontológica da Quarta Colônia, Universidade Cabreira et al., 2011; Pretto et al., 2018), as well as the ornithosu-
Federal de Santa Maria, São João do Polêsine, Brazil; GPIT, chid Dynamosuchus collisensis (Müller et al., 2020b). Further-
Institut und Museum für Geologie und Paläontologie, Universi- more, rhynchosaur remains attributed to Hyperodapedon sp.
tat Tübingen, Tübingen, Germany; ISI, Geological Studies Unit were collected (Langer et al., 2007a), however, with the discov-
of the Indian Statistical Institute, Calcutta, India; MB, Museum ery of the taxon Teyumbaita sulcognathus (Montefeltro et al.,
für Naturkunde, Berlin, Germany; MCN, Museu de Ciências 2010) occurring in stratigraphic levels above Hyperodapedon in
Naturais, Fundaçãoo Zoobotânica do Rio Grande do Sul, several sites from both Brazil and Argentina (Desojo et al.,
Porto Alegre, Brazil; MCP, Museu de Ciências e Tecnologia Pon- 2020), the assignment of these remains to Hyperodapedon
tifícia Universidade Católica do Rio Grande do Sul, Porto should be considered carefully until they are properly reviewed.
Alegre, Brazil; MCZ, Museum of Comparative Zoology,
Harvard University, Cambridge, Massachusetts, U.S.A.; MLP,
Museo de La Plata, La Plata, Argentina; MNHN Muséum MATERIAL AND METHODS
National d’Histoire Naturelle, Paris, France; PVL, Instituto Material
Miguel Lillo, Universidad de Tucumán, San Miguel de
Tucumán, Argentina; PVSJ, Instituto y Museo de Ciencias Nat- The specimen here described is housed at the Centro de Apoio
urales, Universidad Nacional de San Juan, San Juan, Argentina; à Pesquisa Paleontológica da Quarta Colônia/Universidade
TMM, Vertebrate Paleontology Laboratory, University of Texas Federal de Santa Maria (CAPPA/UFSM), under specimen
at Austin, Austin, TX, U.S.A.; UFRGS, Paleovertebrate Collec- number CAPPA/UFSM 0276 (Fig. 1C−L). It was prepared
tion of the Laboratório de Paleovertebrados da Universidade using air scribe tools, scalpels, and needles. Finally, the specimen
Federal do Rio Grande do Sul, Porto Alegre, Brazil; UFSM, was stabilized with Paraloid B-72 (Koob, 1986).
Laboratório de Estratigrafia e Paleobiologia, Universidade
Federal de Santa Maria, Santa Maria, Brazil; ULBRA, Universi- Anatomical and Taxonomic Terminology
dade Luterana do Brasil, Coleção de paleovertebrados, Canoas,
Brazil; UMMP, University of Michigan Museum of Paleontology, This study employs traditional (Romerian) anatomical terms
Ann Arbor, MI, U.S.A. (i.e., ‘anterior’ and ‘posterior’). The nomenclature for vertebral
laminae follows Wilson (1999) and the nomenclature for ver-
tebral fossae follows Wilson et al. (2011).
GEOLOGICAL SETTING The phylogenetic definitions of clades discussed below follow
‘Várzea do Agudo’ also known as ‘Janner’ site is situated about Langer et al. (2019). Therefore, Bagualosauria comprises the
2 km west of the urban area of the Agudo municipality (29°39′ “minimal clade including Bagualosaurus agudoensis and Salta-
10.89″S, 53°17′34.20″W), Rio Grande do Sul, Brazil (Fig. 1A). saurus loricatus” (node-based definition; Langer et al., 2019)
It is exposed in ravines of reddish massive to laminated mud- and Saturnaliidae represents the “maximal sauropodomorph
stones, associated with yellowish cross-bedded sandstones and clade to encompass Saturnalia tupiniquim, but not Plateosaurus
sandstones with mudstone interclasts (Fig. 1B). The reddish engelhardti” (branch-based definition; Langer et al., 2019).
mudstones constitute most of the outcrop and are interpreted
as representative of a distal floodplain accumulation, whereas
Phylogenetic Analysis
the yellowish sandstones are mostly present in the top portion
of the ravines and represent a river channel (Da-Rosa, 2015; The phylogenetic affinities of CAPPA/UFSM 0276 were inves-
Pretto et al., 2018). The mudstones belong to the upper portion tigated with two data matrices. First, the specimen was scored in
of the Alemoa Member of the Santa Maria Formation, lower the data matrix of Müller (2021), which includes a wide range of
to middle Candelária Sequence, Paraná Basin (Zerfass et al., early dinosaurs and related forms. We excluded Nhandumirim
2003; Horn et al., 2014), whereas the overlaying sandstones are waldsangae given its immature ontogenetic status (Marsola
considered part of the Caturrita Formation, upper Candelária et al., 2018). The final data matrix includes 277 characters and
Sequence (Zerfass et al., 2003; Horn et al., 2014). 63 operational taxonomic units (OTUs). It was analyzed in the
Based on biostratigraphy, the site is considered Late Triassic software TNT v.1.1 (Goloboff et al., 2008) following the same par-
(Langer et al., 2007a, 2018). Its faunal content allows its assign- ameters as in the former study (i.e., Müller 2021). Therefore,
ment to the Hyperodapedon Assemblage Zone, comprising the Euparkeria capensis was used to root the most parsimonious
older Hyperodapedon Acme Zone and the younger Exaeretodon trees, which were recovered with a ‘Traditional search’ (random
sub-Assemblage Zone, in which the ‘Várzea do Agudo’ site is addition sequence + tree bisection reconnection) with 1000 repli-
situated (Langer et al., 2007a; Müller and Garcia, 2020; Schultz cates of Wagner trees (with random seed = 0) and using tree
et al., 2020). The distinction between both zones is based on bisection reconnection and branch swapping (holding 10 trees
the relative abundance of the index taxa in the corresponding save per replication). All characters were treated as having
strata, that is corroborated by biostratigraphic data from Argen- equal weight, and the following characters were treated as
tina (Martinez et al., 2013; Desojo et al., 2020). These biostrati- ordered: 4, 13, 18, 25, 63, 82, 84, 87, 89, 109, 142, 166, 174, 175,
graphic data suggest a latest Carnian age for the site, but 184, 186, 190, 201, 203, 205, 209, 212, 225, 235, 236, 239, 250,
radiometric data are lacking (Müller and Garcia, 2020). and 256. Topologies retained in overflowed replications were
The ‘Várzea do Agudo’ paleofauna is dominated by cynodonts branch-swapped for most parsimonious trees using tree bisection
with a few archosauromorph representatives (Müller et al., reconnection. The strict consensus tree was generated using all
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-3)

FIGURE 1. Location and geological context of the ‘Várzea do Agudo’ site (Agudo, Rio Grande do Sul, Brazil) and specimen CAPPA/UFSM 0276.
A, surface distribution of the geological units in the area (modified from Müller, 2021); B, general view of the site; C, reconstruction of the skeleton of
CAPPA/UFSM 0276 depicting the preserved elements; D, left ilium in lateral view; E, caudal series in left lateral view; F, right ulna in medial view;
G, indeterminate metacarpal in dorsal view; H, indeterminate manual phalanx in dorsal view; I, left femur in anterior view; J, left tibia in lateral view;
K, left fibula in lateral view; L, left metatarsal II in anterior view.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-4)

trees recovered in the analysis and all OTUs. Decay indices portions. Based on the absence of duplicated bones and their
(Bremer support values) were also obtained with TNT v. 1.1. relative size, the elements are believed to belong to a single indi-
For the second analysis, CAPPA/UFSM 0276 was scored in the vidual. Most bones are complete, yet some degree of sedimentary
data matrix of Müller et al. (2021). This data matrix was compression is observed (see description for additional com-
employed because of the large sample of sauropodomorphs. ments). Furthermore, most of the elements present a grayish
The final data matrix includes 404 characters and 78 OTUs. Fol- coloration with orange extremities, as is standard for materials
lowing McPhee et al. (2018), characters 199 and 278 were set as collected in the ‘Várzea do Agudo’ locality.
inactive (ambiguous scores and difficulty of interpretation). The Ontogenetic Assessment—The specimen appears to have
data matrix was subjected to an equally weighted parsimony reached skeletal maturity based on the absence of open neurocen-
analysis in TNT v.1.1 (Goloboff et al., 2008) following the same tral sutures (Brochu, 1996) and presence/shape of some muscle
parameters as in the former study (i.e., Müller et al., 2021). attachment structures (e.g., trochanteric shelf, dorsolateral tro-
Hence, Euparkeria capensis was used to root the most parsimo- chanter; Piechowski et al., 2014; Griffin and Nesbitt, 2016).
nious trees, which were recovered with a ‘Traditional search’
(random addition sequence + tree bisection reconnection) with
1000 replicates of Wagner trees (with random seed = 1), and DESCRIPTION
using tree bisection reconnection and branch swapping Axial Skeleton
(holding 10 trees save per replicate). All characters were
treated as having equal weight, and the following characters Caudal Vertebrae—The centra of caudal vertebrae (Figs. 1E, 2)
were treated as ordered: 8, 14, 22, 26, 43, 63, 75, 108, 118, 135, are spool-shaped as in other early dinosaurs. The centra of anterior
141, 151, 154, 167, 170, 172, 173, 180, 185, 190, 193, 200, 207, caudal vertebrae are anteroposteriorly short (Fig. 2C); i.e., the dor-
231, 234, 241, 251, 256, 264, 273, 282, 285, 299, 312, 334, 340, soventral height of the anterior articular facet is greater than the
348, 372, 385, 388, 391, 396, and 404. Topologies retained in over- anteroposterior length of the centrum (height:length ratio = 1.1;
flowed replications were branch-swapped for most parsimonious see Table S1 in Supplemental Data for measurements). The oppo-
trees using tree bisection reconnection. The strict consensus tree site condition is observed in the posterior caudal vertebrae, where
was generated using all trees recovered in the analysis and all the centrum is longer (height:length ratio of the posteriormost pre-
OTUs. Such as in the first analysis, decay indices (Bremer served centrum = 0.74). Anteroposteriorly compressed anterior
support values) were obtained with TNT v. 1.1. caudal vertebral centra is the typical condition in herrerasaurids,
for example as Herrerasaurus ischigualastensis (PVL 2566; Novas,
1993) and Gnathovorax cabreirai (CAPPA/UFSM 0009; Pacheco
Body Mass Estimation et al., 2019). However, this condition is not unique to herrerasaur-
The body mass (BM) of CAPPA/UFSM 0276 was estimated ids, being observed in some sauropodomorphs (e.g., Coloradi-
based on the femoral circumference (Cfem), according to the saurus brevis, PVL 5904; Eucnemesaurus entaxoni, BP/1/6234;
equation by Campione et al. (2014): log10BM = 2.754 × log10
(Cfem) − 0.683. We used this equation, which assumes that
CAPPA/UFSM 0276 was bipedal, because the specimen lacks
the following osteological indicators of quadrupedal stance, a
triradiate proximal ulna (i.e., deep radial fossa); straightening
of the femoral shaft; loss of the anterior trochanter (= lesser tro-
chanter); migration of the fourth trochanter distally and medi-
ally; and a robust fibula (McPhee et al., 2018).

SYSTEMATIC PALEONTOLOGY
DINOSAURIA Owen, 1842
SAURISCHIA Seeley, 1887
SAUROPODOMORPHA Huene, 1932
Gen. et sp. indet
(Figs. 1–9)

Material—Specimen CAPPA/UFSM 0276 (Fig. 1C–L), partial


postcranial skeleton including 15 caudal vertebrae, the right ulna,
one possible metacarpal, one manual phalanx, the partial left
ilium, the partial left hind limb (including femur, tibia, fibula,
and one partial metatarsal II).
Locality and Horizon—‘Várzea do Agudo’ or ‘Janner’ site (29°
39′10.89″S, 53°17′34.20″W), Agudo, Rio Grande do Sul, Brazil. It
belongs to the Santa Maria Formation, Candelária Sequence,
Paraná Basin (Zerfass et al., 2003; Horn et al., 2014; Fig. 1A).
The site is Carnian in age based on biostratigraphic data
(Langer et al., 2007a; Schultz et al., 2020).
Remarks—The bone elements were accumulated in a small
area of about 1 m2 on the same layer that yielded the holotype FIGURE 2. First caudal vertebra of CAPPA/UFSM 0276 in A, anterior,
B, ventral, and C, left lateral views. The hatched areas are matrix,
of Bagualosaurus agudoensis (Pretto et al., 2018), whereas the whereas the black areas are broken. Abbreviations: acdl, anterior centro-
holotype of Pampadromaeus barberenai comes from a lower diapophyseal lamina; c, centrum; dpr, depression; nc, neural canal; ns,
level (Müller et al., 2016). Most of the caudal series was in articu- neural spine; prcdf, prezygapophyseal centrodiapophyseal fossa; prdl,
lation (a segment with nine vertebrae), as was the hind limb. The prezygodiapophyseal lamina; pz, postzygapophysis; tp, transverse
remaining elements were found scattered around the articulated process.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-5)

Apaldetti et al., 2013; McPhee et al., 2015), whereas in the earliest processes (Fig. 3C). The olecranon process is short (Fig. 3F),
sauropodomorphs the centrum is proportionally longer (e.g., Bur- around 7% of the total length of the ulna. Some other dinosaurs
iolestes schultzi, ULBRA-PVT280; Cabreira et al., 2016). There is have a proximodistally enlarged olecranon process. For instance,
a fossa on the lateral surface of the centrum of the anterior the olecranon process is about 10.5% in Gnathovorax cabreirai
caudal vertebrae (Fig. 2C). It becomes shallow to almost impercep- (CAPPA/UFSM 0009; Pacheco et al., 2019), 10% in Pampadro-
tible in the subsequent elements. Unlike Macrocollum itaquii maeus barberenai (ULBRA-PVT016; Langer et al., 2019), and
(CAPPA/UFSM 0001b; Müller et al., 2018b), the ventral surface 23% (estimated) in Saturnalia tupiniquim (MCP 3844-PV;
of the centra lacks any longitudinal ridges and grooves along the Langer et al., 2007b). In contrast, the process is reduced in Bur-
entire preserved series. There are facets for chevron articulation iolestes schultzi (ULBRA-PVT280; Cabreira et al., 2016:7.8%)
in the anterior and posterior ventral corners of the preserved and in Macrocollum itaquii (CAPPA/UFSM 0001a; Müller
centra. The only exception is the probable first caudal vertebra, et al., 2018b:7.5%). Nevertheless, the shape and size of the ole-
which lacks facets for chevron articulation on the ventral margin cranon process presents strong intraspecific variation in some
(Fig. 2B). dinosaurs depending upon the ossification degree of cartilagi-
The neural arch is poorly preserved in all the caudal vertebral nous structures (e.g., Kentrosaurus aethiopicus, GPIT 1424;
elements. It is dorsoventrally shorter than the height of its MB.R.4800.33; Mallison, 2010).
respective centrum. The height of the neural canal is ∼0.3 The ulnar shaft is quite slender, such as in early sauropodo-
times the height of the anterior articular facet of the centrum morphs (e.g., Buriolestes schultzi, ULBRA-PVT280), whereas
in the first caudal vertebra (Fig. 2A). In addition, the neural in sauropodiforms it is usually robust (e.g., Mussaurus patagoni-
canal of the anterior caudal vertebrae is wider than deep. The cus, MLP 68-II-27-1; Pulanesaura eocollum, BP/1/6210; McPhee
prezygapophyses are poorly preserved in anterior elements. In and Choiniere, 2018). The shaft is featureless, whereas in Pampa-
the posteriormost preserved element, the prezygapophyses are dromaeus barberenai (ULBRA-PVT016; Langer et al., 2019)
anteriorly short, barely projecting beyond the anterior articular there is a suite of longitudinal crests and grooves. In lateral
facet of their respective centrum. Therefore, this condition differ- view (Fig. 3A), the anterior margin is concave, whereas the pos-
entiates the specimen from herrerasaurids with elongated prezy- terior margin is sigmoidal. The distal portion of the ulna is
gapophyses (e.g., Herrerasaurus ischigualastensis, PVSJ 53 and slightly rotated with respect to the proximal articular surface.
PVSJ 2566; Gnathovorax cabreirai, CAPPA/UFSM 0009; There is no evidence of a tubercle for the radial-ulnar ligament,
Novas, 1993; Pacheco et al., 2019). There is a prezygodiapophy- which is well developed in Antetonitrus ingenipes (BP/1/4952;
seal lamina on the first and second caudal vertebrae. This McPhee et al., 2014). The distal articular surface is transversely
lamina roofs a prezygapophyseal centrodiapophyseal fossa in convex. In distal view, it is oval (Fig. 3D).
these two vertebrae. Both the lamina and fossa are absent in Indeterminate Metacarpal—The specimen preserves one poss-
the subsequent elements. An anterior centrodiapophyseal ible metacarpal (Fig. 4A–D). However, some portions are
lamina runs from the anterior facet of the centrum to the broken away, which hampers a confident identification. It
ventral portion of the transverse process in the first caudal verte- might also be a phalanx 1, but it is identified as an indeterminate
bra. In the second caudal vertebra this lamina is poorly devel- metacarpal because of the large size, proportions (more than two
oped, whereas in the subsequent vertebrae it is absent. The times longer than transversely wide), presence of a flat proximal
transverse processes of the anterior vertebrae are poorly pre- articular surface (contrasting with the markedly concave surface
served. In the subsequent elements they are laterally oriented, of non-proximal phalanges), absence of a dorsal intercondylar
located in the posterior half of the neural arch, and dorsoven- process, and presence of a deep and wide extensor depression
trally flat. Finally, the transverse processes are reduced to faint on the dorsal surface of the distal portion. The indeterminate
projections in the posterior elements. metacarpal of CAPPA/UFSM 0276 is longer than wide (length:
The articular facet of the postzygapophyses faces ventrally in proximal width = 2.41). All metacarpals are generally longer
the anterior caudal vertebrae, whereas in the subsequent than wide in early sauropodomorphs (Sereno et al., 2013; Cab-
elements it faces ventrolaterally. The postzygapophyses lack reira et al., 2016). The metacarpal I of Eoraptor lunensis (PVSJ
any evidence of projections related to an accessory intervertebral 512; Sereno et al., 2013) is longer than wide (length:proximal
articulation. The neural spine is incomplete in the anterior width = 1.95; based on the left manus of PVSJ 512), whereas in
elements and poorly preserved along the series. In the posterior- some sauropodiforms it is wider than long (e.g., Antetonitrus
most preserved vertebrae, the neural spine is located on the pos- ingenipes, BP/1/4952; McPhee et al., 2014; Aardonyx celestae,
terior portion of the neural arch and is directed posterodorsally. BP/1/5379; Otero et al., 2015). However, the ratio of the metacar-
It is dorsoventrally short, hardly projecting beyond the dorsal pal I of CAPPA/UFSM 0276 differentiates it from the indetermi-
margin of the postzygapophyses. nate metacarpal of Eoraptor lunensis. The distal asymmetry
typical of metacarpal I in early sauropodomorphs is also not
observed here. Additionally, the ratio of the length of the poss-
Forelimb
ible metacarpal relative to the length of the femur (0.14 in
Ulna—The shape of the ulna was affected by sedimentary CAPPA/UFSM 0276) can be compared with that of Eoraptor
compression, resulting in a transversely compressed shaft lunensis. In the latter, the ratios are 0.09, 0.13, 0.14, and 0.10
(Fig. 3). The length of ulna is about 0.42 times that of tibia (see for metacarpals I, II, III, and IV, respectively. Hence the
Table S2 in Supplemental Data for measurements). The ulna element herein described probably belongs to the manual digit
bows medially along its shaft (Fig. 3A, E). The lateral surface II or III, assuming a similar proportion when compared with Eor-
is straight to slightly convex, whereas the medial surface is aptor lunensis (PVSJ 512; Sereno et al., 2013).
concave. The proximal articular surface is sub-triangular in prox- The proximal end is broader than deep and ovoid in proximal
imal view (Fig. 3C), distinct from the triradiate shape present in view (Fig. 4C), with a straight palmar margin and a convex dorsal
sauropodiforms (e.g., Aardonyx celestae, BP/1/5379c; Leduma- margin. Moreover, the proximal articular surface is straight to
hadi mafube, BP/1/7120; Yates et al., 2010; McPhee et al., gently concave. The dorsal surface of the shaft is transversely
2018). The anteromedial process is rounded and poorly devel- convex. As in the metacarpals of Eoraptor lunensis (PVSJ 512,
oped. In contrast to most massopodans (e.g., Mussaurus patago- Sereno et al., 2013) the shaft is straight in lateral view
nicus, MLP 68-II-27-1; Otero and Pol, 2013; Sefapanosaurus (Fig. 4B), whereas in Herrerasaurus ischigualastensis (PVSJ
zastronensis, BP/1/7437; Otero et al., 2015), there is no evidence 373; Sereno, 1993) the shafts are bowed ventrally. There is a
of a radial fossa between the anterolateral and anteromedial deep extensor depression on the dorsal surface of the distal
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-6)

FIGURE 3. Right ulna of CAPPA/UFSM 0276 in A, lateral, B, anterior, C, proximal, D, distal, E, posterior, and F, medial views. The black areas are
broken. Abbreviations: alp, anterolateral process; amp, anteromedial process; op, olecranon process.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-7)

FIGURE 4. Manus of CAPPA/UFSM 0276. Indeterminate metacarpal in A, dorsal, B, lateral or medial, C, proximal, and D, distal views; indetermi-
nate phalanx in E, dorsal, F, proximal, and G, distal views. The hatched areas are matrix, whereas the black areas are broken. Abbreviations: clp,
collateral pit; dc, distal condyle; ded, dorsal extensor depression; ig, intercondylar groove.

portion of the bone (Fig. 4A), which resembles the condition of intercondylar process (e.g., Martinez et al., 2011; Sereno et al.,
herrerasaurids Herrerasaurus ischigualastensis (PVSJ 373; 2013; Ballell et al., 2020), and the first phalanx of digits IV and
Sereno, 1993) and Gnathovorax cabreirai (CAPPA/UFSM V are reduced (e.g., Otero and Pol, 2013; Reiss and Mallison,
0035; Pacheco et al., 2019), whereas in Eoraptor lunensis (PVSJ 2014). Therefore, it is plausible that the isolated phalanx rep-
512, Sereno et al., 2013) the dorsal extensor depression is resents the first phalanx of the digit II or III.
shallow. The distal condyles are not entirely preserved
(Fig. 4D). Nevertheless, the collateral ligament pits are present.
Pelvic Girdle
Indeterminate Phalanx—There is only one isolated phalanx
preserved (Fig. 4E–G). It is longer than wide. The proximal Ilium—The preserved portion of the left ilium (Fig. 5)
articular surface is concave and slightly broader than deep resembles the general morphology of other coeval saurischians,
(Fig. 4F). The palmar margin of the proximal end is not sym- excepting herrerasaurids. There is a preacetabular fossa on the
metrical; one of the corners extends further proximally anterior portion of the bone (Fig. 5A), such as in Pampadro-
(Fig. 4E). The extensor depression on the dorsal surface of the maeus barberenai (ULBRA-PVT016; Langer et al., 2019) and
distal portion is poorly developed. The distal condyles are asym- Thecodontosaurus antiquus (BRSUG 23613; Ballell et al.,
metrical and separated by a shallow intercondylar groove 2020). Most of the iliac blade is missing, therefore its shape is
(Fig. 4D). Its position is uncertain. However, the absence of a uncertain. Posterior to the preacetabular fossa and dorsal to
dorsal intercondylar process on the proximal margin (Fig. 4E) the supracetabular crest there is a concavity on the ventral
suggests that it is a proximal phalanx. In early sauropodomorphs portion of the iliac blade. It corresponds to the attachment
the first phalanx of digit I usually bears a well-developed dorsal area of m. iliofemoralis (see Hutchinson, 2001; Langer, 2003)
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-8)

FIGURE 5. Left ilium of CAPPA/UFSM 0276 in A, lateral, B, posterior, C, dorsal, and D, medial views. The hatched areas are matrix, whereas the
black areas are broken. Abbreviations: bf, brevis fossa; bs, brevis shelf; ib, iliac blade; ip, ischial peduncle; mb, medial blade; mw, medial wall; pa, post-
acetabular ala; pf, preacetabular fossa; sac, supracetabular crest; sr1, first primordial sacral rib articular surface; sr2, second primordial sacral rib articu-
lar surface.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-9)

present in non-dinosaurian dinosauromorphs and plesiomorphic usually straight in anterior/posterior views in larger sauropodo-
in dinosaurs such as Saturnalia tupiniquim (MCP 3844-PV, MCP morphs (e.g., Coloradisaurus brevis, PVL 5904; Meroktenos tha-
3846-PV; Langer, 2003; Garcia et al., 2019), Herrerasaurus ischi- banensis, MNHN.F.LES16c; Apaldetti et al., 2013; Peyre de
gualastensis (PVL 2566; Novas, 1993), and Caseosaurus crosbyen- Fabrègues and Allain, 2016). The proximal articular surface is
sis (UMMP 8870; Baron and Williams, 2018). The supracetabular poorly preserved. Thus, the presence or absence of a proximal
crest is well developed (Fig. 5C), roofing the acetabulum and pro- groove is uncertain. The posteromedial and anteromedial tuber-
jecting ventrolaterally, differing from the laterally oriented crest cles are subequal in size (Fig. 6C). In Macrocollum itaquii
described for Bagualosaurus agudoensis (UFRGS-PV-1099-T; (CAPPA/UFSM 0001b; Müller et al., 2018b), the posteromedial
Pretto et al., 2018). The posterior portion of the crest does not tubercle is the largest one, whereas in the herrerasaurid Gnatho-
merge with the anterior portion of the brevis shelf. The medial vorax cabreirai (CAPPA/UFSM 0009; Pacheco et al., 2019), the
wall of the acetabulum is extended and, as preserved, it bears a posteromedial tubercle is absent.
slightly concave ventral margin (Fig. 5D), which is the condition The femoral head is expanded and bears a marked concave
observed in similarly aged sauropodomorphs such as Buriolestes surface where it merges with the femoral shaft (Fig. 6E). This
schultzi (ULBRA-PVT280; Cabreira et al., 2016), Saturnalia concave surface is limited to a restricted area on the posterome-
tupiniquim (MCP 3844-PV, MCP 3846-PV; Garcia et al., 2019) dial aspect of the femur. The dorsolateral trochanter is low and
and Pampadromaeus barberenai (ULBRA-PVT016; Langer rounded (Fig. 6A, B), like that of Saturnalia tupiniquim (MCP
et al., 2019). It differentiates the specimen from herrerasaurids 3844-PV; Langer, 2003), Bagualosaurus agudoensis (UFRGS-
and more advanced sauropodomorphs, where the acetabulum PV-1099-T; Pretto et al., 2018), and Pampadromaeus barberenai
is partially to completely perforated (Novas, 1993; Müller et al., (ULBRA-PVT016; Langer et al., 2019). A scar apparently
2018b; Pacheco et al., 2019; Ballell et al., 2020). related to ontogeny (see Griffin and Nesbitt, 2016) and associ-
The postacetabular ala is not completely preserved. However, ated with the dorsolateral trochanter in some specimens of Bur-
it seems elongated, distinct from the short ala of some saur- iolestes schultzi (e.g., CAPPA/UFSM 0035; Müller et al., 2018a)
ischians (Caseosaurus crosbyensis, UMMP 8870; Herrerasaurus and Gnathovorax cabreirai (CAPPA/UFSM 0009; Pacheco
ischigualastensis, PVL 2566; Baron and Williams, 2018; Novas, et al., 2019) is absent in CAPPA/UFSM 0276. The anterior tro-
1993). The ventral margin of the brevis shelf is deep and covers chanter is small and lacks a cleft between the proximal tip and
the medial blade of the postacetabular ala in lateral view the femoral shaft (Fig. 6F), contrasting with the basally branching
(Fig. 5A), such as in Panphagia protos (PVSJ 874; Martinez theropod Erythrovenator jacuiensis (CAPPA/UFSM 0157;
and Alcober, 2009). Conversely, in Bagualosaurus agudoensis Müller, 2021). There is a trochanteric shelf associated with the
(UFRGS-PV-1099-T; Pretto et al., 2018) and Thecodontosaurus anterior trochanter (Fig. 6B). The shelf occurs in several early
antiquus (BRSUG 23613; Ballell et al., 2020), the medial saurischians (e.g., Gnathovorax cabreirai, CAPPA/UFSM 0009;
lamina is deeper than the brevis shelf. The brevis fossa excavates Pacheco et al., 2019; Buriolestes schultzi, ULBRA-PVT280; Cab-
the ventral portion of the postacetabular ala forming a concave reira et al., 2016); however, it is absent in post-Carnian sauropo-
surface (Fig. 5B). Whereas the fossa is ventrally oriented in the domorphs (e.g., Macrocollum itaquii, CAPPA/UFSM 0001b;
specimen, in Bagualosaurus agudoensis (UFRGS-PV-1099-T; Müller et al., 2018b; Thecodontosaurus antiquus, BRSUG
Pretto et al., 2018) it is ventrolaterally oriented. Both the 23615; Ballell et al., 2020). The fourth trochanter is crest-like
brevis shelf and fossa are considered attachment areas for the and slightly asymmetrical. If differs from the strongly symmetri-
m. caudofemoralis brevis as in other dinosaurs (Hutchinson, cal fourth trochanter of theropods (Langer and Benton, 2006).
2001). Unlike in Buriolestes schultzi (ULBRA-PVT280; Cabreira Moreover, the fourth trochanter is located at the posteromedial
et al., 2016), Saturnalia tupiniquim (MCP 3844-PV; Garcia et al., surface and in the proximal half of the femoral shaft, whereas in
2019) and Chromogisaurus novasi (PVSJ 845; Ezcurra, 2010), sauropodiforms it is located more distally (e.g., Meroktenos tha-
there is no evidence of any raised scar (i.e., rugosity) on the banensis, MNHN.F.LES16c; Peyre de Fabrègues and Allain,
lateral surface of the preserved portion of the postacetabular 2016; Antetonitrus ingenipes, BP/1/4952; McPhee et al., 2014).
ala. Nevertheless, the postacetabular ala and the iliac blade are The distal portion of the femur is transversely expanded
not entirely preserved, hence, the absence of scars is ambiguous. (Fig. 6D). Unlike theropods (Pinheiro, 2016) and several post-
The ischial peduncle extends posteroventrally, being ovoid in Carnian sauropodomorphs (e.g., Coloradisaurus brevis, PVL
cross-section, and its convex anterior surface merges anteriorly 5904; Apaldetti et al., 2013), the anterior surface of the distal
with the acetabulum wall. This region corresponds to the antitro- portion of the bone lacks an extensor groove/depression. On
chanter and does not bear the raised process seen in Guaiba- the opposite side, the popliteal fossa extends longitudinally and
saurus candelariensis (MCN PV2355, UFRGS PV0725T; separates the distal condyles (Fig. 6E). The fossa is much
Langer et al., 2011) and Ixalerpeton polesinensis (ULBRA- shorter proximodistally than that of silesaurids and aphanosaurs
PVT059; Cabreira et al., 2016). Unlike Coloradisaurus brevis (Nesbitt et al., 2010, 2017). Conversely, the fossa is transversely
(PVL 5904; Apaldetti et al., 2013) and Antetonitrus ingenipes wide, such as in Pampadromaeus barberenai (ULBRA-
(NM QR1545; McPhee et al., 2014), the specimen lacks a pos- PVT016; Langer et al., 2019), whereas it is usually narrow in saur-
terior projection (‘heel’) at the distal end of the ischial peduncle. opodomorphs. The lateral and medial condyles are subequal in
Likewise, there is no evidence of a small foramen (or pair of for- size, whereas the crista tibiofibularis is slightly smaller than the
amina) between the supracetabular crest and the brevis shelf condyles. This condition differs from the poorly defined crista
(Garcia et al., 2019). Medially, the bone preserves the scars tibiofibularis of Staurikosaurus pricei (MCZ 1669; Bittencourt
from the attachment of the two primordial sacral vertebrae and Kellner, 2009).
(Fig. 5D); however, the total number of sacral vertebrae is Tibia—The tibia (Fig. 7) is slightly longer than the femur (see
uncertain. Table S3 in Supplemental Data for measurements), such as in
other early sauropodomorphs. The reverse condition (i.e.,
femur longer than the tibia) occurs in most herrerasaurids and
Hind Limb
post-Carnian sauropodomorphs, (e.g., Gnathovorax cabreirai,
Femur—The femur (Fig. 6) is mainly straight in lateral/medial CAPPA/UFSM 0009; Pacheco et al., 2019; Macrocollum itaquii,
and markedly sigmoid in anterior/posterior views, such as in CAPPA/UFSM 0001b; Müller et al., 2018b). The tibia is straight
other early sauropodomorphs (e.g., Saturnalia tupiniquim, (Fig. 7A, B) with an anteroposteriorly expanded proximal
MCP 3844-PV; Buriolestes schultzi, ULBRA-PVT280; Langer, portion. In proximal view (Fig. 7C), the bone is subtriangular
2003; Cabreira et al., 2016). On the other hand, the femora are and the cnemial crest is directed anteriorly. The latter extends
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-10)

FIGURE 6. Left femur of CAPPA/UFSM 0276 in A, lateral, B, anterior, C, proximal, D, distal, E, posterior, and F, medial views. The black areas are
broken. Abbreviations: 4t, fourth trochanter; alt, anterolateral tubercle; amt, anteromedial tubercle; at, anterior trochanter; ctf, crista tibiofibularis;
dlt, dorsolateral trochanter; faa, facies articularis antitrochanterica; lc, lateral condyle; mc, medial condyle; pf, popliteal fossa; pmt, posteromedial
tubercle; ts, trochanteric shelf.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-11)

FIGURE 7. Left tibia of CAPPA/UFSM 0276 in A, lateral, B, anterior, C, proximal, D, distal, E, posterior, and F, medial views. The black areas are
broken. Abbreviations: cc, cnemial crest; fta, foramen for the nutritive tibial artery; it, insisura tibialis; lc, lateral condyle; ln, lateral notch; mc, medial
condyle; plf, posterolateral flange; rdg, ridge.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-12)

for approximately one fourth of the tibial length. There is a prox- schultzi (CAPPA/UFSM 0035; Müller et al., 2018a), Pampadro-
imodistal sulcus (= insisura tibialis sensu Butler, 2010) posterior maeus barberenai (ULBRA-PVT016; Langer et al., 2019), and
to the cnemial crest in the lateral surface of the proximal region Gnathovorax cabreirai (CAPPA/UFSM 0009; Pacheco et al.,
of the bone. Immediately posterior to the sulcus, there is a faint 2019). Also, in the proximal third of the fibula, there is no evi-
ridge, which differs from the well-developed fibular crest of dence of a longitudinally oriented medial sulcus, which is
neotheropods (e.g., Powellvenator podocitus, PVL 4414; present in Staurikosaurus pricei (MCZ 1669, Bittencourt and
Ezcurra, 2017). The medial surface is gently concave. On the Kellner, 2009) and Gnathovorax cabreirai (Garcia et al., 2021).
proximal articular surface, the medial condyle is set posterior The midshaft is ‘D’-shaped in cross-section, where the lateral
to the lateral condyle, whereas in Jaklapallisaurus asymmetrica margin is convex and the medial is flat. The transverse width of
(ISI R274; Novas et al., 2011) and Staurikosaurus pricei (MCZ the midshaft is 0.56 times the width of the tibia. In lateral view,
1669; Bittencourt and Kellner, 2009) both condyles lie near the the long axis of the shaft is not entirely straight, and the distal
posterior margin. Distinct from Eoraptor lunensis (PVSJ 559; third projects anteriorly (Fig. 8A). This condition resembles
Sereno et al., 2013), the specimen lacks a notch between the pos- that of Pampadromaeus barberenai (ULBRA-PVT016; Langer
terior portion of the condyles, and both condyles are poorly et al., 2019); however, it is probably related to taphonomic defor-
delimited. There is no projection from the medial condyle, as mation in both specimens (Langer et al., 2019). The distal portion
in Buriolestes schultzi (ULBRA-PVT280; Cabreira et al., 2016). of the bone expands anteroposteriorly, resulting in an oval extre-
The midshaft is ovoid in cross-section. The foramen for the mity in distal view (Fig. 8D). The anterior and posterior margins
nutritive tibial artery lies on the proximal third of the lateral of the distal portion are sharp, forming a ridge on each side. In
portion of the bone (Fig. 7A). The distal end of the tibia is sub- lateral/medial view, the posterior corner of the distal end
quadrangular in distal view (Fig. 7D), resembling the condition extends more distally than the anterior corner. Furthermore,
of other early sauropodomorphs (e.g., Bagualosaurus agudoensis, the distal margin is convex, whereas it is sigmoid in the theropod
UFRGS-PV-1099-T; Pretto et al., 2018; Macrocollum itaquii, Lepidus praecisio (TMM41936-1.3; Nesbitt and Ezcurra, 2015).
CAPPA/UFSM 0001c; Müller et al., 2018b) and some herrera- In contrast to Sefapanosaurus zastronensis (BP/1/7447; Otero
saurids (e.g., Herrerasaurus ischigualastensis, PVSJ 373; Novas, et al., 2015), the distal portion lacks any evidence of an enlarged
1993). In contrast, some other herrerasaurids have a rounded/ anteromedial projection.
ovoid outline in distal view (Garcia et al., 2021). The lateral Metatarsal II—The element preserves its proximal portion and
notch is longitudinally short. The posterior margin of the distal part of the shaft (Fig. 9). The general anatomy resembles that of
end is deeper than the anterior margin. In addition, the posterior the metatarsal II of other early sauropodomorphs, such as Satur-
margin is straight in distal view, whereas in Panphagia protos nalia tupiniquim (MCP 3844-PV; Langer, 2003) and Macrocol-
(PVSJ 874; Martinez and Alcober, 2009) it is concave. In contrast lum itaquii (CAPPA/UFSM 0001a; Müller et al., 2018b), where
with Macrocollum itaquii (CAPPA/UFSM 0001c; Müller et al., the proximal articular surface is flat, dorsoplantarly expanded
2018b) and Panphagia protos (PVSJ 874; Martinez and (Fig. 9B, D), and subrectangular in proximal view (Fig. 9C).
Alcober, 2009), the posterolateral flange of the new specimen This subrectangular surface differentiates the specimen from
does not exceed the lateral margin of the bone in distal view. some sauropodomorphs with an hourglass-shaped proximal
This condition is usually seen in herrerasaurids (e.g., Herrera- surface (e.g., Coloradisaurus brevis, PVL 5904; Apaldetti et al.,
saurus ischigualastensis, PVSJ 373; Novas, 1993; Staurikosaurus 2013). Moreover, the transversely compressed condition of the
pricei, MCZ 1669; Bittencourt and Kellner, 2009). The postero- proximal articular surface (Fig. 9A) differs from the transversely
lateral flange projects beyond the distal limit of the astragalar expanded proximal articular surface of Pampadromaeus barber-
articular surface, a trait common in post-Carnian sauropodo- enai (ULBRA-PVT016; Langer et al., 2019). However, the latter
morphs (Langer, 2003; Garcia et al., 2021), but also present inci- was strongly affected by sedimentary compression, therefore, its
piently in Carnian forms such as Eoraptor lunensis (PVSJ 512; anatomy should be carefully considered. The lateral margin of
Sereno et al., 2013) and Saturnalia tupiniquim (MCP 3844-PV; the proximal surface is concave, whereas in early sauropodo-
Langer, 2003). There is no evidence of a posteromedial notch morphs it is usually straight (e.g., Saturnalia tupiniquim, MCP
(Fig. 7E, F), which occurs in Guaibasaurus candelariensis 3844-PV; Langer, 2003). The medial margin is straight to slightly
(MCN-PV2356; Langer et al., 2011) and Bagualosaurus agudoen- sigmoid, such as in Saturnalia tupiniquim (MCP 3844-PV;
sis (UFRGS-PV-1099-T; Pretto et al., 2018). Langer, 2003) and Macrocollum itaquii (CAPPA/UFSM 0001a;
Fibula—The fibula (Fig. 8) lacks part of the proximal surface. Müller et al., 2018b), whereas in Unaysaurus tolentinoi
The bone is slender and expands in the extremities (Fig. 8A). In (UFSM11069; Leal et al., 2004) and Thecodontosaurus antiquus
proximal view (Fig. 8C), the lateral margin of the proximal (BRSUG 26627; Ballell et al., 2020) it is concave. There are no
articular surface is convex, whereas the medial margin is straight. ventral flanges projecting from the proximal articular surface.
The latter margin is concave in Mussaurus patagonicus (MLP 68- The shaft is slender and transversely compressed (Fig. 9E).
II-27-1; Otero and Pol, 2013) and Powellvenator podocitus (PVL
4414-4; Ezcurra, 2017). The proximal portion of the bone
expands anteroposteriorly and is transversely compressed. RESULTS
There is a gently sloping ridge on the medial surface of the prox-
Phylogenetic Analysis
imal portion (Fig. 8F), similar to that of Saturnalia tupiniquim
(MCP 3844-PV; Langer, 2003) and Buriolestes schultzi The first analysis (i.e., employing the data matrix of Müller,
(CAPPA/UFSM 0035; Müller et al., 2018a). In the proximal 2021) recovered 72 most parsimonious trees (MPTs) of 981
third of the bone, there is a proximodistally oriented ridge steps (consistency index = 0.321; retention index = 0.668). Near
(∼20 mm in length, see Fig. 8A, E) for the iliofibularis muscle the base of the tree, Saturnaliidae is composed of Chromogi-
insertion (Langer, 2003), such as in Buriolestes schultzi saurus novasi and Saturnalia tupiniquim. The clade is supported
(CAPPA/UFSM 0035; Müller et al., 2018a), Chromogisaurus by an enlarged olecranon process of ulna (149 [0→1]) and the
novasi (PVSJ 845; Ezcurra, 2010), and Bagualosaurus agudoensis pubic peduncle of the ilium forming an angle of less than 45
(UFRGS-PV-1099-T; Pretto et al., 2018). This ridge is absent in degrees to the long axis of the iliac blade (186 [1→2]).
Macrocollum itaquii (CAPPA/UFSM 0001b; Müller et al., CAPPA/UFSM 0276 nests within Bagualosauria in all MPTs
2018b). The specimen lacks any evidence of a tubercle on the (Fig. 10A). This clade is supported by three unambiguous syna-
anteromedial margin of the shaft (Fig. 8B). A protruding tuber- pomorphies: (1) ratio between the dorsoventral height and the
cle occurs in the proximal third of the fibula of Buriolestes anteroposterior length of the dentary is greater than 0.2 (73
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-13)

FIGURE 8. Left fibula of CAPPA/UFSM 0276 in A, lateral, B, anterior, C, proximal, D, distal, E, posterior, and F, medial views. The black areas are
broken. Abbreviations: ifi, m. iliofibularis insertion; rdg, ridge.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-14)

FIGURE 9. Left metatarsal II of CAPPA/UFSM 0276 in A, anterior, B, lateral, C, proximal, D, medial, and E, posterior views. The black areas are
broken. Abbreviations: mt I cs, metatarsal I contact surface; mt III cs, metatarsal III contact surface.

[0→1]); (2) short caudoventral process of dentary (75 [0→1]); the tibia (242 [1→0]). The node joining CAPPA/UFSM 0276
and (3) serrations of maxillary/dentary teeth forming oblique and Bagualosaurus agudoensis is the sister taxon of a clade
angles with the margin of the tooth (88 [1→2]). Given the composed of post-Carnian sauropodomorphs. The remaining
absence of cranial elements, none of these features are pre- topology of the strict consensus tree is equivalent to that of
served in CAPPA/UFSM 0276. The specimen is positioned as Müller (2021). The Bremer support value is 1 for all the
the sister taxon of Bagualosaurus agudoensis, and the clade nodes within Sauropodomorpha. CAPPA/UFSM 0276 nests
they form is supported by two unambiguous synapomorphies: basal to Bagualosauria with one extra step. A possible sister-
(1) vertically oriented ischial peduncle of the ilium (179 taxon relationship between CAPPA/UFSM 0276 and Pampa-
[1→0]); and (2) absence of a distinct proximodistally oriented dromaeus barberenai requires at least one extra step, as does
ridge on the posteromedial surface of the distal portion of placement of CAPPA/UFSM 0276 within Saturnaliidae.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-15)

FIGURE 10. Result of the phylogenetic analyses. A, abbreviated strict consensus tree of the first analysis (employing the data matrix of Müller, 2021)
depicting the phylogenetic position of CAPPA/UFSM 0276; B, abbreviated strict consensus tree of the second analysis (employing the data matrix of
Müller et al., 2021) depicting the phylogenetic position of CAPPA/UFSM 0276. Circles and black arrows indicate node- and branch-based clade
names, respectively.

Placement within Theropoda or Herrerasauridae requires at Bagualosauria nests as the sister taxon of an unresolved Saturna-
least five extra steps. liidae. CAPPA/UFSM 0276 nests as the sister taxon of Bagual-
The second heuristic search (i.e., employing the data matrix of osauria in all MPTs (Fig. 10B), and this node preceding
Müller et al., 2021) recovered 24 MPTs of 1555 steps (consistency Bagualosauria is supported by two unambiguous synapomor-
index = 0.304; retention index = 0.673). Saturnaliidae, which phies: (1) concave lateral margin of the proximal surface of the
includes Saturnalia tupiniquim, Chromogisaurus novasi, Pampa- metatarsal II (377 [0→1]); and (2) femoral length between 200
dromaeus barberenai, and Panphagia protos, is supported by and 399 mm (404 [0→1]). Bagualosauria is supported by a ventro-
three synapomorphies: (1) a bone sheet between the anterior laterally oriented brevis fossa (287 [1→2]). The Bremer support
and ventral processes of the prefrontal (47 [0→1]); (2) posterior value is 1 for all the nodes within Sauropodomorpha that are
margin of the acromion process rises from the scapular blade at basal to Plateosauria. A possible sister-taxon relationship
an angle that is greater than 65° from the long axis of the scapula, between CAPPA/UFSM 0276 and Bagualosaurus agudoensis
at its steepest point (226 [0→1]); and (3) a greatly enlarged ole- requires one extra step, whereas a sister taxon relationship
cranon process of the ulna (243 [0→2]). In the strict consensus between CAPPA/UFSM 0276 and Pampadromaeus barberenai
tree, the node supporting CAPPA/UFSM 0276 plus requires two extra steps. Placement within Saturnaliidae or
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-16)

FIGURE 11. Chrono- and biostratigraphy of the Candelária Sequence from southern Brazil, showing the pelvic girdle and hind limb of sauropodo-
morphs from each Assemblage Zone (AZ). A, Macrocollum itaquii (CAPPA/UFSM 0001b); B, indeterminate sauropodomorph (CAPPA/UFSM
0276; pes from the holotype of Bagualosaurus agudoensis); C, Buriolestes schultzi (ULBRA-PVT280; CAPPA/UFSM 0035). The radiometric ages
of 233.2 and 225.4 Ma are according to Langer et al. (2018). Silhouettes are not to scale.

Theropoda requires at least two extra steps. Placement within DISCUSSION


Herrerasauridae requires at least three extra steps.
Taxonomic Status
CAPPA/UFSM 0276 adds to the paleofauna of the ‘Várzea do
Agudo’ site and expands the knowledge of uncommon portions
of the skeleton of Carnian dinosaurs, such as the forelimb.
Prior to CAPPA/UFSM 0276, only Eoraptor lunensis (Sereno
Body Mass Estimation
et al., 2013) and Buriolestes schultzi (Cabreira et al., 2016) pre-
Following the equation of Campione et al. (2014), which is served manual elements among Carnian sauropodomorphs, the
based on the relation between body mass and femoral circumfer- latter being limited to a single metacarpal. It also expands the
ence ( = 66 mm in CAPPA/UFSM 0276), the resulting estimated record of relatively large sauropodomorphs (femoral length >
body mass for the specimen described here is 21 kg, with a range 200 mm) from the late Carnian. It is the second Carnian-aged
of 15–26 kg applying a 25% prediction error (Campione et al., specimen known to date, the first being the holotype of Bagual-
2014). osaurus agudoensis (Pretto et al., 2018; Table S4). Another
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-17)

potential large specimen was recently reassessed and is now con- in CAPPA/UFSM 0276. The relationship between the absence
sidered a herrerasaurid instead of a sauropodomorph (Garcia of the trochanteric shelf and the body size increase in sauropodo-
et al., 2021). This raises some relevant taxonomic questions morphs is clear; i.e., there are no large sauropodomorphs with
necessary for a better understanding of the diversity of this ‘tran- trochanteric shelf. However, as observed in CAPPA/UFSM
sitional’ paleofauna of the ‘Várzea do Agudo’ site. 0276, the loss of this feature did not occur.
The new specimen comes from the same fossiliferous site that Another signature of increased body size that might be
yielded two other early sauropodomorphs: Pampadromaeus bar- expected to occur in the hind limb of CAPPA/UFSM 0276 is
berenai and Bagualosaurus agudoensis. CAPPA/UFSM 0276 may an increase in limb robustness, which is observed in sauropodo-
be referrable to one of these taxa, or pertain to a new taxon. Ana- morphs from younger Upper Triassic beds (Ezcurra and Apal-
tomical and size differences exist among CAPPA/UFSM 0276, detti, 2012). However, the Robustness Index (RI, sensu Wilson
Pampadromaeus, and Bagualosaurus, but taphonomic distortion and Upchurch, 2003; i.e., average of the greatest widths of the
and ontogenetic uncertainty surrounding the specimens of proximal end, midshaft, and distal end of the element divided
concern mean that more material and histological analyses are by the length of the element) obtained for CAPPA/UFSM 0276
necessary to further refine the taxonomic affinities of the new does not differ from those of smaller sauropodomorphs (Table
skeleton. S6). Both the tibia and fibula are quite slender with indices com-
parable to those of Eoraptor lunensis. Even the indices obtained
for Macrocollum itaquii are not distinct from those of smaller
The Early Increase in Size in Sauropodomorpha
forms. Conversely, the RI of the tibia of PULR 136, a large
The holotype of Bagualosaurus agudoensis is particularly (tibia length = 40.5 cm) indeterminate sauropodomorph from
interesting because it is substantially larger than coeval or puta- upper Norian–Rhaetian beds of Argentina (Ezcurra and Apal-
tive older sauropodomorphs. Its estimated femoral length is 22.4 detti, 2012), is 0.33, which is roughly twice as great as the RI
cm (Pretto et al., 2018), similar to the femoral length of CAPPA/ for any Carnian–early Norian sauropodomorph. Hence, body
UFSM 0276 (20.4 cm). Indeed, the femoral lengths of both speci- size and hind limb robustness do not appear to have increased
mens are 40 and 30% (respectively) larger than the next-largest in a coordinated fashion in the early evolutionary history of saur-
sauropodomorphs of similar age (excluding skeletally immature opodomorphs; only at larger body size thresholds is the hindlimb
specimens; see Table S4), though some of those may not have robustness of early sauropodomorphs markedly increased above
died at their asymptotic size. Therefore, both represent the plesiomorphic values. The gracile construction of the pes (pha-
oldest evidence of body size increase within the sauropodomorph langes much longer than wide) of both Bagualosaurus agudoensis
lineage. (Pretto et al., 2018) and Macrocollum itaquii (Müller et al.,
The earliest branching sauropodomorph Buriolestes schultzi 2018b) reinforces this point (Fig. 11). In larger sauropodo-
(Cabreira et al., 2016) is estimated to weigh ∼7 kg based on the morphs, such as Eucnemesaurus entaxonis (BP/1/6234; McPhee
femoral circumference (Müller et al., 2021). Applying the same et al., 2015) and Coloradisaurus brevis (PVL 5904; Apaldetti
approach, the estimated mass of CAPPA/UFSM 0276 is three et al., 2013), the phalanges are stouter.
times heavier at ∼21 kg. Below, we explore how this mass
increase did or did not affect various aspects of skeletal
CONCLUSIONS
anatomy within Sauropodomorpha.
The hindlimb zeugopodium/stylopodium length ratio has been Despite its indeterminate taxonomic status, the new early saur-
traditionally employed in phylogenetic studies of early dinosaurs opodomorph specimen described herein (CAPPA/UFSM 0276)
(e.g., Gauthier, 1986; Nesbitt, 2011; Cabreira et al., 2016; Baron expands the fossil record between the mid-to-late Carnian, fauni-
et al., 2017). The typical condition in the earliest sauropodo- vorous, small-bodied, and short-necked members of the clade
morphs is to have hindlimb stylopodia and zeugopodia similar (e.g., Buriolestes schultzi) and the early Norian sauropodo-
in length (Fig. 11C; Table S6). Conversely, more derived sauropo- morphs with herbivorous/omnivorous adaptations, relatively
domorphs usually have hindlimb zeugopodia shorter than stylo- large body sizes, and elongated necks (e.g., Macrocollum
podia (Fig. 11A). Curiously, despite its relatively larger size, itaquii). The new specimen represents one of the oldest examples
CAPPA/UFSM 0276 retains the plesiomorphic condition, of body increase size within Sauropodomorpha. Whereas the
where the hindlimb zeugopodium is subequal in length to the sty- skeleton of CAPPA/UFSM 0276 is significantly larger than that
lopodium (Fig. 11B; Table S6). The ratio is unknown in Bagual- of nearly all coeval or older forms, its proportions (hindlimb zeu-
osaurus. In large herrerasaurids, the hindlimb stylopodium is gopodium subequal to stylopodium in length; low robustness
longer than the zeugopodium (Novas, 1993; Pacheco et al., indices) and anatomy (presence of a trochanteric shelf on the
2019; Table S6), whereas in the smaller herrerasaurid Stauriko- femur) resemble that of these earliest forms. Hence, despite its
saurus pricei (MCZ 1669), the hindlimb stylopodium is shorter proportionally large body size, the new specimen retains a con-
than the zeugopodium. The latter condition is frequently servative postcranial anatomy. Alternatively, CAPPA/UFSM
regarded as indicative of cursoriality (Persons and Currie, 2016). 0276 may exist at the large extreme of its ontogenetic trajectory,
The presence or absence of a trochanteric shelf on the femur is whereas other known specimens of coeval sauropodomorphs
another trait that has commonly been employed in phylogenetic may represent not fully grown individuals. Such a scenario
studies of early dinosaurs (e.g., Gauthier, 1986; Nesbitt, 2011; would indicate a larger ancestral body size for Sauropodomorpha
Cabreira et al., 2016). This structure is usually regarded as the than is presently hypothesized.
insertion point of m. iliofemoralis externus (Hutchinson, 2001;
Langer, 2003; Piechowski and Tałanda, 2020) and occurs in
several non-dinosaur dinosauriforms and saurischian dinosaurs
ACKNOWLEDGMENTS
(Nesbitt, 2011). In Saurischia, it is present in some herrerasaurids
(Novas, 1993; Alcober and Martinez, 2010; Pacheco et al., 2019), We thank C. Peyre de Fabrègues and an anonymous reviewer
some theropods (Griffin, 2018), and sauropodomorphs from the for corrections and comments that greatly improved this manu-
Carnian (Langer, 2003; Ezcurra, 2010; Cabreira et al., 2016). script. We also extend our gratitude to the editor
Conversely, the trochanteric shelf is absent in post-Carnian saur- M. D. D’Emic for corrections and suggestions. We thank the
opodomorphs, such as Macrocollum itaquii (Fig. 11) and Theco- Coordenação de Aperfeiçoamento de Pessoal de Nível Superior
dontosaurus antiquus (Müller et al., 2018b; Ballell et al., 2020). (CAPES) scholarship for M.S.G. We also thank the Willi
Therefore, its presence is regarded as a plesiomorphic feature Henning Society, for the gratuity of TNT software. This work
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-18)

was carried out with the aid of the Fundação de Amparo à Pes- Desojo, J. B., L. E. Fiorelli, M. D. Ezcurra, A. G. Martinelli, J. Ramezani,
quisa do Estado do Rio Grande do Sul (FAPERGS 21/2551- Á. A. S. Da-Rosa, M. B. V. Baczko, M. J. Trotteyn, F. C. Montefeltro,
0000680-3) to R. T. Müller. M. Ezpeleta, and M. C. Langer. 2020. The Late Triassic
Ischigualasto Formation at Cerro Las Lajas (La Rioja,
Argentina): fossil tetrapods, high-resolution chronostratigraphy,
and faunal correlations. Scientific Reports 10:12782.
Ezcurra, M. D. 2010. A new early dinosaur (Saurischia:
LITERATURE CITED Sauropodomorpha) from the Late Triassic of Argentina: a reassess-
Alcober, O. A., and R. N. Martinez. 2010. A new herrerasaurid ment of dinosaur origin and phylogeny. Journal of Systematic
(Dinosauria, Saurischia) from the Upper Triassic Ischigualasto Palaeontology 8:371–425.
Formation of northwestern Argentina. ZooKeys 63:55–81. Ezcurra, M. D. 2017. A new early coelophysoid neotheropod from the
Apaldetti, C., D. Pol, and A. Yates. 2013. The postcranial anatomy of Late Triassic of northwestern Argentina. Ameghiniana 54:506–538.
Coloradisaurus brevis (Dinosauria: Sauropodomorpha) from the Ezcurra, M. D., and C. Apaldetti. 2012. A robust sauropodomorph speci-
late Triassic of Argentina and its phylogenetic implications. men from the Upper Triassic of Argentina and insights on the diver-
Palaeontology 56:277–301. sity of the Los Colorados Formation. Proceedings of the Geologists’
Apaldetti, C., R. N. Martínez, I. A. Cerda, D. Pol, and O. Alcober. 2018. Association 123:155–164.
An early trend towards gigantism in Triassic sauropodomorph dino- Garcia, M. S., F. A. Pretto, S. Dias-da-Silva, and R. T. Müller. 2019. A
saurs. Nature Ecology and Evolution 2:e1227. dinosaur ilium from the Late Triassic of Brazil with comments on
Ballell, A., E. J. Rayfield, and M. J. Benton. 2020. Osteological redescrip- key-character supporting Saturnaliinae. Anais da Academia
tion of the Late Triassic sauropodomorph dinosaur Brasileira de Ciências 91:e20180614.
Thecodontosaurus antiquus based on new material from Garcia, M. S., R. T. Müller, F. A., Pretto, Á. A. S. Da-Rosa, and S. Dias-
Tytherington, southwestern England. Journal of Vertebrate Da-Silva. 2021. Taxonomic and phylogenetic reassessment of a
Paleontology 40:e1770774. large-bodied dinosaur from the earliest dinosaur-bearing beds
Baron, M. G., and M. E. Williams. 2018. A re-evaluation of the enigmatic (Carnian, Upper Triassic) from southern Brazil. Journal of
dinosauriform Caseosaurus crosbyensis from the Late Triassic of Systematic Palaeontology 19:1–37.
Texas, USA and its implications for early dinosaur evolution. Gauthier, J. A. 1986. Saurischian monophyly and the origin of birds.
Acta Palaeontologica Polonica 63:129–145. Memoirs of the California Academy of Sciences 8:1–55.
Baron, M. G., D. B. Norman, and P. M. Barrett. 2017. A new hypothesis of Goloboff, P. A., J. S. Farris, and K. C. Nixon. 2008. TNT, a free program
dinosaur relationships and early dinosaur evolution. Nature for phylogenetic analysis. Cladistics 24:774–786.
543:501–506. Griffin, C. T. 2018. Developmental patterns and variation among early
Benson, R. B. J., N. E. Campione, M. T. Carrano, P. D. Mannion, C. theropods. Journal of Anatomy 232:604–640.
Sullivan, P. Upchurch, and D. C. Evans. 2014. Rates of dinosaur Griffin, C. T., and S. J. Nesbitt. 2016. The femoral ontogeny and long bone
body mass evolution indicate 170 million years of sustained ecologi- histology of the Middle Triassic (? late Anisian) dinosauriform
cal innovation on the avian stem lineage. PLoS Biology 12: Asilisaurus kongwe and implications for the growth of early dino-
e1001853. saurs. Journal of Vertebrate Paleontology 36:e1111224.
Bittencourt, J. D. S., and A. W. Kellner. 2009. The anatomy and phyloge- Horn, B. L. D., T. M. Melo, C. L. Schultz, R. P. Philipp, H. P. Kloss, and K.
netic position of the Triassic dinosaur Staurikosaurus pricei Colbert, Goldberg. 2014. A new third-order sequence stratigraphic frame-
1970. Zootaxa 2079:1–56. work applied to the Triassic of the Paraná Basin, Rio Grande do
Brochu, C. A. 1996. Closure of neurocentral sutures during crocodilian Sul, Brazil, based on structural, stratigraphic and paleontological
ontogeny: implications for maturity assessment in fossil archosaurs. data. Journal of South American Earth Sciences 55:123–132.
Journal of Vertebrate Paleontology 16:49–62. Huene, F. von. 1932. Die fossile Reptil-Ordnung Saurischia, ihre
Bronzati, M., R. T. Müller, and M. C. Langer. 2019. Skull remains of the Entwicklung und Geschichte. Monographien zur Geologie und
dinosaur Saturnalia tupiniquim (Late Triassic, Brazil): with com- Paläontologie 4:1–361.
ments on the early evolution of sauropodomorph feeding behav- Hutchinson, J. R. 2001. The evolution of pelvic osteology and soft tissues
iour. PLoS ONE 14:e0221387. on the line to extant birds (Neornithes). Zoological Journal of the
Bronzati, M., O. W. M. Rauhut, J. S. Bittencourt, and M. C. Langer. 2017. Linnean Society 131:123–168.
Endocast of the Late Triassic (Carnian) dinosaur Saturnalia tupini- Koob, S. P. 1986. The use of Paraloid B-72 as an adhesive: its application
quim: implications for the evolution of brain tissue in for archaeological ceramics and other materials. Studies in conser-
Sauropodomorpha. Scientific Reports 7:11931. vation 31:7–14.
Butler, R. J. 2010. The anatomy of the basal ornithischian dinosaur Lacovara, K. J., M. C. Lamanna, L. M. Ibiricu, J. C. Poole, E. R.
Eocursor parvus from the lower Elliot Formation (Late Triassic) Schroeter, P. V. Ullmann, K. K. Voegele, Z. M. Boles, A. J.
of South Africa. Zoological Journal of the Linnean Society Carter, E. K. Fowler, V. M. Egerton, A. E. Moyer, C. L.
160:648–684. Coughenour, J. P. Schein, J. D. Harris, R. D. Martínez, and F. E.
Cabreira, S. F., C. L. Schultz, J. S. Bittencourt, M. B. Soares, D. C. Fortier, Novas. 2014. A gigantic, exceptionally complete titanosaurian saur-
L. R. da Silva, and M. C. Langer. 2011. New stem-sauropodomorph opod dinosaur from southern Patagonia, Argentina. Scientific
(Dinosauria, Saurischia) from the Triassic of Brazil. Reports 4:6196.
Naturwissenschaften 98:1035–1040. Langer, M. C. 2003. The pelvic and hindlimb anatomy of the stem-saur-
Cabreira, S. F., A. W. A. Kellner, S. Dias-da-Silva, L. R. da Silva, M. opodomorph Saturnalia tupiniquim (Late Triassic, Brazil).
Bronzati, J. C. A. Marsola, R. T. Müller, J. S. Bittencourt, B. J. Paleobios 23:1–40.
Batista, T. Raugust, R. Carrilho, A. Brodt, and M. C. Langer. Langer, M. C., and M. J. Benton. 2006. Early dinosaurs: a phylogenetic
2016. A unique Late Triassic dinosauromorph assemblage reveals study. Journal of Systematic Palaeontology 4:309–358.
dinosaur ancestral anatomy and diet. Current Biology 26:3090– Langer, M. C., J. S. Bittencourt, and C. L. Schultz. 2011. A reassessment
3095. of the basal dinosaur Guaibasaurus candelariensis, from the Late
Campione, N. E., D. C. Evans, C. M. Brown, and M. T. Carrano. 2014. Triassic Caturrita Formation of south Brazil. Earth and
Body mass estimation in non-avian bipeds using a theoretical con- Environmental Science Transactions of the Royal Society of
version to quadruped stylopodial proportions. Methods in Edinburgh 101:301–332.
Ecology and Evolution 5:913–923. Langer, M. C., M. A. G. Franca, and S. Gabriel. 2007b. The pectoral girdle
Carballido, J. L., D. Pol, A. Otero, I. A. Cerda, L. Salgado, A. C. Garrido, and forelimb anatomy of the stem-sauropodomorph Saturnalia tupi-
J. Ramezani, N. R. Cúneo, and J. M. Krause. 2017. A new giant tita- niquim (Upper Triassic, Brazil). Special Papers in Palaeontology
nosaur sheds light on body mass evolution among sauropod dino- 77:113–137.
saurs. Proceedings of the Royal Society B: Biological Sciences Langer, M. C., J. Ramezani, and Á. A. S. Da-Rosa. 2018. U-Pb age con-
284:20171219. straints on dinosaur rise from south Brazil. Gondwana Research
Da-Rosa, Á. A. S. 2015. Geological context of the dinosauriform-bearing 57:133−140.
outcrops from the Triassic of Southern Brazil. Journal of South Langer, M. C., F. Abdala, M. Richter, and M. J. Benton. 1999. A sauro-
American Earth Sciences 61:108–119. podomorph dinosaur from the Upper Triassic (Carnian) of southern
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-19)

Brazil. Comptes Rendus de l’Académie des Sciences, Series IIA, Müller, R. T., B. M. V. Baczko, J. B. Desojo, and S. J. Nesbitt. 2020b. The
Earth and Planetary Science 329:511–517. first ornithosuchid from Brazil and its macroevolutionary and phy-
Langer, M. C., A. M. Ribeiro, C. L. Schultz, and J. Ferigolo. 2007a. The logenetic implications for Late Triassic faunas in Gondwana. Acta
continental tetrapod bearing Triassic of south Brazil. Bulletin of Palaeontologica Polonica 65:1–10.
the New Mexico Museum of Natural History and Science 41:201 Müller, R. T., M. C. Langer, S. F. Cabreira, and S. Dias-da-Silva. 2016. The
−218. femoral anatomy of Pampadromaeus barberenai based on a new
Langer, M. C., B. W. McPhee, J. C. A. Marsola, L. R. da Silva, and S. F. specimen from the Upper Triassic of Brazil. Historical Biology
Cabreira. 2019. Anatomy of the dinosaur Pampadromaeus barbere- 28:656−665.
nai (Saurischia–Sauropodomorpha) from the Late Triassic Santa Müller, R. T., J. D. Ferreira, F. A. Pretto, M. Bronzati, and L. Kerber.
Maria Formation of southern Brazil. PLoS ONE 14:e0212543. 2021. The endocranial anatomy of Buriolestes schultzi
Leal, L. A., S. A. K. Azevedo, A. W. A. Kellner, and Á. A. S. Da-Rosa. (Dinosauria: Saurischia) and the early evolution of brain tissues in
2004. A new early dinosaur (Sauropodomorpha) from the Caturrita sauropodomorph dinosaurs. Journal of Anatomy 238(4):809–827.
Formation (Late Triassic), Paraná Basin, Brazil. Zootaxa 690:1–24. Müller, R. T., M. C. Langer, M. Bronzati, C. P. Pacheco, S. F. Cabreira,
Liparini, A., T. V. Oliveira, F. A. Pretto, M. B. Soares, and C. L. Schultz. and S. Dias-da-Silva. 2018a. Early evolution of sauropodomorphs:
2013. The lower jaw and dentition of the traversodontid anatomy and phylogenetic relationships of a remarkably well-pre-
Exaeretodon riograndensis Abdala, Barberena & Dornelles, from served dinosaur from the Upper Triassic of southern Brazil.
the Brazilian Triassic (Santa Maria 2 Sequence, Hyperodapedon Zoological Journal of the Linnean Society 184:1187–1248.
Assemblage Zone). Alcheringa 37:331−337. Nesbitt, S. J. 2011. The early evolution of archosaurs: relationships and
Mallison, H. 2010. The digital Plateosaurus II: an assessment of the range the origin of major clades. Bulletin of the American Museum of
of motion of the limbs and vertebral column and of previous recon- Natural History 352:1–292.
structions using a digital skeletal mount. Acta Palaeontologica Nesbitt, S. J., and M. D. Ezcurra. 2015. The early fossil record of dinosaurs
Polonica 55:433–458. in North America: a new neotheropod from the base of the Upper
Marsola, J. C., J. S. Bittencourt, R. J. Butler, Á. A. S. Da-Rosa, J. M. Triassic Dockum Group of Texas. Acta Palaeontologica Polonica
Sayão, and M. C. Langer. 2018. A new dinosaur with theropod affi- 60:513–526.
nities from the Late Triassic Santa Maria Formation, South Brazil. Nesbitt, S. J., C. A. Sidor, R. B. Irmis, K. D. Angielczyk, R. M. H. Smith,
Journal of Vertebrate Paleontology 38:e1531878. and L. A. Tsuji. 2010. Ecologically distinct dinosaurian sister group
Martinez, R. N., and O. A. Alcober. 2009. A basal sauropodomorph shows early diversification of Ornithodira. Nature 464:95–98.
(Dinosauria: Saurischia) from the Ischigualasto Formation Nesbitt, S. J., R. J. Butler, M. D. Ezcurra, P. M. Barrett, M. R. Stocker, K.
(Triassic, Carnian) and the early evolution of Sauropodomorpha. D. Angielczyk, R. M. H. Smith, C. A. Sidor, G. Niedźwiedzki, A. G.
PLoS ONE 4:e4397. Sennikov, and A. J. Charig. 2017. The earliest bird-line archosaurs
Martinez, R. N., P. C. Sereno, O. A. Alcober, C. E. Colombi, P. R. Renne, and the assembly of the dinosaur body plan. Nature 544:484–487.
I. P. Montañez, and B. S. Currie. 2011. A basal dinosaur from the Novas, F. E. 1993. New information on the systematics and postcranial skel-
dawn of the dinosaur era in southwestern Pangaea. Science eton of Herrerasaurus ischigualastensis (Theropoda:
331:206–210. Herrerasauridae) from the Ischigualasto Formation (Upper Triassic)
Martinez, R. N., C. Apaldetti, O. A. Alcober, C. E. Colombi, P. C. Sereno, of Argentina. Journal of Vertebrate Paleontology 13:400–423.
E. Fernandez, P. S. Malnis, G. A. Correa, and D. Abelin. 2013. Novas, F. E., M. D. Ezcurra, S. Chatterjee, and T. S. Kutty. 2011. New
Vertebrate succession in the Ischigualasto Formation. Journal of dinosaur species from the Upper Triassic Upper Maleri and
Vertebrate Paleontology 32(6, Supplement):10–30. Lower Dharmaram formations of central India. Earth and
McPhee, B. W., and J. N. Choiniere. 2018. The osteology of Pulanesaura Environmental Science Transactions of the Royal Society of
eocollum: implications for the inclusivity of Sauropoda Edinburgh 101:333–349.
(Dinosauria). Zoological Journal of the Linnean Society 182:830– Novas, F. E., F. L. Agnolin, M. D. Ezcurra, R. T. Müller, M. Martinelli,
861. and M. Langer. 2021. Review of the fossil record of early dinosaurs
McPhee, B. W., J. N. Choiniere, A. M. Yates, and P. A. Viglietti. 2015. A from South America, and its phylogenetic implications. Journal of
second species of Eucnemesaurus Van Hoepen, 1920 (Dinosauria, South American Earth Sciences 110:103341.
Sauropodomorpha): new information on the diversity and evolution Oliveira, T. V., C. L. Schultz, and M. B. Soares. 2007. O esqueleto pós-cra-
of the sauropodomorph fauna of South Africa’s lower Elliot niano de Exaeretodon riograndensis Abdala et al. (Cynodontia,
Formation (latest Triassic). Journal of Vertebrate Paleontology 35: Traversodontidae), Triássico do Brasil. Revista Brasileira de
e980504. Paleontologia 10:79–94.
McPhee, B. W., A. M. Yates, J. N. Choiniere, and F. Abdala. 2014. The Oliveira, T. V., M. B. Soares, and C. L. Schultz. 2010. Trucidocynodon rio-
complete anatomy and phylogenetic relationships of Antetonitrus grandensis gen. nov. et sp. nov. (Eucynodontia), a new cynodont
ingenipes (Sauropodiformes, Dinosauria): implications for the from the Brazilian Upper Triassic (Santa Maria Formation).
origins of Sauropoda. Zoological Journal of the Linnean Society Zootaxa 2382:1−71.
171:151–205. Otero, A., and D. Pol. 2013. Postcranial anatomy and phylogenetic
McPhee, B. W., R. B. Benson, J. Botha-Brink, E. M. Bordy, and J. N. relationships of Mussaurus patagonicus (Dinosauria,
Choiniere. 2018. A giant dinosaur from the earliest Jurassic of Sauropodomorpha). Journal of Vertebrate Paleontology 33:1138–
South Africa and the transition to quadrupedality in early 1168.
Sauropodomorphs. Current Biology 28:3143–3151. Otero, A., J. L. Carballido, L. Salgado, J. I. Canudo, and A. C. Garrido.
Montefeltro, F. C., M. C. Langer, and C. L. Schultz. 2010. Cranial 2021. Report of a giant titanosaur sauropod from the Upper
anatomy of a new genus of hyperodapedontine rhynchosaur Cretaceous of Neuquén Province, Argentina. Cretaceous
(Diapsida, Archosauromorpha) from the Upper Triassic of Research 122:104754.
southern Brazil. Earth and Environmental Science Transactions of Otero, A., E. Krupandan, D. Pol, A. Chinsamy, and J. Choiniere. 2015. A
the Royal Society of Edinburgh 101:27–52. new basal sauropodiform from South Africa and the phylogenetic
Müller, R. T. 2021. A new theropod dinosaur from a peculiar Late relationships of basal sauropodomorphs. Zoological Journal of the
Triassic assemblage of southern Brazil. Journal of South Linnean Society 174:589–634.
American Earth Sciences 107:103026. Owen, R. 1842. Report on British fossil reptiles, Part II. Reports of the
Müller, R. T., and M. S. Garcia. 2020. Rise of an empire: analysing the British Association for the Advancement of Science 11:60–204.
high diversity of the earliest sauropodomorph dinosaurs through Pacheco, C., R. T. Müller, M. C. Langer, F. A. Pretto, L. Kerber, and S. D.
distinct hypotheses. Historical Biology 32:1334–1339. da Silva. 2019. Gnathovorax cabreirai: a new early dinosaur and the
Müller, R. T., M. S. Garcia, and F. A. Pretto. 2020a. Comments on origin and initial radiation of predatory dinosaurs. PeerJ 7:e7963.
additional dinosaur specimens from the Janner site (Upper Persons IV, W. S., and P. J. Currie. 2016. An approach to scoring cursorial
Triassic of the Paraná Basin), southern Brazil. Revista Brasileira limb proportions in carnivorous dinosaurs and an attempt to
de Paleontologia 23:171–184. account for allometry. Scientific Reports 6:19828.
Müller, R. T., M. C. Langer, and S. Dias-da-Silva 2018b. An exceptionally Peyre de Fabrègues, C., and R. Allain. 2016. New material and revision of
preserved association of complete dinosaur skeletons reveals the Melanorosaurus thabanensis, a basal sauropodomorph from the
oldest long-necked sauropodomorphs. Biology Letters 14:20180633. Upper Triassic of Lesotho. PeerJ 4:e1639.
Müller and Garcia—An early sauropodomorph from Brazil (e2002879-20)

Peyre de Fabrègues, C., and R. Allain. 2019. Kholumolumo ellenberger- Sereno, P. C. 1993. The pectoral girdle and forelimb of the basal theropod
orum, gen. et sp. nov., a new early sauropodomorph from the Herrerasaurus ischigualastensis. Journal of Vertebrate Paleontology
lower Elliot Formation (Upper Triassic) of Maphutseng, Lesotho. 13:425–450.
Journal of Vertebrate Paleontology 39:e1732996. Sereno, P. C. 2007. The phylogenetic relationships of early dinosaurs: a
Piechowski, R., and M. Tałanda. 2020. The locomotor musculature and comparative report. Historical Biology 19:145–155.
posture of the early dinosauriform Silesaurus opolensis provides a Sereno, P. C., R. N. Martínez, and O. A. Alcober. 2013. Osteology of
new look into the evolution of Dinosauromorpha. Journal of Eoraptor lunensis (Dinosauria, Sauropodomorpha). Journal of
Anatomy 236:1044–1100. Vertebrate Paleontology 32(6, Supplement):83–179.
Piechowski, R., M. Tałanda, and J. Dzik. 2014. Skeletal variation and Stefanello, M., R. T. Müller, L. Kerber, R. N. Martínez, and S. Dias-Da-
ontogeny of the Late Triassic dinosauriform Silesaurus opolensis. Silva. 2018. Skull anatomy and phylogenetic assessment of
Journal of Vertebrate Paleontology 34:1383–1393. a large specimen of Ecteniniidae (Eucynodontia:
Pinheiro, F. L. 2016. A fragmentary dinosaur femur and the presence of Probainognathia) from the Upper Triassic of southern Brazil.
Neotheropoda in the Upper Triassic of Brazil. Revista Brasileira de Zootaxa 4457:351−378.
Paleontologia 19:211–216. Wilson, J. A. 1999. A nomenclature for vertebral laminae in sauropods
Pol, D., and J. E. Powell. 2007. New information on Lessemsaurus sauro- and other saurischian dinosaurs. Journal of Vertebrate
poides (Dinosauria: Sauropodomorpha) from the Upper Triassic of Paleontology 19:639–653.
Argentina. Special Papers in Palaeontology 77:223–243. Wilson, J. A., and P. Upchurch. 2003. A revision of Titanosaurus
Pol, D., A. Otero, C. Apaldetti, and R. N. Martínez. 2021. Triassic sauro- Lydekker (Dinosauria-Sauropoda), the first dinosaur genus with a
podomorph dinosaurs from South America: The origin and diversi- ‘Gondwanan’ distribution. Journal of Systematic Palaeontology
fication of dinosaur dominated herbivorous faunas. Journal of South 1:125–160.
American Earth Sciences 107:103145. Wilson, J. A., D. D. Michael, T. Ikejiri, E. M. Moacdieh, and J. A.
Pretto, F. A., M. C. Langer, and C. L. Schultz. 2018. A new dinosaur Whitlock. 2011. A nomenclature for vertebral fossae in sauropods
(Saurischia: Sauropodomorpha) from the Late Triassic of Brazil and other saurischian dinosaurs. PLoS ONE 6:e17114.
provides insights on the evolution of sauropodomorph body plan. Yates, A. M., M. F. Bonnan, J. Neveling, A. Chinsamy, and M. G.
Zoological Journal of the Linnean Society 185:388–416. Blackbeard. 2010. A new transitional sauropodomorph dinosaur
Reiss, S., and H. Mallison. 2014. Motion range of the manus of Plateosaurus from the Early Jurassic of South Africa and the evolution of sauro-
engelhardti von Meyer, 1837. Palaeontologia Electronica 17:1–19. pod feeding and quadrupedalism. Proceedings of the Royal Society
Riga, B. J. G., M. C. Lamanna, L. D. O. David, J. O. Calvo, and J. P. Coria. B: Biological Sciences 277:787–794.
2016. A gigantic new dinosaur from Argentina and the evolution of Zerfass, H., E. L. Lavina, C. L. Schultz, A. J. V. Garcia, U. F. Faccini, and
the sauropod hind foot. Scientific Reports 6:19165. F. Jr. Chemale. 2003. Sequence stratigraphy of continental Triassic
Schultz, C. L., A. G. Martinelli, M. B. Soares, F. L. Pinheiro, L. Kerber, B. strata of Southernmost Brazil: a contribution to Southwestern
L. Horn, F. A. Pretto, R. T. Müller, and T. P. Melo. 2020. Triassic Gondwana palaeogeography and palaeoclimate. Sedimentary
faunal successions of the Paraná Basin, southern Brazil. Journal Geology 161:85−105.
of South American Earth Sciences 104:102846.
Seeley, H. G. 1887. On the classification of the fossil animals commonly Submitted April 15, 2021; revisions received October 14, 2021;
named Dinosauria. Proceedings of the Royal Society of London accepted October 17, 2021.
43:165–171. Handling Editor: Michael D’Emic.

You might also like