palaeogeography, palaeoclimatology, palaeoecology · be linked to the evolution of larger body...

18
Contents lists available at ScienceDirect Palaeogeography, Palaeoclimatology, Palaeoecology journal homepage: www.elsevier.com/locate/palaeo Feeding ecology of Tragelaphini (Bovidae) from the Shungura Formation, Omo Valley, Ethiopia: Contribution of dental wear analyses Cécile Blondel a, , John Rowan b , Gildas Merceron a , Faysal Bibi c , Enquye Negash d , W. Andrew Barr d , Jean-Renaud Boisserie a,e a Laboratoire Paléontologie Évolution Paléoécosystèmes Paléoprimatologie: UMR 7262, Bât. B35 TSA 51106, 6 rue M. Brunet, 86073 Poitiers Cedex9, France b Institute of Human Origins, School of Human Evolution and Social Change, Arizona State University, Tempe, AZ 85282, USA c Museum für Naturkunde, Leibniz Institute for Evolution and Biodiversity Science, Invalidenstrasse 43, 10115 Berlin, Germany d Center for the Advanced Study of Human Paleobiology, Department of Anthropology, the George Washington University, Washington, DC 20052, USA e Centre Français des Études Éthiopiennes, USR 3137 (CNRS & MEAE, Ambassade de France en Éthiopie), P. O. Box 5554, Addis Ababa, Ethiopia ARTICLE INFO Keywords: Paleoecology Diet Dental mesowear Dental microwear Plio-Pleistocene Africa ABSTRACT To better understand the environmental conditions that prevailed in the Plio-Pleistocene Shungura Formation, Lower Omo Valley, Ethiopia, we analyze the feeding preferences of Tragelaphini, the third most common tribe of Shungura bovids. Molar mesowear and dental microwear texture analyses were applied to three species (Tragelaphus rastafari, T. nakuae, and T. gaudryi) and body mass estimates were calculated for the T. rastafari- nakuae lineage to test whether dietary shifts were linked to body mass changes. We compare our results with previous work on stable carbon isotopes (δ 13 C) of enamel, which indicated that tragelaphins from the Shungura Formation possessed dietary exibility. We found that the both the T. rastafari-nakuae lineage and T. gaudryi maintained a mixed feeding dietary niche, with varying proportions of C 3 versus C 4 inputs, from ~3.6 to > 2 Ma. Our results show that T. rastafari consumed more browse than its descendant, T. nakuae, which was a mixed feeder consuming a greater proportion of C 4 dicots ~ 2.8 Ma. The T. rastafari-T. nakuae dietary shift may reect environmental changes in the Shungura Formation during this time, but appears to be oset from body mass increases in this lineage, which occurred gradually through the Plio-Pleistocene. This study highlights the im- portance of a multi-proxy approach to precisely determine the dietary ecologies of extinct bovids and points to how each proxy oers a slightly dierent perspective on the ecology of fossil organisms. 1. Introduction 1.1. Background The Shungura Formation of the Lower Omo Valley, Ethiopia, con- stitutes the most complete stratigraphic and paleontological record for the upper Pliocene lower Pleistocene of Africa (Fig. 1). This con- tinental sequence corresponds to ~800 m of interstratied volcanic tus and uviatile to uvio-lacustrine sediments, deposited almost continuously between ~3.6 Ma and 1.0 Ma (de Heinzelin, 1983). Through decades of eldwork by Arambourg (1934, 1943, 1947), the Omo Research Expedition (Howell, 1968), the International Omo Re- search Expedition (IORE, Coppens et al., 1976), and the Omo Group Research Expedition (OGRE, Boisserie et al., 2008, 2010), the Shungura Formation has yielded a large collection of fossil vertebrates including representatives of Hominidae (Arambourg and Coppens, 1967; Howell and Coppens, 1976; Suwa, 1988) and Bovidae (Gentry, 1985). Bovids constitute 42% of the Shungura faunal remains and the three most common tribes are Reduncini, Tragelaphini, and Aepycerotini (Alemseged et al., 2007). Thus, bovids have been extensively used as paleoecological and paleoenvironmental indicators for the Shungura Formation. The relative abundance of taxa through time (Bobe and Eck, 2001), ecomorphology of astragali (Barr, 2015; Plummer et al., 2015), and stable carbon isotopes (δ 13 C) of enamel (Bibi et al., 2013; Negash et al., 2015) generally suggest that bovids lived in an environment composed of a mosaic of habitats ranging from forests to open wood- lands between ~3.4 Ma and 2 Ma, marked by a trend towards a greater abundance of C 4 vegetation through time. According to these previous studies, the proportions of each habitat diered in space and in time, but reconstructions of habitat proportions often dier across proxies. To improve our understanding of the habitat distribution and overall en- vironmental conditions that prevailed in the Shungura Formation, we https://doi.org/10.1016/j.palaeo.2018.01.027 Received 10 February 2017; Received in revised form 16 January 2018; Accepted 22 January 2018 Corresponding author. E-mail addresses: [email protected] (C. Blondel), [email protected] (J. Rowan), [email protected] (G. Merceron), [email protected] (F. Bibi), [email protected] (E. Negash), [email protected] (W.A. Barr), [email protected] (J.-R. Boisserie). Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120 Available online 31 January 2018 0031-0182/ © 2018 Elsevier B.V. All rights reserved. T

Upload: others

Post on 11-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology

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

Feeding ecology of Tragelaphini (Bovidae) from the Shungura Formation,Omo Valley, Ethiopia: Contribution of dental wear analyses

Cécile Blondela,⁎, John Rowanb, Gildas Mercerona, Faysal Bibic, Enquye Negashd,W. Andrew Barrd, Jean-Renaud Boisseriea,e

a Laboratoire Paléontologie Évolution Paléoécosystèmes Paléoprimatologie: UMR 7262, Bât. B35 TSA 51106, 6 rue M. Brunet, 86073 Poitiers Cedex9, Franceb Institute of Human Origins, School of Human Evolution and Social Change, Arizona State University, Tempe, AZ 85282, USAcMuseum für Naturkunde, Leibniz Institute for Evolution and Biodiversity Science, Invalidenstrasse 43, 10115 Berlin, Germanyd Center for the Advanced Study of Human Paleobiology, Department of Anthropology, the George Washington University, Washington, DC 20052, USAe Centre Français des Études Éthiopiennes, USR 3137 (CNRS & MEAE, Ambassade de France en Éthiopie), P. O. Box 5554, Addis Ababa, Ethiopia

A R T I C L E I N F O

Keywords:PaleoecologyDietDental mesowearDental microwearPlio-PleistoceneAfrica

A B S T R A C T

To better understand the environmental conditions that prevailed in the Plio-Pleistocene Shungura Formation,Lower Omo Valley, Ethiopia, we analyze the feeding preferences of Tragelaphini, the third most common tribe ofShungura bovids. Molar mesowear and dental microwear texture analyses were applied to three species(Tragelaphus rastafari, T. nakuae, and T. gaudryi) and body mass estimates were calculated for the T. rastafari-nakuae lineage to test whether dietary shifts were linked to body mass changes. We compare our results withprevious work on stable carbon isotopes (δ13C) of enamel, which indicated that tragelaphins from the ShunguraFormation possessed dietary flexibility. We found that the both the T. rastafari-nakuae lineage and T. gaudryimaintained a mixed feeding dietary niche, with varying proportions of C3 versus C4 inputs, from ~3.6 to> 2Ma.Our results show that T. rastafari consumed more browse than its descendant, T. nakuae, which was a mixedfeeder consuming a greater proportion of C4 dicots ~ 2.8Ma. The T. rastafari-T. nakuae dietary shift may reflectenvironmental changes in the Shungura Formation during this time, but appears to be offset from body massincreases in this lineage, which occurred gradually through the Plio-Pleistocene. This study highlights the im-portance of a multi-proxy approach to precisely determine the dietary ecologies of extinct bovids and points tohow each proxy offers a slightly different perspective on the ecology of fossil organisms.

1. Introduction

1.1. Background

The Shungura Formation of the Lower Omo Valley, Ethiopia, con-stitutes the most complete stratigraphic and paleontological record forthe upper Pliocene – lower Pleistocene of Africa (Fig. 1). This con-tinental sequence corresponds to ~800m of interstratified volcanictuffs and fluviatile to fluvio-lacustrine sediments, deposited almostcontinuously between ~3.6Ma and 1.0Ma (de Heinzelin, 1983).Through decades of fieldwork by Arambourg (1934, 1943, 1947), theOmo Research Expedition (Howell, 1968), the International Omo Re-search Expedition (IORE, Coppens et al., 1976), and the Omo GroupResearch Expedition (OGRE, Boisserie et al., 2008, 2010), the ShunguraFormation has yielded a large collection of fossil vertebrates includingrepresentatives of Hominidae (Arambourg and Coppens, 1967; Howell

and Coppens, 1976; Suwa, 1988) and Bovidae (Gentry, 1985). Bovidsconstitute 42% of the Shungura faunal remains and the three mostcommon tribes are Reduncini, Tragelaphini, and Aepycerotini(Alemseged et al., 2007). Thus, bovids have been extensively used aspaleoecological and paleoenvironmental indicators for the ShunguraFormation. The relative abundance of taxa through time (Bobe and Eck,2001), ecomorphology of astragali (Barr, 2015; Plummer et al., 2015),and stable carbon isotopes (δ13C) of enamel (Bibi et al., 2013; Negashet al., 2015) generally suggest that bovids lived in an environmentcomposed of a mosaic of habitats ranging from forests to open wood-lands between ~3.4Ma and 2Ma, marked by a trend towards a greaterabundance of C4 vegetation through time. According to these previousstudies, the proportions of each habitat differed in space and in time,but reconstructions of habitat proportions often differ across proxies. Toimprove our understanding of the habitat distribution and overall en-vironmental conditions that prevailed in the Shungura Formation, we

https://doi.org/10.1016/j.palaeo.2018.01.027Received 10 February 2017; Received in revised form 16 January 2018; Accepted 22 January 2018

⁎ Corresponding author.E-mail addresses: [email protected] (C. Blondel), [email protected] (J. Rowan), [email protected] (G. Merceron), [email protected] (F. Bibi),

[email protected] (E. Negash), [email protected] (W.A. Barr), [email protected] (J.-R. Boisserie).

Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

Available online 31 January 20180031-0182/ © 2018 Elsevier B.V. All rights reserved.

T

Page 2: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

use a multi-proxy approach to precisely determine the feeding pre-ferences of an abundant bovid tribe, Tragelaphini.

1.2. Aim of the study

Tragelaphini are abundant in the Plio-Pleistocene hominin sites ofeastern Africa and, interestingly, show significant intra-tribal dietaryvariability among fossil bovids, ranging from predominantly C3

browsing to C4 grazing (Bedaso et al., 2013; Bibi et al., 2013; Negashet al., 2015). This contrasts sharply with extant tragelaphins that haveC3-dominated diets and subsist mainly on browse (Cerling et al., 2003;Sponheimer et al., 2003). The dietary variability of fossil tragelaphinsacross space and time provides a significant opportunity to investigatepaleoenvironmental dynamics related to human evolution, as previouswork has suggested changes in tragelaphin diets and paleoenviron-ments may be correlated (Bibi et al., 2013).

In the Shungura Formation, Tragelaphini are the third mostcommon tribe of bovids and, in contrast to other tribes, their dentalremains are identifiable to the species-level (Gentry, 1985; Bibi, 2011).This circumvents the need to solely rely on dental specimens associatedwith diagnostic elements (i.e., horn cores and skulls) to make com-parisons across species. The three most common species in the Shun-gura Formation are Tragelaphus rastafari and T. nakuae, an anageneti-cally evolving lineage that is the most common, and T. gaudryi. Othertragelaphin species (Tragelaphus pricei and T. strepsiceros) are rare in the

sequence (Gentry, 1985; Bobe and Eck, 2001). Bibi (2011) showed thatthe replacement of T. rastafari by T. nakuae occurred between ~2.95and 2.74Ma and that T. nakuae exhibits a highly derived cranial morphafter ~2.3Ma. The morphological changes occurring in this lineagemay or may not be correlated with a dietary change as indicated byavailable δ13C data. For example, Bibi et al. (2013) demonstrated a shiftin δ13C values towards greater C4 consumption ~ 2.8Ma, whereas δ13Cdata from Negash et al. (2015) indicate that this tribe was characterizedby a wide spectrum of C3- to C4-dominated diets throughout theShungura sequence.

Here we aim to clarify this discrepancy and provide new pa-leoenvironmental data using a multi-proxy approach to determine thedietary preferences of Tragelaphus rastafari-T. nakuae and T. gaudryifrom the Shungura Formation. To describe the feeding traits of theseextinct bovids accurately, we combine two complementary approaches:dental mesowear analyses and dental microwear textural analyses(DMTA). Dental mesowear has been shown to be an effective way ofcharacterizing the long-term signal of feeding habits for extant andextinct herbivores (e.g., Fortelius and Solounias, 2000; Kaiser andFortelius, 2003; Rivals et al., 2007; Blondel et al., 2010; Mihlbachleret al., 2011; Kaiser et al., 2013). Dental microwear provides a short-term dietary signal (a few days to a few weeks) of the physical prop-erties of the last food items consumed by an individual (Teaford andOyen, 1989). DMTA is based on the automated quantification of 3Dsurfaces by using a scale-sensitive fractal analysis (Scott et al., 2006) or

Fig. 1. The Shungura Formation. A, location in Africa (southwestern Ethiopia); B, outcrops of the Shungura Formation (in orange) in the Lower Omo Valley; C, chronostratigraphicsequence of the Shungura Formation after de Heinzelin (1983). Dated horizons are marked by red triangles and dates in Ma provided at member boundaries after Feibel et al. (1989),McDougall and Brown (2006, 2008), and McDougall et al. (2012). D, number of molars analyzed in each member of the Shungura Formation. MSW, mesowear score; DMTA, dentalmicrowear texture analysis.

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

104

Page 3: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

surface texture analysis using ISO parameters (Schulz et al., 2010,2013; Winkler et al., 2016). DMTA has been shown to be an effectiveway to discriminate dietary differences between species in primates(Calandra et al., 2012) and bovids (Rozzi et al., 2013, Scott, 2012;Winkler et al., 2013). Likewise, this method is an appropriate proxy fordetecting variations in diet as a seasonal signal of feeding preference orfood availability (Merceron et al., 2010, 2014; see Calandra andMerceron, 2016 for a recent review).

When data were available, we compared the dental wear patterns ofindividuals with their published δ13C values of tooth enamel. Thecarbon isotopic composition of tooth enamel distinguishes the propor-tion of C3 to C4 plants consumed during the period of enamel miner-alization, with more negative (or depleted) values indicating greater C3

consumption and positive (or enriched) values indicating greater C4

consumption. The combination of published δ13C data with our newmesowear and microwear data provides a strong basis for inferring thediets of fossil tragelaphin species and the habitats they occupiedthrough time in the Shungura Formation. Furthermore, combiningthese three proxies permits investigation of how these proxies, formedon very different time scales, confirm, contrast, or complement oneanother (Davis and Pineda-Munoz, 2016).

Finally, we test whether any potential dietary shifts in the T. ras-tafari-T. nakuae lineage are related to changes in body size. We predictthat if this lineage began to consume a greater proportion of C4 vege-tation through time, as suggested by Bibi et al. (2013), then this shouldbe linked to the evolution of larger body mass. Conversely, if diets ofthe T. rastafari-T. nakuae indicate mixed-feeding through time as sug-gested by Negash et al. (2015), then we expect body mass to be stablethrough the Plio-Pleistocene. To test these hypotheses, we re-constructed body mass for the T. nakuae lineage from the mid-Pliocene(~3.6Ma) to the early Pleistocene (~2.0Ma) using complete lowermolar rows following (Janis, 1990).

2. Materials

2.1. Fossil data

Most of the dental remains are comprised of isolated molars,mandibles, or fragments of maxillae (Fig. 1, S1) from members B, C, E,F, and G of the Shungura Formation. From Member B to the lower partof Member G (units 1 to 13, from 3.44Ma to 2.05Ma), the Shungurasediments were deposited in a fluviatile context, with a sequence ofsands and sandstones at bottom and silts and clays/clayey silts at top.These were interpreted as channel deposits, levee deposits, and flood-plain deposits, respectively (de Heinzelin, 1983). These sediments al-ternate with volcanic tuff deposits, which sedimentary facies stronglysuggest were deposited underwater in a fluviatile context. The speci-mens studied here were essentially found in sandy horizons with theexception of the specimens from OMO 33, which come from a sandy-tuffitic deposit, again in a fluviatile context. The upper part of MemberG (2.05Ma–1.90Ma) is characterized by a significant change in de-positional context and is dominated by lacustrine facies, retrogradinginto deltaic facies towards the top. Only three specimens are derivedfrom this context (localities OMO 55 and OMO 78), and were used formesowear scoring only. All specimens used here displayed few in-dicators of fluvial transportation (e.g., rolling) and are well-preserved.

The Shungura Formation also includes a large number of archae-ological occurrences in Member F and lower part of Member G (Maurinet al., 2014) that are dominated by quartz lithics (see Delagnes et al.,2011, and references cited herein). However, most of these occurrencesare in silts or clays, do not co-occur with fossil concentrations, and arenot associated with significant faunal assemblages. The material stu-died here was not discovered in such archaeological contexts, but innatural fossil assemblages deposited by the paleo-Omo River or theupper Member G lake.

The Tragelaphus rastafari-nakuae lineage is represented in all

targeted members, and is easily identified based on its large dental size.Tragelaphus gaudryi (the probable ancestor of extant T. imberbis, Gentry,1985) is a smaller tragelaphin known from Member C and younger.Other tragelaphins in the Shungura Formation are T. pricei and T.strepsiceros. Tragelaphus pricei is noticeably smaller than the speciesanalyzed here and both it and T. strepsiceros are extremely rare in theShungura sequence (Gentry, 1985; Bibi, 2011). We therefore consider ithighly unlikely that either of these species are included in our dentalsamples, which can be reliably assigned to the three species of interest.All fossil specimens are curated by the Authority for Research andConservation of Cultural Heritage (ARCCH) in the National Museum ofEthiopia (Addis Ababa). The material in this study largely belongs tothe 1967–1976 IORE collections, but also includes new material fromthe OGRE collections.

Though the Shungura has abundant bovid dental specimens, manyspecimens display postmortem damage of cusp tips (e.g., 35% of buccalcusp tips of upper molars and lingual cuspid tips of lower molars inMember C) and could not be scored for mesowear analysis. A Mann-Whitney U test indicates no significant differences between mesowearscores of upper and lower molars (S4) and thus we combine them in thepresent study. Likewise, taphonomic alteration results in poor pre-servation of wear facets at microscopic scales, limiting the number ofavailable specimens for DMTA (Fig. 5). These specimens were not in-cluded. In total, we analyzed 161 specimens for mesowear and 88specimens for DMTA (Fig. 1; S1, S2).

2.2. Comparative dataset

Mesowear patterns of fossil species were compared with those ofextant species with known feeding preferences and behaviors (Table 1).All extant species and specimens used here for comparison come fromthe dataset published by Fortelius and Solounias (2000). Their datasetoriginally comprised 64 species of herbivorous mammals representing2200 adult individuals. However, the number of specimens and speciesis here deliberately limited in comparison to the original publication(Fortelius and Solounias, 2000). Following their recommendations, allsamples constituting a comparative dataset should have a reasonablesample size, and so samples having less than fourteen individuals areexcluded (Fortelius and Solounias, 2000). The original dataset is alsotruncated to ruminants and relatives represented in four dietary cate-gories: grazers, mixed feeders, browsers, and frugivores. The dietaryclassification used by Fortelius and Solounias (2000) is a compilationfrom ecological sources. Our final comparative dataset of extant speciesexclusively derived from Fortelius and Solounias (2000) comprises1324 wild-shot specimens representing 33 extant species (Table 1).

DMTA was applied to 106 wild-shot individuals of six extant speciesof ruminants (Table 3, S2) representing the main dietary poles. TheAfrican savanna buffalo (Syncerus caffer caffer, n= 6) and hartebeest(Alcelaphus buselaphus, n= 28) are C4 grazers. The semi-wild non-fedHeck cattle (Bos taurus, n= 6) from the Netherlands are C3 grazers(Vulink and Drost, 1991; Estes, 1991). Our sample of the mixed feedingred deer (Cervus elaphus, n= 29) derives from a deciduous woodlandnear Chateauroux (Region Centre-Val de Loire, France). In such habi-tats, this species' diet is known to include about 30% herbaceousmonocots (C3 grasses, sedges, rushes; Gebert and Verheyden-Tixier,2001). The giraffe (Giraffa camelopardalis, n= 12) is a browser, with aC3-dominated diet of leaves (Leuthold and Leuthold, 1978; Estes,1991). The yellow-backed duiker (Cephalophus silvicultor, n= 25) has afrugivorous diet supplemented by leaves (Lumpkin and Kranz, 1984).All of these specimens were scanned at the Laboratoire PaléontologieÉvolution Paléoécosystèmes Paléoprimatologie(PALEVOPRIM), Uni-versity of Poitiers, France. We also discuss our results by drawingcomparisons to the larger dataset of bovids from Scott (2012) using theclassification of Gagnon and Chew (2000): grazers (Gr: variable in-cluding 60%–90% monocots, and obligate including> 90% of mono-cots); mixed feeders (Mf: browser-grazer intermediates including

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

105

Page 4: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

Table 1Frequency of extant and extinct specimens in different mesowear categories and mesowear score (MWS). n, specimen number; LS, low relief with sharp, LR low relief with round, LB, lowrelief with blunt cusp(id)s; HS, high relief with sharp, HR high relief with round cusp(id)s; m, mean; sd, standard deviation; sem, standard deviation of the mean. For modern species,Fortelius and Solounias (2000).

Samples n LS LR LB HS HR MWS

m sd sem

Browsers 394 1.27 1.27 0.51 58.12 38.83 0.47 0.64 0.03Alces alces 30 0.00 0.00 0.00 100.00 0.00 0.00 0.00 0.00Antilocapra americana 43 4.65 0.00 0.00 67.44 27.91 0.37 0.58 0.09Capreolus capreolus 68 1.47 0.00 2.94 70.59 25.00 0.40 0.79 0.10Camelus dromedarius 16 0.00 0.00 0.00 31.25 68.75 0.69 0.48 0.12Giraffa camelopardalis 61 3.28 3.28 0.00 72.13 21.31 0.38 0.71 0.09Litocranius walleri 69 0.00 4.35 0.00 33.33 62.32 0.75 0.67 0.08Odocoileus hemionus 33 0.00 0.00 0.00 48.48 51.52 0.52 0.51 0.09Tragelaphus eurycerus 27 0.00 0.00 0.00 40.74 59.26 0.59 0.50 0.10Tragelaphus scriptus 47 0.00 0.00 0.00 48.94 51.06 0.51 0.51 0.07

Frugivores 142 0.00 9.86 4.93 19.01 66.20 1.15 1.01 0.08Cephalophus dorsalis 28 0.00 0.00 7.14 14.29 78.57 1.07 0.90 0.17Cephalophus niger 31 0.00 6.45 3.23 35.48 54.84 0.87 0.96 0.17Cephalophus nigrifrons 44 0.00 13.64 4.55 27.27 54.55 1.14 1.11 0.17Cephalophus silvicultor 39 0.00 15.38 5.13 0.00 79.49 1.46 0.94 0.15

Grazers 497 2.21 7.65 11.67 10.66 67.81 1.42 1.15 0.05Alcelaphus buselaphus 76 5.26 15.79 22.37 11.84 44.74 1.92 1.41 0.16Bison bison 15 0.00 26.67 73.33 0.00 0.00 3.73 0.46 0.12Cervus duvaucelii 50 2.00 8.00 24.00 10.00 56.00 1.80 1.41 0.20Connochaetes taurinus 52 5.77 13.46 23.08 15.38 42.31 1.87 1.46 0.20Hippotragus equinus 26 0.00 15.38 0.00 3.85 80.77 1.27 0.78 0.15Hippotragus niger 20 0.00 0.00 15.00 0.00 85.00 1.45 1.10 0.25Kobus ellipsiprymnus 22 0.00 4.55 0.00 0.00 95.45 1.09 0.43 0.09Ourebia ourebi 128 1.56 1.56 0.78 20.31 75.78 0.87 0.58 0.05Redunca redunca 77 1.30 5.19 2.60 5.19 85.71 1.14 0.70 0.08Syncerus caffer 31 0.00 0.00 0.00 0.00 100.00 1.00 0.00 0.00

Mixed feeders 291 3.09 3.44 3.44 38.83 51.20 0.81 0.91 0.05Aepyceros melampus 17 0.00 0.00 0.00 35.29 64.71 0.65 0.49 0.12Antidorcas marsupialis 26 3.85 0.00 0.00 69.23 26.92 0.35 0.56 0.11Axis axis 43 0.00 2.33 20.93 6.98 69.77 1.60 1.31 0.20Budorcas taxicolor 38 2.63 2.63 0.00 36.84 57.89 0.71 0.65 0.11Cervus unicolor 21 4.76 0.00 4.76 9.52 80.95 1.10 0.77 0.17Ovibos moschatus 52 9.62 9.62 0.00 44.23 36.54 0.85 0.96 0.13Ovis canadensis 29 3.45 10.34 0.00 51.72 34.48 0.72 0.96 0.18Tragelaphus oryx 14 0.00 0.00 0.00 50.00 50.00 0.50 0.52 0.14Tragelaphus angasi 20 0.00 0.00 0.00 35.00 65.00 0.65 0.49 0.11

Tragelaphus imberbis 31 0.00 0.00 0.00 31.94 58.06 0.42 0.50 0.09Fossil taxaTragelaphini total 161 2.48 1.86 0.00 39.13 56.52 0.68 0.62 0.05

Lower molar 64 1.56 3.13 0.00 31.25 64.06 0.78 0.63 0.08Upper molar 97 3.09 1.03 0.00 44.33 51.55 0.61 0.60 0.06

Tragelaphus gaudryi 49 2.04 0.00 0.00 40.82 57.14 0.61 0.53 0.08Lower molar 15 0.00 0.00 0.00 26.67 73.33 0.73 0.46 0.12Upper molar 34 2.94 0.00 0.00 47.06 50.00 0.56 0.56 0.10

Tragelaphus nakuae 102 1.96 2.94 0.00 39.22 55.88 0.70 0.66 0.07Lower molar 44 0.00 4.55 0.00 31.82 63.64 0.80 0.67 0.10Upper molar 58 3.45 1.72 0.00 44.83 50.00 0.62 0.64 0.08

Tragelaphus rastafari 6 16.67 0.00 0.00 33.33 50.00 0.83 0.75 0.31Lower molar 5 20.00 0.00 0.00 40.00 40.00 0.80 0.84 0.37Upper molar 1 0.00 0.00 0.00 0.00 100.00 1.00

Tragelaphini indet. 4 0.00 0.00 0.00 25.00 75.00 0.75 0.50 0.25Member BTragelaphus rastafari 6 16.67 0.00 0.00 33.33 50.00 0.83 0.75 0.31

Lower molar 5 20.00 0.00 0.00 40.00 40.00 0.80 0.84 0.37Upper molar 1 0.00 0.00 0.00 0.00 100.00 1.00

Member CTragelaphus gaudryi 1 0.00 0.00 0.00 0.00 100.00 1.00Tragelaphus nakuae 23 0.00 8.70 0.00 30.43 60.87 0.87 0.81 0.17

Lower molar 17 0.00 11.76 0.00 29.41 58.82 0.94 0.90 0.22Upper molar 6 0.00 0.00 0.00 33.33 66.67 0.67 0.52 0.21

Tragelaphini indet. 3 0.00 0.00 0.00 33.33 66.67 0.67 0.58 0.33Member ETragelaphus gaudryi 7 0.00 0.00 0.00 42.86 57.14 0.57 0.53 0.20

Lower molar 5 0.00 0.00 0.00 20.00 80.00 0.80 0.45 0.20Upper molar 2 0.00 0.00 0.00 100.00 0.00 0.00 0.00 0.00

Tragelaphus nakuae 11 0.00 0.00 0.00 27.27 72.73 0.73 0.47 0.14Lower molar 3 0.00 0.00 0.00 0.00 100.00 1.00 0.00 0.00Upper molar 8 0.00 0.00 0.00 37.50 62.50 0.63 0.52 0.18

(continued on next page)

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

106

Page 5: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

30%–70% of mono- and dicots plus< 20% fruits, and generalists in-cluding>20% of all food type); browsers (Br:> 70% dicots), andfrugivores (Fr:> 70% fruits).

3. Methods

3.1. Mesowear analysis

Although the entire surface of a tooth is affected by wear, mesowearanalysis is focused on the topography of cutting edges of the molarenamel surfaces. Mesowear characterizes the occlusal relief and cuspshape on the cutting edge of either the labial cusps of upper molars orthe lingual cuspids of lower molars (Fortelius and Solounias, 2000;Franz-Odendaal and Kaiser, 2003). Because tooth wear makes cuspsblunter and not sharper, both mesial and distal cusp(id)s were ex-amined and only the sharpest cusp(id) was scored (Fortelius andSolounias, 2000). The mesowear method originally examined only thesecond molar, because previous studies demonstrated that a change inmesowear signal existed along the toothrow in relation to the degree oftooth wear (Fortelius and Solounias, 2000; Louys et al., 2011; see Kaiseret al., 2013 for a recent review). However, we also included thirdmolars (M3) and undetermined lower and upper molars (LM or UM) forwhich carbon isotope data were available (Bibi et al., 2013; Negashet al., 2015) in order to correlate the dental wear signal with isotopicsignal, thus providing complementary information on the ecology ofthese specimens.

In mesowear analysis, occlusal relief (OR) captures the height of thecusp(id)s with respect to the central valley between them. OR wasscored as a categorical variable with two levels: high (H) or low (L) (formethod details, see Fortelius and Solounias, 2000; Merceron et al.,2007; Blondel et al., 2010). The second mesowear variable, cusp(id)shape (CS), included three scoring attributes: sharp (S), round (R), orblunt (B) according to the degree of facet development (see detailedmethod in Fortelius and Solounias, 2000; Blondel et al., 2010). We useda combined occlusal relief and cusp(id) (OR-CS) score, a univariatemeasure representing a continuum of mesowear stages from 0 to 4(Fig. 2), calculated following Kaiser et al. (2009; see also Kaiser et al.,2013). We refer to this as the mesowear score (MWS). For MWS, a scoreof 0 is given to high and sharp cusps (HS), 1 to high and rounded cusps(HR), 2 to low and sharp cusps (LS), 3 to low and rounded cusps (LR),and 4 to low and blunt cusps (LB). MWS were then averaged for eachspecies (Table 1).

3.2. Dental microwear textural analysis

DMTA was preferentially performed on the paracone lingual enamel

facets and the protoconid buccal enamel facets of the second upper andlower molars respectively, but the analysis was extended to isolatedthird or first molars to increase the sample size (Schulz et al., 2010).Teeth were cleaned with acetone to remove dust and glue used in thefield or in the lab for specimen stabilization. After cleaning, dental fa-cets were molded with a silicone dental molding material, poly-vinylsiloxane (Coltène Whaledent, President Regular Body; Goodallet al., 2015). Rather than producing a transparent resin-based cast, themolds were directly scanned using a white-light confocal DCM8 LeicaMicrosystems surface profilometer (housed at PALEVOPRIM) producinginverse (negative) 3D models (point clouds) of wear surfaces. Theprofilometer was equipped with 100× lens (NA=0.90; working dis-tance= 0.9mm) and a 333× 251 μm area was scanned for each spe-cimen. From this, four sub-adjacent surfaces (140×100 μm;1088×776 pixels) were generated and saved as .plμ files. This proce-dure allows comparisons with Scott's (2012) comparative dataset. Thelateral sampling interval was 0.129 μm and that of vertical spacing was0.002 μm. Using LeicaMap 7.4 (Leica Microsystems), scans mirrored onthe z-axis render the positive surface of the dental wear facets. For>95% of the fossil specimens, non-measured points represent< 1% ofthe surface. All other scans have fewer than 3% missing points. Thesemissing points were replaced with a smooth shape calculated fromneighboring points. Artifacts such as aberrant peaks, adhering dust, orsediment particles on the enamel surface (S3) were removed using a3 μm-diameter eraser (or a user-defined contour eraser in case of largerartifacts) and then replaced with a smooth shape calculated from theneighboring points. A final leveling was then performed on the re-sulting surfaces and the result was saved as .sur files (Figs. 3, 4 and 5,S3). Further analysis to generate DMTA variables was conducted withToothfrax and Sfrax (Surfract, www.surfract.com) following proceduresand settings shown in Scott et al. (2006).

Three microwear variables are used in this study (Table 3, S2).Complexity (Asfc or Area-scale fractal complexity) is a measure of theroughness at a given scale. Anisotropy (epLsar or exact proportion oflength-scale anisotropy of relief) measures the orientation concentra-tion of surface roughness. Heterogeneity of complexity (HAsfc or het-erogeneity of area-scale fractal complexity) quantifies the variation ofcomplexity observed within each scan. HAsfc is calculated in each140× 100 μm sub-surface through 81 cells (HAsfc81-cells). All variableshave been described in further details in Scott et al. (2006).

3.3. Statistical analysis of dental wear

Statistical tests were used to analyze differences in mesowear scores(MWS) and microwear textural parameters between groups (both spe-cies and Shungura Formation members). As microwear textural

Table 1 (continued)

Samples n LS LR LB HS HR MWS

m sd sem

Member FTragelaphus gaudryi 12 0.00 0.00 0.00 25.00 75.00 0.75 0.45 0.13

Lower molar 5 0.00 0.00 0.00 20.00 80.00 0.80 0.45 0.20Upper molar 7 0.00 0.00 0.00 28.57 71.43 0.71 0.49 0.18

Tragelaphus nakuae 16 0.00 0.00 0.00 56.25 43.75 0.44 0.51 0.13Lower molar 8 0.00 0.00 0.00 50.00 50.00 0.50 0.53 0.19Upper molar 8 0.00 0.00 0.00 62.50 37.50 0.38 0.52 0.18

Member GTragelaphus gaudryi 29 3.45 0.00 0.00 48.28 48.28 0.55 0.57 0.11

Lower molar 4 0.00 0.00 0.00 50.00 50.00 0.50 0.58 0.29Upper molar 25 0.00 0.00 4.00 48.00 48.00 0.56 0.58 0.12

Tragelaphus nakuae 52 3.85 1.92 0.00 40.38 53.85 0.69 0.64 0.09Lower molar 16 0.00 0.00 0.00 31.25 68.75 0.75 0.45 0.11Upper molar 36 5.56 2.78 0.00 44.44 47.22 0.67 0.72 0.12

Tragelaphini indet. 1 0.00 0.00 0.00 0.00 100.00 1.00

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

107

Page 6: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

parameters violated conditions for parametric tests, they were rank-transformed before each analysis (Conover and Iman, 1981; Sokal andRohlf, 1969). One-way analysis of variance (ANOVA) for each para-meter was used to determine the sources of significant variation. Dif-ferences between groups were then highlighted using the combinationof the conservative HSD test (Tukey's Honest Significant Differences)together with the less conservative LSD test (Fisher's Least SignificantDifferences; Tables 2 and 4). All statistics were computed using Sta-tistica v.7.

3.4. Body mass estimates for the Tragelaphus nakuae lineage

We used the equations for lower molar rows provided by Janis(1990) to estimate body mass for fossil specimens in kilograms (kg).

Molar rows were measured using digital calipers on specimens from theShungura and Hadar Formations, the Middle Awash, and Woranso-Mille (data from Geraads et al., 2009). These sites encompass all of themajor Plio-Pleistocene fossil-bearing localities where dental remains ofthe T. rastafari-T. nakuae lineage have been identified. The mid-Plio-cene Hadar Formation at Hadar, Ethiopia, includes T. rastafari (Reedand Bibi, 2011), while slightly younger sediments in the Middle Awash,Ethiopia, have yielded remains of T. nakuae (Bibi, 2011). Early to mid-Pliocene sediments at Woranso-Mille, Ethiopia, have yielded T. saraitu,the likely ancestor of T. rastafari (Geraads et al., 2009), and thereforewe include these here as the earliest phase of the T. rastafari-T. nakuaelineage.

Fig. 2. Examples of cusps corresponding to mesowear scores (MWS, Kaiser et al., 2009) on upper molars of Tragelaphus nakuae. A. MWS=0, high and sharp (HS), L 625-25; B. MWS=1,high and round (HR), OMO 47-1968-2536; C. MWS=2, low and sharp (LS), OMO 50/3-1968-1625; D. MWS=3, low and round (LR), OMO 47-1968-2612.

Fig. 3. Examples of photosimulations of dental microwear textures on second molars of modern species were generated through the LeicaMap 7.4.: the grazing hartebeest Alcelaphusbuselaphus (A; PALEOPRIM ART-8-21), the mixed feeder red deer Cervus elaphus (B; PALEVOPRIM-9-4152), the frugivorous-browser yellow back duiker Cephalophus silvicultor (C; NHMB-Z3347) and the leaf browsing giraffe Giraffa camelopardalis (D; SNG-36628).

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

108

Page 7: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

4. Results

The results presented below are based on fossil material frommultiple localities that sample a homogeneous geological context (i.e.,fluviatile, with a predominance of fossils from sandstone deposits),which was continuous through most of the Shungura Formation. Theresults do not include fossil material found in anthropogenic contexts(i.e., archaeological assemblages). Both macro- and micro-scopic as-pects of the fossil teeth suggest no difference in preservation throughtime or across space. Therefore, the interpretation of these results isunlikely to be biased by differing geologic or taphonomic effects.

4.1. Dental mesowear score (MWS) results

4.1.1. Comparisons between extant taxa with known dietsANOVAs detect significant differences in MWS between the four

dietary categories (Table 2A). Post-hoc tests indicate that browsers,frugivores, mixed feeders, and grazers significantly differ from eachother (Table 2B). The lowest mesowear score (MWS) is obtained bybrowsing species, followed by mixed feeders, frugivores, and grazers(Fig. 6, Table 1). Extant tragelaphins differ from modern grazers andfrugivores in having lower MWS. The MWS of Tragelaphus eurycerus andT. scriptus (Table 1: 0.59 and 0.51, respectively) reflect their browsingpreferences (Gagnon and Chew, 2000 and citations therein). Tragela-phus angasii, T. imberbis, and T. oryx have a mixed diet (Gagnon andChew, 2000) but their MWS does not differ from that of the two speciesmentioned above (Table 1: 0.65, 0.42, and 0.50, respectively).

Fig. 4. Examples of photosimulations of dental microwear textures on molars of extinct tragelaphins were generated through the LeicaMap 7.4.: OMO20-2-10003 from Member B (A), L444-10003 from Member C (B), OMO174-10009 from Member F (C), L 881-2b2 from Member G (D).

Fig. 5. 3D simulation and photosimulation of dental microwear textures on molars of OMO tragelaphins with taphonomic alterations (sediments).

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

109

Page 8: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

4.1.2. Comparisons between extant species and tragelaphins from theShungura Formation

The average MWS of the entire Tragelaphini sample from ShunguraFormation (n=161; mean= 0.68 ± 0.05; Table 1) falls between thatof extant browsers (0.47) and mixed feeders (0.81, Table 1, Fig. 6). Thisvalue is slightly higher than the average MWS of the entire extanttragelaphin sample (n= 139; mean= 0.53 ± 0.04), which are

classified as browsers and mixed feeders. No significant difference isdetected between T. rastafari, T. nakuae, and T. gaudryi (Table 2B) whenall members are combined. The post-hoc tests show that T. rastafaridoes not differ significantly from any of the four dietary categoriesbased on data from the extant species. Tragelaphus gaudryi (n= 49) andT. nakuae (n= 102) differ significantly from modern grazers and fru-givores (Table 2B). Tragelaphus nakuae differs significantly frommodern browsers with the LSD test (Table 2B).

Post-hoc tests indicate no significant differences betweenTragelaphini from Member B through Member G in MSW (Table 2C). InMember C, tragelaphins (almost all identified as T. nakuae) show sig-nificant differences from extant browsers and grazers but do not differfrom mixed feeders and frugivores. In Member E, F and G, tragelaphinsshow significant differences in MSW compared to extant grazers andfrugivores (Table 2C). Overall, the mesowear data indicate that trage-laphins from Member C through Member G were likely browsers andmixed feeders (Fig. 6), as are their living counterparts. The results for T.rastafari from Member B should be considered inconclusive due to thesmall sample size for this member.

4.2. Dental microwear texture analysis (DMTA) results

4.2.1. Comparisons between extant species with known dietsANOVAs indicate significant differences for the three dental mi-

crowear texture parameters, complexity (Asfc), anisotropy (epLsar),heterogeneity of complexity (HAsfc81cells) (Table 4A, Fig. 7). No sig-nificant differences in complexity and in anisotropy exist between thethree (C4 and C3) grazers Alcelaphus buselaphus, Bos taurus, and Synceruscaffer (Table 4B, Fig. 7). Alcelaphus buselaphus differs in heterogeneityof complexity (HAsfc81cells) from S. caffer, and A. buselaphus and B.taurus have significant lower complexity (Asfc) than Cervus elaphus,Giraffa camelopardalis, and Cephalophus silvicultor (Fig. 8A, Table 4B).Alcelaphus buselaphus and S. caffer differ from the three latter taxa inhaving higher anisotropy, and B. taurus from the two browsing speciesin having higher anisotropy (Fig. 8B). Syncerus caffer also has lower

Table 2A. ANOVA comparison of diet categories and extinct species and between diet categories andTragelaphini in each member. B. Results of pairwise comparisons of extant diet categories andextinct species. C. Results of pairwise comparisons of extant diet categories and Tragelaphini in eachmember. (df, degrees of freedom; SS, sum of square; MS, mean square; F, Fisher value; p, p-value.)Significance at p < 0.05 is indicated in normal black for Fisher's LSD and in bold for both Tukey'sHSD and Fisher's LSD tests.

A

MWS

ANOVA df SS MS F pDiet

categories/species

Effect 6 223 37 43.48 <0.001 Error 1474 1258 1

Diet

categories/member

Effect 8 224 28 32.82 <0.001 Error 1476 1258 1

B

Br Mf Gr Fr Tg Tn Tr Browser Br

Mixed feeder MWS Mf

Grazer MWS MWS Gr

Frugivore MWS MWS MWS Fr

T. gaudryi MWS MWS Tg T. nakuae MWS MWS MWS Tn T. rastafari Tr

C

Br Mf Gr Fr TMbB TMbC TMbE TMbF TMbG

Browser Br

Mixed Feeder MWS Mf

Grazer MWS MWS Gr

Frugivore MWS MWS MWS Fr

Tragelaphini Member B TMbB

Tragelaphini Member C MWS MWS TMbC

Tragelaphini Member E MWS MWS TMbE

Tragelaphini Member F MWS MWS TMbF

Tragelaphini Member G MWS MWS TMbG

Fig. 6. Box-and-whiskers plots of mesowear scores (MWS) for each dietary category (Br,browsers; Fr, frugivores; Gr, grazers; Mf, mixed feeders, Gagnon and Chew, 2000) andTragelaphini (Tg, Tragelaphus gaudryi; Tn, Tragelaphus nakuae; Tr, Tragelaphus rastafari)from the Formation Shungura members (Member B, C, E, F and G). Boxplots represent thestandard deviation and whiskers indicate the confidence interval of the mean at 95%.

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

110

Page 9: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

complexity than C. silvicultor. Cervus elaphus has higher anisotropy thanG. camelopardalis and C. silvicultor; furthermore, C. elaphus differs sig-nificantly in four parameters from A. buselaphus, and differs in hetero-geneity of complexity (HAsfc81cells in complement to Asfc) from B.taurus. Giraffa camelopardalis significantly differs from all other speciesin having lower anisotropy (epLsar, Fig. 8B and Table 4B) and differsfrom C. elaphus in having lower heterogeneity of complexity. Cephalo-phus silvicultor differs from the three grazers and the mixed feeder inhaving higher complexity and lower anisotropy (Fig. 8A, B). The onlydifference between C. silvicultor and G. camelopardalis is in the loweranisotropy of C. silvicultor (Fig. 8, Table 4B).

Complexity and anisotropy are the most effective for discriminatingthe five extant species from one another based on their diets, as shownin Fig. 7. This bivariate plot indicates that grazers have higher aniso-tropy and lower complexity than the browser and frugivore. The mixedfeeder displays intermediate values between grazers and browsers. Thefrugivore has higher complexity than grazers, mixed feeders, andbrowsers. These results are congruent with previous DMTA of extantAfrican bovids (Ungar et al., 2007; Scott, 2012).

4.2.2. Comparisons between extant species and Tragelaphini from theShungura Formation

No significant differences are detected between T. rastafari, T. na-kuae, and T. gaudryi (Table 4B) for all dental microwear texture para-meters, with the exception that T. rastafari differs from T. gaudryi inhigher heterogeneity according to the LSD test (Table 4B). All threetragelaphin species do not differ significantly from the leaf browsing G.camelopardalis (Table 4B). They have significantly lower anisotropyvalues than the grazers (A. buselaphus, B. taurus, S. caffer) and the mixedfeeder (C. elaphus). T. nakuae and T. gaudryi have significantly lowercomplexity values than the frugivorous Cephalophus silvicultor (Fig. 7,Table 4B).

Member B tragelaphins, represented by T. rastafari, are not sig-nificantly different from those in members C, F, and G, but the LSD testshows that they differ from T. nakuae in Member E in anisotropy,complexity, and heterogeneity of complexity (Table 4C, Fig. 8).Member C tragelaphins (T. nakuae) differ from those in members E, F,

and G in having a higher heterogeneity of complexity. Moreover, theyhave a higher complexity than tragelaphins in members E and G and alower anisotropy than those in Member E (Table 4C, Fig. 8A, B). Nosignificant difference is detected between tragelaphins from membersE, F, and G (Table 4C, Fig. 8).

Tragelaphins in members B, C, and F are not significantly differentfrom G. camelopardalis (Table 4C). Member E tragelaphins differ fromthe leaf browsing G. camelopardalis and frugivorous C. silvicultor inhaving lower complexity and lower heterogeneity of complexity(Fig. 8A, C). They differ from grazers in having lower anisotropy(epLsar) and from the mixed feeder in anisotropy and heterogeneity ofcomplexity (Table 3, Fig. 8B, C). Based on the LSD test, the only sig-nificant difference between tragelaphins represented by T. nakuae inMember G and G. camelopardalis is in complexity (Table 4C). T. nakuaein Member G has lower anisotropy than the three extant grazers anddiffers from C. elaphus in anisotropy and heterogeneity of complexity(Fig. 8B, C).

4.3. Body mass evolution in the T. nakuae lineage

There is a significant positive trend (r=0.36, p < 0.001) in bodymass (kg) estimated from molar rows for the Tragelaphus nakuae lineagefrom ~3.6 to 2.0 Ma across eastern Africa (Fig. 10). The smallest spe-cimens are those of the early to mid-Pliocene T. saraitu(187 ± 19.8 kg), followed by those of mid-Pliocene T. rastafari(229 ± 38.6 kg) and late Pliocene and Pleistocene T. nakuae(269 ± 46.7 kg). Among living tragelaphins, the most similar in bodymass to T. saraitu are the mountain nyala T. buxtoni (~175–240 kg) andgreater kudu T. strepsiceros (~160–250 kg). The later T. rastafari and T.nakuae are larger tragelaphins (Fig. 10) and are similar in body mass tomale T. strepsiceros and the bongo T. eurycerus (~230–360 kg), the latterof which is closely related to the T. rastafari-nakuae lineage (Bibi,2011). Some Pleistocene specimens of T. nakuae are exceptionally large(e.g., L 460-10002, estimated at 380.5 kg) and fall within the range ofeland T. oryx (~344–494 kg).

Fig. 7. Distribution (mean, 95% confidence interval of the mean)of the extant species (Ab, Alcelaphus buselaphus; Bt, Bos taurus;Cs, Cephalophus silvicultor; Ce, Cervus elaphus; Gc, Giraffa came-lopardalis; Sc, Syncerus caffer) and T, Tragelaphini; Tn,Tragelaphus nakuae; Tg, Tragelaphus gaudryi; Tr, Tragelaphus ras-tafari of the Shungura Formation depending on the complexity(Asfc) and anisotropy (epLsar). Orange: extant grazers; lightgreen: mixed feeders; dark green: extant browsers; red: extantfrugivores, black: extinct species. (For interpretation of the re-ferences to color in this figure legend, the reader is referred tothe web version of this article.)

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

111

Page 10: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

5. Discussion

5.1. Feeding preferences of Shungura Tragelaphini

Our combination of dental mesowear and microwear textural ana-lyses indicates that the Shungura Formation tragelaphins Tragelaphusrastafari, T. gaudryi, and T. nakuae had diets dominated by browseoverall. However, MWS data were more variable and indicated a rangeof dietary preferences from browsing, such as that of extant T. eurycerus(Hillman and Gwynne, 1987; Gagnon and Chew, 2000), to mixedfeeding like extant T. angasii (Hofmann, 1973; Gagnon and Chew,2000). DMTA data, conversely, suggest that the Shungura fossil trage-laphins were browsers when compared to the modern species analyzedhere. The slight discrepancy between mesowear and microwear sig-natures may be explained by differences in the temporal resolution ofthese proxies (see Section 1.2; Davis and Pineda-Munoz, 2016). Indeed,Merceron et al. (2007) showed that short-term (i.e., microwear) andlong-term (i.e., mesowear) signals can vary in mixed feeders. However,the discrepancy may be due to the choice of extant species in our

comparative dataset or the dietary categories we used. For example,using the dietary categories for African bovids defined by Gagnon andChew (2000), the Shungura tragelaphins fall among the modernbrowser-grazer intermediates, generalists, and browsers based on theextant bovid microwear analyses of Scott (2012). This is consistent withour mesowear analyses and suggests that the slight discrepancy be-tween mesowear and microwear results here may be more analyticalthan biological. Overall, Shungura tragelaphins had similar diets tothose of living tragelaphins classified by Gagnon and Chew (2000).

In Member B Tragelaphus rastafari has high values for complexityand low values of anisotropy, and is similar to browsers (this study;Scott, 2012) that consume>70% dicots, such as T. eurycerus (Gagnonand Chew, 2000). Extant bongos have a diet of high-protein vegetation,such as young leaves and flowers with some seasonal grasses ac-counting for up to 20% of their diet (Estes, 1991; Elkan and Smith,2013). The heterogeneity of T. rastafari is higher than that of T. gaudryiand the significant difference between T. nakuae from Member E and T.rastafari indicates that this species likely consumed more browse thanT. nakuae and T. gaudryi. Tragelaphus rastafari was similar in size (see

Fig. 8. Box-and-whiskers plots of microwear texture data of extant species (Ab, Alcelaphus buselaphus; Bt, Bos taurus; Cs, Cephalophus silvicultor; Ce, Cervus elaphus; Gc, Giraffa camelo-pardalis; Sc, Syncerus caffer) and Tragelaphini for each member of the Shungura Formation (Member B, C, E, F and G). A. mean complexity (Asfc); B. mean anisotropy (epLsar); C. meanheterogeneity 9× 9 (HAsfc81). Boxplots represent the standard deviation and whiskers indicate the confidence interval of the mean at 95%. (For interpretation of the references to colorin this figure legend, the reader is referred to the web version of this article.)

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

112

Page 11: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

below) and thus may have been ecologically similar to T. eurycerus, aselective browser. This conclusion is based solely on microwear dataand awaits confirmation from a larger mesowear sample.

T. rastafari is considered to be the immediate ancestor of T. nakuae,with the replacement of the former by the latter in the Omo-TurkanaBasin occurring between ~2.95Ma and 2.74Ma (Unit B10-C4; Bibi,2011). Based on MWS statistical tests, T. nakuae does not significantlydiffer from mixed feeders such as T. angasii (Table 1), which is de-scribed as a generalist species (Hofmann, 1973; Gagnon and Chew,2000). Among mixed feeders, generalists differ from browser-grazerintermediates by feeding on more fruits and seeds, while mixed feeders(browser-grazer intermediates) focus more on dicots and monocot fo-liage (Gagnon and Chew, 2000 and citations therein). In contrast tomesowear data, the dental microwear textures of T. nakuae fromMember C through G suggest browse-dominated diets. Combined to-gether, the dental mesowear and dental microwear data suggest thatthe Shungura specimens of T. nakuae were mixed feeders preferringbrowse, such as extant T. angasii and T. strepsiceros. This differs from itsclosest living relative, T. eurycerus, which is more similar to the geo-logically older T. rastafari in diet.

In members E, F, and G, T. gaudryi does not differ from mixed fee-ders and browsers based on MWS. DMTA indicates that T. gaudryi fromMember F (n=8) falls within browsers but has lower complexity va-lues and higher anisotropy values than those of T. nakuae and T. ras-tafari (Fig. 4). T. gaudryi is probably closely related to the modern T.imberbis, the lesser kudu (Gentry, 1985; Bibi, 2011). Observational dataindicate that the lesser kudu is a mixed feeding species, classified within“browser-grazer intermediates” by Gagnon and Chew (2000), andmainly consumes leaves of trees and shrubs with about 30% of her-baceous monocots (grasses, sedges; Hofmann and Stewart, 1972). Ourdata indicate that the dietary niche of T. gaudryi would have been closeto that of its descendant T. imberbis. However, it should be noted thatcarbon stable isotope work (n= 4, Cerling et al., 2003) suggests that T.imberbis is an obligate browser and therefore T. gaudryi would haveconsumed a greater proportion of monocots compared to its descen-dant. Considering the small size and limited geographic extent of theisotope sample, we prefer to rely on the observational data for the time

being.

5.2. Comparisons with enamel δ13C values

The carbon isotopic composition of fossil tooth enamel (δ13C) in-dicates the proportion of C3 versus C4 plants an individual consumedduring their life. C3 plants comprise almost all tree species, most shrubs,and temperate/cold or high-altitude herbaceous monocots (Vogel,1978). C4 species comprise 79% of monocots and include grasses(Poaceae) and sedges (Cyperaceae), as well as 21% of dicots such asAmaranthaceae ss, Chenopodiaceae ss and Euphorbiaceae (Sage andMonson, 1999; Sage et al., 2011). C4 monocots have a wide geo-graphical distribution and dominate warm-climate grasslands and sa-vannas, whereas the less abundant C4 dicots are present in warm, semi-arid and arid regions (Cerling et al., 1997; Ehleringer et al., 1997;Edwards et al., 2010; Sage and Monson, 1999; Sage et al., 2011).

Stable carbon isotopic values of tooth enamel of the T. rastafari-nakuae lineage from the Shungura Formation (Bibi et al., 2013; Negashet al., 2015) are higher than recorded for any living tragelaphin sam-pled from southern Africa (Sponheimer et al., 2003) or Kenya (Cerlinget al., 2003; Cerling et al., 2015). δ13C values for living tragelaphinssuggest a diet strongly dominated by C3 plants (−13.4‰ and −11.5‰in Cerling et al., 2003 and in Sponheimer et al., 2003, respectively),whereas mean δ13C values for T. rastafari-nakuae (−4‰ and −5‰ inBibi et al., 2013 and Negash et al., 2015, respectively) indicate a dietwith a significantly higher component of C4 plants. The fossil data,however, show a great range of variation (> 10‰), spanning the rangefrom C3 to C4 diets. Negash et al. (2015) found that tragelaphins fromMember B alone range from −11.6‰ to −1.8‰ (n= 4), although thisresult was not found in the Member B specimens (n= 7) analyzed byBibi et al. (2013).

The great variation in δ13C values among individuals belonging tothe same species found within a single locality (e.g., OMO 28, OMO 4;Table 5) is unexpected. We may hypothesize that such inter-individualvariation results from the analysis of different molars with unknownposition along the tooth row. M1, M2, and M3 mineralize sequentiallyat different time intervals (notably before, during, and after weaning

Table 3Descriptive statistics (m, mean; med, median; sd, standard deviation and sem, standard deviation of the mean) of dental microwear texture parameters for extant species and extinct taxa.

Samples n Asfc epLsar (10−3) Hasfc81cells

m med sd sem m med sd sem m med sd sem

Alcelaphus buselaphus (Gr) 28 1.60 1.41 0.77 0.14 5.80E−03 5.60E−03 1.6E−03 3.11E−04 0.66 0.62 0.18 0.03Bos taurus (Gr) 6 1.37 1.26 0.53 0.22 5.77E−03 5.67E−03 1.4E−03 5.58E−04 0.70 0.57 0.37 0.15Cephalophus silvicultor (Fr) 25 4.03 3.06 3.25 0.65 3.07E−03 2.88E−03 1.6E−03 3.17E−04 1.16 0.77 1.38 0.28Cervus elaphus(Mf) 29 2.12 1.86 0.80 0.15 4.59E−03 4.91E−03 1.9E−03 3.55E−04 1.07 0.90 0.44 0.08Giraffa camelopardalis (Br) 12 2.51 2.01 1.48 0.43 2.10E−03 1.75E−03 1.6E−03 4.54E−04 0.84 0.82 0.36 0.10Syncerus caffer (Gr) 6 1.89 2.07 0.58 0.24 6.02E−03 5.53E−03 2.0E−03 8.07E−04 1.08 1.12 0.44 0.18Fossil taxaTragelaphini 88 2.28 1.73 1.96 0.21 2.43E−03 2.01E−03 1.4E−03 1.44E−04 0.74 0.63 0.43 0.05

Tragelaphus gaudryi (Tg) 10 1.94 1.64 1.01 0.32 2.75E−03 2.56E−03 1.3E−03 4.16E−04 0.53 0.55 0.13 0.04Tragelaphus nakuae (Tn) 57 2.12 1.77 1.58 0.21 2.43E−03 2.03E−03 1.4E−03 1.87E−04 0.76 0.63 0.44 0.06Tragelaphus rastafari (Tr) 17 3.14 2.43 3.20 0.78 2.09E−03 1.82E−03 1.0E−03 2.48E−04 0.84 0.87 0.47 0.12

Member B (Tr) 17 3.14 2.43 3.20 0.78 2.09E−03 1.82E−03 1.0E−03 2.48E−04 0.84 0.87 0.47 0.12Member C 22 2.64 2.12 2.06 0.44 1.90E−03 1.59E−03 1.2E−03 2.45E−04 0.96 0.77 0.57 0.12

Tragelaphus gaudryi (Tg) 1 0.50 4.93E−03 0.42Tragelaphus nakuae (Tn) 20 2.59 1.96 2.16 0.48 1.96E−03 1.71E−03 1.2E−03 2.66E−04 0.96 0.77 0.58 0.13Tragelaphini 1 3.45 1.29E−03 1.34

Member E 18 1.59 1.40 0.87 0.20 3.26E−03 2.79E−03 1.6E−03 3.85E−04 0.52 0.50 0.15 0.04Tragelaphus gaudryi (Tg) 1 0.50 4.93E−03 0.42Tragelaphus nakuae (Tn) 15 1.72 1.50 0.89 0.23 3.07E−03 2.53E−03 1.6E−03 4.16E−04 0.54 0.53 0.16 0.04Tragelaphini 2 1.14 1.14 0.01 0.01 3.87E−03 3.87E−03 2.2E−03 1.57E−03 0.41 0.41 0.06 0.04

Member F 15 2.26 1.75 1.40 0.36 2.33E−03 1.90E−03 9.9E−04 2.55E−04 0.62 0.63 0.24 0.06Tragelaphus gaudryi (Tg) 8 2.02 1.64 0.96 0.34 2.66E−03 2.56E−03 1.1E−03 3.91E−04 0.53 0.55 0.13 0.05Tragelaphus nakuae (Tn) 6 2.80 2.15 1.85 0.75 1.71E−03 1.64E−03 4.0E−04 1.61E−04 0.78 0.79 0.28 0.11Tragelaphini 1 0.98 3.40E−03 0.44

Member G (Tn) 16 1.66 1.48 0.81 0.20 2.68E−03 2.61E−03 1.5E−03 3.72E−04 0.70 0.61 0.37 0.09

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

113

Page 12: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

stage for M1) and therefore we expect their carbon isotope compositionto reflect changes in diet across years or within years (Kohn, 2004;Zazzo et al., 2010). Identifying the position of the analyzed tooth is thuscritical, albeit overlooked, in isotope studies and in particular for ex-tinct tragelaphins whose mixed feeding diet varies more than those ofeither strictly grazing or browsing tribes.

Table 5 combines the meso- and micro-wear data of individuals withtheir published enamel δ13C values. In most cases the wear signature isconsistent with the isotopic signal in the diet that would be inferred,although this is not always the case. For instance, the OMO 28-1967-463 specimen of T. rastafari has a δ13C value of−1.8‰ suggesting a C4-dominated diet whereas its dental wear pattern indicates purebrowsing. Although the dental wear analyses of T. rastafari and T. na-kuae indicate they were browsers, the stable isotope analysis reveals

that they also ingested significant quantities of C4 plants (most likelygrasses) when their teeth formed.

It is often considered that in modern intertropical zones, browsersprimarily consume C3 dicots while grazers consume C4 monocots (Lee-Thorp, 1989; Lee-Thorp and van der Merwe, 1987; Cerling and Harris,1999). Our findings for the tragelaphins from the Shungura Formationare consistent with previous studies of fossil tragelaphins from the mid-to late Pliocene of Laetoli, Tanzania (Upper Laetolil and Upper Ndo-lanya Beds). At Laetoli, the mesowear scores indicate a browsing dietwhereas carbon isotope values suggest a C3–C4 diet (Kaiser, 2011;Kingston, 2011). This apparent discrepancy may be attributable todifferences in dental wear and biogeochemical approaches. First, wearpattern and enamel isotopes of a single individual rarely record asynchronous dietary signal (Grine et al., 2012; Davis and Pineda-

Table 4A. ANOVA comparison of extant species and extinct taxa. B. Results of pairwise comparisons of extant species and extinct taxa. C. Results of pairwise comparisons of extant species andTragelaphini in each member. Significance at p < 0.05 is indicated in normal black for Fisher's LSD and in bold for both Tukey's HSD and Fisher's LSD tests.

A

DMTA

ANOVA species df SS MS F p

Asfc Effect 8 109923 13740 5.16 <0.001Error 179 476911 2664

epLsar Effect 8 264328 33041 18.06 <0.001Error 179 327522 1830

HAsfc81 Effect 8 96472 12059 4.47 <0.001Error 179 482743 2697

ANOVA member df SS MS F p

Asfc Effect 10 135657 13566 5.22 <0.001Error 181 470741 2601

epLsar Effect 10 282926 28293 16.09 <0.001Error 181 318229 1758

HAsfc81 Effect 10 134890 13489 5.22 <0.001Error 181 467538 2583

B

Ab Bt Sc Ce Gc Cs Tg Tn TrAlcelaphus buselaphus AbBos taurus BtSyncerus caffer HAsfc Sc

Cervus elaphusAsfc-epLsar-Hasfc Asfc-HAsfc epLsar Ce

Giraffa camelopardalis Asfc-epLsarAsfc-epLsar epLsar epLsar-HAsfc Gc

Cephalophus silvicultor Asfc-epLsar-HAsfc Asfc-epLsar Asfc-epLsar Asfc-epLsar-HAsfc epLsar Cs

T. gaudryi epLsar epLsarepLsar-HAsfc epLsar-HAsfc Asfc-HAsfc Tg

T. nakuae epLsar epLsarepLsar-HAsfc epLsar-HAsfc Asfc-HAsfc Tn

T. rastafari Asfc-epLsarAsfc-epLsar epLsar epLsar-HAsfc Asfc-epLsar HAsfc Tr

C

Ab Bt Sc Ce Gc Cs TMbB TMbC TMbE TMbF

Alcelaphus buselaphus AbBos taurus BtSyncerus caffer HAsfc Sc

Cervus elaphus

Asfc-epLsar-HAsfc Asfc-HAsfc epLsar Ce

Giraffa camelopardalis Asfc-epLsar Asfc-epLsar epLsar epLsar-HAsfc Gc

Cephalophus silvicultor

Asfc-epLsar-HAsfc Asfc-epLsar Asfc-epLsar

Asfc-epLsar-HAsfc epLsar Cs

Tragelaphini Member B Asfc-epLsar Asfc-epLsar epLsar epLsar-HAsfcAsfc-epLsar TMbB

Tragelaphini Member C

Asfc-epLsar-HAsfc Asfc-epLsar epLsar epLsar

Asfc-epLsar TMbC

Tragelaphini Member EepLsar-HAsfc epLsar

epLsar-HAsfc epLsar-HAsfc Asfc-epLsar-HAsfc

Asfc-HAsfc

Asfc-epLsar-HAsfc

Asfc-epLsar-HAsfc TMbE

Tragelaphini Member F epLsar epLsarepLsar-HAsfc epLsar-HAsfc Asfc-HAsfc Hasfc TMbF

Tragelaphini Member G epLsar epLsarepLsar-HAsfc

Asfc-epLsar-HAsfc Asfc

Asfc-HAsfc Asfc-HAsfc

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

114

Page 13: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

Munoz, 2016), possibly due to the sampling protocol used. In isotopestudies the enamel may be sampled all along the tooth growth axis(Negash et al., 2015), which effectively averages the input signal(Balasse, 2003; Passey and Cerling, 2002). Such protocols diminish theintra-individual variations in δ13C along the mesodont molar crowns oftragelaphins that reflect seasonal changes in their diet (Table 5). Inaddition, dental microwear textures are known to have a ‘last suppereffect’ in which they reflect diet over weeks or days (Merceron et al.,2010; Teaford and Oyen, 1989). Second, dental microwear textures arecorrelated with the physical properties (hardness, toughness, abra-siveness) of food items ingested by individuals (Lucas, 2004; Scott,2012), whereas the stable carbon isotopic composition of enamel re-flects the proportion of C3 vs. C4 photosynthetic pathways among plantsconsumed. Both signals are complementary when inferring paleodietsand need not be viewed as contradictory as they relate to differentaspects of plant anatomy and/or composition.

Finally, previous studies (Kaiser, 2011; Bibi et al., 2013) have sug-gested that the C4 signal in enamel of Pliocene tragelaphins may be dueto C4 dicots rather than C4 monocots. Pollen data indicate that theproportion of C4 dicots (e.g. Amaranthaceae and Chenopodiaceae)

fluctuated throughout the Plio-Pleistocene in Africa with a peak at ca2.6 Ma corresponding to increased aridity (Bonnefille and Dechamps,1983; Bonnefille, 2010). Amaranthaceae comprises herbs, shrubs andsome succulents. Due to a shoot Si concentration half as high asmonocots (Hodson et al., 2005), Amaranthaceae (which are C4 dicots)have structural support and physical properties differing from those ofC4 monocots. This contributes to different dental wear patterns, andnotably a weaker “grazing” signature, meaning lower anisotropy thanC4 grass eaters. Extant tragelaphins are known to consume some suc-culents of Amaranthaceae and shoots of herbaceous monocots (Estes,1991; Kingdon, 2013). We hypothesize that these succulents (C4 dicots)during dry seasons and peak aridity and fresh C4 grass shoots during thewet season may have been consumed in higher proportions by extincttragelaphins. This potentially accounts for the discrepancy between theisotopic data and dental microwear textures.

5.3. From feeding preferences to habitats of Tragelaphini through theShungura Formation

While most δ13C values of tooth enamel show a dietary shift (from

Table 5Comparison of mesowear score (MWS), dental microwear texture analysis (DMTA), and carbon stable isotopes (δ13C) S1 (Bibi et al., 2013) S2 (Negash et al., 2015) in Shungura Formationspecimens.

Specimen Tooth position Mb Species MWS DMTA δ13C S1 δ13C S2

L 35-40e UM3 G T. nakuae Mf-Gr −5.5L 621-2a LM2 G T. nakuae Mf GrL 881-2a LM3 G T. nakuae Mf BrL 881-2b LM3 G T. nakuae Mf BrOMO 115-1973-4764 UM G T. nakuae Mf BrOMO 222-1973-2755 LM3 G T. nakuae Br −5.1OMO 257-1973-5341 UM3 G T. nakuae Mf −2.4OMO 47-1968-2536 UM2 G T. nakuae Mf −9.8OMO 47-1968-2771 UM3 G T. nakuae Br −10.8OMO 47-1970-2030 UM2 G T. nakuae Br −6OMO 75/N-1971-1211 LM2 G T. nakuae Mf Mf-GrOMO 33-1972-1 UM2 F T. nakuae Br −6.8OMO 33-1973-6084 UM3 F T. nakuae Br −4.7OMO 33-1974-3933 LM2 F T. nakuae Br −6.2OMO 33-1974-6306 UM2 F T. nakuae Br BrOMO 33-1974-6592 UM3 F T. nakuae Mf-Gr BrOMO 42-1969-1866 UM2 F T. gaudryi Mf BrOMO 42-1969-1881 UM2 F T. gaudryi Mf Mf-GrOMO 76-1969-1591 UM3 F T. gaudryi Mf Br-FrOMO 76-1969-1620 UM3 F T. gaudryi Br MfL 128-2 LM2 E T. gaudryi Mf GrOMO 166-1973-793 UM2 E T. nakuae Mf GrOMO 305-1976-380 UM3 E T. nakuae Br MfOMO 57/5-1972-375 UM2 E T. nakuae Mf Mf-BrOMO 57-1968-2860 UM2 E T. nakuae Mf MfL 220-4 LM C T. nakuae Br BrL 335-20 LM3 C T. nakuae Mf BrL 444-1 LM3 C T. nakuae Br −5.6L 444-10003 UM2 C Tragelaphini Mf BrL 460-10002 LM2 C T. nakuae Mf MfL 547-5 UM C T. nakuae Br BrL 572-1f LM3 C T. nakuae Br −3.1L 807-10002 UM3 C T. nakuae Mf BrL 831-10003 LM2 C T. nakuae Mf BrOMO 18-1968-2892 UM3 C T. nakuae Br −2.8OMO 56/sup-10008 UM2 C T. nakuae Mf GrOMO P79I/S-1970-2985 UM3 C T. nakuae Mf −4.8OMO18/sup10035 LM3 C T. nakuae Br −4.5B-116 LM3 B T. rastafari Br −8L 745-1 UM3 B T. rastafari Mf −9.4OMO 20/2-10003 LM3 B T. rastafari Mf Br −9.2OMO 28-1967-463 LM2 B T. rastafari Br Fr −1.8OMO 28-1967-508 LM B T. rastafari Br −11.6OMO 28-1968-2463 LM3 B T. rastafari Mf Br −6.6 −4.2OMO 28-1970-1853 UM3 B T. rastafari Br Br −9.8 −9OMO 3-1967-875 LM2 B T. rastafari Mf FrOMO 3-1968-2341 LM2 B T. rastafari Br Fr

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

115

Page 14: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

C3- to C4-dominated diets) between Member B and Member C, there isno dental microwear texture indicative of grazing for Tragelaphinithrough the Shungura sequence (Fig. 9E). The discrepancy betweendental wear proxies and stable isotope data for Member C tragelaphinssuggests that the increase in C4 plant consumption ~ 2.8Ma (Fig. 9)might not reflect an increase of herbaceous monocots abundance, butrather the presence of C4 dicots (see Section 5.2). Another alternative isthat tragelaphins fed seasonally on young, less tough, and Si-depletedC4 grass shoots as suggested by Bibi et al. (2013). They may have grazedon green and tender grasses during the early wet season and thenswitched to foliage when grasses became mature, tough, and scarceduring the dry season.

Among the larger tribes of African bovids, tragelaphins are the mostdependent on tree cover for shelter (although widespread in term oftheir feeding habits). Thus, they inhabit a wide variety of closed ha-bitats including forests (Tragelaphus eurycerus), swamps (T. spekii),forest-savanna mosaic (T. angasii), open woodlands, arid lowlandthornbush (T. imberbis), and scrubland for T. oryx (Kingdon, 2013;Estes, 1991) throughout sub-Saharan Africa. As suggested above, T.rastafari might have been ecologically analogous to T. eurycerus, occu-pying forested habitats, whereas the two other fossil species occupiedmore open habitats. T. nakuae might have inhabited forest-savannamosaics similar to mixed feeding T. angasii, while T. gaudryi might haveoccupied open woodlands and bushland as does present-day T. imberbis.

This suggests that multiple habitats in varying proportions character-ized the Lower Omo Valley through the Plio-Pleistocene.

Botanical proxies also suggest the presence of various habitatsthrough time in the Shungura Formation, although there was a per-manent presence of woody vegetation. In Members B and C, Bonnefilleand Dechamps (1983) show that trees comprise> 30% of pollen as-semblages. This value drops to< 5% in Member E and while there is asteady increase in grass pollen, from 43% to 47% through members Band C to 72% in Member E. In members F and G, grass abundance dropsto 57% and 51%, respectively, suggesting a slight encroachment ofwoody vegetation onto the grass-dominated environments (Bonnefilleand Dechamps, 1983). Using stable isotope analyses of pedogenic car-bonates, Levin et al. (2011) suggest a dominance of C3 vegetationthrough time but an increase of C4 plants in Member G after 2.0Ma.Cerling et al. (2011) also suggest a predominance of woodland/bush-land/shrubland to wooded grassland, indicating a dominance of closedhabitats. Ecomorphological analyses of astragali by Barr (2015) andPlummer et al. (2015) also exclude the presence of a monotypic habitatin the Shungura Formation. Although the habitat predictions differbetween the two studies, the results show a broad range of habitatsfrom forest/heavy cover open/light cover habitats. Both studies detect ashift towards more open habitats and the decrease of the woody coverthrough time, which is also supported by a recent ecometric study ofmammalian fauna of the Omo-Turkana Basin (Fortelius et al., 2016).

Fig. 9. A. Relative abundances of bovid tribes from the Shungura formation (IORE collection). B. Chronostratigraphic sequence of the Shungura Formation after de Heinzelin (1983).Dates in Ma provided at member boundaries after Feibel et al. (1989), McDougall and Brown (2006, 2008), and McDougall et al. (2012). C. Dental microwear textural parameters(mean ± 1.96 sem) on tragelaphins in different members of the Shungura Formation; black square represent mean complexity (Asfc); white square represent mean anisotropy (epLsar). D.Mesowear score (MWS; mean ± 1.96 sem) of tragelaphins in different members of the Shungura Formation. E. Enamel carbon stable isotopic composition of tragelaphins (δ13C ‰VPDB); white square after Bibi et al. (2013) and black triangle after Negash et al. (2015), the dispersions represent standard error of the mean.

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

116

Page 15: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

Collectively, the botanical and ecomorphological proxies, althoughsuggesting different proportions of habitat types, attest to the presenceof available browse resources for tragelaphins through the Shungurasequence. The browsing ecology proposed for T. rastafari, which re-presents 16% of all specimens in Member B (Fig. 9), is consistent withthe presence of closed habitats before 2.9Ma. After 2.9 Ma, tragela-phins increased in abundance (Fig. 9A) and included more C4 grasses orC4 succulent dicots in their diet. In either case, this would suggest thepresence of more open habitats compared with earlier tragelaphinsfrom Member B. It is worth mentioning that dental microwear textureanalysis suggest that tragelaphins, which represent 43% of specimens inMember E (Fig. 9A), likely included more grasses in their diet thanthose in lower members (higher anisotropy combined with lowercomplexity; Fig. 9C). The relative abundance of tragelaphins then dropsto 27% and 24% in members F and G, respectively (Fig. 9A). This mayindicate a decrease of their preferred habitat, presumably with woodycover.

5.4. Synthetic ecological view of the Tragelaphus nakuae lineage

The new dietary data for Shungura specimens of the Tragelaphusrastafari-nakuae lineage presented here can be combined with thosefrom other sites to construct a synthetic ecological view of this lineage'sdietary evolution. Sediments from Woranso-Mille in the Afar ofEthiopia slightly predate those of the Shungura Formation and haveproduced the only known remains of T. saraitu, the likely ancestor of T.rastafari (Geraads et al., 2009). Recently published isotope data of theWoranso-Mille specimens indicate mixed feeding with a preference forbrowse as isotope values range from −7.0 to −4.0‰ (n= 4), althoughseveral other Tragelaphus specimens, possibly belonging to T. saraitu,span the range of obligate browsing (−12.1‰) to C4-dominated mixedfeeding or grazing (−2.4‰) (Levin et al., 2015). Mesowear analysis ofthe Woranso-Mille Tragelaphini indicate attrition-dominated wear si-milar to modern tragelaphins, suggesting a diet of browse or mixedfeeding with a significant browse component (Curran and Haile-Selassie, 2016).

Tragelaphus rastafari is common in the Pliocene Hadar Formationand the Pliocene-age sediments of the Koobi Fora Formation (Harris,1991; Reed and Bibi, 2011; Geraads et al., 2012). Isotope data fromboth formations indicates C3–C4 mixed feeding (Cerling et al., 2015;Wynn et al., 2016), although some of the Koobi Fora sample may re-present T. kyaloi. Regardless, these data contrast sharply with the bulkof Member B specimens of T. rastafari that indicate C3-dominated diets.This may reflect differences in habitat composition between these areas,as the Shungura Formation is generally reconstructed as having agreater proportion of woody cover than either Hadar or Koobi Fora(Bobe et al., 2007; Levin et al., 2011). Tragelaphin specimens, pre-sumably including T. nakuae, from younger sediments in the Koobi ForaFormation (> 2.0Ma) also indicate mixed feeding, which is consistentwith the evidence from the Shungura T. nakuae specimens. Thus,overall it appears that the T. saraitu-rastafari-nakuae lineage maintaineda mixed feeding dietary niche, with varying proportions of C3 versus C4

inputs, from ~3.6 to>2.0Ma.In contrast to the dietary stability in the T. saraitu-rastafari-nakuae

lineage through time, these species show increases in body massthrough the Plio-Pleistocene (Fig. 10). Although sample sizes are small,most Pliocene specimens older than 3Ma are reconstructed as havingbody masses < 250 kg, whereas there is a shift to greater body massafter this time (average body masses between 250 and 300 kg). A fewexceptionally large individuals (350–400 kg) are found during the earlyPleistocene (~2.6–2.4Ma). The trend towards larger body mass in thislineage cannot be explained by dietary changes as all three species seemto have been engaged in mixed feeding through time. However, it ispossible that the increase in body mass occurred with the fragmentationof closed habitats and the spread of open ones through time, as sug-gested for the Shungura Formation. It has been shown in extant African

bovids that habitat preference is significantly linked to body size (Bro-Jørgensen, 2008). Open habitats would have exposed individuals togreater predation risk from a diverse Plio-Pleistocene guild of felids andhyaenids, and thus selection might have favored greater body mass as apredator deterrent.

Visually, there is no clear bimodality to the body mass data at anypoint in time that would differentiate males from females. This is not asurprising result, given the fact that the body mass estimates are basedon molar row lengths and bovids are rarely dentally dimorphic.However, it has been shown that there is clear bimodality in the horncore sizes of both T. rastafari and T. nakuae, which indicates that thesetaxa likely had horned females and were sexually dimorphic in horncore size, as in their extant relative the bongo T. eurycerus (Reed andBibi, 2010; Bibi, 2011). Bongos are also strongly dimorphic in bodymass – males range from 335 to 400 kg, whereas females range182–276 kg (Elkan and Smith, 2013). Perhaps the increased variance inthe body mass estimates seen in Pleistocene specimens of T. nakuaeindicates the evolution towards increased body mass dimorphism (butsee above), but this hypothesis is confounded by the small sample sizesfor the Pliocene-age specimens. If the T. nakuae lineage does becomemore sexually dimorphic over time, this too may be related to thefragmentation of closed habitats and the spread of open ones, as pro-posed above for the general increase in body mass through time. Bovidsin open habitats are not only larger, but they also rely on group living asan antipredator strategy, which in turn produces polygynous socialsystems and increases the intensity of sexual selection (Jarman, 1974;Perez-Barberia et al., 2002). Therefore, a shift in the social organizationof T. nakuae coincident with the spread of open habitats could explainthe evolution of body mass dimorphism in this lineage.

6. Conclusions

We analyzed dental mesowear and microwear texture ofTragelaphini from the Shungura Formation and combined these datawith previously published isotope data to reconstruct the diet and ha-bitat preference of these species through time. Our results stronglysupport the usefulness of a multi-proxy approach when studying thedietary preferences of extinct herbivores, as each approach differs in thedietary information provided. Overall, we find that the ShunguraFormation tragelaphins had flexible diets ranging from browsing tomixed feeding. It appears that the T. rastafari-nakuae lineage main-tained a mixed feeding dietary niche, with varying proportions of C3

versus C4 inputs, from ~3.6 to>2.0Ma. Our results suggest that T.rastafari might be more engaged in browsing than T. nakuae, a mixedfeeder that may have begun to consume C4 dicots ~ 2.8Ma. The slightchange of diet in the T. rastafari-nakuae lineage was reflective ofbroader ecological changes in the Shungura Formation during this time,as suggested by Bibi et al. (2013), but is seemingly decoupled frombody mass evolution. A third fossil species T. gaudryi, related to theextant lesser kudu T. imberbis, was also a mixed feeder from Member Ctimes onwards. Our study represents a first step towards a better un-derstanding of paleoenvironmental change through time in the Shun-gura Formation using a multi-proxy approach of herbivore diets. Itwould be fruitful for future studies to analyze other portions of theherbivore assemblage from the Shungura Formation as a way oftracking the shifting proportions of habitats in the Omo-Turkana Basinduring the Plio-Pleistocene.

Acknowledgements

We thank the Authority for Research and Conservation of CulturalHeritage (ARCCH) for authorizing our field research and for granting uspermission to sample bovid dental remains from the Omo collections.We thank the Ministère Français de l'Education Nationale (UFR SFA,Université de Poitiers). This research was funded in the framework ofthe Omo Group Research Expedition (Agence Nationale de la

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

117

Page 16: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

Recherche: projects EVAH ANR-09-BLAN-0238 and OLD ANR-CE27-0009-02; Ministère de l'Europe et des Affaires Étrangères; CNRS; CentreFrançais des Etudes Ethiopiennes; and Laboratoire PaléontologieÉvolution Paléoécosystèmes Paléoprimatologie; Fondation Fyssen).This study was also financed by the Project ANR TRIDENT (ANR-13-JSV7-0008-01, PI: G. Merceron, http://anr-trident.prd.fr/). J. Rowanwas supported by a National Science Foundation Graduate ResearchFellowship. We also thank all the other people of OGRE who con-tributed to field work and fossil preparation. We are grateful to SabineRiffaut for illustrations. We acknowledge five anonymous reviewers fortheir constructive comments.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.palaeo.2018.01.027.

References

Alemseged, Z., Bobe, R., Geraads, D., 2007. Comparability of fossil data and its sig-nificance the interpretation of hominin environments. In: Bobe, R., Alemseged, Z.,Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: AnAssessment of the Faunal Evidence, pp. 159–181.

Arambourg, C., 1934. Le Dinotherium des gisements de l'Omo (Abyssinie). Bull. Soc. Geol.Fr. 5, 305–310.

Arambourg, C., 1943. Contribution à l'étude géologique et paléontologique du bassin dulac Rodolphe et de la basse vallée de l'Omo. Première partie: géologie. In: Arambourg,C. (Ed.), Mission Scientifique de l'Omo, 1932–1933. Géologie, Anthropologie, Tome 1Editions du Muséum, pp. 157–230.

Arambourg, C., 1947. Contribution à l'étude géologique et paléontologique du bassin dulac Rodolphe et de la basse vallée de l'Omo. Deuxième partie: paléontologie. In:Arambourg, C. (Ed.), Mission Scientifique de l'Omo, 1932–1933. Géologie,Anthropologie, Tome 1 Editions du Muséum, Paris, pp. 231–562.

Arambourg, C., Coppens, Y., 1967. Sur la découverte, dans le Pléistocène inférieur de lavallée de l'Omo (Éthiopie), d'une mandibule d'australopithécien. C. R. Acad. Sci.,Paris Ser. D. 265, 589–590.

Balasse, M., 2003. Potential biases in sampling design and interpretation of intra-toothisotope analysis. Int. J. Osteoarchaeol. 13, 3–10.

Barr, W.A., 2015. Paleoenvironments of the Shungura Formation (Plio-Pleistocene:Ethiopia) based on ecomorphology of the bovid astragalus. J. Hum. Evol. 88, 97–107.

Bedaso, Z., Wynn, J.G., Alemseged, Z., Geraads, D., 2013. Dietary and paleoenviron-mental reconstruction using stable isotopes of herbivore tooth enamel from middlePliocene Dikika, Ethiopia: implication for Australopithecus afarensis habitat and foodresources. J. Hum. Evol. 64, 21–38.

Bibi, F., 2011. Tragelaphus nakuae: evolutionary change, biochronology, and turnover inthe African Plio-Pleistocene. Zool. J. Linn Soc-Lond. 162, 699–711.

Bibi, F., Souron, A., Bocherens, H., Uno, K., Boisserie, J.R., 2013. Ecological change in thelower Omo Valley around 2.8Ma. Biol. Lett. 9, 20120890.

Blondel, C., Merceron, G., Likius, A., Mackaye, H.T., Vignaud, P., Brunet, M., 2010. Dentalmesowear analysis of the late Miocene Bovidae from Toros-Menalla (Chad) and earlyhominid habitats in Central Africa. Palaeogeogr. Palaeoclimatol. Palaeoecol. 292,184–191.

Bobe, R., Eck, G.G., 2001. Responses of African bovids to Pliocene climatic change.Paleobiology 27, 1–47.

Bobe, R., Behrensmeyer, A.K., Eck, G.G., Harris, J.M., 2007. Patterns of Abundance andDiversity in Late Cenozoic Bovids from the Turkana and Hadar Basins, Kenya andEthiopia. In: Bobe, R., Alemseged, Z., Behrensmeyer, A.K. (Eds.), HomininEnvironments in the East African Pliocene. Springer, Dordrecht, pp. 129–157.

Boisserie, J.-R., Guy, F., Delagnes, A., Hlusko, L.J., Bibi, F., Beyene, Y., Guillemot, C.,2008. New palaeoanthropological research in the Plio-Pleistocene Omo Group, LowerOmo Valley, SNNPR (Southern Nations, Nationalities and People Regions), Ethiopia.C.R. Palevol. 7, 429–439.

Boisserie, J.R., Delagnes, A., Beyene, Y., Schuster, M., 2010. Reconstructing the Africanbackground to human expansions in Eurasia: new research in the ShunguraFormation, Ethiopia. Quat. Int. 223-224, 426–428.

Bonnefille, R., 2010. Cenozoic vegetation, climate changes and hominid evolution intropical Africa. Glob. Planet. Chang. 72, 390–411.

Bonnefille, R., Dechamps, R., 1983. Data on fossil flora. In: de Heinzelin, J. (Ed.), TheOmo Group. Annales-Série Sciences Géologiques, vol. 85. Musée Royale de l'AfriqueCentral, Tervuren, Belgique, pp. 191–207.

Bro-Jørgensen, J., 2008. Dense habitats selecting for small body size: a comparative studyon bovids. Oikos 117, 729–737.

Calandra, I., Merceron, G., 2016. Dental microwear texture analysis in mammal ecology.Mammal Rev. 46, 215–228.

Calandra, I., Schulz, E., Kaiser, T.M., 2012. Teasing apart the contributions of hard dietaryitems on 3d dental microtextures in primates. J. Hum. Evol. 63, 8598.

Cerling, T.E., Harris, J.M., 1999. Carbon isotope fractionation between diet and bioapa-tite in ungulate mammals and implications for ecological and paleoecological studies.Oecologia 120, 347–363.

Cerling, T.E., Harris, J.M., MacFadden, B.J., Leakey, M.G., Quade, J., Eisenmann, V.,Ehleringer, J.R., 1997. Global vegetation change through the Miocene/Plioceneboundary. Nature 389, 153–158.

Cerling, T.E., Harris, J.M., Passey, B.H., 2003. Diets of East African Bovidae based onstable isotopic analysis. J. Mammal. 84, 456–470.

Cerling, T.E., Wynn, J.G., Andanje, S.A., Bird, M.I., Korir, D.K., Levin, N.E., Mace, W.,Macharia, A.N., Quade, J., Remien, C.H., 2011. Woody cover and hominin environ-ments in the past 6 million years. Nature 476, 51–56.

Cerling, T.E., Andanje, S.A., Blumenthal, S.A., Brown, F.H., Chritz, K.L., Harris, J.M.,Hart, J.A., Kirera, F.M., Kaleme, P., Leakey, L.N., Leakey, M.G., Levin, N.E., Manthi,F.K., Passey, B.H., Uno, K.T., 2015. Dietary changes of large herbivores in theTurkana Basin, Kenya from 4 to 1Ma. Proc. Natl. Acad. Sci. U. S. A. 112,11467–11472.

Conover, W.J., Iman, R.L., 1981. Rank transformations as a bridge between parametricand nonparametric statistics. Am. Stat. 35, 124–129.

Coppens, Y., Howell, F.C., Isaac, G.L., Leakey, R.E., 1976. Earliest Man and Environments

Fig. 10. Body mass evolution in the Tragelaphus saraitu-rastafari-nakuae lineage from the mid-Pliocene (~3.6Ma) until the earlyPleistocene (~2.1Ma). Data are shown in roughly 250,000-yearbins plotted by the bin's midpoint age (based on published siteages). These data document a significant size increase in thelineage through time, with early Pleistocene specimens fallingwithin the range of extant male bongos (Tragelaphus eurycerus).

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

118

Page 17: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

in the Lake Rudolf Basin: Stratigraphy, Paleoecology, and Evolution. University ofChicago Press, Chicago.

Curran, S.C., Haile-Selassie, Y., 2016. Paleoecological reconstruction of hominin-bearingmiddle Pliocene localities at Woranso-Mille, Ethiopia. J. Hum. Evol. 96, 97–112.

Davis, M., Pineda-Munoz, S., 2016. The temporal scale of diet and dietary proxies. Ecol.Evol. 6 (6), 1883–1897.

Delagnes, A., Boisserie, J.-R., Beyene, Y., Chuniaud, K., Guillemot, C., Schuster, M., 2011.Archaeological investigations in the Lower Omo Valley (Shungura Formation,Ethiopia): new data and perspectives. J. Hum. Evol. 61, 215–222.

Edwards, E.J., Osborne, C.P., Strömberg, C.A.E., Smith, S.A., the C4 Grasses Consortium,2010. The origins of C4 grasslands: integrating evolutionary and ecosystem science.Science 328, 587–591.

Ehleringer, J.R., Cerling, T.E., Helliker, B.R., 1997. C4 photosynthesis, atmospheric CO2,and climate. Oecologia 112, 285–299.

Elkan, P.W., Smith, J.L.D., 2013. Tragelaphus eurycerus Bongo. In: Kingdon, J., Hoffmann,M. (Eds.), Mammals of Africa. Volume VI, Pigs, Hippopotamuses, Chevrotain,Giraffes, Deer and Bovids. Bloomsbury Publishing, London, pp. 179–185.

Estes, R.D., 1991. Behavior Guide to African Mammals. The University of California Press,Los Angeles.

Feibel, C.S., Brown, F.H., McDougall, I., 1989. Stratigraphic context of fossil Hominidsfrom the Omo group deposits: Northern Turkana Basin, Kenya and Ethiopia. Am. J.Phys. Anthropol. 78, 595–622.

Fortelius, M., Solounias, N., 2000. Functional characterization of ungulate molars usingthe abrasion-attrition wear gradient: a new method for reconstructing paleodiets.Am. Mus. Novit. 3301, 1–36.

Fortelius, M., Zliobaite, I., Kaya, F., Bibi, F., Bobe, R., Leakey, L., Leakey, M., Patterson,D., Rannikko, J., Werdelin, L., 2016. An ecometric analysis of the fossil mammalrecord of the Turkana Basin. Philos. Trans. R. Soc. B 371, 20150232.

Franz-Odendaal, T.A., Kaiser, T.M., 2003. Differential mesowear in the maxillary andmandibular cheek dentition of some ruminants (Artiodactyla). Ann. Zool. Fenn. 40,395–410.

Gagnon, M., Chew, A.E., 2000. Dietary preferences in extant African Bovidae. J. Mammal.8, 490–511.

Gebert, C., Verheyden-Tixier, H., 2001. Variations of diet composition of red deer (Cervuselaphus L.) in Europe. Mammal Rev. 31, 189–201.

Gentry, A.W., 1985. The Bovidae of the Omo deposits, Ethiopia. In: Cahiers dePaléontologie-Travaux de paléontologie est-africaine, vol. 1. pp. 119–191.

Geraads, D., Melillo, S., Haile-Selassie, Y., 2009. Middle Pliocene Bovidae from Hominid-bearing sites in the Woranso-Mille area, Afar region, Ethiopia. Palaeontol. Afr. 44,59–70.

Geraads, D., Bobe, R., Reed, K., 2012. Pliocene Bovidae (Mammalia) from the HadarFormation of Hadar and Ledi-Geraru, Lower Awash, Ethiopia. J. Vertebr. Paleontol.32, 180–197.

Goodall, R.H., Darras, L.P., Purnell, M.A., 2015. Accuracy and precision of silicon basedimpression media for quantitative a real texture analysis. Sci. Rep. 5, 10800.

Grine, F.E., Sponheimer, M., Ungar, P.S., Lee-Thorp, J., Teaford, M.F., 2012. Dental mi-crowear and stable isotopes inform the paleoecology of extinct hominins. Am. J.Phys. Anthropol. 148, 285–317.

Harris, J.M., 1991. Family Bovidae. In: Harris, J.M. (Ed.), Koobi Fora Research Project.The Fossil Ungulates: Geology, Fossil Artiodactyls, and Palaeoenvironments.Clarendon Press, Oxford, pp. 139–320.

de Heinzelin, J., 1983. The Omo Group. Archives of the International Omo Research,Expedition. Musée Royal de l'Afrique Centrale, Tervuren, pp. 365.

Hillman, J.C., Gwynne, M.D., 1987. Feeding of the Bongo, Tragelaphus eurycerus (Ogilby,1837), in south west Sudan. Mammalia 51, 53–64.

Hodson, M.J., White, P.J., Mead, A., Broadley, M.R., 2005. Phylogenetic variation in thesilicon composition of plants. Ann. Bot. London 96, 1027–1046.

Hofmann, R.R., 1973. The Ruminant Stomach. Stomach Structure and Feeding Habits ofEast African Game Ruminants. East African Literature Bureau, Nairobi (354 pp).

Hofmann, R.R., Stewart, D.R.M., 1972. Grazer or browser: a classification based on thestomach structure and feeding habits of East African ruminants. Mammalia 36,226–240.

Howell, F.C., 1968. Omo Research expedition. Nature 219, 567–572.Howell, F.C., Coppens, Y., 1976. An overview of Hominidae from the Omo succession,

Ethiopia. In: Coppens, Y., Howell, F.C., Isaac, G.L., Leakey, R.E.F. (Eds.), Earliest Manand Environments in the Lake Rudolf Basin. Chicago Press, Chicago, pp. 522–532.

Janis, C.M., 1990. Correlation of cranial and dental variables with body size in ungulatesand macropodoids. In: Damuth, J., McFadden, B.J. (Eds.), Body Size in MammalianPaleobiology: Estimation and Biological Implications. Cambridge University Press,Cambridge, pp. 255–299.

Jarman, P., 1974. The social organisation of antelope in relation to their ecology.Behaviour 48, 215–267.

Kaiser, T.M., 2011. Feeding ecology and niche partitioning of the Laetoli ungulate faunas.In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution inContext, Volume 1: Geology, Geochronology, Paleoecology, and Paleoenvironment.Springer, New York, pp. 329–354.

Kaiser, T.M., Fortelius, M., 2003. Differential mesowear in occluding upper and lowermolars: opening mesowear analysis for lower molars and premolars in hypsodonthorses. J. Morphol. 258, 67–83.

Kaiser, T.M., Brasch, J., Castell, J.C., Schulz, E., Clauss, M., 2009. Toothwear in captivewild ruminant species differs from that of free-ranging conspecifics. Mamm. Biol. 74,425–437.

Kaiser, T.M., Müller, D.W.H., Fortelius, M., Schulz, E., Codron, D., Clauss, M., 2013.Hypsodonty and tooth facet development in relation to diet and habitat in herbi-vorous ungulates: implications for understanding tooth wear. Mammal Rev. 43,34–46.

Kingdon, J., 2013. Genus Tragelaphus spiral-horned antelopes. In: Kingdon, J., Hoffmann,M. (Eds.), Mammals of Africa. Volume VI, Pigs, Hippopotamuses, Chevrotain,Giraffes, Deer and Bovids. Bloomsbury Publishing, London, pp. 138–141.

Kingston, J.D., 2011. Stable isotopic analyses of Laetoli fossil herbivores. In: Harrison, T.

(Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context Volume 1:Geology, Geochronology, Paleoecology, and Paleoenvironment. Springer, New York,pp. 293–328.

Kohn, M.J., 2004. Comment: tooth enamel mineralization in ungulates: implications forrecovering a primary isotopic time-series, by B.H. Passey and T.E. Cerling (2002).Geochim. Cosmochim. Acta 68, 403–405.

Lee-Thorp, J.A., 1989. Stable Carbon Isotopes in Deep Time: Diets of Fossil Fauna andHominid. University of Cape Town, pp. 174 (PhD thesis).

Lee-Thorp, J.A., van der Merwe, N.J., 1987. Carbon isotope analysis of fossil bone apatite.S. Afr. J. Sci. 83, 712–715.

Leuthold, B.M., Leuthold, W., 1978. Ecology of the giraffe in Tsavo East National Park,Kenya. East Afr. Wildlife J. 16, 1–20.

Levin, N.E., Brown, F.H., Behrensmeyer, A.K., Bobe, R., Cerling, T.E., 2011. Paleosolcarbonates from the Omo Group: isotopic records of local and regional environmentalchange in East Africa. Palaeogeogr. Palaeoclimatol. Palaeoecol. 307, 75–89.

Levin, N.E., Haile-Selassie, Y., Frost, S.R., Saylor, B.Z., 2015. Dietary change among ho-minins and cercopithecids in Ethiopia during the early Pliocene. Proc. Natl. Acad. Sci.U. S. A. 112, 12304–12309.

Louys, J., Meloro, C., Elton, S., Ditchfield, P., Bishop, L.C., 2011. Mesowear as a means ofdetermining diets in African antelopes. J. Archaeol. Sci. 28, 1485–1495.

Lucas, P.W., 2004. Dental Functional Morphology: How Teeth Work. CambridgeUniversity Press, Cambridge, UK.

Lumpkin, S., Kranz, K.R., 1984. Cephalophus silvicultor. Mamm. Species (225), 1–7.Maurin, T., Delagnes, A., Boisserie, J.-R., 2014. Spatial behaviours of early Oldowan

toolmakers in the Shungura Formation (Lower Omo Valley, Ethiopia): proposal for anintegrated approach. C.R. Palevol. 13, 737–746.

McDougall, I., Brown, F.H., 2006. Precise 40Ar/39Ar geochronology for the upper KoobiFora Formation, Turkana Basin, northern Kenya. J. Geol. Soc. Lond. 163, 205–220.

McDougall, I., Brown, F.H., 2008. Geochronology of the pre-KBS Tuff sequence, OmoGroup, Turkana Basin. J. Geol. Soc. 165, 549–562.

McDougall, I., Brown, F.H., Vasconcelos, P.M., Cohen, B., Thiede, D., Buchanan, M., 2012.New single crystal 40Ar/39Ar ages improve time scale for deposition of the OmoGroup, Omo-Turkana Basin, East Africa. J. Geol. Soc. 169, 213–226.

Merceron, G., Schulz, E., Kordos, L., Kaiser, T.M., 2007. Paleoenvironment of Dryopithecusbrancoi at Rudabánya, Hungary: evidence from dental meso- and micro-wear analysesof large vegetarian mammals. J. Hum. Evol. 53, 331–349.

Merceron, G., Escarguel, G., Angibault, J.-M., Verheyden-Tixier, H., 2010. Can dentalmicrowear textures record inter-individual dietary variations? PLoS One 5, e9542.

Merceron, G., Hofman-Kaminska, E., Kowalczyk, R., 2014. 3D dental microwear textureanalysis of feeding habits of sympatric ruminants in the Białowiezia Primeval Forest,Poland. For. Ecol. Manag. 328, 262–269.

Mihlbachler, M.C., Rivals, F., Solounias, N., Semprebon, G.M., 2011. Dietary change andevolution of horses in North America. Science 331, 1178–1181.

Negash, E.W., Alemseged, Z., Wynn, J.G., Bedaso, Z.K., 2015. Paleodietary reconstructionusing stable isotopes and abundance analysis of bovids from the Shungura Formationof South Omo, Ethiopia. J. Hum. Evol. 88, 127–136.

Passey, B.H., Cerling, T.E., 2002. Tooth enamel mineralization in ungulates: implicationsfor recovering a primary isotopic time-series. Geochim. Cosmochim. Acta 66,3225–3234.

Perez-Barberia, F.J., Gordon, I.J., Pagel, M., 2002. The origins of sexual dimorphism inbody size in ungulates. Evolution 56, 1276–1285.

Plummer, T.W., Ferraro, J.V., Louys, J.C., Hertel, F., Alemseged, Z., Bobe, R., Bishop, L.C.,2015. Bovid ecomorphology and hominin paleoenvironments of the ShunguraFormation, Lower Omo River Valley, Ethiopia. J. Hum. Evol. 88, 108–126.

Reed, K., Bibi, F., 2011. Fossil Tragelaphini (Artiodactyla: Bovidae) from the LatePliocene Hadar Formation, Afar Regional State, Ethiopia. J. Mamm. Evol. 18, 57–69.

Rivals, F., Mihlbachler, M.C., Solounias, N., 2007. Effect on the ontogenetic-age dis-tribution in fossil and modern samples on the interpretation of ungulate paleodietsusing the mesowear method. J. Vertebr. Paleontol. 27, 763–767.

Rozzi, R., Winkler, D.E., de Vos, J., Schulz, E., Palombo, M.R., 2013. The enigmatic bovidDuboisia santeng (Dubois, 1891) from the Early-Middle Pleistocene of Java: a multi-proxy approach to its paleoecology. Palaeogeogr. Palaeoclimatol. Palaeoecol. 377,73–85.

Sage, R.F., Monson, R.K. (Eds.), 1999. C4 Plant Biology. Academic Press, San Diego, CA.Sage, R.F., Christin, P.-A., Edwards, E.J., 2011. The C4 plant lineages of planet Earth. J.

Exp. Bot. 62 (9), 3155–3169.Schulz, E., Calandra, I., Kaiser, T.M., 2010. Applying tribology to teeth of hoofed mam-

mals. Scanning 32, 162–182.Schulz, E., Calandra, I., Kaiser, T.M., 2013. Feeding ecology and chewing mechanics in

hoofed mammals: 3D tribology of enamel wear. Wear 300, 169–179.Scott, J.R., 2012. Dental microwear texture analysis of extant African Bovidae. Mammalia

76, 157–174.Scott, R.S., Ungar, P.S., Bergstrom, T.S., Brown, C.A., Childs, B.E., Teaford, M.F., Walker,

A., 2006. Dental microwear texture analysis: technical considerations. J. Hum. Evol.51, 339–349.

Sokal, S.R., Rohlf, F.J., 1969. Biometry. W.E. Freeman and Company, New York.Sponheimer, M., Lee-Thorp, J.A., DeRuiter, D.J., Smith, J.M., Van der Merwe, N.J., Reed,

K., Grant, C.C., Ayliffe, L.K., Robinson, T.F., Heidelberger, C., Marcus, W., 2003. Dietsof southern African Bovidae: stable isotope evidence. J. Mammal. 84, 471–479.

Suwa, G., 1988. Evolution of the “robust” australopithecines in the Omo succession:evidence from mandibular premolar morphology. In: Grine, F. (Ed.), TheEvolutionary History of the Robust Australopithecines. Aldine, New York, pp.199–222.

Teaford, M.F., Oyen, O.J., 1989. In vivo and in vitro turnover in dental microwear. Am. J.Phys. Anthropol. 80, 447–460.

Ungar, P.S., Merceron, G., Scott, R.S., 2007. Dental microwear texture analysis ofVarswater bovids and Early Pliocene paleoenvironments of Langebaanweg, WesternCape Province, South Africa. J. Mamm. Evol. 14, 163–181.

Vogel, J.C., 1978. Isotopic assessment of the dietary habits of ungulates. S. Afr. J. Sci. 74,298–301.

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

119

Page 18: Palaeogeography, Palaeoclimatology, Palaeoecology · be linked to the evolution of larger body mass. Conversely, if diets of the T. rastafari-T. nakuae indicate mixed-feeding through

Vulink, T.J., Drost, J.T., 1991. Nutritional characteristics of cattle forage plants in theeutrophic nature-reserve Oostvaardersplassen, Netherlands. Neth. J. Agric. Sci. 39,263–272.

Winkler, D.E., van den Hoek Ostende, L.W., Schulz, E., Calandra, I., Gailer, J.P.,Landwehr, C., Kaiser, T.M., 2013. Dietary divergence in space and time – lessons fromthe dwarf-goat Myotragus balearicus (Pleisto-Holocene, Mallorca, Spain). Mamm. Biol.78, 430–437.

Winkler, D.E., Andrianasolo, T.H., Andriamandimbiarisoa, L., Ganzhorn, J.U.,Rakotondranary, S.J., Kaiser, T.M., Schulz-Kornas, E., 2016. Tooth wear patterns inblack rats (Rattus rattus) of Madagascar differ more in relation to human impact than

to differences in natural habitats. Ecol. Evol. 6 (7), 2205–2215.Wynn, J.G., Reed, K.E., Sponheimer, M., Kimbel, W.H., Alemseged, Z., Bedaso, Z.K.,

Campisano, C.J., 2016. Dietary flexibility of Australopithecus afarensis in the face ofpaleoecological change during the middle Pliocene: faunal evidence from Hadar,Ethiopia. J. Hum. Evol. 99, 93–106.

Zazzo, A., Balasse, M., Passey, B.H., Moloney, A.P., Monahan, F.J., Schmidt, O., 2010. Theisotope record of short- and long-term dietary changes in sheep tooth enamel: im-plications for quantitative reconstruction of paleodiets. Geochim. Cosmochim. Acta74, 3571–3586.

C. Blondel et al. Palaeogeography, Palaeoclimatology, Palaeoecology 496 (2018) 103–120

120