<<

EURASIAN MIDDLE AND LATE HOMINOID PALEOBIOGEOGRAPHY

AND THE GEOGRAPHIC ORIGINS OF THE

by

Mariam C. Nargolwalla

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Anthropology University of Toronto

© Copyright by M. Nargolwalla (2009)

Eurasian Middle and Late Miocene Hominoid Paleobiogeography and the Geographic Origins of the Homininae

Mariam C. Nargolwalla Doctor of Philosophy Department of Anthropology University of Toronto 2009

Abstract

The origin and diversification of great and is among the most researched and debated series of events in the evolutionary history of the . A fundamental part of understanding these events involves reconstructing paleoenvironmental and paleogeographic patterns in the Eurasian Miocene; a time period and geographic expanse rich in evidence of lineage origins and dispersals of numerous mammalian lineages, including apes. Traditionally, the geographic origin of the African and lineage is considered to have occurred in , however, an alternative hypothesis favouring a Eurasian origin has been proposed. This hypothesis suggests that that after an initial dispersal from Africa to at ~17Ma and subsequent radiation from to , apes disperse back to Africa at least once and found the African ape and human lineage in the late Miocene. The purpose of this study is to test the Eurasian origin hypothesis through the analysis of spatial and temporal patterns of distribution, in situ evolution, interprovincial and intercontinental dispersals of Eurasian terrestrial in response to environmental factors. Using the NOW and

Paleobiology databases, together with data collected through survey and excavation of middle and late Miocene vertebrate localities in and Romania, taphonomic bias and sampling completeness of Eurasian faunas are assessed. Previous bioprovincial zonations of and Western are evaluated and modified based on statistical analysis of Eurasian faunas and consideration of geophysical, climatic and eustatic events. Within these bioprovinces, occurrences of in situ evolution and directionality of dispersals of land mammals are used as a framework to address and evaluate these same processes in Eurasian apes. The results of this analysis support previous hypotheses regarding first occurrences and phyletic relations among Eurasian apes and propose new ideas regarding the relations of these taxa to previously known and newly discovered late Miocene

African apes. Together with analysis of environmental data, Eurasian mammals support the hypothesis that the descendant of a Eurasian ape dispersed to Africa in the early late Miocene (top of MN7/8 or base of MN9), however the question of whether this taxon founded the African ape and human lineage remains equivocal.

ii

Acknowledgements

I have been so incredibly fortunate during the course of my graduate studies to be surrounded by some of the most brilliant, helpful and supportive colleagues and friends at the University of Toronto and in the field. First and foremost, I am indebted to my doctoral advisor, David Begun, who has patiently answered ten of endless questions and who has always provided me with every opportunity to progress in the field of . I honestly could not have asked for a better advisor. I am grateful to my doctoral committee members, Shawn Lehman, Michael Schillaci and Martin Evison for their very helpful comments and suggestions for improving my dissertation. I am extremely lucky to have had John Fleagle as the external examiner extraordinaire, whose extensive comments will undoubtedly improve the quality of the publications derived from this project. In addition to my committee, I thank Matthew Betts, Dave Bovee, Maia Bukhsianidze, Terry “Captain Handsome” Clark, Andrew “has the Eagle landed?” Deane, János Hír, Laszlo Kordos, Gerald Romme, Matthew Tocheri and Márton Venczel. Each and every one of these individuals has contributed data, verified data, or discussed the methods used in my dissertation. I am also very lucky to have friends who are experts at picking me up when I’m down, brushing me off and sending me back on my way (everyone above and Jennifer Campbell, Emily Court, Lesley Howse, Karen Ryan and Alexandra Sumner). I thank Cleo and Charlie for keeping me company at all hours of the day and night and for their insightful comments on my dissertation. I thank Stephen Spensieri for pointing out that my potential as a play companion could not be reached until my dissertation was completed … and then for taking enough naps to let me finish writing. I thank my parents for their enthusiasm. Finally, I thank Carlo Spensieri, who has supported me since day one and has bent over backwards to help me accomplish everything that I have done.

My research was supported by General Motors Women in Science and Mathematics, the Ontario Graduate Scholarship and University of Toronto Fellowship.

Many years ago, a very wise woman told me to take an introductory course in Anthropology. She told me that it was a fascinating subject and although I didn’t really believe her at the time, it turns out that she was right. I dedicate this dissertation to her memory.

iii

Table of Contents Chapter 1 - Introduction ...... 1 Chronological overview...... 5 Paleoenvironmental overview...... 7 Overview of study...... 11 Chapter 2 – Data Quality...... 14 Introduction...... 14 Loss of Biological Information in the Fossil Record...... 14 Materials and Methods...... 16 Spatio-temporal Distribution of Faunas...... 16 Method of Analysis...... 17 Results...... 19 Relationship between duration of temporal intervals & number of localities/number of localities & CI...... 19 Large Completeness ...... 24 Small Mammal Completeness ...... 29 Identification of “Marker Taxa” ...... 33 Discussion...... 38 Overall Completeness and Data Quality...... 38 Ecology of “Marker Taxa” ...... 40 Distribution of sample localities...... 40 Chapter Summary & Conclusions ...... 44 Chapter 3 – Provinciality, paleoenvironments, in situ evolution and dispersal ...... 45 Introduction...... 45 Zoogeographic Provinces and Purpose of Study ...... 45 Barriers to dispersal & paleoenvironmental events in the Miocene of Eurasia...... 46 Previous method and study of provinciality ...... 53 Materials & Methods ...... 57 Materials ...... 57 Analysis ...... 58 Results...... 69 Algorithms ...... 69 Patterns of locality clustering ...... 70 Discussion...... 148 Resolution of analysis...... 148 Faunal provinces in comparison to previous studies ...... 151 Faunal provinces: In situ evolution and dispersals ...... 157 Paleoenvironmental Influence on Faunal Provinces...... 179 Inter- and Intracontinental Faunal Exchange and Dispersal Pathways...... 180 Chapter Summary & Conclusions ...... 184 Chapter 4 – Eurasian Miocene hominoid biogeography ...... 186 Introduction...... 186 First appearances and relations among Eurasian Miocene apes ...... 186 and cf. Griphopithecus...... 187 / ...... 203

iv

Table of Contents (Continued)

Ouranopithecus...... 216 Summary & Conclusions ...... 223 Chapter 5 - Conclusions ...... 225 Overview of study...... 225 Paleobiogeographic relations of Eurasian apes...... 226 Intercontinental relations ...... 227 References Cited ...... 233

v

List of Tables

Table 1.1: Morphological characters shared with African apes and Dryopithecus ...... 3 Table 1.2: Temporal association of the Eurasian MN Zones...... 6 Table 2.1: Temporal ranges of Eurasian large mammal genera ...... 25 Table 2.2: Temporal ranges of Eurasian small mammal genera...... 30 Table 2.3: Temporal range and paleoecology of well-sampled Eurasian small mammals..... 34 Table 2.4: Temporal range and paleoecology of well-sampled Eurasian large mammals ..... 36 Table 3.1: Localities included in analysis...... 59 Table 3.2: Summary of provincial composition (-level) ...... 140 Table 4.1: Eurasian – Afro-Arabian shared taxa (MN5 and earlier) ...... 192 Table 4.2: FAs and centres of origin of German and Turkish MN5 taxa ...... 193 Table 4.3: European-Turkish shared taxa ...... 200 Table 4.4: Spanish large and small mammal FAs, MN6-MN9 ...... 206 Table 4.5: MN7/8 Spanish – Central European shared taxa...... 210 Table 4.6: Shared Spanish taxa...... 213

vi

List of Figures

Figure 2.1: Distribution of Sample Localities...... 16 Figure 2.2: (a-g) ...... 20 Figure 2.3: Relative completeness in Eurasian large mammal taxa ...... 24 Figure 2.4: Relative completeness in Eurasian small mammal taxa...... 29 Figure 2.5: Regional or bioprovincial zones previously assessed for CI...... 41 Figure 2.6: Regional and total large mammal completeness ...... 43 Figure 2.7: Regional and total small mammal completeness ...... 43 Figure 3.1: Physiographic features in study area...... 53 Figure 3.2: Distribution of localities...... 69 Figure 3.3a: MN5 large mammal species dendrogram...... 76 Figure 3.3b: MN5 large mammal genera dendogram...... 77 Figure 3.3c: MN5 small mammal species dendrogram ...... 78 Figure 3.3d: MN5 small mammal genera dendogram ...... 79 Figure 3.4a: MN6 large mammal species dendrogram...... 84 Figure 3.4b: MN6 large mammal genera dendogram...... 85 Figure 3.4c: MN6 small mammal species dendrogram ...... 86 Figure 3.4d: MN6 small mammal genera dendogram ...... 87 Figure 3.5a: MN7/8 large mammal species dendrogram...... 93 Figure 3.5b: MN7/8 large mammal genera dendogram...... 94 Figure 3.5c: MN7/8 small mammal species dendrogram ...... 95 Figure 3.5d: MN7/8 small mammal genera dendogram ...... 96 Figure 3.6a: MN9 large mammal species dendrogram...... 102 Figure 3.6b: MN9 large mammal genera dendogram...... 103 Figure 3.6c: MN9 small mammal species dendrogram ...... 104 Figure 3.6d: MN9 small mammal genera dendogram ...... 105 Figure 3.7a: MN10 large mammal species dendrogram...... 110 Figure 3.7b: MN10 large mammal genera dendogram...... 111 Figure 3.7c: MN10 small mammal species dendrogram ...... 112 Figure 3.7d: MN10 small mammal genera dendogram ...... 113 Figure 3.8a: MN11 large mammal species dendrogram...... 118 Figure 3.8b: MN11 large mammal genera dendogram...... 119 Figure 3.8c: MN11 small mammal species dendrogram ...... 120 Figure 3.8d: MN11 small mammal genera dendogram ...... 121 Figure 3.9a: MN12 large mammal species dendrogram...... 127 Figure 3.9b: MN12 large mammal genera dendogram...... 128 Figure 3.9c: MN12 small mammal species dendrogram ...... 129 Figure 3.9d: MN12 small mammal genera dendogram ...... 130 Figure 3.10a: MN13 large mammal species dendrogram...... 136 Figure 3.10b: MN13 large mammal genera dendogram...... 137 Figure 3.10c: MN13 small mammal species dendrogram ...... 138 Figure 3.10d: MN13 small mammal genera dendogram ...... 139 Figure 3.11: Geographic distribution of faunal provinces...... 155 Figure 3.12a: Dispersal and in situ evolution in Eurasian faunas (MN5)...... 171 Figure 3.12b: Dispersal and in situ evolution in Eurasian faunas (MN6) ...... 171

vii

List of Figures (Continued)

Figure 3.12c: Dispersal and in situ evolution in Eurasian faunas (MN7/8)...... 171 Figure 3.12d: Dispersal and in situ evolution in Eurasian faunas (MN9) ...... 171 Figure 3.12e: Dispersal and in situ evolution in Eurasian faunas (MN10)...... 171 Figure 3.12f: Dispersal and in situ evolution in Eurasian faunas (MN11) ...... 171 Figure 3.12g: Dispersal and in situ evolution in Eurasian faunas (MN12) ...... 171 Figure 3.12h: Dispersal and in situ evolution in Eurasian faunas (MN13) ...... 171 Figure 3.13: West Asian-Eastern European dispersal routes...... 181 Figure 3.14: Dispersal routes into and trans-Pannonian Basin...... 183 Figure 4.1: Distribution of Griphopithecus localities...... 188 Figure 4.2a: MN3/5 dispersal pathways into Europe (modified from Popov et al. 2004) .. 197 Figure 4.2b: MN4 dispersal pathways into Europe via (modified from Rögl 1999. 199 Figure 4.3: Distribution of Dryopithecus localities (including Pierolapithecus) ...... 204 Figure 4.4: Late Astaracian Eurasian immigrations ...... 206 Figure 4.5: Distribution of localities ...... 218

viii

Chapter 1 - Introduction

On the Birthplace and Antiquity of Man – We are naturally led to enquire, where was the birthplace of man at that stage of descent when our progenitors diverged from the Catarhine stock? The fact that they belonged to this stock clearly shews that they inhabited the Old World; but not Australia nor any oceanic island, as we may infer from the laws of geographical distribution. In each great region of the world the living mammals are closely related to the extinct species of the same region. It is therefore probable that Africa was formerly inhabited by extinct apes closely allied to the and ; and as these two species are now man’s nearest allies, it is somewhat more probable that our early progenitors lived on the African continent than elsewhere… C. Darwin (1871, p182)

Since the publication of Darwin’s The Descent of Man, and Selection in Relation to

Sex in 1871, this passage has been well cited, illuminating Africa as the geographic place of origin and diversification of the African ape and human (hominine) lineage. An African origin hypothesis is logical on many counts. As Darwin (1871) pointed out, because our closest living relatives, the gorilla and chimpanzee, are restricted to Africa, the last common ancestor of African apes and humans most likely evolved there as well. Furthermore, the earliest fossil hominins (species more closely related to than to Pan) are similarly restricted to Africa. African paleoenvironments are thought to have provided the ecological requirements necessary for ancestral hominines to evolve there in the late Miocene (Bernor

1978; Cote 2004; Pilbeam 1997; Pilbeam & Young 2004). However, until recently, very few fossil apes were known from the late Miocene of Africa. Cote (2004) has suggested that this is a consequence of a poor fossil record, riddled with preservational and sampling biases, and

1 2 that of the late Miocene fossil localities that are known, very few sample the type of environment in which fossil apes could thrive.

Darwin’s (1871) speculation of Africa as the birthplace of the hominine lineage was immediately followed in the next sentence by an interesting observation; curiously enough, one that is largely overlooked:

…. But it is useless to speculate on this subject; for two or three anthropomorphous apes, one the Dryopithecus of Lartet, nearly as large as a man, and closely allied to , existed in Europe during the Miocene age; and since so remote a period the earth has undergone many great revolutions, and there has been ample time for migrations on the largest scale. C. Darwin (1871, p182-183)

A Eurasian origin for the Homininae is considered a minority view (Cote 2004; Pilbeam &

Young 2004; Stewart & Disotell 1998). Advocates of this hypothesis suggest that after an initial intercontinental dispersal from Africa to Eurasia at ~17Ma and subsequent radiation from Spain to China, fossil apes disperse into Africa at least once in the late Miocene. It is therefore probable that African hominines are derived from a Eurasian ancestor (Begun 2007,

2005, 2001; Begun et al. 1997; Heizmann & Begun 2001). Support for this hypothesis is derived from cladistic analyses, identifying synapomorphies supporting a sister relationship between late Miocene Eurasian and the African ape and human clade (Table 1.1).

These analyses specifically identify late Miocene Eurasian genera, such as Dryopithecus and

Ouranopithecus, as stem hominines, while the few contemporaneous specimens in Africa are considered to lack synapomorphies linking them to the African ape and human clade (Begun

2007, 2005, 2001, 1994; Begun et al. 1997). A Eurasian origin of the Homininae is also supported by the biogeographic and phylogenetic patterns in other land mammal groups

3

(Begun 2007, 2005; Folinsbee & Brooks 2007; Made 1999; Stewart & Disotell 1998) and is consistent with the molecular divergence dates of extant taxa (Stewart & Disotell 1998), while the African fossil record fails to preserve diagnostic hominines until the latest

Miocene.

Table 1.1: Morphological characters shared with African apes and Dryopithecus (from Begun 2005, p56)

The Eurasian origin hypothesis has been challenged on several grounds. First, there lacks consensus regarding the phylogenetic relations among Eurasian Miocene apes and between these taxa and African hominins, as a consequence of temporal, geographic and morphological unevenness in fossil representation, as well as differences in interpretation of the phylogenetic significance of preserved (Andrews 1992a; Andrews & Bernor

1999; Andrews et al. 1996; Andrews & Martin 1987; Begun 1995, 1994, 1992a; Begun &

4

Kordos 1997; Begun et al. 1997; Harrison & Rook 1997; Martin & Andrews 1993; Moyà-

Solà et al. 2004; Moyà-Solà & Köhler 1993; Pilbeam 1997, 1996; Pilbeam & Young 2004;

Ward et al. 1997). Second, the ancestral hominine morphotype has been reconstructed using what some consider crown ape synapomorphies, as a suspensory, frugivorous, tropical forest- dweller, most similar to Pan (Cote 2004; Pilbeam 1997, 1989; Pilbeam & Young 2004).

According to supporters of the African origin hypothesis, no Eurasian ape fits this morphotype and instead, the last common ancestor of African apes and humans has yet to be discovered in equatorial Africa. Lastly, from an ecological perspective, late Miocene

Eurasian paleoenvironments would not support intercontinental dispersals of Eurasian apes to

Africa, due to the lack of densely forested conditions and the unavailability of ripe fruit along dispersal corridors (Andrews 2007; Cote 2004; Pilbeam & Young 2004). As a result,

Eurasian apes underwent during or shortly after the [mid-] Vallesian Crisis at

~9.6Ma, when most forest-dwelling mammals succumbed to the cooler, drier, and more seasonal climates, and thus no relation to the African ape and human lineage

(McCrossin & Benefit 1997; Pilbeam 1997, 1996; Pilbeam & Young 2004).

Complete chronological records of mammalian migration and diversification patterns in the Eurasian Miocene are obscured by geological conditions, including a rarity of continental basin deposits preserving complete records of fossil mammals (Schmidt-Kittler

1989), the occurrence of isolated localities lacking stratigraphic control (usually due to tectonism), the resulting stratigraphic gaps (Agustí 1999; Fejfar & Heinrich 1989; Sen 1997,

1996) and a rarity of volcanic or datable sediments (Aguilar et al. 2004; Fahlbusch & Mein

1989; Kappelman et al. 1996; Schmidt-Kittler 1989; Sen 1996). As a result, the fossil record of mammalian distribution and evolution in most regions of Eurasia is discontinuous

5

(Fahlbusch & Mein 1989). According to Andrews (2007), this discontinuity of the Miocene fossil record specifically complicates the identification and analysis of biogeographic and morphological trends in fossil primates. One approach to this problem is to study mammalian groups that preserve more complete fossil records and are commonly associated with fossil apes (Andrews 2007).

In this study, the paleobiogeographic patterns of distribution, in situ evolution, interprovincial and intercontinental dispersal in Eurasian land mammals are used as a framework to clarify the spatial and temporal of dispersal and biogeographic relations among the Eurasian apes, with the purpose of 1) evaluating previous hypotheses regarding the relations among Eurasian fossil hominids, and 2) critically assessing the Eurasian origin hypothesis. Fossil mammals from localities encompassing the geographic expanse from

Spain to the Republic of are the focus of this study. However, localities in Saudi

Arabia and Africa are also included for consideration of faunal interchange between continents. Similarly, middle to late Miocene land mammals are the focus of this study, however the relations of these faunas to those of the are also considered. The data for this study are derived mostly from the Neogene of the Old World (NOW) and

Paleobiology (PBDB) databases, and to a lesser extent from survey and excavation of middle and late Miocene vertebrate localities in Hungary and Romania.

Chronological overview

Stratigraphic discontinuity across the Eurasian land mass has impeded magnetostratigraphic correlation (Kappelman et al. 1996), and as a result, the majority of

Eurasian Miocene localities used in this study rely on biochronologic methods to provide a relative chronological position of their constituent faunas (Bernor 1983). The Neogene

6

Mammal Zones (Mein 1975) is a biochronologic system of correlating fossil land mammal localities based on faunal composition that was implemented to systematize the changes occurring in Western European and Eastern Mediterranean Neogene fossil mammals (Agustí et al. 2001a). Each MN interval is designated by a set of criteria relating to its faunal composition, including the first and last appearance of mammal taxa in the fossil record

(FAD and LADs), the recognition of immigrant taxa, and taxa that are characteristic of specific temporal periods (Alroy et al. 1998; Bernor et al. 1996a & b; Kappelman et al. 1996;

Steininger et al. 1989; Swisher 1996; Woodburne et al. 1996). Since its original publication,

Mein has revised the MN zones and now recognizes 17 Neogene zones (13 Miocene, 4

Pliocene) (Table 1.2), two subzones (2a and 2b) and Quaternary zones (Q) (Guérin 1989;

Mein 1989, 1979).

Table 1.2: Temporal association of the Eurasian MN Zones

Time European Land Mammal (Ma) Mega-Zones MN Zones 4 5 6 MN13 7 Turolian MN12 8 MN11 late 9 MN10 10 Vallesian MN9 11 12 MN7/8 13 Astaracian middle 14 MN6 Miocene 15 16 MN5 17 MN4 18 Orleanian

19 MN3 20 early 21 MN2 22 Ageanian 23 MN1 24

7

Paleoenvironmental overview

The paleobiogeography of Eurasian Miocene land mammals, including primates, was largely influenced by the extensive climatic and geophysical change occurring throughout this epoch (Böhme 2003a) and thus requires a concise overview.

The early Miocene experienced warmer climates than the preceding epoch, due to a global excursion of the tropical belt during the late and early Miocene (Agustí &

Antón 2002; Böhme 2003b; Bojar et al. 2004; Pagani et al. 1999; Rögl 1999; Schwartz

1997). In the circum-Mediterranean region, subtropical to tropical temperatures prevailed, with humid conditions expanding throughout the area (Jones 1999; Rögl & Steininger 1984).

Towards the end of the early Miocene, a slight cooling event occurred and is documented through a drop in atmospheric CO2 levels. These levels decreased from five times pre- industrial levels at 20 Ma to one and a half times pre-industrial levels by 15 Ma, as a result of a significant expansion in the Antarctic ice sheet during this time interval (Schwartz 1997).

The most significant geophysical event during the early Miocene was the collision of the Afro-Arabian and Anatolian tectonic plates. This tectonic movement caused an interruption in the Tethys seaway (which joined the North Atlantic to the Indian Ocean), and introduced the first land corridor for continental mammal exchange between Eurasia and

Africa across Arabia and Asia Minor (Agustí & Antón 2002; Görur et al. 1995; Jones 1999;

Krijgsman 2002; Made 1999; Potts & Behrensmeyer 1992; Rögl 1999; Rögl & Steininger

1984; Steininger et al. 1985). The subtropical and largely forested conditions in Eurasia influenced the migration and subsequent radiation of several land mammal groups, including thickly-enamelled hominoid primates. While the African immigrant taxa were drawn to the

8 warmer climates of Europe, fossil assemblages in Africa document European taxa showing a preference for increased aridity (Begun et al. 2001; Görür et al. 1995; Steininger et al. 1985).

In addition to faunal migrations (beginning ~19Ma), the closure of the Tethys disrupted the circum-equatorial oceanic current system. This disruption is considered to have triggered global climatic cooling, as indicated in δ18O values and carbonate levels recovered from deep sea sediment cores (Bicchi et al. 2003; Bradley 1999; Frakes et al. 1992;

Steininger et al. 1985; Tarling 1978). δ18O values from benthic foraminifera, a sensitive indicator of dissolved oxygen levels in ancient sediments, indicate that the closure of the

Tethys corresponded to a global turnover in deep water circulation patterns. ‘Sluggish’ circulation with warmer bottom waters during the early Miocene changed to increased deep- water circulation with the cooling of bottom waters towards the late-early/early-middle

Miocene boundary (Gebhardt 1999).

During the early-middle Miocene, the Tethys seaway closed and re-opened intermittently, as a result of tectonic activity (Jones 1999; Rögl 1999; Steininger et al. 1985).

Sea surface temperatures derived from the δ18O values in benthic foraminifera ranged from a low of 15˚C in North Africa and the Southern Mediterranean to a high of 20˚C in the

Northern Mediterranean and Southern Europe (Jones 1999). Other indicators, such as corals

(indicating average annual temperatures of 16˚C), mangroves (indicating average annual temperatures of <20˚C), paleobotanical, and palynological data (indicating mean annual temperatures between 16-18˚C and mean annual precipitation between 1100 - 1300 mm) all suggest a continuation of the early Miocene tropical and subtropical climates (Ivanov et al.

2002).

9

Non-marine gastropods during this time indicate the establishment and re- establishment of warm conditions, however the absence of a significant number of species common in early Miocene deposits signify a distinct shift in the climate to cooler conditions

(Esu 1999). Palynological evidence indicates a trend towards cooler and drier conditions in the circum-Mediterranean region and throughout Central Europe at this time (Jones 1999).

Also associated with the late middle Miocene cooling trend is an abrupt divergence in temperature in low and high altitude regions, with a dramatic drop in high latitude temperatures (Schwartz 1997).

The cooler, drier climates of the middle Miocene have been attributed to the ongoing expansion of the Antarctic ice sheet (Agustí & Anton 2002; Agustí et al. 1999a; Bojar et al.

2004; Ivanov et al. 2002; Ohta et al. 2003; Pagani et al. 1999). This expansion is documented by both a change in the rate of 87Sr/86Sr increase at 14.8-14.6 Ma and increases in δ18O occurring between 14.5 and 14.0 Ma, with a major increase between 14.1 and 14.05

Ma, due to the development of a permanent Antarctic ice cap (Krijgsman et al. 1994).

From ~15.5 to 5 Ma, distinct shifts occurred in European floras from subtropical to warm-temperate taxa. Warm-temperate mesophytic forests with relatively low seasonality to warm-temperate open country floras with relatively higher seasonality predominated throughout most of Europe. These forests persisted throughout much of the later middle

Miocene in Central Europe (Bernor et al. 1993).

The late Miocene (~12 - 5 Ma) is considered to represent an important interval known as the ‘Cenozoic climatic deterioration’ (Agustí et al. 1997; van Dam & Weltje 1999). From

9 to 5 Ma, European terrestrial ecosystems were characterized by a significant expansion of more open vegetation, with deciduous woodlands beginning to replace the retreating

10 subtropical forests, and grasslands expanding on a global scale (Ivanov et al. 2003; Potts

2004a & b; Potts & Behrensmeyer 1992). Many land mammal groups (particularly those adapted to stable humid or closed environments and /folivores), including

European apes, closely tracked the retreating subtropical humid conditions, contracting their ranges from Western Europe and Eastern Europe/Western Asia from 12 to 9 Ma (Agustí et al. 2003; Agustí et al. 2001b; Andrews et al. 1996; Begun et al. 2001; Eronen et al. 2003;

Jones 1999; Made et al. 2003; Potts 2004a & b; Rook et al. 2000). A significant extinction event [the ‘(mid-) Vallesian Crisis’] is recorded in the fossil record of Eurasian hominids north of the Tropic of Cancer and other land mammal groups, in response to the onset of cooler, drier, and more seasonal climates (Agustí 2003; Agustí et al. 2003; Agustí et al.

1999a; Agustí et al. 1997; Begun et al. 2001; Hay et al. 2002; Heissig 2003; Jones 1999;

Pagani et al. 1999; Rook et al. 2000; van Dam & Weltje 1999). As a result, changes in faunal community structure occurred, with a growing predominance in large-bodied, open-country .

Climatic change at this time resulted from the closure of the Tethys and continued uplift of the Himalaya, Karakoram and Hindu Kush mountains (~9.6 Ma). These orogenies caused large scale changes in continental/marine distribution, global climatic circulation patterns and seasonality (Bradley 1999). The regression of the Paratethys at this time also caused a shift in ecosystems towards even greater seasonality and replacement of evergreen subtropical forests by deciduous woodlands (Rook et al. 2000).

Continued glacial expansion of the Antarctic ice sheet and the resulting drop in global sea levels, together with tectonic events, led to the most significant geophysical event of the late Miocene. The Messinian Salinity Crisis refers to the isolation and subsequent

11 desiccation of the Mediterranean Basin at 5.96 or 5.75 to 5.32Ma. During this event some consider that parts of, or the entire Mediterranean dried up, with water levels dropping perhaps as much as 1500m, thus allowing opportunities for faunal interchange between

Europe and North Africa (Clauzon et al. 1996; Costeur et al. 2007; Esteban 1996; Esteban et al. 1996; Hsü et al. 1977; Martín et al. 1999; Rögl & Steininger 1984; Spezzaferri et al.

1998; Steininger et al. 1985).

Overview of study

As previously mentioned, paleoenvironmental events have affected the temporal and geographic continuity of the fossil records of Eurasian land mammals to varying degrees. In the next chapter, I assess data quality by quantifying sampling completeness of the large and small mammals analysed in this study. The analysis in chapter 2 is necessary to identify temporal intervals, as well as taxa, that are incompletely sampled prior to the large scale analysis in the following chapter. Since primates are rare in the fossil record, the purpose of chapter 3 is to model the paleobiogeographic patterns of middle and late Miocene Eurasian mammalian taxa and changes in their distribution over time in response to environmental factors, with specific regard to climatic, tectonic and eustatic events. The resulting identification of distinct bioprovinces, including their faunal composition and interprovincial dispersals, is imperative for developing a framework to clarify the spatial and temporal dynamics of the Eurasian fossil apes. Using the results of this analysis, I address questions regarding the chronology, biogeography and evolutionary relations among Eurasian Miocene apes in chapter 4. Focusing specifically on Griphopithecus, Dryopithecus (including

Pierolapithecus) and Ouranopithecus, I examine previous ideas regarding the relations of these taxa to each other and to contemporaneous and subsequently occurring African fossils.

12

In chapter 5, my conclusions regarding the relations among Eurasian apes and the geographic origins of the hominine lineage are presented.

The approach used in this study is multidisciplinary in nature, integrating paleoanthropological, paleontological, geological and paleoenvironmental data together in a chronogeographic context. Intercontinental and regional differences in taxa have thus far been limited to descriptive, cladistic and statistical analyses of morphological and paleoenvironment. Although useful, these approaches make it difficult to consider, compare, and contrast multiple variables simultaneously. Mapping is an essential aspect of understanding spatial distribution and movement patterns in response to any variable and recent advances in Geographic Information Systems (herein referred to as GIS) technology have the potential to make significant contributions to paleoanthropological method and analysis.

Currently, GIS provide storage, dynamic mathematical manipulation, rapid retrieval and flexible visual presentation of large quantities of diverse, spatially referenced data (Green

1990; Lake et al. 1998). GIS are essentially spatially referenced computer databases that allow one to control for the distribution of form over space and through time. They are thus more useful than computerized cartography, and an application that is an integral part of mapping and data analysis in paleoanthropological field methods. The ability to present data in a 3- or 4- dimensional context provides a visual display of variables, while enabling pattern recognition within and between various types of data (Allen et al. 1990; Estrada Belli

1999; Koussoulakou & Stylianidis 1999; Potts et al. 1996). Furthermore, multiple data sets can be assessed and compared to each other, rather than as separate entities, allowing the

13 potential for regional, continental, or global analyses of the spatial patterning of specific variables.

GIS are infrequently used in paleoanthropology and paleobiology, despite the vast potential that this methodology offers for complex, multivariate analyses. No study to date has used GIS applications to explore questions pertaining to biogeographic relations among

Miocene apes, and specifically to the origins of the Homininae. However, given the relative rarity of Miocene fossil hominoids, it is only through using such methods that multiple lines of evidence can be considered simultaneously to understand the complexity of factors that led to the origin of living African apes and humans.

Chapter 2 – Data Quality

Introduction

Loss of Biological Information in the Fossil Record

During the transition of an organism from the biosphere to the lithosphere (the fossilization process), pertinent information regarding the organism as well as its surroundings is often lost due to the influence of various biological and biochemical processes. These processes, although responsible for producing concentrations of fossil remains, are also responsible for biasing the representation of species, individuals and skeletal elements, relative to their original community (Andrews 1996; Dreyer 1984;

Efremov 1958; Kidwell & Behrensmeyer 1993; Lyman 1994; Maas et al. 1995; Olson 1980).

The bias introduced by taphonomic processes tends not to affect all organisms to the same extent; the relationship between body mass and likelihood/state of preservation being a case in point. Small mammals (micromammals) weighing less than 5 kg (i.e., , insectivores, lagomorphs) tend to be underrepresented in fossil accumulations since they are easily broken into unidentifiable fragments by predators, excavators and natural processes

(i.e., trampling and weathering), and are more often overlooked for larger more diagnostic specimens (Andrews 1990; Behrensmeyer 1991; Behrensmeyer et al. 1979; Dreyer 1984;

Fernández-Jalvo 1995). With respect to the latter, differential recovery of fossils can be an additional source of bias. According to Lyman (1994), collector bias can significantly influence the type of fossils that are collected and the type of data being recorded. As a result, very few studies of Eurasian land mammals consider both small and large mammals together and those that do tend to focus on large mammals communities (i.e., Fortelius et al.

1996; Made 1999), despite good representation of small mammals in some

14 15 regions of Western ( and Spain) and Central Europe (Austria and Hungary).

Regardless of body size, data quality must be assessed prior to any analysis and evaluation of faunal communities.

In addition to taphonomic processes, other random factors can influence the likelihood of an organism being sampled in a fossil assemblage. Rare taxa (organisms occurring less frequently in the fossil record) and taxa with short temporal ranges are less likely to be sampled than common taxa (organisms occurring more frequently in the fossil record) and those with long temporal ranges, especially if sample sizes are small (Badgley

1990; Badgley & Gingerich 1988; Maas et al. 1995).

Sampling Completeness

Sampling quality is an evaluation of the state of preservation of a fossil assemblage relative to the original community from which it was derived (Krause & Mass 1990; Maas et al. 1995), and thus is a measure, albeit crude, of data quality. The consequences of incomplete sampling from taphonomic, biological or temporal factors include underestimates of both taxonomic richness (total number of taxa), as well as the timing of faunal turnover events. With respect to the latter, first and last occurrence datums not recognized in poorly sampled temporal intervals will appear to have occurred in more thoroughly sampled intervals (Maas et al. 1995), resulting in erroneous estimates of temporal range. In sum, issues stemming from incomplete sampling may lead to the incorrect identification of taxa thought to be involved in episodes of vicariance, or conversely, taxa actually involved may not be recognized as participatory. Either outcome results in dispersal events in which the timing of and taxa involved are potentially unclear or incorrect.

16

The purpose of this chapter is twofold: 1) to assess sampling completeness and thus quality in the dataset being used in this study, and 2) to use the results of this analysis to identify taxa/taxonomic groups that may serve as good ‘marker’ taxa to track patterns of dispersal, due to their overall temporal completeness and clarity in temporal range.

This analysis is necessary to identify both temporal intervals, as well as taxa that may be incompletely sampled, which would subsequently be analyzed with caution.

Materials and Methods

Spatio-temporal Distribution of Faunas

Following Fortelius et al. (2003a), modern geographic references (i.e., country boundaries) were used as neutral landmarks. The study area for this analysis spans geographically over the European and Western Asian landmass, from (~9.62594°W) to the Caspian Sea (~46.919506°E). Localities are distributed latitudinally from

~34.813211°N to ~52.459061°N (Figure 2.1). Temporally, the sample of localities spans from 18Ma (base of MN4) to 4.2Ma (top of MN14). This temporal duration predates the first appearance of primates and postdates the disappearance of non-cercopithecoid primates in

Europe and Western Asia in the late Miocene.

Figure 2.1: Distribution of Sample Localities

17

Method of Analysis

The completeness index (CI) of Krause & Maas (1990) and Maas et al. (1994) was used here to assess data quality. The CI is the percentage of mammal genera inferred to have been present within a temporal interval that have actually been found within that interval

(Fortelius et al. 1996; Krause & Maas 1990; Maas et al. 1994) and is calculated using the following equation:

CI = [Nt / (Nt + Nrt)] * 100

Where Nt = total number of genera known for an interval (MN unit) and Nrt is the number of range-through taxa (taxa found before and after, but not during a temporal interval). The presence of range-through taxa is inferred during these intervals under the assumption that they have not been sampled due to differential preservation or collecting bias, rather than their actual absence from the fossil record (Fortelius et al. 1996; Krause & Maas 1990; Maas et al. 1994). Temporal intervals with an index ≤ 70 are considered to be inadequately sampled, while indices increasing towards 100 are considered to indicate adequate sampling of faunas (Krause & Maas 1990; Maas et al. 1995). Maas et al. (1995) also implemented a strict completeness index (CIbda) using the following equation:

CIbda = [ Nbda / (Nbda + Nrt)] * 100 where Nbda is the number of taxa known before, during and after the interval, excluding first and last occurrences. According to Maas et al. (1995), the CIbda provides a more accurate estimate of completeness because it does not make assumptions about factors (with the exception of sampling) that affect the number of first and last appearances.

18

Data

The majority of the taxa used in this analysis are derived from the NOW Database

(Fortelius 2007; downloaded 02/07), however faunas recorded from survey and excavation efforts in Hungary and Romania from 2002-2007 were also integrated. Initially, faunas from localities spanning only a single MN unit were selected and then grouped according to their temporal interval (i.e., MN4, MN5, MN6, etc.). Taxa from these localities were sorted according to mammalian order, divided into the categories of ‘small’ (rodents, insectivores, lagomorphs) and ‘large’ (remaining orders) mammal and then scored as present or absent for each consecutive interval. I calculated both the regular (CI) and conservative (CIbda) indices of Maas et al. (1995). In addition, I revisited individual range-through taxa and further researched each using the literature as well as the Paleobiology Database to determine whether these temporal gaps were reasonably accurate. In these instances, taxa present at localities spanning one or more MN units, localities straddling the boundary of an MN unit

(i.e., MN9/10), and localities with paleomagnetic dates (having an absolute range falling within, but not including the whole duration of an MN unit) were considered within reason in a third CI calculation, referred to here as CI+u. In the first instance, localities with mixed fauna (i.e., having representatives from more than one MN unit) were in most cases considered, while localities with overall poor temporal resolution (i.e., where the age of the locality was indicated as “late Miocene”) were rejected. A total of 519 genera (222 small and

297 large mammal genera) from 973 localities (882 single unit and 91 + unit localities) were considered here.

Since the duration of each MN unit is unequal, it was necessary to test whether units representing longer temporal durations would also have a higher number of localities.

19

Similarly, it was necessary to test whether a larger number of sampled localities resulted in higher completeness indices. I used linear regression to test the null hypotheses that there was no relationship between variables in both cases. For example, in the latter case, the completeness index within an interval should be statistically independent of the number of localities to demonstrate that the index is indeed a measure reflective of sampling quality, rather than a proportionate measure of localities. In each analysis, the model assumptions were tested to ensure the appropriateness of the resulting model (Quinn & Keough 2002;

Sokal & Rohlf 1995).

Results

Relationship between duration of temporal intervals & number of localities/number of localities & CI

In general, the model assumptions of normality and equality of variance of the error terms were not violated. No significant relationship between the duration of each temporal interval and the number of sample localities per temporal interval was observed, nor was any significant relationship observed between the number of localities and CI (Figure 2.2a-g).

In all cases, the slope of the regression line for the small mammals was greater than in the large mammals (0.15 – 0.25 in the small mammals vs -0.01 – 0.06 in the large mammals), however these slope values were not significantly different from zero in most cases ( p >

0.05), demonstrating the lack of relationship between variables. In two cases, the significance of the small mammal CIbda and CI+u slope was rejected at the 0.05 level. In the small mammals, the relatively larger slopes might be due to the influence of a disproportionately high number of localities for some temporal intervals. For example,

20

MN5, MN7/8 and MN9 all have more than 115 localities, while all other intervals have less

91 localities or less.

Also, even though the co-efficient of determination was consistently higher for the small mammals (0.43-0.57) than the large mammals (0.01-0.09), it was still only explaining about half of the variation in the small mammal CIs. Among large mammals there is clearly no significant relationship between the number of localities and CI, but the number of localities and other factors may exert a weak influence on the CIs of the small mammals.

Figure 2.2: (a-g)

Interval duration vs # of localities

5

4.5

4

3.5

3

2.5

interval duration interval 2

1.5

1

0.5 y = 0.0038x + 1.1155 R2 = 0.0250 0 40 50 60 70 80 90 100 110 120 130 140 # of localities

21

CI lrg vs # of localities

100

90

80

70

60

50 CI lrg

40

30

20

10 y = 0.0630x + 85.2565 R2 = 0.0910 0 40 50 60 70 80 90 100 110 120 130 140 # of localities

CI sm vs # of localities

100

90

80

70

60

50 CI sm CI

40

30

20 y = 0.1586x + 70.4330 10 R2 = 0.4391

0 40 50 60 70 80 90 100 110 120 130 140 # of localities

22

CIbda lrg vs # of localities

100

90

80

70

60

50 CIbda lrg CIbda 40

30

20

10 y = 0.0382x + 81.8822 R2 = 0.0263 0 40 50 60 70 80 90 100 110 120 130 140 # of localities

CIbda sm vs # of localities

100

90

80

70

60

50 CIbda sm CIbda 40

30

20

10 y = 0.2580x + 54.7234 R2 = 0.5768 0 40 50 60 70 80 90 100 110 120 130 140 # of localities

23

CI+u lrg vs # of localities

100

90

80

70

60

50 CI+u lrg

40

30

20 y = -0.0161x + 96.4144 10 R2 = 0.0113

0 40 50 60 70 80 90 100 110 120 130 140 # of localities

CI+u sm vs # of localities

100

90

80

70

60

50 CI+u sm

40

30

20 y = 0.1852+ 68.2436 10 R2 = 0.5561

0 40 50 60 70 80 90 100 110 120 130 140 # of localities

24

Large Mammal Completeness

Using all three measures, the CI, CIbda, CI+u for the sample of large mammals fluctuated, but remained over the arbitrary cut off of 70, thus indicating that this group was well sampled (Figure 2.3 & Table 2.1). CI results for MN 4 and MN 14 were omitted since the preceding (MN 3) and subsequent (MN 15) intervals were not included and thus the CIs for MN 4 and MN 14 cannot be adequately evaluated (CI for both = 100).

Figure 2.3: Relative completeness in Eurasian large mammal taxa

Large Mammal Completeness

100

90

80

70

60 CI+u 50 CI CIbda 40 Completeness Index Completeness

30

20

10

0 MN 5 (17- MN 6 (15.2 - MN 7/8 (12.5- MN 9 (11.2- MN 10 (9.5- MN 11 (9- MN 12 (8.2- MN 13 (7.1- 15.2Ma) 12.5Ma) 11.2Ma) 9.5Ma) 9Ma) 8.2Ma) 7.1Ma) 5.3Ma) MN Unit

25

Table 2.1: Temporal ranges of Eurasian large mammal genera present range-through

Family MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Andegamerycidae Andegameryx Anthracotheriidae Brachyodus Anthracotheriidae Elomeryx Alephis Bovidae Aragoral Bovidae Austroportax Bovidae Caprotragoides Bovidae Criotherium Bovidae Bovidae Gazella Bovidae Graecoryx Bovidae Helicotragus Bovidae Helladorcas Bovidae Hispanodorcas Bovidae Hypsodontus Bovidae Kobus Bovidae Koufotragus Bovidae Maremmia Bovidae Mesembriacerus Bovidae Bovidae Moldoredunca Bovidae Nesogoral Bovidae Nisidorcas Bovidae Oioceros Bovidae Ouzocerus Bovidae Pachytragus Bovidae Palaeoreas Bovidae Palaeoryx Bovidae Parabos Bovidae Paratragocerus Bovidae Plesiaddax Bovidae Plioportax Bovidae Procobus Bovidae Prostrepsiceros Bovidae Protagocerus Bovidae Protoryx Bovidae Protragelaphus Bovidae Protragocerus Bovidae Pseudoeotragus Bovidae Pseudotragus Bovidae Sinotragus Bovidae Tethytragus Bovidae Tossunnoria Bovidae Bovidae Tragoreas Bovidae Turcocerus Bovidae Tyrrhenotragus Bovidae Urmiatherium Bovoidea Amphimoschus Bovoidea Hispanomeryx Cervidae Amphiprox Cervidae Cervavitulus Cervidae Cervidae Cervidae Croizetoceros Cervidae Dicrocerus Cervidae Euprox

26

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Cervidae Heteroprox Cervidae Lagomeryx Cervidae Lucentia Cervidae Neomegaloceros Cervidae Paracervulus Cervidae Paradicrocerus Cervidae Pliocervus Cervidae Procapreolus Cervidae Procervulus Cervidae Stehlinoceros Cervidae Stephanocemas Cervidae Turiacemas Cervoidea Ligeromeryx Cervoidea Orygotherium Cervoidea Palaeoplatyceros Birgerbohlinia Giraffidae Bohlinia Giraffidae Bramatherium Giraffidae Csakvarotherium Giraffidae Decennatherium Giraffidae Georgiomeryx Giraffidae Giraffa Giraffidae Giraffokeryx Giraffidae Giraffidae Orasius Giraffidae Palaeogiraffa Giraffidae Giraffidae Giraffidae Giraffidae Umbrotherium Haplobunodontidae Parabunodon Hippopotamidae Hexaprotodon Moschidae Amphitragulus Moschidae Micromeryx Palaeochoeridae Palaeochoerus Palaeochoeridae Schizochoerus Palaeochoeridae Taucanamo Ampelomeryx Palaeomerycidae Oriomeryx Palaeomerycidae Palaeomerycidae Triceromeryx Sanitheriidae Sanitherium Albanohyus Suidae Aureliachoerus Suidae Barberahyus Suidae Bunolistriodon Suidae Suidae Dicoryphochoerus Suidae Eumaiochoerus Suidae Hippopotamodon Suidae Hyotherium Suidae Kubanochoerus Suidae Suidae Microstonyx Suidae Suidae Parachleuastochoerus Suidae Propotamochoerus Suidae Sivachoerus Suidae Sus Suidae Xenohyus Tragulidae Dorcabune

27

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Tragulidae Dorcatherium Ampictis Amphicyonidae Agnotherium Amphicyonidae Amphicyonidae Amphicyonopsis Amphicyonidae Cynelos Amphicyonidae Euroamphicyon Amphicyonidae Haplocyonoides Amphicyonidae Ictiocyon Amphicyonidae Pseudarctos Amphicyonidae Pseudocyon Amphicyonidae Thaumastocyon Amphicyonidae Ysengrinia Canis Canidae Eucyon Canidae Canidae Vulpes Felidae Felidae Epimachairodus Felidae Felis Felidae Fortunictis Felidae Lynx Felidae Felidae Felidae Miomachairodus Felidae Felidae Felidae Stenailurus Hyaenidae Hyaenidae Belbus Hyaenidae Chasmaporthetes Hyaenidae Hyaenidae Hyaenictis Hyaenidae Hyaenictitherium Hyaenidae Hyaenotherium Hyaenidae Ictitherium Hyaenidae Lycyaena Hyaenidae Miohyaenotherium Hyaenidae Plioviverrops Hyaenidae Hyaenidae incertae sedis Alopecocyon incertae sedis Amphictis incertae sedis Palaeogale incertae sedis Simocyon Anatolictis Mustelidae Baranogale Mustelidae Circamustela Mustelidae Enhydriodon Mustelidae Eomellivora Mustelidae Hadrictis Mustelidae Hoplictis Mustelidae Iberictis Mustelidae Ischyrictis Mustelidae Laphyctis Mustelidae Limnonyx Mustelidae Lutra Mustelidae Marcetia Mustelidae Martes

28

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Mustelidae Meles Mustelidae Melidellavus Mustelidae Mellivora Mustelidae Mesomephitis Mustelidae Mionictis Mustelidae Mustela Mustelidae Palaeomeles Mustelidae Paludolutra Mustelidae Pannonictis Mustelidae Paraenhydriodon Mustelidae Paralutra Mustelidae Parataxidea Mustelidae Plesiogale Mustelidae Mustelidae Plesiomeles Mustelidae Promeles Mustelidae Promephitis Mustelidae Proputorius Mustelidae Rhodanictis Mustelidae Sabadellictis Mustelidae Sivaonyx Mustelidae Stromeriella Mustelidae Taxodon Mustelidae Trocharion Mustelidae Trochictis Mustelidae Trochotherium Mustelidae Tyrrhenolutra Mustelidae Vormela Nimravidae Prosansanosmilus Nimravidae Sansanosmilus Percrocutidae Allohyaena Percrocutidae Dinocrocuta Percrocutidae Percrocuta Procyonidae Simocyon Procyonidae Viretius Ursidae Ursidae Ursidae Ursidae Ursidae Metarctos Ursidae Ursidae Ursidae Herpestes Viverridae Jourdanictis Viverridae Leptoplesictis Viverridae Lophocyon Viverridae Schlossericyon Viverridae Semigenetta Viverridae Sivanasua Viverridae Viverra Viverridae Viverrictis Hyaenodontidae Hyainailouros Chalicotheriidae Chalicotheriidae Chalicotheriidae Macrotherium Chalicotheriidae Metaschizotherium Chalicotheriidae Phyllotillon Equidae Anchitherium Equidae Cremohipparion Equidae

29

Small Mammal Completeness

Using CI and CI+u, the small mammal completeness fluctuated, but remained well above the arbitrary cutoff of 70. However, using the more conservative CIbda, completeness fell below the cutoff in MN 6 (69.1) and MN 12 (65.5) (Figure 2.4 & Table 2.2).

Figure 2.4: Relative completeness in Eurasian small mammal taxa

Small Mammal Completeness

100

90

80

70

60

CI+u 50 CI CIbda 40 Completeness Index

30

20

10

0 MN 5 (17- MN 6 (15.2 - MN 7/8 (12.5- MN 9 (11.2- MN 10 (9.5- MN 11 (9- MN 12 (8.2- MN 13 (7.1- 15.2Ma) 12.5Ma) 11.2Ma) 9.5Ma) 9Ma) 8.2Ma) 7.1Ma) 5.3Ma) MN Unit

30

Table 2.2: Temporal ranges of Eurasian small mammal genera present range-through

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Anchitheriomys Castoridae Castor Castoridae Chalicomys Castoridae Castoridae Eucastor Castoridae Euroxenomys Castoridae Palaeomys Castoridae Steneofiber Castoridae Ctenodactylidae Sayimys Dipodidae Allactaga Dipodidae Eozapus Dipodidae Protalactaga Dipodidae Sarmatosminthus Apeomys Eomyidae Eomyodon Eomyidae Eomyops Eomyidae Estramomys Eomyidae Keramidomys Eomyidae Leptodontomys Eomyidae Ligerimys Eomyidae Megapeomys Eomyidae Pseudotheridomys Eomyidae Rhodanomys Gerbillidae Debruijnimys Gerbillidae Pseudomeriones Gliridae Anthracoglis Gliridae Armantomys Gliridae Bransatoglis Gliridae Carbomys Gliridae Eliomys Gliridae Eomuscardinus Gliridae Glirudinus Gliridae Glirulus Gliridae Glis Gliridae Graphiurops Gliridae Heteromyoxus Gliridae Margaritamys Gliridae Microdryomys Gliridae Miodyromys Gliridae Muscardinus Gliridae Myoglis Gliridae Myomimus Gliridae Paraglirulus Gliridae Peridiromys Gliridae Praearmantomys Gliridae Prodryomys Gliridae Pseudodryomys Gliridae Ramys Gliridae Stertomys Gliridae Tempestia Gliridae Vasseuromys Hystricidae Hystrix Muridae Allocricetus Muridae Anomalomys Muridae Anthracomys Muridae Apocricetus Muridae Apodemus Muridae Blancomys Muridae Bujoromys

31

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Muridae Byzantinia Muridae Calomyscus Muridae Castillomys Muridae Castromys Muridae Celadensia Muridae Centralomys Muridae Collimys Muridae Cotimus Muridae Cricetodon Muridae Cricetulodon Muridae Cricetulus Muridae Cricetus Muridae Dakkamys Muridae Debruijnia Muridae Debruinimys Muridae Democricetodon Muridae Deperetomys Muridae Dicrostonyx Muridae Enginia Muridae Epimeriones Muridae Eumyarion Muridae Fahlbuschia Muridae Hansdebruijnia Muridae Hattomys Muridae Heramys Muridae Hispanomys Muridae Huerzelerimys Muridae Karnimata Muridae Karydomys Muridae Kowalskia Muridae Lartetomys Muridae Megacricetodon Muridae Melissiodon Muridae Mesocricetus Muridae Micromys Muridae Microtia Muridae Microtocricetus Muridae Microtodon Muridae Microtoscoptes Muridae Microtus Muridae Mimomys Muridae Mirabella Muridae Mixocricetodon Muridae Myocricetodon Muridae Neocometes Muridae Occitanomys Muridae Orientalomys Muridae Paraethomys Muridae Parapodemus Muridae Pliospalax Muridae Progonomys Muridae Promimomys Muridae Prosomys Muridae Prospalax Muridae Protatera Muridae Pseudocricetus Muridae Pterospalax Muridae Rhagamys Muridae Rhagapodemus

32

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Muridae Rhodomys Muridae Rotundomys Muridae Ruscinomys Muridae Sinapospalax Muridae Stephanomys Muridae Trilophomys Muridae Turkomys Muridae Valerymys Pedetidae Megapedetes Petauristidae Albanensia Pteromyidae Aliveria Pteromyidae Blackia Pteromyidae Forsythia Pteromyidae Hylopetes Pteromyidae Miopetaurista Pteromyidae Pliopetaurista Sciuridae Atlantoxerus Sciuridae Csakvaromys Sciuridae Heteroxerus Sciuridae Palaeosciurus Sciuridae Ratufa Sciuridae Sciurus Sciuridae Spermophilinus Sciuridae Tamias Spalacidae Debruijina Zapodidae Heterosminthus Pliohyrax Dimylidae Chainodus Dimylidae Metacordylodon Dimylidae Plesiodimylus Amphechinus Erinaceidae Erinaceidae Galerix Erinaceidae Lanthanotherium Erinaceidae Mioechinus Erinaceidae Palerinaceus Erinaceidae Postpalerinaceus Erinaceidae Schizogalerix Heterosoricidae Dinosorex Heterosoricidae Heterosorex Plesiosoricidae Plesiosorex Soricidae Alloblarinella Soricidae Alloscapanus Soricidae Allosorex Soricidae Amblycoptus Soricidae Angustidens Soricidae Anourosorex Soricidae Asoriculus Soricidae Soricidae Cokia Soricidae Soricidae Crusafontina Soricidae Deinsdorfia Soricidae Dinosorex Soricidae Soricidae Florinia Soricidae Hemisorex Soricidae Lartetium Soricidae Limnoecus Soricidae Mafia

33

Family Genus MN 4 MN 5 MN 6 MN 7-8 MN 9 MN 10 MN 11 MN 12 MN 13 MN 14 Soricidae Miosorex Soricidae Neomysorex Soricidae Paenelimnoceus Soricidae Paranourosorex Soricidae Petenyia Soricidae Plesiosorex Soricidae Soricidae Soricella Soricidae Sulimskia Soricidae Zelceina Archaeodesmana Talpidae Asthenoscapter Talpidae Desmana Talpidae Desmanella Talpidae Desmanodon Talpidae Domninoides Talpidae Leptoscaptor Talpidae Mygalea Talpidae Myxomygale Talpidae Parascalops Talpidae Proscapanus Talpidae Ruemkelia Talpidae Scalopoides Talpidae Scapanulus Talpidae Scaptonyx Talpidae Suleimania Talpidae Talpidae Tenuibrachiatum Talpidae Theratiskos Talpidae Urotrichus Leporidae Alilepus Leporidae Hispanolagus Leporidae Hypolagus Leporidae Lepus Leporidae Trischizolagus Ochotonidae Albertona Ochotonidae Alloptox Ochotonidae Amphilagus Ochotonidae Eurolagus Ochotonidae Lagopsis Ochotonidae Ochotona Ochotonidae Paludotona Ochotonidae Prolagus Ochotonidae Proochotona

Identification of “Marker Taxa”

Within each mammalian order there occurred genera with overall complete fossil records and long temporal durations and would thus serve as good “marker” taxa to clarify patterns of primate dispersal. Specifically, many of these genera are considered to have similar niche requirements and/or are associated with non-cercopithecoid primates at fossil localities. These taxa are listed in the following tables (Tables 2.3 & 2.4). Note that for many genera, their last appearance actually occurs in the , however they are listed with a LAD of MN13 since Pliocene taxa are not being considered here.

34

Table 2.3: Temporal range and paleoecology of well-sampled Eurasian small mammals * = co-occurs with a non-cercopithecoid primate (E) = environmental tolerance of extant form/relative

Taxon Ecology Temporal Range Reference Lagomorpha Ochotonidae: MN5-10 Boon-Kristkoiz & Kristkoiz Eurolagus* (1999) Lagopsis * MN4-9 Boon-Kristoiz & Kristkoiz (1999) Prolagus* MN4-13 Boon-Kristoiz & Kristkoiz (1999) Sciuridae: MN4-7/8 Bruijn (1999) Palaeosciurus* Spermophilinus* MN4-13 Bruijn (1999) Tamias* arboreal MN7/8-13 Bruijn (1999) Heteroxerus* MN4-13 Bruijn (1999) Pteromyidae/ Petauristidae: Albanensia* arboreal MN5-10 Bruijn (1999) Miopetaurista* arboreal MN4-11 Bruijn (1999) Hylopetes* arboreal MN7/8-13 Bruijn (1999) Blackia* arboreal MN4-11 Bruijn (1999) Castoridae: Chalicomys* semi-aquatic MN4-12 Hugueney (1999) Steneofiber* semi-aquatic MN4-11 Hugueney (1999) Trogontherium* semi-aquatic MN4-11 Hugueney (1999) Gliridae: Bransatoglis* MN4-10 Daams (1999) Eliomys forest & open (E) MN9-13 Daams (1999) Glirudinus* MN4-9 Daams (1999) Glirulus* forest, arboreal (E) MN4-13 Daams (1999) Microdryomys* MN4-11 Daams (1999) Miodryomys* MN4-9 Daams (1999) Muscardinus* woodland (E) MN4-13 Daams (1999) Myoglis* MN4-10 Daams (1999) Tempestia MN6-10 Daams (1999) Eomyidae: Eomyops* humid forest MN5-11 Daams & Van Der Meulen (1984), Engesser (1999) Keramidomys* humid forest MN4-13 Engesser (1999) Dipodidae/Zapodidae: Eozapus* MN10-13 Daxner-Höck (1999) Muridae: Byzantinia* MN7/8-12 Rummel (1999) Cricetodon* forest MN4-9 Rummel (1999) Microtocricetus* moist riparian MN7/8-10 Fejfar (1999) Eumyarion* wet, wooded MN4-10 Kälin (1999), Weerd & Daams (1978) Megacricetodon* ecologically MN4-10 Kälin (1999) differentiated Democricetodon* wet, wooded MN5-10 Kälin (1999), Weerd & Daams (1978)

35

Taxon Ecology Temporal Range Reference Kowalskia* MN7/8-13 Kälin (1999) Cricetulodon* MN7/8-12 Kälin (1999) Neocometes* arboreal (E) MN4-7/8 Freudenthal & Martin Suárez (1999) Progonomys* MN7/8-12 Freudenthal & Martin Suárez (1999) Occitanomys MN10-13 Freudenthal & Martin Suárez (1999) Muridae/ Anomalomyidae: Anomalomys* later forms prefer MN4-11 Bolliger (1996, 1999) Prospalax later forms prefer steppe MN10-12 Bolliger (1996, 1999) Insectivora Erinacidae: Lanthanotherium* forests, proximity to MN4-11 Ziegler (1999) water (E) Galerix* forests, proximity to MN4-13 Ziegler (1999) water (E) Schizogalerix* forests, proximity to MN4-13 Ziegler (1999) water (E) Amphechinus* MN4-7/8 Ziegler (1999) Mioechinus* MN5-7/8 Ziegler (1999) Plesiosoricidae: Plesiosorex* MN4-9 Ziegler (1999) Talpidae: Archaeodesmana* MN9-13 Ziegler (1999) Desmanella* MN7/8-13 Ziegler (1999) Proscapanus* MN4-9 Ziegler (1999) Talpa* MN5-13 Ziegler (1999) Dimylidae: Plesiodimylus* MN4-11 Ziegler (1999) Metacordylodon* MN5-9 Ziegler (1999) Soricidae: Amblycoptus MN10-13 Ziegler (1999) Blarinella MN10-13 Ziegler (1999) Crusafontina* forest, proximity to water MN7/8-12 Ziegler (1999) Dinosorex* MN4-11 Ziegler (1999) Miosorex* MN4-10 Ziegler (1999) Paenelimnoecus* MN6-13 Ziegler (1999)

36

Table 2.4: Temporal range and paleoecology of well-sampled Eurasian large mammals * = co-occurs with a non-cercopithecoid primate (E) = environmental tolerance of extant form/relative

Taxon Ecology Temporal Range Reference Artiodactyla Bovidae: Eotragus* lightly wooded (lrg body MN4-7/8 Gentry et al. (1999) size), closed Miotragocerus* MN5(6?) -12 Gentry et al. (1999) Palaeoreas* grazer MN9-13 Gentry et al. (1999) Prostrepsiceros* grazer MN9-12 Gentry et al. (1999) Tethytragus* closed MN5-7/8 Gentry et al. (1999) Tragoportax* MN9-13 Gentry et al. (1999) Cervidae: Dicrocerus* forest (E), wooded, MN5-9 Gentry et al., (1999) proximity to water Euprox* MN5-10 Gentry et al. (1999) Lagomeryx* forest, wooded, + MN4-7/8 Gentry et al. (1999) underbrush Procapreolus* MN9-13 Gentry et al. (1999) Stehlinoceros* wooded MN5-7/8 Gentry et al. (1999) Giraffidae: Bohlinia* MN10-13 Gentry et al. (1999) Giraffokeryx* MN5-7/8 Gentry et al. (1999) Helladotherium* associated with open MN10-13 Gentry et al. (1999) fauna Paleotragus* associated with forest/ MN7/8-13 Gentry et al. (1999) steppe/mixed/open fauna Samotherium* mixed feeder/grazer, MN9-13 Gentry et al. (1999) associated with mixed/ open fauna Palaeochoeridae: Taucanamo forest MN4-9 Fortelius, Made & Bernor (1996) Suidae: Aureliachoerus* MN4-6 Hünermann (1999) Bunolistriodon* heterogeneous environment, open MN4-7/8 Fortelius, Made & Bernor (1996), Hünermann (1999) Conohyus* marshy forest, not open MN5-9 Fortelius, Made & Bernor (1996), Hünermann (1999), Thenius (1952) Hippopotamodon* woodland/grassland MN7/8-10 Fortelius, Made & Bernor mosaic, open (1996), Hünermann (1999) Hyotherium* not open MN4-6 Fortelius, Made & Bernor (1996), Hünermann (1999) Listriodon* heterogeneous MN4-9 Fortelius, Made & Bernor environment, open (1996), Hünermann (1999) Microstonyx* woodland/grassland MN9-13 Fortelius, Made & Bernor mosaic (1996), Hünermann (1999) Paraleuasto- not open MN6-11 Fortelius, Made & Bernor choerus* (1996), Hünermann (1999) Propotamo- heterogeneous MN7/8-13 Fortelius, Made & Bernor

37

Taxon Ecology Temporal Range Reference choerus* environment, open (1996), Hünermann (1999) Tragulidae: Dorcatherium* forest, + undergrowth, MN4-13 Gentry et al. (1999) proximity to water, semi- aquatic Moschidae: Micromeryx* forest MN4-11 Gentry et al. (1999) Palaeomerycidae: Palaeomeryx* boggy forest MN4-9 Gentry et al. (1999) Amphicyonidae: Agnotherium* subtropical forest MN5-9 Ginsburg (1999), Werdelin (1996) Amphicyon* subtropical forest MN4-9 Ginsburg (1999), Werdelin (1996) Pseudarctos subtropical forest MN4-9 Ginsburg (1999), Werdelin (1996) Pseudocyon subtropical forest MN4-7/8 Ginsburg (1999), Werdelin (1996) Felidae: Amphi- MN10-13 Ginsburg (1999) machairodus* Machairodus* MN9-12 Ginsburg (1999) Paramachairodus* MN9-13 Ginsburg (1999) Pseudaelurus* MN4-10 Ginsburg (1999) Hyaenidae: Adcrocuta* MN10-13 Ginsburg (1999) Plioviverrops* more terrestrial than MN7/8-13 Ginsburg (1999), Werdelin arboreal (1996) Protictitherium* semi-arboreal MN4-11 Ginsburg (1999), Werdelin (1996) Thalassictic* more open, MN7/8-13 Bernor (1983), Ginsburg diversification follows (1999), Werdelin (1996) increase in middle and late Miocene seasonal woodlands Mustelidae: Ischyrictis* MN4-9 Ginsburg (1999) Martes* MN4-12 Ginsburg (1999) Plesiogulo* MN9-13 Ginsburg (1999) Proputorius* MN5-9 Ginsburg (1999) Nimravidae: Sansanosmilus* MN5-9 Ginsburg (1999) Percrocutidae: Percrocuta* MN5-13 Ginsburg (1999) Ursidae: Hemicyon* MN4-9 Ginsburg (1999) Indarctos* MN7/8-13 Ginsburg (1999) Plithocyon* MN4-7/8 Ginsburg (1999) Ursavus* MN4-9(10?), 11 Ginsburg (1999) Viverridae: Semigenetta* MN4-10 Ginsburg (1999) Sivanusua* MN4-7/8 Ginsburg (1999) Perissodactlya Rhinocerotidae:

38

Taxon Ecology Temporal Range Reference Hoplo- non-grazer MN4-9 Heissig (1999a) * Aceratherium* non-grazer, swamp, MN4-13 Heissig (1999a) proximity to water Brachypotherium* rain forest/steppe, high MN4-7/8 Heissig (1999a) vegetation, co-occurs with * MN4-9 Heissig (1999a) Dicerorhinus* high vegetation MN4-9 Heissig (1999a) Dihoplus* swamp, proximity to MN9-13 Heissig (1999a) water Ceratotherium* savanna MN9-13 Heissig (1999a) Chalicotheriidae: Chalicotherium* woodland MN4-12 Heissig (1999b) Equidae: Anchitherium only co-occurs with MN4-9 Bernor & Armour-Chelu Hipparion where forest (1999) conditions known to occur Cremohipparion* strong browser/mixed MN9-13 Bernor et al. (1996c), Bernor feeder & Armour-Chelu (1999) Hipparion* swamp-browser/savanna- MN9-13 Bernor et al. (1996c), Bernor grazer & Armour-Chelu (1999) * subtropical to warm MN7/8-13 Bernor et al. (1996c), Bernor temperate woodlands, & Armour-Chelu (1999) browser Deinotheriidae: Deinotherium* forest MN4-13 Göhlich (1999) Prodeinotherium* forest MN4-10 Göhlich (1999) Gomphotheriidae: * MN5-13 Göhlich (1999) * MN4-9 Göhlich (1999) Tetralophodon* MN7/8-13 Göhlich (1999) Zygolophodon* MN4-9(10?) Göhlich (1999) Tubulidentata Orycteropodidae: Orycteropus* rainforest-savanna (E) MN4-7/8(9?)10-13 Heissig (1999c) Discussion

Overall Completeness and Data Quality

The large and small mammals included in this analysis can be considered relatively completely sampled, with the exception of the small mammals in MN 6 and MN 12 when using the most conservative completeness method. During these two intervals, it is clear that relative incompleteness has no relation to the number of localities sampled (76 localities in

39

MN 6 and 92 localities in MN 12), since the lowest numbers of localities were found in MN units that were relatively complete (i.e., 54 localities in MN 11 and 62 localities in MN 13).

Therefore, further analysis of small mammals within these particular intervals is focused on the smaller number of taxa with complete temporal ranges. Although the small mammal fossil record is more prone to taphonomic bias, it is clear that a robust subset of taxa from the original sample have complete temporal ranges (Table 2.2 & 2.3).

The motivation to include small mammals in this and subsequent analyses stems from the fact that they provide one of the principle means for interpreting past environments and habitats (Andrews 1990; Avery 2003, 1982; Fernández-Jalvo 1995; Fernández-Jalvo et al.

1998). Extant small mammals are known from a diversity of ecological zones and are very sensitive to environmental change. They demonstrate very strong, sometimes narrow habitat preferences and since they are primary consumers, their distribution closely follows patterns of vegetation. In many cases, small mammals can be much stronger predictors of ecological variables (i.e., temperature, humidity/precipitation, vegetation structure) than the large mammals with which they are often preserved (Andrews 1990; Wesselman 1995).

Furthermore, morphological specialization and rapid evolution in some small mammal taxa make them ideal for biostratigraphic and biogeographic correlations (Benammi et al. 1996;

Ni & Qiu 2002). Therefore, even though small mammals are commonly excluded from large scale analyses of Eurasian land mammal provinciality and biogeography for the reasons mentioned above, the effort to assess sampling quality and include small mammals in subsequent analysis is warranted, due to the potential information they can provide.

40

Ecology of “Marker Taxa”

Although many of the large (and some small) mammals have unknown paleoenvironments or have morphology diagnostic of ecological environments uncharacteristic for primates (i.e., grazers, taxa known from more open areas), it is worth noting that most are also found in association with non-cercopithecoid primate fossils. This typically occurs in mosaic environments (i.e., Rudabánya in MN9, Merceron et al. 2007) or in later Miocene localities in , which happen to sample drier more seasonal environments that are possibly less forested (possibly more wooded) than contemporaneous localities in Western Europe (Andrews et al. 1997; Bonis et al. 1999; Merceron et al. 2007,

2005a & b; Solounias et al. 1999).

Distribution of sample localities

The fossil localities used in this and subsequent analyses are unevenly distributed, both spatially and temporally (Figure 2.1). Therefore, in this study, even though many taxa demonstrate relatively complete temporal ranges, these same ranges may not be complete at the regional or provincial level. Previously, Fortelius et al. (1996) divided Europe into East and West faunal blocks and evaluated CI for large mammals only (Figure 2.5). The CI of

Fortelius et al. (1996) corresponds to the CI+u used here and was assessed on a smaller dataset. Nargolwalla (2007) and Nargolwalla et al. (2006) assessed CI (corresponds to CI+u used here) for the large and small mammals from the Pannonian Basin of Central Europe

(Figure 2.5).

For the large mammals, fauna from the West and East are clearly contributing heavily to the total completeness in Europe and Western Asia in MN7/8, MN10 and MN12, while faunas restricted to the Pannonian Basin are incomplete during these intervals (Figure 2.6).

41

Interestingly, incompleteness in MN7/8 in the Pannonian Basin large mammal faunas is likely due to sampling strategies that focus almost exclusively on small mammal recovery at new localities and renewed excavations in Hungary (i.e., Felsőtárkány) and Romania (i.e.,

Subpiatrǎ) (Nargolwalla et al. 2006). Furthermore, in MN10 and MN12, incompleteness in

Figure 2.5: Regional or Bioprovincial zones previously assessed for CI

Pannonian Basin (Nargolwalla 2006; Nargolwalla et al. 2007) Division of East and West (Fortelius et al. 1996)

the Pannonian Basin faunas is due to a low number of localities during these temporal intervals, coupled with a very low number of genera and a higher number of range-through taxa. Conversely, relative completeness of faunas in the Pannonian Basin, particularly in

MN6, is largely influenced by the presence of rich localities [i.e., Göriach, Austria; Devínska

Nová Ves (Neudorf-Spalte), Slovakia], that are few in number. The occurrence of regional or bioprovincial variation in relative completeness suggests that despite overall relative completeness for the European and Western Asian land mass, the subsequent analysis of

42 provinciality (Chapter 3) should ensure that when considering taxa that disperse between regions or bioprovinces, these taxa demonstrate complete temporal ranges.

As mentioned previously, Fortelius et al. (1996) did not assess the CI of the small mammals within the defined East and West faunal blocks due to sampling and taphonomic issues associated with small mammals. However, comparison of the small mammal CIs for

Europe and Western Asia with the small mammals from the Pannonian Basin (Figure 2.7) reveals that during many intervals, relative completeness in Europe as a whole is being driven by the overall completeness of the small mammals from Western Asia. On the other hand, even though Fortelius et al. (1996) did not include small mammals in their analysis, it is likely that the small mammal faunas from East and/or West are influencing the overall relative completeness in MN6, MN11 and MN12. In MN6 and MN11, it is possible that small mammal faunas from either or both of these regions are relatively less complete (and thus decreasing overall relative completeness), while conversely, in MN12, the opposite is likely, with an increase in relative completeness of these faunas buffering the effect of considerable relative incompleteness in the Pannonian Basin small mammals. Therefore, as with the large mammals, it is important to recognize that although relative generic completeness is adequate for the European and Western Asian small mammal faunas, there is regional or bioprovincial variation in this index.

43

Figure 2.6: Regional and total large mammal completeness

CI comp lrg

100

90

80

70

60 CI total PB lrg 50 W E 40 Completeness Index Completeness

30

20

10

0 MN 5 (17- MN 6 (15.2 - MN 7/8 (12.5- MN 9 (11.2- MN 10 (9.5- MN 11 (9- MN 12 (8.2- MN 13 (7.1- 15.2Ma) 12.5Ma) 11.2Ma) 9.5Ma) 9Ma) 8.2Ma) 7.1Ma) 5.3Ma) MN Unit

Figure 2.7: Regional and total small mammal completeness

sm

100

90

80

70

60

CI total 50 PB sm

40 Completeness Index

30

20

10

0 MN 5 (17- MN 6 (15.2 - MN 7/8 (12.5- MN 9 (11.2- MN 10 (9.5- MN 11 (9- MN 12 (8.2- MN 13 (7.1- 15.2Ma) 12.5Ma) 11.2Ma) 9.5Ma) 9Ma) 8.2Ma) 7.1Ma) 5.3Ma) MN Unit

44

Chapter Summary & Conclusions

o There was no relationship observed between the duration of an MN unit and the

number of localities per unit, nor was there a relationship observed for the number of

localities per interval and the resulting CI measures.

o Overall, large mammal faunas were relatively complete for the time period of interest

using all three completeness indices.

o Overall, small mammal faunas were relatively complete using the CI and CI+u,

although relatively incomplete in MN6 and MN12 using the CIbda.

o Despite these findings, comparison with previous studies demonstrates that relative

completeness across Europe and Western Asia is not homogeneous and that regional

or bioprovincial variation in completeness must be assessed prior to consideration of

dispersal events between these regions or bioprovinces.

Chapter 3 – Provinciality, paleoenvironments, in situ evolution and dispersal

Introduction

Zoogeographic Provinces and Purpose of Study

Zoogeographic or biotic provinces (bioprovinces) are geographic regions occupied by taxonomic communities that differ from those in adjacent regions (Heikinheimo et al. 2007;

Middlemiss & Rawson 1969; Ross 1974). Provincial stability is attained through the adaptation over time of these taxa to their environmental setting (Durden 1974). While bioprovinces were originally defined on the basis of both faunistic and floristic elements

(Vestal 1914), more commonly studies primarily emphasize faunal uniformity, measured through proportions of endemic and cosmopolitan taxa, and then subsequently look to the environmental conditions in which this uniformity occurs. The zoogeographic distribution of mammals and distinction of faunal bioprovinces is thus spatially dependent on environmental factors and the of the component fauna to these factors (Bellier et al. 2007; Ross

1974). These factors can act as potentially restrictive barriers through which all, some or none of the component species can pass and include regional topography (tectonic or marine barriers including uplift of mountain chains and fluctuations in bodies of water) and climate

(differences in temperature, humidity, etc.), both of which affect proximate vegetation structure. Significant faunistic differences among relatively contemporaneous taxa of similar depositional context indicate the presence of a restrictive barrier, while resemblance suggests connectivity among regions (Middlemiss & Rawson 1969; Raup & Crick 1979; Simpson

1940).

Since primates are considered rare in the fossil record, it is difficult to study their paleobiogeographic trends and assess phylogenetic connectivity at the genus and species

45 46 level. Therefore, one alternative is to examine the paleobiogeographic, provincial and evolutionary trends in other mammalian groups often found in association with primates, which have more complete fossil records and clearer patterns of overall distribution and speciation (Andrews 2007).

The purpose of delineating faunal bioprovinces in this chapter is to model the overall distribution of middle and late Miocene Eurasian mammalian taxa and changes in their distribution over time in response to environmental factors, with specific regard to climatic, tectonic and eustatic events. A clear understanding of the chronology and geographic extent of these events will 1) serve to clarify how these environmental factors influenced faunal distribution and provinciality, and 2) allow for the assessment of the spatiotemporal occurrence and nature of dispersal pathways among faunal bioprovinces and directionality of movement. A further goal is to describe the taxic composition of each province during the middle and late Miocene to define patterns of in situ evolution versus dispersal into a region, which will be used as a framework to clarify hominoid occurrences in the following chapter.

Barriers to dispersal & paleoenvironmental events in the Miocene of Eurasia

Of the environmental factors influencing the distribution of Eurasian Miocene land mammals, climate, tectonics, and eustasy with subsequent epicontinental marine fluctuations are considered here. According to Hallam (1974) and Middlemiss & Rawson (1969), the influence of climatic barriers is variable in different taxa, dependent on a species’ tolerance to temperature and humidity change and the resulting shifts in vegetation structure.

Geographic barriers on the other hand, including continental divergence, the uplift of mountain chains and occurrence of large bodies of water, are a more effective and “potent” means of restricting faunal distribution and overall movement between areas (Hallam 1974,

47 p214; Middlemiss & Rawson 1969). However, Fortelius et al. (2003a) more recently suggested that due to the “low” temporal resolution of the formation and cessation of geophysical barriers, in addition to the ability of land mammals to cross seemingly impenetrable barriers, there arises a difficulty in identifying direct causal relationships between dispersal events and specific non-climatic environmental changes. According to

Middlemiss & Rawson (1969), however, evidence of environmental influences, with specific regard to climatic variables, is often not preserved in the fossil record and therefore goes undetected. This study, therefore, includes both climatic and geophysical events.

Climate

Few studies have quantitatively assessed the relationship between Eurasian primate distribution and climate during the Miocene. However, with the advent of GIS in , researchers are now assessing how specific aspects of climate influenced primate distribution. For example, Eronen & Rook (2004) used GIS to quantitatively evaluate the spatiotemporal relationship between the distribution of Eurasian primates and fluctuations in paleoprecipitation during the Neogene, using large crown height as a proxy for humidity. Crown height relates to increases in plant fibrousness and abrasiveness, due to the silica, extraneous grit and decreased nutrient value associated with generalized water stress (Fortelius 1985; Fortelius et al. 2003a; Janis 1988). Eronen & Rook

(2004) plotted humidity values over time on a series of maps, concluding that both the

European hominoids and pliopithecoids were restricted to areas of peak humidity from MN5-

MN9, but then with the subsequent increasing aridity, the hominoids moved southward, potentially tracking the humidity from MN10 to MN12.

48

No study to date has quantitatively associated the movement patterns of Eurasian terrestrial mammals to temperature change, despite many references to qualitative changes

(i.e., warm-tropical, cool-temperate). However, several recent studies have used palynological and paleobotanical data to obtain proxies for paleotemperature at a limited number of localities. For example, Bruch et al. (2007, 2006, 2004) used the co-existence approach to estimate mean annual temperature, mean temperature in the coldest month and mean temperature in the warmest month for a series of early middle Miocene localities and early late Miocene localities in Europe. These authors used an inverse distance weighted

(IDW) spatial interpolator in ArcVIEW GIS to produce maps of spatial and latitudinal gradients in temperature at these localities. Although a potentially rich source of information, these proxy data require validation by another data source, particularly since the distribution maps put forth by these authors are based on very few data points.

In sum, these data sources can be utilized to assess the extent of the influence of certain aspects of climate on the distribution of Eurasian terrestrial mammals in the middle and late Miocene, however they must be regarded as tentative estimates at best.

Tectonics

The continental collision and orogeny occurring in Eurasia prior to and during the

Miocene had profound effects on the biogeographic distribution of terrestrial mammals in

Africa and Europe. While events prior to the Miocene have been well addressed in the literature, the Miocene itself has received very little attention, at least on the continental scale. Prior to the early Miocene, the Tethys seaway, which joined the Atlantic to the Indo-

Pacific Ocean via the Mediterranean, prevented any contact between the African and

European continental land masses. However, the anticlockwise rotation and northern drift of

49 the Afro-Arabian tectonic plate resulted in collision with Eurasia in the early Miocene (~20-

19Ma), establishing the first route for faunal interchange, known as the “Gomphotherium land bridge,” across Arabia and Asia Minor. This led to the first dispersal of African land mammals, including primates, into Europe in MN4 and MN5 (Andrews & Bernor 1999;

Begun et al. 2003a & b; Bernor 1983, 1978; Harrison & Gu 1999; Maridet et al. 2007; Rögl

1999a-c; Rögl & Steininger 1984; Steininger et al. 1985; Thomas 1985; Thomas et al. 1982).

According to Maridet et al. (2007), this tectonic event, which effectively bridged the marine barrier of the Tethys seaway, was coincident with the emergence of clear biogeographic patterns in mammalian distribution and distinct faunal bioprovinces in Europe.

Within Europe, the uplift of mountain chains that could have constituted formidable physical barriers influencing the distribution of mammals included the , ,

Carpathians and Balkan Mountains (Figure 3.1). Although many of these topographic features were formed prior to the Neogene, segments of individual mountain chains still experienced orogenic activity during the Miocene and into the Pliocene. The occurrence and uplift of these mountains in the study area prior to and during the Miocene had varying effects on climate patterns (Bernor 1978; Ruddiman & Prell 1997). They are evaluated here only as potential biogeographic barriers.

Initiating in the Mesozoic due to a collision of the European and Iberian tectonic plates, the Pyrenees experienced maximum uplift in the mid to late (Boillot &

Capdevila 1977; Dèzes et al. 2004; Giese et al. 1982; Maridet et al. 2007; Rosenbaum &

Lister 2002; Schellart 2002; Sibuet et al. 2004; Stampfli et al. 2002; Vergé et al. 2002), erecting a potential physical barrier between the Iberian Peninsula and the rest of Europe

(Maridet et al. 2007). While individual segments of the Alpine chain experienced different

50 tectonic histories, the Alps in general began uplifting in the late and Paleocene due to collision of the European and African tectonic plates in addition to other localized factors, and continued their orogeny during the Miocene, with certain regions continuing to uplift into the Pliocene (Dèzes et al. 2004; Giese et al. 1982; Peresson & Decker 1997;

Rantitsch 1997; Schlunegger & Simpson 2002; Stampfli et al. 2002; Ziegler 1990, 1988).

The Jura Mountains in northwestern Switzerland and eastern France arise north of the western Alpine arc and merge with the Alps in the south. Although these mountains began their uplift in the Oligocene, they became incorporated into the Alpine chain in the Mio-

Pliocene, ~11-3Ma (Stampfli et al. 2002) or ~9-4Ma based on stratigraphic and paleontologic evidence (Becker 2002). Together, the Alps and Jura Mountains potentially served as a north-south dividing barrier across western and central Europe. The uplift of the Carpathians also occurred as a result of Mesozoic to Cenozoic continental collision between Europe and

Africa. These mountains, together with the Eastern Alps and Dinarides almost completely encircle the Pannonian Basin in Central and Eastern Europe. The overall trend of mountain building in the Carpathians initiated from the northwest to southeast during the Neogene

(Csontos 1992; Royden 1988), with the uplift of the Eastern Carpathians occurring during the late middle to early late Miocene (~14 - 12 Ma) and the uplifting of the Southern Carpathians in the latest Miocene and early Pliocene (Golonka 2004; Popov et al. 2004). To the southeast, the Carpathians grade into the Balkan chain, which initiated during the Eocene

(Golonka 2004; Huismans et al. 1997).

Although the following mountain chains would not have served to obstruct pan-

European movement of terrestrial mammals during the Miocene, they could have affected dispersal between Asia, Africa and Europe. Originating in the Mesozoic, the Betic Cordillera

51 in Spain was active into the middle Miocene (Giese et al. 1982; Sánchez-Gómez et al. 2002;

Steininger et al. 1985; Zeck 1996) and together with the North African Maghrebides, served to restrict the Atlantic-Mediterranean gateway. According to Golonka (2004), the continued thrusting of the Moroccan Rifs and Spanish Betics in the late Miocene temporarily joined

Africa and Europe across the Gibraltar Strait, and isolated the Mediterranean from the

Atlantic Ocean. Towards the eastern-most boundary of the study region, the Caucasus

Mountains of Georgia are also significant in posing a potential physical barrier from Eastern

Europe into Asia proper. These mountains uplifted beginning in the Oligocene/early

Miocene and again in the late Miocene (Golonka 2004; Kopp 2007).

Eustasy & Epicontinental Marine Fluctuations

As mentioned previously, the Tethys seaway posed an insurmountable marine barrier to intercontinental terrestrial mammal interchange between Africa and Europe, as well as between Eastern Europe and Southwest Asia for much of the Miocene. However, several opportunities arose in the Miocene for faunal interchange to occur due to the movement of the African tectonic plate. Occurring first in the early and early middle Miocene, land corridors allowed intermittent waves of dispersal between Africa and Eurasia (Bernor 1978;

Esteban 1996; Hsü et al. 1977; Rögl 1999a; Rögl & Steininger 1984; Steininger et al. 1985;

Thomas 1985; Thomas et al. 1982). However, the middle Miocene marine transgression (the

Langhian transgression) reconnected the Indo-Pacific to the Atlantic Ocean through the

Mediterranean gateway (Rögl & Steininger 1984; Steininger et al. 1985). According to Rögl

(1999b), this transgression was short-lived, with intercontinental land bridges re-appearing in

MN6, upper MN7/8 (around the Eastern Paratethys only) and at the base of MN9 with the spread of Hipparion. In the late Miocene (~10.5Ma), the Tethys was once again disrupted in

52 the area of the eastern Mediterranean/Indo-Pacific due to the uplift of the Arabian tectonic plate, after which point, the Mediterranean became an embayment of the Atlantic Ocean. At this time, uplift of the Zagros mountain chain together with geokinematic activity around the

Turkish tectonic plate reestablished the land corridor between Eurasia and Africa. At approximately 6Ma, the Tethys was yet again interrupted due to worldwide sea level regression, as well as by the uplifting Betic Cordillera in Spain and the North African

Maghrebides. This led to the Messinian Salinity Crisis at 5.96 or 5.75 to 5.32Ma, where some consider that parts of, or the entire Mediterranean dried up with water levels dropping perhaps as much as 1500m, thus allowing opportunities for mammal dispersal between the

Europe and North Africa at this time (Costeur et al. 2007; Clauzon et al. 1996; Hsü et al.

1977; Rögl & Steininger 1984; Steininger et al. 1985). Esteban (1996) identifies land bridges across the Betic and Rif Straits, the Sicily Straits, the Suez Straits, and the Balearics to central Mediterranean, which were separated from each other by evaporitic seas.

Therefore, physical bridging of the Tethys and drying of the Mediterranean provided at least three separate opportunities for intercontinental exchange of terrestrial mammals first in the early and early middle Miocene, again at ~10.5Ma and finally by ~5.75Ma.

The biogeographic distribution of middle and late Miocene European terrestrial mammals was also affected by fluctuations in the Paratethys Sea, which represented the northern epicontinental extension of the Tethys and stretched from the Western Alps to the

Central Asian Aral Sea (Figure 3.1) (Bernor 1983; Rögl & Steininger 1984). The uplift of the Alpine chain served to isolate the Paratethys and by ~14Ma, a system of independent basins and land-locked lake systems had developed (Harzhauser et al. 2007; Hsü et al. 1977;

Rögl & Steininger 1984). The paleogeography and lithology of the Paratethys realm is

53 reviewed extensively and figured in Rögl (1999b), Rögl & Steininger (1984) and Popov et al.

(2004), while Magyar et al. (1999) focus on the Central Paratethys and Central European region of the Pannonian Basin. Together, these paleogeographic maps depicting the extent and fluctuation of the Tethys and Paratethys clarify the relationship between the distribution of terrestrial biotas and the spatial and temporal potential for inter-provincial movement within Eurasia.

Figure 3.1: Physiographic features in study area

C C A C D P B Cau

G

P Pyrenees B Balkan Mountains A Alps Cau Caucasus Mountains D Dinarides G Gibraltar Strait C Carpathians Paratethys

Previous method and study of provinciality

Three previous studies have assessed provinciality within Eurasian Miocene land mammals. Bernor (1983 & 1978) analysed 38 Eurasian and African large mammal localities, using Simpson’s index of faunal similarity (Simpson 1960), together with hierarchical cluster

54 analysis at both the genus and species level. Bernor (1983 & 1978) found that while there was a temporal effect in the grouping of his locality sample, geographic proximity exerted a significant influence on locality clustering. From these studies, Bernor (1978) initially proposed four zoogeographic provinces, then later revised his initial findings for a total of six distinct zoogeographic provinces (Bernor 1983). These include: 1) a Western and Southern

European Province, 2) an Eastern and Central European Province, 3) a Romanian and

Western Russian Province, 4) a Sub-Paratethyan Province (Turkey, Greece and Iran), 5) a

North African province and lastly, 6) a Siwalik province. Bernor (1983 & 1978) discussed his findings in relation to hominoid and pliopithecoid occurrences and indicated potential routes for migration between adjacent provinces and between continental land masses.

Fortelius et al. (1996a) also conducted an analysis of species richness, provinciality, turnover and paleoecology in Eurasian Miocene land mammals. Initially dividing Western

Eurasia into two major faunal blocks, “East” and “West,” and further into six pre-defined geographical regions: Western Europe (Portugal, Spain, France and Italy), west Central

Europe (Germany and Switzerland), Austria, the Black Sea region (Hungary, Romania,

Moldova and the Ukraine), the Balkans (Slovenia, Croatia, Bosnia, Serbia, Macedonia,

Albania, Greece, Turkish Thrace and Bulgaria) and Anatolia (Anatolia, Samos and Georgia).

Fortelius et al. (1996a) analysed 490 large mammal and 120 small mammal localities, using the Dice (Sokal & Sneath 1963), Pielou (Pielou 1979) and Simpson indices (Simpson 1943) to measure faunal similarity. Like Bernor (1983 & 1978), these authors also found a correlation between geographic proximity and degree of similarity. Fortelius et al. (1996a) tracked changes in similarity between regions over time and related this to the evolution and migration of specific mammalian groups.

55

Methods of analysis of provinciality, including those mentioned previously, generally utilize one or more similarity indices, which are based on the pair-wise comparisons of presence and absence of genera and/or species (i.e., Johnson 1971, Raup & Crick 1979;

Simpson 1960, etc.). Each having its own advantages and disadvantages, the large number of available indices reflects an effort to eradicate the problems inherent in these methods, which can lead to erroneous results. These problems can be divided into two main areas: those pertaining to sensitivity to differences in sample size and those pertaining the equal weighting of presence vs. absence (Hallam 1974). With regard to the former, researchers often choose to conduct their analysis using a given similarity index, while limiting their sample to localities with equivalent numbers of fauna. While a valid solution to the problem of unequal samples, this alternative usually results in smaller overall samples available for comparison.

Hierarchical cluster analyses using various grouping algorithms and non-metric multidimensional scaling (MDS) are also commonly used as exploratory methods for studying provinciality, either using the resulting indices of a similarity analysis, or the raw presence/absence data itself (Hughes 1973; Manly 2005; Quinn & Keough 2002). According to Quinn & Keough (2002), the resulting dendrogram of a cluster analysis does not represent all pair-wise comparisons in a given sample due to either the nature of the clustering method or the use of poor data. In the latter, different grouping algorithms will produce different topologies when data quality is suspect. As such, cluster analysis should be used in biology to display phylogenetic relationships rather than to analyze species abundance data to determine relationships between sampling units (Quinn & Keough 2002). However, other researchers recognize the advantages and utility of cluster analysis for certain tasks, such as

56 the identifications of internal groupings within a dataset, data reduction, etc. (i.e., Fortin &

Dale 2005; Legendre & Legendre 1998; Manly 2005). Of particular relevance to this study,

Legendre & Legendre (1998) and Legendre (1973) note that cluster analysis is effective for identifying what they term as ‘associated species,’ or those that are reacting in similar ways to environment flux. It would seem that a good awareness of the data and their limitations, together with validation through the use of several grouping algorithms to ensure topology replication, would justify this method for studies of provinciality. As a further caveat, the results of a cluster analysis should only be considered when the co-phenetic correlation, or measure of the goodness of fit of the resulting dendrogram to the original data, are above .8

(a value of 1.0 being a perfect correspondence between the resulting dendrogram and the original data) (Legendre & Legendre 1998). Furthermore, the reliability of a cluster analysis can be determined by including data with known structure to observe whether the linkage algorithm used can produce congruous results (Manly 2005). MDS can be used in conjunction with, or as an alternative to cluster analysis, and is designed specifically to provide visual representation of relationships between data points in multidimensional space

(Hughes 1973; Manly 2005; Quinn & Keough 2002). Unlike cluster analysis, MDS does consider all pair-wise dissimilarities among data points (Quinn & Keough 2002).

The current study differs from those previously mentioned in several significant ways. In contrast to both Bernor (1983 & 1978) and Fortelius et al. (1996a), this analysis was conducted using raw presence/absence data, in addition to similarity indices. When using a similarity index, taxonomic identity is lost, due to the numerical representation of taxa. Therefore, analysis conducted on the raw presence/absence data facilitates the recognition of key taxa influencing the formation, composition and maintenance of

57 bioprovinces. Over successive time slices, these same taxa can be tracked to better understand if, when and where they disperse. In contrast to Fortelius et al. (1996a), who divided their study area into blocks and regions a priori and subsequently assessed similarity, this study uses the fauna themselves to define faunal provinces. This allows for the recognition of bioprovinces in relation to landforms and regional topography. In addition, no study to date has specifically accounted for large scale faunal distribution in relation to the combined effect of specific climatic, tectonic and eustatic factors. This study uses the most recently updated and extensive sample of fossil mammals and integrates new data not previously used. Lastly, the innovative use of GIS to graphically represent environmental effects together with faunal distribution facilitates the recognition and definition of bioprovinces, their barriers and fluctuations therein over time.

Materials & Methods

Materials

Taxon occurrence data from the previous chapter were used here to assess provinciality. Localities were temporally ordered according to their MN designation into a data matrix and their component taxa were scored as present or absent. Localities falling within a single MN unit were considered here, although a number of localities correlated with land mammal ages (i.e., Vallesian, Turolian, etc.) were also included if these localities had geographic, temporal or taxonomic relevance. Furthermore, localities with more than 10 species were selected for analysis as to minimize taxonomic “noise,” as per Alroy (2004).

However, localities located in poorly represented regions, as well as species-poor primate localities were also considered. The sample of localities extends across the same geographic expanse as in the previous chapter, but also includes late Miocene localities in the Middle

58

East and a smaller series of middle and late Miocene African localities to demonstrate the timing and taxonomic composition of intercontinental faunal exchange (Figure 3.2, Table

3.1). Temporally, the sample of localities spans from 17Ma (base of MN5) to 5.3Ma (base of

MN 14) and includes small mammals (rodents, insectivores, lagomorphs) from 270 localities and large mammals from 328 localities.

Analysis

Hierarchical cluster analysis was used to explore and determine groupings at the genus and species level for large and small mammals within individual MN units. I used single linkage, complete linkage, unweighted pair-group average and Ward’s minimum variance to determine which linkage method would result in the highest co-phenetic correlation, as well to observe whether topologies would change when using the different algorithms. Single linkage determines the distance between clusters using nearest neighbours, or the minimum distance between objects in different clusters, while complete linkage performs the opposite way; distances between clusters are determined using the greatest distance between any two objects in the different clusters. The unweighted pair- group average method assesses the mean distance between all pairs of objects in different clusters. Ward’s minimum variance evaluates distance between clusters using an analysis of variance to minimize the sum of squares of each cluster (Hammer et al. 2007; Legendre &

Legendre 1998). I also conducted identical cluster analyses using the Jaccard, Dice and

Raup-Crick similarity indices, again to observe any differences in clustering and dendrogram topology. The Jaccard and Dice index both use pair-wise presence/absence data to produce a coefficient of faunal similarity, however, the latter weights observed joint occurrences more heavily than absences. The Raup-Crick index uses Monte Carlo randomization to compare

59 the observed number of co-occurrences in a pair of localities with the distribution of co- occurrences in 200 random iterations (Hammer et al. 2007). Following the similarity analyses, I verified the clustering patterns with the raw data to ensure that the groups formed were supported by the data. Not wanting to force the data into a specified number of groups a priori, I chose to avoid any partition methods of clustering (i.e., K-means, non-hierarchical linkage). The spatial distribution of fauna and the geographic extent of mountain chains, the

Paratethys and the Tethys, were plotted as individual layers on a series of base maps, using

ArcGIS 9.2.

Table 3.1: Localities included in analysis

MN 4-MN5 (17-15.2Ma)

Locality Country Age Ad Dabtiyah MN3-5 Rusinga (-) Kenya MN4 Rusinga (Gumba) Kenya MN4 Rusinga (Hiwegi west) Kenya MN4 Rusinga (Hiwegi) Kenya MN4 Rusinga (Kathwanga) Kenya MN4 Rusinga (Kiahera Hill) Kenya MN4 Rusinga (Kiyune) Kenya MN4 Rusinga (Nyamsingula) Kenya MN4 Rusinga (R 113) Kenya MN4 Rusinga (R 114) Kenya MN4 Rusinga (R105) Kenya MN4 Rusinga (R106) Kenya MN4 Rusinga (R107) Kenya MN4 Rusinga (R3) Kenya MN4 Rusinga (Wayondo) Kenya MN4 Kalodirr Kenya Napak Uganda Burdigalian Napak (Iriri member) Uganda Burdigalian Napak Member Uganda Burdigalian Napak Rhino Site Uganda Burdigalian Göriach Austria MN5 Grund Austria MN5 Muhlbach am Manhartsberg Austria MN5 Obergänserndorf 1 & 2 Austria MN5 Teiritzberg 1 (T1) Austria MN5 Franzensbad Czech Republic MN5

60

Baigneaux-en Beauce France MN5 Esvres - Marine Faluns France MN5 Faluns of Touraine & Anjou France MN5 La Condoue France MN5 Manthelan France MN5 Pontlevoy France MN5 Poudenas-Peyrecrechen France MN5 Rimbez - Lapeyrie base France MN5 Savigné-sur-Lathan France MN5 Contres MN 5 France MN5 Belometchetskaja Georgia MN5 Affalterbach Germany MN5 Engelswies Germany MN5 Gisseltshausen Germany MN5 Griesbeckerzell Germany MN5 Häder Germany MN5 Hambach 6C Germany MN5 Heggbach Germany MN5 Laimering 3 GE 5 Germany MN5 Massendorf Germany MN5 Puttenhausen Germany MN5 Rothenstein 1 Germany MN5 Sandelzhausen Germany MN5 Schellenfeld 2-4 Germany MN5 Walda 2 Germany MN5 Wannenwaldtobel-2 Germany MN5 Ziemetshausen 1b Germany MN5 Antonios (ANT) Greece MN5 Chios Greece MN5 Belchatow B Poland MN5 La Hidroelectrica Madrid Spain MN5 Montejo de la Vega Spain MN5 Puente de Vallecas Spain MN5 Sant Mamet Spain MN5 Somosaguas-Sur Spain MN5 Chatzloch Switzerland MN5 Frohberg Switzerland MN5 Hüllistein Switzerland MN5 Martinsbrünneli Switzerland MN5 Tobel-Hombrechtikon Switzerland MN5 Vermes 1 Switzerland MN5 Vermes 2 Switzerland MN5 Ardiç-Mordoğan Turkey MN5 Çandır (Loc. 3) Turkey MN5 Çandır 2 Turkey MN5 Karaağaç 1 Turkey MN5 Koçgazi Turkey MN5 Paşalar Turkey MN5 Arrisdrift MN5 Al-Sarrar Saudi Arabia MN5

61

Sinda Congo Congo/Zaire Kipsaramon 1 Kenya Langhian Kipsaramon 2 Kenya Langhian Maboko Kenya Langhian Majiwa Kenya Langhian Ombo Kenya Langhian

MN6 (15.2-12.5)

Locality Country Age Sinda Congo Congo/Zaire Langhian Kipsaramon 1 Kenya Langhian Kipsaramon 2 Kenya Langhian Maboko Kenya Langhian Majiwa Kenya Langhian Ombo Kenya Langhian Klein Hadersdorf Austria MN6 Trimmelkam Austria MN6 Castelnau-d'Arbieu France MN6 Four France MN6 Four (general) France MN6 Liet France MN6 Sansan France MN6 Simorre France MN6 Diessen am Ammersee Germany MN6 Stätzling Germany MN6 Steinberg Germany MN6 Thannhausen Germany MN6 Unterneul Germany MN6 Matraszolos1-2 Hungary MN6 Samsonhaza Hungary MN6 Subpiatrã 2/1R Romania MN6 Devínská Nová Ves - Bonanza Slovakia MN6 Devínská Nová Ves - Fissures Slovakia MN6 Devínská Nová Ves - Sandhill Slovakia MN6 Armantes 7 Spain MN6 Arroyo del Val Spain MN6 Elgg Switzerland MN6 Kreutzlingen Switzerland MN6 Rümikon Switzerland MN6 Schwamendingen Switzerland MN6 Stein am Rhein Switzerland MN6 Wiesholz Switzerland MN6 Zeglingen Switzerland MN6 Bagici Turkey MN6 Catakbagyaka Turkey MN6 Inönü I (Sinap 24A) Turkey MN6 Al Jadidah Saudi Arabia MN6 Sevastopol (Sebastopol) Ukraine Sarmatian

62

Fort Ternan Kenya Fort Ternan 2 (Serek) Kenya Serravallian Ngorora Kenya Serravallian Nyakach 10 (Kaimogool North) Kenya Serravallian Nyakach 10 (Kaimogool North) Kenya Serravallian Nyakach 11 (Kaimogool South) Kenya Serravallian Nyakach 11(Kaimogool South) Kenya Serravallian Nyakach 8 (Kadianga West) Kenya Serravallian Nyakach 8 (Kadianga West) Kenya Serravallian Ngorora Kenya Serravallian-

MN7/8 (12.5-11.2Ma)

Locality Country Age St. Stephan im Lavanttal Austria MN7/8 La Grive St. Alban France MN7/8 Poudenas-Cayron France MN7/8 St. Gaudens France MN7/8 Massenhausen Germany MN7/8 Steinheim Germany MN7/8 Felsotárkány 1 Hungary MN7/8 Felsotárkány 3/2 Hungary MN7/8 Felsotárkány 3/2 (Güdör-kert) Hungary MN7/8 Felsotárkány 3/8 Hungary MN7/8 Felsotárkány-Felnémet Hungary MN7/8 Opole 2 Poland MN7/8 Przeworno 2 Poland MN7/8 Comanesti-1 Romania MN7/8 Subpiatrã 2/2 Romania MN7/8 Can Feliu Spain MN7/8 Can Mata 1 Spain MN7/8 Castell de Barberà Spain MN7/8 Escobosa Spain MN7/8 Hostalets de Pierola Inferior Spain MN7/8 Nombrevilla-2 Spain MN7/8 Sant Quirze Spain MN7/8 Anwil Switzerland MN7/8 Bois de Raube 3 Switzerland MN7/8 Bayraktepe 1 Turkey MN7/8 Pismanköy (Yenidibekkoyu) Turkey MN7/8 Sariçay Turkey MN7/8 Sofca Turkey MN7/8 Yeni Eskihisar 1 Turkey MN7/8 Yenieskihisar Turkey MN7/8 Bled Douarah 12.5-9.5Ma Sevastopol (Sebastopol) Ukraine Sarmatian

63

Fort Ternan Kenya Serravallian Fort Ternan 2 (Serek) Kenya Serravallian Ngorora Kenya Serravallian Nyakach 10 (Kaimogool North) Kenya Serravallian Nyakach 10 (Kaimogool North) Kenya Serravallian Nyakach 11 (Kaimogool South) Kenya Serravallian Nyakach 11(Kaimogool South) Kenya Serravallian Nyakach 8 (Kadianga West) Kenya Serravallian Nyakach 8 (Kadianga West) Kenya Serravallian Ngorora Kenya Serravallian-Tortonian

MN9 (11.2-9.5Ma)

Locality Country Age Götzendorf Austria MN9 Mariathal Austria MN9 Vösendorf Austria MN9 Suchomasty Czech Republic MN9 Doué-la-Fontaine France MN9 Jujurieux France MN9 Priay II France MN9 Udabno I Georgia MN9 Eppelsheim Germany MN9 Esselborn Germany MN9 Hammerschmiede Germany MN9 Höwenegg Germany MN9 Melchingen Germany MN9 Wartenberg Germany MN9 Wissberg Germany MN9 Rudabánya Hungary MN9 Sümeg Hungary MN9 Atavaska MN9 Buzhor 1 Moldova MN9 Kalfa Moldova MN9 Varnitsa Moldova MN9 Belchatow A Poland MN9 Ballestar Spain MN9 Can Llobateres I Spain MN9 Can Ponsic Spain MN9 Can Ponsic I Spain MN9 El Firal Spain MN9 Hostalets de Pierola Superior Spain MN9 Los Valles de Fuentidueña Spain MN9 Nombrevilla Spain MN9 Pedregueras 2A Spain MN9 Pedregueras 2C Spain MN9 Peralejos 5 Spain MN9 Sant Miquel de Taudell Spain MN9 Santiga Spain MN9

64

Seu d'Urgel Spain MN9 Subsol de Sabadell Spain MN9 Charmoille Switzerland MN9 Alkcaköy (1-6) Turkey MN9 Bayraktepe 2 Turkey MN9 Esme Akçaköy Turkey MN9 Middle Sinap Turkey MN9 Sinap 108 Turkey C5N.2N Sinap 4 Turkey C5N.2N Sinap 64 Turkey C5N.2N Sinap 65 Turkey C5N.2N Sinap 91 Turkey C5N.2N Sinap 94 Turkey C5N.2N Udabno II Georgia MN9-MN10 Eldari I Georgia 10.1-9Ma Chorora Fm. Ethiopia 11-10Ma Awash 1 Ethiopia MN9-MN10 Bled Douarah Tunisia 12.5-9.5Ma Bouhanifia 5 10.8-9.8Ma Oued el Atteuch Algeria 10-9.7Ma Bou Hanifia Algeria MN9 Amama 1 Algeria 10Ma Sig 1 Algeria MN9-MN10 Oued Zra Morocco 10.7-10.5Ma Nakali Kenya MN9-MN10

MN10 (9.5-9Ma)

Locality Country Age Eichkogel-upper Austria MN10 Kohfidisch Austria MN10 Ambérieu 2A France MN10 Ambérieu 2C France MN10 Douvre France MN10 Lo Fournas 1993 France MN10 Montredon France MN10 Soblay France MN10 Salmendingen Germany MN10 Nikiti 1 (NKT) Greece MN10 Nikiti 2 (NIK) Greece MN10 Pentalophos 1 (PNT) Greece MN10 Pyrgos Vassilissis Greece MN10 Ravin de la Pluie (RPL) Greece MN10 Xirochori1(XIR) Greece MN10 Poksheshty Moldova MN10 Can Purull Spain MN10 La Cantera Spain MN10 La Roma 2 Spain MN10

65

La Tarumba I Spain MN10 Los Aguanaces 5A & B Spain MN10 Masia de la Roma 11 Spain MN10 Masia de la Roma 3-9 Spain MN10 Masía del Barbo Spain MN10 Masía del Barbo 2 Spain MN10 Masía del Barbo2B Spain MN10 Puente Minero 2 & 8 Spain MN10 Terrassa Spain MN10 Gülpinar Turkey MN10 Yulafli (CY) Turkey MN10 Sinap 49 Turkey C4AR.1N Sinap 12 Turkey C4AR.2N Sinap 84 Turkey C4AR.2R Eldari I Georgia 10.1-9Ma Udabno II Georgia MN9-MN10 Awash 1 Ethiopia MN9-MN10 Nakali Kenya MN9-MN10 Sig 1 Algeria MN9-MN10 Feid El Atteuch R1 Algeria MN10 Tafna Algeria MN10

MN11 (9-8.2Ma)

Locality Country Age Ambérieu 1 France MN11 Ambérieu 3 France MN11 Bernardière France MN11 Dionay France MN11 Lobrieu France MN11 Mollon France MN11 DornDürkheim1 Germany MN11 Halmyropotamos (HAL) Greece MN11 Halmyropotamos (HAL)+A90 Greece MN11 Pikermi Greece MN11 Prochoma Greece MN11 Ravin des Zouaves 5 Greece MN11 Samos Greece MN11 Samos (A-1) Greece MN11 Vathylakkos 3 (VAT) Greece MN11 Csakvar Hungary MN11 Baccinello V1 Italy MN11 MonteBamboli Italy MN11 Alfambra Spain MN11 Crevillente 2 Spain MN11 La Gloria 10 Spain MN11 Los Aguanaces Spain MN11 Los Aguanaces 3 Spain MN11 Masada Ruea 2 Spain MN11

66

Peralejos D Spain MN11 Piera Spain MN11 Puente Minero Spain MN11 Puente Minero 3 Spain MN11 Vivero de Pinos Spain MN11 Çorakyerler Turkey MN11 Garkin Turkey MN11 Karacahasan Turkey MN11 Kayadibi Turkey MN11 Kemiklitepe 1,2 Turkey MN11 Kemiklitepe D Turkey MN11 Küçükçekmece Turkey MN11 Mahmutgazi Turkey MN11 BalaYaylaköy Turkey MN11 Grebeniki Ukraine MN11 Novo-Elizavetovka Ukraine MN11 Kalimanci 3&4 Bulgaria MN11-12 Maragheh Iran MN11 Injana Iraq MN11 Samburu Hills 2 Kenya Turolian Afoud 6 Morocco 8.7Ma

MN12 (8.2-7.1Ma)

Locality Country Age Hadjidimovo-1 Bulgaria MN12 Kalimanci 2 Bulgaria MN12 Kalimanci 3&4 Bulgaria MN11-12 Mt. Luberon France MN12 Ano Metochi 3 Greece MN12 Chomateres Greece MN12 Pikermi-MNHN (PIK) Greece MN12 Vathylakkos 2 (VTK) Greece MN12 Baltavar Hungary MN12 Polgardi Hungary MN12 Middle Maragheh Iran MN12 UpperMaragheh Iran MN12 Baccinello V2 Italy MN12 Lothagam 1 Kenya MN12 Titov Veles Macedonia MN12-13 Chimishlija (Cimislia) Moldova MN12 Chobruchi (Tchobroutchi) Moldova MN12 Gura-Galben Moldova MN12 Taraklia Moldova MN12 Tudorovo Moldova MN12 Aljezar B Spain MN12 Casadel Acero Spain MN12 Cerrodela Garita Spain MN12 ConcudB Spain MN12 Concudbarranco Spain MN12

67

Crevillente 15 Spain MN12 Crevillente 8 Spain MN12 Los Aljezares Spain MN12 Los Mansuetos Spain MN12 Masada Ruea 3 & 4 Spain MN12 Masadadel Valle 2 Spain MN12 Masadadel Valle 5 Spain MN12 Regajo 3 & 4 Spain MN12 Tortajada A Spain MN12 Tortajada C & D Spain MN12 Valdecebro 5 Spain MN12 Villalba Baja 2B/2C Spain MN12 Belka Ukraine MN12 Novaja Emetovka Ukraine MN12 Akgedik-Bayir Turkey MN12 Akkasdagi Turkey MN12 Çobanpinar(Sinap42) Turkey MN12 Duzyayla Turkey MN12 Gelibolu Bayirköy Turkey MN12 Kemiklitepe A-B Turkey MN12 Salihpasalar Turkey MN12 Salihpasalar 1-3,6 Turkey MN12 Sandikli Kinik Turkey MN12 Serefköy Turkey MN12 Sinap 26 Turkey C4R.1R Sinap 33 Turkey C4R.1R Sinap 72 Turkey C5N.1R Sinap 8A Turkey C5N.1R Sinap 8B Turkey C5N.1R Molayan Afghanistan MN12 Shuwaihat Saudi Arabia MN12 Hamra Saudi Arabia MN12-13 Jebel Dhannah Saudi Arabia MN12-13 Samburu Hills 2 Kenya Turolian Khendekel Ouaich Morocco 7.4Ma Khendekel Ouaich Morocco 7.4Ma

MN13 (7.1-5.3Ma)

Locality Country Age Lissieu France MN13 Dytiko 1-3 (DTK) Greece MN13 Maramena Greece MN13 Monasteri Greece MN13 Samos Main Bone Beds Greece MN13 Silata Greece MN13 Hatvan Hungary MN13 Baccinello V3 Italy MN13

68

Brisighella Italy MN13 Casino Italy MN13 Gargano Italy MN13 Gravitelli Italy MN13 Velona Italy MN13 Titov Veles Macedonia MN12-13 Almenara-Casablanca M Spain MN13 Arenasdel Rey Spain MN13 Arquillo 1 Spain MN13 Bacochas-1 Spain MN13 Celadas 2 Spain MN13 El Arquillo 1 Spain MN13 La Alberca Spain MN13 La Gloria 5 Spain MN13 La Gloria 6 Spain MN13 Las Casiones Spain MN13 Las Casiones superior Spain MN13 Layna Spain MN13 Librilla Spain MN13 Milagros Spain MN13 Rambla de Valdecebro 3 & 6 Spain MN13 Venta del Moro Spain MN13 Villastar Spain MN13 Crevillente 14 Spain Messinian Crevillente 22 Spain Messinian Kangal 1 Turkey MN13 Süleimanli 2 Turkey MN13 Taskinpasa 1 Turkey MN13 Ananjev Ukraine MN13 Amama Algeria MN13 Menacer Algeria 7.1-5.3Ma Toros-Menalla (TM-266) Chad MN13 Albertine 14 Congo/Zaire Messinian Wadi Natrun Egypt MN13 Awash 3 Ethiopia MN14 Awash 5 Ethiopia MN14 Lothagam 2 Kenya MN13 Lukeino 3-6 Kenya Messinian Lothagam 5 Kenya Messinian-Piacenzian Lukeino Kenya Messinian Mpesida Kenya 6-7Ma Sahabi Libya MN14 Afoud 1 Morocco 6.1Ma Afoud 2 Morocco 5.8Ma Afoud 5 Morocco MN13 Afoud 8 Morocco 5.3Ma Lissasfa Morocco MN13 Langebaanweg 2 (PPM) South Africa Messinian Langebaanweg 3 (QSM) South Africa Messinian Manonga 2 Tanzania Zanclian

69

Manonga 1 Tanzania Messinian Inolelo 1 Tanzania Messinian-Zanclian Douaria Tunisia 7.1-5.3Ma Albertine 1 Uganda Messinian Nkondo Uganda 6-5Ma Hamra Saudi Arabia MN12-13 Jebel Dhannah Saudi Arabia MN12-13

Figure 3.2: Distribution of localities

Results

Algorithms

On both the raw data and the similarity indices, the unweighted group average algorithm consistently produced the highest co-phenetic correlation, however the dendrogram topologies did not change considerably using other methods of linkage. Based on this finding, the results of the unweighted group average are discussed here.

Raw data vs. similarity indices

Every dendrogram constructed using the raw presence/absence data produced topologies that were, to varying degrees, inconsistent with the raw data. Localities were

70 frequently clustered despite having no common taxa and those with nearly identical faunal lists were placed in different clusters. This is likely the result of the hierarchical and sequential nature of the clustering methodology as a whole being less flexible with large data sets consisting of 1s and 0s. Clustering based on the similarity indices produced dendrograms that were consistent with the raw data. In all cases, clustering based on the

Dice and Jaccard indices resulted in identical topologies. Clustering based on the Raup-

Crick index was nearly identical to the previous measures, differing only in the placement of a very small number of localities (usually one or two), which remained in relation to other localities from the same geographic region. Therefore, validating the replicability of similar topologies using several different measures of faunal similarity (Legendre & Legendre 1998), together with well-supported clusters and the clustering of localities with known structure

(Manly 2005), all suggests that the methodology used here was producing realistic results.

Patterns of locality clustering

MN 5

At the species level, the MN5 large mammals cluster into five clear provinces, composed of Turkish and Georgian, Greek, Central European (German, Austrian, French),

Spanish, and African faunas (Figure 3.3a). The Turkish cluster was supported by a large number of taxa that are endemic to the region, including several species of bovid

(Caprotragoides stehlini, Hypsodontus pronaticornis, Tethytragus koehlerae, Turcocerus gracilis), suid (Bunolistriodon meidamon, Listriodon splendens), palaeochoerid

(Schizochoerus anatoliensis, Taucanamo inonuensis), rhino (Beliajevina grimmi,

Procoelodonta tekkayai), hyaenid (Protictitherium intermedium), mustelid (Anatolictis laevicaninus, Ischyrictis anatolicus), and individual species of cervid (Heteroprox

71 anatoliensis), giraffid (Giraffokeryx anatoliensis), gomphotheriid (Gomphotherium pasalarensis), hominoid (Griphopithecus alpani), and the , Orycteropus seni. The

Turkish and Georgian faunas cluster based on a single shared endemic taxon, the mustelid

Ischyrictis anatolicus. Therefore, although localities from these two countries group together, the degree of actual faunal similarity between the Georgian locality

(Belometchetskaja) and the Turkish cluster (Paşalar, Çandır, Ardiç-Mordoğan) is low. The cluster of Greek localities is supported by four shared endemic taxa, including gomphotheriid

(Choerolophodon chioticus), giraffid (Georgiomeryx georgialasi), tayassuid (peccary)

(Sanitherium schlagintweiti) and viverrid (Lophocyon paraskevaidisi) species. The third grouping includes fauna that support a larger Central European (Germany, France, Austria) cluster, as well as smaller Germany-France and Germany-Austria clusters. Although France has a considerable number of its own endemic taxa, it shares significantly more large mammal species in common with Germany (~22) than with Spain (~3). Large mammals that support this third cluster include several species of tragulid (mouse-/)

(Dorcatherium naui, D. guntianum, D. peneckei, D. vindobonense), cervid (Dicrocerus elegans, Heteroprox larteti), bovid (Eotragus clavatus), palaeomerycid (pecoran )

(Palaeomeryx bojani, P. kaupi), rhino (Lartetotherium sansaniensis, Prosantorhinus germanicus), mustelid (Ischyrictis zibethoides, Martes munki, Trocharion albanense), in addition to individual species of ursid (Plithocyon stehlini), felid (Pseudaelurus romieviensis), amphicyonid (Pseudocyon steinheimensis), moschid ()(Micromeryx flourensianus), palaeochoerid (Taucanamo sansaniense) and pliopithecid ( antiquus). Although there are several species known to Spain that are also found in adjacent regions, these taxa are mostly cosmopolitan. The clustering of Spanish localities (with the

72 exception of Sant Mamet) is supported by a series of endemic species shared by these localities, including cervid (Lagomeryx meyeri), bovid (Eotragus artenensis), suid

(Bunolistriodon adelli), palaeomerycid (Palaeomeryx garsonnini and Triceromeryx pachecoi), rhino (Alicornops simorrensis), mustelid (Martes laevidens), gomphotheriid

(Gomphotherium olisiponensis) and mammutid (Zygolophodon pyrenaicus) species. The grouping of African localities very clearly differentiates them from the Eurasian localities, due to the endemic nature of the African faunas. However, the fauna from the Namibian locality of Arrisdrift, while having similarities to other African localities, also share species with France (Amphicyon giganteus and Hyainailouros sulzeri), as well as Spain, Germany and Turkey (Amphicyon major).

At the genus level, although the provinces identified above are still observable in

MN5, there are fewer genera to support each grouping, since at this taxonomic level, taxa are inevitably more widespread (Figure 3.3b). For example, while at the species level, only a few species were common between Eurasian and African localities, at the genus level, numerous genera are shared between Turkey, Spain, France and Africa. These include the rhino, Aceratherium, and the aardvark, Orycteropus, known from Turkey, Kenya and

Uganda; the gomphotheriid, Platybelodon, and suid, Kubanochoerus, known from Georgia and Kenya; the gomphotheriid, Gomphotherium, known from Germany, France, Spain and

Kenya; the rhino, Brachypotherium, known from Germany, Turkey, Zaire, Uganda and

Kenya; the suid, Bunolistriodon, known from Germany, Spain and Kenya; the chalicotheriid,

Chalicotherium, known from Georgia, Turkey, France, Kenya and Uganda; the hyaenodontid, Hyainailouros, known from France, Kenya, Uganda and Namibia; the rhino

Dicerorhinus, known from Spain, Kenya and Uganda; the sanitheriid, Sanitherium, known

73 from Greece and Kenya; and the hominoid, , known from Turkey (MN5) and

Kenya (MN6). While the broader distribution of genera is still very informative about specific patterns of shared taxa, this taxonomic level seemingly lacks the resolution to define more precise bioprovincial distinctions in the distribution of large mammal taxa.

The MN5 small mammal species cluster into four main provincial groupings, similar to the large mammals (Figure 3.3c). The first Turkish group is supported by several species of murid (rodents including Old World mice, rats and gerbils) (including Byzantinia pasalarensis, Cricetodon candirensis, Democricetodon brevis, Megacricetodon andrewsi,

Pliospalax marmarensis), glirid (dormice) (Glirulus daamsi, Peridyromys lavocati), and pteromyid rodents (flying squirrels) (Albanensia sansaniensis, Forsythia gaudryi), an ochotonid lagomorph (pikas) (Alloptox anatoliensis), an insectivorous erinaceid ()

(Schizogalerix pasalarensis) and a talpid (mole) insectivore (Desmanodon minor). The

Georgian locality of Belometchetskaja does not form part of this cluster, due to the absence of shared species, which may or may not be influenced by poor species richness. The lack of adequate small mammal faunas from Greece during MN5 prevents any provincial recognition. The single locality of Antonios clusters weakly with localities with Austria based on one shared taxon that is cosmopolitan and one taxon shared exclusively by these localities (Cricetodon meini). A second provincial grouping is formed by the small mammal species, which is very similar to the Central European cluster formed by the large mammals.

Within this larger cluster, Germany forms smaller groupings with the Czech Republic and

Poland (based on murids, Anomalomys minor, Neocometes similis; glirids, Glirudinus undosus, Miodyromys hamadryas; and the soricid insectivore, Plesiosorex germanicus),

France and Switzerland (based on murid, Eumyarion bifidus, Megacricetodon germanicus,

74

Miodyromys aegercii; glirid, Bransatoglis cadeoti; and eomyid rodents, Keramidomys carpathicus; as well as a talpid insectivore, Proscapanus sansaniensis), and Austria (based on the glirid rodents, Muscardinus sansaniensis and Prodryomys satus; and the erinaceid insectivore, Lanthanotherium sansaniensis), in addition to numerous endemic species exclusive to the region. The larger grouping is supported by an abundance of widespread rodents, including murids, glirids, pteromyids, and a sciurid (ground squirrel). This grouping is also supported by several insectivore species, including erinaceids, a talpid, soricid () and dimylid (insectivore with no living relatives). With the exception of the ochotonids,

Lagopsis penai and L. verus, from Sant Mamet that are shared with the Swiss and French faunas, a number of endemic small mammal species support a Spanish bioprovince. These species include several murid (Democricetodon darocensis, Fahlbuschia darocensis), glirid

(Armantomys aragonensis, Glirudinus gracilis, Microdyromys monspeliensis, Miodyromys biradiculus) and sciurid rodents (Heteroxerus grivensis, H. rubricati), in addition to an erinaceid insectivore (Mioechinus butleri). The final province recognized in the small mammal species is comprised of faunas from African localities. This province lacks any shared species with Eurasian localities.

At the genus level, there is considerably less resolution in the definition of small mammal faunal provinces due to the preponderance of cosmopolitan genera (Figure 3.3d).

For example, there are many small mammal genera that are endemic to Turkey, including murid, glirid, pteromyid, sciurid, zapodid (jumping mice, birch mice and jerboas), and ctenodactylid rodents. Although these endemics serve to group Paşalar and Çandır, other

Turkish localities are not as tightly clustered (i.e., Koçgazi and Karaağaç 1). A similar condition is evident in the Spanish localities, which are dispersed throughout the dendrogram

75 because of a larger proportion of cosmopolitan taxa and only two endemic genera

(Fahlbuschia and Armantomys). The more central European province, which in other analyses shared common taxa with Germany, France, Austria, Switzerland, the Czech

Republic and Poland, is maintained by the MN5 small mammal genera and is well-supported by several castorid (), pteromyid, glirids, murid and sciurid rodents, as well as soricid, erinaceid and dimylid insectivores. Similarly, the African province is well maintained by a large number of endemic genera, however, the cosmopolitan erinaceid insectivore, Galerix, which is widespread in Europe, is also known from two of the Rusinga Island localities in

Kenya. The erinaceid insectivore, Amphechinus, is common to both localities in Eurasia

(Georgia and Germany), Kenya (several of the Rusinga localities) and Namibia. The ctenodactylid , Sayimys, is also present at localities in Turkey and Saudi Arabia.

76

Figure 3.3a: MN5 large mammal species dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Ad Dabtiyah SA Napak (IR) KE Arrisdrift NA Kipsaramon KE Kipsaramon KE Maboko KE Majiwa KE 8 Ombo KE Kalodirr KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE 16 Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE 24 Rusinga KE Rusinga KE Sinda Congo CO/ZA Napak UG La Condoue FR Chios GR Antonios GR Belometchetskaja RG 32 Çandır 3 TU Paşalar TU Ardiç-Mordoğan TU Walda 2 GE Griesbeckerzell GE Hambach 6C GE Ziemetshausen 1b GE Savigné-sur-Lathan FR 40 Contres MN5 FR Wannenwaldtobel-2 GE Baigneaux-en Beauce FR Pontlevoy FR Esvres - Marine Faluns FR Göriach AU Heggbach GE Rothenstein 1 GE 48 Poudenas FR Rimbez FR Sant Mamet SP Engelswies GE Sandelzhausen GE Häder GE Puente de Vallecas SP La Hidroelectrica SP 56 Montejo de la Vega SP Grund AU Faluns of T & A FR Manthelan FR Laimering 3 GE Gisseltshausen GE Napak KE cc=0.91

77

Figure 3.3b: MN5 large mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Grund AU Laimering 3 GE Walda 2 GE Griesbeckerzell GE Wannenwaldtobel-2 GE Heggbach GE 8 Contres MN5 FR Sant Mamet SP Savigné-sur-Lathan FR Poudenas FR Rimbez FR Häder GE Puente de Vallecas SP Engelswies GE 16 La Hidroelectrica SP Rothenstein 1 GE Göriach AU Hambach 6C GE Baigneaux-en Beauce FR Esvres - Marine Faluns FR Pontlevoy FR Sandelzhausen GE 24 Faluns of T & A FR Montejo de la Vega SP Antonios GR Al Sarrar SA Ad Dabtiyah SA Çandır 3 TU Paşalar TU Belometchetskaja RG Ardiç-Mordoğan TU 32 Karaağaç 1TU Gisseltshausen GE Chios GR Kipsaramon KE Maboko KE Majiwa KE Ombo KE Napak KE 40 Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE 48 Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Kalodirr KE 56 Rusinga KE Kipsaramon KE Arrisdrift NA Napak UG Ziemetshausen 1b GE La Condoue FR

Manthelan FR 64 Napak KE Sinda Congo CO/ZA

cc=0.81

78

Figure 3.3c: MN5 small mammal species dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Arrisdrift NA Kipsaramon KE Kipsaramon KE Kalodirr KE Rusinga KE Rusinga KE Rusinga KE 8 Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Napak UG 16 Rusinga KE Rusinga KE Rusinga KE Napak UG Napak UG Al Sarrar SA Sant Mamet SP Montejo de la Vega SP 24 Esvres - Marine Faluns FR Contres MN5 FR Vermes 2 SW Çandır 3 TU Paşalar TU Somosaguas-Sur SP Teiritzberg 1 AU Obergänserndorf 1 & 2 AU 32 Schellenfeld 2-4 GE Engelswies GE Vermes 1 SW Massendorf GE Rothenstein 1 GE Franzenbad CZ Belchatow PO Sandelzhausen GE 40 Affalterbach GE Gisseltshausen GE Puttenhausen GE Laimering 3 GE Hambach 6C GE Tobel-Hombrechtikon SW Martinsbrünneli SW Chatzloch SW 48 Hüllistein SW Belometchetskaja RG Ziemetshausen 1b GE Wannenwaldtobel-2 GE Frohberg SW Muhlbach AU Grund AU Antonios GR 56 Göriach AU Häder GE Heggbach GE Manthelan FR Savigné-sur-Lathan FR Pontlevoy FR Rimbez FR cc=0.90

79

Figure 3.3d: MN5 small mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Al Sarrar SA Rusinga KE Rusinga KE Rusinga KE Napak UG Napak UG Rusinga KE 8 Kalodirr KE Rusinga KE Rusinga KE Rusinga KE Rusinga KE Napak UG Rusinga KE 16 Rusinga KE Rusinga KE Rusinga KE Rusinga KE Arrisdrift NA Kipsaramon KE Kipsaramon KE Heggbach GE 24 Manthelan FR Savigné-sur-Lathan FR Çandır 3 TU Paşalar TU Karaağaç 1 TU Muhlbach AU Grund AU Antonios GR Belometchetskaja RG 32 Montejo de la Vega SP Ziemetshausen 1b GE Wannenwaldtobel-2 GE Sandelzhausen GE Laimering 3 GE Teiritzberg 1 AU Obergänserndorf 1 & 2 AU Massendorf GE 40 Tobel-Hombrechtikon SW Puttenhausen GE Vermes 1 SW Engelswies GE Gisseltshausen GE Martinsbrünneli SW Schellenfeld 2-4 GE 48 Franzenbad CZ Hambach 6C GE Belchatow PO Chatzloch SW Esvres - Marine Faluns FR Affalterbach GE Sant Mamet SP Vermes 2 SW 56 Frohberg SW Rimbez FR Contres MN5 FR Rothenstein 1 GE Hüllistein SW Somosaguas-Sur SP Pontlevoy FR Göriach AU Häder GE 64 Koçgazi TU cc=0.91

80

MN6

In contrast to MN5, the MN6 large mammal species do not form a distinct Turkish cluster (Figure 3.4a). Although there are a number of species endemic to the region

(including several bovid species, a suid and palaeochoerid, a rhino and a mammutid), the lack of association in the Turkish localities is due to the combined influence of low species richness and a lack of shared species among localities, coupled with the presence of cosmopolitan taxa. The MN6 Spanish faunas are similar in this respect. Although several species from this region are shared with France and Slovakia, the low richness of these faunas and presence of cosmopolitan taxa complicates the placement of the Spanish locality in the large mammal species dendrogram. This locality is grouped with the single

Turkish locality based on the presence of one shared rhino, Alicornops simorrensis. A

Greek cluster is not recognized for MN6 due to a lack of large mammal genera known from this region during the MN6 interval. The Central European province of MN5 continues to be supported during MN6, although by a smaller number of species (cervids and suids).

However, smaller clusters of localities are evident. These include clusters formed by

Slovakian faunas (supported by species of suid, Aureliachoerus aurelianensis; deinotheriid,

Deinotherium laevius; tragulid, Dorcatherium vindobonense; bovid, Eotragus haplodon; pliopithecid, vindobonensis; mustelid, Lartetictis dubia and Trocharion albanense; and ursid, Ursavus brevirhinus), French-German-Swiss faunas (supported by species of bovid, Eotragus clavatus; cervid, Euprox furcatus; moschid, Micromeryx flourensianus; rhino, Lartetotherium sansaniensis; pliopithecid, Pliopithecus antiquus; and deinotheriid, Deinotherium giganteum) and Austrian faunas (supported by species of bovid,

Tethytragus langai; and pliopithecid, Plesiopliopithecus lockeri). The large mammal species

81 from the Saudi Arabian group more closely with the Eurasian faunas, due to the shared gomphotheriid, Gomphotherium angustidens, which is known from France and Turkey.

There are no common taxa between this region and Africa, nor are there any shared large mammal species between Africa and Eurasia during this temporal interval.

The Turkish large mammal genera in MN6 form weak clusters with the Central

European group, because of very few endemic genera (the bovid, Turcocerus, and rhino,

Beliajevina) (Figure 3.4b). The Turkish localities are in large part comprised of more widely- ranging taxa, which are shared with faunas in Central Europe (particularly France), Kenya and Zaire. The Raup-Crick similarity analysis weakly groups the Zaire and Turkish faunas, while the Jaccard and Dice analyses do not recognize the single shared widely-ranging rhino,

Brachypotherium, and group the Zaire locality with other African localities. The two

Turkish localities analysed for MN6 do not cluster together because although they share common genera, these genera are widespread, and no endemic taxa are shared between these localities. Large mammal genera from Greece are poorly sampled for this temporal interval and are thus excluded in the MN6 analysis. The Spanish fauna weakly groups with French and Turkish localities based on the co-occurrence of suid (Listriodon) and rhino (Alicornops) taxa, in addition to several carnivore genera (including the felid, Pseudaelurus, also known from Slovakia; hyaenid, Protictitherium; nimravid, Sansanosmilus, also known from

Slovakia; and ursid, Plithocyon). The Central European province is maintained and well- supported by a diversity of large mammals, however again includes smaller sub-groupings supported by their constituent taxa. These include a French-German-Swiss-Austrian grouping and a Slovakian grouping. The former is supported by several rhino genera

(Lartetotherium, Plesiaceratherium, Prosantorhinus), a bovoid (Amphimoschus), cervid

82

(Stehlinoceros), cainotheriid (Cainotherium) and pliopithecid (Pliopithecus). The latter is supported by mustelid genera (Trocharion, Lartetictis), an ursid (Ursavus), rhino

(Hoploaceratherium), equid (Hippotherium) and pliopithecid (Epipliopithecus). The African localities share the tragulid, Dorcatherium; gomphotheriid, Gomphotherium; suid,

Albanohyus; and rhinos, Aceratherium and Dicerorhinus, with European faunas, as well as the rhino, Brachypotherium, and the hominoid, Kenyapithecus, with Eurasian faunas. These localities share the gomphotheriid, Choerolophodon, with Turkish faunas. The African localities also share several giraffid and bovid genera exclusively with faunas from Saudi

Arabia.

The MN6 small mammal species create clusters that are very similar to the large mammals (Figure 3.4c). Due to the lack of small mammal faunas at this time period, the

Turkish province is again poorly differentiated and groups with Swiss faunas based on the shared castorid, Chalicomys jaegeri (also known from Germany). Similarly, the Spanish small mammal species are represented in MN6 at a single locality. While comprised of numerous endemic species, including murid, sciurid and glirid rodents, soricid insectivores, an ochotonid lagomorph and a castorid, this locality clustered weakly with the Central

European group, based on a single shared soricid insectivore, Miosorex grivensis, which is also found in France. The lack of small mammals from Greece excludes this region from analysis. The small mammal species during MN6 form a broad Central European cluster, which subdivide into well-supported smaller groupings. These include a group formed by

Slovakian species, which despite having many endemics (glirid, murid and pteromyid rodents; erinaceid, soricid and talpid insectivores), also shares taxa with Hungary (murid,

Anomalomys gaudryi, Eumyarion latior; and glirid rodents, Muscardinus sansaniensis; and

83 dimylid insectivore, Plesiodimylus chantrei), Romania (glirid, Muscardinus sansaniensis),

Germany (murid, Eumyarion weinfurteri; and eomyid rodent, Keramidomys carpathicus),

France (murid, Cricetodon sansaniensis, Democricetodon gaillardi, Eumyarion latior,

Megacricetodon schaubi, Neocometes brunonis; and glirid rodents, Bransatoglis astaracensis and Muscardinus sansaniensis; dimylid insectivore, Plesiodimylus chantrei; erinaceid,

Lanthanotherium sansaniensis; and talpid, Talpa minuta) and Switzerland (murid rodent,

Democricetodon gaillardi; and dimylid insectivore, Plesiodimylus chantrei). Similarly, the smaller French cluster, supported by glirid, murid and pteromyid rodents, as well as erinaceid, heterosoricid and soricid insectivores, shares species with Slovakia, Hungary,

Romania, Switzerland and Germany. The Swiss cluster is formed by fewer endemic species, including murid, eomyid and sciurid rodents and a soricid insectivore, however the raw data indicates that the majority of fauna from this region are also shared with France, Germany,

Slovakia, Hungary and Romania. Similar to the large mammal species, the African small mammals have no species in common with Eurasia or Saudi Arabia.

The MN6 small mammal genera broadly support the species-level observations

(Figure 3.4d). The Turkish localities do not cluster together because although they each share common taxa with regions to the west, they lack any common genera with each other.

Greek small mammal genera are not known for MN6. The Central European province is still recognizable and the fauna still support a Slovakian, French and Swiss cluster, however few endemic genera support these groupings and more commonly, localities in these regions are comprised mostly of genera known to the Central European region. The African faunas share no common taxa with either Eurasia or Saudi Arabia, although the latter shares a sciurid rodent, Atlantoxerus, with Turkish faunas.

84

Figure 3.4a: MN6 large mammal species dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Diessen am Ammersee GE Kreutzlingen SW Stein am Rhein SW

4 Four FR

Trimmelkam AU

Steinberg GE

Sansan FR

8 Simorre FR Castelnau-d'Arbieu FR

Inönü I TU

Arroyo del Val SP

12 Catakbagyaka TU

Liet FR Rümikon SW

Stätzling GE

Thannhausen GE 16 Klein Hadersdorf AU Devínská Nová Ves SL

Devínská Nová Ves SL

Devínská Nová Ves SL 20 Elgg SW

Al Jadidah SA

Fort Ternan KE

Fort Ternan KE 24 Ngorora KE

Kipsaramon KE

Kipsaramon KE

Maboko KE 28 Majiwa KE Ombo KE

Nyakach KE

32 Nyakach KE

Nyakach KE

Sinda Congo ZA/CO cc=0.89

85

Figure 3.4b: MN6 large mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Al Jadidah SA

Ngorora KE

Kipsaramon KE Maboko KE 4 Majiwa KE

Ombo KE Nyakach KE

8 Nyakach KE Nyakach KE

Fort Ternan KE

Fort Ternan KE 12 Kipsaramon KE

Sinda Congo ZA/CO Devínská Nová Ves SL

Elgg SW

16 Trimmelkam AU Sansan FR

Devínská Nová Ves SL Devínská Nová Ves SL

20 Liet FR

Rümikon SW Stätzling GE

Thannhausen GE 24 Klein Hadersdorf AU

Castelnau-d'Arbieu FR

Simorre FR

28 Arroyo del Val SP Inönü I TU

Catakbagyaka TU Four FR

Wissholz SW 32 Steinberg GE

Diessen am Ammersee GE Kreutzlingen SW Stein am Rhein SW

36 Armantes 7 SP cc=0.87

86

Figure 3.4c: MN6 small mammal species dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Devínská Nová Ves SL

Devínská Nová Ves SL 2.5 Four FR

Four FR

Matraszolos 1-2 HU 5

Schwamendingen SW

Rümikon SW

7.5 Zeglingen SW

Samsonhaza HU

10 Wissholz SW

Steinberg GE

Unterneul GE 12.5 Sansan FR

Armantes 7 SP

15 Liet FR

Subpiatrã 2/1R RO

Stätzling GE 17.5

Catakbagyaka TU

Elgg SW

Al Jadidah SA 20

Kipsaramon KE

Kipsaramon KE 22.5 cc=0.92

87

Figure 3.4d: MN6 small mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Sansan FR

Devínská Nová Ves SL 2.5 Four FR

Four FR

Rümikon SW 5 Wissholz SW

Zeglingen SW

7.5 Samsonhaza HU

Unterneul GE

Steinberg GE 10

Liet FR

Devínská Nová Ves SL

12.5 Matraszolos 1-2 HU

Schwamendingen SW

15 Subpiatrã 2/1R RO

Bagici TU

Stätzling GE 17.5 Elgg SW

Catakbagyaka TU

Al Jadidah SA 20 Kipsaramon KE

cc=0.95 Kipsaramon KE 22.5

88

MN7/8

During this temporal interval, the large mammal species support a weak cluster of some, but not all Turkish localities (Figure 3.5a). The Yeni Eskihisar and Yeni Eskihisar 1 localities cluster on the basis of few shared species, but interestingly other Turkish localities group with the Tunisian locality of Bled Douarah on the basis of the shared gomphotheriid,

Gomphotherium angustidens, and the bovid, Protoryx solignaci. Although these localities preserve a number of endemic species, including mustelids, felids, hyaenids, rhinos, and bovids, they also contain a number of more cosmopolitan taxa. The other Turkish localities such as Sofca cluster with those in Germany, while Sariçay clusters with Spain, due to shared species, including the rhino, Brachypotherium brachypus, in the former and the hyaenid,

Protictitherium crassum, in the latter. Both localities share the chalicotheriid,

Chalicotherium grande, with Germany, France and Spain. Sofca and Sariçay lack any common large mammal species. Spanish large mammal species in MN7/8 are represented at more localities than in the preceding intervals. These faunas support clusters of Spanish localities, in addition to clustering with Turkey (mentioned above), weakly with Switzerland, based on the moschid, Micromeryx flourensianus, and Austria, based on the cervid, Euprox furcatus (also known from Poland and the Ukraine) and suid species, Listriodon splendens

(also known from Turkey). The Spanish clusters are supported by bovid (Miotragocerus monacensis, Protragocerus chantrei), suid (Albanohyus castellensis, Parachleuastochoerus huenermanni, Propotamochoerus palaeochoerus), mustelid (Palaeomeles pachecoi,

Semigenetta grandis, Trocharion albanense), hyaenid (Protictitherium gaillardi), nimravid

(Sansanosmilus jourdani), amphicyonid (Amphicyon major), rhino (Hoploaceratherium tetradactylum, Lartetotherium sansaniensis) and deinotheriid (Deinotherium giganteum), that

89 although are endemic to this region, are found at all localities, hence the clustering pattern.

Like MN6, Greek faunas are not sampled for this temporal interval. The Central European faunas divide into a more eastern Polish-Ukrainian group and a more central-western French-

German-Swiss group, however these regions continue to share a number of common species.

The Polish-Ukraine group is only weakly supported by the common occurrence of a widely ranging Western and Central cervid species, Euprox furcatus, and in fact, the Polish faunas have more common taxa with Central Europe, than do the Ukrainian fauna, which are comprised mostly of taxa endemic to that region. The French-Swiss-German grouping is supported by palaeochoerid (Taucanamo grandaevum), mustelid and viverrid (Semigenetta sansaniensis, Trochotherium cyamoides; Leptoplesictis aurelianensis), as well as hominid species (Dryopithecus fontani, also known from Austria). Fauna supporting the African province are almost all endemic, with the exception of the gomphotheriids, Gomphotherium angustidens and Tetralophodon longirostris, shared between Tunisia and Germany (former) and Tunisia, Germany and Spain (latter), in addition to the bovid, Protoryx solignaci, shared between Tunisia and Turkey.

The MN7/8 large mammal genera from Turkey support the previous clusters of the

Yeni Eskihisar localities, to the exclusion of Sariçay and Sofca, which again, form groupings based on shared genera with Austria, Germany, France and Spain (Figure 3.5b). Similar to the species-level analysis, the large mammal genera from Spain form a cluster of localities, while also grouping weakly with Austria, Poland, and Turkey and the Ukraine. The Spain-

Austria cluster is based on the shared cosmopolitan suid, Listriodon, as well as the widespread Western and Central European hominid, Dryopithecus. The Spain-Poland grouping is again based on shared widespread genera, including the cervid, Euprox, and the

90 ursid, Ursavus. The Spain-Turkey-Ukraine grouping is supported exclusively by the hyaenid, Protictitherium, also known from France. Thus these clusters are based on very few widespread genera, and also happen to be notably species-poor. Large mammal genera from

Central Europe continue to support a large France-Germany-Austria-Poland grouping, which broadly corroborates the species-level findings. In contrast to the species-level analysis, the African province shares an appreciable number of genera in common with

Eurasia. These include bovid (Gazella, Pachytragus, Protoryx, Protragocerus), tragulid

(Dorcatherium), suid (Albanohyus), giraffid (Palaeotragus), gomphotheriid

(Choerolophodon, Gomphotherium, Tetralophodon), deinotheriid (Deinotherium), chalicotheriid (Chalicotherium), felid (Machairodus) and hyaenid (Percrocuta) genera, which are all to varying degrees widespread.

The MN7/8 small mammal species from Turkey support a distinct clustering of localities (Figure 3.5c). This cluster is supported by several murid rodent species (Byzantinia bayraktepensis, B. eskihisarensis, Myocricetodon eskihisarensis), talpid (Desmanella cingulata, Desmanodon major) and erinaceid insectivores (Mioechinus tobieni, Schizogalerix anatolica). Moreover, this cluster shares very few taxa in common with European faunas, with only two murid (Megacricetodon similes, Neocometes brunonis) and one pteromyid rodent species (Albanensia albanensis) in common with Switzerland. One exclusion to the

Turkish cluster is Bayraktepe 1, which has only one species in common with the Turkish cluster of localities, the murid, Byzantinia bayraktepensis. Bayraktepe 1 groups with Spain and Poland (Jaccard and Dice) and Poland (Raup-Crick), based on the common occurrence of the castorids, Chalicomys jaegeri (also known from Germany), and Trogontherium minutum

(also known from Switzerland and Hungary). The Spanish small mammal species in MN7/8

91 form a much more distinct cluster than in the previous interval. This cluster is extremely well supported by murid (Cricetodon lavocati, Democricetodon crusafonti, D. nemoralis,

Eumyarion leemanni, Hispanomys dispectus, Megacricetodon debruijni, M. ibericus) and glirid rodents (Myomimus dehmi, Muscardinus hispanicus, M. vallesiensis, Tempestia hartenbergeri), in addition to several soricid (Alloscapanus lehmani, Crusafontina excultus), erinaceid (Amphechinus golpeae), talpid (Domninoides santafei, Talpa minuta, T. vallesensis) and heterosoricid insectivore species (Dinosorex sansaniensis). Greek small mammals are not sampled for this interval. The larger more central European province of

French-Swiss-German-Hungarian-Romanian faunas support smaller divisions of Hungarian-

Romanian localities and French-Swiss-German-Hungarian localities. The former group is supported by murid (Eumyarion medius) and glirid (Glirulus lissiensis) rodents, in addition to a larger number of more cosmopolitan species. The latter group is supported by eomyid

(Eomyops oppligeri), glirid (Microdyromys complicatus, Miodyromys aegercii, M. hamadryas) and murid (Democricetodon brevis) rodents (as well as by murids and pteromyids also found in Romania), however, still shares many species in common with the larger Central European cluster. No African small mammals were included in the MN7/8 species-level analysis due to lack of adequate samples for this time period.

The MN7/8 small mammal genera support clear clusters of both Turkish and Spanish localities (Figure 3.5d). The Turkish cluster is supported by several murids and sciurids, a glirid rodent, in addition to an erinaceid and talpid insectivore. Similarly, the Spanish localities are supported by murid, sciurid and glirid rodent, together with soricid and talpid insectivore genera. No Greek small mammal genera were available from this temporal interval. The previously clustered Hungary-Romania group also includes Swiss localities,

92 supported by pteromyid genera, although Germany and Hungary also share a number of common small mammal genera. Collectively, this more central region of Europe shares a larger number of wide-ranging genera in common with the surrounding regions (ie. Spain,

Turkey), rather than having its own exclusively endemic forms. There are no MN7/8 small mammal genera from Africa included for this interval.

93

Figure 3.5a: MN7/8 large mammal species dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Fort Ternan KE

Nyakach KE

3 Nyakach KE

Nyakach KE

Fort Ternan KE

6 Ngorora KE

Opole 2 PO

Przeworno PO

Sevastopol UK 9 Can Feliu SP

Nombrevilla 2 SP

12 St. Stephan AU

Sofca TU

Massenhausen GE

Poudendas-Cayron FR 15 Sariçay TU

Can Mata 1 SP

18 Sant Quirze SP

Hostalets de Pierola In SP

Castell de Barberà SP

La Grive St. Alban FR 21 Steinheim GE

Anwil SW

24 Escobosa SP

St. Gaudens FR

Yeni Eskihisar TU

Yeni Eskihisar TU 27 Bled Douarah TUN cc=0.86

94 Figure 3.5b: MN7/8 large mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Felsotárkány HU

Can Feliu SP

St. Stephan AU

4 Can Mata 1 SP

St. Gaudens FR

Nombrevilla 2 SP

Opole 2 PO

8 Anwil SW

Sariçay TU

Massenhausen GE

Przeworno PO

12 La Grive St. Alban FR

Steinheim GE

Sant Quirze SP

Castell de Barberà SP 16

Hostalets de Pierola In SP

Poudendas-Cayron FR

Sofca TU

20 Fort Ternan KE

Fort Ternan KE

Nyakach KE

Nyakach KE

24 Nyakach KE

Ngorora KE

Bled Douarah TUN

Escobosa SP 28 Yeni Eskihisar TU

Yeni Eskihisar TU

Sevastopol UK cc=0.86

95 Figure 3.5c: MN7/8 small mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Comanesti-1 RO

Felsotárkány HU 2.5 Bois de Raube 3 SW

Felsotárkány HU

5 Steinheim GE

La Grive St. Alban FR

Anwil SW 7.5

Sant Quirze SP

Hostalets de Pierola In SP

10 Castell de Barberà SP

Escobosa SP

Nombrevilla 2 SP 12.5

Can Feliu SP

Bayraktepe 1 TU

15 Opole 2 PO

Felsotárkány HU

Felsotárkány HU 17.5

Subpiatrã 2/2 RO

Yeni Eskihisar TU

20 Yeni Eskihisar TU

Pismanköy TU

Sofca TU 22.5

Sariçay TU cc=0.86

96 Figure 3.5d: MN7/8 small mammal genera dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Felsotárkány HU

Felsotárkány HU 2.5 Subpiatrã 2/2 RO

Sant Quirze SP

5 Castell de Barberà SP

Anwil SW

Felsotárkány HU 7.5 Sariçay TU

Opole 2 PO

10 Hostalets de Pierola In SP

Escobosa SP

Nombrevilla 2 SP 12.5

Can Feliu SP

Comanesti-1 RO

15 Bois de Raube 3 SW

Steinheim GE

Felsotárkány HU 17.5 La Grive St. Alban FR

Bayraktepe 1 TU

20 Yeni Eskihisar TU

Yeni Eskihisar TU

Pismanköy TU 22.5 cc=0.76 Sofca TU

97

MN9

The MN9 Turkish localities form a distinct cluster, supported by equid

( sinapensis), giraffid (Palaeotragus coelophrys, also known from Georgia, and P. roueni), rhino (Ceratotherium neumayri), and hyaenid species (Ictitherium intuberculatum), as well as the bovid, Protoryx solignaci, which is also known from Tunisia

(Figure 3.6a). Large mammal faunas also support a Moldovan cluster of localities on the basis of shared equid (Hipparion sarmaticum), felid (Machairodus laskarevi) and percrocutid species (Dinocrocuta robusta). However, the Moldovan locality of Varnitsa clusters with the

Georgian localities of Udabno 1 and Eldari 1, based exclusively on the shared presence of the bovid, Tragoportax leskewitschi, and all localities within this cluster are notably species poor. No Greek faunas are sampled for this time period. The Spanish large mammal species for this temporal interval support well-defined clusters, based on deinotheriid (Deinotherium laevius), bovid (Protragocerus chantrei), cervid (Euprox minimus), suid

(Parachleuastochoerus crusafonti), tragulid (Dorcatherium jourdani), equid (Hipparion catalaunicum), chalicotheriid (Chalicotherium grande), rhino (Lartetotherium sansaniensis), mustelid (Limnonyx sinerizi), amphicyonid (Thaumastocyon dirus), hyaenid (Protictitherium gaillardi), and ursid species (Indarctos vireti). The Central European group is divided into a

France-Germany-Switzerland-Austria group, as well as a Hungary-Austria group. However, while pliopithecid and hominid primates ( hernyaki and Dryopithecus brancoi) potentially unite the Hungary-Austria cluster (if the species in both countries are the same), the faunas from these countries as well as the others in the Central European group generally lack species endemic to the region (although Germany has several endemic forms). As a whole, the Central European region is comprised largely of species whose ranges extend

98 westward into Spain and less commonly, eastward, and is therefore lacking in provincial definition during MN9. Faunas from Africa remain mainly endemic during this interval, with the exception of the bovid, Protoryx solignaci, shared between Tunisia and Turkey, the equid, Hippotherium primigenium, shared between Algeria, Kenya, Tunisia and Spain, and the gomphotheriid, Gomphotherium angustidens, shared between Tunisia, Spain and

Germany.

At the genus level, the Turkish cluster remains intact, and although several localities stray from this cluster, a number of large mammal genera still support this grouping, including equid, bovid, and rhino taxa (Figure 3.6b). Interestingly, this cluster also has several genera in common with Georgia and Moldova, including the bovid, Tragoportax, and the percrocutid, Dinocrocuta. These regions, including Turkey, also share the gomphotheriid genera, Choerolophodon, with Kenya, the bovid, Gazella (also known from Tunisia) and the percrocutid, Percrocuta, with Kenya and Algeria. The Georgian localities, in addition to sharing these taxa, also share genera in common with Spain, including the suid, Microstonyx, as well as the bovid, Tragoportax (also known from Austria and Moldova) and the giraffid,

Palaeotragus (also known from Moldova and Kenya). Although Spanish localities are fairly dispersed throughout the dendrogram topology, this region has several endemic genera, including amphicyonids, mustelids, bovids (Protragocerus, also known from Kenya) and moschids. A similar case is evident with the Central European localities, however unlike

Spain, this region lacks large mammal genera that are exclusive to it, instead, producing small groups that usually include Spain. Continuing the trend observed in MN7/8, the

African localities share an extensive list of genera in common with Eurasian localities, including the bovids, Prostrepsiceros (Turkey, Algeria), Protoryx (Turkey, Tunisia) and

99

Protragocerus (Spain, Kenya); the suids, Sivachoerus (Turkey, Kenya) and Albanohyus

(France, Spain, Kenya); the tragulid, Dorcatherium (Turkey, Spain, Switzerland, Germany,

Hungary, Kenya); the equid, Hipparion (ubiquitous); the rhinos, (Turkey,

Georgia, Kenya) and Dicerorhinus (Spain, Germany, Moldova, Ethiopia, Algeria); the chalicotheriid, Ancylotherium (Spain, Ethiopia); the giraffids, Samotherium (Turkey,

Algeria, Kenya) and Palaeotragus (Spain, Moldova, Turkey, Georgia, Kenya); the deinotheriid, Deinotherium (Spain, Switzerland, Germany, Austria, Hungary, Moldova,

Georgia, Kenya); the gomphotheriids, Choerolophodon (Turkey, Moldova, Georgia, Kenya),

Gomphotherium (Spain, Germany, Tunisia, Kenya) and Tetralophodon (Spain, Switzerland,

Germany, Hungary, Moldova, Georgia, Tunisia); the percrocutid, Percrocuta (Turkey,

Algeria, Kenya); and finally, the felid, Machairodus (Spain, Switzerland, Germany,

Moldova, Turkey, Ethiopia). As a result of this great number of intercontinental co- occurrence, the African localities are likewise interspersed within the dendrogram topology for the MN9 interval. This is an interesting finding in light of the timing of potential hominid dispersals into Africa during this interval. Eurasian large mammals that are incompletely sampled for this interval include the bovids, Ouzoceros and Pachytragus, which are only known from Tunisia in MN9.

The results of the small mammal species level analysis are strikingly more resolved than the large mammal analysis (Figure 3.6c). The Turkish small mammals support distinct clusters, well-supported by murid rodents (Byzantinia bayraktepensis, B. dardanellensis, B. nikosi, B. ozansoyi, Heramys anatolicus, Myocricetodon eskihisarensis, Progonomys minus), erinaceid insectivores (Schizogalerix anatolica, S. intermedia, S. sinapensis), and an ochotonid lagomorph (Bellatonoides eroli). The murid, Progonomys cathalai, is also

100 common to Moldova, Algeria and Morocco. The Moldovan localities cluster together, well- supported by murid (Bujoromys olarensis, Byzantinia orientalis, Cricetulodon bugesiensis,

Kowalskia moldavica, Pseudocricetus polgardiensis, Ruscinomys orientalis), sciurid

(Spermophilinus turoliensis), dipodid (Sarmatosminthus genuine), and castorid rodents

(Castor neglects), erinaceid (Galerix exiles, G. sarmaticum, Schizogalerix sarmaticum) and soricid insectivores (Chemisorbed schaubi), as well as an ochotonid lagomorph

(Proochotona kalfaense). The Spanish cluster of localities is similarly well supported by murid (Cricetodon lavocati, Cricetulodon sabadellensis, Democricetodon crusafonti, D. sulcatus, Eumyarion leemanni, Hispanomys dispectus, H. peralensis, H. thaleri, Kowalskia seseae, Megacricetodon debruijni, M. ibericus, Progonomys hispanicus), glirid (Eliomys truci, Miodyromys hamadryas, Tempestia hartenbergeri), sciurid (Heteroxerus rubricate), eomyid (Keramidomys carpathicus), erinaceid (Postpalerinaceus vireti), and pteromyid

(Miopetaurista crusafonti) rodents, as well as talpid (Talpa vallesensis) and soricid insectivores (Crusafontina excultus), and ochotonid lagomorphs (Eurolagus fontannesi,

Prolagus crusafonti). Small mammal species from Greece are not sampled for this interval.

Localities from the Central European region are again divided into smaller groupings, including a Czech-Austrian-Hungarian-French group and a German-French-Swiss group, however both sub-groups share more common taxa exclusively to this larger region, than either does to itself. These species include murid (Anomalomys gaudryi, Democricetodon freisingensis, Kowalskia fahlbuschi, Microtocricetus molassicus), pteromyid (Albanensia grimmi), and glirid rodents (Glis minor, Glirulus lissiensis). The African small mammal species known from this temporal interval are restricted to localities in Algeria and Morocco and are entirely endemic to Africa.

101

The MN9 genus-level analysis supports the previously defined Turkish cluster

(Figure 3.6d). Localities in this region continue to share a number of common genera, but incidentally also share genera with Algeria and Morocco (the murid rodent, Myocricetodon), as well as some wide-ranging rodents, insectivores and lagomorphs with Europe. Although the Moldovan localities share a number of taxa in common with regions to the west and east, these localities cluster together based on a number of shared genera, including murid

(Bujoromys, Pseudocricetus, Ruscinomys), dipodid (Sarmatosminthus), and castorid rodents

(Palaeomys), talpid (Proscapanus) and soricid insectivores (Hemisorex) and ochotonid lagomorphs (Lagopsis). The Spanish localities also cluster based on murid (Cricetodon), sciurid (Tamias), and glirid rodents (Tempestia) and heterosoricid insectivores (Heterosorex).

This region also shares the sciurid rodent, Atlantoxerus, with Algeria, Tunisia and Turkey.

The Central European region is again divided up into smaller clusters, however, as in the species-level analysis, the larger group is better supported as a whole than the smaller clusters are as individual groups. Genera supporting this cluster include the murid,

Parapodemus; the glirid, Glirulus; the dipodid, Eozapus; the insectivorous talpid, Desmana; and the soricids, Paenelimnoecus and Anourosorex. Although the African localities contain several endemic genera, they also share genera with Eurasia, including the murids,

Myocricetodon (Turkey) and Progonomys (Turkey, Spain, Hungary and Moldova), the glirid,

Microdyromys (Turkey, Czech Republic, Hungary and Poland), the sciurid, Atlantoxerus

(Turkey) and the erinaceid insectivore, Galerix (Spain, France, Czech Republic, Hungary and

Moldova). Of the small mammals sampled for MN9, only the murid, Paraethomys, sampled in Algeria, is incompletely sampled in Eurasia for this interval.

102

Figure 3.6a: MN9 large mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Sümeg HU Buzhor 1 MO Atavaska MO Kalfa MO 5 Vösendorf AU Höwenegg GE Melchingen GE Priay II FR Wissberg GE Esselborn GE 10 Eppelsheim GE Charmoille SW Can Llobateres SP Can Ponsic SP 15 Santiga SP Los Valles de F SP El Firal SP Hostelats de Pierola Su SP Ballestar SP 20 Rudabánya HU Subsol de Sabadell SP San Miquel SP Hammerschmiede GE Wartenberg GE 25 Nombrevilla SP Götzendorf AU Mariathal AU Seu d’Urgel SP Varnitsa MO 30 Eldari 1 RG Udabno 1 RG Middle Sinap TU Esme Akçaköy TU Alkcaköy TU 35 Sinap 4 Sinap 72 Sinap 94 Sinap 8A Sinap 91 40 Sinap 8B Doué-la-Fontaine FR Pedregueras SP Nakali KE Bou Hanifia 5 AL 45 Bled Douarah TUN Ngorora KE cc=0.85

103 Figure 3.6b: MN9 large mammal genera dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Oued el Atteuch AL Sinap 8A Götzendorf AU Mariathal AU Chorora Frm ET 6 Los Valles de F SP Buzhor 1 MO Kalfa MO Atavaska MO Vösendorf AU Höwenegg GE 12 Ballestar SP Wissberg GE El Firal SP Eppelsheim GE Subsol de Sabadell SP Charmoille SW 18 Esselborn GE Can Llobateres SP Rudabánya HU Hammerschmiede GE Can Ponsic SP Santiga SP 24 Hostelats de Pierola Su SP Wartenberg GE Nombrevilla SP Melchingen GE Doué-la-Fontaine FR Priay II FR 30 Pedregueras SP Bou Hanifia 5 AL Awash 1 ET Varnitsa MO Sinap 72 TU Sümeg HU 36 Sinap 91 Sinap 8B Alkcaköy TU Nakali KE Ngorora KE San Miquel SP 42 Udabno 1 RG Eldari 1 RG Bled Douarah TUN Middle Sinap TU Esme Akçaköy TU Sinap 4 TU 48 Sinap 94 TU Sinap 108 TU Sinap 84 TU cc= 0.75

104 Figure 3.6c: MN9 small mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Amama 1 AL

Sinap 84 TU

Sinap 85 TU

4 Sümeg HU

Middle Sinap TU

Bayraktepe 2 TU

Sinap 108 TU

8 Sig 1 AL

Oued Zra MOR

Bou Hanifia 5 AL

Suchomasty CZ

12 Götzendorf AU

Rudabánya HU

Jujurieux FR

Belchatow A PO

Hammerschmiede GE 16

Can Llobateres SP

Can Ponsic SP

Hostelats de Pierola Su SP

20 Buzhor 1 MO

Kalfa MO

San Miguel SP

Charmoille SW 24 Doué-la-Fontaine FR

Vösendorf AU

Priay II FR

Pedregueras SP

Pedregueras SP 28 Peralejos 5 SP

cc = 0.88 Sinap 4 TU

105 Figure 3.6d: MN9 small mammal genera dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Priay II FR Middle Sinap TU Sinap 4 TU

4 Bayraktepe 2 TU Sinap 4 TU

Sinap 84 TU Sinap 85 TU

8 Sinap 8A TU Sinap 94 TU Sinap 108 TU

Sig 1 AL

12 Oued Zra MOR

Bou Hanifia 5 AL Bou Hanifia AL

Oued el Atteuch AL

16 Amama 2 AL Amama 1 AL Ballestar SP

Charmoille SW 20 Buzhor 1 MO Kalfa MO

Vösendorf AU Belchatow A PO 24 Can Llobateres SP

Rudabánya HU Götzendorf AU Suchomasty CZ Jujurieux FR 28 Hammerschmiede GE

Sümeg HU Can Ponsic SP

32 Hostelats de Pierola Su SP

Pedregueras SP Pedregueras SP

Peralejos 5 SP 36 San Miguel SP Doué-la-Fontaine FR Awash 1 ET

cc = 0.86

106

MN10 Large mammal species in Turkey support clusters based on gomphotheriid

(Choerolophodon anatolicus), rhino (Acerorhinus zernowi) and giraffid species

(Palaeotragus coelophrys), in addition to several endemic taxa known from individual localities (Figure 3.7a). However, the Turkish faunas share more commonalities with those in Greece than within the Turkish region. Taxa supporting the clustering of Turkish with

Greek localities include the rhinos, Ceratotherium neumayri and Chilotherium kiliasi; the giraffid, Bohlinia attica and Helladotherium duvernoyi; the bovid, Tragoportax amalthea; and the gomphotheriid, Choerolophodon pentelici. Taxa supporting the clusters of Greek localities include the hominid, Ouranopithecus macedoniensis; the bovid, Oioceros praecursor; and the giraffid, Decennatherium macedoniae. The Greek faunas share few common taxa with regions to the west; only the hyaenid, Protictitherium crassum, and the equid, Cremohipparion macedonicum, with French faunas, and the equid, Cremohipparion mediterraneum, and the bovid Tragoportax gaudryi, with Spain. The Spanish faunas cluster on the basis of only two shared species (the moschid, Micromeryx flourensianus, and the giraffid, Decennatherium pachecoi), however many hyaenid, mustelid, felid, ursid, tragulid, cervid, bovid, suid, deinotheriid, hominid, equid and chalicotheriid species are also unique to individual localities in this region. The Central European region is represented by very few localities during MN10. Like Spain, this region lacks species known from all the localities and instead is comprised of species restricted to individual countries (i.e., France, Austria,

Germany), together with more cosmopolitan taxa. The African faunas are poorly represented in MN10 and although there do not appear to be any shared taxa with Europe, this could be an artefact of sampling in this region.

107

The clusters of Turkish localities are weakly supported by only two genera: the rhino,

Acerorhinus and the bovid, Palaeoreas, the latter of which is also shared with Kenya (Figure

3.7b). However, there are a number of genera that are known exclusively from single

Turkish localities. Genera are also shared between Turkey and neighbouring Greece, including the bovid, Prostrepsiceros; the giraffids, Bohlinia and Helladotherium; the rhino,

Ceratotherium; the chalicotheriid, Ancylotherium; the aardvark, Orycteropus; and the gomphotheriid, Choerolophodon. The Spanish localities form distinct clusters based on their shared genera. These taxa include the suid, Propotamochoerus and Schizochoerus; the cervid, Euprox; and the felid, Felis. The Central European region is again represented by very few localities and although it has several genera exclusive to individual localities, this region contains only one shared taxon, the hyaenid, Ictitherium, with the remaining fauna being more cosmopolitan. Localities in Africa share a number of endemic genera, however they also share several taxa with Eurasia, including the previously mentioned bovid,

Palaeoreas, as well as Gazella; the giraffid, Palaeotragus; the chalicotheriid,

Chalicotherium; the equid, Hipparion; the gomphotheriid, Tetralophodon and the deinotheriid, Deinotherium, which are all widespread in Eurasia. The only incompletely sampled taxon in Eurasia is the bovid, Pachytragus, sampled only in Kenya for this interval.

The MN10 small mammal species from Turkey are very poorly sampled, and as a result, the Turkish localities, despite individually having several endemic species, cluster weakly with the Algerian locality of Feid El Atteuch based on the murid, Progonomys cathalai, which has a very widespread distribution (Figure 3.7c). The cluster of Spanish localities, in contrast, is extremely well supported by numerous small mammal species, including the murid rodents, Cricetulodon hartenbergeri, Hispanomys peralensis,

108

Hispanomys thaleri and Progonomys hispanicus; the glirid, Tempestia hartenbergeri; the castorid, Eucastor adroveri; the erinaceid insectivore, Galerix socialis, the soricids,

Crusafontina excultus and Crusafontina endemica; the talpid, Talpa vallesensis; and the lagomorph, Eurolagus fontannesi. Many endemic species are also known from individual localities in this region. Within the Central European region, localities in France group together, as do localities in Austria. The sole German locality of Salmendingen remains isolated due to the presence there of only two small mammal species, Hystrix parvae and

Dipoides problematicus, the latter of which is only shared with Spain. Clustering of the

French localities is well supported by eomyid (Keramidomys pertesunatoi), glirid (Myoglis meini, Paraglirulus werenfelsi), murid (Hispanomys mediterraneus; Kowalskia ambarrensis,

Rotundomys montisrotundi), pteromyid (Albanensia grimmi) and sciurid rodents

(Heteroxerus grivensis), as well as the talpid insectivore, Archaeodesmana vinea. Although the French localities cluster together, they also share a significant number of species in common with Austria (n=16), while localities in the latter region are grouped based on two shared species, the glirid, Glirulus lissiensis; and murid, Epimeriones austriacus. The

African faunas for MN10 are represented exclusively at localities in Algeria. These localities cluster on the basis of several shared species endemic to the region, together with the previously mentioned cosmopolitan murid rodent, Progonomys cathalai.

The MN10 small mammal genera broadly support the pattern found in the species- level analysis (Figure 3.7d). Although individual Turkish localities contain small mammal genera that are endemic to the region, these localities cluster on the basis of three shared rodent genera: the gerbillid, Pseudomeriones; and the murids Byzantinia and Pliospalax.

Similarly, the Spanish localities cluster together, but on the basis of few shared genera (the

109 castorid, Eucastor; glirid, Tempestia; and murid Democricetodon) and mostly on the presence of endemic, locality-specific taxa. Unlike the species-level analysis, the Spanish faunas share the majority of their genera with France and other regions of Europe. In the

Central European region, Austria and France each have endemic genera; the murid,

Epimeriones; soricid, Blarinella; and talpid, Desmana, in Austria; and the glirids, Myoglis and Paraglirulus; pteromyid, Albanensia; heterosoricid, Dinosorex; and talpid,

Archaeodesmana, in France. Like the species-level analysis, these two regions share more common genera than they do genera exclusive to their individual regions. These include glirid (Glirulus, Graphiurops, Paraglirulus, Vasseuromys), murid (Prospalax), and pteromyid rodents (Blackia, Hylopetes, Pliopetaurista), as well as soricid insectivores

(Anourosorex and Petenyia). The African small mammal genera are represented from localities in Algeria. These localities share no small mammal genera in common with

Eurasia, with the exception of the murid, Progonomys.

110

Figure 3.7a: MN10 large mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Nkondo UG

Samburu KE

3 Poksheshty MO

Gülpinar TU

Nikiti 2 GR

6 Ravin de la Pluie GR

Xirochori GR

Nikiti 1 GR

9 Pyrgos Vassilissis GR

Pentalophos 1 GR

Sinap 12 TU

12 Sinap 49 TU

Kohfidisch AU

Udabno II RG

15 Masía del Barbo SP Masía del Barbo SP

Masía del Barbo SP

18 La Roma 2 SP

Terrassa SP

Montredon FR

21 Yulafli TU

Can Purull SP

La Tarumba SP

24 Soblay FR

cc = 0.89 Salmendingen GE

\

111 Figure 3.7b: MN10 large mammal genera dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Samburu KE

Samburu KE

3 Gülpinar TU

Sinap 49 TU

Nikiti 2 GR

6 Pyrgos Vassilissis GR

Sinap 12 TU

Pentalophos 1 GR

9 Poksheshty MO

Nikiti 1 GR

Xirochori GR

12 Çorakyerler TU

Ravin de la Pluie GR

Kohfidisch AU

15 Udabno RG

Udabno RG

Can Purull SP

La Tarumba SP 18

Terrassa SP

Montredon FR

Soblay FR 21

Masía del Barbo SP Masía del Barbo SP

Masía del Barbo SP 24

La Cantera SP

La Roma 2 SP

27 Yulafli TU

Salmendingen GE cc = 0.83

112 Figure 3.7c: MN10 small mammal species dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Can Purull SP

Terrassa SP

3 Montredon FR

Lo Fournas FR

Masía del Barbo SP

6 Masía del Barbo SP

Masía del Barbo SP

Masía de la Roma SP

9 Puente Minero SP

Masía de la Roma SP

La Roma 2 SP

12 Los Aguanaces SP

La Cantera SP

Ambérieu FR 15 Ambérieu FR

Soblay FR

Douvre FR

18 Kohfidisch AU

Eichkogel AU

Sinap 49 TU

21 Feid El Atteuch AL

Sinap 84 TU

Sig 1 AL

24 Tafna AL

Salmendingen GE cc = 0.88

113 Figure 3.7d: MN10 small mammal genera dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Sinap 49 TU

Sinap 84 TU

3 Feid El Atteuch AL

Sig 1 AL

Can Purull SP

6 Masía del Barbo SP

Masía del Barbo SP

Masía de la Roma SP

9 Puente Minero SP

Masía del Barbo SP

Terrassa SP

12 Montredon FR

Lo Fournas FR

La Cantera SP

15 La Roma 2 SP

Los Aguanaces SP

Masía de la Roma SP

18 Ambérieu FR

Ambérieu FR

Kohfidisch AU

21 Eichkogel AU

Soblay FR

Douvre FR

24 Feid El Atteuch AL

Tafna AL

Salmendingen GE

cc = 0.93

114

MN11

The large mammal species in MN11 support clusters of Turkish localities. These localities share the hyaenid, Ictitherium hipparionum; the rhino, Chilotherium habereri; the bovids, Gazella gaudryi and Helicotragus rotundicornis; and the giraffid, Palaeotragus germaini (the latter also known from Kenya), in addition to a larger number of endemic species known only from individual localities (Figure 3.8a). The Turkish localities share a significant number of species in common with Greece, including the tragulid, Dorcatherium puyhauberti; the bovids, Criotherium argalioides, Gazella capricornis, Oioceros rothi, O. wegneri and Tragoportax amalthea; the rhinos, Ceratotherium neumayri and Chilotherium samium; the gomphotheriid, Choerolophodon pentelici, and the pliohyracid, Pliohyrax graecus. These two regions also share a number of species in common with Iran and Iraq, including the giraffid, Palaeotragus coelophrys; the bovid, Nisidorcas planicornis; the hyaenid, Ictitherium robustum and the mustelid, Parataxidea maraghana. The Greek large mammal species strongly support their own distinct clusters of localities, based on numerous shared species. These include equid (Cremohipparion macedonicum, C. proboscideum,

Hipparion dietrichi), rhino (Dihoplus pikermiensis, Prostrepsiceros rotundicornis), chalicotheriid (Ancylotherium pentelicum), bovid (Pachytragus crassicornis, P. laticeps,

Pseudotragus capricornis, P. parvidens), cervid (Pliocervus pentelici), pliohyracid species

(Pliohyrax kruppii) and several carnivore taxa, including the hyaenid, Plioviverrops orbignyi; the felid, Metailurus parvulus and Simocyon primigenius (incertae sedis). Greek localities also cluster weakly with the Bulgarian locality of Kalimanci 3 and 4, based mostly on the presence of shared cosmopolitan taxa, but also because of the equid, Cremohipparion matthewi, that is shared among these localities. Although the Spanish localities share many

115 common species with regions to the east, these localities cluster on the basis of shared equid

(Hippotherium primigenium), giraffid (Birgerbohlinia schaubi), cervid (Lucentia pierensis), hyaenid (Plioviverrops guerini) and mustelid (Baranogale adroveri) species. The cluster of

Central European localities, restricted to Germany and Hungary, is supported by cervid

(Cervavitulus mimus), hyaenid (Allohyaena kadici), and mustelid species (Eomellivora wimani), the latter of which is also known from the Ukraine. The Ukrainian localities cluster on the basis of a small number of fauna, including cervid (Cervavitus variabilis), bovid

(Gazella schlosseri) and mustelid species (Promephitis maeotica), and although this region also shares species in common with Turkey and Greece due to its easterly location, it also shares taxa to the west, as far as Spain. The cluster of Italian localities is not surprisingly supported by species that are completely endemic to this region, including bovid (Etruria viallii, Maremmia hauptii, M. lorenzi, Tyrrhenotragus gracillimus), suid (Eumaiochoerus etruscus), mustelid (Mustela majori, Paludolutra campanii, P. maremmana, Tyrrhenolutra helbingi) and ursid species (Indarctos anthracitis), as well as the enigmatic ape,

Oreopithecus bambolii. The African localities for this interval are notably species-poor, and although they share no common species between them, they share the giraffid, Palaeotragus germaini, with Turkey.

Although the genus-level analysis groups localities from Turkey, there are very few genera shared exclusively by localities in this region (Figure 3.8b). These include the bovids,

Helicotragus and Plesiaddax, in addition to the giraffid, Giraffa. In fact, the faunas from

Turkey share more genera in common with Greece, and to a lesser extent with adjacent regions, such as Iran. Although the Greek localities contain genera endemic to the region, there are similarly very few taxa that actually support the clusters. These taxa include the

116 bovid, Pseudotragus; the suid, Propotamochoerus; and the mustelid, Plesiogulo. Like

Turkey, the Greek faunas share more taxa in common with surrounding regions. The same is true of the Spanish large mammal genera. Although the Spanish localities cluster together, this is on the basis of two shared genera, the cervid, Lucentia, and the bovid, Birgerbohlinia, together with endemic genera known from single localities and cosmopolitan taxa. Like the species-level analysis, the Italian faunas support a cluster of localities based on the endemic taxa, however this region shares the mustelid, Mustela, with Turkey and the more wide- spread ursid, Indarctos, with Eurasia. Localities in the Ukraine cluster based on the shared cervid, Cervavitus, in addition to large mammal genera shared with other Eurasian localities.

The MN11 small mammal species support clusters of Turkish and Greek localities

(Figure 3.8c). These groupings are supported by murid (Byzantinia pikermiensis, Karnimata provocator) and hystricid species (Hystrix primigenia), in addition to endemic taxa known from individual localities. The small mammal species from Spain support a large cluster of localities, based on the murids, Hispanomys freudenthali and Kowalskia occidentalis. Like the large mammal analysis, fauna from this region tend to be dominated by more widespread species. In the Central European region, there is a clear cluster of French and French-

German localities. The French cluster is well supported by eomyid (Eomyops catalaunicus,

Graphiurops austriacus), glirid (Muscardinus austriacus), murid (Kowalskia skofleki,

Prospalax petteri, Rotundomys bressanus), and pteromyid rodents (Blackia miocaenica,

Pliopetaurista bressana), as well as erinaceid (Lanthanotherium sanmigueli), soricid

(Crusafontina kormosi, Petenyia dubia), and talpid insectivores (Archaeodesmana vinea,

Talpa gilothi).

117

The genus level analysis of small mammal faunas again support the cluster of Turkish and Greek localities, based on the gerbillid, Pseudomeriones; the glirid, Myomimus; and the murid, Byzantinia; in addition to genera with more widespread distribution patterns (Figure

3.8d). The small mammal genera continue to support a large cluster of Spanish localities.

These localities group on the basis of the glirid rodent, Eliomys, in addition to genera endemic to the region known only from individual localities, and more widespread taxa. The

Central European small mammal genera support a cluster of French localities. These localities cluster on the basis of glirid (Glirulus, Graphiurops), murid (Rotundomys) and sciurid rodents (Tamias), and a soricid insectivore (Petenyia). Despite the clustering of these localities, the French small mammals have more genera in common with Germany, than within the sample of French localities. Shared genera include the castorid, Trogontherium; eomyid, Eomyops; murid, Cricetulodon and Prospalax; and pteromyid rodents, Blackia and

Pliopetaurista; and dimylid, Plesiodimylus; erinaceid, Lanthanotherium; and soricid insectivores, Crusafontina.

118

Figure 3.8a: MN11 large mammal species dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Samburu KE

Baccinello IT Monte Bamboli

4 Halmyropotamos GR

Samos GR

Samos GR Pikermi GR

8 Ravin des Zouaves GR

Prochoma GR Vathylakkos 3 GR

Maragheh IR

12 Halmyropotamos GR

Kalimanci BU

Küçükçekmece TU BalaYaylaköy TU

16 Karacahasan TU

Garkin TU Mahmutgazi TU

Çorakyerler TU

20 Kemiklitepe TU Kemiklitepe TU

Piera SP Crevillente 2 SP 24 Puente Minero SP

Grebeniki UK Novo-Elizavetkovka UK

Vivero de Pinos SP

28 Dorn Dürkheim1 GE Csakvar HU

Kayadibi TU

Injana IR 32 Halmyropotamos GR cc = 0.85

119 Figure 3.8b: MN11 large mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Baccinello IT

Monte Bamboli

Injana IR 3

Kayadibi TU

Küçükçekmece TU

6 BalaYaylaköy TU

Karacahasan TU

Mahmutgazi TU

9 Çorakyerler TU

Ravin des Zouaves GR

Prochoma GR

12 Vathylakkos 3 GR

Garkin TU

Pikermi GR

15 Halmyropotamos GR

Samos GR

Samos GR

Kemiklitepe TU 18

Kemiklitepe TU

Maragheh IR

21 Grebeniki UK

Novo-Elizavetkovka UK

Vivero de Pinos SP

24 Piera SP

Crevillente 2 SP

Puente Minero SP

27 cc = 0.88 Dorn Dürkheim1 GE

120 Figure 3.8c: MN10 small mammal species dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Afoud 6 MOR

Kemiklitepe TU 2.5 Samos GR

Pikermi GR

Çorakyerler TU 5

Crevillente SP

Vivero de Pinos SP

7.5 Los Aguanaces SP

Puente Minero SP

10 Alfambra SP

La Gloria 10 SP Los Aguanaces SP 12.5 Masada Ruea 2 SP

Puente Minero SP

15 Peralejos D SP

Dorn Dürkheim1 GE

Dionay FR 17.5 Mollon FR

Ambérieu FR

Ambérieu FR 20

Bernardière FR

Lobrieu FR 22.5

Csakvar HU

Baccinello IT

cc = 0.97

121 Figure 3.8d: MN11 small mammal genera dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Mollon FR

Lobrieu FR

2.5

Crevillente SP

Los Aguanaces SP

Alfambra SP 5

La Gloria 10 SP

7.5 Los Aguanaces SP Puente Minero SP

Masada Ruea 2 SP

10 Puente Minero SP

Vivero de Pinos SP

Peralejos D SP 12.5

Baccinello IT

Pikermi GR

15 Samos GR

Kemiklitepe TU

Çorakyerler TU 17.5

Dionay FR

Ambérieu FR

20 Dorn Dürkheim1 GE cc = 0.94

122

MN12 The MN12 large mammals species support a cluster of Turkish localities, based on the shared occurrence of bovid (Gazella gaudryi and Sinotragus occidentalis), equid

(Hipparion matthewi and H. mediterraneum) and hyaenid (Ictitherium hipparionum and I. robustum) species in this region (Figure 3.9a). A small cluster of Turkish and Iranian localities also occurs and is supported by species of bovid (Pachytragus laticeps), equid

(Cremohipparion matthewi) and felid (Machairodus giganteus). This region also clusters broadly with areas of Eastern Europe, based on shared taxa with Greece (Hyaenotherium wongi, Pliohyrax graecus, Prostrepsiceros rotundicornis, Hipparion dietrichi,

Tetralophodon atticus), Macedonia (Nisidorcas planicornis), and Moldova (Pachytragus crassicornis, Palaeoryx majori, Chilotherium schlosseri, Orycteropus gaudryi), as well as sharing a number species broadly with the Eastern Mediterranean region, including the suid,

Microstonyx erymanthius (Bulgaria, Greece, Iran); the bovid, Palaeoreas lindermayeri

(Greece, Moldova); the giraffid, Palaeotragus roueni (Bulgaria, Greece, Macedonia,

Moldova, Ukraine); the rhinos, Ceratotherium neumayri (Bulgaria, Greece, Moldova, Iran) and Stephanorhinus pikermiensis (Bulgaria and Moldova); the equid, Cremohipparion mediterraneum (Greece, Bulgaria, Moldova); the hyaenid, Ictitherium viverrinum

(Macedonia, Moldova, Ukraine, Iran); and the gomphotheriid, Choerolophodon pentelici

(Greece, Macedonia, Ukraine, Iran). A number of large mammal species support the distinct grouping of localities in Spain. These taxa include species of bovid (Hispanodorcas torrubiae), cervid (Turiacemas concudensis), giraffid, (Birgerbohlinia schaubi), equid

(Hippotherium primigenium, Hipparion concudense), mustelid (Baranogale adroveri,

Enhydriodon lluecai, Martes basilii, Sivaonyx lehmani, S. lluecai), felid (Metailurus major), hyaenid (Plioviverrops guerini), and canid (Canis cipio), in addition to numerous endemic

123 taxa known only from individual localities in this region. Large mammal species from

Central Europe for this interval are known from France and Hungary. With the exception of cosmopolitan taxa, fauna from these regions share no common endemic species. Instead each shares species with Spain, and Hungary shares a number of taxa in common with regions to the east, including the giraffid, Helladotherium duvernoyi (shared with Turkey,

Greece, Macedonia, Moldova, Bulgaria and Iran); the deinotheriid, Deinotherium gigantissimum (shared with Moldova); the hyaenid, Ictitherium pannonicum (shared with the

Ukraine and Moldova) and the cercopithecid, Mesopithecus pentelicus (shared with

Afghanistan, Greece, Macedonia and Bulgaria). Like MN11, the Italian faunas are completely endemic at the species level and in fact, remain largely unchanged from the preceding interval. A cluster of localities known from Saudi Arabia is supported by several endemic species, including the bovids, Pachyportax latidens and Tragoportax cyrenaicus; the equid, Hipparion abudhabiense; the elephantid syrticus; and the mustelid, Plesiogulo praecocidens. The African localities do not cluster due to an absence of shared fauna among localities.

The MN12 large mammal genera corroborate the patterns found in the species-level analysis (Figure 3.9b). Localities towards the eastern boundary of the study area tend to cluster into smaller groupings including Turkish, Turkish-Moldovan, Turkish-Iranian,

Ukrainian-Moldovan and Bulgarian-Greek clusters. Together, these localities share a large number of common genera including bovids, Miotragocerus, Oioceros, Pachytragus (the latter shared with Kenya), Protragelaphus, Pseudotragus, Prostrepsiceros (the latter shared with Saudi Arabia); cervid, Pliocervus; giraffids, Helladotherium (the latter shared with

Hungary), Samotherium and Palaeotragus (the latter shared with Kenya); rhinos,

124

Ceratotherium (the latter shared with Kenya), Chilotherium; chalicotheriid, Ancylotherium; equid, Cremohipparion; and gomphotheriid, Choerolophodon. Localities in Spain cluster together based shared genera, including the bovid, Hispanodorcas; the cervid, Turiacemas; the giraffid, Birgerbohlinia; the mustelids, Baranogale, Enhydriodon (shared with France),

Plesiogulo (shared with Saudi Arabia) and Sivaonyx; and the canid, Canis. Although

Spanish and African localities are observed to cluster on the MN12 large mammal dendrogram, this grouping is based on shared genera that are cosmopolitan in nature, rather than genera that are shared exclusively between regions. Similarly, Spain clusters with

France on the basis of cosmopolitan genera, in addition to the previously mentioned mustelid, Enhydriodon. The Hungarian localities also cluster on the basis of shared taxa that are largely cosmopolitan. The Italian large mammal localities included in the genus-level analysis do not cluster due to the absence of shared genera with surrounding regions. The

African and Arabian localities share several genera in common, including the suid,

Nyanzachoerus; the hippo, Hexaprotodon; and the elaphantid, Stegotetrabelodon. The

African localities also share taxa in common with localities from Eastern Europe, as previously mentioned. Saudi Arabia shares a number of taxa with Eastern Europe and

Western Asia, including the bovid, Prostrepsiceros; and the suid, Propotamochoerus (shared also with Spain). Both regions share a number of cosmopolitan Eurasian genera, including the bovid, Tragoportax; the equid, Hipparion; and the deinotheriid, Deinotherium.

The small mammals in MN 12 support small clusters of Turkish localities (Figure

3.9c). Although these localities cluster on the basis of only one shared species, the hystricid,

Hystrix primigenia, several murid, glirid, eomyid and pteromyid rodents, as well as soricid and erinaceid insectivores are endemic to this region, but are known only from individual

125 localities. In contrast to the large mammal species for this interval, the Turkish small mammals share no species in common with surrounding localities in Greece, Moldova, or

Afghanistan. Instead, the small mammals known from these latter regions are also endemic, with the exception of the murid rodent, Occitanomys adroveri, and the ochotonid lagomorph,

Prolagus sorbinii (known from Moldova, France and Spain). The clusters of Spanish localities are very well supported by small mammal species including murid (Hispanomys adroveri, H. freudenthali, Huerzelerimys turoliensis, Kowalskia occidentalis), sciurid

(Atlantoxerus adroveri and Spermophilinus turoliensis), and castorid rodents (Dipoides problematicus); an ochotonid lagomorph (Prolagus crusafonti) and a soricid insectivore

(Blarinella dubia). Spain also shares several small mammals with France, including glirid

(Eliomys truci) and murid rodents (Parapodemus barbarae and Ruscinomys schaubi), in addition to the talpid insectivore, Desmanella crusafonti. Although the small mammal species from Africa are all endemic to the region, they are not shared among the African localities, and instead are known exclusively from single localities.

At the genus level, the Turkish localities cluster on the basis of a single shared genus,

Byzantinia (Figure 3.9d). However, like the species-level analysis, a much greater number of endemic species known from individual localities is known from this region. Turkish localities also cluster with those in Greece, on the basis of the shared murid, Schizogalerix, in addition to a number of more widely ranging genera. The localities from Moldova cluster on the basis of few genera (the castorid, Castor, and the ochotonid lagomorph, Proochotona), in addition to sharing more widely-ranging taxa. The cluster of Spanish localities is supported by castorid (Dipoides), glirid (Muscardinus), murid (Hispanomys, Huerzelerimys,

Kowalskia, Ruscinomys) and sciurid rodents (Atlantoxerus, Heteroxerus, Spermophilinus), as

126 well as soricid insectivores (Blarinella, Paenelimnoecus). However, this region shares small mammal genera with France (the glirid rodent, Eliomys, and the talpid insectivore, Talpa), in addition to sharing more cosmopolitan taxa with localities in Eastern Europe, Western Asia and Africa. The small mammal genera from Italy remain almost entirely endemic, with the exception of the murid, Parapodemus. The locality from Saudi Arabia shares the murids,

Myocricetodon, with Morocco, and Parapelomys with Afghanistan, as well as having its own endemic taxa. Similarly, the African localities (Morocco) share the hystricid, Hystrix, with

Turkey and Afghanistan, and the more widely ranging erinaceid insectivore, Galerix, with localities in Eurasia. The leporid lagomorph, Alilepus, is shared among localities in Kenya,

Turkey and Spain. Eurasian small mammals incompletely sampled for this interval include the murids, Myocricetodon and Paraethomys, and the glirid, Microdyromys, all known only from Morocco in MN12.

127

Figure 3.9a: MN12 large mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Samburu KE Lothagam KE Hamra SAr Shuwaihat SAr 5 Baccinello V2 IT Concud B SP Aljezar B SP Valdecebro 5 SP Mt Luberon FR

10 Concud SP Cerrodela Garita SP Los Mansuetos SP Chomateres GR Casadel Acero SP 15 Crevillente SP Concudbarranco SP Belka UK Novaja Emetovka UK Gura-Galben MO 20 Chimishlija MO Tudorovo MO Taraklia MO Chobruchi MO Polgardi HU 25 Kemiklitepe TU Middle Maragheh IR Upper Maragheh IR Akgedik-Bayir TU Salinpaşalar TU 30 Salinpaşalar TU Serefköy TU Sandikli Kinik TU Duzyayla TU Akkasdagi TU 35 Titov Veles MA Kalimanci BU Hadjidimovo BU Pikermi GR Molayan AF 40 Baltavar HU Los Aljezares SP Çobanpinar TU Sinap 26 TU Sinap 33 TU 45 Vathylakkos GR Masadadel Valle SP cc = 0.82

128 Figure 3.9b: MN12 large mammal genera dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Baccinello V2 IT Casadel Acero SP Masadadel Valle SP

Samburu KE 5 Jebel Dhannah SAr Hamra SAr Shuwaihat SAr Valdecebro 5 SP Cerrodela Garita SP 10 Los Mansuetos SP Polgardi HU Baltavar HU Belka UK

Taraklia MO 15 Novaja Emetovka UK Kemiklitepe TU Akkasdagi TU Upper Maragheh IR Middle Maragheh IR 20 Salinpaşalar TU

Kalimanci BU Pikermi GR Hadjidimovo BU Chobruchi MO 25 Titov Veles MA Tudorovo MO Akgedik-Bayir TU Serefköy TU Sandikli Kinik TU

Duzyayla TU 30 Molayan AF Salinpaşalar TU Crevillente SP Concudbarranco SP

35 Mt Luberon FR Chomateres GR

Vathylakkos GR Sinap 26 TU Gura-Galben MO 40 Çobanpinar TU Lothagam KE Sinap 33 TU

Los Aljezares SP cc = 0.84 Aljezar B SP 45

129 Figure 3.9c: MN12 small mammal species dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Lothagam KE

Khendekel Ouaich MOR

3 Afoud 6 MOR

Molayan AF

Duzyayla TU

Ano Metochi 3 GR 6

Cerrodela Garita SP

Concud 8 SP

9 Los Mansuetos SP

Casadel Acero SP

Concudbarranco SP

12 Concud B SP

Valdecebro 5 SP

Los Aljezares SP

15 Regajo 3 & 4 SP

Villalba Baja 2B/2C SP

Masadadel Valle SP

Tortajada C & D SP 18 Masadadel Valle SP

Masada Ruea 3 & 4 SP

Aljezar B SP 21 Mt. Luberon FR

Crevillente SP

24 Tortajada A SP

Taraklia MO

Chimishlija MO

27 Polgardi HU

Çobanpinar TU

Gelibolu Bayirköy TU cc = 0.97

130 Figure 3.9d: MN12 small mammal genera dendrogram similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Afoud 6 MOR

Baccinello IT

Duzyayla TU

4 Ano Metochi 3 GR

Concudbarranco SP Cerrodela Garita SP

Tortajada C & D SP

8 Crevillente SP

Los Aljezares SP

Valdecebro 5 SP

Tortajada A SP

12 Masadadel Valle SP Masadadel Valle SP

Los Mansuetos SP

Aljezar B SP

16 Villalba Baja 2B/2C SP

Regajo 3 & 4 SP Concud SP

Çobanpinar TU

Kemiklitepe TU 20 Gelibolu Bayirköy TU

Crevillente SP

Casadel Acero SP

24 Concud B SP

Masada Ruea 3 & 4 SP Mt. Luberon FR

Polgardi HU

28 Shuwaihat SAr

Khendekel Ouaich MOR Taraklia MO

Chimishlija MO

32 Molayan AF

Lothagam KE cc = 0.95

131

MN13 Although the MN13 large mammal species support a cluster of Turkish and Greek localities, this grouping is supported by only three taxa: the rhinos, Ceratotherium neumayri and Chilotherium schlosseri, and the suid, Microstonyx erymanthius (Figure 3.10a).

However it should be noted that the sole locality representing the Turkish faunas during this interval is species-poor. The Greek localities also cluster on the basis of few shared species, including the equid, Cremohipparion matthewi, and the gomphotheriid, Choerolophodon pentelici, however these localities also contain a large number of species endemic to the region and known only from individual localities. Spanish faunas are well represented for this interval and cluster on the basis of cervid (Pliocervus turolensis), bovid (Hispanodorcas torrubiae), hippo (Hexaprotodon crusafonti), equid (Hipparion gromovae, Hippotherium primigenium), gomphotheriid (Tetralophodon longirostris), hyaenid (Thalassictis hipparionum), and mustelid species (Sivaonyx lehmani), in addition to a number of taxa known only from individual localities. Spain shares a number of taxa in common with localities in Italy, including mammutid (Zygolophodon turicensis), hippo (Hexaprotodon pantanellii), and suid species (Propotamochoerus provincialis, also known from Hungary).

Interestingly, Spain also shares the mustelid, Plesiogulo monspessulanus, with the South

African locality of Langebaanweg 3. The Central European region is represented by a single, species-poor locality in Hungary. This locality lacks any species endemic to it, yet interestingly shares the giraffid, Helladotherium duvernoyi, with the Tunisian locality of

Douaria. Although the Italian localities group only on the basis of one shared taxon, the rhino, Stephanorhinus megarhinus, they also continue to contain a large number of species endemic to the region. However, Italy also shares taxa with Greece (the aardvark,

Orycteropus gaudryi; cercopithecid, Mesopithecus pentelicus) and Hungary (the

132 cercopithecid, Mesopithecus pentelicus). The African faunas remain almost entirely endemic to the continent, with the exception of the mustelid and giraffid mentioned previously.

In the genus-level analysis, the MN13 Turkish localities fail to cluster due to the absence of shared genera, with the exception of the equid, Hipparion, which is cosmopolitan

(Figure 3.10b). Moreover, at this level of analysis, the Turkish localities also lack any genera that are endemic to the region. Instead, these localities share genera with Greece (the giraffid, Samotherium; suid, Microstonyx; bovid, Tragoportax; and rhinos, Ceratotherium and Chilotherium), Hungary (the giraffid, Helladotherium), and Spain (the bovid,

Tragoportax; percrocutid, Percrocuta; and gomphotheriid, Tetralophodon).

Intercontinentally, Turkey shares genera with Tunisia (the giraffid, Helladotherium), Kenya

(the bovid, Tragoportax; and rhino, Ceratotherium), South Africa, Ethiopia, Tanzania,

Uganda (the rhino, Ceratotherium), and Libya (the giraffid, Samotherium; percrocutid,

Percrocuta). Shared large mammal genera in Greece support a cluster in this region based on the mustelids, Promephitis and Promeles, in addition to the gomphotheriid,

Choerolophodon. This region also contains endemic genera known from individual localities.

Greece shares large mammal genera with Italy and Hungary (the cercopithecid,

Mesopithecus) and the Ukraine (the ursid, Indarctos) and intercontinentally with Algeria

(Indarctos, also in the Ukraine), South Africa (the giraffid, Palaeotragus; and hyaenid,

Chasmaporthetes), Kenya (Palaeotragus, and the aardvark, Orycteropus), the Congo and

Zaire (the tragulid, Dorcatherium), and Chad (Orycteropus). The Spanish localities cluster supported by the bovid, Parabos; the equid, Hippotherium; the pliohyracid, Pliohyrax; and the canid, Nyctereutes, in addition to endemic genera known only from individual localities.

Like Greece, the Spanish localities share taxa within Eurasia and intercontinentally. These

133 genera include the bovid, Gazella (Greece, Italy, Kenya, South Africa, Libya); rhino,

Aceratherium (Congo/Zaire); hippo, Hexaprotodon (Italy, Uganda, Libya, Chad, Tanzania,

Kenya, Ethiopia); equid, Cremohipparion (Libya, Greece); gomphotheriid, (Kenya,

Chad, South Africa, Tanzania, Ethiopia, Uganda); ursid, Agriotherium (Uganda, South

Africa, Libya); the canid, Vulpes (South Africa) and the felid, Dinofelis (Kenya, South

Africa). The locality from Hungary again contains no endemic fauna. Although Italy maintains some endemic genera during this interval mostly known from individual localities, it shares a number of widely-ranging taxa with Europe and Africa. With the latter, Italy shares the suid, Nyanzachoerus (Kenya, Chad, Congo/Zaire, Tanzania, Ethiopia, Uganda,

Libya); the bovids, Kobus (Kenya, Chad, South Africa, Tanzania, Egypt, Uganda, Ethiopia) and Miotragocerus (Ethiopia, Libya); the felid, Machairodus (Chad, South Africa, Tanzania,

Libya); and the viverrid, Viverra (South Africa, Libya). Although the African localities share many genera with Eurasia, there remain a considerable number of endemic taxa known from the individual regions.

No small mammal localities from Turkey were included in the species-level analysis due to a lack of adequate samples from this region during MN13 (Figure 3.10c). The cluster of Greek localities is supported by glirid (Myomimus maritsensis), murid (Apodemus gorafensis), and pteromyid rodents (Pliopetaurista dehneli), as well as soricid insectivore species (Asoriculus gibberodon, Deinsdorfia kerkhoffi). This cluster is also supported by a significant number of endemic species known only from individual localities, but shares a number of taxa with regions to the west. Interestingly, the Greek localities share the ochotonid lagomorph, Prolagus michauxi, both with Spain, as well as intercontinentally with the Afoud localities in Morocco. The Spanish localities dominate the those sampled for this

134 interval and are extremely well supported by shared murid (Apocricetus alberti, Blancomys sanzi, Castillomys crusafonti, Castromys inflatus, Huerzelerimys turoliensis, Hispanomys adroveri, Occitanomys alcalai, Paraethomys meini – latter known from Italy, Ruscinomys lasallei, R. schaubi, Stephanomys ramblensis –latter known from France), glirid (Eliomys intermedius, E. truci) and castorid rodents (Dipoides problematicus – known from Italy), soricid (Blarinella dubia) and talpid insectivores (Archaeodesmana adroveri, Desmanella crusafonti –latter known from France), and ochotonid lagomorphs (Prolagus sorbinii, P. crusafonti –latter known from France). The Spanish small mammal faunas also document a large number of endemic species that are known only from individual localities. Although the Italian localities cluster together, this grouping is based mostly on endemic species known from individual localities, rather than species shared among localities. These localities share few species with the rest of Eurasia and none with Africa. The African small mammal faunas are represented by species in Tanzania and Morocco. These regions do not share any common species, nor do they share any species with Eurasia, with the exception of the previously mentioned ochotonid lagomorph, Prolagus michauxi, known from Morocco,

Greece and Spain.

The cluster of MN13 Turkish localities is solely supported by the shared presence of the murid, Pliospalax (Figure 3.10d). These localities share more genera in common with

Greece, than they do among themselves. Genera supporting the broader cluster of Turkish and Greek localities include the murids, Allocricetus and Micromys; the gerbillid,

Pseudomeriones (also known from Spain); the sciurid, Tamias; and the erinaceid insectivore,

Schizogalerix. The cluster of Greek localities is supported by the shared presence of the pteromyid rodents, Hylopetes (also known from Italy) and Pliopetaurista, as well as the

135 soricid insectivores, Asoriculus and Deinsdorfia. Endemic genera known from individual localities are also known for this region. The large cluster of Spanish localities is supported by the shared murids, Castromys, Hispanomys, Parapodemus; and sciurid, Heteroxerus; in addition to soricid insectivores, Blarinella and Paenelimnoecus. Spain also shares a large number of genera in common with regions to the East, as well as intercontinentally with

Morocco and Libya. These genera include murid, Apocricetus (shared with Morocco),

Castillomys (shared with Turkey and Morocco), Huerzelerimys (shared with France),

Myocricetodon (shared with Morocco), Paraethomys (shared with Italy and Morocco),

Protatera (shared with Libya and Morocco), Ruscinomys (shared with Italy), Stephanomys

(shared with France and Italy); glirid, Eliomys (shared with Italy); and sciurid rodents,

Atlantoxerus (shared with Italy and Morocco); as well as the erinaceid insectivore, Galerix

(shared with France and Italy) and the ochotonid lagomorph, Prolagus, which is cosmopolitan in Eurasia and also known from Morocco. Although the Moroccan and Libyan localities share genera with Eurasia, the African localities as a whole are also comprised of genera that are regionally endemic.

136

Figure 3.10a: MN13 large mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Samos Main GR Amasya 2 TU Sahabi 14 LI Taskinpasa 1 TU 5 El Arquillo 1 SP Arquillo 1 SP Las Casiones SP Milagros SP Villastar SP 10 Rambla de V SP La Gloria 5 SP Venta del Moro SP Arenas del Rey SP Casino IT 15 Hatvan HU Gravitelli IT Maramena GR Dytiko 1-3 GR La Alberca SP

20 Anajev UK Menacer AL Douaria TUN

Lothagam 2 KE Lothagam KE 25 Inolelo 1 TA Manonga 1 TA Lukeino 3-6 KE Lukeino KE Mpesida KE 30 Albertine 1 UG Nkondo UG Toros-Menalla CH Awash 5 ET Manonga 2 TA 35 Awash 3 ET Langebaanweg SA Langebaanweg SA Albertine 14 CO/ZA Brisighella IT 40 Baccinello IT Gargano IT Layna SP cc = 0.92

137 Figure 3.10b: MN13 large mammal genera dendrogram

similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Gargano IT Brisighella IT Baccinello IT Casino IT 5 Hatvan HU Douaria TUN Gravitelli IT Layna SP Lothagam 2 KE 10 Lothagam 5 KE Manonga 2 TA Awash 5 ET Toros-Menalla CH Nkondo UG 15 Lukeino KE Inolelo 1 TA Manonga 1 TA Awash 3 ET Mpesida KE

20 Lukeino 3-6 KE Langebaanweg SA Langebaanweg SA Sahabi 14 LI Wadi Natrun EG Menacer AL 25 La Alberca SP Amasya 2 TU Librilla SP Villastar SP 30 Rambla de V SP La Gloria 5 SP Taskinpasa 1 TU El Arquillo 1 SP Arquillo 1 SP 35 Milagros SP Samos Main GR Venta del Moro SP Maramena GR Dytiko 1-3 GR 40 Las Casiones SP Arenas del Rey SP Albertine 1 UG Albertine 14 CO/ZA Anajev UK 45 Almenara-Casablanca SP cc = 0.83

138 Figure 3.10c: MN13 small mammal species dendrogram similarity 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Afoud 5MOR

Inolelo 1 TA Manonga 1 TA

4 Lissasfa MOR

Maramena GR

Silata GR

Monasteri GR

8 Afoud 1 MOR Afoud 2 MOR

Afoud 8 MOR

Baccinello IT

12 Layna SP

Brisighella IT La Gloria 1 SP

Crevillente SP

16 Crevillente SP

El Arquillo 1 SP

Villastar SP

La Gloria 5 SP

20 Rambla de V SP

Las Casiones SP

Las Casiones SP

Celadas 2

Arquillo 1 SP 24

El Arquillo 4 SP

La Gloria 8 SP

Almenara-Casablanca SP 28 Arenas del Rey SP

Librilla SP

Venta del Moro SP

La Alberca SP Lissieu FR 32 Bacochas 1 SP

Gargano IT cc = 0.93

Figure 3.10d: MN13 small mammal genera dendrogram 139

similarity

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Inolelo 1 TA Manonga 1 TA Lukeino 3-6 KE Langebaanweg SA Lissasfa MOR 5 Sahabi 14 LI

Baccinello IT Gargano IT Maramena GR 10 Lissieu FR

Layna SP Brisighella IT Librilla SP

La Alberca SP 15 Afoud 8 MOR Almenara-Casablanca SP Venta del Moro SP El Arquillo 1 SP Arquillo 1 SP Villastar SP 20 La Gloria 5 SP

Rambla de V SP La Gloria 88 Las Casiones SP 25 Celadas 2 SP La Gloria 1 SP

El Arquillo 4 SP

Crevillente SP Crevillente SP 30 Bacochas 1 SP Silata GR

Monasteri GR Kangal 1 TU Süleimanli TU

35 Amasya 2 TU Velona IT

Afoud 1 MOR Afoud 2 MOR

Afoud 5 MOR 40 Las Casiones SP Arenas del Rey SP cc = 0.86

Table 3.2: Summary of provincial composition (species-level) Eastern Europe/West Asia Central Europe Spain MN5 TU GR Dorcatherium naui Lagomeryx meyeri Caprotragoides stehlini Choerolophodon chioticus D. guntianum Eotragus artenensis Hypsodontus pronaticornis Georgiomeryx georgialasi D. peneckei Bunolistriodon adelli Tethytragus koehlerae Sanitherium schlagintweiti D. vindobonense Palaeomeryx garsonnini Turcocerus gracilis Lophocyon paraskevaidisi Dicrocerus elegans Triceromeryx pachecoi Bunolistriodon meidamon Heteroprox larteti Alicornops simorrensis Listriodon splendens Eotragus clavatus Martes laevidens Schizochoerus anatoliensis Palaeomeryx bojani Gomphotherium olisiponensis Taucanamo inonuensis P. kaupi Zygolophodon pyrenaicus Beliajevina grimmi Lartetotherium sansaniensis Procoelodonta tekkayai Prosantorhinus germanicus Democricetodon darocensis Protictitherium intermedium Ischyrictis zibethoides Fahlbuschia darocensis Anatolictis laevicaninus Martes munki Armantomys aragonensis Ischyrictis anatolicus Trocharion albanense Glirudinus gracilis Heteroprox anatoliensis Plithocyon stehlini Microdyromys monspeliensis Giraffokeryx anatoliensis Pseudaelurus romieviensis Miodyromys biradiculus Gomphotherium pasalarensis Pseudocyon steinheimensis Heteroxerus grivensis Griphopithecus alpani Micromeryx flourensianus H. rubricati Orycteropus seni Taucanamo sansaniense Mioechinus butleri TU-RG Pliopithecus antiquus Ischyrictis anatolicus GE-CZ-PO Byzantinia pasalarensis Anomalomys minor Cricetodon candirensis Neocometes similes Democricetodon brevis Glirudinus undosus Megacricetodon andrewsi Miodyromys hamadryas Pliospalax marmarensis Plesiosorex germanicus Glirulus daamsi GE-FR-SW Peridyromys lavocati Eumyarion bifidus Albanensia sansaniensis Megacricetodon germanicus Forsythia gaudryi Miodyromys aegercii Alloptox anatoliensis Bransatoglis cadeoti Schizogalerix pasalarensis Keramidomys carpathicus Desmanodon minor Proscapanus sansaniensis

140

Eastern Europe/West Asia Central Europe Spain MN6 TU GE-AU Cricetodon jotae Bunolistriodon latidens Muscardinus sansaniensis Democricetodon darocensis Hispanotherium grimmi Prodryomys satus Heteroxerus grivensis Lanthanotherium sansaniensis Hypsodontus pronaticornis Limnoecus truyolsi

Megacricetodon crusafonti Protoryx enanus SL Taucanamo inonuensis Aureliachoerus aurelianensis Tethytragus koehlerae Deinotherium laevius Turcocerus gracilis Dorcatherium vindobonense Zygolophodon tapiroides Eotragus haplodon Epipliopithecus vindobonensis Lartetictis dubia Trocharion albanense Ursavus brevirhinus GE-FR-SW Eotragus clavatus Euprox furcatus Micromeryx flourensianus Lartetotherium sansaniensis Pliopithecus antiquus Deinotherium giganteum AU Tethytragus langai Plesiopliopithecus lockeri

SL-HU Anomalomys gaudryi Eumyarion latior Muscardinus sansaniensis Plesiodimylus chantrei SL-RO Muscardinus sansaniensis SL-GE Eumyarion weinfurteri Keramidomys carpathicus

141

Eastern Europe/West Asia Central Europe Spain MN7/8 TU SL-FR Miotragocerus monacensis Byzantinia bayraktepensis Cricetodon sansaniensis Protragocerus chantrei B. eskihisarensis Democricetodon gaillardi Albanohyus castellensis Myocricetodon eskihisarensis Eumyarion latior Parachleuastochoerus Desmanella cingulata Megacricetodon schaubi huenermanni Desmanodon major Neocometes brunonis Propotamochoerus Mioechinus tobieni Bransatoglis astaracensis palaeochoerus Schizogalerix anatolica Muscardinus sansaniensis Palaeomeles pachecoi Plesiodimylus chantrei Semigenetta grandis, Lanthanotherium sansaniensis Trocharion albanense Talpa minuta Protictitherium gaillardi SL-SW Sansanosmilus jourdani Democricetodon gaillardi Amphicyon major Plesiodimylus chantrei Hoploaceratherium FR-SW-GE tetradactylum Taucanamo grandaevum Lartetotherium sansaniensis Semigenetta sansaniensis Deinotherium giganteum Trochotherium cyamoides Leptoplesictis aurelianensis Cricetodon lavocati Dryopithecus fontani Democricetodon crusafonti D. nemoralis HU-RO Eumyarion leemanni Eumyarion medius Hispanomys dispectus Glirulus lissiensis Megacricetodon debruijni FR-SW-GE-HU M. ibericus Eomyops oppligeri Myomimus dehmi Microdyromys complicates Muscardinus hispanicus Miodyromys aegercii M. vallesiensis M. hamadryas Tempestia hartenbergeri Democricetodon brevis Alloscapanus lehmani Crusafontina excultus Amphechinus golpeae Domninoides santafei Talpa minuta T. vallesensis Dinosorex sansaniensis

142

Eastern Europe/West Asia Central Europe Spain MN9 TU Anomalomys gaudryi Deinotherium laevius Cormohipparion sinapensis Democricetodon freisingensis Protragocerus chantrei Palaeotragus roueni Kowalskia fahlbuschi Euprox minimus Ceratotherium neumayri Microtocricetus molassicus Parachleuastochoerus Ictitherium intuberculatum Albanensia grimmi crusafonti TU-RG Glis minor Dorcatherium jourdani Palaeotragus coelophrys Glirulus lissiensis Hipparion catalaunicum Chalicotherium grande Byzantinia bayraktepensis Lartetotherium sansaniensis B. dardanellensis Limnonyx sinerizi B. nikosi Thaumastocyon dirus B. ozansoyi Protictitherium gaillardi Heramys anatolicus Indarctos vireti Myocricetodon eskihisarensis Progonomys minus Cricetodon lavocati Schizogalerix anatolica Cricetulodon sabadellensis S. intermedia Democricetodon crusafonti S. sinapensis D. sulcatus Bellatonoides eroli Eumyarion leemanni Hispanomys dispectus H. peralensis H. thaleri Kowalskia seseae Megacricetodon debruijni M. ibericus Progonomys hispanicus Eliomys truci Miodyromys hamadryas Tempestia hartenbergeri Heteroxerus rubricate Keramidomys carpathicus Postpalerinaceus vireti Miopetaurista crusafonti Talpa vallesensis

143

Eastern Europe/West Asia Central Europe Spain MN10 TU GR FR Crusafontina excultus Choerolophodon anatolicus Ouranopithecus macedoniensis Keramidomys pertesunatoi Eurolagus fontannesi Acerorhinus zernowi Oioceros praecursor Myoglis meini Prolagus crusafonti Palaeotragus coelophrys Decennatherium macedoniae Paraglirulus werenfelsi Micromeryx flourensianus Hispanomys mediterraneus Decennatherium pachecoi TU-GR Kowalskia ambarrensis Ceratotherium neumayri Rotundomys montisrotundi Cricetulodon hartenbergeri Chilotherium kiliasi Albanensia grimmi Hispanomys peralensis Bohlinia attica Heteroxerus grivensis Hispanomys thaleri Helladotherium duvernoyi Archaeodesmana vinea Progonomys hispanicus Tragoportax amalthea FR-AU Tempestia hartenbergeri Choerolophodon pentelici Cricetulodon bugesiensis Eucastor adroveri Anourosorex kormosi Galerix socialis Crusafontina kormosi Crusafontina excultus Paenelimnoecus repenningi Crusafontina endemica Petenyia dubia Talpa vallesensis Schizogalerix zapfei Eurolagus fontannesi Talpa gilothi Eozapus intermedius Graphiurops austriacus Kowalskia skofleki Muscardinus austriacus Parapodemus lugdunensis Pliopetaurista bressana Prospalax petteri Blackia miocaenica Vasseuromys pannonicus AU Glirulus lissiensis Epimeriones austriacus MN11 TU GR GE-HU Hippotherium primigenium Ictitherium hipparionum Cremohipparion Cervavitulus mimus Birgerbohlinia schaubi Chilotherium habereri macedonicum Allohyaena kadici Lucentia pierensis Gazella gaudryi C. proboscideum Plioviverrops guerini Helicotragus rotundicornis Hipparion dietrichi FR Baranogale adroveri Dihoplus pikermiensis Eomyops catalaunicus Prostrepsiceros Graphiurops austriacus Hispanomys freudenthali rotundicornis Muscardinus austriacus Kowalskia occidentalis GR Kowalskia skofleki

144

Eastern Europe/West Asia Central Europe Spain Ancylotherium Prospalax petteri pentelicum Rotundomys bressanus Pachytragus crassicornis Blackia miocaenica P. laticeps Pliopetaurista bressana Pseudotragus capricornis Lanthanotherium sanmigueli P. parvidens Crusafontina kormosi Pliocervus pentelici Petenyia dubia MN11 Pliohyrax kruppii Archaeodesmana vinea Plioviverrops orbignyi Talpa gilothi Metailurus parvulus Simocyon primigenius TU-GR Dorcatherium puyhauberti Criotherium argalioides Gazella capricornis Italy Oioceros rothi Etruria viallii O. wegneri Maremmia hauptii M. Tragoportax amalthea lorenzi Tyrrhenotragus Ceratotherium neumayri gracillimus Chilotherium samium Eumaiochoerus etruscus Choerolophodon pentelici Mustela majori Pliohyrax graecus Paludolutra campanii P. maremmana Byzantinia pikermiensis Tyrrhenolutra helbingi Karnimata provocator Indarctos anthracitis Hystrix primigenia bambolii TU-GR-IR-IQ Palaeotragus coelophrys Nisidorcas planicornis Ictitherium robustum Parataxidea maraghana MN12 TU Hispanodorcas torrubiae Gazella gaudryi Turiacemas concudensis Sinotragus occidentalis Birgerbohlinia schaubi Hipparion matthewi Hippotherium primigenium H. mediterraneum Hipparion concudense Ictitherium hipparionum Baranogale adroveri I. robustum Enhydriodon lluecai TU Martes basilii Hystrix primigenia Sivaonyx lehmani

145

Eastern Europe/West Asia Central Europe Spain S. lluecai TU-GR Metailurus major Hyaenotherium wongi Plioviverrops guerini Pliohyrax graecus Canis cipio Prostrepsiceros rotundicornis Hipparion dietrichi Hispanomys adroveri Tetralophodon atticus H. freudenthali TU-IR Huerzelerimys turoliensis Pachytragus laticeps Kowalskia occidentalis Cremohipparion matthewi Atlantoxerus adroveri Machairodus giganteus Spermophilinus turoliensis MN12 TU-MAC Dipoides problematicus Nisidorcas planicornis Prolagus crusafonti TU-MOL Blarinella dubia Pachytragus crassicornis Palaeoryx majori Chilotherium schlosseri Orycteropus gaudryi Italy TU-BU-GR-IR Etruria viallii Microstonyx erymanthius Maremmia hauptii M. TU-GR-MOL lorenzi Tyrrhenotragus Palaeoreas lindermayeri gracillimus TU-BU-GR-MAC-MOL-UK Eumaiochoerus etruscus Palaeotragus roueni Mustela majori TU-BU-GR-MOL-IR Paludolutra campanii Ceratotherium neumayri P. maremmana TU-BU-MOL Tyrrhenolutra helbingi Stephanorhinus pikermiensis Indarctos anthracitis TU-GR-BU-MOL Oreopithecus bambolii Cremohipparion mediterraneum TU-MAC-MOL-UK-IR Ictitherium viverrinum TU-GR-MAC-UK-IR Choerolophodon pentelici

MN13 GR Cremohipparion matthewi Pliocervus turolensis Hispanodorcas torrubiae Choerolophodon pentelici Myomimus maritsensis Hexaprotodon crusafonti

146

Eastern Europe/West Asia Central Europe Spain Apodemus gorafensis Hipparion gromovae , Pliopetaurista dehneli Hippotherium primigenium Asoriculus gibberodon Tetralophodon longirostris Deinsdorfia kerkhoffi Thalassictis hipparionum Sivaonyx lehmani TU-GR Ceratotherium neumayri Apocricetus alberti Blancomys Chilotherium schlosseri sanzi Castillomys crusafonti Microstonyx erymanthius Castromys inflatus Huerzelerimys turoliensis Hispanomys adroveri Occitanomys alcalai Ruscinomys lasallei R. schaubi Eliomys intermedius MN13 E. truci Blarinella dubia Archaeodesmana adroveri Prolagus sorbinii

Italy Stephanorhinus megarhinus

147 148

Discussion

Resolution of analysis

Genus vs species level analysis

Although previous studies include analysis at both the genus and species level, results are reported for the genus level analysis only. Bernor (1978, 1983) found that clustering at the genus level was more strongly supported and produced logical results, while clusters at the species level tended to be weakly supported and often produced different dendrogram topologies than the genus-level analysis. Bernor (1978) reasoned that analyses at the genus level were more useful for studies of faunal correlation and zoogeography at the inter- provincial scale, since paleospecies were subject to personal interpretation and had shorter temporal and geographic ranges. Interestingly, the results of this study demonstrate the opposite; in almost all cases, the species level analyses provided similar, if not identical topologies to the genus-level analyses (differences were usually negligible). Furthermore, faunal provinces were more clearly defined at the species level. Although few species were found to support clusters during several intervals, genus-level analysis also often failed to provide adequate support for clusters of localities. At the species level, the sample of taxa available for analysis was decreased, due to the omission of indeterminate or cf. species designations (i.e., Dorcatherium indet., Dorcatherium cf. D. naui). However, species level analyses provided a precise definition of provincial boundaries, in addition to measure of species diversity. In this analysis, clustering at the genus level was more inclusive and often supported by taxa that were geographically widespread continentally and intercontinentally.

Although the results of this analysis at both taxonomic levels were broadly in agreement with

149 each other (and with previous studies, to varying degrees), the provincial signal from particular taxa can be very different depending on the level of analysis. For instance, while

Dryopithecus laietanus is endemic to Spain, the genus, Dryopithecus, is considered cosmopolitan, ranging from Spain to Georgia. Overall, the results here suggest that the choice of genus or species level analysis depends on the specific goals of the study.

Inclusion of small mammal taxa

Previous studies of provinciality either excluded small mammals altogether (Bernor

1978, 1983) or emphasized them to a lesser degree due to potential sampling bias, in comparison to their large mammal counterparts (Fortelius et al. 1996a). In this study, the small mammals supported clusters of localities better than the large mammals, particularly in

MN7/8 and MN9, but also in MN10 and MN11. These results suggest that when small mammals are adequately sampled, they may in fact provide a better indication of provinciality than their large mammal counterparts. The small mammals generally revealed similar, but more distinct distribution patterns than the large mammals, perhaps due to an increased sensitivity to paleoenvironmental barriers. Sensitivity to climatic variables was discussed in the previous chapter, but it is logical that physiogeographic barriers may also pose a more stringent restriction on the dispersal capabilities of most small mammal taxa.

The more rapid generation and turnover rates of small mammals may also provide clues as to why they appear to be more provincially informative. A quicker response to paleoenvironmental change may be more visible in these taxa than in large mammals.

However, Liow et al. (2008) recently documented higher origination and extinction rates

(and thus shorter temporal ranges) in large mammal genera and species and argue that the

150

“sleep-or-hide” behaviours common in small mammals (including hibernation, torpor, burrowing) serve as an environmental buffer contributing to higher mean survivorship.

Non-uniform geographic representation

Jernvall and Fortelius (2004) noted that the distribution of Eurasian Miocene fossil vertebrate localities was spatially and temporally non-uniform, affecting the reliability of geographic range estimates of fossil mammals. In the study described here, areas of Eastern

Europe are poorly represented in comparison to other regions of Europe and Western Asia.

For example, although a single Greek locality is considered during MN5, localities from this region are not well represented again until MN10 and onwards. While only further sampling will assist in better representation of faunas from these regions, this non-uniformity complicates provincial divisions, particularly between Central Europe and more eastern regions.

Terrestrial faunas within the Pannonian Basin are one example of this. Surrounded by the Alps, Dinarides and Carpathians, the Pannonian Basin of Central Europe is a relatively closed system and corridors for faunal interchange during the middle and late Miocene were limited by topography and fluctuating paleoenvironments. In a previous study, Nargolwalla

(2006) found that although the Pannonian Basin represents a fairly restricted geographic area, its constituent faunas are not more similar to each other than to faunas outside of the basin system, contrary to expectations. Overall, the Pannonian Basin faunas did not cluster as expected, with few localities demonstrating any similarity (i.e., in MN7/8). The bioprovinces recognized in MN6, MN9 and MN12 in this study were consistent with previous research, with the Austrian, Slovakian and Hungarian localities more similar to the Central and

151

Western Europe province. MN7/8 Romanian localities were more similar to Central and

Western Europe than the South Eastern province of Fortelius et al. (1996a). However,

during this interval, the Pannonian lake had regressed slightly from the Vienna Basin, potentially introducing a corridor for faunal interchange. Similarly, faunas from the MN11

Hungarian locality of Csakvar were most similar to those from the Ukraine, perhaps due to the presence of an interchange corridor in the region of the Eastern Carpathians, which had not yet attained their full elevation and were still experiencing lowland conditions (Popov et

al. 2004). Overall, this may explain the similarity between Pannonian Basin localities and

those in Central and Western Europe and South Eastern province in MN7/8 and MN11,

respectively, and demonstrates that although the Pannonian Basin faunas are broadly

consistent with the bioprovincial concept outlined by Bernor (1983) and Fortelius et al.

(1996a), there are several exceptions that further sampling of taxa could clarify. It is

important to note that these conclusions are influenced by the nature of the faunal data in the

Pannonian Basin, where uneven distribution of terrestrial vertebrate localities in this region is

related to a number of factors. The intensity of research has varied over the region, and

suitable outcrops are also unevenly distributed. Further sampling could support the previous

findings, or indicate stronger provincial affinities with either regions to the west or east.

Faunal provinces in comparison to previous studies

Bernor (1983) recognized a Western and Southern European Province, comprised of

late Miocene large mammal faunas from Spain, France and Italy. Fortelius et al. (1996a)

also recognized a similar province, termed “Western Europe,” which also included Portugal.

At the species level, this analysis suggests that Spanish large and small mammal faunas are

152

distinct from those to the east from MN5 to MN9 (Figure 3.11a-c). In MN10, the species

level analysis of large mammals lacks resolution due to the considerable number of taxa endemic to the region, but known only from single localities (i.e., not shared). During this

interval though, the small mammal species demonstrate a distinct provincial association

within Spain. In MN11, Spanish large mammals again fail to cluster with localities from the

region, but in this case it is due to the presence of cosmopolitan taxa. In fact, only the MN12

large mammal genera and the MN12 and MN13 small mammal species demonstrate distinct

similarities to France and Italy specifically. At the genus level, the similarities that the

Spanish faunas have to other regions are on the basis of more widespread taxa, rather than to

those found exclusively in France or Italy.

Next, Bernor (1983) recognized an Eastern and Central European province, composed

of late Miocene large mammals from Switzerland, Germany, Austria, Hungary, the Czech

Republic, Slovakia and Poland. In contrast, Fortelius et al. (1996a) divide west Central

Europe (Germany and Switzerland), from Austria and the Black Sea region (Hungary,

Romania, Moldova and the Ukraine). Overall, the results of this study tend to more strongly

corroborate those of Bernor (1983), however faunas from Romania, albeit poorly sampled,

are included within the Central European group (Figure 3.11a-c). Although the regions

comprising the larger Central European province cluster into small groupings, these regions

tend to share more taxa in common with the Central European province than they do with

their smaller groupings. In addition, the French faunas, which both Bernor (1983) and

Fortelius et al. (1996a) group with those from Spain and Italy (and Portugal – Fortelius et al.

1996a), share more large and small mammals in common with the rest of Central Europe

than they do with Spain for most of the time period of interest, with the exception of MN12

153 and 13. The Central European province (and smaller groupings) remains distinct from MN5 to MN7/8, after which point, the influx of more cosmopolitan taxa coupled with a low incidence of shared endemic fauna in this region hinders provincial distinction. After MN9, sampling becomes an issue since Central European faunas are represented at few, patchily distributed localities. However after this point, localities from the region tend to have more similarities to localities from Eastern Europe, likely related to drying of the fore-Carpathian basin, which was flooded until at least the top of MN9 (Popov et al. 2004).

Mammal faunas are known from Italy from MN11 onwards. Although Bernor (1983) and Fortelius et al. (1996a) group the Italian faunas with those from Spain and France, this study suggests that these groupings are based on very few shared genera in MN11 and MN12

(Figure 3.11a-c). During these intervals, the Italian faunas are completely endemic at the species level and share more taxa amongst themselves than they do with either Spain or

France at the genus level. Furthermore, the geographic and geological evidence of insular environments in Italy (Moyà-Solà et al. 1999; Harrison & Rook 1997; Rook et al. 1999;

Rook et al. 1996) further support the recognition of these faunas belonging to their own province. In MN13, however, this previously isolated region is invaded by more cosmopolitan taxa. Italy shares faunas with Spain, Hungary and Greece at the species level, but also shares even more fauna intercontinentally with Africa at the genus level.

Bernor (1983) proposed a Romanian and Western Russian province, composed of localities in Romania, Moldova, the Ukraine and Georgia, as well as a Sub-Paratethyan province of localities in Turkey, Greece and Iran. Fortelius et al. (1996a) described similar

Balkan (Slovenia, Croatia, Bosnia, Serbia, Macedonia, Albania, Greece and Bulgaria) and

Anatolian provinces (Anatolia, Samos and Georgia). The results of my analysis generally

154

support those of Bernor (1983) (Figure 3.11a-c). Despite geographic proximity, Turkish and

Georgian faunas rarely cluster and only weakly when they do. Turkish faunas tended to be

more similar to themselves, however often displayed low speciosity, few endemics during

certain intervals and a higher proportion of more widespread taxa (i.e., MN6, MN7/8 large

mammals). However, during MN5, MN7/8 (small mammals) and MN9, the Turkish faunas

support a provincial distinction, strongly supported at the species level. At the genus level,

particularly in MN9, this region shares cosmopolitan taxa continentally with the rest of

Europe, as well as sharing taxa with Algeria and Morocco. Similar to Bernor (1983), the

Georgian and Moldovan localities cluster in MN9, but only weakly (based on species-poor localities). From MN10 to MN12, the Turkish large mammals share more similarities to

Greece, which is finally becoming well sampled during this and successive intervals.

Although these regions share large mammals with those in Bernor’s (1983) Sub-Paratethyan

province, MN11 localities in Moldova (Bernor’s Romanian and Western Russian province)

also share common taxa with Turkey and Greece, as well as regions to the west. However,

MN11 small mammals support an exclusive Greek and Turkish cluster. In MN12, Turkish and Greek faunas support both a Sub-Paratethyan province, as well as sharing taxa with the

Romanian and Western Russian province (Moldova and Bulgaria). Interestingly, the MN12 small mammal species from Turkey share no taxa in common with Greece or Moldova, although at the genus level, these faunas group very weakly with those in Greece. Turkish and Greek faunas cluster only weakly in MN13 at the species level of large mammals, and at the genus level of large mammals, the lack of endemics, coupled with the preponderance of cosmopolitan taxa, link this region continentally with the rest of Europe and

155 intercontinentally with localities in Africa. At the species level, the small mammals weakly group the Turkish and Greek localities.

Lastly, while Bernor (1983) recognized a North African province, Fortelius et al.

(1996a) limited their analysis to Eurasian localities. For the purposes of this analysis, localities from the expanse of the African continent were lumped into a single province since they were not the prime focus of this study (Figure 3.11a-c).

Figure 3.11: Geographic distribution of faunal provinces A. Modified from Bernor (1983)

2 3

1

4

5

1 Western and Southern European Province 2 Eastern and Central European Province 3 Romanian and Western USSR Province 4 Sub-Paratethyan Province 5 North African Province

156

B. Bioprovinces recognized by Fortelius et al. (1996)

2 3 4

5 1 6

1 Western Europe 4 the Black Sea 2 west Central Europe 5 the Balkans 3 Austria 6 Anatolia

C. Provinces recognized in this study

2

3 1 4

1 Spanish province 3 Italian province 2 Central European province 4 East European/West Asian Province

157

Faunal provinces: In situ evolution and dispersals

This study, together with previous research (i.e., Agustí et al. 1999a; Benammi et al.

1996; Bernor 1983; Made 1999; Steininger et al. 1985; Thomas 1985; Thomas et al. 1982), identifies the taxa involved and temporal occurrences of intercontinental dispersals between

Eurasia and Africa at various intervals during the late early, middle and late Miocene.

Similarly, patterns of in situ evolution and dispersal have been recognized for specific mammalian groups and for specific geographic regions in Europe and Western Asia (i.e.,

Agustí et al. 1996b; Bernor et al. 1996c; Fortelius et al. 2003b; Lindsay 1989; Rössner &

Heissig 1999). However, a large-scale, comprehensive account of bioprovincial evolutionary events is currently lacking. The faunal provinces, as defined in this analysis, provide the opportunity for evaluation of their constituent faunas and changes in these faunas over time, allowing for a characterization of continental trends in faunal evolution. These trends are described below and episodes of dispersal and in situ evolution are summarized in figures

3.12a-h. These figures do not include taxa previously known to a region. Location of origin,

FA or distribution in previous intervals is included in brackets (see map legends).

1. The Spanish Province: MN5 Spanish proboscideans, although African in origin, are known in Western

Europe prior to this interval (FA MN3/4) (Göhlich 1999). The palaeomerycids are both known from earlier intervals in Eurasia, as are the cervids and bovids, both of which have

Asiatic origins (MN3 and MN4, respectively) (Gentry et al. 1999). Although the small mammals known from this region originate in Asia or Eurasia, all taxa are known from previous intervals in Spain (Daams 1999; de Bruijn 1999; Kälin 1999; Ziegler 1999).

158

In MN6, Agustí (1999) notes several important events in Spain coincident with this interval, including the replacement of the suid, Bunolistriodon, by the Asian, Listriodon, in addition to the first appearances of taxa arriving from the east, including the hyaenid,

Protictitherium, bovids, Tethytragus and Hispanomeryx, and the cervid, Euprox.

During MN7/8, several taxa comprising the Spanish province are already known to the region from previous intervals. These include the previously mentioned, Protictitherium, and the deinotheriid, Deinotherium. The bovid, Miotragocerus, may also be included in this group, however while Agustí (1999) recognizes this taxon from MN5 in Spanish deposits,

Gentry et al. (1999) proposes MN7/8 for the entry of this taxon into the region. MN7/8 also marks the immigration of the bovid Protragocerus (Gentry et al. 1999), the nimravid,

Sansanosmilus jourdani (Ginsburg 1999) and mustelids, Palaeomeles and Trocharion.

Interestingly, this interval documents the arrival in Spain of taxa known from previous intervals in Europe. These include the suid, Albanohyus (Hünermann 1999); the rhinos,

Hoploaceratherium and Lartetotherium (Heissig 1999a), and hominid and pliopithecid primates. This pattern of arrival suggests that the Pyrenees were acting as a filter-type barrier that restricted these taxa from entering prior to MN7/8, particularly since some of these taxa were known adjacent to the mountains from MN5. Of the small mammals known for this interval, all of the cricetid rodents (Cricetodon, Megacricetodon, Democricetodon,

Eumyrarion) are previously known in Spain, with the one exception of Hispanomys dispectus having entered during this interval (Agustí 1999; Kälin 1999). The glirid rodents (i.e.,

Myomimus, Muscardinus) exclusive to this province also are known previously, and although

Tempestia hartenbergeri first appears in this interval in Spain, it is thought to succeed the occurrence of T. ovilis in MN5. The erinaceid insectivore, Amphechinus golpae, is known

159

from MN7/8 but the genus is known from the region since MN4 (Ziegler 1999). Similarly,

Talpa minuta is first known from MN7/8 in Spain, but T. vallesiensis and Dinosorex

sansaniensis are both known from earlier intervals in Western Europe (Ziegler 1999).

The MN9 large mammal faunas comprising the Spanish province are almost entirely

known from previous intervals in the region (Agustí 1999). Immigrants to the region,

however, include the North American Hipparion catalaunicum, entering from the east; and

the Limnonyx sinerizi, Indarctos vireti and Thaumastocyon dirus all making their

first appearances during MN9 (although Thaumastocyon is known elsewhere in Europe from

MN5) (Ginsburg 1999). Like the large mammals, the MN9 small mammals from Spain are

mostly known from previous intervals. However, Progonomys hispanicus and potentially

Cricetulodon sabadellensis both have first appearances during this interval in Spain. It is

disagreed upon, however, whether the latter taxon originated from Democricetodon, or

immigrated from the East (Agustí 1999; Agustí et al. 1997; Kälin 1999). According to

Agustí (1999), the late Aragonian (MN7/8) and earliest Vallesian (MN9) deposits in the

Vallès-Penedès Basin share such a similarity in large and small mammal faunas that they are biostratigraphically indistinguishable, in the absence of Hippotherium.

MN10 marks the mid-Vallesian Crisis, involving the extinction of forest-dwelling

large and small mammals and the replacement of these forms by more arid/open adapted

taxa. In Spain, small suids (i.e., Conohyus), cervids (Amphiprox), bovids (Miotragocerus

and Protragocerus), rhinos (Lartetotherium sansaniense), larger nimravid and amphicyonid carnivores, as well as smaller carnivores (i.e., Protictitherium) undergo extinction. These also affected most cricetids and glirid rodents that previously were abundant in

Spain. Some taxa, however, persist to the end of this interval, including hominids, larger

160

suids, Listriodon and Paraleuastochoerus, and the rodent, Cricetulodon (Agustí 1999). As a result of these immediate and prolonged extinctions, the Spanish large mammals exclusive to the region are known from individual localities (rather than being shared), while the small mammals remain almost identical to previous successions.

The MN11 large mammals from Spain contain a significant number of cosmopolitan taxa, however they also include the first appearances of the giraffid, Bigerbohlinia schaubi, the cervid, Lucentia pierensis (Agustí 1999), and the mustelid, Baranogale adroveri.

Localities continue to preserve Hippotherium primigenium from previous intervals, as well as the hyaenid, Plioviverrops guerini, which is considered to have evolved from the MN7/8 P. gaudryi (Ginsburg 1999). The small mammals for MN11 are likewise mainly composed of more cosmopolitan taxa. The two taxa shared among the Spanish localities are common in earlier intervals (Hispanomys freudenthali and Kowalskia occidentalis) and one of these taxa

(Kowalskia), most likely evolved from a Democricetodon species (Kälin 1999). Agustí

(1999) notes that both of these taxa represent the remaining members of a previously diverse cricetid radiation. Agustí (1999) also includes Occitanomys sondaari as a taxon common to this province, having evolved from the late Vallesian O. hispanicus.

MN12 marks the first appearance of the bovid, Hispanodorcas torrubiae (Made

1999), the cervid, Turiacemas concudensis (Gentry et al. 1999), as well as the North

American canid, Canis cipio (Ginsburg 1999). Agustí (1999) also notes the appearances of

Palaeoryx and Gazella during this interval. These taxa combine with the large mammals that continue their occurrences from previous intervals, including giraffid, equid, mustelid and hyaenid species. The Spanish small mammals remain similar to the preceding interval, however include the first appearance of the North American castorid, Dipoides, which

161

although known from Central European localities since MN11, becomes common only in

Spain in MN12 and MN13 (Hugueney 1999).

MN13 marks the entry of the cervid, Pliocervus turolensis, and bovid, Parabos

(Agustí 1999); the hippo, Hexaprotodon crusafonti (Made 1999); and the North American

canid, Nyctereutes (Ginsburg 1999). The remainder of the large mammals exclusive to the

Spanish province during this interval are known previously from the region. Small mammal

first appearances include Blancomys sanzi (Fejfar 1999) and Castillomys crusafonti. In

addition to the latter taxon, Agustí (1999) also notes the first appearances of lineages that

persist into the Pliocene, including Apodemus, Stephanomys, and ‘Cricetus.’ Interestingly,

Agustí (1999) also notes that although most of these taxa can be related to previous genera,

these taxa are considered as new immigrants, rather than evolving in situ from previous forms.

2. The Central European Province: All of the large mammals exclusive to the Central European province first appear from Africa or Asia, either at the genus or species level, prior to MN5 in this region (Gentry et al. 1999; Ginsburg 1999; Fortelius et al. 1996b; Made 1999), with a few notable

exceptions. These include the cervid, Dicrocerus elegans, and moschid, Micromeryx

flourensianus, both of which possibly descended from earlier forms in the region (Gentry et

al. 1999); the mustelid, Trocharion albanense, whose origins are uncertain since it is known

from Europe, Asia and at this time; and the primates, Pliopithecus antiquus

and cf. Griphopithecus, having either an Asian origin (in the former) or African origin (both),

but both arriving from the east. The Central European small mammal faunas are similar in

162 that a number of taxa exclusive to the region are known from previous intervals (Bolliger

1999; Daams 1999; Fejfar 1999). Although the insectivores, Plesiosorex germanicus and

Proscapanus sansaniensis first appear in MN5, both might be descended from species known earlier in the region (Ziegler 1999). Both Keramidomys carpathicus and Muscardinus sansaniensis are known to have first appearances in MN5 as immigrants arriving from the east. Bernor et al. (1996b) also note the number of small and large mammal lineages

“holding over” from previous intervals, and particularly recognize the prevalence of relictual small mammals in Central and Western Europe.

The MN6 large and small mammals are similarly comprised of a number of taxa known from previous intervals. Interestingly, two of the small mammals possibly evolved from earlier forms. Anomalomys gaudryi either migrated more westward or alternatively evolved from A. minor in the southern Alpine region (Bolliger 1999) and Neocometes brunonsis evolved from N. similis, known from the previous interval (Fejfar 1999). The

MN6 faunas also document a number of first appearances, including the mustelid, Lartetictis dubia of uncertain origin (Ginsburg 1999); the cervid, Euprox furcatus, of European origin

(Gentry et al. 1999); the bovid, Tethytragus from Anatolia; and the pliopithecids,

Epipliopithecus and Plesiopliopithecus. Bernor (1983) and Made (1999) note a significant immigration of Sub-Paratethyan and African faunas into this region at the base of MN6, which included primates, proboscideans (Deinotherium and Platybelodon), bovids

(Protragocerus) and suids (Kubanochoerus and Listriodon).

The MN7/8 large and small mammal faunas document fewer first appearances than

MN6 or MN9. Only the mustelid, Trochotherium cyamoides, of uncertain origin and the hominid, Dryopithecus fontani, first appear during this interval. Made (1999) also

163

recognizes the arrival of the suid, Propotamochoerus, from Asia during this interval. Among

the small mammals, Eomyops oppligeri, and Miodyromys hamadryas, have first appearances

in MN7/8 although both genera are previously known (Daams 1999; Engesser 1999).

The MN9 Central European large and small mammal faunas lack provincial distinction, however, Bernor (1983) notes that this interval begins with a large scale immigration of large mammals from the Sub-Paratethyan region and Asia. According to

Steininger et al. (1985), the first appearance of the equid, Hipparion, in MN9 coincided with

the immigration of Asian small mammals into Europe, including soricids, (Blarinella,

Anourosorex, Petenyiella), cricetids (Microtocricetus and Kowalskia), as well as the bovid,

Tragocerus. These authors also recognize the influence of African immigrants, including the rhino, Diceros; hyaena, Ictitherium; and saber-toothed felid, Machairodus.

The MN10 Mid-Vallesian Crisis in Central Europe records the extinction of many of the same forest-dwelling small and large mammal taxa as in Spain. This faunal turnover event perhaps contributes to the lack of provincial definition in the MN10 large mammal faunas, also exacerbated by the small number of localities sampled for this interval. Bernor

et al. (1996b) also note a significant drop in faunal diversity that likely had an effect on

provinciality in MN10. Bernor et al. (1996b) have observed a decrease in the number of

entries into this region, recognizing only the occurrence of the hyaenid, Adcrocuta eximia,

and large suid, Microstonyx, during this interval. Made (1999) additionally recognizes the

first occurrence of the African Pliohyrax in MN10. The MN10 small mammals retain a

number of taxa known from earlier intervals, however, Archaeodesmana vinea and

Graphiurops, both of uncertain origin, and Prospalax, possibly of Asian origin (Bolliger

1999), all have their first appearances during this interval. Bernor et al. (1996b) noticed that

164

the MN11 faunas from Central and Western Europe are relictual, due to the retention of more forest adapted taxa. This is demonstrated through the first appearance of only two large mammals, Cervavitulus mimus (Gentry et al. 1999) and Eomellivora wimani, although the

latter genus is known previously in Spain (Ginsburg 1999). The small mammal faunas from

this region are known from the previous interval and none of the taxa that are exclusive to

this region have a first appearance during MN11.

Although the Central European province lacks any endemic forms during MN12 and

MN13, Bernor et al. (1996b) indicate that the trend observed in MN11 continues into the

following intervals, however also recognize a lack of data in this region. In MN12, these

authors recognize the first appearance of the murid, Stephanomys, and in MN13, they note

the expansion of more eastern “Pikermian” faunas into Central and Western Europe (Bernor

et al. 1996b&d).

3. The Italian Province: Large mammal taxa known from Italy document the presence of three distinct

bioprovinces, two of which (Abruzzi-Apulia and Tusco-Sardinian) preserve endemic

mammals. The third province, Calabria-Sicily, preserves land mammals also known from

North Africa (Rook et al. 2006). Although taxa from the Abruzzi-Apulia province do not

appear to have relations to other known forms, taxa from the Tusco-Sardinian province have

affinities predominantly with European mammals (Moyà-Solà et al. 1999; Rook et al. 2006,

1999). According to Rook et al. (2006), these faunas are indicative of land connections with

Europe at the beginning of the late Miocene, occurring in the region of the south-western

Alps and the northern Apennines. The non-endemic Baccinello V3 fauna also indicate

migration pathways from the northern Apennines into southern Tuscany, which are supported

165 by the distribution of localities along this route (Fine Valley, Casino, Velona and Baccinello

V3) (Rook et al. 2006).

4. The Eastern European-South Asian Province The Eastern European/South Asian faunas from MN5 are dominated by Turkish large and small mammals during this interval. As noted by previous researchers, the large mammal faunas from Turkey are derived from Europe and possibly Asia, while many are endemic to Turkey. Of the taxa endemic to the Turkish localities in MN5, the bovids,

Hypsodontus and Tethytragus are both known from Europe (MN5-MN6 in the former; MN6-

MN7/8 in the latter, Gentry et al. 1999), with the former having an East Asian origin (Gentry

& Heizmann 1996). Among the suids, although both Listriodon and Bunolistriodon are present, Made (1999) notes that the former taxon evolved from the latter in the Indian subcontinent. This would suggest that the replacement of Listriodon by Bunolistriodon within Europe was not the result of in situ evolution, but rather of an immigration event.

Fortelius et al. (1996b), however, suggests that an evolutionary transition between the two genera is unlikely, but that Listriodon also has a South Asian origin in MN5. Made (2003) recognizes the suid, Schizochoerus, from the Paratethys area in MN5. The mustelid,

Ischyrictis, and the bovid, Turcocerus, have their first appearances in Turkey in MN5 (Begun et al. 2003a-b; Nagel 2003). The giraffid, Giraffokeryx, originating in Africa (Gentry &

Heizmann 1996) is known from MN6 in Europe and is considered to be dentally advanced from Georgiomeryx, occurring contemporaneously in Greece (Gentry et al. 1999). The hyaenid, Protictitherium, is known previously in Western Europe (MN4) (Ginsburg 1999;

Made 1999), as is the cervid, Heteroprox (MN5-MN7/8) (Gentry et al. 1999). According to

Bernor et al. (1996b) the former taxon appears in Turkey after extending its initial range

166

eastward. Although Gomphotherium is known previously from Europe (MN4 in Germany,

Spain), this taxon is also known from Turkey during this interval. Begun et al. (2003a & b)

additionally observe that Griphopithecus and Orycteropus are represented in MN5 in Europe

after an initial African migration. These authors also recognize a close relationship between

Aceratherium and Plesiaceratherium from MN4-MN5, and Beliajevina and Hispanotherium,

which disappears in Europe in late MN5-early MN6. Heizmann & Begun (2001) observe a

larger number of taxa that first occur in Europe (MN4/5) likely before their occurrences at

Paşalar, including Amphicyon major, Plithocyon, felids, Pseudaelurus quadridentatus and P.

larteti, suid, Conohyus, and tayassuid, Taucanamo. Fortelius et al. (1996b) also recognize

Taucanamo from MN4 and note that the Turkish species (T. inonuensis) is derived relative to

the MN5 T. sansaniensis known from Europe. Begun et al. (2003a & b) make the same

observation and further suggest the European Bunolistriodon lockharti to be also be more

primitive than those occurring in Turkey. The Turkish small mammals are derived from a

number of regions. For example, Cricetodon, Megacricetodon, Glirulus and Schizogalerix

are all known previously in the region (de Bruijn et al. 2006; Mein 2003). Pliospalax,

Albanensia and Forsythia all have their first appearances during MN5, while the European

Oligocene genus Peridyromys (Daams 1999) and European MN3 genus, Desmanodon

(Ziegler 1999), represent the rare occurrence of taxa that disperse from west to east into

Turkey (de Bruijn et al. 2003; de Bruijn & Ünay 1996; Engesser & Ziegler 1996).

The MN6 Turkish faunas lack the provincial distinction evident in MN5. Despite this, the bovid species remain unchanged from the previous interval, as does the palaeochoerid, Taucanamo inonuensis. According to Made (1999), the first appearance of

Protoryx in the Turkish deposits in this interval is the result of evolution from Tethytragus

167

(which is also still present from the previous interval). Bunolistriodon splendens known

from the previous interval is replaced by B. latidens, however no phylogenetic relationship

has been suggested between these taxa. The mammutid, Zygolophodon, is known from MN4

in Turkey (Madden 1980). Similarly, the castorid, Chalicomys jaegeri is known from the

previous interval in the region, however this taxon is cosmopolitan and is known as early as

MN4 in Central Europe (Hugueney 1999).

Most of the MN7/8 large mammals shared among the Turkish localities are

previously known in the region. However, the hyaenid, Protictitherium cingulatum, first appears in this interval (Werdelin & Solounias 1991), as does the machairodont ,

Miomachairodus pseudailuroides, the latter originating in Africa (Hoek Ostende et al. 2006).

An indeterminate species of the palaeomerycid, Triceromeryx, also has a first appearance in

MN7/8 in Turkey, however this taxon is known from much earlier deposits in Spain (MN4)

(Janis & Lister 1985). Among the small mammals from MN7/8, two species of Byzantinia have first appearances during this interval after possibly originating in Asia (Bernor et al.

1996b; de Bruijn & Ünay 1996; Rummel 1999). According to Wessels (1999) and Wessels et al. (2003), the origin and migration patterns of Myocricetodon eskihisarensis are not clear, but this taxon may have originated on the Arabian Peninsula. Although the genus,

Desmanella, is known from the late Oligocene of Europe, this taxon makes a first appearance in Turkey in MN7/8 and likely evolved from a Central European congener and dispersed into

Turkey, together with Desmanodon (Engesser & Ziegler 1996). According to Engesser &

Ziegler (1996), Schizogalerix anatolica known from this interval forms a lineage with the previously known S. pasalarensis from Paşalar.

168

The early Vallesian marks a number of immigrations into Turkey from Asia, Africa

and North America, in addition to genera previously known from the region. Bernor et al.

(2003) recognize the first appearance of the North American equid, Cormohipparion, at

10.692Ma, which they consider to be distinct from Hippotherium primigenium from Central

and Western Europe. The rhino, Ceratotherium neumayri, also has a first appearance in

Turkey in MN9 after immigrating from (Bernor et al. 1996b; Heissig 1999a).

The hyaenid, Ictitherium intertuberculatum, first appears in MN9, although the origins of this

taxon are unclear (Bernor et al. 1996b; Werdelin & Solounias 1996). The bovid, Protoryx, is

previously known from the region and the giraffid species, Palaeotragus coelophrys and P.

roueni are thought to succeed Giraffokeryx. Gentry & Heizmann (1996) also note that the

former giraffid is the likely ancestor of the latter taxon. Among the small mammals, a

number of taxa are previously known in Turkey. Byzantinia bayraktepensis and B. ozansoyi

are previously known, while B. dardanellensis and B. nokosi have their first appearances in

MN9 (de Bruijn & Ünay 1996; Ünay et al. 2003). Similarly, Schizogalerix anatolica is

previously known, while S. intermedia and S. sinapensis have their first appearances.

Selänne (2003) notes a gradual evolutionary change among these taxa. Among the immigrant taxa from Asia are Progonomys minus and Bellatonoides eroli (Bernor et al.

1996b; Made 1999; Sen 2003). The latter taxon is considered to occur later than the

Hipparion datum (base of MN9) and is contemporaneous with the appearance of

Progonomys (Sen 2003).

Most of the MN10 large mammal faunas from the regions of Turkey and Greece are

known previously. However, first appearances in the region include Chilotherium kiliasi, which originated in Asia and occurs towards the top of MN9 (Bernor et al. 1996b; Heissig

169

1996). The giraffids Bohlinia attica and Helladotherium duvernoyi are considered to have an

MN10 entry into the region (Gentry & Heizmann 1996), however Gentry (2003) suggests that the ancestors of the former taxon are perhaps known from the Eurasian middle Miocene.

The Asian chalicotheriid, Ancylotherium, also makes a first appearance during this interval, however entered Europe no later than the middle Miocene (Made 1999). Although the aardvark, Orycteropus, is previously known from the region (O. seni), Heissig (1999b) recognizes a second immigration during the Vallesian of O. gaudryi. Most of the small mammals are previously known to the region, however the gerbillid, Pseudomeriones, is considered the first of its kind to appear in this region after an Asian origin (Agustí &

Casanovas-Vilar 2003; Ünay et al. 2003).

Among the MN11 large mammals known from this region, most are previously known. Giraffa has a first appearance during this interval, however Gentry et al. (1996) suggest that this taxon is probably related to the earlier Bohlinia. Similarly, although the genus is known from the previous interval, Chilotherium samium makes its first appearance in MN11. All of the small mammals known for this interval in Eastern Europe and Turkey are previously known to the region.

The MN12 large mammals are dominated by ungulates, specifically bovids, which continue their trend of expansion and diversification from the early Turolian (Gentry &

Heizmann 1996). The large mammals are all known previously either from the region or from more central areas of Europe. An exception is the first appearance of the bovid,

Sinotragus, which is known previously from China (Geraads et al. 2002). The felid,

Machairodus giganteus, is also known from the previous interval, but is thought to have descended from the Spanish MN9 species, M. alberdiae (Ginsburg 1999). The hyaenid,

170

Hyaenotherium wongi, is known previously from MN9 in Germany (Bernor et al. 1996b).

The MN12 small mammals known to this region are also all previously known, however

Rhagapodemus and Micromys are both considered to have descended from Apodemus, also known to the region (de Bruijn et al. 1996).

The MN13 large mammal faunas from the region are previously known, mostly from the same region, but some from more central regions of Europe. The small mammals are similar in this respect, however the insectivore, Deinsdorfia, makes a first appearance in

MN13 from an unknown origin (Ziegler 1999). Allocricetus also makes a first appearance, however according to Kälin (1999) this taxon is not recognized until MN14.

Figure 3.12a: Dispersal and in situ evolution in Eurasian faunas (MN5) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Keramidomys carpathicus (EU) Muscardinus sansaniensis (EU) Trocharion albanense (EU?) Dicrocerus elegans (EU?) Pliopithecus antiquus (AS? AF?) cf. Griphopithecus (AF) Pseudaelurus quadridentatus (EU) Micromeryx flourensianus (in situ) P. larteti (EU) Plesiosorex germanicus (in situ?) Amphicyon major (EU) Proscapanus sansaniensis (in situ?) Protictitherium (EU)

Heteroprox (EU) Plithocyon (EU) Conohyus (EU) Taucanamo (EU) Peridyromys (EU) Desmanodon (EU) * Turcocerus (EU?) Ischyrictis (EU?) Albanensia (EA) Forsythia (EA) Hypsodontus (AS) Tethytragus (AS) Listriodon (AS) Bunolistriodon (AS) Pliospalax (AS) Schizochoerus (AS?) Griphopithecus (AF) Giraffokeryx (AF) * All faunas in MN5 Spanish province known from previous intervals Gomphotherium (AF – EU) EU Europe EA Eurasia AS Asia AF Africa

171

Figure 3.12b: Dispersal and in situ evolution in Eurasian faunas (MN6) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Euprox furcatus (EU) Tethytragus (AS) Listriodon (AS) Protragocerus (AS) Anomalomys gaudryi (AS, EU) Deinotherium (AF) Platybelodon (AF) Kubanochoerus (AF, AS) Neocometes brunonsis (in situ) Lartetictis dubia (?) Epipliopithecus (?) Plesiopliopithecus (?)

Hispanomeryx (EA) Listriodon (AS) Tethytragus (AS) Chalicomys jaegeri (EU?) Protictitherium (AS, EU) Protoryx (in situ) Bunolistriodon latidens (in situ?) Euprox (EU)

EU Europe EA Eurasia AS Asia AF Africa

172

Figure 3.12c: Dispersal and in situ evolution in Eurasian faunas (MN7/8) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Propotamochoerus (AS) Eomyops oppligeri (in situ?) Miodyromys hamadryas (in situ?) Trochotherium cyamoides (?) Dryopithecus fontani (?)

Albanohyus (EU) Hoploaceratherium (EU) Protictitherium cingulatum (EU) Lartethotherium (EU) Pliopithecus (EU) Triceromeryx (EU) Desmanella (EU) Sansanosmilus jourdani (EU) Hispanomys dispectus (EU) Byzantinia x 2 (AS) Talpa minuta (EU) Miomachairodus pseudailuroides (AF) Myocricetodon eskihisarensis (Ar) Miotragocerus (EU, AS) Schizogalerix anatolica (in situ) Amphechinus golpae (EA) Protragocerus (AS)

Tempestia hartenbergeri (in situ) Dryopithecus (?) Palaeomeles Trocharion

EU Europe EA Eurasia AS Asia AF Africa Ar Saudi Arabia

173

Figure 3.12d: Dispersal and in situ evolution in Eurasian faunas (MN9) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Diceros (EA?) Blarinella (AS) Anourosorex (AS) Petenyiella (AS) Microtocricetus (AS) Kowalskia (AS) Tragocerus (AS) Ictitherium (AF) Machairodus (?) Hippparion (NA)

Thaumastocyon dirus (EU) Ictitherium intertuberculatum (EU?) Indarctos vireti (EU?) Progonomys minus (AS) Progonomys hispanicus (AS) Bellatonoides eroli (AS) Hipparion catalaunicum (NA) Ceratotherium neumayri (AF) Cricetulodon sabadellensis (?) Cormohipparion (NA) Limnonyx sinerizi Palaeotragus coelophrys & P. roueni (in situ) Schizogalerix intermedia & S. sinapensis (in situ) Byzantinia dardanellensis & B. nokosi (in situ?)

EU Europe EA Eurasia AS Asia AF Africa NA North America

174

Figure 3.12e: Dispersal and in situ evolution in Eurasian faunas (MN10) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Archaeodesmana vinea (EU) Graphiurops (EU) Prospalax (AS) Microstonyx (AS) Pliohyrax (AF) Adcrocuta eximia

Bohlinia attica (EU?) * Chilotherium kiliasi (AS) Pseudomeriones (AS) Ankarapithecus (AS?) Ancylotherium (AF) Orycteropus gaudryi (AF) Helladotherium duvernoyi (?)

* mainly extinctions in Spain in MN10 EU Europe EA Eurasia AS Asia AF Africa

175

Figure 3.12f: Dispersal and in situ evolution in Eurasian faunas (MN11) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Eomellivora wimani (EU) Cervavitulus mimus (EU?)

Indarctos anthracitis (EU?) Oreopithecus bambolii (EA? AF?) Lucentia pierensis (EU?) Maremmia hauptii (EA? AF?) Baranogale adroveri (EU?) M. lorenzi (EA? AF?) Pliovivverops guerini (in situ) Tyrrhenotragus gracillimus (AF?) Mustela majori (EA?) Giraffa (in situ?) Kowalskia occidentalis (in situ) Chilotherium samium (in situ?) Occitanomys sondaari (in situ) Paludolutra campanii (AF?) Bigerbohlinia schaubi P. maremmana, (AF?) Tyrrhenolutra helbingi (AF?) Etruria viallii Eumaiochoerus etruscus

EU Europe EA Eurasia A Eurasia AS Asia AF Africa

176

Figure 3.12g: Dispersal and in situ evolution in Eurasian faunas (MN12) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Stephanomys (EU?)

Palaeoryx (EA) Hyaenotherium wongi (EU) Hispanodorcas torrubiae (EA) Gazella (AS) Machairodus giganteus (*EU) Canis cipio (NA) Sinotragus (AS) Dipoides (NA - EU) Rhagapodemus & Micromys (in situ) Turiacemas concudensis

* in situ from EU M. alberdiae EU Europe EA Eurasia AS Asia AF Africa

177

Figure 3.12h: Dispersal and in situ evolution in Eurasian faunas (MN13) (Taxa previously known to a region are not included. Location of origin, FA or distribution in previous interval is included in brackets)

Hippopotamidae (AF) Agriotherium (AS?) Macaca (AF) Parabos (AS) Orycteropus (EU, AF)

Altantoxerus (EU) Galerix (EU) Apodemus (EU?) Ruscinomys (EU) Stephanomys (EU?) Eliomys (EU) Paraethomys meini (*EU) Deinsdorfia (?) Cricetus (EU?) Parabos (AS) Prolagus (EU, AF) Allocricetus (EU?)

Hexaprotodon crusafonti (AF) Mesopithecus pentelicus (EU, AF) Nyctereutes (NA) Orycteropus gaudryi (EU, AF) Pliocervus turolensis Miotragocerus (EU, AF) Blancomys sanzi Machairodus (EU, AF) Castillomys crusafonti Nyanzachoerus (AF) Kobus (AF) Viverra (AF)

* genus known from AF & EU, P. meini from EU only EU Europe EA Eurasia AS Asia AF Africa

178 179

Paleoenvironmental Influence on Faunal Provinces

According to the paleoprecipitation maps of Eronen & Rook (2004), most of the

Spanish region is distinguished from regions to the east in terms of being drier during MN5.

Precipitation is estimated during this interval at 830-990mm/a in the region of the localities sampled, with the interesting exception of San Mamet; the one locality that shares taxa and precipitation rates with Central Europe (+1320mm/a). These drier conditions are also indicated in the Spanish small mammal succession of this interval, and also through the paleobotanical and palynological record (Agustí 1999). After this interval, Spain and the remainder of Europe is uniform in terms of paleoprecipitation until MN10, with annual rates falling after MN9. Interestingly, although eastern Turkey is depicted as significantly drier than regions to the west in MN9, the Georgian localities of Udabno and Eldari both fall within a zone of higher precipitation (~1320mm/a for this region of Georgia, in comparison to 830-990mm/a just to the south), reinforcing the fact that the Caucasus Mountains created a refugium in this region during the later Miocene. In MN10, Spain and Central Europe are both notably more humid than Eastern Europe and Turkey, while this band of aridity shifts westward from MN11-MN12. In MN12, Spain and northwestern France remain humid

(~1320mm/a), while rates in the immediate east have dropped off (~990mm/a) (Eronen &

Rook 2004).

The distinctiveness of the Spanish faunas suggest that the Pyrenees acted as a physiogeographic filter for some taxonomic groups (particularly small mammals), from MN5 to MN9. Thereafter, erosion in areas of the range perhaps facilitated the ability of other large and small mammals to more readily traverse this barrier. The more cosmopolitan nature of the faunas from Central Europe is reflective of the lack of restrictive physical barriers within

180

this region. Localities are bounded in the south by the Alpine chain, which suggests that this

feature did not act as a physical barrier to dispersal within Central Europe, however together

with the Western Paratethys, served to isolate the Italian faunas until their late Miocene

invasion of the Tusco-Sardinian and Abruzzi-Apulia regions (Rook et al. 2006). As

previously mentioned, although the mountains encircling the Pannonian Basin appear to

constitute a formidable physical barrier, localities are restricted to the peripheries of the basin

system and areas of high topographic relief in response to transgression of the Central

Paratethys and subsequent Pannonian lake. To the south, however, both the southern

Dinarides and Balkan mountains, together with the Eastern Paratethys most certainly acted to

restrict faunas from the sub-Paratethyan province of Bernor (1983). Towards the east, the lack of similarities in Georgian and Turkish faunas may suggest that together with the

Eastern Paratethys, the Caucasus Mountains may have posed a physical barrier between these

regions. However, Georgia in particular is poorly sampled and further description and

analysis of the unpublished faunas there will certainly clarify the likelihood of faunal

relations with Turkey.

Inter- and Intracontinental Faunal Exchange and Dispersal Pathways

Prior to the Messinian, faunal exchange between Africa and Eurasia occurred via

Turkey and Saudi Arabia. In the Messinian, dispersals occurred in various areas across the

Mediterranean Basin, but particularly in the region of the Iberian Peninsula (Agustí et al.

2006, 2003; Bernor 1983; Esteban et al. 1996; Steininger et al. 1985; Wessels 1999). Within

Europe, East to West migrations typify the vast majority of dispersals, however according to

181

Mein (2003), West to East migrations to Asia occurred infrequently in few mammals,

including rodent (Mein 2003; Ünay et al. 2003), felid and hyaenid taxa (Bernor et al. 1996b).

Trans-Eurasian dispersal routes in either direction would have passed either between

the Eastern Paratethys and the fore-Carpathian basin (joining Eastern Europe/Western Asia

with regions to the west) or north of the Eastern Paratethys. The paleogeographic

reconstruction of the Langhian/early Badenian (16-15Ma) Paratethys realm by Popov et al.

(2004) suggests that the former scenario is likely due to areas of subareal exposure in MN5,

however thereafter, the Eastern Paratethys merges with the fore-Carpathian basin, restricting

this route. This reconstruction contrasts to that of Rögl (1999b), which does not indicate

areas of terrestrial exposure between the fore-Carpathian basin and the Eastern Paratethys until the early Serravallian (MN6) (Figure 3.13a-b).

Figure 3.13: West Asian-Eastern European dispersal routes A. Modified from Popov et al. (2004) - Langhian/early Badenian (16-15Ma)

182

B. Modified from Rögl (1999b) – early Serravallian (MN6 - ~15Ma)

Carpathians

Eastern Paratethys

In the region of the Pannonian Basin, fluctuations in the Central Paratethys inhibited passage into the basin system until approximately 9.5Ma, when the Pannonian lake withdrew entirely from the Vienna Basin (Magyar pers. comm.). The incomplete orogeny of the

Eastern Carpathians, together with drying of the fore-Carpathian basin after MN9 would have also served as a migration route into the basin system (Nargolwalla et al. 2006). Until this point, trans-continental migrations would certainly have passed north of the Pannonian

Basin (Figure 3.14a-b).

183

Figure 3.14: Dispersal routes into and trans-Pannonian Basin A. Extent of Lake Pannon and Eastern Paratethys pre-9.5Ma

Lake Pannon

Eastern Paratethys

B. Regression of Lake Pannon and Eastern Paratethys post-9.5Ma – dispersal routes

in/out of Pannonian Basin

According to Bernor (1983), his Western bioprovince (Spain, France, Italy) was periodically segregated from the rest of Europe in response to the Paratethys marine barrier.

184

Bernor (1983) cites Antunes (1979) who suggested a “Sub-Alpine Arch,” migration pathway

linking Spain, Turkey and Asia during regression of the Tethys. In support of this

hypothesis, Bernor (1983) indicated that although a more northerly route around the

Paratethys is plausible, the presence of deciduous hardwood forests would have posed a

vegetational barrier for African immigrants until the Vallesian disconnection of the Eastern

Paratethys. He considered the Sub-Alpine Arch to have been actively utilized into the

Turolian. Szalay & Delson (1979) also recognize the Sub-Alpine Arch as the pathway used by Pliopithecus into Western Europe. In contrast, Steininger et al. (1985) suggest that the

Sub-Alpine route was highly improbable. The paleogeographic maps of Popov et al. (2004)

would support this, however as previously mentioned, Rook et al. (2006) indicate that Sub-

Alpine terrestrial conditions were responsible for faunal colonization of the Italian

paleobioprovinces.

Chapter Summary & Conclusions

o The results of the present analysis delineate four distinct paleobioprovinces, which

differ from those outlined in previous studies. 1) Spanish faunas are considered to be

distinct from those in France and Italy. 2) A Central European and 3) Eastern

European/Western Asian Province broadly corresponds to that suggested by Bernor

(1979, 1983). 4) Endemic faunas from Italy are distinct enough to warrant identity in

their own province.

o The strength and definition of these provinces fluctuates over time in response to

sampling and proportions of cosmopolitan and endemic taxa. Further sampling in

185

Eastern Europe and Georgia is necessary to more firmly establish provincial affinities

of these regions.

o Significant barriers to faunal migration include the Pyrenees (isolating Spain), the

Alps (contributing to the isolation of Italy), the Central Paratethys (restricting

migrations across the Pannonian Basin) until the top of MN 9, the Eastern Paratethys

(bounding the Eastern European/Western Asian province), and potentially the

Caucasus Mountains. Although climate would have certainly played a role in

provinciality and migration patterns, currently available data requires validation and

lacks resolution to identify any causal links between patterns of provinciality,

migration and climate.

o Small and large mammal continental migrations occur most frequently from East to

West, however instances of migrations in the opposite direction occur as well. Faunal

interchange between African and Europe occurs only through Turkey and Saudi

Arabia until the Messinian, when exchange across the Mediterranean Basin occurs.

o Patterns of in situ evolution and dispersal observed here provide a framework to

evaluate the same evolutionary events in hominoid primates.

Chapter 4 – Eurasian Miocene hominoid biogeography

Introduction

First appearances and relations among Eurasian Miocene apes

The Eurasian Miocene fossil record preserves an incredible diversity of apes, in terms of their adaptation, and geographical and temporal expanse. The relations among these apes, however, is complicated due to the fragmentary nature and the resulting morphological, temporal and biogeographic discontinuities in their fossil record, as well as a lack of consensus in the interpretation of the phylogenetic significance of preserved anatomy

(Andrews 1992a; Andrews & Bernor 1999; Andrews et al. 1996; Andrews & Martin 1987;

Martin & Andrews 1993; Begun 1995, 1994, 1992a; Begun & Kordos 1997; Begun et al.

1997; Pilbeam 1997, 1996; Pilbeam & Young 2004; Ward et al. 1997). The vast majority of

Eurasian Miocene hominoids are known from isolated occurrences, with only very few lengthy stratigraphic sequences being preserved (i.e., Spain, Italy, the Siwaliks) (Andrews &

Bernor 1999). As a result, any examination of the evolution of morphological characters, evolutionary relationships between genera and species, and dispersal patterns is forced to rely on incomplete data. According to Andrews (2007), one approach to the latter problem is to study other mammalian groups that preserve more complete fossil records and are commonly associated with apes. Brooks & McLennan (2002) point out that the initiation of speciation due to vicariance (i.e., a geological event) predicts that because ancestral communities become fragmented in the same way, all members of that community could theoretically speciate because the mechanism (i.e., orogeny) initiating speciation is independent of the community members themselves. Therefore, the expected outcome of vicariance is similar biogeographic distribution patterns in several different biotic groups (Brooks & McLennan

186 187

2002). The purpose of this chapter is to address questions regarding the chronology and

biogeographic patterns of first appearances and relatedness among Eurasian hominoids, using

occurrences of in situ evolution and dispersal of non-primate land mammals as a framework.

The focus here is primarily on fossil apes that have been recognized as potential relatives of

the African ape and human lineage and their predecessors (Griphopithecus, Dryopithecus

[including Pierolapithecus] and Ouranopithecus).

Griphopithecus and cf. Griphopithecus

Griphopithecus is known from MN5 localities in Turkey (Paşalar, Çandır) and MN6

localities in the Pannonian Basin (Klein Hadersdorf, Austria; Devínská Nová Ves – Sandhill,

Slovakia)(Figure 4.1). A fragmentary M3 from the MN5 German locality of Engelswies is

also attributed to cf. Griphopithecus and is considered to be the first appearance of this taxon

in Eurasia (Begun 2002; Heizmann & Begun 2001). Although Engelswies and the Turkish

localities belong to the same temporal interval, the former is considered to slightly predate

the latter localities (Andrews & Bernor 1999; Begun 2001; Güleç & Begun 2003). The significance of this taxon lies in its biogeographic link to preceding Afro-Arabian taxa and potential phyletic relations to subsequent African and European taxa.

188

Figure 4.1: Distribution of Griphopithecus localities

KH E DNV

PC

E Engelswies DNV Devínská Nová Ves C Çandır KH Klein Hadersdorf P Paşalar

The invasion of Eurasia by Griphopithecus is thought to have involved a descendant of or Heliopithecus after ~17Ma (Heizmann & Begun 2001). Establishing a link between these taxa and the first occurrence of Griphopithecus is complicated by three issues. First, Afropithecus is known from localities (Kalodirr, Buluk and Locherangan,

Kenya) that have been assigned to a range of dates: +17.2Ma (Leakey & Walker 1997, 1985),

18-16Ma (Leakey et al. 1988) and 17.5 + 0.2Ma (Boschetto et al. 1992). The Arabian locality of Ad Dabtiyah that preserves Heliopithecus is dated faunally and has yielded dates of 17-15Ma (Hamilton et al. 1978), ?19-16Ma (Whybrow & Clements 1999) and ?19-17Ma

(Kingston & Hill 1999). The lack of resolution in the dating of these localities creates some difficulties when attempting to establish a chronology of first appearances and biogeographic relationships. Second, although Afropithecus is fairly well known, particularly in respects to its cranio-dental anatomy, Heliopithecus is represented only by a partial maxilla containing

189 four teeth and an additional four isolated teeth, two of which are deciduous (Harrison 2002).

As a result, there is little consensus on the interpretation of the preserved morphology in the latter taxon and subsequent phylogenetic relations between these taxa. While some recognize shared dental similarities, some of which are plesiomorphic, as evidence of a close relationship (Andrews 1992b; Andrews & Martin 1987; Leakey et al. 1988), others consider these taxa to be congeneric (Andrews 1992b; Andrews & Martin 1987). In contrast, Harrison

(2002) interprets the morphological differences between these taxa as an indication that

Heliopithecus is more primitive than Afropithecus, although recognizes that further analysis is necessary. Lastly, none of the localities containing Afropithecus, and particularly

Heliopithecus, are tremendously species rich and thus further complicate the recognition of biogeographic continuity between them and the MN5 Eurasian Griphopithecus localities.

The fauna from these early Afro-Arabian localities, however, can be used to propose testable hypotheses regarding the chronology of hominoid occurrences and biogeographic relations.

At the species level, there is only one taxon shared between the early Miocene Afro-

Arabian localities and the first occurrences of Griphopithecus (including cf. Griphopithecus) in Eurasia. The gomphothere, Gomphotherium angustidens, is known from localities at

Rusinga and possibly Buluk (Gomphotherium cf. G. angustidens), and also from Engelswies.

Interestingly, in their description of the proboscideans from Wadi Moghara, Egypt, Sanders

& Miller (2002) noted the similarities in occlusal morphology and molar proportions between

Gomphotherium from this locality (G. angustidens libycum) and G. cooperi from Ad

Dabtiyah, which according to these authors, is morphologically dissimilar to the type of G. cooperi from the Bugti Beds in . In fact, based on molar similarities, Sanders &

Miller (2002) suggest that the specimens from Wadi Moghara and Ad Dabtiyah represent the

190

same subspecies of gomphothere. According to these authors, G. angustidens libycum is most similar to European G. angustidens and this subspecies represents a return to

Africa from Europe, rather than a member of the original Proboscidean Datum dispersal from

Africa. What can this tell us about patterns of hominoid migration in the early Miocene?

According to Sanders and Miller (2002), the Wadi Moghara faunas, dated biochronologically to 18-17Ma, are most similar to those from Rusinga (~17.8Ma). If their scenario about a dispersal back to Africa in the early Miocene is correct, if the Ad Dabtiyah gomphothere is actually con(sub)specific with G. angustidens libycum, and if this taxon represents a more basal member in the lineage, then Ad Dabtiyah is likely towards the older range of its age estimate (closer to 18Ma) and potentially predates Kalodirr and Buluk. This is chronologically possible, given the range of ages for both Kalodirr and Buluk. In this case,

Heliopithecus occurs earlier and could be ancestral to Afropithecus or alternatively have no relation at all. The former scenario would lend support to Harrison’s (2002) observations that Heliopithecus is more primitive than Afropithecus. Therefore, it is possible that a descendant of Afropithecus, rather than Heliopithecus, may have perhaps been the first hominoid to disperse into Eurasia in the early middle Miocene due to its temporal proximity to the first occurrence of Griphopithecus.

At the genus level, although many more shared taxa are known to provide a biogeographic link between East Africa, Saudi Arabia and Eurasia, this level of taxonomic resolution only provides broad associations at the continental and intercontinental level.

Some taxa known from these regions are so cosmopolitan and/or information is lacking regarding their centre of origin (i.e., Dorcatherium), that they are unfortunately largely uninformative. Regardless, some interesting patterns do emerge. The scenario outlined in

191 the previous paragraph may be supported by the fact that, of the genera shared between Afro-

Arabia and Eurasia between MN4 and MN5, Ad Dabtiyah shares more taxa in common with the earlier Rusinga Island faunas (six common genera), than it does with either Kalodirr or

Buluk (three common genera) (Table 4.1) and although close in age, perhaps predates both.

This pattern may be an artefact of the number of localities and species richness at Rusinga in comparison with the two Afropithecus localities, although Ad Dabtiyah is less speciose than either. Nevertheless, the hypothesis that Heliopithecus is more primitive than Afropithecus and that a descendant of Afropithecus was likely involved in the initial dispersal of hominoids to Eurasia, can be subjected to further scrutiny, either through more detailed character analysis of these two taxa or through continued analysis and comparison of the fauna from the associated localities.

Table 4.1: Eurasian – Afro-Arabian shared taxa (MN5 and earlier)

Eurasia UAE Africa Ad Al Genus Engelswies Paşalar Çandır other Dabtiyah Sarrar Rusinga Kalodirr Buluk Maboko Napak Arrisdrift Canthumeryx X X X X Bunolistriodon X X Germany X X X Kubanochoerus Belometchetskaja X Dorcatherium X X X X X X X Eotragus France, Austria X Aceratherium X X? X X X? X? X Brachypotherium X X Germany, France X X X Lartetotherium X X Chalicotherium X X France X X Amphicyon X X Germany, France X X Ysengrinia France X Hyainailouros France X Pseudaelurus X X Germany, France X X X Orycteropus X Deinotherium X X France (1) X X Platybelodon Belometchetskaja X Gomphotherium X X X X

cosmopolitan Galerix Europe X Amphechinus Germany, Georgia X X Sayimys X X

192 193

As previously mentioned, the Eurasian Griphopithecus localities date to MN5 and

MN6, however Engelswies is considered to be slightly younger and thus marks the first appearance of Griphopithecus (Andrews & Bernor 1999; Begun 2001; Güleç & Begun

2003). Is there a biogeographic relationship between these early Griphopithecus localities?

The chronology of appearance suggests two possible alternatives, both of which are based on the current state of the fossil record: first, although very close in age, Griphopithecus alpani from Turkey represents an eastern range extension of the previously occurring Central

European cf. Griphopithecus; or second, that Griphopithecus from Turkey represents a separate immigration from Africa. It is also possible that Griphopithecus was present in

Turkey prior to the FA at Engelswies, yet is not being sampled.

The eastern range extension alternative can be addressed through analysis of the first appearances and centres of origin of the taxa known from Engelswies, Paşalar and Çandır.

The results of this analysis are summarized below (Table 4.2):

Table 4.2: FAs and centres of origin of German and Turkish MN5 taxa FA FA Europe Turkey* Uninformative** Total genera (shared & unshared) large mammals (39) 19 (49%) 11 (28%) 8 (21%) small mammals (23) 9 (39%) 7 (30%) 7 (30%)

Total genera (shared only) large mammals (10) 9 (90%) 1 (10%) small mammals (6) 2 (33%) 2 (33%) 2 (33%) * occurs at Paşalar OR Çandır, or both ** FA occurs simultaneously in both regions or has limited range (endemic)

At a first glance, these results suggest that more taxa occur first in Central Europe in comparison to Turkey, thus supporting an eastward dispersal of Griphopithecus. These taxa include Palaeomeryx and Pseudaelurus, which are both known from MN3 in Europe and not

194 until MN5 in Turkey. Both of these taxa appear in MN4/5 in Greece, perhaps on their way to

Turkey. Similarly, both Amphicyon and Ursavus have their first occurrences in MN3, prior to their appearance at Paşalar and Çandır (Bernor & Tobien 1990). Among the small mammals, Keramidomys thaleri is known earlier in Europe than Turkey, however according to de Bruijn et al. (2003), the absence of this taxon at Paşalar is potentially due to differences in collecting techniques, since many of the small mammals from this locality demonstrate the opposite pattern; occurring first in Turkey before extending their ranges into Europe.

However, as noted in the previous chapter, a pair of European insectivores, Peridyromys and

Desmanodon, known from previous intervals also disperse from west to east into Turkey in

MN5 (Daams 1999; de Bruijn et al. 2003; de Bruijn & Ünay 1996; Engesser & Ziegler 1996;

Ziegler 1999), and thus directionality of dispersal to the east is established in other taxa, regardless of whether Keramidomys is being sampled in Turkey during this interval or not.

Although an eastward range extension of Griphopithecus after its first appearance at

Engelswies is well-supported by other fauna, the presence of mammals of African origin

(mostly large mammals) that would have had to pass through Turkey en route to Central

Europe, but perhaps are not being preserved or have not been recovered, indicate a taphonomic bias in the Turkish faunas. 16-30% of the large mammals in Table 4.2 that first occur in Europe have previous records in Africa and would have passed through Turkey on their way to Central Europe, yet appear in Turkey after their FA in Europe. For example, a first appearance of Griphopithecus in Turkey is more parsimonious geographically, since the dispersal pathway to Central Europe from Africa would certainly have passed through

Turkey, yet this taxon is documented first in Central Europe. Although it is certainly possible, and perhaps even likely, that Griphopithecus is present but remains unsampled at

195 earlier intervals, this same pattern of occurrence (having an African origin, FA in Europe and subsequent appearance in Turkey) is evident in a number of other MN5 large mammal genera, including Bunolistriodon and Chalicotherium, and possibly Dorcatherium,

Aceratherium, Brachypotherium and Lartetotherium (the FAs of the latter three taxa in

Africa are based on land mammal ages and thus are lacking in temporal resolution), and reveals a taphonomic bias affecting large mammal taxa.

The presence of bias is also supported by the previous study of provinciality (chapter

3). Both Dorcatherium and Bunolistriodon are known earlier in Greece (MN3/4 in the former and MN4 in the latter), while not occurring in Turkey until MN5 even though they belong to the same biogeographic province. It is thus possible that these taxa did occur in

Turkey before MN5, yet are not being preserved and/or sampled. Conversely, it is also possible that these taxa simply do not appear in Turkey until MN5.

In addition, bias against large mammal preservation and/or sampling in the Turkish fossil record is clearly apparent in earlier intervals. MN3 is represented by 18 Turkish localities in the NOW Database. Large mammals are recorded at five of these localities but all are taxonomically indeterminate. In contrast, only half of the small mammals are indeterminate at seven localities and thus are better represented in MN3 than their large mammal counterparts. In MN4, 22 localities are known in Turkey. While the small mammals are better known during this interval (only indeterminate at six localities), only one large mammal is known and is taxonomically indeterminate. Therefore, although many pre-

MN5 localities are known in Turkey, the faunas from them (in MN3) and large mammals (in both intervals) are poorly represented. This suggests that perhaps the same phenomenon is affecting taxonomic representation of large mammals in MN5, but to a lesser degree. Begun

196

(pers comm) has suggested that a possible cause of this pattern is the scarcity of MN5 localities in Turkey. Furthermore, he suggests that it is likely that the age designations for many MN6 Turkish localities are underestimated and that in actuality, these localities may possibly be attributable to MN5.

A significant number of the small mammals known in Europe (and Turkey) are thought to have originated in Asia and may also have passed through Turkey prior to their occurrences in Central Europe. Alternatively, these Asian small mammal groups may have accessed Europe via a more lengthy northern route around the Black Sea and Eastern

Paratethys, however this region is poorly sampled and evidence for such a route is currently lacking.

One further consideration is the dispersal pathway of the into Central

Europe. If this group originated and dispersed directly from Africa (Begun 2002; Harrison et al. 1991; Made 1999; Rossie & MacLatchy 2006; Thomas 1985) before appearing in Central

Europe in MN5, this casts some doubt on the taphonomy argument. Unlike the previously mentioned taxa, which do not appear prior to MN5 in Turkey (i.e., Bunolistriodon,

Dorcatherium, Chalicotherium), pliopithecoids, like Old World monkeys, are not known from any interval in Turkey and serve as interesting exceptions to the trend of taxa originating in Africa and appearing first in Europe, and subsequently in Turkey. The earliest pliopithecoids out of Africa first occur in Asia at ~18-17Ma (Harrison 2004; Harrison & Gu

1999; Zhang & Harrison 2008), afterwhich a westward dispersal route north of the Black Sea into Europe is plausible for this group. The paleogeographic maps of Popov et al. (2004) and Rögl (1999) support this hypothesis, since both depict the separation of Saudi Arabia and

Turkey by the Tethys in MN3 and MN5, as well as the separation of Turkish land mass from

197

Iran and Iraq by the Eastern Paratethys during the same intervals (Figure 4.2a). As a result, dispersal into Europe during these intervals could not have passed through Turkey without passing north of the Eastern Paratethys. Passage of the pliopithecids into Europe at this time, either directly from Africa or from Asia, was only possible during MN4 (Figure 4.2b).

Figure 4.2a: MN3/5 dispersal pathways into Europe (modified from Popov et al. 2004)

water barriers between Saudi Arabia and Turkey, and the Middle East and Turkey

alternative pathway into Europe

198

Figure 4.2b: MN4 dispersal pathways into Europe via Turkey (modified from Rögl 1999)

Based on the current state of the fossil record, the majority of fauna from

Engelswies, Paşalar and Çandır support the FA of Griphopithecus in Central Europe, followed by an eastward range extension of this taxon into Turkey. However, this scenario poses some geographic and taphonomic inconsistencies. After a direct dispersal from Africa, a FA of Griphopithecus in Turkey is also possible, although this scenario is not as well supported faunally. Neither the limited fauna from Engelswies, nor the fragmentary and temporally mixed fauna from Turkey can likely resolve this issue. However, discovery of additional MN5 localities preserving Griphopithecus could provide insights into the biogeographic patterns of dispersal and first appearances of this taxon.

How, if at all, is MN5 Griphopithecus related biogeographically to subsequent G. darwini in Slovakia and Austria? It is most parsimonious in terms of geographic proximity

199 to assume that the later occurring Central European species is derived from cf.

Griphopithecus from Engelswies. Phylogenetically, the likelihood of this relationship could be established through a comparison of very limited specimens: one partial molar

(Engelswies) to five isolated teeth (Devínská Nová Ves) to the postcranial material from

Klein Hadersdorf. Despite this, Andrews & Kelley (2007) note similarities of the Devínská

Nová Ves specimens to both Engelswies and Paşalar, while Begun et al. (2006a) more conservatively maintain that the relations of the Pannonian Basin G. darwini to those from either the German or Turkish fossils remains unclear.

As demonstrated in the previous chapter, the Central European bioprovince during

MN6 is comprised almost entirely of taxa known previously in the region, with several occurrences of in situ evolution within the small mammals and few FAs of large mammal taxa arriving from the east. These findings, together with those of Bernor (1983), indicate that the Griphopithecus occurrences in Central Europe are all located within the same bioprovince, and together with few first appearances in this region, suggest that the presence of Griphopithecus in the Pannonian Basin is most likely biogeographically linked to cf.

Griphopithecus from Engelswies, rather than G. alpani from Turkey. However, the Eastern

European/Western Asian province recognized in the previous chapter lacks distinction in

MN6 due to an influx of cosmopolitan taxa, together with low species richness. Upon direct faunal comparison of the MN5 and MN6 Griphopithecus localities, it appears that despite a large number of unshared taxa between intervals, that the MN6 Pannonian Basin localities share more taxa in common with the Turkish localities than with Engelswies (Table 4.3).

200

Table 4.3: European-Turkish shared Astaracian taxa Devínská Nová Klein Combined Ves Hadersdorf Sample Çandır/Paşalar shared species 2 (7%) 1 (14%) 3 (8%) shared genera 8 (31%) 3 (43%) 11 (33%)

Engelswies shared species 2 (7%) 0 (0%) 2 (8%) shared genera 2 (8%) 0 (0%) 2 (6%)

Çandır/Paşalar + Engelswies shared species 1 (3%) 0 (0%) 1 (3%) shared genera 5 (19%) 2 (29%) 7 (21%)

Unshared shared species 24 (83%) 6 (86%) 30 (83%) shared genera 11 (42%) 2 (29%) 13 (39%)

Total species 29 7 36 genera 26 7 33

This pattern is likely influenced by the low species richness at some of the localities, particularly Klein Hadersdorf, and when considering the centres of origin of the taxa represented at the MN6 Griphopithecus localities, this pattern ceases to exist almost entirely.

Of the MN6 large and small mammals at Devínská Nová Ves and Klein Hadersdorf, only two to four taxa first appear in Turkey, while 10-13 first appear in Europe. 15 additional taxa known from these localities (including Griphopithecus) are indeterminate, either due to simultaneous first occurrences in Turkey and Europe or endemism, and thus determination of their centres of origin would certainly clarify this issue. From the current faunal evidence, a biogeographic association between cf. Griphopithecus from Engleswies and G. darwini from the Pannonian Basin is more plausible than a close relation of the latter taxon and

Griphopithecus from Turkey.

201

A further consideration is the relationship between Griphopithecus and

Kenyapithecus, the latter recognized now as the second (less common) species (K. kizili) at

Paşalar and slightly later at Fort Ternan (K. wickeri) (Andrews & Kelley 2007; Begun 2001;

Humphrey & Andrews 2008; Kelley 2008; Kelley et al. 2008). Although the presence of a second hominoid at Paşalar is accepted based on morphological and metric variation in the dental sample from this locality (i.e., Andrews & Kelley 2007; Humphrey & Andrews 2007;

Kelley 2008; Kelley et al. 2008; Kelley & Alpagut 1999; Martin & Andrews 1993; Waddle et al. 1995; Ward et al. 1999), this represents the only occurrence of sympatric large bodied apes at a Eurasian Miocene locality. This is curious, considering that the norm is occurrence of single ape taxa and rarely, an ape with a pliopithecoid. Previously, Begun (2002, 2001),

Begun et al. (2006a, 2003) and Heizmann & Begun (2001) suggested that K. wickeri from

Fort Ternan bridged the temporal gap between previously occurring Griphopithecus and subsequent European taxa and that this taxon represented an early Serravallian return of hominoids to Africa during a cycle of sea level lowering, contemporaneous with the appearance of Griphopithecus in the Pannonian Basin. Dispersal back to Africa has been previously shown to be supported faunally by the appearance of rodent (Tong & Jaeger

1993), ruminant (Gentry & Heizmann 1996), suid (Bunolistriodon – although this taxon is previously known in Africa) and carnivore taxa (Thomas 1985) at Fort Ternan that are considered to be Eurasian in origin. Interestingly, Begun (2002) and Begun et al. (2003) also suggest that Limnopithecus legetet from Fort Ternan may also have affinities to Eurasian pliopithecoids. From the study of bioprovinciality in chapter 3, the pattern of faunal association between Turkey and Kenya is supported by Brachypotherium, Aceratherium and

202

Dicerorhinus. Fort Ternan and Devínská Nová Ves also share Albanohyus pygmaeus, although this taxon is cf. at Fort Ternan.

More recently, Andrews & Kelley (2007) observed that, with the exception of minor morphological details of the anterior , the degree of similarity between the species of

Kenyapithecus from Turkey and Kenya makes distinguishing between the two taxa difficult.

Kelley et al. (2008) and Andrews & Kelley (2007) recognize a suite of synapormorphies which, according to them, are clear evidence of a phylogenetic and paleobiogeographic link between the earlier occurrence at Paşalar and Fort Ternan at ~14Ma. These authors suggest that after an initial dispersal from Africa of the more primitive Griphopithecus, the more derived Kenyapithecus dispersed back to Africa.

In sum, the results of this analysis together with previous research suggest that a descendant of Afropithecus extended its range into Eurasia ~17Ma. The current state of the fossil record supports a first appearance of cf. Griphopithecus at Engelswies in MN5, where it most likely extended its range eastward thereafter. Slightly later in MN5, Griphopithecus is recognized at Çandır and Paşalar, together with Kenyapithecus kizili at the latter locality.

Since the fossil assemblage at Paşalar is considered to represent a restricted temporal period and approximates a natural assemblage (Andrews 1995), the occurrence of two large bodied hominoids at this locality were likely coincident and thus represents the only occurrence of sympatric apes in Eurasia. There is some support for the hypothesis that Griphopithecus from Turkey represents the first appearance of this taxon in Eurasia (rather than Engelswies), however the eastern range extension hypothesis is more supported faunally. Similarly, the majority of fauna support a biogeographic relation link between the Central European cf.

203

Griphopithecus and G. darwini from the Pannonian Basin. However, a small number of taxa also support an association between the latter and G. alpani from Turkey.

Dryopithecus/Pierolapithecus

Several hypotheses have been put forth with the purpose of explaining the phyletic relations of Griphopithecus to later European hominids, specifically Dryopithecus. The last occurrence of Griphopithecus (Klein Hadersdorf, top of MN6) immediately precedes the first appearance of Dryopithecus (D. fontani) in the Pannonian Basin (St. Stephan, MN7/8) and despite morphological differences in these taxa, an ancestor-descendant relationship is possible (Agustí et al. 1999; Andrews & Bernor 1999; Andrews et al. 1996; Begun et al.

2006a). Alternatively, the appearance of Dryopithecus in Europe could represent a separate immigration event from Africa (Agustí et al. 1999; Andrews & Bernor 1999; Made 1999).

Agustí et al. (1999) suggest an additional hypothesis, proposing that Griphopithecus and

Dryopithecus immigrated to Europe together and perhaps due to taphonomic processes,

Dryopithecus does not appear in the fossil record until much later on. The recent discovery of Pierolapithecus catalaunicus from MN7/8 deposits in Spain must also be considered with these hypotheses for the appearance of middle and late Miocene apes. According to Begun

(2006) and Begun et al. (2008, 2006b), this taxon is conspecific with D. fontani in France and Austria and thus the nomen, Pierolapithecus, is a junior subjective synonym of

Dryopithecus. Begun et al. (2008) suggested that this taxon diversified into at least two genera, one which is more endemic and one which is more cosmopolitan, following the pattern of non-primate land mammals in their respective regions. A biogeographic association between the latest occurring Griphopithecus and the first appearance of

204

Dryopithecus and/or Pierolapithecus can be evaluated through faunal continuity between localities (Figure 4.3). Lastly, Begun suggested that a relative of a western Eurasian hominid, such as Dryopithecus or Ouranopithecus, dispersed back to Africa in the late

Miocene and represents the sister group to the African ape and human lineage (Begun 2005,

2002, 2001; Begun et al. 1997). Review of the taxa involved in intercontinental dispersals during the temporal range of Dryopithecus can potentially clarify this issue.

Figure 4.3: Distribution of Dryopithecus localities (including Pierolapithecus)

W M S*S, ,*Me *Me, , *T G R SS LGL SG E*E H de P, LT U H de P CL,*CV , *SQ,*C de B CP

* lack of concensus Late Astaracian (MN7/8) SS St. Stephan (Austria) Dryopithecus fontani LG La Grive St. Alban (France) Dryopithecus fontani SG St. Gaudens (France) Dryopithecus fontani H de P Hostalets de Pierola Inferior (Spain) Pierolapithecus catalaunicus CV Can Vila Dryopithecus laietanus ? SQ Sant Quirze Dryopithecus laietanus ? C de B Castell de Barberà Dryopithecus laietanus ? Vallesian (MN9 - MN10) R Rudabánya Dryopithecus brancoi M Mariathal Dryopithecus brancoi? Me Melchingen Dryopithecus brancoi? T Trochtelfingen Dryopithecus brancoi? W Wissberg Dryopithecus brancoi? EF El Firal Dryopithecus fontani/crusafonti? CL Can Llobateres Dryopithecus laietanus CP Can Ponsic Dryopithecus crusafonti G Götzendorf ? S Salmendingen Dryopithecus brancoi? LT La Tarumba Dryopithecus laietanus P Polinýa Dryopithecus laietanus U Udabno Dryopithecus garedziensis

205

In Central Europe, Made (1999) noted considerably fewer immigrations in MN7/8 than in the preceding and subsequent intervals, which would support the argument that

Dryopithecus is derived from locally occurring Griphopithecus, rather than the result of a separate immigration event. Although relatively few taxa are shared between the MN6

Griphopithecus and MN7/8 Dryopithecus/Pierolapithecus localities in comparison to the total number represented at these localities (13 of 109 genera, eight of 128 species), this is almost certainly a reflection of the overwhelming taxonomic richness of the MN7/8 reference locality, La Grive, in comparison to earlier Griphopithecus localities or Hostalets.

Nevertheless, these localities do in fact share taxa. The occurrence of Albanohyus in Spain in

MN7/8 is the result of a westward immigration into the region, which also likely included

Dryopithecus/Pierolapithecus. Despite the high degree of endemism in Spain from MN5 to

MN9, the MN7/8 Spanish faunas include approximately double the immigrants in either the preceding or subsequent intervals (Table 4.4), and therefore, if Dryopithecus/

Pierolapithecus immigrated prior to MN7/8, it is not preserved in the MN6 deposits. Central

Europe, in contrast, shows the opposite pattern; fewer immigrations in MN7/8 (only three) than MN6 and MN9 and instead the MN7/8 Central European province is formed largely of taxa already known to the region (Figure 4.4). This latter observation also supports the hypothesis that the MN7/8 occurrence of Dryopithecus in Central Europe is most likely the result of in situ evolution from a taxon previously known to the region.

206

Table 4.4: Spanish large and small mammal FAs, MN6-MN9 MN 6 FAs MN7/8 FAs MN9 FAs Listriodon Protragocerus Hipparion catalaunicum Tethytragus Miotragocerus* Limnonyx sinerizi Hispanomeryx Albanohyus Indarctos vireti Euprox Hoploaceratherium Thaumastocyon dirus Protictitherium Lartetotherium Progonomys hispanicus Dryopithecus/ Cricetulodon sabadellensis** Pierolapithecus Pliopithecus Sansanosmilus jourdani Palaeomeles Trocharion Hispanomys dispectus * uncertain FA – Genry et al. (1999) FA = MN7/8, Agustí (1999) FA = MN5 ** no concensus on whether this taxon represents in situ evolution from Democricetodon or whether it is an immigrant from the east (Agustí 1999, Agustí et al. 1997, Kälin 1999)

Figure 4.4: Late Astaracian Eurasian immigrations

FA Central Europe Trochotherium cyamoides (?) Propotamochoerus (As) Dryopithecus fontani Eomyops oppligeri Miodyromys hamadryas FA Spain Protragocerus FA E Europe/W Asia ? Miotragocerus Miomachairodus pseudailuroides (Af) Sansanosmilus jourdani Protictitherium cingulatum (Eu) Palaeomeles Triceromeryx (Eu) Trocharion Byzantinia (As) Albanohyus Desmanella (Eu) Hoploaceratherium Desmanodon (Eu) Lartetotherium Pierolapithecus Pliopithecus Hispanomys dispectus Amphechinus golpae

The second hypothesis of a dispersal from Africa for Dryopithecus is also supported, albeit weakly, by the faunal data. Made (1999) identified two dispersal events during

MN7/8. Although this temporal interval was not coincident with a regression, sea levels

207 were already very low and likely did not restrict intercontinental exchange pathways. The first, the “Tethytragus event,” at 12.5Ma involved Albanohyus and possibly Dryopithecus entering Europe, while the hyaenid, Percrocuta, appears in Africa. Towards the end of this interval, Made (1999) noted a “Propotamochoerus event,” in which Protoryx disperses to

Africa. In the previous chapter, Protoryx solignaci was found to be shared between Turkish localities and the Tunisian locality of Bled Douarah, however the Turkish occurrence of this taxon predates (MN6-MN7/8) the Tunisian occurrence (MN7/8-MN9). The proboscidean,

Tetralophodon longirostris, also possibly demonstrates the same pattern as Protoryx, although the temporal range of this taxon is not terribly well constrained. The earliest occurrence of T. longirostris in Europe is at Nombrevilla 9 (Spain) during the Aragonian (18-

11.2Ma), but is later known at Sant Quirze (Spain) with a more definite age of MN7/8. In

Africa, Tetralophodon cf. T. longirostris is known from Djebel Krechem el Artsouma

(Tunisia), however this locality, like Nombrevilla 9, is dated from the Vallesian to the

Turolian (11.2Ma-5.3Ma). Tetralophodon sp. is similarly known from the Chorora

Formation in Ethiopia from the Serravallian to the Tortonian (13.65-7.25Ma). Therefore, it is possible that this taxon, together with Protoryx, dispersed to Africa towards the top of

MN7/8, however the temporal resolution of the associated localities must be refined. Despite the significant number of taxa shared between Africa and Eurasia during MN7/8, the vast majority are previously known in both regions and are thus uninformative in clarifying the appearance of Dryopithecus. In addition, Central Europe during MN7/8 documents very few immigrations (only three), supporting the hypothesis that Dryopithecus is derived from a taxon already known to the region. It is still possible that the occurrence of Dryopithecus in

208

Europe is the result of an immigration from Africa, although very few taxa support this hypothesis.

The third hypothesis, of an earlier immigration of Dryopithecus along with

Griphopithecus but lack of retrieval for taphonomic reasons, is particularly interesting, but difficult to test. However, the results of the completeness analysis in Chapter 2 lend some interesting insights. The appearance of Griphopithecus in MN5 corresponds to an interval that is extremely well sampled with CIs between 92.5 and 98 for the large mammals and between 90.7 and 96.6 for the small mammals, both well above the cutoff of 70. These results would suggest that the likelihood of Dryopithecus being unsampled during this interval following an immigration with Griphopithecus is unlikely, particularly since the small mammals, who are more prone to taphonomic bias, are extremely complete. However, if Dryopithecus immigrated from Africa at the base of MN6, during the extensive interchange identified by Bernor (in Steininger et al. 1996), there is slightly more of a possibility that it is not being sampled during this interval. Although most of the CIs remain above the cutoff of 70, they have noticeably decreased from both the preceding and subsequent interval (76.3-86.2 for the large mammals and 68.1-77.5 for the small mammals).

The results of the completeness analysis therefore suggest that it is unlikely that

Dryopithecus immigrated with Griphopithecus in MN5 and is not being sampled. Although there is a slight possibility of this taxon going undetected in MN6, all measures of completeness suggests that although there are more sampling gaps, this interval still remains relatively well sampled, with the exception of the strict index for the small mammals.

Furthermore, in Spain, few entries (except in MN7/8) and the endemic nature of Spanish faunas would also suggest that Dryopithecus and/or Pierolapithecus were not present in this

209 region until the influx of immigrations in MN7/8. In sum, of the three hypotheses put forth to explain the appearance of Dryopithecus in Central Europe and Spain in MN7/8, the most well supported scenario is that this taxon evolved from a species previously known to the region.

In 2004, a large bodied ape was described from the upper MN7/8 locality of Els

Hostalets de Pierola and a new nomen, Pierolapithecus catalaunicus, was erected to distinguish this taxon from other early, middle and late African and Eurasian Miocene hominoids (Moyà-Solà et al. 2004). Based on their analysis of the preserved morphology of this taxon, Moyà-Solà et al. (2004) proposed that Pierolapithecus represents a stem hominid.

Begun et al. (2006b) and Begun and Ward (2005), however, suggested that Pierolapithecus is instead most likely a stem hominine and also that this taxon is conspecific with D. fontani from Austria and France based on dental similarities, temporal and geographic proximity, and the paleobiogeographic trends in other terrestrial mammals (Begun 2006; Begun et al.

2008, 2006b). With respect to the latter, these authors found that despite the endemism of the Spanish faunas, 59% of the total Hostalets mammals are also shared with La Grive, including an 82% overlap in carnivore taxa (Table 4.5). Furthermore, in the previous chapter, Hostalets, together with Sant Quirze, was found to cluster broadly with La Grive and

Steinheim. Within Spain, Agustí (1999) identified MN8 in the Vallès-Penedès Basin through the entry of Hispanomys, Palaeotragus, Protragocerus and Tetralophodon. Interestingly, of these immigrant taxa, Hostalets shares Palaeotragus with other MN7/8 primate-bearing localities, such as Castell de Barberà, and Protragocerus with Castell de Barberà and Sant

Quirze. Therefore, it seems unlikely that while having these immigrants in common,

Hostalets would receive a different genus of ape. These data, together with the influx of

210 first appearances in MN7/8 in Spain, clearly indicate faunal continuity between the Spanish and Central European bioprovinces and therefore support the hypothesis that Pierolapithecus is at least congeneric with Dryopithecus.

Table 4.5: MN7/8 Spanish – Central European shared taxa Locality Hostalets La Grive St. Gaudens St. Stephan (Spain) (France) (France) (Austria) Taxon Chalicotherium grande X X X Listriodon splendens X X X Euprox furcatus X X X Albanohyus pygmaeus X X Protragocerus chantrei X X Plithocyon armagnacensis X X Pseudaelurus quadridentatus X X Pseudaelurus lorteti X X Hemicyon goriachensis X X Sansanosmilus jourdani X X Semigenetta sansaniensis X X Protictitherium crassum X X Thalassictis X X Alicornops simorrensis X X Lartetotherium sansaniensis X X Deinotherium giganteum X X

The results of the previous chapter support the endemic nature of the Spanish lineage from

MN7/8 to MN9, with very few immigrations and the virtually unchanged nature of the large and small mammal faunas. Agustí (1999) also noted that the MN7/8 to MN9 Spanish faunas are biostratigraphically indistinguishable in the absence of Hippotherium. Across the MN9

- MN10 transition, the Spanish faunas continue to support a single endemic ape lineage in

Spain. The results presented in the previous chapter indicate that the MN10 Spanish localities (including La Tarumba) continue to cluster together. Furthermore, comparison of the MN10 Spanish Dryopithecus locality, La Tarumba, (Polinýa II has only a single taxon), reveals that this locality shares 67% of its constituent species and 73% of its genera with

MN9 Spanish localities.

211

Following a re-analysis of the Spanish apes, Begun (1992b) described a new species of Dryopithecus, D. crusafonti, from Can Ponsic in the Vallès-Penedès Basin, northeastern

Spain. Similar in age to Can Llobateres (MN9), Can Ponsic is considered to be slightly older based on its rodent faunas. This locality is also similar to Can Llobateres in terms of its faunal composition, although slightly less abundant, but with similar depositional and paleoecological settings (Begun 1992b and references therein). Begun (1992b, 1991) recognized this new species based on dental differences with other Spanish and European

Dryopithecus, including distinctive molar occlusal morphology, large and broad upper molars and very high crowned upper central with well-developed lingual pillars.

Begun (1992b) also noted similarities in two lower molars from Can Ponsic (IPS 1813 and

1816) to the El Firal mandible, which he attributed to D. cf. D. crusafonti. Begun (1992b) observed that D. crusafonti is most similar in dental characters to D. laeitanus and dissimilar from D. fontani and the Central European D. brancoi.

In contrast, Ribot et al. (1996) reviewed the morphology of Dryopithecus from

Vallès-Penedès, as well as El Firal, and using the range of variation observed in extant apes, concluded that sufficient morphological or metrical distinction was currently lacking to justify the recognition of two separate species in the samples from Can Ponsic and other sites in Vallès-Penedès. These authors attributed all material from Vallès-Penedès to D. laietanus and recognized D. crusafonti as a junior subjective synonym of D. laietanus. Specifically,

Ribot et al. (1996) found that lower molar morphology of El Firal distinguished this specimen from the entire sample from Vallès-Penedès and were unable to find any morphological features that El Firal shared uniquely with Can Ponsic, to the exclusion of other specimens from Vallès-Penedès. However, like Begun (1992b), Ribot et al. (1996)

212 also observed differences between El Firal and St. Gaudens, but noted a number of distinctive features that the El Firal mandible shares with D. fontani that are not seen in other

Dryopithecus specimens in Spain. In conclusion, these authors concur with the majority of prior research (i.e., Andrews et al. 1996) in recognizing the El Firal mandible as belonging to

D. fontani.

As previously mentioned, there is little change and very few immigrations from

MN7/8 to MN9 in Spain as a whole. The results from chapter 3 indicate that for large and small mammal species, Can Ponsic groups with Can Llobateres, and for large and small mammal genera, Can Ponsic groups overall with Spanish localities. Although the degree of similarity between Can Ponsic and Can Llobateres is not tremendous, this is likely due to the size of the Can Llobateres faunal list (79 versus 52 taxa). Upon closer examination, the similarity between these localities is much more evident. Of the fauna present at Can Ponsic,

69% of species and 83% of the genera are shared with Can Llobateres (Table 4.6). El Firal, shares 62% of its species and 75% of its genera with Can Llobateres. Although the latter finding is almost certainly being influenced by the considerably smaller faunal list at El Firal

(16 taxa), there are still a number of taxa at El Firal that are not known at either Can Ponsic or Can Llobateres, including Dicrocerus elegans, Amphicyon pyrenaicus and

Gomphotherium angustidens. From chapter 3, the large mammals from El Firal clustered broadly with Spain at the species level, but with Wissberg and the Central European cluster at the genus level. Dicrocerus elegans is shared between El Firal and Wissberg. Amphicyon pyrenaicus is unique to El Firal and the only other occurrence of Gomphotherium angustidens in Spain during MN9 is at Santiga. These findings indicate that the faunas from

Can Ponsic and to a lesser extent, El Firal, are similar enough to Can Llobateres to suggest

213 that, coupled with the very low immigration in MN9, it is perhaps unlikely that an additional species of Dryopithecus (i.e., D. crusafonti) dispersed into the region. However, the replacement of four other mammalian species perhaps due to in situ evolution (Semigenetta sansaniensis or S. grandis to S. ripola, Dorcatherium crassum to D. jourdani or D. naui,

Hispanomys dispectus to H. thaleri and Ischyrictis mustelinus to I. petteri) suggests that perhaps D. crusafonti evolved from Dryopithecus previously known. The presence of several taxa that are shared with other localities in Central Europe also indicates that perhaps

El Firal is somewhat distinct from the Vallès-Penedès faunas.

Table 4.6: Shared Spanish taxa Locality Can Ponsic I El Firal Can Llobateres Taxon Listriodon splendens X X Parachleuastochoerus steinheimensis X X Parachleuastochoerus huenermanni X Propotamochoerus palaeochoerus X X X Dorcatherium jourdani X Micromeryx flourensianus X X Euprox dicranoceros X X Miotragocerus pannoniae X X Palaeotragus indet. X X Sansanosmilus jourdani X X Amphicyon cf. major X X Ursavus primaevus X X Protictitherium gaillardi X X Machairodus aphanistus X X Promeles indet X Indarctos vireti X X Mesomephitis medius X X Limnonyx sinerizi X Martes cf. andersoni X X Thalassictis montadai X Plesiodimylus chantrei X X Galerix socialis X X Lanthanotherium sanmigueli X X Dinosorex sansaniensis X X Talpa minuta X X Talpa vallesensis X Postpalerinaceus vireti X X Prolagus crusafonti X X Hoploaceratherium tetradactylum X

214

Locality Can Ponsic I El Firal Can Llobateres Taxon Alicornops simorrensis X X Aceratherium incisivum X X X Dicerorhinus steinheimensis X Lartetotherium sansaniensis X X cf. Hippotherium catalaunicum X X X Chalicotherium grande X X X Tapirus priscus X X X Deinotherium giganteum X X Cricetulodon hartenbergeri X Miodyromys hamadryas X Myoglis meini X X Muscardinus hispanicus X X Albanensia cf. grimmi X X Heteroxerus cf. grivensis X Heteroxerus rubricati X Spermophilinus bredai X X Hispanomys thaleri X X Trogontherium minutum X X Miopetaurista crusafonti X X Keramidomys carpathicus X Chalicomys jaegeri X X X Dicrocerus elegans X Amphicyon pyrenaicus X Dihoplus schleiermacheri X X Gomphotherium angustidens X Tetralophodon longirostris X X Paraleuastochoerus crusafonti X

Dryopithecus brancoi is known from the type locality of Salmendingen, Germany, and also from the large sample from Rudabánya, Hungary. According to Begun (2002) isolated molars that may belong to this species are also known from Melchingen, Ebingen,

Trochtelfingen and Wissberg (Germany), Mariathal (Austria) and Udabno (Georgia).

Unfortunately no faunal list for Ebingen could be located and the list for Trochtelfingen includes only the primate occurrence. In chapter 3, these localities lacked any clear association with each other and upon further comparison of their component faunas to

Rudabánya (due to its taxonomic diversity), each had very few taxa in common. This is perhaps due in part to the lack of species richness at these localities (i.e., only Wissberg has

215 an appreciable number of taxa [18] in comparison to the 72 taxa at Rudabánya). Therefore, the lack of faunal associations between these localities prevents any conclusions regarding the taxonomic identity of the specimens from Germany, Austria and particularly Georgia.

However, with regard to the latter locality, the late Astaracian marks the dispersal of several

European taxa into Eastern Europe/Western Asia, including Protictitherium cingulatum,

Triceromeryx, Desmanella and Desmanodon (MN7/8). It is therefore possible that the later primate occurrence at Udabno 2 in MN11 (Gabunia et al. 2001) is the result of this earlier dispersal.

The relations of Dryopithecus to living African apes have been suggested by some researchers (Begun 2005, 2002, 2001; Begun et al. 1997; Stewart & Disotell 1998).

Interestingly, two intercontinental dispersal events occur during the temporal range of

Dryopithecus, either or both of which certainly could have involved this taxon. First, the

Astaracian (MN6-MN7/8) marks the appearance of Protictitherium, Hypsodontus,

Protragocerus, Protoryx solignaci, Percrocuta and possibly Machairodus in Africa.

Although the first three of these taxa have Asian origins, they are known from Eurasia before they appear in Africa. Similarly, Percrocuta also has Asian origins but follows the same pattern as the previous three taxa. The origins of Machairodus is uncertain, however this taxon is known from Europe (MN7/8-MN9) slightly before it is known in Africa (MN9).

The early Vallesian marks a second significant and diverse dispersal to Africa (Made 1999), which included a founding population of hipparionine derived from Eurasian

Hippotherium primigenium (Bernor & Harris 2003), tetraconodonine suids such as

Conohyus (represented in Africa as Nyanzachoerus) (Harris & Leakey 2003), and possibly

Sivachoerus, Progonomys, Myocricetodon and Atlantoxerus. According to Bernor & Harris

216

(2003), the first occurrence of hipparionines in Africa is at Chorora between 10.5 and 9.3Ma.

The eastern range extension of Dryopithecus into Georgia, together with two dispersal events, both including numerous other mammalian groups, indicates that Dryopithecus or a close relative most certainly had the opportunity to disperse along with these other groups to

Africa.

In sum, the patterns in faunal dynamics during the middle and late Miocene more fully support the hypothesis that Dryopithecus evolved from a taxon previously known in

Central Europe, rather than arising from a dispersal into the region. Pierolapithecus is most likely a junior subjective synonym of Dryopithecus, although this analysis was unable to determine whether this taxon is conspecific with D. fontani. Trends in the Spanish faunas do not support a dispersal of an additional species, D. crusafonti, into the region, however the possible occurrence of in situ evolution in other mammalian groups does not preclude the evolution of this species from previously occurring Dryopithecus. This analysis was unable to provide further insight into the species allocation of isolated molars from Melchingen,

Ebingen, Trochtelfingen and Wissberg (Germany), Mariathal (Austria) and Udabno

(Georgia), due to a lack of correlation between faunas at these localities. However, it is possible that the latter specimen is the result of an eastward migration of Central European

Dryopithecus in the late Astaracian, together with other large and small mammals. It is also possible that Dryopithecus or a close relative dispersed to Africa during a significant migration event in the late Astaracian and/or earliest Vallesian.

Ouranopithecus

217

Ouranopithecus macedoniensis was originally recognized from the late Vallesian of

Greece. More recently, a new species was identified in the Turolian of Turkey, O. turkae

(Güleç et al. 2007). An isolated P4 from the Chirpan District of Bulgaria at ~7Ma is also considered to resemble Ouranopithecus (Spassov & Geraads 2008) (Figure 4.5). According to Bernor et al. (1996b), the first appearance of Ouranopithecus is likely to be the result of an

African immigration or alternatively a vicariant lineage of MN8-MN9 Dryopithecus. The mandible attributed to freybergi has also been considered to be synonomous with Ouranopithecus by some (Andrews et al. 1996; Made 1999; Martin & Andrews 1984;

Pilbeam 1996; Szalay & Delson 1979), while others consider features of the corpus and molar dentition to signify a separate taxon (i.e., Begun 2002; Koufos & de Bonis 2004,

2005). The evolutionary relations of this taxon are contentious. de Bonis & Koufos (2004, p257) consider that “if Ouranopithecus occurred in Africa, it would even be a plausible ancestor for ,” due to similarities with the latter, which they consider to be synapomorphies (de Bonis & Koufos 1997). Others, however, consider these similarities to be the result of parallel evolution in response to environmental conditions favouring hard object consumption. Recently, two large bodied hominoids were described from the late

Miocene of Africa, nakayamai and abyssinicus, of which the morphology of the former has been likened to Ouranopithecus (Kunimatsu et al. 2007).

Through the previous analysis of the non-primate mammalian faunas from these localities, the following questions will be addressed: is Ouranopithecus the result of an African immigration or a vicariant lineage of earlier Dryopithecus; do faunal differences exist between Pyrgos Vassilissis and the other Greek localities to suggest that Graecopithecus and

218

Ouranopithecus are distinct genera; and lastly, what, if any, relationship does

Ouranopithecus have with late Miocene African species?

Figure 4.5: Distribution of Ouranopithecus localities

B RP X C N

P

RP Ravin de la Pluie N Nikiti 1 X Xirochori 1 P Pyrgos Vassilissis B Chirpan, Bulgaria C Çorakyerler

The Greek and Turkish Ouranopithecus localities share 25% of their large mammal faunas and also share taxa in common with Iran, Iraq, and to a lesser extent, Bulgaria.

Establishing a biogeographic link between these faunas and those occurring earlier in Europe is complicated by the Vallesian Crisis, such that almost all of the forest-dwelling taxa that characterize the primate localities in MN9 have gone extinct by the late Vallesian (Agustí et al. 2003). However, the giraffid, Decennatherium?, from Ravin de la Pluie and

Tragoportax from Nikiti 1 have close affinities with the Spanish Vallesian species (de Bonis

219

& Koufos 1999). In addition, the MN9 primate localities of Can Llobateres and Rudabánya share a number of large mammal genera and species in common with later MN10 and MN11

Greek and Turkish localities, including Aceratherium incisivum, Chalicotherium goldfussi,

Deinotherium giganteum (Turkey only), Dihoplus schleiemacheri (Turkey only),

Hippotherium primigenium, Tetralophodon longirostris, Protictitherium crassum,

Dorcatherium, Indarctos (Turkey only) and Propotamochoerus, although these taxa are all fairly cosmopolitan. Can Llobateres also shares Keramidomys with Çorakyerler. With the presence of Dryopithecus in Georgia, it is therefore possible that the occurrence of

Ouranopithecus represent a southern excursion in the eastward range extension of

Dryopithecus. Such a speciation event, however, would require considerable morphological change in craniodental anatomy, and particularly the dentition, from Dryopithecus to

Ouranpithecus.

The MN10 and MN11 Greek and Turkish localities also share large mammal genera with the localities of Chorora, Ethiopia (10.5-10.0Ma) and Nakali, Kenya (9.88-9.80Ma), including Choerlophodon, Palaeotragus, Hipparion, Samotherium, Deinotherium and

Dorcatherium. All of these taxa are known previously in both regions, however Leakey et al. (1996) consider Palaeotragus to be a Eurasian immigrant in the later Lothagam faunas and one of the earliest occurrences of a giraffine in Africa (although this taxon is preserved several million years earlier at Samburu, 9.5Ma, as well). Made (1999), on the other hand, proposed that this taxon immigrated to Africa from Asia as early as 16.5Ma (what he refers to as Giraffokeryx/Palaeotragus). In their description of Nakalipithecus, Kunimatsu et al.

(2007) indicate that if the dental similarities between Nakalipithecus and Ouranopithecus are the result of a close evolutionary relationship and not of convergence, then the direction of

220 dispersal was from Africa to southeastern Europe, in order to accommodate the temporal disparity between Nakali (latest early Vallesian) and the earliest occurrence of

Ouranopithecus (late Vallesian, MN10, 9.7-8.7Ma), as well as the more primitive dental characteristics (more strongly developed lingual and buccal cingula) in Nakalipithecus.

Made (1999) identified two Vallesian intercontinental dispersals between Africa and Europe, first at 11Ma, due to regression in sea level, and again at 9.6Ma. Both events involve very few taxa and the latter postdates the first occurrence of Ouranopithecus. A large bodied hominoid from Nakali therefore definitely had the opportunity to disperse to Europe 11Ma ago or earlier. The question of whether the first appearance of Ouranopithecus represents a southeastern lineage related to Dryopithecus or whether this taxon is the result of an African dispersal therefore remains unclear, since the mammalian faunas lend support to both alternatives. Spassov and Geraads (2008) consider Nakalipithecus to be morphologically more similar to Ouranopithecus than to any other fossil hominoid. If this is true and

Ouranopithecus is more similar to Nakalipithecus than it is to Dryopithecus, then such a finding would support the occurrence of Ouranopithecus as the result of an intercontinental dispersal, in addition to supporting an evolutionary relationship between the two.

The Graecopithecus locality of Pyrgos Vassilissis is not species rich and shares only

7% (two taxa) of its component large mammal fauna with other MN10 Ouranopithecus localities in Greece (Ravin de la Pluie, Xirochori 1, Nikiti 1) and the MN11 locality of

Çorakyerler in Turkey. Furthermore, there appears to be a considerable temporal gap between Ouranopithecus from Greece (MN10) and Pyrgos (MN12), however Koufos & de

Bonis (2005) note that age of the latter locality is far from concrete due to its faunal paucity.

Andrews et al. (1996) cite von Koenigswald (1972) in providing a possible date of MN10.

221

Therefore, if the faunal and temporal differences between the Greek primate localities are real and not an artifact of poor preservation, this would lend support to a taxonomic distinction between Graecopithecus and Ouranopithecus.

The temporal and geographic range of Ouranopithecus has expanded very recently and these findings provide intriguing considerations for the evolutionary relations of this taxon, particularly with respect to subsequent African taxa. What is most significant about

Ouranopithecus is that until these new discoveries in Turkey and possibly Bulgaria, it was considered that, with the exception of Oreopithecus, late Miocene Eurasian apes had gone extinct by the end of the Vallesian. Instead, the three individuals from Çorakyerler, Turkey

(MN11), and specimen from Bulgaria (~7Ma) extend the temporal range of European and

Western Eurasian apes considerably past the Vallesian Crisis and well into the Turolian.

Furthermore, and of utmost importance, all of these occurrences were found within the context of more open, savannah-like and seasonal environments. Made (1999) identified terminal Miocene intercontinental dispersals at 6.9Ma and 6.3Ma. In addition to sharing taxa with Nakali and Chorora, the Ouranopithecus localities also share taxa with the later occurring African hominin localities, including Tragoportax and Gazella (Middle Awash,

Ethiopia; Lothagam, Kenya), Ceratotherium (Middle Awash; Lukeino, Kenya) and

Prostrepsiceros (Middle Awash). Werdelin (2003) also recognizes a significant number of carnivore taxa within the 7-5Ma Lothagam (Kenya) deposits that are Eurasian in origin.

These include Amphicyonidae sp. A, Viverra and Ictitherium. Hyaenictitherium, and

Hyaenictis are known from both regions, however the more primitive species are Eurasian and the more derived taxa are African and likely descended from the former.

Lokotunjailurus and cf. Metailurus demonstrate a similar pattern. Lastly, Ekorus ekakeran

222 may have affinities with large European mustelids (Werdelin 2003). Leakey and Harris

(2003) further suggest that both the Leporidae and Muridae from Lothagam are the first known in the late Miocene of Africa and are evidently descended from Eurasian ancestors.

According to Leakey et al. (1996) this association of Eurasian faunal elements, together with

African taxa, is most clearly visible at Lothagam, although also occurs as far south as

Langebaanweg, South Africa. The tchadensis locality of Toros Menalla,

Chad, preserves several large mammals of Eurasian origin, including Machairodus cf. M. giganteus, Nyanzachoerus, Hyaenictitherium, Ictitherium and Hystrix. It is therefore evident that the Greek and Turkish Ouranopithecus localities share an appreciable number of large mammals with localities in Africa. Furthermore, many African late Miocene hominin localities include taxa of Eurasian origin in their faunal assemblages. Benammi et al. (1996) have also noted that many Turolian and early Messinian fossiliferous localities in north

Africa preserve small mammal species which have affinities to taxa found in southwestern

Europe. These localities include Aïn Guettara in Morocco, La Voie Ferrée in Tunisia,

Argoub Kemallal 1 in Algeria and Afoud (1, 2, 5 & 8) in Morocco. Based on the chronology of first appearances, these authors suggest that the faunal interchanges occurred prior to the

Messinian Salinity Crisis (Benammi et al. 1996)

In sum, the faunal composition East European/West Asian province does not clarify whether the first appearance of Ouranopithecus was the result of speciation from a

Dryopithecus ancestor or a separate intercontinental immigration. However, further analysis of Nakalipithecus may indicate an ancestor-descendant relationship between the two. The lack of overlap in fauna, together with the temporal gap between the Greek localities and

Pyrgos Vassilissis lend support to the idea that Graecopithecus is a distinct taxon, rather than

223 being synonomous with Ouranopithecus. However, the Graecopithecus locality is significantly lacking in comparable faunal completeness and diversity. Most importantly, new discoveries of Ouranopithecus in Turkey and potentially Bulgaria extend the temporal and geographic range of this taxon well past the Vallesian Crisis and into the Turolian in paleoenvironments not associated with densely forested subtropical conditions. The

Ouranopithecus localities also share many large mammal taxa in common with African hominin localities, indicating significant faunal interchange in the late Miocene.

Summary & Conclusions

From the study of faunal provinciality in the preceding chapter, the following hominoid dispersals and biogeographic relations are well supported, based on the current state of the fossil record:

o After the first appearance of Griphopithecus (cf. Griphopithecus) at Engelswies,

Germany, in MN5, this taxon most likely extended its range eastward and appears

slightly later in MN5 in Turkey. MN6 G. darwini from the Pannonian Basin is

biogeographically linked to cf. Griphopithecus from Engelswies, more so than to G.

alpani from Turkey.

o The first appearance of Dryopithecus is most likely the result of in situ evolution

from a Central European ancestor (Griphopithecus). Pierolapithecus is congeneric

with Dryopithecus from Spain. Faunal analysis was unable to clarify the association

among the isolated dental remains from Central Europe that are tentatively attributed

224

to Dryopithecus. There is considerable evidence for dispersal of land mammals,

including forest-dwellers, to Africa at the top of MN7/8 and/or earliest MN9.

o Faunal analysis was not able to clarify whether Ouranopithecus is a descendant of

Dryopithecus or an African immigrant, however, more detailed comparison with

Nakalipithecus may indicate that these taxa share a close evolutionary relationship.

Graecopithecus is a distinct taxon. New discoveries in Turkey and Bulgaria extend

the temporal and geographic range Ouranopithecus into the Turolian. These

localities collectively share many large mammal taxa in common with African late

Miocene hominin localities.

o The analysis of dispersals of non-primate mammals provided some interesting

insights into the biogeographic patterns among Eurasian Miocene apes, however the

results varied according to the quality of the fossil record. Directionality of dispersal

events in non-primate mammals were clarified, however, in most cases, the results

produced more than one alternative for the ape associated with that temporal interval

and geographic location. Regardless, this study refines what is currently known about

the biogeographic relationships among Eurasian Miocene apes. Furthermore, in

conjunction with phylogenetic analyses, the results of this study will contribute to our

understanding of the evolutionary relations among Eurasian Miocene apes.

Chapter 5 - Conclusions

Overview of study

The origin and diversification of great apes and humans is among the most researched and debated series of events in the evolutionary history of the order Primates. A fundamental part of understanding these events involves reconstructing paleoenvironmental and paleogeographic patterns in the Eurasian Miocene; a time period and geographic expanse rich in evidence of lineage origins and dispersals of numerous mammals, including apes. The purpose of this study is to clarify the paleobiogeographic relations among Eurasian Miocene apes and to test the hypothesis that all African hominines are descended from a Eurasian ancestor. To this end, the patterns of distribution, in situ evolution, interprovincial and intercontinental dispersal of non-primate terrestrial mammals were used as a framework to elucidate these same patterns in Eurasian apes.

In chapter 2, I test for sampling completeness in middle and late Miocene large and small mammals to identify taxa and temporal intervals affected by sampling bias. Large mammals were found to be relatively complete across all temporal intervals, while small mammals were sensitive to the stringency of the index used and were thus incomplete in

MN6 and MN12 using the most conservative measures. The analysis presented in this chapter differs from previous studies of completeness in using a larger dataset and explicitly identifying genera with incomplete temporal ranges.

In chapter 3, I use cluster analysis to explore whether Eurasian Miocene large and small mammals differentiate into identifiable faunal bioprovinces. The results of this analysis support the recognition of a Central European and East European/West Asian province as found in previous studies, however, both the Spanish and Italian faunas were

225 226

considered distinct enough to warrant inclusion into their own exclusive zoogeographic regions. The temporal disparity in sampling completeness and the proportion of cosmopolitan to endemic taxa during some intervals were found to affect the exclusivity of the bioprovinces. The analysis presented in this chapter differs from previous studies again in using a larger, more recent dataset and providing a thorough explanation of the statistical strength of the individual bioprovinces recognized in each temporal interval.

In chapter 4, I draw on the findings of the previous chapters to clarify patterns of biogeographic association among middle and late Miocene Eurasian apes and their relation to contemporaneous and subsequent African hominines. Using the patterns of distribution, in situ evolution and dispersal of non-primate large and small mammals, I present alternatives for the chronology of first appearances and biogeographic links among Eurasian apes and between these taxa and their African counterparts. It is important to note that in all cases, more than one alternative was apparent, however, the explanation supported by the majority of non-primate fauna was favoured. In some cases, the favoured alternative relied on very few taxa. Therefore, these results are certainly subject to interpretation and could indeed be falsified (or further supported) by future faunal analysis. Most importantly, the findings presented in chapter 4 contribute a finer degree of resolution to what is known of Eurasian hominid paleobiogeography.

Paleobiogeographic relations of Eurasian apes

Based on the current state of the Eurasian hominoid fossil record, the results of this study suggest that the FA of Griphopithecus in Central Europe in early MN5 is most likely the result of dispersal from an Afropithecus-like ancestor. Griphopithecus appears in Turkey slightly later in MN5, most likely as an eastward range extension of a Central European

227 taxon. Taphonomic bias in the MN5 Turkish faunas is clear and therefore these results may be subject to change in light of new discoveries. The occurrence of G. darwini in the

Pannonian Basin is most likely the result of in situ evolution from a taxon already present in the Central European province, rather than a separate dispersal into the region.

Similarly, the first appearance of Dryopithecus in the Pannonian Basin is also likely the result of in situ evolution from a taxon known previously from Central Europe. The recently discovered Pierolapithecus appears to be congeneric with Spanish Dryopithecus.

This study was unable to clarify the nature of the first appearance of Ouranopithecus and specifically whether this taxon is the result of a southern excursion of the Dryopithecus lineage or alternatively, the result of an African immigration. However, further comparison with Nakalipithecus may clarify the relationship of this taxon to the Greek Ouranopithecus.

New discoveries in Turkey and Bulgaria extend the temporal and geographic range of

Ouranopithecus well into the more open paleoenvironments of the Turolian. This has significant implications for our understandings of the context and timing of the extinction of the Eurasian apes and will be discussed further.

Intercontinental relations

In 1871, considered Africa the most likely geographic place of origin of the last common ancestor of African apes and humans. However, Darwin also cautioned such a conclusion, due to the presence of the “anthropomorphous” apes in Europe, together with the likelihood of intercontinental exchange of terrestrial mammals during the Miocene.

Since Darwin’s time, opinions on this topic have become polarized into two mutually exclusive alternatives: an African origin, where all Eurasian apes bear no relation to living

African apes and humans, and a Eurasian origin, where a descendant of Dryopithecus or

228

Ouranopithecus dispersed back to Africa in the late Miocene. Late Miocene intercontinental faunal exchange is well established (i.e., Made 1999; Bernor et al. 1996b; Leakey & Harris

2003; Leakey et al. 1996; Werdelin 2003), but despite this, many researchers are adamantly opposed to the idea of a Eurasian origin for the Homininae. This is largely due to a lack of consensus regarding the phylogenetic relations of middle and late Miocene apes and their relations to living taxa (Andrews 1992a; Andrews & Bernor 1999; Andrews et al. 1996;

Andrews & Martin 1987; Begun 1995, 1994, 1992a; Begun & Kordos 1997; Begun et al.

1997; Harrison & Rook 1997; Martin & Andrews 1993; Moyà-Solà et al. 2004; Moyà-Solà

& Köhler 1993; Pilbeam 1997, 1996; Pilbeam & Young 2004; Ward et al. 1997), as well as the opinion that no Eurasian ape conforms to the ancestral hominine morphotype of a suspensory, frugivorous, tropical forest-dweller, most similar to Pan (Cote 2004; Pilbeam

1997, 1989; Pilbeam & Young 2004). In addition, it is claimed that Eurasia fails to preserve the tropical/subtropical conditions along possible dispersal routes to African in the late

Miocene (Andrews 2007; Cote 2004; Pilbeam & Young 2004).

This study, together with previous research, has identified two and as many as three episodes of faunal interchange between Eurasia and Africa, beginning in the Astaracian.

Some taxa involved in these interchanges are reconstructed as arboreal frugivores (i.e.,

Protictitherium, Werdelin & Solounias 1991), some are known from subtropical to warm temperate woodland environments (i.e., Hippotherium, Bernor et al. 1996c, Bernor &

Armour-Chelu 1999), or closed marshy forests (i.e., Conohyus, Fortelius, Made & Bernor

1996, Hünermann 1999, Thenius 1952), while some are known from mixed forest conditions

(Amphicyonidae, Palaeotragus, Gentry et al. 1999; Ginsberg 1999, Werdelin 1996). These are all ape-friendly environments. A number of the taxa involved in these interchanges lack

229 clear environmental signals, particularly the carnivores, but are commonly found in association with Eurasian apes (i.e., Percrocuta, Machairodus, Viverra, Ictitherium,

Hyaenictitherium, Metailurus, and the murid, Progonomys).

Although these episodes of faunal exchange do not prove an associated intercontinental dispersal of a fossil ape, they do clarify the timing of movement patterns and importantly demonstrate a definitive opportunity for movement back to Africa, along with taxa having similar niche requirements and/or common associations with Eurasian fossil apes. This study cannot identify whether a late Miocene Eurasian hominid (or descendant thereof) founded the African ape and human lineage, but the results suggests that perhaps the two competing hypotheses surrounding this event are not as mutually exclusive as previously thought. Late Miocene Eurasian apes certainly had the potential and opportunity to disperse to Africa, however, it is seems equally likely that they or perhaps an undiscovered African taxon founded the Homininae.

As noted, Eurasian hominids are commonly excluded from African ape and human ancestry due to traditional conceptions of which fossils conform to the ancestral hominine morphotype. Specifically, Cote (2004), Pilbeam (1997, 1989) and Pilbeam & Young (2004) specify that no Eurasian ape fits the model of a large-bodied, suspensory , most similar to Pan and dependent on tropical rainforest habitats. This conclusion is inconsistent with reconstructions of the functional anatomy of Eurasian fossil apes and their paleoenvironments. First, all Eurasian apes are large bodied and many preserve anatomy indicative of suspension, such as Dryopithecus, Pierolapithecus, Oreopithecus and

Lufengpithecus (Andrews & Harrison 2005; Begun 2007, 2002, 1994, 1992a; Begun et al.

2006b; Begun & Ward 2005; Deane & Begun 2008, Harrison & Rook 1997; Harrison &

230

Rook 1997; Moya-Sola & Kohler 1996; Sarmiento 1987; Schwartz 1997; Susman 1985).

Some Eurasian taxa, such as Dryopithecus, have even been reconstructed as frugivorous

(Deane 2007; Kay & Ungar 1997; Ungar 2004, 1998, 1996; Ungar & Kay 1995).

Although Dryopithecus is known from subtropical swamp-forest conditions (Andrews et al. 1997; Kordos & Begun 2002), some have argued that the increasingly seasonal, more open conditions of late Miocene Eurasia would not have provided hospitable habitats along dispersal corridors to Africa. If this taxon dispersed prior to the Vallesian Crisis, the argument that forested conditions did not exist is irrelevant since these conditions are known to have been widespread at that time. Following the Vallesian Crisis, many studies have documented the shift in vegetation structure from pre-Vallesian subtropical closed forest conditions to more open deciduous woodlands in the late Vallesian and Turolian (Agustí et al. 2003; Agustí et al. 1997; Ivanov et al. 2003; Potts 2004a & b; Potts & Behrensmeyer

1992; Solounias et al. 1999; van Dam & Weltje 1999). However, the presence of seemingly isolated pockets of closed forest refugia is also well known (i.e., Dorn-Dürkheim, Germany,

Franzen 1997, Franzen & Storch 1999; Udabno, Georgia, Gabunia et al. 1999). A recent study from Turkey has documented humid, forested landscapes in the latest Vallesian, extending along the Aegean Coast, possibly as far south as Crete (Geraads et al. 2005).

Therefore, despite the general shift in climate and vegetation structure, forested conditions are known to have occurred in areas considered as “crossroads” for mammalian dispersal

(Turkey, Georgia). Conversely, past and recent discoveries of Ouranopithecus have demonstrated the persistence of Eurasian apes considerably beyond the Vallesian Crisis and their expansion into more open habitats, clearly demonstrating the capacity to survive in open conditions.

231

The locality-specific nature and subsequent geographic discontinuity of paleoecological reconstruction must also be considered. Many regions where exchange corridors most likely existed are either unevenly sampled in time and space or unsampled altogether in terms of their paleoecology (i.e., Saudi Arabia). This, coupled with the lack of reliability of most quantitative paleoclimatological and paleoecological studies, indicates that little is known regarding the spatial distribution of forested to open conditions in the late

Vallesian and Turolian.

The strong emphasis on tropical rainforest conditions as a necessity for dispersal to

Africa is exaggerated. Although all African apes are known to reside in dense tropical forests, inhabit a spectrum of different ecotypes. They are not confined exclusively to rainforests and are known to inhabit woodlands, grasslands and wooded grasslands (Bogart & Pruetz 2008; Pruetz 2007; Russak & McGrew 2008; Sponheimer et al.

2006). Furthermore, although hominines, and specifically Pan, demonstrate a clear preference for soft fruit consumption, reliance on fallback foods in times of food scarcity is documented in many populations (Hladik 1977; Stanford et al. 1994; Tutin et al. 1997;

Yamakoshi 1998). Therefore, dismissing the dispersal of Eurasian apes into Africa in the late

Miocene on the basis of a lack of required habitats is clearly unwise.

Nargolwalla & Begun (2005) noted an interesting trend in a significant number of large and small terrestrial mammals that persist beyond the Vallesian crisis. Many small and large mammal taxa were characterized by an increase in body size and/or dental complexity, in response to the shift in ecological conditions. Therefore, perhaps thick enamel and megadonia were key adaptations that allowed a descendent of a Eurasian ape to disperse back to Africa in the late Miocene, in much the same way that these adaptations perhaps allowed

232

Griphopithecus to disperse and radiate in Eurasia in the early middle Miocene (Begun 2002).

Although the circumstances surrounding the origins of the Homininae remain unclear, the late Miocene fossil record in Africa and Eurasia continues to develop. Survey and excavation in the Black Sea region, Eastern Mediterranean, Saudi Arabia and North Africa will undoubtedly unearth new fossils and further opportunity for more spatially and temporally continuous paleoenvironmental reconstruction along intercontinental dispersal crossroads. These, combined with advances in analytical methods promise fascinating insights into questions of African ape and human origins.

References Cited

Aguilar J-P, WA Berggren, M-P Aubry, DV Kent, G Clauzon, M Benammi & J Michaux. 2004. Mid-Neogene Mediterranean marine-continental correlations: an alternative interpretation. Palaeogeogr Palaeoclimatol Palaeoecol 24:165-186.

Agustí J. 1999. A critical re-evaluation of the Miocene mammal units in Western Europe: dispersal events and problems of correlation. In: Agustí J, L Rook & P Andrews, (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p84-112.

Agustí J, M Garcés & W Krijgsman. 2006. Evidence for African-Iberian exchanges during the Messinian in the Spanish mammalian record. Palaeogeogr Palaeoclimatol Palaeoecol 238:5-14.

Agustí J, I Casanovas-Vilar & M Furió. 2005. Rodents, insectivores and chiropterans (Mammalia) from the late Aragonian of Can Missert (Middle Miocene, Vallès- Penedès, Spain). Geobios 38: 575-583.

Agustí J, A Sanz de Siria & M Garcés. 2003. Explaining the end of the hominoid experiment in Europe. J Hum Evol 45:145-153.

Agustí J & I Casanovas-Vilar. 2003. Neogene gerbils from Europe. In: Reumer JWF & W Wessels (Eds) Distribution and Migration of Tertiary Mammals in Eurasia. A Volume in Honour of Hans DeBruijn. Deinsea 10:13-22.

Agustí J & M Antón. 2002. Mammoths, sabertooths, and hominids: 65 million years of mammalian evolution in Europe. New York: Columbia University Press.

Agustí J, L Cabrera, M Garcés, W Krijgsman, O Oms & JM Parés. 2001a. A calibrated mammal scale for the Neogene of Western Europe. State of the art. Earth-Sci Rev 52:247-260.

Agustí J, L Cabrera & M Garcés. 2001b. Chronology and zoogeography of the Miocene hominoid record in Europe. In: Bonis L de, GD Koufos & P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 2: Phylogeny of the Neogene Hominoid Primates of Eurasia. Cambridge: Cambridge University Press, p2-18.

Agustí J, L Cabrera, M Garcés & M Llenas. 1999a. Mammal turnover and global climate change in the late Miocene terrestrial record of the Vallès-Penedès basin. In: Agustí J, L Rook & P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p397-412.

233 234

Agustí J, L Cabrera, M Garcés & JM Parés, 1997. The Vallesian mammal succession in the Vallès-Penedès basin (northeast Spain): Paleomagnetic calibration and correlation with global events. Palaeogeogr Palaeoclimatol Palaeoecol 133:148-180.

Allen KMS. 1990. Interpreting Space. In: Allen KMS, SW Green & EBW Zubrow (Eds) Interpreting Space: GIS and Archaeology. New York: Taylor and Francis, p383-386.

Alroy J. 1994. Appearance event ordination: a new biochronological method. Paleobiology 20:191-207.

Alroy J, RL Bernor, M Fortelius & L Werdelin. 1998. The MN System: Regional or Continental? Mitt Bayer Staatsslg Paläont hist Geol 38: 243-258.

Andrews P. 2007. The biogeography of hominid evolution. J Biogeogr 34:381-382.

Andrews P. 1996. Palaeoecology and Hominoid Palaeoenvironments. Biological Reviews 71:257-300.

Andrews P. 1995. Time resolution in the Miocene fauna from Paşalar. J Hum Evol 28:343- 358.

Andrews P. 1992a. Evolution and environment in the Hominoidea. Nature 360:641-646.

Andrews P. 1992b. An ape from the south. Nature 356:106.

Andrews P. 1990. Owls, Caves and Fossils: Predation, Preservation, and Accumulation of Small Mammal Bones in Caves, with an Analysis of the Cave Faunas from Westbury-sub-Mendip, Somerset, UK. London: British Museum of Natural History.

Andrews P & J Kelley. 2007. Middle Miocene dispersals of apes. Folia Primatol 78:328- 343.

Andrews P & T Harrison. 2005. The Last Common Ancestor of Apes and Humans. In: Lieberman DE, RJ Smith & J Kelley (Eds) Interpreting the Past: Essays on Human, Primate, and Mammal Evolution. Boston: Brill Academic Publishers, Inc, p103-121.

Andrews P & RL Bernor. 1999. Vicariance Biogeography and Paleoecology of Eurasian Miocene Hominoid Primates. In: Agustí J, L Rook & P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p454- 487.

Andrews P, DR Begun & M Zylstra. 1997. Paleoecology of Miocene hominoids. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny and Fossils: Miocene Hominoid Origins and Adaptations. New York: Plenum Press, p29-58.

235

Andrews P, T Harrison, E Delson, RL Bernor & L Martin. 1996. Distribution and Biochronology of European and Southwest Asian Miocene Catarrhines. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p168-207.

Andrews P & LB Martin. 1987. Cladistic relationships of extant and fossil hominoids. J Hum Evol 16:101-118.

Andrews P, WR Hamilton & PJ Whybrow. 1978. Dryopithecines from the Miocene of Saudi Arabia. Nature 274:249-250.

Antunes MT. 1979. Hispanotherium fauna in Iberian Middle Miocene, its importance and paleogeographical meaning. Ann Géol Pays Hellén 1:19-26.

Avery DM. 2003. Early and middle Pleistocene environments and hominid biogeography: micromammalian evidence from Kabwe, Twin Rivers and Mumbwa Caves in central Zambia. Palaeogeogr Palaeoclimatol Palaeoecol 189:55-69.

Badgley C. 1990. A statistical assessment of last appearances in the Eocene record of mammals. In: Bown TM & KD Rose (Eds) Dawn of the Age of Mammals in the northern part of the Rocky Mountain Interior. Geological Society of America Special Papers 243:153-167.

Badgley C & PD Gingerich. 1988. Sampling and faunal turnover in early Eocene mammals. Palaeogeogr Palaeoclimatol Palaeoecol 63:141-157.

Becker A. 2002. The Jura Mountains—an active foreland fold-and-thrust belt? Tectonophysics 321:381– 406.

Begun DR. 2007. Fossil record of Miocene hominoids. In: Henke W & I Tattersall (Eds) Handbook of Paleoanthropology, vol II: Primate Evolution and Human Origins. New York: Springer-Verlag, p921-977.

Begun DR. 2006. Revision of the Dryopithecini. J Vert Paleontol 26(S3):40A.

Begun DR. 2005. is east and Dryopithecus is west, and never the twain shall meet. Anthrop Sci 113:53-64.

Begun DR. 2002. European Hominoids. In: Hartwig WC (Ed) The Primate Fossil Record. Cambridge: Cambridge University Press, p339-368.

Begun DR. 2001. African and Eurasian Miocene hominoids and the origins of the . In: de Bonis L, GD Koufos & P Andrews (Eds) Hominoid Evolution and Climate Change in Europe, vol 2: Phylogeny of the Neogene Hominoid Primates of Eurasia. Cambridge: Cambridge University Press, p231-253.

236

Begun DR. 2002. European Hominoids. In: Hartwig WC (Ed) The Primate Fossil Record. Cambridge: Cambridge University Press, p339-368.

Begun DR. 2001. African and Eurasian Miocene hominoids and the origins of the Hominidae. In: de Bonis L, GD Koufos & P Andrews (Eds) Hominoid Evolution and Climate Change in Europe, vol 2: Phylogeny of the Neogene Hominoid Primates of Eurasia. Cambridge: Cambridge University Press, p231-253.

Begun DR. 1995. Late Miocene European orang-utans, , humans, or none of the above? J Hum Evol 29:169-180.

Begun DR. 1994. Relations Among the Great Apes and Humans: New Interpretations Based on the Fossil Great Ape Dryopithecus. Yearb phys Anthrop 37:11-63.

Begun DR. 1992a. Phyletic Diversity and Locomotion in Primitive European Hominids. Am J Phys Anthrop 87:311-340.

Begun DR. 1992b. Dryopithecus crusafonti sp. nov., a new Miocene hominoid species from Can Ponsic (Northeastern Spain). Am J phys Anthrop 87:291-307.

Begun DR. 1991. European Miocene catarrhine diversity. J Hum Evol 20:521-526.

Begun DR, MC Nargolwalla & L Kordos. 2008. Revision of the Dryopithicinae: Phylogenetic and paleobiogeographic implications. Am J Phys Anthrop 135(S46):66.

Begun DR, MC Nargolwalla & MP Hutchison. 2006a. Primate diversity in the Pannonian Basin: In situ evolution, dispersals, or both. Beiträge zur Paläontologie 30:43-56.

Begun DR, CV Ward, AS Deane, TL Kivell, MC Nargolwalla & ND Taylor. 2006b. Stem hominine or hominid? The phylogeny and functional anatomy of Pierolapithecus catalaunicus. Am J Phys Anthrop 129(S42).

Begun DR & CV Ward. 2005. Comment on “Pierolapithecus catalaunicus,” a New Middle Miocene Great Ape from Spain.” Science 308:203c.

Begun DR, E Güleç, D Geraads. 2003a. Dispersal patterns of Eurasian hominoids: implications from Turkey. In: Reumer JWF & W Wessels (Eds) Distribution and Migration of Tertiary Mammals in Eurasia. A Volume in Honour of Hans DeBruijn. Deinsea 10:23-39.

Begun DR, D Geraads & E Güleç. 2003b. The Çandır hominoid locality: Implications for the timing and pattern of hominoid dispersal events. In: Güleç E, DR Begun & D Geraads (Eds) and Vertebrate Paleontology of the Middle Miocene Hominoid Locality Çandır. Cour Forsch –Inst Senckenberg 240:251-265.

Begun DR, E Güleç, D Geraads. 2001. Dispersal pattern of Eurasian hominoids:

237

implications from Turkey. Distribution and Migration of Tertiary Mammals in Eurasia. The University of Utrecht, 17-19 May, 2001. p8-9.

Begun DR & L Kordos. 1997. Phyletic Affinities and Functional Convergence in Dryopithecus and Other Miocene and Living Hominids. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p291-316.

Begun DR, CV Ward, MD Rose. 1997. Events in Hominoid Evolution. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p389-416.

Behrensmeyer AK. 1991. Terrestrial Vertebrate Accumulations. In: Allison PA & DEG Briggs (Eds) Taphonomy: Releasing the Data Locked in the Fossil Record. New York: Plenum Press, p291-335.

Behrensmeyer AK, D Western & D Dechant-Boaz. 1979. New Perspectives in Vertebrate Paleoecology from a Recent Bone Assemblage. Paleobiology 5 (1):12-21.

Bellier E, P Monestiez, J-P Durbec & J-N Candau. 2007. Identifying spatial relationships at multiple scales. Ecography 30:385-399.

Benammmi M, M Calvo, M Prévot & J-J Jaeger. 1996. Magnetostratigraphy and paleontology of Aït Kandoula Basin (High Atlas, Morocco) and the Africa-European late Miocene terrestrial fauna exchanges. Earth Planet Sci Lett 145:15-29.

Bernor RL. 1983. Geochronology and Zoogeographic Relationships of Miocene Hominoidea. In: Ciochon RL & RS Corruccini (Eds) New Interpretations of Ape and Human Ancestry. New York: Plenum Press, p21-64.

Bernor RL. 1978. The Mammalian Systematics, Biostratigraphy and Biochronology of Maragheh and its Importance for Understanding Late Miocene Hominoid Zoogeography and Evolution (PhD thesis). University of , Los Angeles.

Bernor RL & JM Harris. 2003. Systematics and Evolutionary Biology of the Late Miocene and Early Pliocene Hipparionine Equids from Lothagam, Kenya. In: Leakey MG & JM Harris (Eds) Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, p388-438.

Bernor RL, RS Scott, M Fortelius, J Kappelman & S Sen. 2003. Equidae (Perissodactyla). In: Fortelius M, J Kappelman, S Sen & RL Bernor (Eds) Geology and Paleontology of the Miocene Sinap Formation, Turkey. New York: Columbia University Press, p220-281.

Bernor RL & M Armour-Chelu. 1999. Family Equidae. In: Rössner GE & K Heissig

238

(Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p193-202.

Bernor RL, V Fahlbusch, H-W Mittmann & S Rietschel. 1996a. The Evolution of Western Eurasian Neogene Mammal Faunas: The 1992 Schloss Reisensburg Workshop Concept. In: Bernor RL, V Fahlbusch, H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press. p1-4.

Bernor RL, V Fahlbusch, P Andrews, D de Bruijn, M Fortelius, F Rögl, FF Steininger & L Werdelin. 1996b. The Evolution of Western Eurasian Neogene Mammal Faunas: A Chronologic, Systematic, Biogeographic, and Paleoenvironmental Synthesis. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p449-469.

Bernor RL, GD Koufos, MO Woodburne & M Fortelius. 1996c. The Evolutionary History and Biochronology of European and Southwest Asian Late Miocene and Pliocene Hipparionine Horses. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p307-338.

Bernor RL, N Solounias, CC Swisher III & JA Van Couvering. 1996d. The Correlation of Three Classical “Pikermian” Mammal Faunas – Maragheh, Samos, and Pikermi – with the European MN Unit System. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p137-156.

Bernor RL, V Fahlbusch & S Rietschel. 1993. Evolution of Neogene continental biotopes in Central Europe and the eastern Mediterranean (15.5-5Ma). Immendingen and Schloss Reisensburn, 5-11th July, 1992. J Hum Evol. 24: 169-171.

Bernor RL & H Tobien. 1990. The mammalian geochronology and biogeography of Paşalar (Middle Miocene, Turkey). J Hum Evol 19:551-568.

Bicchi E, E Ferrero, M Gonera. 2003. Palaeoclimatic interpretation based on Middle Miocene planktonic foraminifera: the Silesia Basin (Paratethys) and Monferrato (Tethys) records. Palaeogeogr Palaeoclimatol Palaeoecol 196:265-303.

Bogart SL & JD Pruetz. 2008. Ecological context of savanna chimpanzees (Pan troglodytes verus) termite fishing at Fongoli, Senegal. Am J Primatol 70:605-612.

Böhme M. 2003a. The Miocene Climatic Optimum: evidence from ectothermic vertebrates of Central Europe. Palaeogeogr Palaeoclimatol Palaeoecol 195:389-401.

Böhme M. 2003b. Migration history of air-breathing fishes reveal Neogene atmospheric

239

circulation pattern. EEDEN – Environments and Ecosystem Dynamics of the Eurasian Neogene, Birth of the New World. Stará Lesná, Slovakia. November 12- 16th, 2003. p14.

Boillot G & R Capdevila. 1977. The Pyrenees: Subduction and collision. Earth Planet Sci Lett 35:151-160.

Bojar A-V, H Hiden, A Fenninger, F Neubauer. 2004. Middle Miocene seasonal temperature changes in the Styrian basin, Austria, as recorded by the isotopic composition of pectinid and brachiopod shells. Palaeogeogr Palaeoclimatol Palaeoecol 203:95-105.

Bolliger T. 1999. Family Anomalomyidae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p411-420.

Bolliger T. 1996. A Current Understanding about the Anomalomyidae (Rodentia): Reflections on Stratigraphy, Paleobiogeography, and Evolution. In: RL Bernor, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p235-245.

Boon-Kristkoiz E & AR Kristkoiz. 1999. Order Lagomorpha. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p259-262.

Boschetto HB, FH Brown & I McDougall. 1992. Stratigraphy of the Lothidok Range, northern Kenya, and K/Ar ages of its Miocene primates. J Hum Evol 22:47-71.

Bradley RS. 1999. Paleoclimatology (2nd Edition). Toronto: Academic Press, Ltd.

Brooks DR & DA McLennan. 2002. The Nature of Diversity: An Evolutionary Voyage of Discovery. Chicago: University of Chicago Press.

Bruijn H de. 1999. Superfamily Sciuroidea. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p271-280.

Bruch AA, D Uhl & V Mosbrugger. 2007. Miocene climate in Europe – Patterns and evolution – A first synthesis of NECLIME. Palaeogeogr Palaeoclimatol Palaeoecol 253:1-7.

Bruch AA, T Utescher, V Mosbrugger, I Gabrielyan & DA Ivanov. 2006. Late Miocene climate in the circum-Alpine realm – a quantitative analysis of terrestrial palaeofloras. Palaeogeogr Palaeoclimatol Palaeoecol 238:270-280.

Brusch AA. T Utescher, CA Olivares, N Dolakova, D Ivanov & V Mosbrugger. 2004. Middle and late Miocene spatial temperature patterns and gradients in Europe – preliminary results based on palaeobotanical climate reconstructions. Cour Forsch – Inst Senckenberg 249:15-27.

240

Clauzon G, J-P Suc, F Gautier, A Berger & M-F Loutre. 1996. Alternate interpretation of the Messinian Salinity Crisis: Controversy resolved. Geology 24:363-366.

Costeur L, S Montuire, S Legendre & O Maridet. 2007. The Messinian event: What happened to the peri-Mediterranean mammalian communities and local climate? Geobios 40:423-431.

Cote SM. 2004. Origins of the African hominoids: an assessment of the paleobiogeographical evidence. CR Palevol 3:321–338.

Csontos L. 1992. Tertiary tectonic evolution of the Intra-Carpathian area: a review. Acta Vulcanologica 7: 1-13.

Daams R. 1999. Family Gliridae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p301-318.

Daams R & AJ Van Der Meulen. 1984. Paleoenviromental and paleoclimatic interpretations of micromammal faunal successions in the Upper Oligocene and Miocene of North Central Spain. Paléobiologie continentale 14(2):241-257.

Darwin C. 1871. The Descent of Man, and Selection in Relation to Sex. London: John Murray.

Daxner-Höck G. 1999. Family Zapodidae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p337-342.

Deane AS. 2007. Inferring Dietary Behaviour for Miocene Hominoids: A High-Resolution Morphometric Approach to Incisal Crown Curvature (PhD Thesis), University of Toronto.

Deane AS, EP Kremer & DR Begun. 2008. Broken fingers: retesting locomotor hypotheses for fossil hominoids using fragmentary proximal phalanges and high-resolution polynomial curve fitting (HR-PCF). J Hum Evol. de Bonis L & GD Koufos. 1997. The Phylogenetic and Functional Implications of Ouranopithecus macedoniensis. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p317-326. de Bonis L & GD Koufos. 2004. Ouranopithecus et la date de separation des hominoids modernes. CR Palevol 3:255-262. de Bonis L & GD Koufos. 1999. The Miocene large mammal succession in Greece. In: Agustí J, L Rook & P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p205-237.

241 de Bruijn H, S Mayda, L vd Hoek Ostende, T Kaya & G Saraç. 2006. Small mammals from the Early Miocene of Sabuncubeli (Manisa, SW Anatolia, Turkey). Beiträge zur Paläontologie 30:57-87. de Bruijn H, L vd Hoek Ostende, E Kristkoiz-Boon, M Rummel, C Theocharopoulos & E Ünay. 2003. Rodents, lagomorphs and insectivores, from the middle Miocene hominoid locality of Çandır. In: Güleç E, DR Begun & D Geraads (Eds) Geology and Vertebrate Paleontology of the Middle Miocene Hominoid Locality Çandır. Cour Forsch –Inst Senckenberg 240:51-88. de Bruijn H, E Ünay & L vd Hoek Ostende. 1996. The Composition and Diversity of Small Mammal Associations from Anatolia Through the Miocene. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p266-270. de Bruijn H & E Ünay. 1996. On the Evolutionary History of the Cricetodontini from Europe and Asia Minor and Its Bearing on the Reconstruction of Migrations and the Continental Biotope during the Neogene. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p227-234.

Dèzes P, SM Schmid & PA Ziegler. 2004. Evolution of the European Cenozoic Rift System: interaction of the Alpine and Pyrenean orogens with their foreland lithosphere. Tectonophysics 389:1-33.

Dreyer SK. 1984. The Theory and Use of Methods for the Study of Mammalian Paleoecology (Ph.D. thesis). University College, London.

Durden CJ. 1974. Biomerization: An ecologic theory of provincial differentiation. In: CA Ross (Ed) Paleogeographic Provinces and Provinciality. Society of Economic Paleontologists and Mineralogists, Special publication No. 21:18-53.

Efremov JA. 1940. Taphonomy: A New Branch of Paleontology. Pan-America Geologist. 74 (2):81-93.

Engesser B. 1999. Family Eomyidae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p319-336.

Engesser B & R Ziegler. 1996. Didelphids, Insectivores, and Chiropterans from the Later Miocene of France, Central Europe, Greece, and Turkey. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p157-167.

Eronen J & L Rook. 2004. The Mio-Pliocene European primate fossil record: dynamics and habitat tracking. J Hum Evol 47:323-341.

242

Eronen J, L Rook & M Fortelius. 2003. Mammal community structure and primate dynamics during late Neogene. EEDEN – Environments and Ecosystem Dynamics of the Eurasian Neogene, Birth of the New World. Stará Lesná, Slovakia. November 12-16th, 2003. p17.

Esteban M. 1996. An overview of Miocene reefs from Mediterranean areas: general trends and facies models. In: Franseen EK, M Esteban, WC Ward & J-M Rouchy (Eds) Models for Carbonate Stratigraphy from Miocene Reef Complexes of Mediterranean Areas. Tulsa: Society for Sedimentary Geology, Concepts in Sedimentology and Paleontology 5:3-53.

Esteban M, JC Braga, J Martín, C De Santisteban. 1996. Western Mediterranean Reef Complexes. In: Franseen EK, M Esteban, WC Ward, J-M Rouchy (Eds) Models for Carbonate Stratigraphy from Miocene Reef Complexes of Mediterranean Areas. Tulsa: Society for Sedimentary Geology, Concepts in Sedimentology and Paleontology Volume 5, p55-72.

Estrada Belli F. 1999. The Archaeology of Complex Societies in South Eastern Pacific Coastal Guatemala: A Regional GIS Approach. England: British Archaeological Reports.

Esu D. 1999. Contribution to the knowledge of Neogene climatic changes in western and central Europe by means of non-marine mollusks Agustí J, L Rook & P Andrews (Eds) The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p328-354.

Fahlbusch V & P Mein. 1989. New Perspectives: The Past, the Present, and the Future. In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:625-628.

Fejfar O. 1999. Microtoid Cricetids. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p365-372.

Fejfar O, W-D Heinrich. 1989. Muroid Rodent Biochronology of the Neogene and Quaternary in Europe. In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:91-117.

Fernández-Jalvo Y. 1995. Small Taphonomy at La Trinchera de Atapuerca (Burgos, Spain): A Remarkable Example of Taphonomic Criteria Used for Stratigraphic Correlations and Palaeoenvironment Interpretations. Palaeogeogr Palaeoclimatol Palaeoecol 114 (2-4):167-195.

Fernández-Jalvo Y, C Denys, P Andrews, T Williams, Y Dauphin & L Humphrey. 1998. Taphonomy and palaeoecology of Olduvai Bed-I (Pleistocene, Tanzania). J Hum Evol 34:137-172.

243

Folinsbee KE & DR Brooks. 2007. Miocene hominoid biogeography: pulses of dispersal and differentiation. J Biogeogr 34:383-397.

Fortelius M. 1985. Ungulate cheek teeth: developmental, functional and evolutionary interrelations. Acta Zoologica Fennica 180:1-76.

Fortelius M, J Eronen, LP Liu, D Pushkina, A Tesakov, I Vislobokova & ZQ Zhang. 2003a. Continental-scale hysodonty patterns, climatic paleobiogeography and dispersal of Eurasian Neogene land mammal herbivores. In: Reumer JWF & W Wessels (Eds) Distribution and Migration of Tertiary Mammals in Eurasia. A Volume in Honour of Hans de Bruijn. Deinsea 10:1-11.

Fortelius M, J Kappelman, S Sen & RL Bernor. 2003b. Geology and Paleontology of the Miocene Sinap Formation, Turkey. New York: Columbia University Press,

Fortelius M, J van der Made & RL Bernor. 1996. Middle and Late Miocene Suoidea of Central Europe and the Eastern Mediterranean: Evolution, Biogeography, and Paleoecology. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p348-377.

Fortelius M, L Werdelin, P Andrews, RL Bernor, A Gentry, L Humphrey, H-W Mittmann & S Viranta. 1996a. Provinciality, Diversity, Turnover, and Paleoecology in Land Mammal Faunas of the Later Miocene of Western Eurasia. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p414-448.

Fortelius M, J Van der Made & RL Bernor. 1996b. Middle and Late Miocene Suoidea of Central Europe and the Eastern Mediterranean: Evolution, Biogeography, and Paleoecology. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press. p348-377.

Fortin M-J & M Dale. 2005. Spatial Analysis: A Guide for Ecologists. Cambridge: Cambridge University Press.

Frakes LA, JE Francis, JI Syktus. 1992. Climate Models of the Phanerozoic. Cambridge: Cambridge University Press.

Franzen JL. 1997. Die Säugetiere aus dem Turolium von Dorn-Dürkheim (Rheinhessen, Deutschland), Teil 1: Carnivora, Proboscidea (Tetralophodontidae), Perissodactyla (Rhinocerotidae, Equidae), Artiodactyla (Suidae). Cour Forsch –Inst Senckenberg, Band 197, Senckenbergische Naturforschende Gesellschaft, Frankfurt am Main.

Franzen JL & G Storch. 1999. Late Miocene mammals from Central Europe. In: Agustí J, L

244

Rook, P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p165-190.

Freudenthal M & E Martin Suárez. 1999. Family Muridae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p401-410.

Gabunia L, E Gabashvili, A Vekua & D Lordkipanidze. 2001. The late Miocene hominoid from Georgia. In: Bonis L de, GD Koufos & P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 2: Phylogeny of the Neogene Hominoid Primates of Eurasia. Cambridge: Cambridge University Press, p316-325.

Gebhardt H. 1999. Middle to Upper Miocene benthonic foraminiferal palaeoecology of the Tap Marls (Alicante Province, SE Spain) and its palaeoceanographic implications. Palaeogeogr Palaeoclimatol Palaeoecol 145:141-156.

Gentry AW, GE Rössner & EPJ Heizmann. 1999. Suborder Ruminantia. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p225-258.

Gentry A & EPJ Heizmann. 1996. Miocene of the Central and Eastern Paratethys. 1996. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p378-394.

Geraads D, T Kaya & S Mayda. 2005. Late Miocene large mammals from Yulaflı, Thrace region, Turkey, and their biogeographic implications. Acta Palaeontological Polonica 50:523-544.

Geraads D, E Güleç & T Kaya. 2002. Sinotragus (Bovidae, Mammalia) from Turkey, and the Late Miocene Middle Asiatic Province. Neues Jahrbuch für Geologie und Paläontologie Monatschefte 8:477-489.

Giese P, K-J Reutter, V Jacobshagen & R Nicolich. 1982. Explosion seismic crustal studies in the Alpine Mediterranean region and their implications to tectonic processes. In: Berckhemer H & K Hsü (Eds), Alpine-Mediterranean Geodynamics. Colorado: American Geophysical Union 7:39-74.

Ginsburg L. 1999. Order Carnivora. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p109-148.

Göhlich U. 1999. Order Proboscidea. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p157-168.

Golonka J. 2004. Plate tectonic evolution of the southern margin of Eurasia in the Mesozoic and Cenozoic. Tectonophysics 381:225-273.

245

Görür N, M Sakinç, A Barka, R Akkök. 1995. Miocene to Pliocene palaeogeographic evolution of Turkey and its surroundings. J Hum Evol 28:309-324.

Green SW. 1990. Approaching Archaeological Space. In: Allen KMS, SW Green, EBW Zubrow (Eds) Interpreting Space: GIS and Archaeology. New York: Taylor and Francis, p3-8.

Guérin C. 1989. Biozones or Mammal Units? Methods and Limits in Biochronology. In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:119-130.

Güleç E, A Sevim, C Pehlevan & F Kaya. 2007. A new great ape from the late Miocene of Turkey. Anthrop Sci 115:153-158.

Güleç E & DR Begun. 2003. Functional morphology and the affinities of the hominoid mandible from Çandır. Geology and Vertebrate Paleontology of the Middle Miocene Hominoid Locality Çandır (Central Anatolia, Turkey). Cour Forsch –Inst Senckenberg 240:89-112.

Hallam A. 1974. Changing patterns of provinciality and diversity of fossil in relation to plate tectonics. J Biogeog 1:213-225.

Hamilton WR, PJ Whybrow & HA McClure. 1978. Fauna of fossil mammals from the Miocene of Saudi Arabia. Nature 274:248-249.

Hammer Ø, DAT Harper & PD Ryan. 2007. PAST – Palaeotological Statistics, ver. 1.72. http://folk.uio.no/ohammer/past/

Harris JM & MG Leakey. 2003. Lothagam Suidae. In: Leakey MG & JM Harris (Eds), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, p485-519.

Harrison T. 2004. The zoogeographic and phylogenetic relationships of early catarrhine primates in Asia. Anthrop Sci 113:43-51.

Harrison T. 2002. Late Oligocene to middle Miocene catarrhines from Afro-Arabia. In: Hartwig WC (Ed) The Primate Fossil Record. Cambridge: Cambridge University Press, p311-338.

Harrison T. 1991. Some observations on the Miocene hominoids from Spain. J Hum Evol 20:515-520.

Harrison T & Y Gu. 1999. and phylogenetic relationships of early Miocene catarrhines from Sihong, China. J Hum Evol 37:225-277.

Harrison T & L Rook. 1997. Enigmatic anthropoid or misunderstood ape? The phylogenetic

246

status of Oreopithecus bambolii reconsidered. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptations. New York: Plenum, p327-362.

Harzhauser M, C Latal & WE Piller. 2007. The stable isotope archive of Lake Pannon as a mirror of Late Miocene climate change. Palaeogeogr Palaeoclimatol Palaeoecol 249:335-350.

Heikinheimo H, M Fortelius, J Eronen & H Mannila. 2007. Biogeography of European land mammals show environmentally distinct and spatially coherent clusters. J Biogeogr 34:1053-1064.

Heissig K. 1999a. Family Rhinocerotidae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p175-188.

Heissig K. 1999b. Family Chalicotheriidae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p189-192.

Heissig K. 1999c. Family Tubulidentata. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p155-156.

Heissig K. 1996. The Stratigraphical Range of Fossil in the Late Neogene of Europe and the Eastern Mediterranean. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p339-347.

Heizmann EPJ & DR Begun. 2001. The oldest Eurasian hominoid. J Hum Evol 41:463- 481.

Hladik CM. 1977. Chimpanzees of Gabon and chimpanzees of Gombe: Some comparative data in the diet. In: Clutton-Brock TH (Ed) Primate Ecology: Studies of Feeding and Ranging Behaviour in Lemurs, Monkeys and Apes. London: Academic Press, p481- 501.

Hoek Ostende L vd, M Morlo & D Nagel. 2006. Magestic Killers: the sabre-toothed . Geology Today 22:150-157.

Hsü KJ, L Montadert, D Bernoulli, MB Cita, A Erickson, RE Garrison, RB Kidd, F Mèliéres, C Müller & R Wright. 1977. History of the Mediterranean salinity crisis. Nature 267:399-404.

Hughes CP. 1973. Analysis of past faunal distributions. In: Tarling DH & SK Runcorn, (Eds) Implications of Continental Drift to the Earth Sciences, vol. 1. New York: Academic Press, p219-230.

Hugueney M. 1999. Family Castoridae. In: Rössner GE & K Heissig (Eds) The Miocene

247

Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p281-300.

Huismans RS, Bertotti G, D Ciulavu, CAE Sanders, S Cloetingh & C Dinu. 1997. Structural evolution of the Transylvanian Basin (Romania): A sedimentary basin in the bend zone of the Carpathians. Tectonophysics 272:249-268.

Humphrey L & P Andrews. 2008. Metric variation in the postcanine teeth from Paşalar, Turkey. J Hum Evol 54:503-517.

Hünermann KA. 1999. Superfamily Suoidea. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p209-216.

Ivanov D, AR Ashraf, V Mosbrugger & E Palamarev. 2003. Climate change about 10Ma in Paratethyan realm of Balkan Peninsula. EEDEN – Environments and Ecosystem Dynamics of the Eurasian Neogene, Birth of the New World. Stará Lesná, Slovakia. November 12-16th, 2003. p 48.

Ivanov D, AR Ashraf, V Mosbrugger & E Palamarev. 2002. Palynological evidence for Miocene climate change in the Forecarpathian Basin (Central Paratethys, NW Bulgaria). Palaeogeogr Palaeoclimatol Palaeoecol 178:19-37.

Janis C. 1988. An estimation of tooth volume and hypsodonty indices in ungulate mammals, and the correlation of these factors with dietary preferences. Mémoirs du Muséum national d’Histoire naturelle, Paris 53:367-387.

Janis CM & A Lister. 1985. The morphology of the lower fourth as a taxonomic character in the Ruminantia (Mammalia; Artiodactyla), and the systematic position of Triceromeryx. Journal of Paleontology 59:405-410.

Jernvall J & M Fortelius. 2004. Maintenance of trophic structure in fossil mammal communities: site occupancy and taxon resilience. The American Naturalist 164(5):614-624.

Johnson JG. 1971. A quantitative approach to faunal province analysis. Am J Sci 270:257- 280.

Jones RW. 1999. Marine invertebrate evidence for the palaeogeography of the Oligocene- Miocene of western Eurasia, and consequences for terrestrial vertebrate migration. In Agustí J, L Rook & P Andrews (Eds) The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p274-308.

Kälin D. 1999. Tribe Cricetini. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p373-388.

Kappelman J, S Sen, M Fortelius, A Duncan, B Alpagut, J Crabaugh, A Gentry, JP Lunkka,

248

F McDowell, N Solounias, S Viranta & L Werdelin. 1996. Chronology and Biostratigraphy of the Miocene Sinap Formation in Central Turkey. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p78-95.

Kay RF & PS Ungar. 1997. Dental evidence for diets in some Miocene catarrhines with comments on the effects of phylogeny on the interpretation of adaptation. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p131-151.

Kelley J. 2008. Identification of a single birth cohort in Kenyapithecus kizili and the nature of sympatry between K. kizili and Griphopithecus alpani at Paşalar. J Hum Evol 54:530-537.

Kelley J, P Andrews, B Alpagut. 2008. A new hominoid species from the middle Miocene site of Paşalar, Turkey. J Hum Evol 54:455-479.

Kelley J & B Alpagut. 1999. Canine sexing and species number in the Paşalar large hominoid sample. J Hum Evol 36:335-341.

Kingston JD & A Hill. 1999. Late Miocene palaeoenviornments in Arabia: A synthesis. In: Whybrow PJ & A Hill (Eds) Fossil Vertebrates of Arabia. New Haven: Yale University Press, p389-409.

Kidwell SM & AK Behrensmeyer. 1993. Taphonomic Approaches to Time Resolution in Fossil Assemblages. 16th Annual Short Course of the Paleontological Society, Boston, Massachusetts.

Kordos L & DR Begun. 2002. Rudabánya: A late Miocene subtropical swamp desposit with evidence of the origin of the African apes and humans. Evol Anthropol 11:45-57.

Koufos GD & de Bonis L. 2005. The Late Miocene hominoids Ouranopithecus and Graecopithecus. Implications about their relationships and taxonomy. Annls Paléont 91:227-240.

Koufos GD & L de Bonis. 2004. The late Miocene hominoids Ouranopithecus and Graecopithecus. Implications about their relationships and taxonomy. 5th International Symposium on Eastern Mediterranean Geology, Thessaloniki 1:322- 325.

Koussoulakou A & E Stylianidis. 1999. The Use of GIS for the Visual Exploration of Archaeological Spatio-Temporal Data. Cart Geog Info Sci 26:152-160.

Kopp ML. 2007. The late alpine structure of the Greater Caucasus as an element of the Peri- Arabian collisional area. European Geosciences Union Geophysical Research Abstracts, Volume 9.

249

Krause DW & MC Maas. 1990. The biogeographic origins of late Paleocene-early Eocene mammalian immigrants to the Western Interior of North America. In: Bown TM & KD Rose (Eds) Dawn of the Age of Mammals in the northern part of the Rocky Mountain Interior. Geological Society of American Special Papers 243:71-105.

Krijgsman W. 2002. The Mediterranean: Mare Nostrum of Earth Sciences. Earth Planet Sci Lett 205:1-12.

Krijgsman W, CG Langereis, R Daams & AJ van der Meulen. 1994. Magnetostratigraphic dating of the middle Miocene climate change in the continental deposits of the Aragonian type area in the Calatayud-Teruel basin (Central Spain). Earth Planet Sci Lett 128:513-529.

Kunimatsu Y, M Nakatsukasa, Y Sawada, T Sakai, M Hyodo, H Hyodo, T Itaya, H Nakaya, H Saegusa, A Mazurier, M Saneyoshi, H Tsujikawa, A Yamamoto & E Mbua. 2007. A new Late Miocene great ape from Kenya and its implications for the origins of African great apes and humans. Proc natl Acad Sci 104:19220-19225.

Lake MW, PE Woodman & SJ Mithen. 1998. Tailoring GIS Software for Archaeological Applications. J Arch Sci 25:27-38.

Leakey MG & JM Harris. 2003. Lothagam: Its Significance and Contributions. In: Leakey MG & JM Harris (Eds) Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, p 625-670.

Leakey MG & A Walker. 1997. Afropithecus: Function and Phylogeny. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p225-240.

Leakey MG, CS Feibel, RL Bernor, JM Harris, TE Cerling, KM Stewart, GW Storrs, A Walker, L Werdelin & A Winkler. 1996. Lothagam: A record of faunal change in the late Miocene of East Africa. J Vert Paleontol 16:556-570.

Leakey RE, MG Leakey & AC Walker. 1988. Morphology of Afropithecus turkanensis from Kenya. Am J phys anthrop 76:289-307.

Leakey REF & A Walker. 1985. New higher primates from the early Miocene of Buluk, Kenya. Nature 318:173-175.

Lindsay EH. 1989. European Neogene Mammal Chronology. NATO ASI Series A 180.

Lyman RL. 1994. Vertebrate Taphonomy. Cambridge: Cambridge University Press.

Maas MC, MRL Anthony, PD Gingerich, GF Gunnell & DW Krause. 1995. Mammalian

250

generic diversity and turnover in the Late Paleocene and Early Eocene of the Bighorn and Crazy Mountains Basins, Wyoming and Montana (USA). Palaeogeogr Palaeoclimatol Palaeoecol 115:181-207.

Madden CT. 1980. Zygolophodon from Subsaharan Africa, with Observations on the Systematics of Palaeomastodontid Proboscideans. Journal of Paleontology 54:57-64.

Made J van der. 2003. Suoidea (pigs) from the Miocene hominoid locality of Çandır in Turkey. In: Güleç E, DR Begun & D Geraads (Eds) Geology and Vertebrate Paleontology of the Middle Miocene Hominoid Locality Çandır. Cour Forsch –Inst Senckenberg 240:149-180.

Made J van der. 1999. Intercontinental relationship Europe - African and the Indian Subcontinent. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p457-472.

Made J van der, J Agustí & J Morales. 2003. The Vallesian Crisis in the Iberian and European mammal records. EEDEN – Environments and Ecosystem Dynamics of the Eurasian Neogene, Birth of the New World. Stará Lesná, Slovakia. November 12-16th, 2003. p61-62.

Magyar I, DH Geary & P Müller. 1999. Paleogeographic evolution of the Late Miocene Lake Pannon in Central Europe. Palaeogeogr Palaeoclimatol Palaeoecol 147:151- 167.

Manly BJF. 2005. Multivariate Statistical Methods: A Primer. Florida: Chapman & Hall/CRC.

Maridet O, G Escarguel, L Costeur, P Mein, M Hugueney & S Legendre. 2007. Small mammal (rodent and lagomorphs) European biogeography from the Late Oligocene to the Pliocene. Global Ecol. Biogeogr 16:529-544.

Martin L & P Andrews. 1993. Renaissance of Europe’s ape. Nature 365:494.

Martin L & P Andrews. 1984. The phyletic position of Graecopithecus freybergi von Koenigswald. Cour Forsch –Inst Senckenberg 69:25-40.

Martin L & P Andrews. 1983. Species recognition in middle Miocene hominoids. In: Kimbel WH & L Martin (Eds) Species, Species Concepts and Primate Evolution. p393-427.

Martín JM, JC Braga, I Sánchez-Almazo. 1999. The Messinian record of the outcropping marginal Alboran Basin deposits: significance and implications. In: Zahn R, MC Comas & A Klaus (Eds) Proceedings of the Ocean Drilling Program, Scientific Results, Volume 161 (Sites 974-979 Mediterranean II). Texas: Ocean Drilling Program. p543-551.

251

McCrossin ML & BR Benefit. 1997. On the relationships and adaptations of Kenyapithecus, a large-bodied hominoid from the middle Miocene of eastern Africa. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptations. New York: Plenum Press, p241-267.

Mein P. 2003. On Neogene rodents in Eurasia: distribution and migrations. In: Reumer JWF & W Wessels (Eds) Distribution and Migration of Tertiary Mammals in Eurasia. A Volume in Honour of Hans DeBruijn. Deinsea 10:407-418.

Mein P. 1989. Updating of MN Zones. In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:73-90.

Mein P. 1979. European Mammal Correlations. In: Lindsay EH, V Fahlbusch & P Mein (Eds) European Mammal Geochronology. New York: Plenum Press, p73-90.

Mein P. 1975. Résultats du Groupe de Travail des Vertébrés. In: Senes J (Ed) Report on Activity of RCMNS Working Groups 6. Congress of the Regional Committee of Mediterranean Neogene Stratigraphy, Proceedings 1:78-81.

Merceron G, E Schulz, L Kordos & TM Kaiser. 2007. Palaeoenvironment of Dryopithecus brancoi at Rudabánya, Hungary: evidence from dental meso and microwear analyses of large herbivorous mammals. J Hum Evol 53:331-349.

Merceron G, L de Bonis, L Viriot & C Blondel. 2005a. Dental microwear of fossil bovids from northern Greece: paleoenvironmental conditions in the eastern Mediterranean during the Messinian. Palaeogeogr Palaeoclimatol Palaeoecol 217:173-185.

Merceron G, L de Bonis, L Viriot & C. Blondel. 2005b. Dental microwear of the late Miocene bovids of northern Greece: Vallesian/Turolian environmental changes and disappearance of Ouranopithecus macedoniensis? Bull Soc geol. Fr 176:475-484.

Middlemiss FA & PF Rawson. 1969. Faunal provinces in Space and Time – some general considerations. In: Middlemiss FA, PF Rawson & G Newall (Eds) Faunal Provinces in Space and Time. Liverpool: Seel House Press, p199-210.

Moyà-Solà S, M Köhler, DM Alba, I Casanovas-Vilar & J Galindo. 2004. Pierolapithecus catalaunicus, a New Middle Miocene Great Ape from Spain. Nature 306:1339-1344.

Moyà-Solà S, J Quintana, J Antonio Alcover & M Köhler. 1999. Endemic Island Faunas of the Mediterranean Miocene. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p435-442.

Moyà-Solà S & M Köhler. 1993. Recent discoveries of Dryopithecus shed new light on evolution of great apes. Nature 365:543-545.

252

Nagel D. 2003. Carnivora from the middle Miocene Hominoid locality of Çandır. In: Güleç E, DR Begun & D Geraads (Eds) Geology and Vertebrate Paleontology of the Middle Miocene Hominoid Locality Çandır. Cour Forsch –Inst Senckenberg 240:113-132.

Nargolwalla MC. 2007. Middle and late Miocene terrestrial mammals from the Pannonian Basin, Central Europe. J Vert Paleontol 26(3):104a.

Nargolwalla MC, MP Hutchison & DR Begun. 2006. Middle and late Miocene terrestrial vertebrate localities and paleoenvironments in the Pannonian Basin. Beiträge zur Paläontologie. 30:347-360.

Nargolwalla MC & DR Begun. 2005. Late Miocene hominid biogeography and extinction patterns. Am J Phys Anthrop 126(S40):155.

Ni X & Z Qiu. 2002. The micromammalian fauna from the Leilao, Yuanmou hominoid locality: Implications for biochronology and paleoecology. J Hum Evol 42:535-546.

Ohta S, K Kaiho & T Takei. 2003. Relationship between surface-water temperature and ice- sheet expansion during the middle Miocene. Palaeogeogr Palaeoclimatol Palaeoecol 201:307-320.

Olson EC. 1980. Taphonomy: Its History and Role in Community Evolution. In Behrensmeyer AK & A Hill (Eds) Fossils in the Making: Vertebrate Taphonomy and Paleoecology. Chicago: University of Chicago Press, p5-19.

Pagani M, KH Freeman, MA Arthur. 1999. Concentrations and the Expansion of C4 Grasses. Science 285:876-880.

Peresson H & K Decker. 1997. The Tertiary dynamics of the northern Eastern Alps (Austria): changing palaeostresses in a collisional plate boundary. Tectonophysics 272(2-4):125-157.

Pickford M. 1989. Dynamics of Old World Biogeographic Realms During the Neogene: Implications for Biostratigraphy. In: In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:413-442.

Pielou EC. 1979. Biogeography. New York: Wiley

Pilbeam D. 1997. Research on Miocene Hominoids and Hominid Origins: The last three decades. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p13-28.

Pilbeam D. 1996. Genetic and morphological records of the Hominoidea and hominid origins: A synthesis. Mol Phylogenet Evol 5(1):155-168.

253

Pilbeam D. 1989. Human fossil history and evolutionary paradigms. In: Hecht MK (Ed) Evolutionary Biology at the Crossroads. New York: Queens College Press, p117- 138.

Pilbeam D & N Young. 2004. Hominoid evolution: synthesizing disparate data. CR Palevol 3:303-319.

Pilbeam D, M Morgan, JC Barry & L Flynn. 1996. European MN Units and the Siwalik Faunal Sequence of Pakistan. In: Bernor RL, V Fahlbusch, H-W Mittmann (Eds) TheEvolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p96-105.

Popov SV, F Rögl, AY Rozanov, FF Steininger, IG Shcherba M & Kovac. 2004. Lithological-Paleogeographic maps of Paratethys. Cour Forsch –Inst Senckenberg 250: 1-46.

Potts R. 2004a. Paleoenvironmental Basis of Cognitive Evolution in Great Apes. Am J Primatol 62:209-228.

Potts R. 2004b. Paleoenvironments and the Evolution of Adaptability in Great Apes. In: Russon AE & DR Begun (Eds) The Origins of Thought: The Evolutionary Origins of Great Ape Intelligence. Cambridge: Cambridge University Press, p78-127.

Potts R, T Jorstad & D Cole. 1996. The Role of GIS in Interdisciplinary Investigations at Olorgesailie, Kenya, at a Pleistocene Archaeological Locality. In: Aldenderfer M & HDG Maschner (Eds) Anthropology, Space, and Geographic Information Systems. New York: Oxford University Press, p202-213.

Potts R & AK Behrensmeyer. 1992. Late Cenozoic Terrestrial Ecosystems. In: Behrensmeyer AK, JD Damuth, WA DiMichele, R Potts, H-D Sues & SL Wing (Eds) Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. Chicago: University of Chicago Press, p419-541.

Pruetz JD. 2007. Evidence of cave use by savanna chimpanzees (Pan troglodytes verus) at Fongoli, Senegal: Implications for thermoregulatory behaviour. Primates 48:316- 319.

Quinn GP & MJ Keough. 2002. Experimental Design and Data Analysis for Biologists. Cambridge: Cambridge University Press.

Rantitsch G. 1997. Thermal history of the Carnic Alps (Southern Alps, Austria) and its palaeogeographic implications. Tectonophysics 272:213-232.

Ribot F, J Gibert & T Harrison. 1996. A reinterpretation of the taxonomy of Dryopithecus from Vallès-Penedès. J Hum Evol 31:129-141.

254

Rögl F. 1999a. Mediterranean and Paratethys palaeogeography during the Oligocene and Miocene. In: Agustí J, L Rook & P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p8-23.

Rögl F. 1999b. Circum-Mediterranean Miocene Paleogeography. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. Germany: Verlag Dr. Friedrich Pfeil, p39-48.

Rögl F. 1999c. Oligocene and Miocene Palaeogeography and Stratigraphy of the Circum- Mediterranean Region. In: Whybrow PJ & A Hill (Eds) Fossil Vertebrates of Arabia. New Haven: Yale University Press, p485-500.

Rögl F & FF Steininger. 1984. Neogene Paratethys, Mediterranean and Indo-pacific Seaways. In: Brenchley P (Ed) Fossils and Climate. New York: John Wiley and Sons, p171-200.

Rook L, G Gallai & D Torre. 2006. Lands and endemic mammals in the Late Miocene of Italy: Constraints from paleogeographic outlines of Tyrrhenian area. Palaeogeogr Palaeoclimatol Palaeoecol 238:263-269.

Rook L, P Renne, M Benvenuti & M Papini. 2000. Geochronology of Oreopithecus-bearing successions at Baccinello (Italy) and the extinction pattern of European Miocene hominoids. J Hum Evol 39:577-582.

Rook L, Abbazzi L & Engesser B. 1999. An overview of the Italian Miocene land mammal faunas. In: Agustí J, L Rook, P Andrews (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p191-204.

Ross CA. 1974. Paleogeography and Provinciality. In: Ross CA (Ed) Paleogeographic Provinces and Provinciality. Society of Economic Paleontologists and Mineralogists, Special publication No. 21:1-17.

Royden LH. 1988. Late Cenozoic Tectonics of the Pannonian Basin System. In: Royden LH & F Horváth (Eds) The Pannonian Basin – A Study in Basin Evolution, AAPG Memoir 45, 27-48.

Ruddiman WF & WL Prell. 1997. Introduction to the Uplift-Climate Connection. In: Ruddiman WF (Ed) Tectonic Uplift and Climate Change. New York: Plenum, pp3- 15.

Rummel M. 1999. Tribe Cricetodontini. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p359-364.

255

Russak SM & WC McGrew. 2008. Chimpanzees as fauna: Comparisons of sympatric large mammals across long-tem study sites. Am J Primatol 70:402-409.

Sánchez-Gómez M, JC Balanyá, V García-Dueñas & JM Azañón. 2002. Intracrustal tectonic evolution of large lithosphere mantle slabs in the western end of the Mediterranean orogen (Gibraltar arc). Journal of the Virtual Explorer 8.

Sanders WJ & ER Miller. 2002. New proboscideans from the early Miocene of Wadi Moghara, Egypt. J Vert Paleontol 22:388-404.

Sarmiento E. 1987. The phylogenetic position of Oreopithecus and its significance in the origin of the Hominoidea. Am Mus Novitates 2881:1-44.

Schellart W. 2002. Alpine deformation at the western termination of the Axial Zone, Southern Pyrenees. Journal of the Virtual Explorer 8.

Schlunegger F & G Simpson. 2002. Possible erosional control on lateral growth of the European Central Alps. Geology 30(10):907-910.

Schmidt-Kittler N. 1989. A Biochronologic Subdivision of the European Paleogene Based on Mammals. In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:-50.

Schwartz T. 1997. Lateritic bauxite in central Germany and implications for Miocene palaeoclimate. Palaeogeogr Palaeoclimatol Palaeoecol 129:37-50.

Selänne L. 2003. Genus Schizogalerix (Insectivora). In: Fortelius M, J Kappelman, S Sen & RL Bernor (Eds) Geology and Paleontology of the Miocene Sinap Formation, Turkey. New York: Columbia University Press, p69-89.

Sen S. 2003. Muridae and Gerbillidae (Rodentia). In: Fortelius M, J Kappelman, S Sen & RL Bernor (Eds) Geology and Paleontology of the Miocene Sinap Formation, Turkey. New York: Columbia University Press, p125-140.

Sen S. 1997. Magnetostratigraphic calibration of the European Neogene mammal chronology. Palaeogeogr Palaeoclimatol Palaeoecol 133:181-204.

Sen S. 1996. Present State of Magnetostratigraphic Studies in the Continental Neogene of Europe and Anatolia. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p56-63.

Sibuet J-C, SP Srivastava, W Spakman. 2004. Pyrenean orogeny and plate kinematics. Journal of Geophysical Research 109:1-18.

Simpson GG. 1960. Notes on the measurement of faunal resemblance. Am J Sci 258a:300-

256

311.

Simpson GG. 1940. Mammals and Land Bridges. J Wash Acad Sci 30:137-163.

Sokal RR & FJ Rohlf. 1995. Biometry. New York: WH Freeman and Company.

Sokal RR & PHA Sneath. 1963. The Principles of Numerical Taxonomy. San Francisco: Freeman & Co.

Solounias N, JM Plavcan, J Quade & L Witmer. 1999. The paleoecology of the Pikermian Biome and the savanna myth. In: Agustí J, L Rook & P Andrews, (Eds) Hominoid Evolution and Climatic Change in Europe, vol 1: The Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, p436- 453.

Spassov N & D Geraads. 2008. The latest prehumans of Europe: discovery of a late Miocene hominoid in Bulgaria of about 7Ma. Bulgarian Academy of Sciences News 2(54):2-4.

Spezzaferri S, Cita MB & McKenzie JA. 1998. The Miocene/Pliocene boundary in the Eastern Mediterranean: results from sites 967 and 969. In: Robertson AHF, K-C Emeis, C Richter & A Camerlenghi (Eds) Proceedings of the Ocean Drilling Program, Scientific Results, vol 160 (Sites 963-973 Mediterranean I). Texas (Ocean Drilling Program). p9-28.

Sponheimer M, JE Loudon, D Codron, ME Howells, JD Pruetz, J Codron, DJ de Ruiter & JA Lee-Thorp. 2006. Do “savanna” chimpanzees consume C4 resources? J Hum Evol 51:128-133.

Stampfli GM, GD Borel, R Marchant & J Mosar. 2002. Western Alps geological constraints on western Tethyan reconstructions. Journal of the Virtual Explorer 7:75-104.

Stanford C, J Wallis, H Matama & J Goodall. 1994. Patterns of predation by chimpanzees on red colobus monkeys in Gombe National Park, Tanzania, 1982-1991. Am J Phys Anthrop 94:213-228.

Steininger FF, RL Bernor & V Fahlbusch. 1989. European Neogene Marine/Continental Chronologic Correlations. In: Lindsay E, V Fahlbusch & P Mein (Eds) European Neogene Mammal Chronology. NATO ASI Series A 180:15-46.

Steininger FF, G Rabeder & F Rögl. 1985. Land Mammal Distribution in the Mediterranean Neogene: A Consequence of Geokinematic and Climatic Events. In: Stanley DJ & F- C Wezel (Eds) Geological Evolution of the Mediterranean Basin: New York: Springer-Verlag, p560-571.

Stewart C-B & TR Disotell. 1998. Primate evolution – in and out of Africa. Current

257

Biology 8:R582-R588.

Susman RL. 1985. Functional morphology of the Oreopithecus hand. Am J Phys Anthrop 66:235.

Swisher III CC. 1996. New 40Ar/39Ar Dates and Their Contribution Toward a Revised Chronology for the Late Miocene of Europe and West Asia. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p64-77.

Suwa G, RT Kono, S Katoh, B Asfaw & Y Beyene. 2007. A new species of great ape from the late Miocene epoch in Ethiopia. Nature 448:921-924.

Szalay FS & E Delson. 1979. Evolutionary History of the Primates. New York: Academic Press.

Tarling DH. 1978. The geological-geophysical framework of ices ages. In: Gribbin J (Ed) Climatic Change. Cambridge: Cambridge University Press, p3-24.

Thenius E. 1952. Die Säugetierfauna aus dem Torton von Neudorf an der March (CSR). Neues Jahrbuch für Geologie und Paläontologie 96:27-136.

Thomas H. 1985. The Early and Middle Miocene Land Connections of the Afro-Arabian Plate and Asia: A Major Event for Hominoid Dispersal? In: Delson E (Ed) Ancestors, the Hard Evidence. New York: Liss, p42-50.

Thomas H, RL Bernor & J-J Jaeger. 1982. Origines du peuplement mammalien en Afrique due nord durant le Miocène terminal. Geobios 15 :283-297.

Tong H & J-J Jaeger. 1993. Muroid rodents from middle Miocene Fort Ternan locality (Kenya) and their contribution to the phylogeny of muriods. Palaeontographica 229:51-73.

Tutin CEG, R Ham, LJT White & MJS Harrison. 1997. The primate community of Lope’ reserve in Gabon: Diets, responses to fruit scarcity, and effects on biomass. Am J Primatol 42:1-24.

Ünay E, H de Bruijn & G Saraç. 2003. A preliminary zonation of the continental Neogene of Anatolia based on rodents. In: Reumer JWF & W Wessels (Eds) Distribution and Migration of Tertiary Mammals in Eurasia. A Volume in Honour of Hans DeBruijn. Deinsea 10:539-548.

Ungar PS. 2005. Dental evidence for the diets of fossil primates from Rudabánya, northeastern Hungary with comments on extant primate analogs and “noncompetitive” sympatry. Palaeontographica Italica 90:97-111.

258

Ungar PS. 1998. Dental allometry, morphology and wear as evidence for diet in fossil primates. Evol Anthropol 6:205-217.

Ungar PS. 1996. Dental microwear of European Miocene catarrhines: Evidence for diets and tooth use. J Hum Evol 31:335-366.

Ungar PS & RF Kay. 1995. The dietary adaptations of European Miocene catarrhines. Proc Nat Acad Sci USA 92:5479-5481. van Dam JA, L Alcalá, A Alonso Zarza, JP Calvo, M Garcés & W Krijgsman. 2001. The upper Miocene mammal record from the Teruel-Alfambra region (Spain). The MN system and continental stage/age concepts discussed. J Vert Paleontol 21(2):367-385. van Dam JA & GJ Weltje. 1999. Reconstruction of the Late Miocene climate of Spain using rodent palaeocommunity successions: an application of end-member modelling. Palaeogeogr Palaeoclimatol Palaeoecol 151:267-305.

Vergé J, M Fernàndez & A Martínez. 2002. The Pyrenean orogen: pre-, syn-, and post- collisional evolution. Journal of the Virtual Explorer 8.

Vestal AG. 1914. Internal relations of terrestrial associations. Am Naturalist 48:413-445.

Waddle DM, LB Martin, DA Stock. 1995. Sexing isolated hominoid canines with special reference to the Middle Miocene specimens from Paşalar, Turkey. J Hum Evol 28:385-403.

Ward CV, DR Begun & MD Rose. 1997. Function and Phylogeny in Miocene Hominoids. In: Begun DR, CV Ward & MD Rose (Eds) Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptation. New York: Plenum Press, p1-12.

Ward SC, B Brown, A Hill, J Kelley & W Downs. 1999. Equatorius: a new hominoid genus from the middle Miocene of Kenya. Science 285:1382-1386.

Weerd A van de & R Daams. 1978. Quantitative composition of rodent faunas in the Spanish Neogene and paleoecological implications. Proceedings Koninklijke Nederlandse Akademie van Wetenschappen 81(4):448-473.

Werdelin L. 2003. Mio-Pliocene Carnivora from Lothagam, Kenya. In: Leakey MG & JM Harris (Eds), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, p261-328.

Werdelin L. 1996. Carnivores, exclusive of the Hyaenidae, from the later Miocene of Europe and Western Asia. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p271-289.

259

Werdelin L & N Solounias. 1996. The Evolutionary History of in Europe and Western Asia During the Miocene. In: Bernor RL, V Falhbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p290-306.

Werdelin L & N Solounias. 1991. The Hyaenidae: taxonomy, systematics and evolution. Fossils and Strata 30:1-104.

Wesselman HB. 1995. Of mice and almost men. In: Vrba ES, GH Denton, TC Partridge & LH Burckle (Eds) Paleoclimate and evolution with emphasis on human origins. New Haven: Yale University Press, p356–368.

Wessels W, O Fejfar, P Peláez-Campomanes, A van der Meulen & H de Bruijn. 2003. Miocene small mammals from Jebel Zelten, Libya. Coloquois de Paleontologia 1:699-715.

Wessels W. 1999. Family Gerbillidae. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Pfeil, p395-400.

Whybrow PJ & D Clements. 1999. Arabian Tertiary fauna, flora, and localities. In: Whybrow PJ & A Hill (Eds), Fossil Vertebrates of Arabia. New Haven: Yale University Press, p460-475.

Woodburne MO, RL Bernor, CC Swisher III. 1996. An appraisal of the Stratigraphic and Phylogenetic Bases for the “Hipparion” Datum in the Old World. In: Bernor RL, V Fahlbusch & H-W Mittmann (Eds) The Evolution of Western Eurasian Neogene Mammal Faunas. New York: Columbia University Press, p124-136.

Yamakoshi G. 1998. Dietary responses to fruit scarcity of wild chimpanzees at Boussou, Guinea: Possible implications for ecological importance of tooth use. Am J Phys Anthrop 106:283-295.

Zeck HP. 1996. Betic-Rif orogeny: subduction of Mesozoic Tethys lithosphere under eastward drifting Iberia, slab detachment shortly before 22 Ma, and subsequent uplift and extensional tectonics. Tectonophysics 254:1-16.

Zhang Z & T Harrison. 2008. A new middle Miocene pliopithecid from Inner Mongolia, China. J Hum Evol. 54:444-447.

Ziegler R. 1999. Order Insectivora. In: Rössner GE & K Heissig (Eds) The Miocene Land Mammals of Europe. München: Verlag Dr. Friedrich Pfeil, p53-74.