<<

Pan-African Climate Variability since the Plio- and Possible Implications for Hominin Evolution

Item Type text; Electronic Dissertation

Authors Billingsley, Anne Louise

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction, presentation (such as public display or performance) of protected items is prohibited except with permission of the author.

Download date 01/10/2021 17:15:21

Link to Item http://hdl.handle.net/10150/634339 1

PAN-AFRICAN CLIMATE VARIABILITY SINCE THE PLIO-PLEISTOCENE AND

POSSIBLE IMPLICATIONS FOR HOMININ EVOLUTION

By

Anne Louise Billingsley

______

Copyright © Anne Louise Billingsley 2019

A Dissertation Submitted to the Faculty of the

DEPARTMENT OF GEOSCIENCES

In Partial Fulfillment of the Requirements

For the Degree of

DOCTOR OF PHILOSOPHY

In the Graduate College

THE UNIVERSITY OF ARIZONA

2019

2

3

ACKNOWLEDGEMENTS

I would like to express my gratitude for my family and friends who inspire and encourage me no matter what crazy decisions I decide to make. In particular, my mother, Mary, and my sister, Mary Lynn. I would like to express how thankful I am for my husband, Andrew

Cunningham, who I met while working on this dissertation, and who actually enjoyed my company enough to want to spend the rest of his life with me in spite of the fact that I was working on this dissertation.

To all my friends who have been on this journey with me from the beginning when we were just baby geologists, meeting you all and growing as scholars together has been the best part about this experience.

My collaborators and labmates: Chad Yost, Jordan Bright, Benjamin Mohler, Kevin

Ortiz, Mathew Fox, Jennifer Kielhofer, Kayla Worthey, Brendan Fenerty, Zachary Naiman,

David Dettman, Jessie Pearl, Nate Rabideaux, Jeffery Stone, Karlyn Westover, Jose Joordens,

Christopher Campisano and the entirety of the HSPDP team. It was a real joy to work with such great people.

My committee members have been ever so helpful with advice and guidance: Andrew S.

Cohen, Jay Quade, Peter N. Reinthal, Matt Grove and Kevin Anchukaitis.

The National Science Foundation (EAR1123000, EAR1338553, BCS124185) and the

International Continental Drilling Program provided support for this research.

4

DEDICATION

This manuscript is dedicated to my father, Thad Hoffman Billingsley, who inspired my love of the sciences and instilled in me the value of an education.

5

TABLE OF CONTENTS

ABSTRACT………………………………………………………………………………………7

INTRODUCTION ………………………………………………………………………………..8

PRESENT STUDY ……………………………………………………………………………...15

REFERENCES ………………………………………………………………………………….22

APPENDIX A: δ13C RECORDS FROM FISH FOSSILS AS PALEO-INDICATORS OF ECOSYSTEM RESPONSE TO LEVELS IN THE PLIO-PLEISTOCENE OF TUGEN HILLS, KENYA ……………………………………………………………………….27

Key Points ….……………………………………...…………………………………….28 Abstract ……………………………………………………………………………….…28 1. Introduction……………………………...………………..………………………….29 2. Methods …………………………………..…………………………………….……34 3. Results ……………………………..………………………….……………………..36 4. Discussion ……………………………..…………………….……………………....38 5. Conclusions ………………………..……………...…………………………………40 Acknowledgements ……………….……………………………………………………...41 References ……………………………………………………………………………….41 Figures …………………………………………………………………………………..47 Tables……………………………………………………………………………….……54

APPENDIX B: ORBITAL AND HIGH-LATITUDE CONTROLS OF PLIO-PLEISTOCENE CLIMATE VARIABLITY IN , IMPLICATIONS FOR EARLY HOMININS AND THE RISE OF THE GENUS HOMO …………………………………………………………...56

Key Points ……………………………………………………………………………….57 Abstract ………………………………………………………………………………….57 1. Introduction …..…………………...…………………………………………………58 2. Methods ..…………………………………………………………………………….61 3. Results ……..………………………………………………………………………...64 4. Discussion ..………………………………………………………………………….66 5. Conclusions ………..………………………………………………………………...75 Acknowledgements ………………………………………………………………………77 References ……………………………………………………………………………….77 Figures …………………………………………………………………………………..95 6

TABLE OF CONTENTS – Continued

APPENDIX C: NEW APPROACHES TO QUANTIFYING PALEOENVIRONMENTAL VARIABILITY AND APPLICATION TO UNDERSTANDING THE EVOLUTION AND DISPERSAL OF ANATOMICALLY MODERN HUMANS OVER THE PAST 400 KA …....99

Key Points ……………………………………………………………………………...100 Abstract ………………………………………………………………………………...100 1. Introduction ………..……………………………………………………………….101 2. Methods ……..……………………………………………………………………...109 3. Results …………………………..………………………………………………….113 4. Discussion ………………………………………..………………………………...115 5. Conclusions …………………………………..…………………………………….129 Acknowledgements ……………………………………………………………………..130 References ……………………………………………………………………………...131 Figures …………………………………………………………………………………145 Tables …………………………………………………………………………………..153

7

ABSTRACT

Three studies were performed in order to ascertain how climate variability in Africa may have influenced the , extinction and dispersals of our early hominin ancestors since the

Plio-Pleistocene. δ13C values from fish fossils from a Hominin Sites and Paleolakes Drilling

Project Core (HSPDP) from Tugen Hills, Kenya were used to determine that local responses to climate shifts between 3.29 – 2.57 Ma were abrupt enough to limit the habitats of fish in the paleolake. Singular spectral analysis and correlation analysis on five paleoenvironmental records from Africa and one from NE Russia indicate that orbital parameters can explain most of the variance in climate records and that there are far reaching high-latitude teleconnections to

African climate. An evolutionary algorithm run on reconstructed time-series from western, eastern and allowed for environmental context of hominin evolutionary events to be made. Comparisons between the evolutionary model and the paleoanthropological record showed that during a period of relative stability in eastern Africa, 2.9 – 2.7 Ma, a drastic change in environment at 2.8 Ma would have elicited a response from a population of increased plasticity. This event coincides with a major turnover in the faunal assemblages as well as the first appearance of the genus Homo. New methods were developed in order to aggregate and compare paleorecords of different type and thereby quantify periods of high and low climate variability. Both the Standard Deviation Variability Method and the Accumulated Weighted Rate of Change Variability Method proved valuable in determining environmental variability in and around Africa over the last 400-ka. When the new methods were applied to environmental records and compared to the genetic, archaeological and fossil record, they indicated that the earliest appearance of the MSA occurred during a period of high variability in northern and eastern Africa. Furthermore, these methods showed that migrations within and out of Africa 8 occurred after the source switched from a period of high variability to a stable environment.

INTRODUCTION

The questions of if and how environment and climate shaped the process of hominin evolution have been hotly debated over the course of the last century. Many hypotheses have been put forth to explain the environmental forcing factors which could have given rise to the features that define the hominin lineage such as increased brain size, bipedalism and the use of tool technologies. Two of the leading hypotheses focus on the increased climate variability early hominins would have experienced. The variability selection hypothesis claims that a continuously changing environment would have selected for early hominins to evolve adaptive strategies, and the accumulated plasticity hypothesis adds to this by suggesting that species experiencing increased climate variability would have evolved phenotypic plasticity, a suite of characteristics and technology which allows for adaptation to a consistently changing environment. During periods of climate stability, these species might be outcompeted by species more adapted to one particular environmental regime (Potts, 1996, 1998; Grove, 2011, 2012,

2014, 2015). In this case, the species with accumulated plasticity would be able to disperse into other where they would be able to survive in a variety of novel environments (Grove,

2015).

Relating climate variability to human evolution has been difficult because of the lack of environmental records that are at the temporal and spatial resolution needed to make direct comparisons to known evolutionary milestones such as speciation, extinctions and dispersal events. The problem becomes further complicated as there are no direct measures of what constitutes a more variable environment, how different environments will react to various global 9 forcing mechanisms or exactly what sort of amplitude or temporal change in climate is required to elicit an evolutionary response in species. The purpose of the three studies appended to this manuscript is to address some of these concerns.

The Plio-Pleistocene is a period of major global climate reorganization as the earth moved from the warm Pliocene into the much colder, more variable Pleistocene at the onset of

Northern Hemisphere glaciation. Records from deep sea drill cores as well as multiple terrestrial records indicate that during this period, the African became much more arid and climate became increasingly unstable, with fluctuations between wet/dry phases (deMenocal,

1995; Deino et al., 2006; Kingston, 2007; Kingston et al., 2007; Maslin and Trauth, 2009; Trauth et al., 2009). During this period, the Valley had large lakes which varied in water levels through time (Trauth et al., 2009; Shultz and Maslin, 2013). However, the timing, extent and synchronicity of most of these lake-level pulses is still weakly constrained.

In 2013, a 227 m drill core was retrieved from the Chemeron Formation in the Tugen

Hills, Baringo Basin, Kenya as part of the Hominin Sites and Paleolakes Drilling Project

(HSPDP-BTB13-1A). A robust age model dates this core from 2.57 – 3.29 Ma (Deino et al.,

2019). The site was selected because of its close proximity to many known fossil vertebrate

(including hominin) sites, and because previous studies of outcrops in the area had shown that the lake fluctuated on precessional and millennial timescales (Deino et al., 2006; Kingston, 2007;

Wilson et al., 2014; Cohen et al., 2016; Campisano et al., 2017). One of the defining characteristics of the BTB13-1A core is a series of five deep lake phases denoted in the sedimentary record by diatomites (Deino et al., 2006; Kingston et al., 2007; Westover et al., in review). Outcrop observations previously showed that these deposits lacked the abundant fish 10 fossils seen elsewhere in the paleolake record, raising questions about the timing and extent of the lake level high-stands and the nature of these lakes (Kingston et al., 2007).

The first study appended to this manuscript seeks to determine how fish communities in the Tugen Hills paleolakes responded to lake-level pulses by using the δ13C values obtained from the bulk organic matter preserved in fish fossils. δ13C values from fish reflect trophic resources and can be used to infer fish habitats (Bootsma et al., 1996; Reinthal et al., 2011). Benthic, shallow water species typically have much higher δ13C values than species with a pelagic food source allowing for inferences to be made about relative lake depth (Bootsma et al., 1996;

Reinthal et al., 2011). We also compare our fish fossil record to the δ13C of modern (ca 1978) fish bones obtained from as well as the values seen in a Lake drill core.

The δ13C values of the fish from BTB13-1A (‒20 ‰ and ‒27 ‰ VPDB) are much more negative than those seen in Lake Turkana (‒ 16 ‰ and ‒ 24.6 ‰ VPDB) but overlap with the Lake

Malawi values (‒7.2 ‰ to ‒27.5 ‰ VPDB). The lack of benthic, shallow water species in the

BTB13-1A core suggests that the lake was deep during maximum transgressions and that the responses in the Tugen Hills paleolakes to climate change were abrupt enough to prevent shallow habitats from becoming established. The abruptness of environmental response may have greatly influenced hominins living in the region as resources on the landscape would have been affected by the presence or absence of a large lake.

The first objective of the second study appended to this manuscript is to determine how climate variability in Africa responded to variations in orbital parameters around the time of the onset of glaciation. Many previous studies have shown that there are both orbital and high-latitude teleconnections in African climate; however, the amount of control these forcing have on the regional climate is still debated (deMenocal, 1995; Trauth et al., 2007; 11

Tierney et al., 2008; Wilson et al., 2014). Singular spectral analysis was performed on five climate records from Africa as well as one record from NE Russia in order to determine the defining empirical orthogonal functions (EOFs) above noise. This analysis indicated that a high percentage (61 % ‒ 86 %) of climate variability in all six records can be attributed to orbital parameters on Milankovitch timescale cycles of eccentricity, obliquity and precession. A reconstructed time-series of each record was created using the EOFs which were significant above noise. Correlation analysis of these reconstructed series indicated that there was significant positive correlation between the temperature record from NE Russia and the Tugen

Hills record from Kenya, with eastern Africa becoming more arid during warm periods in the high latitude Northern Hemisphere.

The second objective of the second study appended to this manuscript was to determine how climate variability during the Plio-Pleistocene may have influenced hominin evolution. An evolutionary algorithm was run on the reconstructed time-series from eastern, western and southern Africa. The algorithm output is of the mean assigned environment, standard deviation around a mean environment and fitness to a particular environment of a constant population of

1000 individuals (Grove, 2011, 2015). The standard deviation of the model population is related to the plasticity accumulated in response to environmental change, whereas fitness is related to the mean environment to which the species is adapted (Grove, 2011, 2015). In this sense, increased standard deviation in the population represents adaptations to more variable environments, whereas more stable environments lead to increased fitness and specialization

(Grove, 2011, 2015).

The evolutionary algorithm showed that all three reconstructed time-series elicited both fitness and standard deviation responses in the population. When compared to the 12 paleoanthropological record, these predicted responses aligned well with known fossil and archaeological hominin milestones as well as species turnover events seen in both eastern and southern Africa.

Between 3.29 Ma and 3.05 Ma relatively stable environments in eastern and western

Africa correlate with increased model fitness in the hypothetical population. At 3.05 Ma, there is a shift in both records to more climate variability and a response in the modeled population of increased standard deviation. During this time interval, Australopithecus bahrelghazali (Chad) falls out of the fossil record, and Australopithecus afarensis () experienced an increase in body size (Bonnefille et al., 2004; Campisano and Feibel, 2007; Lee-Thorp et al., 2007). At

2.9 Ma, eastern African climate entered a more stable state, which initiated a model response of fitness. During this period, Australopithecus afarensis disappeared from the fossil record, perhaps suggesting that the new, stable environment was not suitable for this species. At 2.8 Ma, there was a brief but rapid climate transition in eastern Africa. This elicits a peak in standard deviation from the modeled population, which coincides with a massive faunal turnover in eastern Africa as well as the first appearance of our genus, Homo (Bobe and Behrensmeyer,

2004; Bobe and Eck, 2001; DiMaggio, 2018; deMenocal, 2004). At 2.7 Ma, the climate in eastern Africa again became much more variable. In the model this elicits a standard deviation response, which coincides with the timing of the first appearance of Paranthropus aethiopicus and Oldowan tool technologies (Grove, 2012; Braun et al., 2019).

In opposition to the dynamic changes seen throughout western and eastern Africa, southern Africa remains relatively stable between 3.27 – 2.7 Ma. Between 2.7 – 2.6 Ma, however, there was a drastic change in climate, which elicits a standard deviation response in the model population. This event coincides with a faunal turnover in South Africa (Vrba, 1993). The 13 relatively stable environment and the consistent response of the modeled population in southern

Africa may lend credence to the theory that southern Africa was a refugium for various species throughout the Plio-Pleistocene, and could help explain why the massive faunal turnover seen in eastern Africa at 2.8 Ma is not recorded in the fossil record of southern Africa (Reynolds, 2007;

Faith and Behrensmeyer, 2013). Comparisons between the predicted population response to environmental variability calculated by the evolutionary algorithm and known faunal and hominin events in Africa indicate that climate variability may be a driving factor in hominin and other mammalian evolution, supporting the concepts put forth by both the variability selection and accumulated plasticity hypotheses.

Although we have shown that climate variability in Africa may elicit modeled responses in adaptive traits that correlate to known hominin milestones, what constitutes a high-variability climate has not previously been clearly defined. The third study appended to this manuscript aims to quantify paleoenvironmental variability. To do this, two new metrics were developed, which can be applied to any type of environmental paleorecord. The first method is the Standard

Deviation Variability Method (SDV). In this method the variance of the detrended z-score data in any given time bin determines the range of variability within that bin. The second method is the

Accumulated Weighted Rate of Change Variability Method (AWRCV). In this method, it is the summed absolute values of the duration-normalized slopes (or accumulated change) between detrended z-scores of data points in any particular time bin that defines variability. Explicit instructions for the application of these methods are included in Appendix C. When applied to an evenly spaced data record like insolation, both methods produce similar records of variability.

By comparing values from individual bins and aggregating records from sites within regions, periods of higher climate variability for a given region can be determined. 14

These methods were then applied to 31 different paleoenvironmental records from 23 different locations in and around Africa covering the last 400 ka. The purpose of this exercise was to get the full range of spatial and temporal environmental variability experienced by early anatomically modern humans and determine whether variability was likely to have been a driving factor in speciation, technology development and dispersals within and out of Africa.

The individual records were then divided into five different regions based on prolonged climate similarities between the records (i.e. Blome et al., 2012). This analysis indicated that all five regions experienced periods of climate stability and climate instability, with periods of instability rarely overlapping between regions. Climate variability indicated by the new methods coincided with known environmental changes seen in other studies adding confidence to the methods.

Comparisons between the records of variability and the human record show interesting correlations. First, the first appearance of Middle Stone Age technology at both Jebel Irhoud,

Morocco and Olorgesailie, Kenya occur during periods of increased climate variability (Richter et al., 2017; Deino et al., 2018). Second, the appearance of multiple fossils containing anatomically modern features appear in the records shortly after periods of high climate variability in all five regions. Lastly, dispersals known and implied through genetic, archaeological and fossil records all occur as the source region of the dispersal shifts from a period of climate instability into a period of greater stability. Most importantly, the timing of plausible dispersal events out of Africa during MIS 5 occurred during or directly after a period of variability in northern Africa, while the 65-ka dispersal out of Africa occurred after a period of higher variability in northern Africa and an extended period of higher variability in eastern

Africa. These patterns support the notions put forward by the accumulated plasticity hypothesis. 15

The three studies appended to this manuscript are aimed at determining climate variability in Africa since the Plio-Pleistocene in the context of human evolution. The first study shows that local responses to climate change can be relatively abrupt affecting the ability for species in any location to adapt to the changes. The second study indicates that during the Plio-

Pleistocene, high latitude and orbital forcing greatly influenced the pacing of climate variability in Africa. This study also demonstrated how using an evolutionary algorithm may help put hominin evolutionary milestones into an environmental context. The third study provides methods for quantifying environmental variability in paleorecords. Although the SVD and

AWRCV metrics were developed here to address specific questions regarding anatomically modern humans and climate variability, they are equally applicable to addressing questions regarding environmental change and species’ responses for other taxa and/or time periods.

Furthermore, these studies support the basic notions put forth by the variability selection and accumulated plasticity hypotheses, and the author hopes that they contribute to the conversation regarding how climate may have influenced human evolution.

PRESENT STUDY

The papers appended to this document (A-C) are formatted for individual publication in scientific journals. Each paper gives a detailed description of the individual study’s motivation, location, methods, results and conclusions. Appendix A has been submitted for publication while appendices B and C are prepared for submission. I am the lead author and contributor on these multi-authored manuscripts. The following is a brief overview of each individual project.

Appendix A

In this study we use the δ13C values from fish fossils in order to better constrain the timing and extent of lake level fluctuations in a Hominins Sites and Paleolakes Drilling Project 16

(HSPDP) core collected in the Tugen Hills region of Kenya. The 227-meter core from part of the

Chemeron Formation dates from 3.29 – 2.56 Ma, encompassing the Plio-Pleistocene climate transition and the earliest appearance of Paranthropus sp. and Homo sp. (Deino et al., 2019).

Previous studies of the region have shown that lake levels fluctuated significantly on precessional and millennial time-scales signifying periods of high climate variability (Deino et al., 2006; Kingston et al., 2007; Wilson et al., 2014). One of these periods of high climate instability (~2.58 – 2.7 Ma) is marked by the appearance of a series of diatomites formed during deep-lake phases, which were separated by fluvial and exposure surface deposits during lake- level low stands (Deino et al., 2006; Kingston et al., 2007; Wilson et al., 2014). However, these deep lake phases lack the abundance of fish fossils seen in other depositional environments in outcrops, which raises the question of how fish communities may have responded to the timing and extent of deep lake-level pulses.

δ13C values of fish fossils reflect trophic resources and can be used to infer habitats of fish in the paleolakes (Bootsma et al., 1996; Reinthal et al., 2011). For this study, we compare the carbon isotope ratios of organic matter of fish fossils obtained from the Tugen Hills core to a previously published record of a drill core (139 ka – present) and core top samples

(modern ca 1978) from the water/sediment boundary of Lake Turkana, Kenya. This study shows that both Lake Malawi (‒7.2 ‰ to ‒27.5 ‰ VPDB) and Lake Turkana (‒16 ‰ to ‒24.6 ‰

VPDB) have δ13C values ranging from those found in near-shore, littoral species to those of deep-water pelagic species. However, the Tugen Hills fossil assemblage vary only between ‒20

‰ and ‒27 ‰ VPBD, lacking the more positive values seen in near-shore, benthic fish. The absence of shallow water, benthic lacustrine fish fossils through the Tugen Hills lake cycles suggests that the rate of change from low-lake stands to deeper lake phases was very rapid, 17 preventing near-shore fish communities from becoming established. These results strongly suggest that lake level responses to climate variability in the Baringo Basin of the East African

Rift were very abrupt during the Plio-Pleistocene transition.

Appendix B

The Plio-Pleistocene was an important time in the earth’s history both climatically and in terms of human evolution. During this period, global climate shifted from a warm Pliocene to the cooler, more variable Pleistocene. As the world got colder, Africa experienced an increase in aridity and climate variability. There are many questions about how high-latitude climate change would influence African climates and to what extent orbital parameters would control the pacing of climate variability. Furthermore, how these changes influenced human evolution has been a constant point of discussion in paleoanthropology for more than a century. The unidirectional hypothesis suggests that we adapted to more arid and savannah-dominated landscapes, whereas others suggest that the features that define hominins evolved in response to more mosaic environments and increasing variability in the climate (Dart, 1925; Vrba, 1993; Potts, 1998;

Kingston, 2007; Maslin and Trauth, 2009; Grove, 2011). Testing the links between hominin evolution and climate have proven difficult as it is unclear what responses climate may elicit in early hominins.

In this study, we looked at five environmental records from Africa and one record from

NE Russia spanning 3.29 – 2.57 Ma. We use singular spectral analysis and correlations to determine the pacing of climate variability and the influence of high latitude teleconnections. We used the significant explained variance from our analysis to create environmental time-series above noise. Correlations between the reconstructed time-series were used to determine relationships between regions in Africa and the Northern Hemisphere. An evolutionary algorithm 18 was run on the reconstructed time-series from eastern, southern and western Africa in order to determine possible species’ responses to the variability in each record (Grove, 2011, 2014,

2015). The predictions were then compared to the fossil and paleoanthropological records of

Africa to determine if evolutionary milestones corresponded to the predicted responses to variability.

Our study shows that a high percentage of climate variability in each record can be explained by orbital parameters of Milankovitch cycles of eccentricity, obliquity and precession.

We also show through correlation analysis that when the Northern Hemisphere was warm during the Plio-Pleistocene, our eastern Africa site experienced more arid conditions. We also conclude that the environmental changes across the continent are not synchronous, with periods of aridity and moisture varying between regions.

The variability seen within the climate records is reflected in the modeled hominin responses created by the evolutionary algorithm. Periods of extreme environmental shifts elicit a response of increased standard deviation, i.e. a wider range of environments a species is adapted to, in the modeled population while relatively stable periods select for increased fitness in populations. Various hominin species’ populations from eastern, western and southern Africa would have been subjected to different abiotic forcing events, and the differences in the populations responses between each region indicates this.

Eastern Africa experienced five periods where the environment was variable enough trigger model selection towards plasticity. The first of two of these periods, 3.05 – 2.9 Ma, correlates with a time period of increasing body size for Australopithecus afarensis in Ethiopia and the Last Appearance Datum (LAD) of Australopithecus bahrelghazali in Chad. The switch from high variability into a low stable environment coincides with the LAD of Australopithecus 19 afarensis. A drastic change in environment at 2.8 Ma occurred simultaneous with a massive faunal turnover in eastern Africa as the environment went from relatively arid to much wetter.

This peak in standard deviation also correlates with the first appearance of our genus Homo.

Between 2.7 -2.57, eastern Africa experienced increased variability which is associated with an increased standard deviation of the population. During this time, Paranthropus aethiopicus and

Oldowan technology appear in the paleoanthropological record.

Southern African climate did not vary as much as eastern or western Africa did. Although there were multiple fluctuations in the climate record, only one of these occurred quickly enough to elicit an increased standard deviation response in the model population. This one period, 2.7 –

2.6 Ma, coincides with a pulse turnover event in South African bovids (Vrba, 1993). This stable climate and consistent response of the model population lends credence to the idea that southern

Africa may have served as a refugium for many mammal species during the Plio-Pleistocene when conditions elsewhere in Africa were much more unfavorable.

Our analysis has indicated that climate variability and the abiotic forcing it creates hold some explanatory power for understanding the timing of speciation and extinction events for hominins and other mammals in Africa during the Plio-Pleistocene transition. The correlations between climate variability, increased standard deviation in environmental range of a modeled population, and changes in the hominin and fossil record support the fundamental ideas outlined by the variability selection and the accumulated plasticity hypotheses.

Appendix C

One of the enduring questions in paleoanthropology has been how climate and environment influenced the evolution and dispersal of our early human ancestors around and out of Africa. There have been many proposed hypotheses linking the environmental history and 20 human evolution. Some, such as the savanna hypothesis and the pulse-turnover hypothesis, focus on a unidirectional model of increased aridity and grasslands as a driving factor (Vrba, 1993; deMenocal, 1995) whereas others, such as the variability selection hypothesis, climate-pulse hypothesis and the accumulated plasticity hypothesis, focus on the heterogeneity of past environments and specifically the role of climate variability as selective agents for evolutionary novelty (Potts, 1996; Kingston, 2007; Grove, 2011, 2014; Maslin et al., 2014).

Testing these hypotheses has been difficult because many paleoenvironmental records lack the spatial and temporal range to allow for direct comparison between past environmental changes and the paleoanthropological record on time-scales relevant to human evolutionary events. The problem is further complicated by the fact that it has proven difficult to determine exactly what constitutes a more variable environment. In this study, we develop two new methods for quantifying environmental variability and comparing those quantifications among different types of paleoenvironmental records. The first method, the Standard Deviation

Variability Method (SDV), allows for the variance in detrended z-scores of any given time bin to be determined by the range of data for that time bin. The second, the Accumulated Weighted Rate of Change Variability Method (AWRCV), uses the summed and duration-weighted rates of change between sequential detrended z-score data points in a given time bin to determine the variability. Higher SDV and AWRCV values would indicate high climate variability, allowing for periods of climate/environmental instability to be distinguished from those showing little to no change.

We applied these new methods to 31 different paleoenvironmental records obtained from

23 locations in Africa, through the Levant, southwest and southern . These records cover the past 400 ka, a time period encompassing the evolution of anatomically modern humans 21

(AMH), the developments of Middle and Later Stone Age tool technology and the dispersal of

AMHs out of Africa. The records were then divided into five ( and southwest Asia, northern Africa and southern Levant, eastern Africa, and southern Africa) in order to better compare our climate variability record to known dispersals and evolutionary events of AMHs. Our study shows that all five of these regions experienced periods of high and low environmental variability, and that these periods are not synchronous though the study area. This suggests important regional responses to external and global climate forcing. Although most periods of regional climate variability can be related to global climate events, such as the interglacial/glacial cycles and orbital forcing, the heterogeneous responses of each region to these forcing mechanisms indicates that regional topography, continentality and sea surface temperatures are also influencing regional variability.

When our records are compared to known and implied hominin evolutionary events and dispersals, we see that the first appearance of the Middle Stone Age tool assemblages in

Morocco and Kenya occur during periods of high climate variability in both northern Africa and the southern Levant and eastern Africa. Similarly, anatomically modern features in fossil human morphology begin appearing across the study area after periods of high variability in all regions.

Dispersals events within and out of Africa seem to be timed when the source region for the population switches from periods of high variability to relative climate stability. Specifically, the dispersals out of Africa of 125 ka – 90 ka suggested by archaeological records occurred as northern Africa and the southern Levant experienced a period of extended variability from 130 –

110 ka, whereas the 65 ka dispersal out of Africa inferred from molecular phylogenetic data occurred after an extended period of climate instability in eastern Africa, 90 – 60 ka, and northern Africa and the southern Levant, 70 – 60 ka. These loose correlations between our 22 climate variability records and known hominin events offers some support to both the accumulated plasticity hypothesis and the variability selection hypothesis.

REFERENCES

Blome, M.W., Cohen, A.S., Tryon, C.A., Brooks, A.S., Russell, J., 2012. The environmental

context for the origins of modern human diversity: A synthesis of regional variability in

African climate 150,000-30,000 years ago. Journal of Human Evolution. 62, 563–592.

Bonnefille, R., Potts, R., Chalie, F., Jolly, D., Peyron, O., 2004. High-resolution vegetation and

climate change associated with Pliocene Australopithecus afarensis. Proceedings of the

National Academy of Sciences. 101, 33, 12125–12129.

Bootsma, H.A., Hecky, R.E., Hesslein, R.H., Turner, G.F., 1996. Food partitioning among Lake

Malawi nearshore fishes as revealed by stable isotope analyses. Ecology. 77, 1286–1290.

Braun, D.R., Aldeias, V., Archer, W., Arrowsmith, J.R., Baraki, N., Campisano, C.J., Deino,

A.L., DiMaggio, E.N., Dupont-Nivet, G., Engda, B., Feary, D.A., Garello, D.I.,

Kerfelew, Z., McPherron, S.P., Patterson, D.B., Reeves, J.S., Thompson, J.C., Reed,

K.E., 2019. Earliest known Oldowan artifacts at 2.58 Ma from Ledi-Geraru, Ethiopia,

highlight early technological diversity. Proceedings of the National Academy of

Sciences.116, 24,11712–11717.

Campisano, C.J., Cohen, A.S., Arrowsmith, J.R., Asrat, A., Behresmeyer, A.K., Brown, E.T.,

Deino, A.L., Deocampo, D., Feibel, C.S., Kingston, J., Lamb, H.F., Lowenstein, T.K.,

Noren, A.J., Olago, D.O., Owen, R.B., Pelletier, J.D., Potts, R., Reed, K.E., Renaut,

R.W., Russell, J.M., Russell, J.L., Shäbitz, F., Stone, J.R., Trauth, M.H., Wynn, J.G.,

2017. New high-resolution paleoclimate records from the East African Rift System and 23

their implications for understanding the environmental context of hominin evolution.

PaleoAnthropology. 1–43.

Campisano, C.J., Feibel, C.S., 2007. Connecting local environmental sequences to global climate

patterns: evidence from the hominin-bearing Hadar Formation, Ethiopia. Journal of

Human Evolution.53, 5, 515–527.

Cohen, A., Campisano, C., Arrowsmith, R., Asrat, A., Behrensmeyer, A.K., Deino, A., Feibel,

C., Hill, A., Johnson, R., Kingston, J., Lamb, H., Lowenstein, T., Noren, A., Olago, D.,

Owen, R.B., Potts, R., Reed, K., Renaut, R., Schäbitz, F., Tiercelin, J.J., Trauth, M.H.,

Wynn, J., Ivory, S., Brady, K., O’Grady, R., Rodysill, J., Githiri, J., Russell, J., Foerster,

V., Dommain, R., Rucina, S., Deocampo, D., Russell, J., Billingsley, A., Beck, C.,

Dorenbeck, G., Dullo, L., Feary, D., Garello, D., Gromig, R., Johnson, T., Junginger, A.,

Karanja, M., Kimburi, E., Mbuthia, A., McCartney, T., McNulty, E., Muiruri, V.,

Nambiro, E., Negash, E.W., Njagi, D., Wilson, J.N., Rabideaux, N., Raub, T., Sier, M.J.,

Smith, P., Urban, J., Warren, M., Yadeta, M., Yost, C., Zinaye, B., 2016. The Hominin

Sites and Paleolakes Drilling Project: Inferring the environmental context of human

evolution from eastern African deposits. Scientific Drilling. 21, 1–16.

Dart, R.A., 1925. Australopithecus africanus: The man-ape of South Africa. Nature. 115, 195.

Deino, A.L., Behrensmeyer, A.K., Brooks, A.S., Yellen, J.E., Sharp, W.D., Potts, R., 2018.

Chronology of the Acheulean to Middle Stone Age transition in eastern Africa. Science.

360, 95–98.

Deino, A.L., Kingston, J.D., Glen, J.M., Edgar, R.K., Hill, A., 2006. Precessional forcing of 24

lacustrine sedimentation in the late Cenozoic Chemeron Basin, Central Kenya Rift, and

calibration of the Gauss/Matuyama boundary. Earth and Planetary Science Letters. 247,

41–60.

Deino, A.L., Sier, M., Garello, D., Keller, B., Kingston, J., Scott, J., Dupont-Nivet, G., Cohen,

A. 2019. Chronostratigraphy of the Baringo-Tugen-Barsemoi (HSPDP-BTB-1A Core -

40Ar/39Ar dating, magnetostratigraphy, tephrostratigraphy, sequence stratigraphy and

Bayesian age modeling. Palaeogeogr. Paleoclimatol.Palaeoeco. In press. deMenocal, P., 1995. Plo-Pleistocene African Climate. Science. 270, 53–59.

Faith, J.T., Behrensmeyer, A.K., 2013. Climate change and faunal turnover: testing the

mechanics of the turnover-pulse hypothesis with South African fossil data. Paleobiology.

39, 609–627.

Grove, M., 2011. Speciation, diversity, and Mode 1 technologies: The impact of variability

selection. Journal of Human Evolution. 61, 306–319.

Grove, M., 2012. Orbital dynamics, environmental heterogeneity, and the evolution of the human

brain. Intelligence. 40, 404–418.

Grove, M., 2014. Evolution and dispersal under climatic instability: A simple evolutionary

algorithm. Adaptive Behavior.22, 235–254.

Grove, M., 2015. Palaeoclimates, plasticity, and the early dispersal of Homo sapiens.

Quaternary International. 369, 17–37.

Kingston, J.D., 2007. Shifting adaptive landscapes: Progress and challenges in reconstructing

early hominid environments. Yearbook of Physical Anthropology. 134, 20–58

Kingston, J.D., Deino, A.L., Edgar, R.K., Hill, A., 2007. Astronomically forced climate change 25

in the Kenyan Rift Valley 2.7-2.55 Ma: implications for the evolution of early hominin

ecosystems. Journal of Human Evolution. 53, 487–503.

Lee-Thorp, J.A., Sponheimer, M., Luyt, J., 2007. Tracking changing environments using stable

carbon isotopes in fossil tooth enamel: an example from the South African hominin sites.

Journal of Human Evolution. 53, 595–601.

Maslin, M.A., Brierley, C.M., Milner, A.M., Shultz, S., Trauth, M.H., Wilson, K.E., 2014. East

African climate pulses and early human evolution. Quaternary Science Reviews. 101, 1–

17.

Maslin M.A., Trauth M.H., 2009. Plio-Pleistocene East African Pulsed Climate Variability and

Its Influence on Early Human Evolution. In: Grine F.E., Fleagle J.G., Leakey R.E. (eds)

The First Humans – Origin and Early Evolution of the Genus Homo. Vertebrate

Paleobiology and Paleoanthropology. Springer, Dordrecht, 151–158.

Potts, R., 1996. Evolution and climate variability. Science. 273, 922–923

Potts, R., 1998. Variability selection in hominid evolution. Evolutionary Anthropology.

Reinthal, P.N., Cohen, A.S., Dettman, D.L., 2011. Fish fossils as paleo-indicators of

ichthyofauna composition and climatic change in Lake Malawi, Africa. Palaeogeography,

Palaeoclimatology, Palaeoecology. 303, 126–132.

Reynolds, S.C., 2007. Mammalian body size changes and Plio-Pleistocene environmental shifts:

implications for understanding hominin evolution in eastern and southern Africa. Journal

of Human Evolution. 53, 528–548.

Richter, D., Grün, R., Joannes-Boyau, R., Steele, T.E., Amani, F., Rué, M., Fernandes, P., 26

Raynal, J.P., Geraads, D., Ben-Ncer, A., Hublin, J.J., McPherron, S.P., 2017. The age of

the hominin fossils from Jebel Irhoud, Morocco, and the origins of the Middle Stone Age.

Nature. 546, 293.

Shultz, S., Maslin, M., 2013. Early Human Speciation, Brain Expansion and Dispersal

Influenced by African Climate Pulses. PLoS ONE. 8. e76750.

Tierney, J.E., Russell, J.M., Huang, Y., Sinninghe Damsté, J.S., Hopmans, E.C., Cohen, A.S.,

2008. Northern hemisphere controls on tropical southeast African climate during the past

60,000 years. Science. 322, 252–255

Trauth, M.H., Larrasoaña, J.C., Mudelsee, M., 2009. Trends, rhythms and events in Plio-

Pleistocene African climate. Quaternary Science Reviews. 28, 399–411.

Trauth, M.H., Maslin, M.A., Deino, A.L., Strecker, M.R., Bergner, A.G.N., Dühnforth, M.,

2007. High- and low-latitude forcing of Plio-Pleistocene East African climate and human

evolution. Journal of Human Evolution. 53, 475–486.

Vrba, E.S., 1993. Turnover-pulses, the Red Queen, and related topics. American Journal of

Science.293-A, 418–452.

Wilson, K.E., Maslin, M.A., Leng, M.J., Kingston, J.D., Deino, A.L., Edgar, R.K., Mackay,

A.W., 2014. East African lake evidence for Pliocene millennial-scale climate variability.

Geology. 42, 955–958.

27

APPENDIX A:

δ13C RECORDS FROM FISH FOSSILS AS PALEO-INDICATORS OF ECOSYSTEM

RESPONSES TO LAKE LEVELS IN THE PLIO-PLEISTOCENE LAKES OF TUGEN HILLS,

KENYA

BILLINGSLEY, Anne L.1, REINTHAL, Peter2, DETTMAN, David L.1, KINGSTON, John D.3,

DEINO, Alan L.4, ORTIZ, Kevin5, MOHLER, Benjamin1, and COHEN, Andrew S.1

1. Department of Geosciences, University of Arizona, Tucson, AZ

2. Department of Ecology and Evolutionary Biology, University of Arizona, Tucson, AZ

3. Department of Anthropology, University of Michigan, Ann Arbor, MI

4. Berkeley Geochronology Center, Berkeley, CA

5. Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, MI

Accepted to the Journal of Palaeogeography, Palaeoclimatology, Palaeoecology

28

Key Points 1. The paleolakes of Baringo Basin, Kenya responded quickly to changes in climate during the Plio-Pleistocene transition. 2. Lake level changes were abrupt enough in the Baringo Basin, Kenya through the Plio- Pleistocene to prevent near-shore, shallow ecosystems 3. δ13C values of fish fossils from HSPDP-BTB13-1A (2.56 – 3.29 Ma) indicate dependency on pelagic food sources 4. Fish communities may have been extremely vulnerable to environmental changes, such as algal blooms. δ13C values of fish bones from Lake Turkana suggest that differentiation between benthic and planktonic feeders can be made in lakes with a shallowed photic zone

Abstract

The carbon isotopic ratios of organic matter in fish fossils from diatomites and other lake beds in the HSPDP drill core from Tugen Hills, Kenya (2.56 – 3.29 Ma) reflect trophic resource uses, and can indicate the dietary habitats of fish in the paleolake. This information offers insight into how fish communities responded to lake-level fluctuations during the Plio-Pleistocene in the

East African Rift Valley. We have compared this record with fish fossil isotopes from both a previously published study of a Lake Malawi drill core (139 ka - present) and core top (modern ca 1978) samples collected at the water/sediment boundary from Lake Turkana (Kenya) of known environmental provenance. Both the Lake Malawi drill core fossils (‒7.2 ‰ to ‒27.5 ‰

VPDB) and modern Lake Turkana samples (‒16 ‰ to ‒24.6 ‰ VPDB) have δ13C values indicating a mix of near-shore and deep-water pelagic species. In contrast, the δ13C values for the

Tugen Hills core fossils vary only between ‒20 ‰ and ‒27 ‰ VPDB. The absence of δ13C values greater than ‒19‰, suggests none of these fossils are derived from near-shore benthic habitats. The lack of shallow water, benthic lacustrine fish fossils through the Tugen Hills lake cycles may indicate that the rate of change from low-lake stands to deeper lake phases was very rapid, and shallow water communities were not established for long enough to leave a fish fossil record at the core site. These results strongly suggest that lake level responses to climate 29 variability in the Baringo basin of the East African Rift were very abrupt during the Plio-

Pleistocene transition.

1. Introduction

Isotopic records from fish have been used to determine many aspects of the environment such as the source water of lacustrine system, the functional redundancy and vulnerability of fish populations, the food partitioning within communities, and migration patterns (Bootsma et al.,

1996; Zazzo et al., 2006; Joordens et al., 2011; Reinthal et al., 2011; Gownaris et al., 2015). The isotopic composition of both extant fish and fish fossils can offer insight into their ecological strategies and how they responded to climate and environmental changes. As the biological processes that control the isotopic values of fish communities operate similarly in most lacustrine environments, comparisons can be made between basins in different geographical locations as well as through time.

In 2013, a 227-meter core spanning the Plio-Pleistocene transition was taken from the

Chemeron Formation of the Baringo Basin in the Tugen Hills region of Kenya as part of the

Hominin Sites Paleolake Drilling Project (HSPDP-BTB13-1A, hereafter BTB13-1A) (Cohen et al., 2016). The drill site was chosen because of the paleoanthropological significance of the basin, which includes the recovery of several hominin specimens as well as relevant paleoenvironmental data(Cohen et al., 2016; Campisano et al., 2017). Prior detailed outcrop studies of lacustrine deposits provide a well-dated context for the investigation of this core

(Deino et al., 2006; Kingston, 2007; Wilson et al., 2014). This work had previously indicated a series of relatively large lakes occupying this portion of the central Kenyan rift during the Plio-

Pleistocene, which underwent lake level fluctuations driven by both precessional forcing and millennial-scale climate variability. 40Ar/39Ar dating of BTB13-1A core shows the core ranges in 30 age from approximately 2.56 – 3.29 Ma, encompassing the Plio-Pleistocene climate transition and the apparent in hominin evolution, including the emergence of the

Paranthropus and Homo lineages (Deino et al., this volume). Further details on the coring, project objectives and core stratigraphic context can be found in Cohen et al. (2016) and

Kingston et al., this volume. Various records suggest that prior to ~3.05 Ma the deposystem was dominated by , fluvial, and small-lake deposystems, whereas lacustrine conditions became more prevalent after that time (Scott et al., this volume; Westover et al., this volume;

Yost et al., this volume). Gamma density, magnetic susceptibility, and total organic matter content records from the post 3.05 Ma part of the core show that the lake went through multiple transgressions and regressions, some of which were geologically rapid, interpreted to represent relative climate instability (Fig. 1).

One of these periods of instability is represented in the drill core by a series of five diatomaceous units between 10 and 55 m (~2.70 – 2.58 Ma). Prior investigations indicated that these diatomites formed during deep-lake phases, which were separated by fluvial deposits and paleosols deposited when the lake floor was exposed for extended periods during lake-level low stands (Deino et al., 2006; Kingston et al., 2007; Wilson et al., 2014). The five diatomite layers are dominated by two planktonic diatom genera, Stephanodiscus and Aulacoseira (Kingston et al., 2007; Westover et al., this volume). The assemblages and the condition of the diatoms suggest the lake was at least 40 m deep during these periods (Kingston et al., 2007). However, sediments from these deep-lake phases lack the abundance of fish bones seen within other depositional environments in outcrops, raising interesting questions about the responses of fish communities to the timing and extent of deep lake-level pulses. 31

For this study, we have investigated the record of δ13C changes in organic carbon isolated from fish fossils from the BTB13-1A core to determine the paleo-environments that the fish inhabited in the Plio-Pleistocene lakes of the central Kenyan rift. Species or genus level identifications of fossil fish material preserved in the cores is generally impractical because most of the fossil remains are too small. Fish fossils isotopic compositions have previously proven useful in establishing paleo-food web dynamics of aquatic environments (Bootsma et al., 1996;

Genner et al., 1999; Reinthal et al., 2011). Fish communities of African rift valley lakes are often highly diverse, especially among members of the family Cichlidae, and have exhibited a remarkable capacity for speciation and adaptation to new ecological niches (Salzburger et al.,

2014). These morphological and behavioral adaptations often reflect evolutionary convergence in feeding specializations between lake systems (Burress, 2014; Salzburger et al., 2014). A variety of fish taxa occupy both benthic (demersal) and pelagic environments. In large lakes comparable to the Tugen Hills paleolakes at their maximum extent, benthic algal production (i.e. algae growing on the bottom of the lake or on rock surfaces) is an important trophic resource for many and non-cichlid fish species from 0 – 30 m water depth, whereas many offshore species obtain their food either directly from planktonic algae (i.e. algae grown within the photic zone of the water column), or by feeding at higher trophic levels on planktivores (Reinthal, 1990;

Duponchelle et al., 2005).

Benthic and pelagic algae typically differ in their δ13C compositions. Carbon in lakes is limited, and algae preferentially incorporate 12C during photosynthesis compared to 13C (Pardue et al., 1976). The amount of carbon available is influenced by the boundary layer of the algae.

Pelagic algae have a smaller boundary layer and much more carbon available because of their smaller size and larger surface area/volume ratio, and grow in zones of the lake most influenced 32 by surface water/atmosphere interactions. Benthic algae grow on the bottom of the lake where the influence of atmospheric carbon dioxide is more limited. Benthic algae also preferentially take up 12C, but 12C becomes proportionately depleted within the boundary layer through its uptake. As time progresses, 13C is increasingly included in the photosynthetic process and benthic algae becomes enriched in 13C relative to pelagic algae (Bootsma et al., 1996). The difference between the two is substantial. Modern studies from Lake Malawi, a large

Afrotropical lake, show δ13C values of benthic algae as positive as ‒6 ‰ whereas planktonic species may be as negative as ‒26 ‰ (Fig. 2) (Bootsma et al., 2016).

There is little fractionation of carbon isotope values within a food chain and thus, isotopic values of higher trophic levels are reflective of the original autotrophic food source (Reinthal,

1990; Bootsma et al., 1996). The difference in carbon uptake between pelagic and benthic algae allows for a spatial determination of the living environment of the fish (Melville and Connolly,

2003; Devlin et al., 2013). Benthic algae are restricted by available light and relegated to the shallower portions of the lake, a pattern that is accentuated in more turbid lakes. Higher (less negative) δ13C values in fish fossils would suggest a benthic resource and a shallow habitat.

Lower (more negative) δ13C values would suggest a phytoplankton resource and relatively deeper-water habitat (Fig. 2). Lake depth fluctuations through time would cause a migration of the benthic/pelagic boundary through the BTB13-1A core site. These migrations would then be recorded by the δ13C values of fish communities represented in the core. As the lake gets deeper, the food resources would become more pelagic with lower δ13C values. As the lake regresses and becomes shallower, the food resources become more benthic, reflected in higher δ13C (Reinthal,

1990; Bootsma et al., 1996). 33

If the diatomites were deposited rapidly, we hypothesize that they would overwhelm the fossil record, producing fish fossil isotopes with similar values before and after deposition, most likely reflecting the lower values of pelagic food sources. If the diatomites were deposited in a series of algal blooms, similarly we would expect to observe lower δ13C values both before and after the diatomites; however, we would expect to observe an increase in abundance of fish fossil prior to deposition. If the fish community did not have time to respond to the lake level pulses, then there would only be a small range in δ13C values representing one preferred habitat type.

13 For this study, we have compared δ Corg values of fish fossils (bones, teeth and scales) from the BTB13-1A core with bones, teeth, scales, and flesh of modern fish samples as well as core top fish bones from Lake Turkana (northern Kenya) and Pleistocene samples from a Lake

Malawi drill core (Fig. 3). Lake Turkana is a potential modern analog for the Chemeron

Formation paleolakes, given its proximity and possible similar range of ecological conditions.

Lake Turkana has experienced large scale fluctuations in lake levels, with approximately 100 m difference between high- and low-stands during the Quaternary (Owen and Renaut, 1986; Garcin et al., 2009). Modern trophic studies from Lake Turkana show some functional redundancy in the fish communities, but most species have relatively low isotopic overlap (Gownaris et al.,

2015). The lake is also turbid, which controls the depth of the photic zone, and therefore, may shift the δ13C boundary to shallower and more nearshore positions.

In 2005, core MAL-05 1C, was collected from Lake Malawi. This drill core, which penetrated to 89.9 mblf (6.51 m overpenetration) contains numerous fish fossils at all depths and spans from ~139 – 10 ka (Scholz et al., 2011; Ivory et al., 2016). Lake Malawi has the most diverse freshwater fish fauna of any lake in the world, overwhelmingly dominated by cichlid species. Many studies have focused on habitat and food partitioning in Lake Malawi, which has 34 generated an extensive data set relating δ13C values from its fish and their associated habitats and food webs (Reinthal, 1990; Bootsma et al., 1996). δ13C values of fish bones from the Lake

Malawi drill core show that lower values are associated with deeper lake phases through the late

Pleistocene, as would be predicted from modern habitat and trophic relationships (Reinthal et al.,

2011). This Malawi study indicates that despite their small size and thus our inability to precisely identify the BTB13-1A fish bones, the isotopic record can potentially provide insight into the rate of ecological and depositional changes within the paleolake’s history.

2. Methods

Magnetic susceptibility and gamma density were determined using a GEOTEK multi- sensor core logger (MSCL) at the National Lacustrine Core Facility, University of Minnesota

(LacCore). Gamma density data were screened to remove spurious low values caused by core voids (densities < 1.0 g/cm3). Subzero MS values were also removed from that data set because of the difficultly in distinguishing if these points represent drill water, organic material or a void in the core. Total organic carbon was determined using standard loss-on-ignition techniques at the University of Arizona (Bengtsson and Enell, 1986).

The BTB13-1A core was sampled every 32 cm. The samples were then disaggregated by freezing and thawing or by using a kerosene treatment. There is no indication that the kerosene treatment influences the outcome of the isotopic analysis. The samples were screen washed using distilled water on a 120-micron sieve, and then examined using a Leica M165 stereomicroscope.

Fish bones, teeth and scales were counted and calculated per dry gram weight. Close examination of the fossils showed that there was no apparent overgrowth or contamination of the bone structure. The fossils were selected for isotope analyses from different sediment compositions ranging from clays to coarse sands to ensure that all possible environments were 35 tested. The fossils were mainly bone fragments, but some teeth and complete vertebrae were also processed. Multiple fish fossils were run for three horizons in order to determine homogeneity of the δ13C values.

We determined δ13C values of modern fish bones from 12 samples taken along three water depth transects in Lake Turkana, Kenya. The samples were taken from modern core tops

(collected in 1978), sampling the water/sediment interface. The shallowest depth is 2 m whereas the deepest is 38 m. (Table 2). The samples were processed using the same procedures used for the BTB13-1A samples.

Modern fish samples of niloticus ( tilapia) were raised in common culture at the University of Arizona Aquaponics Greenhouse. This modern analog was chosen because Oreochromis niloticus baringoensis is an endemic subspecies found in , a freshwater lake with seven fish species near the core site. Bones, teeth, scales and flesh were removed and pretreated twice with a 2:1 chloroform/methanol solution in order to remove excess lipids which would influence the δ13C values.

Stable isotope data were collected on a continuous-flow gas-ratio spectrometer (Finnigan

Delta PlusXL). Samples were combusted using an elemental analyzer (Costech) coupled to the mass spectrometer. Standardization is based on acetanilide for elemental content and NBS-22 and USGS-24 for δ13C. The following isotope ratios are assigned to these standards (δ13C

VPDB): USGS24, ‒16.05 ‰; NBS 22, ‒30.03 ‰. Analytical precision is better than ± 0.08 for

δ13C (1σ), based on repeated internal standards. Samples that were extremely small have a higher uncertainty: standards of equivalent size had a variability of ±0.25 ‰ (1σ). Because the amount of organic carbon in these samples was extremely small, extra care was taken to eliminate inorganic carbon contamination. On average, fossilized fish material contains 0.25 – 0.5 times 36 the amount of organic carbon as modern samples of similar size. Bone and tooth samples were acidified with sulfurous acid (H2SO3) in silver foil capsules to remove structural carbonate from the bone apatite. Two drops of acid were used followed by overnight drying in 60 ⁰C oven. This procedure was repeated once. The silver foil capsules were heated to ~500 ⁰C.

3. Results

3.1 BTB13-1A core samples

Approximately 1,200 fish scales, teeth, and bone were counted from 592 samples taken throughout the BTB13-1A core. Most of the material is unidentifiable fragments; however, the fossils that can be identified come from the families Cichlidae and Cyprinidae. There are very few fish bones in the dominantly fluvial-wetland-shallow pond deposits in the lower portions of the core, below 132 mbs (3.04 Ma) (Fig. 1). Only two segments in the lower portion, at 183.6 -

181.9 and 174.6 mbs (3.17 and 3.15 Ma respectively), contain fish material. Above 132 mbs, the samples that contain an abundance of fish fossils coincide with higher percentages of total organic content and lower values for gamma density and magnetic susceptibility. These intervals of the core are often clayey or silty, and are associated with lacustrine phases as inferred from both litho- and bio-facies analysis. A few samples towards the top are associated with sands which have been interpreted as being deposited in a delta front within the lake (Scott et al., this volume) (Fig. 1).

δ13C values were determined for 36 fish fossils throughout the Tugen Hills drill core

(Table 1). Values range between ‒20 ‰ to ‒27 ‰. The low δ13C values correlate with zones of low magnetic susceptibility and deeper water environments (Fig. 4).

37

3.2 Carbon isotopic fractionation in modern fish

To supplement the interpretive isotopic framework for the fossil fish analyses, δ13C values for scales, bones, oral teeth and muscle (flesh) from the cichlid, Oreochromis niloticus

() raised in common culture at the University of Arizona Aquaponics Greenhouse were determined. Four specimens were raised in the same habitat. Fish 1 and 2 are juveniles whereas fish 3 and 4 are adults. The δ13C values ranged from ‒17.0 ‰ to ‒23.1 ‰, with an uncertainty of ± 0.33 ‰ for all samples but the flesh from fishes 2,3 and 4 have an uncertainty of

± 0.10 ‰. Fractionation seems to occur in individual specimens as well as different sample types. However, the range of δ13C average values between the four specimens are similar to ranges observed among modern from Lake Malawi, where fish populations can vary by approximately 1.5 ‰ (Bootsma et al., 1996) (Fig. 5). The fish flesh is significantly more negative than bone, teeth and scales because fish oils are depleted in carbon-13.

3.3 Modern Lake Turkana

δ13C values of core top samples range from ‒24.6 ‰ to ‒16.8 ‰. Higher values are found in samples from water depths shallower than 5 m. The lowest value occurs at 11.5 m (‒24.6 ‰), but this sample was collected relatively close to the shore, which may indicate it was reworked from previously deposited sediments from the nearby and much higher Early Holocene lake stand terraces that surround the lake today (Owen and Renaut, 1986; Garcin et al., 2009) (Fig. 6).

δ13C values of samples from sediments deposited in deeper water are between ‒20.1 ‰ and ‒

21.6 ‰, which suggest a more planktonic food source than found at shallower water depths. The isotope values from this study are similar to those found in the living fish species of Lake

Turkana, which range between ‒20.18 ‰ to ‒16.88 ‰ (Gownaris et al., 2015).

38

4. Discussion

There are no δ13C values for fossil fish from the BTB13-1A core within the range of shallow lacustrine environments (i.e. benthic algivores) determined through modern fish δ13C ranges (‒7.0 ‰ to ‒19 ‰ VPDB (Fig.2) (Bootsma et al., 1996; Reinthal et al., 2011). It is possible, though unlikely, that this uniformity of results is a consequence of low sampling frequency in this study, meaning that not all environments were tested. Higher resolution sampling is needed to determine if benthic species are in other sedimentary units. A more probable explanation is that fluctuations between the core site exposure and deep lake conditions were sufficiently rapid at the core location that the lake was rarely shallow. This may be consistent with the observation that exposure surfaces with pedogenesis throughout the core typically lie directly below silty-clay profundal lacustrine flooding surfaces (Scott et al., this volume).

When compared with the diatom record from the core, we see that fish fossils are present in five horizons with an abundance of diatoms: 8.5 – 10.95 mbs, 17.08 – 22 mbs, 25.29 – 39.27 mbs, 40.98 – 43.48 mbs and 50.67 – 52.29 mbs (Westover et al., this volume). However, the abundance of fish fossils is not consistent throughout the diatomaceous units within the core, and often large segments within these units have no fish material. This may suggest that the lack of fish fossils in the diatomaceous units of the outcrop may have resulted from an increased sedimentation rates of diatoms overwhelming the deposition of fish material.

The range of δ13C values found in the fish fossils from BTB13-1A are within the range of values seen in modern fish communities seen at Lake Malawi (Bootsma et al., 1996) (Fig. 7).

The BTB13-1A values, when compared to modern fish communities studied in Lake Malawi, indicate pelagic food sources, suggesting that even during low-lake stands, the lake was deep 39 enough to preclude benthic algae as a significant food resource or a benthic fish community.

However, published modern fish values were obtained from analysis of tissue rather than bone, and a number of factors could have influenced these values such as water chemistry, terrestrial input, post-depositional alterations, and life-cycles (Genner et al., 1999; Wurster and Patterson,

2001; Bootsma et al., 2016). Reinthal et al. (2011) demonstrate δ13C isotope values in MAL 05-

1C follow the trend of modern fish communities of lower values coinciding with deeper lakes and higher values coinciding with shallower lakes despite the difference in material used. Our modern fish samples showed that a range of δ13C values can occur not only throughout fish communities but within different tissue types from individual specimens as well. However, the flesh exhibits significantly lower δ13C values than either the bone or the teeth. This means that the bone and teeth of the BTB13-1A fossil fish represent the highest δ13C values of the fish communities. Furthermore, there is little overlap between the δ13C values of the BTB13-1A samples and those from modern Lake Turkana (one-way ANOVA shows no substantial overlap, p = 7.96 x 10-10). This indicates that the fish living in the Tugen Hills paleolakes were more dependent on pelagic food sources than those currently living in Lake Turkana (Fig. 7).

Environmental interpretations can become further complicated in situations where sediment and terrestrial carbon may be transported into the system from elsewhere. This can be seen in one of our Lake Turkana samples, where a nearshore, shallow environment contained a bone most likely deposited from the shoreline where fossils from an older, deeper lake can be found. However, our study does show that even in an environment such as Lake Turkana where the photic zone may be restricted in depth by increased turbidity, there is still a pattern of higher

δ13C values in environment shallower than 5 m and lower values in deeper waters (Table 2;

Figure 6). 40

Fluvial systems may be difficult to differentiate from deep-water habitats solely on the

δ13C values of fish fossils. In a fluvial or deltaic system, the water is usually turbulent, and both pelagic and benthic algae would have their boundary layer continuously replenished in carbon.

Fish living in a river or near a delta environment have lower δ13C values than shallow lacustrine species at the same depth (Hoffman et al., 2010). Significant fluvial inputs from a nearby delta may explain the anomalous combination of low δ13C values and relatively high MS values for sediment horizon HSPDP-BTB13-1A-4Q-1, 32-34 cm. The sample is a relatively coarse-grained sandy unit, which has been determined through other lithologic methods (trace fossils, sedimentary structures and composition) to be a deltaic deposit (Scott et al., this volume).

5. Conclusions

The isotopic analysis of fish fossils can be an important tool when determining how fish communities respond to climate and environmental variability, especially when the changes are dramatic and quick. We have investigated fossil fish abundances and carbon isotopes from a drill core record of paleolakes that occupied the central Kenyan Rift Valley during the Late Pliocene-

Early Pleistocene. This study of the HSPDP-BTB13-1A fossil record is the first application of

δ13C analyses to fish fossils of this age in . The δ13C values are within the range of those seen in both modern fish community assemblages and fish fossils of Lake Malawi. The

BTB13-1A values, however, are lower than those of modern Lake Turkana, which suggests that the fish of the paleolake were more dependent on pelagic food sources than those currently present in Lake Turkana. The lack of shallow benthic δ13C values for the fish fossils indicates that the transitions between shallow and deep lake conditions were abrupt. The narrow range in

δ13C values also suggests high dependency on a phytoplanktonic food source that would make the fish communities more vulnerable to sudden changes in the pelagic community. This study 41 shows that the δ13C of fish fossils may be used in tandem with other environmental indicators as a tool to better understand the timing and extent of lake-level pulses as East African climate became more highly variable during the Plio-Pleistocene transition.

Acknowledgements

The National Science Foundation (EAR1123000, EAR1338553, BCS124185) and the

International Continental Drilling Program provided support for this research. MSCL core logging was done at the National Lakes Core Repository (LacCore). We thank LacCore, Kristina

Brady, Jessamy Doman, Sarah Ivory, Beau Marshall, Dennis Njagi, Anders Noren, Ryan

O’Grady, Steve Rucina, Rich Szenmiklosi, Gladys Tuitoek, Chad Yost, Jacob Marker, and Julia

Richter for their assistance with this work. This is publication 20 of the Hominin Sites and

Paleolakes Drilling Project (HSPDP).

References

Bengtsson, L.; Enell, M.., 1986. Chemical analysis. In: Berglund, B.E. (Ed.), Handbook of

Holocene Palaeoecology and Palaeohydrology. John Wiley and Sons, pp. 423–451.

Bootsma, H.A., Hecky, R.E., Hesslein, R.H., Turner, G.F., 1996. Food partitioning among Lake

Malawi nearshore fishes as revealed by stable isotope analyses. Ecology. 77, 1286–1290.

Bootsma, H.A., Hecky, R.E., Hesslein, R.H., Turner, G.F., Turner, G.F., 2016. Food Partitioning

Among Lake Malawi Nearshore Fishes as Revealed by Stable Isotope Analyses Linked

references are available on JSTOR for this article : 77, 1286–1290.

Burress, E.D., 2014. Cichlid fishes as models of ecological diversification: patterns, mechanisms,

and consequences. Hydrobiologia. 7–27.

Campisano, C.J., Cohen, A.S., Arrowsmith, J.R., Asrat, A., Behresmeyer, A.K., Brown, E.T., 42

Deino, A.L., Deocampo, D., Feibel, C.S., Kingston, J., Lamb, H.F., Lowenstein, T.K.,

Noren, A.J., Olago, D.O., Owen, R.B., Pelletier, J.D., Potts, R., Reed, K.E., Renaut, R.W.,

Russell, J.M., Russell, J.L., Shäbitz, F., Stone, J.R., Trauth, M.H., Wynn, J.G., 2017. New

high-resolution paleoclimate records from the East African Rift System and their

implications for understanding the environmental context of hominin evolution.

PaleoAnthropology. 1–43.

Cohen, A., Campisano, C., Arrowsmith, R., Asrat, A., Behrensmeyer, A.K., Deino, A., Feibel,

C., Hill, A., Johnson, R., Kingston, J., Lamb, H., Lowenstein, T., Noren, A., Olago, D.,

Owen, R.B., Potts, R., Reed, K., Renaut, R., Schäbitz, F., Tiercelin, J.J., Trauth, M.H.,

Wynn, J., Ivory, S., Brady, K., O’Grady, R., Rodysill, J., Githiri, J., Russell, J., Foerster, V.,

Dommain, R., Rucina, S., Deocampo, D., Russell, J., Billingsley, A., Beck, C., Dorenbeck,

G., Dullo, L., Feary, D., Garello, D., Gromig, R., Johnson, T., Junginger, A., Karanja, M.,

Kimburi, E., Mbuthia, A., McCartney, T., McNulty, E., Muiruri, V., Nambiro, E., Negash,

E.W., Njagi, D., Wilson, J.N., Rabideaux, N., Raub, T., Sier, M.J., Smith, P., Urban, J.,

Warren, M., Yadeta, M., Yost, C., Zinaye, B., 2016. The Hominin Sites and Paleolakes

Drilling Project: Inferring the environmental context of human evolution from eastern

African rift lake deposits. Scientific Drilling. 21, 1–16.

Deino, A.L., Kingston, J.D., Glen, J.M., Edgar, R.K., Hill, A., 2006. Precessional forcing of

lacustrine sedimentation in the late Cenozoic Chemeron Basin, Central Kenya Rift, and

calibration of the Gauss/Matuyama boundary. Earth and Planetary Science Letters. 247, 41–

60.

Deino, A.L., Sier, M., Garello, D., Keller, B., Kingston, J., Scott, J., Dupont-Nivet, G., Cohen,

A. 2019. Chronostratigraphy of the Baringo-Tugen-Barsemoi (HSPDP-BTB-1A Core - 43

40Ar/39Ar dating, magnetostratigraphy, tephrostratigraphy, sequence stratigraphy and

Bayesian age modeling. Palaeogeogr. Paleoclimatol.Palaeoeco. In press.

Devlin, S.P., Vander Zanden, M.J., Vadeboncoeur, Y., 2013. Depth-specific variation in carbon

isotopes demonstrates resource partitioning among the littoral zoobenthos. Freshwater

Biology. 58, 2389–2400.

Duponchelle, F., Ribbink, A.J., Msukwa, A., Mafuka, J., Mandere, D., Bootsma, H., 2005. Food

partitioning within the species-rich benthic fish community of Lake Malawi, East Africa.

Canadian Journal of Fisheries and Aquatic Sciences. 62, 1651–1664.

Garcin, Y., Junginger, A., Melnick, D., Olago, D.O., Strecker, M.R., Trauth, M.H., 2009. Late

Pleistocene-Holocene rise and collapse of , northern Kenya Rift. Quaternary

Science Reviews. 911–925.

Genner, M.J., Turner, G.F., Barker, S., Hawkins, S.J., 1999. Niche segregation among Lake

Malawi cichlid fishes? Evidence from stable isotope signatures. Ecology Letters. 2, 185–

190.

Gownaris, N.J., Pikitch, E.K., Ojwang, W.O., Michener, R., Kaufman, L., 2015. Predicting

species’ vulnerability in a massively perturbed system: The fishes of Lake Turkana, Kenya.

PLoS ONE. 10(5):e0127027 doi10.1371/journal.pone.0127027

Hoffman, J.C., Peterson, G.S., Cotter, A.M., Kelly, J.R., 2010. Using Stable Isotope Mixing in a

Great Lakes Coastal Tributary to Determine Food Web Linkages in Young Fishes. Estuaries

and Coasts. 33, 1391–1405.

Ivory, S.J., Blome, M.W., King, J.W., McGlue, M.M., Cole, J.E., Cohen, A.S., 2016. 44

Environmental change explains cichlid adaptive radiation at Lake Malawi over the past 1.2

million years. Proceedings of the National Academy of Sciences. 113, 11895–11900.

Joordens, J.C.A., Vonhof, H.B., Feibel, C.S., Lourens, L.J., Dupont-Nivet, G., van der Lubbe,

J.H.J.L., Sier, M.J., Davies, G.R., Kroon, D., 2011. An astronomically-tuned climate

framework for hominins in the Turkana Basin. Earth and Planetary Science Letters. 307, 1–

8.

Kingston, J.D., 2007. Shifting adaptive landscapes: Progress and challenges in reconstructing

early hominid environments. Yearbook of Physical Anthropology. 20–58.

Kingston, J.D., Deino, A.L., Edgar, R.K., Hill, A., 2007. Astronomically forced climate change

in the Kenyan Rift Valley 2.7-2.55 Ma: implications for the evolution of early hominin

ecosystems. Journal of Human Evolution. 53, 487–503.

Melville, A.J., Connolly, R.M., 2003. Spatial analysis of stable isotope data to determine primary

sources of nutrition for fish. Oecologia. 136, 499–507.

Owen, R.B., Renaut, R.W., 1986. Sedimentology, stratigraphy and palaeoenvironments of the

Holocene Galana Boi Formation, NE Lake Turkana, Kenya. In: FROSTICK, L.E. et al.

(Ed.), Sedimentation in the African Rifts. Geological Society, 311–322.

Pardue, J.W., Scalan, R.S., Van Baalen, C., Parker, P.L., 1976. Maximum carbon isotope

fractionation in photosynthesis by blue-green algae and a green alga. Geochimica et

Cosmochimica Acta. 40, 309–312.

Reinthal, P.N., 1990. The feeding habits of a group of herbivorous rock-dwelling cichlid fishes

(Cichlidae: Perciformes) from Lake Malawi, Africa. Environmental Biology of Fishes. 27, 45

215–233.

Reinthal, P.N., Cohen, A.S., Dettman, D.L., 2011. Fish fossils as paleo-indicators of

ichthyofauna composition and climatic change in Lake Malawi, Africa. Palaeogeography,

Palaeoclimatology, Palaeoecology. 303, 126–132.

Salzburger, W., Van Bocxlaer, B., Cohen, A.S., 2014. Ecology and Evolution of the African

Great Lakes and Their Faunas. Annual Review of Ecology, Evolution, and Systematics. 45,

519–545.

Scholz, C.A., Cohen, A.S., Johnson, T.C., King, J., Talbot, M.R., Brown, E.T., 2011. Scientific

drilling in the : The 2005 Lake Malawi Scientific Drilling Project - An

overview of the past 145,000years of climate variability in East

Africa. Palaeogeography, Palaeoclimatology, Palaeoecology. 303, 3–19.

Westover, K.S., Stone, J.R., Yost, C.L., Scott, J.J., Cohen, A.S., Rabideaux, N., Stockhecke, M.,

Kingston, J.D. 2019. Diatom paleolimnology of late Pliocene Baringo Basin (Kenya)

paleolakes. Palaeogeogr. Paleoclimatol.Palaeoeco. In review.

Wilson, K.E., Maslin, M.A., Leng, M.J., Kingston, J.D., Deino, A.L., Edgar, R.K., Mackay,

A.W., 2014. East African lake evidence for Pliocene millennial-scale climate variability.

Geology. 42, 955–958.

Wurster, C.M., Patterson, W.P., 2001. Late holocene climate change for the eastern interior

united states: Evidence from high-resolution δ18O values of sagittal otoliths.

Palaeogeography, Palaeoclimatology, Palaeoecology. 170, 81–100.

Yost, C.L., Ivory, S.J., Deino, A.L., Rabideaux, N., Kingston, J.D., Cohen, A.S., 2019. 46

Phytoliths, pollen, and microcharcoal from the Baringo Basin, Kenya reveal savanna

dynamics during the Plio-Pleistocene transition. Palaeogeogr. Palaeoclimatol. Palaeoecol. In

review.

Zazzo, A., Smith, G.R., Patterson, W.P., Dufour, E., 2006. Life history reconstruction of modern

and fossil sockeye salmon (Oncorhynchus nerka) by oxygen isotopic analysis of otoliths,

vertebrae, and teeth: Implication for paleoenvironmental reconstructions. Earth and

Planetary Science Letters. 249, 200–215.

47

Figures

Figure 1: Summary stratigraphy, physical properties and fish fossil abundance from a portion of the Chemeron Formation recovered in the BTB13-1A drill core from the Tugen Hills, Kenya.

Increased magnetic susceptibility and gamma density coincide with periods of increased terrestrially derived sediments, such as siliciclastic sands and gravels, whereas higher total organic content (mostly in fine-grained silts and clays) is associated with more lacustrine phases 48 of the core record. Terrestrial and lacustrine depositional environments were determined using sediment and trace fossil analysis as well as other indicators such as diatoms (Scott et al., in prep; Westover et al., in review).

Figure 2: δ13C values of potential food sources for fish (modified from Bootsma et al., 1996)

49

Figure 3: Lake Turkana (3.3830 N; 36.1167 E) and Lake Malawi (11.2933 S; 34.4358 E) are both positioned in the African rift system allowing for a good comparison between the Plio-

Pleistocene record of BTB13-1A (.5546 N; 35.9375 E) and modern lacustrine systems. (Source:

Esri, DigitalGlobe, Geoeye, Earthstar Geographics, CNES/Airbus DS, USDA, USGS,

AeroGRID, IGN and the GIS User Community) 50

Figure 4: δ13C values of fish fossils from the BTB13-1A plotted along with magnetic susceptibility. 63 % of our isotope values are at depths where the magnetic susceptibility falls below the mean (377.65) for the whole record.

51

Figure 5: δ13C values, with a mean of ‒19.05 ‰ ± 1.12, from four living, tank reared

Oreochromis niloticus reveal that the isotopic values of material can vary over both individual specimens and sample types. The dark line represents median value, boxes represent the upper and lower quartiles, and whiskers represent the upper and lower values of each data set excluding outliers.

52

Figure 6: δ13C values relative to VPDB for modern Lake Turkana fish bones are plotted against lake depth (top) and distance from shore (bottom).

53

13 Figure 7: Summary plot of all δ Corg values determined for the three lake/paleolake data sets:

MAL-05 1C (Reinthal et al., 2011), Lake Turkana and the Plio-Pleistocene lakes of BTB13-1A.

The dark line represents median value, boxes represent the upper and lower quartile, and whiskers represent the upper and lower values of each data set excluding outliers.

54

Tables

Table 1: δ13C data from fish fossils from the HSPDP-BTB13A core.

55

Table 2: δ13C VPBD values were determined for 12 samples from Lake Turkana.

56

APPENDIX B:

ORBITAL AND HIGH-LATITUDE CONTROLS OF PLIO-PLEISTOCENE CLIMATE

VARIABILITY IN AFRICA, IMPLICATIONS FOR EARLY HOMININS AND THE RISE OF

THE GENUS HOMO

BILLINGSLEY, Anne L.1; GROVE, Matt2; ANCHUKAITIS, Kevin3, KINGSTON, John D.4;

DEINO, Alan L.5 and COHEN, Andrew S.1

1. Department of Geosciences, University of Arizona, Tucson, AZ 2. Department of Archaeology, Classics and Egyptology, University of Liverpool, Liverpool, UK 3. School of Geography and Development, University of Arizona, Tucson, AZ 4. Department of Anthropology, University of Michigan, Ann Arbor, MI 5. Berkeley Geochronology Center, Berkeley, CA

Prepared for submission to the Journal of Human Evolution 57

Key Points 1. High percentages of climate variability in Africa can be attributed to orbital parameters on Milankovitch cycles of eccentricity, obliquity and precession 2. A Northern Hemisphere temperature record (NE Russia) and an aridity record from Eastern Africa are significantly correlated 3. Reconstructed time-series above noise are able to elicit modeled evolutionary responses 4. Modeled evolutionary responses of plasticity align well to the paleoanthropological record, specifically at 2.8 Ma, where a peak in predicted plasticity occurs during the first occurrence of the genus Homo 5. Southern Africa could have served as a refugium for many mammals when the environment was unfavorable in other regions

Abstract

A ~230 m drill core was collected from the Tugen Hills, Kenya as part of the Hominin Sites and Paleolakes Drilling Project (HSPDP). 40Ar/39Ar dates indicate that the core record spans the

3.29 – 2.57 Ma interval, providing a rare high-resolution African terrestrial record of the important Plio-Pleistocene transition, when global climate began to cool and glaciation intensified in the Northern Hemisphere. This period was marked by C4 grass expansion in eastern

Africa as the climate became more arid and variable. This time period brackets important hominin milestones such as the first appearance of both Paranthropus sp. and Homo sp., and the development of the Lomekwian and Oldowan technologies. Singular spectral analysis was performed on high resolution magnetic susceptibility from this core as well as four other African paleoenvironmental records and a temperature record from NE Russia. This analysis indicates that the eastern Africa record shows aspects of both high-latitude forcing and orbitally driven cycles, and that a high percentage of variability in all African records can be explained associated with orbital timescales. The resulting empirical orthogonal functions from this analysis were used to create a reconstructed time-series above noise. The reconstructed time- series from eastern Africa, western Africa and southern Africa were used as the environmental input for an evolutionary algorithm. Comparisons of the modeled hominin responses to the 58 paleoanthropological record suggest that climate variability may have explanatory power for important hominin milestones during the Plio-Pleistocene transition, specifically the first known occurrence of our genus Homo at 2.8 Ma.

1. Introduction

The debate regarding whether and how climate influenced the evolution of early hominins has been ongoing in paleoanthropology for over a century and has generated many hypotheses. The savanna hypothesis states that hominins evolved to inhabit a more arid and grassland dominated environment (Dart, 1925). A unidirectional move from woodlands to grasslands could have created a step-wise turnover in faunal assemblages in which species go extinct and are replaced with those more adapted to the new environmental regime, referred to as the pulse-turnover hypothesis (Vrba, 1985). This unidirectional model has been updated to include increased variability between wet and dry phases while grasslands expanded, trends which are evident in both offshore paleoenvironmental records as well as faunal fossil assemblages (Bobe and Behrensmeyer, 2004; deMenocal, 2004; Trauth et al., 2009)

The variability selection hypothesis (VSH) suggests that hominin evolution is driven by consistent changes in the environment which create conditions requiring the multiple adaptive strategies seen in hominins, such as bipedalism, increased brain size and technological advancement (Potts, 1996, 1998). Similarly, the stable/instable hypothesis suggests that both instable and stable climates would have influenced hominin evolution (Grove, 2014; Maslin et al., 2015). The stable/instable hypothesis fits well with the accumulated plasticity hypothesis

(APH), which postulates that a species that experiences increased temporal variation within its environment is selected to have more adaptive strategies, and that an increase in adaptive strategies would allow for dispersal into more novel environments (Grove, 2011, 2014, 2015, 59

2017; Grove and Burke, 2015). The suite of biological and technological adaptive strategies is referred to as plasticity.

Previous work relating hominin evolution to possible environmental forcing has mainly been focused in the East African Rift Valley. A multitude of hominin fossil sites and the geology of the region has allowed for environmental interpretation to be garnered from both outcrops, and more recently, paleolake drill cores (Bobe and Behrensmeyer, 2004; deMenocal, 2004;

Deino et al., 2006; Campisano and Feibel, 2007; Kingston et al., 2007; Aronson et al., 2008;

Magill et al., 2013; Shultz and Maslin, 2013; DiMaggio et al., 2015; Trauth et al., 2015; Cohen et al., 2016; Campisano et al., 2017). However, the full terrestrial extent of early hominins on the

African continent is poorly known. Consideration has to be taken that any one locality may not be representative of the full geographical and environmental range of our hominin ancestors

(Behrensmeyer, 2006). Also, many of these records may have been influenced by local processes, which may swamp the signal of regional to global climatic shifts (Behrensmeyer,

2006). It is therefore beneficial to take a Pan-African approach when addressing climate in the context of human evolution.

A critical interval in human evolution occurred towards the end of the Pliocene and into the early Pleistocene. During this period, we see the last appearances of Australopithecus afarensis and Australopithecus bahrelghazali, and the first appearances of Paranthropus sp. and

Homo sp. as well as the Lomekwian and Oldowan tool assemblages (Johanson and White, 1979;

Brunet, 1997; Harmand et al., 2015; Villmoare et al., 2015; Campisano et al., 2017). The Plio-

Pleistocene has been defined in the global climate records by a decrease in global temperatures and the onset of Northern Hemisphere glaciation, with orbitally forced stadials and interstadials

(Lisiecki and Raymo, 2005). It is important to determine the effects of these global climate 60 changes on the regional environments of Africa in order to understand the environmental context of hominin evolution, migration and extinction events. Dust records from deep sea drill cores have been interpreted by some researchers to indicate that this transition is marked by increased aridity across the African continent (deMenocal, 1995). This transition also represents an increase in climate variability punctuated by wet-dry transition phases which are not synchronous across regions of different latitudes (Maslin and Trauth, 2009; Blome et al., 2012;

Shultz and Maslin, 2013; Maslin et al., 2014). Pollen records from deep-sea drill cores also show variability in the vegetation on the continent, with transitions occurring between woodlands, grasslands and arid/semi-arid flora being paced on orbital cycles (Dupont et al., 2011).

Most of the high-resolution environmental records from the Plio-Pleistocene have been developed using offshore drill cores. However, these may lack the spatial and temporal resolution needed to put climate in the context of human evolution. High resolution terrestrial records like those provided by paleolake drill cores are much more useful.

In 2013, a 227-meter drill core was taken from the Baringo Basin in Tugen Hills, Kenya as part of the Hominin Sites and Paleolake Drilling Project (HSPDP) (Cohen et al., 2016;

Campisano et al., 2017). The site was chosen because of the paleoanthropological significance of the basin as both hominin fossils and ancient tool assemblages spanning this time period have been found in the region (Deino and Hill, 2002; Hill, 2002; Kingston et al., 2007; Cohen et al.,

2016; Campisano et al., 2017). Furthermore, lacustrine deposits in the region that are mapped in detail and well-dated previously indicated precessional forcing as a pacing mechanism for lake- level fluctuations (Deino et al., 2006; Kingston et al., 2007; Wilson et al., 2014). Detailed analysis of the Tugen Hills core show that the lake went through multiple transgressions and regressions, with deep-lake phases and exposure surfaces being present, and that these lake level 61 fluctuations were most likely rather abrupt (Deino et al., 2019; Billingsley et al., in press; Yost et al., submitted; Westover et al., submitted). The Tugen Hills core ranges in age from approximately 3.29 – 2.57 Ma, encompassing the important Plio-Pleistocene transition (Deino et al., 2019). This offers a rare, high-resolution terrestrial record of paleoenvironment during this period. There is a major environmental shift at approximately 3.05 Ma, which may suggest a teleconnection to high latitude forcing during the lead up to Northern Hemisphere glaciation.

2. Methods

2.1 Records Used

Six paleoenvironmental records are used in this study (Figure 1). A high-resolution magnetic susceptibility record from the Tugen Hills, Kenya core is used as a terrestrial paleoenvironmental record for Eastern Africa. Magnetic susceptibility has been determined to be a good indicator of aridity through the core as low values correlate to deep-lake horizons and higher values indicate more terrestrial input and exposure surfaces (Billingsley et al., in press; Chen et al., 2013;

Tamuntuan et al., 2015). The age model of this core is untuned and based on 40Ar/39Ar, magnetostratigraphy, tephrostratigraphy, sequence startigraphy and Bayesian age modeling

(Deino et al., 2019). Magnetic susceptiblity measurements were taken every 0.5 cm with an average time-step of ~17yrs using a Geotek XYZ multisensor core scanner at the National

Lacustrine Core Facility (University of Minnesota). A highly continuous, continental pollen- derived temperature record obtained from Lake El’gygytgyn, NE Russia was used as an indicator of Northern Hemisphere climate changes (Brigham-Grette et al., 2013). The age model for the

Lake El’gygytgyn core was based on magnetostratigraphic tie points and the tuning of various proxy data with marine isotope stratigraphy and summer insolation at 67.5° N (Brigham-Grette et al., 2013). The average time-step for this record is ~4.67 ka. Dust flux records from offshore 62 drill sites ODP 659, ODP 721, and ODP 967 were used as aridity indices for western Africa, eastern Africa and the Mediterranean, respectfully (Tiedemann et al., 1994; deMenocal, 1995;

Larrasoana et al., 2003). The ODP 659 age model was based on magnetostratigraphy and tuning the δ18O record to marine isotope stages (Tiedemann et al., 1994). The average time-step for

ODP 659 is ~4.26 ka. The ODP 721 age model was derived from integrated oxygen isotopic, biotstratigraphic and magnetic polarity data, which was tuned to orbital insolation (deMenocal,

1995). The dust flux of ODP 721 was calculated every 0.5 cm, with an average time-step of 2.1 ka. The age model for ODP 967 is based on the tuning of the sapropel occurrences to orbital precession and then constrained by nannofossil datums and oxygen isotope values (Kroon et al.

1998; Larrasoana et al., 2003). ODP 967 dust flux was calculated at 1 cm intervals, with an average time-step of 494 yrs. A pollen derived precipitation record from ODP 1082 serves a paleoenvironmental record for southern Africa (Dupont, 2006). The age model for ODP 1082 is derived from stable isotopes of benthic and pelagic foraminifers and assignment to marine isotope stages as well as comparison tie points of other proxies between this core and the ODP

1083 (Dupont et al., 2005; Dupont et al., 2006). Palynological analyses between 2.57 Ma and 2.9

Ma were performed every 50 cm with an average time-step of 5.35 ka. Between 2.9 Ma and 3.27

Ma, analyses were performed between 50 cm and 3 m with an average time-step of 13.36 ka

(Dupont et al., 2006).

2.2 Singular Spectral Analysis

Original data was first normalized to z-scores and then linearly interpolated to 50-year time- steps. A trajectory matrix was created for the data using a target cycle length of 100 kyr in order to encompass all possible orbital parameters of the climate record, specifically Milankovitch cycles of precession (23 kyr), obliquity (41 kyr) and eccentricity (100 kyr). A covariance matrix 63 was then created for the trajectory matrix. Single Value Decomposition (SVD) was performed on the covariance matrix. Explained variance was calculated by dividing the singular value by the sum of all singular values. The 95 % significance of empirical orthogonal functions (EOFs) above noise was determined by using 1000 Monte Carlo simulations. Signals were created by applying the left eigenvector (U) output to the original trajectory matrix. Reconstructed components were then calculated by convolution of the left eigenvector outputs and the signal matrix. Reconstructed time-series were created by summing all of the reconstructed components represented by EOFs above noise. Dominant periodicities in the time series were determined by spectral analysis using the smoothed periodogram. Correlations between records were determined using the correlation method first outlined by Ebisuzaki (1997) in order to determine statistically significant correlation between autocorrelated time-series. All analyses were done using MATLAB® 2018a.

2.3 Hominin Response Model

Hominin responses to our climate records were modeled using an evolutionary algorithm originally published by Grove (2014). Details about the approach can be found in that publication. The model operates on a fixed population size of 1000, with individuals assigned at random two values meant to simulate a single chromosome (Grove, 2014, 2015). One value is meant to represent the optimum environment for the individual. This is a value chosen from the range of environmental input (i.e. an individual z-score from our reconstructed time-series). The other assigned value is a determination of the individual’s adaptability, which is a randomly assigned standard deviation around the optimum environmental value (Grove, 2014, 2015).

Larger standard deviation values imply that an individual has higher plasticity. For any given environmental input, an individual’s fitness is determined by the normal probability density 64 function and the individual’s assigned mean and standard deviation (Grove, 2014, 2015).

Survival in any environment is determined by the individual’s fitness. In any given time-step, the

500 weakest individuals “die” off and are replaced by the offspring of the 500 fittest individuals

(Grove, 2014, 2015). Mutation is simulated in the model by adding a random variate uniformly at each interval of time to the standard deviation and mean of each offspring (Grove, 2014,

2015). This process is repeated for each generational time-step. In order to apply this model to our records, the reconstructed time-series were first linearly interpolated to 20-year time-steps to create a generation time axis, with generation 1 being at the oldest date in any given record.

3. Results

3.1 Reconstructed Time-Series and Variance Explained

The reconstructed time-series track the main variability in each of our climate records well, with major transitions in the original data being recorded (Figure 2). The Lake El’gygytgyn temperature record has 6 EOFs determined above noise allowing for a reconstructed time-series representing 80.93 % of the variance in the record. 39.86 % of this variance is within a broadband eccentricity signal, 25.72 % represents heterodynes between eccentricity and obliquity, and 15.33 % is obliquity.

The dust record of ODP 659 has 9 EOFs above noise allowing for a reconstructed time-series representing 84.61 % of the variance in the record. 45.49 % of this variance can be explained by eccentricity, 5.46 % can be explained by heterodynes between eccentricity and obliquity, 12.8 % can be explained by obliquity, 3.70% can be explained by heterodynes between obliquity and precession and 17.15 % can be explained by precession. 65

The dust record from ODP 721/722 has 5 EOFs above noise allowing for a reconstructed time-series representing 76.69 % of the variance in the record. 62.69 % of this variance can be explained by eccentricity while 14.01% can be explained by precession.

The dust record of ODP 967 has 7 EOFs above noise allowing for a reconstructed time-series representing 86.82 % of the variance in the record. 37.85% of the variance can be explained by eccentric, 3.93 % explained by heterodynes between eccentricity and obliquity, 29.29 % by obliquity, and 15.75 % can be explained by precession.

The pollen record of precipitation from ODP 1082 8 EOFs above noise allowing for a reconstructed time-series representing 82.70 % of the variance in the record. 43.46 % of this variance can be explained by eccentricity, 22.05 % can be explained by heterodynes of eccentricity and obliquity, 13.77 % can be explained by obliquity and 3.42 % can be explained by precession.

The magnetic susceptibility record from Tugen Hills, Kenya has 9 EOFs above noise allowing for a reconstructed time-series representing 61.36 % of the variance in the record. 23.96

% of the variance can be explained by eccentricity, 11.97 % can be explained by heterodynes between eccentricity and obliquity, 16.43 % can be explained by obliquity and 13.82 % can be explained by precession.

3.2 Hominin Responses to Climate Records

The reconstructed time-series from ODP 659 (western Africa), ODP 1082 (southern Africa) and Tugen Hills Kenya (eastern Africa) were used as the environmental input for Grove’s (2014) evolutionary algorithm. These three records were used because they represent the environments closest to the known hominin sites located in southern Africa, the East African Rift Valley and

Chad. 66

In western Africa, there were eight periods when the standard deviation of the population increases: 3.2 Ma, 3.1 Ma, 3.06 Ma, 3.05 – 2.96 Ma, 2.95 – 2.92 Ma, 2.78 – 2.74 Ma, 2.73 – 2.69

Ma and 2.66 – 2.60 Ma. The population’s fitness increases during nine intervals throughout the record: 3.29 – 3.19 Ma, 3.17 – 3.14 Ma, 3.12 – 3.08 Ma, 3.05 – 3.04 Ma, 2.97 – 2.94 Ma, 2.92 –

2.77 Ma, 2.74 Ma, 2.68 – 2.66 Ma, and 2.61- 2.57 Ma. In southern Africa, there were two periods when the standard deviation of the population increases: 2.71 – 2.65 Ma and 2.61 Ma.

There are three intervals when the population fitness increases: 3.29 – 2.71 Ma, 2.65 – 2.61 Ma, and 2.6 – 2.57 Ma. There is a substantial peak in population fitness at 2.84 Ma. There are six periods when population standard deviation increases in eastern Africa: 3.05 – 2.99 Ma, 2.96 –

2.92 Ma, 2.82 – 2.80 Ma, 2.69 – 2.67 Ma, and 2.66 – 2.57 Ma. There are five periods when population fitness increases: 3.29 – 3.05 Ma, 2.99 – 2.96 Ma, 2.93 – 2.82 Ma, 2.81 – 2.69 and

2.67 – 2.65 Ma.

4. Discussion

4.1 Orbital and High-Latitude Controls on Plio-Pleistocene Climate of Eastern Africa

Previous studies of African climate during the Plio-Pleistocene showed increasing aridity as well as an increase in variability over time (deMenocal, 1995; Trauth et al., 2009). The controls on variability have been linked to orbital parameters such as eccentricity modified precession, teleconnections to Northern Hemisphere glaciation processes, and changes in ocean circulation and temperatures inferred to be related to the restriction of the Indonesian throughflow

(deMenocal, 2004; Campisano and Feibel, 2007; Donges et al., 2011). Although connections have been made to orbital and global climate forcing in multiple climate records throughout

Africa, the degree of influence global changes remains unclear (deMenocal, 1995; Leroy and

Dupont, 1997; deMenocal, 2004; Cohen et al., 2007; Kingston et al., 2007; Maslin and 67

Christensen, 2007; Scholz et al., 2007; Donges et al., 2011; Rose et al., 2016). Regional influences such as tectonic uplift of in eastern Africa and local convection processes may influence how these global changes are represented in different records around Africa

(Behrensmeyer, 2006; Maslin and Christensen, 2007). This matter is further complicated by the time scales and lags between terrestrial environmental changes and marine climate records

(Behrensmeyer, 2006; Trauth et al., 2009).

Lake El’gygytgyn reconstructed components suggests a long-term trend towards cooler temperature. The cooling at Lake El’gygytgyn should be expected over this period as the global climate cooled and Northern Hemisphere glaciation (NHG) initiated (Lisiecki and Raymo,

2005). This trend is supported by analysis of the Lake El’gygytgyn drill core which shows a lack of calcite being produced in the lake after 3.3 Ma in response to permafrost developing in the region (Wennrich et al., 2014). Pollen records indicate that beginning between 3.1 – 3.03 Ma there is a sharp decrease in coniferous tree pollen and an increase tundra-steppe vegetation

(Andreev et al., 2014). The ODP 967 dust flux record has a long-term trend towards increased aridity with an overlaid eccentricity signal. The trend of increased terrigenous dust in the

Mediterranean is expected because became more arid during the onset of NHG

(Larrasoana et al., 2003; Grant et al., 2017). Previous studies have also found long eccentricity in this record (Larrasoana et al., 2003). The Tugen Hills and ODP 721 display a 409 ka, a pattern seen in other Eastern Africa paleoclimate records (Trauth et al., 2009; Nie, 2018).

The main variance seen within all records can be explained using orbital parameters of eccentricity, obliquity and precession (Thomas et al., 2016; Clemens et al., 2018). Each record has a different amount of variance explained by these parameters. This may suggest that latitudinal position and regionally complex relationships between topography, convection 68 processes and sea surface temperatures may play a role in regional responses to different orbital forcing (Thomas et al., 2016). For instance, higher insolation in the Northern Hemisphere during precessional minima has been shown to influence the extent of the West African Monsoonal system (Rohling et al., 2002; Tierney et al., 2008, 2011; Junginger et al., 2014). Higher sea surface temperatures in both the North Atlantic and the have also been correlated to increased rainfall in Africa (Dupont, 2006; Weldeab et al., 2007; Blome et al., 2012; Ivory et al., 2016). The increase in higher frequency variability in in tropical Africa may be an indication of a more precession-controlled climate, bimodal rainfall and/or interhemispheric climate influences (Deino et al., 2006; Kingston et al., 2007; Verschuren et al., 2009; Clemens et al.,

2010; Singarayer and Burrough, 2015). Glacial expansion may limit the latitudinal extent of the tropical rain belt making these regions more susceptible to precession which may create mixed- frequency peaks (Trauth et al., 2007; Ivory et al., 2016).

Correlation analysis shows that the Tugen Hills, Kenya record is significantly (p=.03) and positively (r=0.33) correlated with the temperature record from Lake El’gygytgyn. This suggests teleconnections between the Northern Hemisphere and eastern Africa climate, which may have intensified during the lead up to and onset of NHG. NHG would have altered atmospheric and ocean interactions (Clark et al., 1999; Alley, 2002; Woodard et al., 2014). Ice sheets are topographic features which have been shown to change atmospheric circulation (Clark and Mix, 2000; Beghin et al., 2015). Expansive ice sheets in may have altered the placement of the Atlantic jet in the Northern Hemisphere (Löfverström et al., 2014). Antarctic ice sheet expansion and retraction have similar controls on atmospheric circulation in the

Southern Hemisphere (Hall and Visbeck, 2002; Ogura et al., 2004). Temperatures and ice volume in Antarctica have also been linked to variations in atmospheric pCO2, perhaps caused by 69 the slowing or quickening of Southern Ocean ventilation paced by glacial extent or by the reduced formation of North Atlantic Deep-Water (NADW) during glaciation (Driscoll and Haug,

1998; Ogura et al., 2004; Rohling et al., 2013; Brook and Buizert, 2018). The formation of

NADW has also been linked to increased upwelling off the coast of Africa and increased climate variability (Rohling et al., 2013). Albedo changes caused by fluctuations of extensive ice sheets also influences global atmospheric circulation by increasing or decreasing latitudinal temperature gradients (Ogura and Abe-Ouchi, 2001; Blackport and Kushner, 2017). Furthermore, ice sheet volume controls sea level which also influences local climate and coastal topography (Raymo,

2006; Cook et al., 2013; McKay, 2014; Wündsch et al., 2016).

It is clear that the onset of NHG had significant influences on the reorganization of the global climate patterns. Our correlation analysis shows a positive correlation between the temperature record of Lake El’gygytgyn and the aridity record from Tugen Hills, Kenya. This correlation suggests that while the Northern Hemisphere was warmer Eastern Africa became more arid. This supports previous studies suggesting high latitude influences on eastern Africa climate (deMenocal et al., 1993; deMenocal, 1995; Hopley et al., 2007; Tierney et al., 2008;

Caley et al., 2011, 2018).

4.2 Modeled Hominin Responses to Climate

Hominins during the Plio-Pleistocene experienced environments that were highly variable in both time and space. These different environments between regions in Africa would have allowed for larger populations to be separated by habitat fragmentation allowing for vicariance, genetic drift and allopatric speciation (Kingston, 2007; Reynolds, 2007; Blome et al., 2012;

Maslin et al., 2015). Each environment would have provided differences in both abiotic and biotic forcing (Grove, 2011; Levin, 2015; Maslin et al., 2015; Potts and Faith, 2015). 70

Abiotic and biotic forcing have been invoked to explain speciation and extinction events

(Benton, 2010; Grove, 2012; Voje et al., 2015; Žliobaitė et al., 2017). In the absence of extreme climate or other environmental changes (abiotic forcing), speciation and extinction would be controlled by biotic forcing mechanisms such as competition or predation (i.e. the Red Queen

Hypothesis) (Van Valen, 1973). Species would be consistently evolving to adapt to biotic interactions they experience. This might lead to specialization in the context of a uniform environmental regime. This hypothesis has been shown to prove explanatory in fossil records as well as ecological experiments, in which environmental factors are stable or inconsequential

(Kerfoot and Weider, 2004; Liow et al., 2011; Gibson and Fuentes, 2015; Voje et al., 2015).

However, statistical analysis of the Red Queen scenario has determined that this evolutionary process would be difficult to distinguish from a stationary model where a species exhibit no evolutionary changes except in the cases of extreme physical environmental changes (Stenseth and Maynard Smith, 1984). In other words, biotic forcing may be a driving cause of evolution primarily in the absence of major perturbation to the environment (Barnosky, 2001; Benton,

2009; Venditti et al., 2010). This scenario also has support in the fossil record where speciation and extinction can be correlated with extreme environmental overturn (Bobe and Eck, 2001;

Bobe et al., 2002; Hernández Fernández and Vrba, 2006). Abiotic and biotic forces can both be drivers of evolution; however, these two processes may work on separate timescales and influence populations differently (Barnosky, 2001; Grove, 2012; Žliobaitė et al., 2017). In the presence of consistent abiotic forcing, a species may evolve primarily in response to environmental variability, as put forth by both the VSH and APH hypotheses. Population in these circumstances would be selected to favor generalists, whereas specialists would be favored 71 during periods of stable environments. Specialists may be ill-suited for extreme environmental variability and may suffer extinction in such scenarios (Grove, 2012).

In order to predict the responses of hominins to the environmental change seen in Africa, we have applied an evolutionary algorithm to the reconstructed time-series from eastern, western and southern Africa (Figure 3). This algorithm has been used to demonstrate species responses to environmental variability in both synthetic and empirical climate data (Grove, 2011, 2014,

2015). Using synthetic data the model indicates that directional shifts in the environment will elicit an increased standard deviation in the population (plasticity), whereas an invariant environment causes the mean of the population (fitness) to increase (Grove, 2011, 2014, 2015).

An increase in the fitness value would correspond to an increase in specialization during a period of environmental stability. When applied to an environmental record from , Ethiopia, the algorithm showed that APH could be used to explain known early Homo sapiens dispersals from eastern Africa into the Levant (Grove, 2015).

The eastern African model results show that from 3.29 – 3.05 Ma, a given species’ population’s fitness would have remained relatively high. The climate during this period was relatively stable, tending to be more arid than the mean of the whole time-series. This period coincided with a period of peak species richness (less extinction and more origination) occurring in eastern Africa (Werdelin and Lewis, 2005; Lewis and Werdelin, 2007; Macho, 2014). This interval also covers the last known appearance of (LAD) Australopithecus bahrelghazali (Chad), which disappears from the fossil record approximately 3 Ma (Grove, 2012; Lee-Thorp et al.,

2012). An environmental forcing of the disappearance of Australopithecus bahrelghazali may have been related to the extremely narrow diet inferred for this species (Macho, 2015). 72

At 3.05 Ma, the Tugen Hills record suggests there was rapid change in the environmental conditions in eastern Africa, which results in a model prediction of increased plasticity for the population. Although this record is from Tugen Hills, Kenya, high amplitude oscillations in climate have also been recorded in Hadar, Ethiopia between 3.15 – 3.05 Ma (Campisano and

Feibel, 2007). This period coincides with a period of increasing body size for Australopithecus afarensis, possibly an adaptation to a changing environment (Lockwood et al., 2000; Drapeau et al., 2005; Campisano and Feibel, 2007).

Between 2.9 and 2.67 Ma, the Tugen Hills record suggests a relatively stable climate in eastern Africa, and the model population fitness, or increased specialization, output reflects this.

This time period also coincides with the LAD of Australopithecus afarensis. Interpretations of the environment of Australopithecus afarensis indicate that this species experienced many rapid environmental shifts over time (Bonnefille et al., 2004). The disappearance of a generalist during a period of climate stability may suggest that the new environment was outside the range of possible adaptations or Australopithecus afarensis may have been outcompeted by a more specialized species. However, interpretation of the precise timing of this extinction is hampered by the presence of a disconformity spanning 2.9 – 2.7 Ma at Hadar, where many of these fossils have been recovered. (Quade et al., 2004).

2.9 – 2.67 Ma was a period of relatively stable climate in eastern Africa, except for a rapid transition from a relatively arid climate to one much wetter at 2.8 Ma. The modeled population experienced an increase in plasticity. The timing of this peak coincides with a major faunal turnover in eastern Africa among fossil mammals observed at Omo, Koobi Fora, Baringo, Hadar and the Middle Awash (Bobe and Behrensmeyer, 2004; Bibi, 2011; Reed and Geraads, 2012; 73

DiMaggio et al., 2015). Most importantly, this peak in standard deviation correlates to the earliest known fossil of our genus, Homo (Villmoare et al., 2015).

Eastern Africa returns to a relatively stable and slightly more arid phase before becoming much wetter and more variable between 2.7 – 2.57 Ma. For this time period, the model predicts an increased plasticity of the population, similar to the earlier predictions by Grove (2012) using global climate records. This period coincides with high variability and deep lake phases seen in other records throughout Eastern Africa (deMenocal, 1995; Deino et al., 2006; Kingston et al.,

2007; Trauth et al., 2007, 2009; Donges et al., 2011; Shultz and Maslin, 2013; Maslin et al.,

2014; Maslin et al., 2015). It has been suggested that increased environmental variability during this period would have created a mosaic of environments, which in turn would have fragmented communities leading to dispersals and speciation (Kingston, 2007). Other noteworthy events in hominin evolutionary history during this interval include an increase in cephalization among hominin species, the FAD of Paranthropus aethiopicus and the first appearance of Oldowan tool technologies (Grove, 2012; Shultz et al., 2012; Braun et al., 2019). There is an amplified speciation pulse and possible turnover among fossil mammals between 2.7 – 2.5 Ma in eastern

Africa as well (Reed, 1997; Behrensmeyer et al., 1997; Hernández Fernández and Vrba, 2006;

McKee, 2017; Reed and Geraads, 2012; Rannikko et al., 2017; Maxwell et al., 2018).

For southern Africa, the model results predict very little change in fitness or plasticity in reaction to the climate signal. The lack of any predicted relative response in the population during this time coincides with other observations that the expansions and contractions of woodland environments may not have been protracted enough to elicit a response in mammals on the landscape (Thackeray and Reynolds, 1997). This may also explain why an extreme faunal turnover has not been observed at ~2.8 Ma in southern Africa as is seen in eastern Africa (Maslin 74 and Christensen, 2007; Faith and Behrensmeyer, 2013). Similarly, it has been noted that, as a group, the pattern of mammalian species evolution differed between southern and eastern Africa, with changes in mammalian body sizes being much more extreme in eastern Africa ( Reynolds,

2007). This may support the claim that southern Africa may have served as a climatically-stable refugium during the Plio-Pleistocene transition, while climate instabilities occurred elsewhere in

Africa (Reynolds, 2007; Maslin et al., 2012). The persistence of Australopithecus africanus may also be a reflection of this climate stability.

At 2.7 Ma, however, there are transitions in southern African climate between arid and wet environments. This yields three peaks of modeled plasticity between 2.7 – 2.6 Ma. The climate shift represents a period of much higher climate variability in the region as well as a lowering of sea level during the intensification of NHG, which may have controlled coastal biomes (Maslin et al., 2012; Stollhofen et al., 2014). This is also the period of time outlined as a possible pulse- turnover event in African bovines (Vrba, 1993). However, it is important to note that this record of increasing aridity off the coast of Namibia differs from the climate reconstruction from ODP

1085, where southern African climate has been interpreted as getting much wetter during the same time interval (Maslin et al., 2012). This may be the result of the two drill cores’ different locations in relation to both the Benguela current and the northern extension of the Southern

Hemisphere westerlies, which would control the amount of precipitation on the landscape. The differences in the core records suggest Australopithecus africanus experienced regional spatial variability caused by local processes and topography which are difficult to determine from a single precipitation record. This variability in spatial environment may be the reason the species did not go extinct between 2.7 – 2.6 Ma as it was well adapted to multiple environments

(Reynolds et al., 2011). 75

The amplitude of environmental change seen in southern Africa between 2.7 – 2.6 Ma is similar in range to another amplitude change between 3.05 – 3.10 Ma. However, the 2.7 – 2.6

Ma generates a response in the population plasticity whereas the earlier one does not. The main reason for this relates to the rate of change in the environment. The 3.05 Ma change occurred at a slow enough pace that the model population was able to adapt to it without requiring any increase in plasticity (Grove, 2014). This implies that the duration or rate of environmental change is just as important, if not more so, than the extremes of amplitude change when considering climate variability’s impact on species.

5. Conclusions

Our study shows that most of the variance seen within the five climate records from Africa and one from NE Russia can be explained by orbital parameters. The similarities between the records as well as the significant correlation between Lake El’gygytgyn and eastern Africa suggest that the onset of Northern Hemisphere cooling may have had far reaching teleconnections, and that this change predates the earliest Pleistocene onset of extensive northern hemisphere glaciation. Pan-African climate was highly variable 3.29 – 2.57 Ma, and consistent with prior investigations covering much shorter time periods, that this variability was not synchronous across the continent.

The variability seen within the climate records is reflected in the modeled hominin responses of population fitness and plasticity created by the evolutionary algorithm. Periods of extreme environmental shifts elicit a response of increased plasticity, while relatively stable periods select for a more specialized population. The populations of eastern, west and southern Africa would have been subjected to different abiotic forcing events, and the differences in the populations responses between each region indicates this. 76

Eastern Africa experienced five periods when the environment was variable enough to create strong selection for plasticity in our model. Periods of high environmental variability and increased modeled plasticity between 3.05 – 2.9 Ma correlate with increasing body size of

Australopithecus afarensis in Ethiopia and the LAD of Australopithecus bahrelghazali in Chad.

The increasing body size of Australopithecus afarensis may reflect an increase in plasticity as it evolved in response to an ever more variable environment. Australopithecus bahrelghazali, with its more specialized dietary limitations, may have been victim of rapid environmental changes.

A switch from high variability into more stable environment coincides with the LAD of

Australopithecus afarensis, which may have been ill-suited for the new environment or outcompeted by a yet unknown species more adapted to the new stable environment. A drastic change in environment at 2.8 Ma occurred coincident with a massive faunal turnover in eastern

Africa as the environment transitioned from relatively arid to wetter conditions. This peak in plasticity also correlates with the first appearance of our genus Homo. Between 2.7 -2.57 Ma, eastern Africa experienced increased environmental variability, which is associated with an increased plasticity of the modeled population. During this time, Paranthropus aethiopicus and

Oldowan technology appear in the record.

The records we examined suggest that southern African climate did not vary as much as eastern or western Africa did over the study interval. Although there were multiple fluctuations in the climate record, only one of these occurred on the temporal scale rapidly enough and with sufficient magnitude to elicit increased plasticity in the modeled population. This one period, 2.7

– 2.6 Ma, coincides with a pulse turnover event in African bovids (Vrba, 1993). The otherwise more stable climatic conditions and consistent model response of the population’s fitness lends 77 credence to the idea that southern Africa could have served as a refugium for many species during the Plio-Pleistocene while faunal turnover was occurring elsewhere on the continent.

Our analysis has indicated that climate variability and the abiotic forcing it creates hold important explanatory power for understanding the speciation and extinction of hominins and other mammals in Africa. The increased plasticity in a modeled population in response to real environmental variability aligning with known evolutionary events in the hominin and faunal fossil record lends support to the fundamental ideas outlined by the variability selection and the accumulated plasticity hypotheses.

Acknowledgements

This research was funded by grants from the National Science Foundation (EAR1123000,

EAR1338553, BCS124185) and the International Continental Drilling Program. We thank the other members of HSPDP and DOSECC Exploration Services field and initial core description teams (Kristina Brady, Jessamy Doman, Sarah Ivory, Beau Marshall, Dennis Njagi, Anders

Noren, Ryan O’Grady, Steve Rucina, Rich Szenmiklosi, Gladys Tuitoek, Chad Yost, for assistance with core collection). MSCL core logging was done at the National Lakes Core

Repository (LacCore). Thank you to Jessie Pearl for her insight and assistance on the time series analysis. This is publication xx of the Hominin Sites and Paleolakes Drilling Project.

References

Alley, R.B., 2000. Ice-core evidence of abrupt climate changes. Proceedings of the National

Academy of Sciences. 97, 1331–1334.

Andreev, A.A., Tarasov, P.E., Wennrich, V., Raschke, E., Herzschuh, U., Nowaczyk, N.R.,

Brigham-Grette, J., Melles, M., 2014. Late Pliocene and Early Pleistocene vegetation

history of northEasternern Russian inferred from the Lake El’gygytgyn pollen 78

record. Climate of the Past. 10, 1017–1039.

Aronson, J.L., Hailemichael, M., Savin, S.M., 2008. Hominid environments at Hadar from

paleosol studies in a framework of Ethiopian climate change. Journal of Human Evolution.

55, 532–550.

Barnosky, A., 2001. Distinguishing the Effects of the Red Queen and Court Jester on Miocene

Mammal Evolution in the Northern Rocky Mountains. Journal of Vertebrate Paleontology.

21, 172–185.

Beghin, P., Charbit, S., Dumas, C., Kageyama, M., Ritz, C., 2015. How might the North

American ice sheet influence the northwestern Eurasian climate? Climate of the Past. 11,

1467–1490.

Behrensmeyer, A.K., Todd, N.E., Potts, R., McBrinn, G.E., 1997. Late Pliocene faunal trunover

in the Turkana Basin, Kenya and Ethiopa. Science. 278, 1589–1594.

Behrensmeyer, A.K., 2006. Climate change and human evolution. Science. 11, 1467–1490.

Benton, M.J., 2009. The Red Queen and the Court Jester: Species diversity and the role of biotic

and abiotic factors through time. Science. 323, 728–732.

Benton, M.J., 2010. New take on the Red Queen. Nature. 463, 306–307.

Bibi, F. 2011. Tragelaphus nakuae: evolutionary change, biochronology, and turnover in the

African Plio-Pleistocene. Zoological Journal of the Linnean Society. 162, 699–711.

Billingsley, A.L., Reinthal, P., Dettman, D.L., Kingston, J.D., Deino, A.L., Ortiz, K., Mohler, B.,

Cohen, A.S. 2019. δ13C records from fish fossils as paleo-indicators of ecosystem response

to lake levels in the Plio-Pleistocene lakes of Tugen Hills, Kenya. Palaeogeography,

Palaeoclimatology, Palaeoecology. In press.

Blackport, R., Kushner, P.J., 2017. Isolating the atmospheric circulation response to arctic sea 79

ice loss in the coupled climate system. Journal of Climate. 30, 2163–2185.

Blome, M.W., Cohen, A.S., Tryon, C.A., Brooks, A.S., Russell, J., 2012. The environmental

context for the origins of modern human diversity: A synthesis of regional variability in

African climate 150,000-30,000 years ago. Journal of Human Evolution. 62, 563–592.

Bobe, R., Behrensmeyer, A.K., 2004. The expansion of grassland ecosystems in Africa in

relation to mammalian evolution and the origin of the genus Homo. Palaeogeography,

Palaeoclimatology, Palaeoecology. 207, 399–420.

Bobe, R., Behrensmeyer, A.K., Chapman, R.E., 2002. Faunal change, environmental variability

and late Pliocene hominin evolution. Journal of Human Evolution. 42, 475–497.

Bobe, R., Eck, G.G., 2001. Responses of African bovids to Pliocene climatic change.

Paleobiology. 27, 1–47.

Bonnefille, R., Potts, R., Chalie, F., Jolly, D., Peyron, O., 2004. High-resolution vegetation and

climate change associated with Pliocene Australopithecus afarensis. Proceedings of the

National Academy of Sciences. 101, 12125–12129.

Braun, D.R., Aldeias, V., Archer, W., Arrowsmith, J.R., Baraki, N., Campisano, C.J., Deino,

A.L., DiMaggio, E.N., Dupont-Nivet, G., Engda, B., Feary, D.A., Garello, D.I., Kerfelew,

Z., McPherron, S.P., Patterson, D.B., Reeves, J.S., Thompson, J.C., Reed, K.E., 2019.

Earliest known Oldowan artifacts at 2.58 Ma from Ledi-Geraru, Ethiopia, highlight early

technological diversity. Proceedings of the National Academy of Sciences. 116, 11712–

11717.

Brigham-Grette, J., Melles, M., Minyuk, P., Andreev, A., Tarasov, P., DeConto, R., Koenig, S.,

Nowaczyk, N., Wennrich, V., Rosen, P., Haltia, E., Cook, T., Gebhardt, C., Meyer-Jacob,

C., Snyder, J., Herzschuh, U., 2013. Pliocene Warmth, Polar Amplification, and Stepped 80

Pleistocene Cooling Recorded in NE Arctic Russia. Science. 340, 1421–1427.

Brook, E.J., Buizert, C., 2018. Antarctic and global climate history viewed from ice cores.

Nature. 558, 200–208.

Brunet, M. 1997. Origine des hominides: east side sotry…west side story. Geobios. 30. 79–83.

Caley, T., Extier, T., Collins, J.A., Schefuß, E., Dupont, L., Malaizé, B., Rossignol, L., Souron,

A., McClymont, E.L., Jimenez-Espejo, F.J., García-Comas, C., Eynaud, F., Martinez, P.,

Roche, D.M., Jorry, S.J., Charlier, K., Wary, M., Gourves, P.Y., Billy, I., Giraudeau, J.,

2018. A two-million-year-long hydroclimatic context for hominin evolution in

southEasternern Africa. Nature. 560, 76–79.

Caley, T., Malaizé, B., Revel, M., Ducassou, E., Wainer, K., Ibrahim, M., Shoeaib, D., Migeon,

S., Marieu, V., 2011. Orbital timing of the Indian, Eastern Asian and African boreal

monsoons and the concept of a “global monsoon.” Quaternary Science Reviews. 30, 3705–

3715.

Campisano, C.J., Cohen, A.S., Arrowsmith, J.R., Asrat, A., Behresmeyer, A.K., Brown, E.T.,

Deino, A.L., Deocampo, D., Feibel, C.S., Kingston, J., Lamb, H.F., Lowenstein, T.K.,

Noren, A.J., Olago, D.O., Owen, R.B., Pelletier, J.D., Potts, R., Reed, K.E., Renaut, R.W.,

Russell, J.M., Russell, J.L., Shäbitz, F., Stone, J.R., Trauth, M.H., Wynn, J.G., 2017. New

high-resolution paleoclimate records from the Eastern African Rift System and their

implications for understanding the environmental context of hominin evolution.

PaleoAnthropology. 1–43.

Campisano, C.J., Feibel, C.S., 2007. Connecting local environmental sequences to global climate

patterns: evidence from the hominin-bearing Hadar Formation, Ethiopia. Journal of Human

Evolution. 53, 515–527. 81

Chen, F., Liu, J., Xu, Q., Li, Y., Chen, J., Wei, H., Liu, Q., Wang, Z., Cao, X., Zhang, S., 2013.

Environmental magnetic studies of sediment cores from Gonghai Lake: Implications for

monsoon evolution in North China during the late glacial and Holocene. Journal of

Paleolimnology. 49, 447–464.

Clark, P.U., Alley, R.B., Pollard, D., 1999. Northern hemisphere ice-sheet influences on global

climate change. Science. 286, 1104–1111.

Clark, P.U., Mix, A.C., 2000. Ice sheets by volume. Nature.406, 689–690.

Clemens, S.C., Prell, W.L., Sun, Y., 2010. Orbital-scale timing and mechanisms driving Late

Pleistocene Indo-Asian summer monsoons: Reinterpreting cave speleothem δ18O.

Paleoceanography. 25, doi:10.1029/2010PA001926.

Clemens, S.C., Holbourn, A., Kubota, Y., Lee, K.E., Liu, Z., Chen, G., Nelson, A., Fox-Kemper

B., 2018. Precession-band variance missing form East Asian monsoon runoff. Nature

Communications. 9:3364. doi:10.1038/s41467-018-05814-0.

Cohen, A.S., Campisano, C., Arrowsmith, R., Asrat, A., Behrensmeyer, A.K., Deino, A., Feibel,

C., Hill, A., Johnson, R., Kingston, J., Lamb, H., Lowenstein, T., Noren, A., Olago, D.,

Owen, R.B., Potts, R., Reed, K., Renaut, R., Schäbitz, F., Tiercelin, J.J., Trauth, M.H.,

Wynn, J., Ivory, S., Brady, K., O’Grady, R., Rodysill, J., Githiri, J., Russell, J., Foerster, V.,

Dommain, R., Rucina, S., Deocampo, D., Russell, J., Billingsley, A., Beck, C., Dorenbeck,

G., Dullo, L., Feary, D., Garello, D., Gromig, R., Johnson, T., Junginger, A., Karanja, M.,

Kimburi, E., Mbuthia, A., McCartney, T., McNulty, E., Muiruri, V., Nambiro, E., Negash,

E.W., Njagi, D., Wilson, J.N., Rabideaux, N., Raub, T., Sier, M.J., Smith, P., Urban, J.,

Warren, M., Yadeta, M., Yost, C., Zinaye, B., 2016. The Hominin Sites and Paleolakes

Drilling Project: Inferring the environmental context of human evolution from Easternern 82

African rift lake deposits. Scientific Drilling. 21, 1–16.

Cohen, A. S., Stone, J.R., Beuning, K.R.M., Park, L.E., Reinthal, P.N., Dettman, D., Scholz,

C.A., Johnson, T.C., King, J.W., Talbot, M.R., Brown, E.T., Ivory, S.J., 2007. Ecological

consequences of early Late Pleistocene megadroughts in tropical Africa. Proceedings of the

National Academy of Sciences .104, 16422–16427.

Cook, C.P., Van De Flierdt, T., Williams, T., Hemming, S.R., Iwai, M., Kobayashi, M., Jimenez-

Espejo, F.J., Escutia, C., González, J.J., Khim, B.K., McKay, R.M., Passchier, S., Bohaty,

S.M., Riesselman, C.R., Tauxe, L., Sugisaki, S., Galindo, A.L., Patterson, M.O., Sangiorgi,

F., Pierce, E.L., Brinkhuis, H., Klaus, A., Fehr, A., Bendle, J.A.P., Bijl, P.K., Carr, S.A.,

Dunbar, R.B., Flores, J.A., Hayden, T.G., Katsuki, K., Kong, G.S., Nakai, M., Olney, M.P.,

Pekar, S.F., Pross, J., Röhl, U., Sakai, T., Shrivastava, P.K., Stickley, C.E., Tuo, S., Welsh,

K., Yamane, M., 2013. Dynamic behaviour of the Eastern Antarctic ice sheet during

Pliocene warmth. Nature Geoscience. 6, 765–769.

Dart, R.A., 1925. Australopithecus africanus: The man-ape of South Africa. Nature. 115, 195–

199.

Deino, A.L., Hill, A., 2002. 40Ar/39Ar dating of Chemeron Formation strata encompassing the

site of hominid KNM-BC 1, Tugen Hills, Kenya. Journal of Human Evolution. 42, 141–

151.

Deino, A.L., Kingston, J.D., Glen, J.M., Edgar, R.K., Hill, A., 2006. Precessional forcing of

lacustrine sedimentation in the late Cenozoic Chemeron Basin, Central Kenya Rift, and

calibration of the Gauss/Matuyama boundary. Earth and Planetary Science Letters. 247, 41–

60.

Deino, A.L., Sier, M., Garello, D., Keller, B., Kingston, J.D., Scott, J., Dupont-Nivet, G., Cohen, 83

A.S. 2019. Chronostratigraphy of the Baringo-Tugen-Barsemoi (HSPDP-BTB-1A Core) -

40Ar/39Ar dating, manetostratigraphy, tephrostratigraphy, sequence startigraphy and

Bayesian age modeling. Palaeogeogr. Palaeoclimatol. Palaeoeco. In press. deMenocal, P.B., 1995. Plo-Pleistocene African Climate. Science. 270, 53–59. deMenocal, P.B., 2004. African climate change and faunal evolution during the Pliocene-

Pleistocene. Earth and Planetary Science Letters. 220, 3–24. deMenocal, P.B., Ruddiman, W.F., Pokras, E.M., 1993. Influences of High‐ and Low‐Latitude

Processes on African Terrestrial Climate: Pleistocene Eolian Records from Equatorial

Atlantic Ocean Drilling Program Site 663. Paleoceanography. 8, 209–242.

DiMaggio, E.N., Campisano, C.J., Rowan, J., Dupont-Nivet, G., Deino, A.L., Bibi, F., Lewis,

M.E., Souron, A., Garello, D., Werdelin, L., Reed, K.E., Arrowsmith, J.R., 2015. Late

Pliocene fossiliferous sedimentary record and the environmental context of early Homo

from Afar, Ethiopia. Science. 347, 1355–1359.

Donges, J.F., Donner, R. V., Trauth, M.H., Marwan, N., Schellnhuber, H.-J., Kurths, J., 2011.

Nonlinear detection of paleoclimate-variability transitions possibly related to human

evolution. Proceedings of the National Academy of Sciences. 108, 20422–20427.

Drapeau, M.S.M., Ward, C.V., Kimbel, W.H., Johanson, D.C., Rak, Y. 2005. Associated cranial

and forelimb remains attributed to Australopithecus afarensis from Hadar, Ethiopia. Journal

of Human Evolution. 48, 593–642.

Driscoll, N.W., Haug, G.H., 1998. A short circuit in thermohaline circulation: A cause for

Northern Hemisphere glaciation? Science. 282, 436–438.

Dupont, L.M., Donner, B., Vidal, L., Perez, E.M., Wefer, G. 2005. Linking desert evolution and

coastal upwelling: Pliocene climate change in Namibia. Geology. 33, 461–464; 84

doi:10.1130/G21401.1.

Dupont, L.M., 2006. Late Pliocene vegetation and climate in Namibia (southern Africa) derived

from palynology of ODP Site 1082. Geochemistry, Geophysics, Geosystems. 7,

doi:10.1029/2005GC001208.

Dupont, L.M., Linder, H.P., Rommerskirchen, F., Schefuß, E., 2011. Climate-driven rampant

speciation of the Cape flora. Journal of Biogeography. 38, 1059–1068.

Ebisuzaki, W. 1997. A method to estimate the statistical significance of a correlation when the

data are serially correlated. Journal of Climate. 10, 2147-2153.

Faith, J.T., Behrensmeyer, A.K., 2013. Climate change and faunal turnover: testing the

mechanics of the turnover-pulse hypothesis with South African fossil data. Paleobiology.

39, 609–627.

Gibson, A.K., Fuentes, J.A., 2015. A phylogenetic test of the Red Queen Hypothesis:

Outcrossing and parasitism in the Nematode phylum. Evolution. 69, 530–540.

Grant, K.M., Rohling, E.J., Westerhold, T., Zabel, M., Heslop, D., Konijnendijk, T., Lourens, L.,

2017. A 3 million year index for North African humidity/aridity and the implication of

potential pan-African Humid periods. Quaternary Science Reviews. 171, 100–118.

Grove, M., 2011. Speciation, diversity, and Mode 1 technologies: The impact of variability

selection. Journal of Human Evolution. 61, 306–319.

Grove, M., 2012. Orbital dynamics, environmental heterogeneity, and the evolution of the human

brain. Intelligence. 40, 404–418.

Grove, M., 2014. Evolution and dispersal under climatic instability: A simple evolutionary

algorithm. Adaptive Behavior. 22, 235–254.

Grove, M., 2015. Palaeoclimates , plasticity , and the early dispersal of Homo sapiens. 85

Quaternary International. 369, 17–37.

Grove, M., 2017. Environmental complexity, life history, and encephalisation in human

evolution. Biology and Philosophy. 32, 395–420.

Grove, M., Burke, A., 2015. Environmental variability and hominin dispersal. Journal of Human

Evolution. 87, 1–4.

Hall, A., Visbeck, M., 2002. Synchronous variability in the Southern Hemisphere atmosphere,

sea ice, and ocean resulting from the annular mode. Journal of Climate. 15, 3043–3057.

Harmand. S., Lewis, J.E., Feibel, C.S., Lepre, C.J., Prat, S., Lenoble, A., Boes, X., Quinn, R.L.,

Brenet, M., Arroyo, A., Taylor, N., Clement, S., Dave, G., Brugal, J-P., Leakey, L.,

Mortlock, R.A., Wright, J.D., Lokorodi, S., Kirwa, C., Kent, D.V., and Roche, H. 2015. 3.3-

million year old stone tools from Lomekwi 3, West Turkana, Kenya. Nature. 521, 310–315.

Hernández Fernández, M., Vrba, E.S., 2006. Plio-Pleistocene climatic change in the Turkana

Basin (Eastern Africa): Evidence from large mammal faunas. Journal of Human Evolution.

50, 595–626.

Hill, A., 2002. Paleoanthropological research in the Tugen Hills, Kenya. Introduction. Journal of

human evolution. 42, 1–10.

Hopley, P.J., Weedon, G.P., Marshall, J.D., Herries, A.I.R., Latham, A.G., Kuykendall, K.L.,

2007. High- and low-latitude orbital forcing of early hominin habitats in South Africa. Earth

and Planetary Science Letters.256, 419–432.

Ivory, S.J., Early, R., Sax, D.F., Russell, J., 2016. Niche expansion and temperature sensitivity of

tropical African montane forests. Global Ecology and Biogeography. 25, 693–703.

Johanson, D.C. and White, T.D. A systematic assesment of early African hominids. Science.

203, 321-330. 86

Junginger, A., Roller, S., Olaka, L.A., Trauth, M.H., 2014. The effects of solar irradiation

changes on the migration of the Congo Air Boundary and water levels of paleo-Lake

Suguta, Northern Kenya Rift, during the (15-5ka BP).

Palaeogeography, Palaeoclimatology, Palaeoecology. 396, 1–16.

Kerfoot, W.C., Weider, L.J., 2004. Experimental paleoecology (resurrection ecology): Chasing

Van Valen’s Red Queen hypothesis. Limnology and Oceanography. 49, 1300–1316.

Kingston, J.D., 2007. Shifting adaptive landscapes: Progress and challenges in reconstructing

early hominid environments. Yearbook of Physical Anthropology. 134, 20–58.

Kingston, J.D., Deino, A.L., Edgar, R.K., Hill, A., 2007. Astronomically forced climate change

in the Kenyan Rift Valley 2.7-2.55 Ma: implications for the evolution of early hominin

ecosystems. Journal of Human Evolution. 53, 487–503.

Kroon, D., Alexander, I., Little, M., Lourens, L.J.Matthewson, A., Robertson, A.H.F., Sakamoto,

T. 1998. Oxygen isotope and saporopel stratigraphy in the eastern Mediterranean during the

last 3.2 million years. In: Robertson, A.H.F., Emeis, K.-C., Camerlenghi, A., eds.

Proceedings of the Ocean Drilling Program, Scientific Results. 160, 181–189.

Larrasoana, J.C., Roberts, A.P., Rohling, E.J., Winklhofer, M., Wehausen, R., 2003. Three

million years of monsoon variability over the northern . Climate Dynamics. 21, 689–

698.

Lee-Thorp, J., Likius, A., Mackaye, H.T., Vignaud, P., Sponheimer, M., Brunet, M., 2012.

Isotopic evidence for an early shift to C4 resources by Pliocene hominins in Chad.

Proceedings of the National Academy of Sciences. 109, 20369–20372.

Leroy, S.A.G., Dupont, L.M., 1997. Marine palynology of the ODP site 658 (N-W Africa) and

its contribution to the stratigraphy of Late Pliocene. Geobios. 30, 351–359. 87

Levin, N.E., 2015. Environment and Climate of Early Human Evolution. Annual Review of

Earth and Planetary Sciences. 43, 405–429.

Lewis, M.E., Werdelin, L., 2007. Patterns of change in the Plio-Pleistocene carnivorans of

Easternern Africa: Implications for hominin evolution. In: Vertebrate Paleobiology and

Paleoanthropology. 77–105.

Liow, L.H., Van Valen, L., Stenseth, N.C., 2011. Red Queen: From populations to taxa and

communities. Trends in Ecology and Evolution. 26, 349–358.

Lisiecki, L.E., Raymo, M.E., 2005. A Pliocene-Pleistocene stack of 57 globally distributed

benthic δ 18O records. Paleoceanography. 20, 1–17.

Lockwood, C.A., Kimbel, W.H., Johanson, D.C. 2000. Temporal trends and metric variation in

the mandibles and dentition of Australopithecus afarensis. Journal of Human Evolution. 39,

23–55.

Löfverström, M., Caballero, R., Nilsson, J., Kleman, J., 2014. Evolution of the large-scale

atmospheric circulation in response to changing ice sheets over the last glacial cycle.

Climate of the Past. 10, 1441–1451.

Macho, G.A., 2014. An ecological and behavioural approach to hominin evolution during the

Pliocene. Quaternary Science Reviews. 96, 23–31.

Macho, G.A., 2015. Pliocene hominin biogeography and ecology. Journal of Human Evolution.

87, 78–86.

Magill, C.R., Ashley, G.M., Freeman, K.H., 2013. Water, plants, and early human habitats in

Easternern Africa. Proceedings of the National Academy of Sciences. 110, 1175–1180.

Maslin, M.A., Brierley, C.M., Milner, A.M., Shultz, S., Trauth, M.H., Wilson, K.E., 2014.

Eastern african climate pulses and early human evolution. Quaternary Science Reviews. 88

101, 1–17.

Maslin, M.A., Christensen, B., 2007. Tectonics, orbital forcing, global climate change, and

human evolution in Africa: introduction to the African paleoclimate special volume. Journal

of Human Evolution. 53, 443–464.

Maslin, M.A., Pancost, R.D., Wilson, K.E., Lewis, J., Trauth, M.H., 2012. Three and half million

year history of moisture availability of South : Evidence from ODP site 1085

biomarker records. Palaeogeography, Palaeoclimatology, Palaeoecology. 317-318, 41–47.

Maslin, M. A., Shultz, S., Trauth, M.H., 2015. A synthesis of the theories and concepts of early

human evolution. Philosophical Transactions of the Royal Society B: Biological Sciences.

370, 20140064–20140064.

Maslin, M.A., Trauth, M.H., 2009. Plio-Pleistocene East African pulsed climate variability and

its influence on early human evolution. In: Vertebrate Paleobiology and Paleoanthropology.

pp. 151–158.

Maxwell, S.J., Hopley, P.J., Upchurch, P., Soligo, C., 2018. Sporadic sampling, not climatic

forcing, drives observed early hominin diversity. Proceedings of the National Academy of

Sciences. 115, 4891–4896.

McKay, R., 2014. Did Antarctica initiate the ice age cycles? Science. 346, 812–813.

McKee, J.K., 2017. Correlates and catalysts of hominin evolution in Africa. Theory in

Biosciences. 136, 123–140.

Nie, J., 2018. The Plio-Pleistocene 405-kyr climate cycles. Palaeogeography, Palaeoclimatology,

Palaeoecology. 510, 26–30.

Ogura, T., Abe-Ouchi, A., 2001. Influence of the Antarctic ice sheet on southern high latitude

climate during the Cenozoic: Albedo vs topography effect. Geophysical Research Letters. 89

28, 587–590.

Ogura, T., Abe-Ouchi, A., Hasumi, H., 2004. Effects of sea ice dynamics on the Antarctic sea ice

distribution in a coupled ocean atmosphere model. Journal of Geophysical Research C:

Oceans. 109, doi: 10.1029/2003JC002022

Potts, R., 1996. Evolution and climate variability. Science. 273, 922–923.

Potts, R., 1998. Variability selection in hominid evolution. Evolutionary Anthropology. 7, 81–

96.

Potts, R., Faith, J.T., 2015. Alternating high and low climate variability: The context of natural

selection and speciation in Plio-Pleistocene hominin evolution. Journal of Human

Evolution. 87, 5–20.

Quade, J., Levin, N., Semaw, S., Stout, D., Renne, R., Rogers, M., Simpson, S., 2004.

Paleoenvironments of the earliest stone toolmakers, Gona, Ethiopia. Geol. Soc. Am. Bull.

116, 1529–1544.

Rannikko, J., Žliobaitė, I., Fortelius, M., 2017. Relative abundances and palaeoecology of four

suid genera in the Turkana Basin, Kenya, during the late Miocene to Pleistocene.

Palaeogeography, Palaeoclimatology, Palaeoecology. 487, 187–193.

Raymo, M.E., 2006. Plio-Pleistocene Ice Volume, Antarctic Climate, and the Global 18O

Record. Science. 313, 492–495.

Reed, D.N., Geraads, D., 2012. Evidence for a late Pliocen faunal transition based on new roden

assemblages from Oldowan locality Hadar A.L. 894, Afar Region, Ethiopia. Journal of

Human Evolution. 62, 328–337.

Reed, K.E., 1997. Early hominid evolution and ecological. Journal of human evolution. 32, 289–

322. 90

Reynolds, S.C., 2007. Mammalian body size changes and Plio-Pleistocene environmental shifts:

implications for understanding hominin evolution in East and southern Africa. Journal of

Human Evolution. 53, 528–548.

Reynolds, S.C., Bailey, G.N., King, G.C.P., 2011. Landscapes and their relation to hominin

habitats: Case studies from Australopithecus sites in eastern and southern Africa. Journal of

Human Evolution. 60, 281–298.

Rohling, E.J., Grant, K.M., Roberts, A.P., Larrasoaña, J.-C., 2013. Paleoclimate Variability in

the Mediterranean and Regions during the Last 500,000 Years. Current

Anthropology. 54, S183–S201.

Rose, C., Polissar, P.J., Tierney, J.E., Filley, T., de Menocal, P.B., 2016. Changes in northeast

African hydrology and vegetation associated with pliocene-pleistocene sapropel cycles.

Philosophical Transactions of the Royal Society B: Biological Sciences. 371,

doi: 10.1098/rstb.2015.0243.

Scholz, C.A., Johnson, T.C., Cohen, A.S., King, J.W., Peck, J.A., Overpeck, J.T., Talbot, M.R.,

Brown, E.T., Kalindekafe, L., Amoako, P.Y.O., Lyons, R.P., Shanahan, T.M., Castaneda,

I.S., Heil, C.W., Forman, S.L., McHargue, L.R., Beuning, K.R., Gomez, J., Pierson, J.,

2007. East African megadroughts between 135 and 75 thousand years ago and bearing on

early-modern human origins. Proceedings of the National Academy of Sciences. 102,

16416–16421.

Shultz, S., Maslin, M., 2013. Early Human Speciation, Brain Expansion and Dispersal

Influenced by African Climate Pulses. PLoS ONE. 8, e76750

Shultz, S., Nelson, E., Dunbar, R.I.M., 2012. Hominin cognitive evolution: Identifying patterns

and processes in the fossil and archaeological record. Philosophical Transactions of the 91

Royal Society B: Biological Sciences. 367, 2130–2140.

Singarayer, J.S., Burrough, S.L., 2015. Interhemispheric dynamics of the African rainbelt during

the late Quaternary. Quaternary Science Reviews. 124,

dx.doi.org/10.1016/j.quascirev.2015.06.021

Stenseth, N. C., Maynard Smith, J., 1984. Coevolution in ecosystems: Red Queen evolution or

stasis? Evolution. 38, 870–880.

Stollhofen, H., Stanistreet, I.G., Von Hagke, C., Nguno, A., 2014. Pliocene-Pleistocene climate

change, sea level and uplift history recorded by the Horingbaai fan-delta, NW Namibia.

Sedimentary Geology. 309, 15–32.

Tamuntuan, G., Bijaksana, S., King, J., Russell, J., Fauzi, U., Maryunani, K., Aufa, N.,

Safiuddin, L.O., 2015. Variation of magnetic properties in sediments from Lake Towuti,

Indonesia, and its paleoclimatic significance. Palaeogeography, Palaeoclimatology,

Palaeoecology. 420, 163–172.

Thackeray, J.F., Reynolds, S., 1997. Variability in Plio-Pleistocene climates, habitats, and

ungulate biomass in southern Africa. South African Journal of Science. 93, 171–172.

Thomas, E.K., Clemens, S.C., Sun, Y., Prell, W.L., Huang, Y., Gao, L., Loomis, S., Chen, G.,

Liu, Z., 2016. Heterodynes dominate precipitation isotopes in the East Asian monsoon

region, reflecting interaction of multiple climate factors. Earth and Planetary Science

Letters. 455, 196–206.

Tiedemann, R., Sarnthein, M., Shackleton, N.J., 1994. Astronomic timescale for the Pliocene

Atlantic δ18O and dust flux records of Ocean Drilling Program Site 659. Paleoceanography.

9, 619–638.

Tierney, J.E., Russell, J.M., Huang, Y., Sinninghe Damsté, J.S., Hopmans, E.C., Cohen, A.S., 92

2008. Northern hemisphere controls on tropical southeastAfrican climate during the past

60,000 years. Science. 322, 252–255.

Tierney, J.E., Russell, J.M., Sinninghe Damsté, J.S., Huang, Y., Verschuren, D., 2011. Late

Quaternary behavior of the East African monsoon and the importance of the Congo Air

Boundary. Quaternary Science Reviews. 30, 798–807.

Trauth, M.H., Bergner, A.G.N., Foerster, V., Junginger, A., Maslin, M.A., Schaebitz, F., 2015.

Episodes of environmental stability versus instability in Late Cenozoic lake records of East

Africa. Journal of Human Evolution. 87, 21–31.

Trauth, M.H., Larrasoaña, J.C., Mudelsee, M., 2009. Trends, rhythms and events in Plio-

Pleistocene African climate. Quaternary Science Reviews. 28, 399–411.

Trauth, M.H., Maslin, M.A., Deino, A.L., Strecker, M.R., Bergner, A.G.N., Dühnforth, M.,

2007. High- and low-latitude forcing of Plio-Pleistocene East African climate and human

evolution. Journal of Human Evolution. 53, 475–486.

Van Valen, L., 1973. A New Evolutionary Law. Evolutionary Theroy. 1, 1–30.

Venditti, C., Meade, A., Pagel, M., 2010. Phylogenies reveal new interpretation of speciation and

the Red Queen. Nature. 463, 349–352.

Verschuren, D., Sinninghe Damsté, J.S., Moernaut, J., Kristen, I., Blaauw, M., Fagot, M., Haug,

G.H., 2009. Half-precessional dynamics of monsoon rainfall near the East African Equator.

Nature. 462, 637–641.

Villmoare, B., Kimbel, W.H., Seyoum, C., Campisano, C.J., DiMaggio, E.N., Rowan, J., Braun,

D.R., Arrowsmith, J.R., Reed, K.E., 2015. Paleoanthropology. Early Homo at 2.8 Ma from

Ledi-Geraru, Afar, Ethiopia. Science. 347, 1352–1355.

Voje, K.L., Holen, O.H., Liow, L.H., Stenseth, N.C., 2015. The role of biotic forces in driving 93

macroevolution: beyond the Red Queen. Proceedings of the Royal Society B: Biological

Sciences.282: 20150186, dx.doi.org/10.1098/rspb.2015.0186.

Vrba, E.S., 1985. Ecological and Adaptive Changes Associated With Early Hominid Evolution.

In: Delson, E., ed. Ancestors: The Hard Evidence: Proceedings of the Symposium Held at

the American Museum of Natural History Agpril 6-10, 1984 to Mark the Opening of the

Exhibition "Ancestors, Four Million Years of Humanity". New York, A.R. Liss. 366 p.

Vrba, E.S., 1993. Turnover-pulses, the Red Queen, and related topics. American Journal of

Science. 293, 418–452.

Weldeab, S., Lea, D.W., Schneider, R.R., Andersen, N., 2007. 155,000 Years of West African

monsoon and ocean thermal evolution. Science. 316, 1303–1307.

Wennrich, V., Minyuk, P.S., Borkhodoev, V., Francke, A., Ritter, B., Nowaczyk, N.R.,

Sauerbrey, M.A., Brigham-Grette, J., Melles, M., 2014. Pliocene to pleistocene climate and

environmental history of Lake El’gygytgyn, Russian Arctic, based on high-

resolution inorganic geochemistry data. Climate of the Past. 10, 1381–1399.

Werdelin, L., Lewis, M.E., 2005. Plio-Pleistocene Carnivora of East Africa: species richness and

turnover patterns. Zoological Journal of the Linnean Society. 144, 121–144.

Westover, K.S., Stone, J.R., Yost, C.L., Scott, J.J., Cohen, A.S., Rabideaux, N., Stockhecke, M.,

Kingston, J.D. 2019. Diatom paleolinology of late Pliocen Baringo Basin (Kenya)

paleolakes. Palaeogeogr. Paleoclimatol. Palaeoeco. In review.

Wilson, K.E., Maslin, M.A., Leng, M.J., Kingston, J.D., Deino, A.L., Edgar, R.K., Mackay,

A.W., 2014. East African lake evidence for Pliocene millennial-scale climate variability.

Geology. 42, 955–958.

Woodard, S.C., Rosenthal, Y., Miller, K.G., Wright, J.D., Chiu, B.K., Lawrence, K.T., 2014. 94

Paleooceanography. Antarctic role in northern hemisphere glaciation. Science. 346, 847–

851.

Wündsch, M., Haberzettl, T., Kirsten, K.L., Kasper, T., Zabel, M., Dietze, E., Baade, J., Daut,

G., Meschner, S., Meadows, M.E., Mäusbacher, R., 2016. Sea level and climate change at

the southern Cape coast, South Africa, during the past 4.2 kyr. Palaeogeography,

Palaeoclimatology, Palaeoecology. 446, 295–307.

Yost, C.L., Ivory, S.J., Deino, A.L., Rabideaux, N., Kingston, J.D., Cohen, A.S. 2019.

Phytoliths, pollen, and microcharcoal from the Baringo Basin, Kenya reveal savana

dynamics during the Plio-Pleistocen transition. Palaeogeogr. Palaeoclimatol. Palaeoecol. In

review.

Žliobaitė, I., Fortelius, M., Stenseth, N.C., 2017. Reconciling taxon senescence with the Red

Queen’s hypothesis. Nature. 552, 92–95.

95

Figures

Figure 1: Locations of the paleoenvironmental records used in this study. 96

Figure 2: Z-scores of the original data set plotted with the reconstructed time series developed using reconstructed components calculated through singular spectral analysis. 97

Figure 3: Results of the evolutionary algorithm run on reconstructed time-series from eastern, western and southern Africa plotted with the known span of human fossils and FADs of tool 98 technologies. The grey dashed line denotes the 2.8 Ma pulse-turnover event seen in the fossil record of Eastern Africa as well as the earliest occurrence of the genus Homo (Lewis and

Werdelin, 2007; Werdelin and Lewis, 2005; Bobe and Behrensmeyer, 2004; Villmoare et al.,

2015). Increased standard deviation of a population denotes phenological plasticity while increased fitness indicates specialization.

99

APPENDIX C:

NEW APPROACHES TO QUANTIFYING PALEOENVIRONMENTAL VARIABILITY AND

APPLICATION TO UNDERSTANDING THE EVOLUTION AND DISPERSAL OF

ANATOMICALLY MODERN HUMANS OVER THE PAST 400 KA

BILLINGSLEY, Anne L.1, YOST, Chad L.1, FOX, Mathew L.2, KIELHOFER, Jennifer1,

WORTHEY, Kayla2, FENERTY, Brendan1, NAIMAN, Zachary1, GROVE, Matt 3, KUHN,

Steven L.2 and COHEN, Andrew S.1

1. Department of Geosciences, University of Arizona, Tucson, Arizona

2. School of Anthropology, University of Arizona, Tucson, Arizona

3. Department of Archaeology, Classics and Egyptology, University of Liverpool,

Liverpool, UK

Prepared for submission to the Journal of Human Evolution 100

Key Points

1. New metrics were developed to determine paleoenvironmental variability using relatively continuous time series data on a temporal and spatial scale relevant to human speciation, technological advancements and dispersals and which can be compared and aggregated directly between various paleorecord types. 2. Both the Standard Deviation Variability Method and the Accumulated Weighted Rate of Change Method recorded amplitude changes in detrended z-score time series of paleoenvironmental records which correlated with shifts seen in previous studies. 3. A record of climate variability developed for five regions in and around Africa over 400 ka shows that variability is not synchronous across the continent and region. 4. The first appearance of the Middle Stone Age is during a period of high climate variability in North Africa and Eastern Africa. 5. Dispersals in and out of Africa occur when the source region switches from high variability into a period of low variability.

Abstract

Numerous hypotheses have correlated key events in human evolution and dispersals with environmental changes. Some of these hypotheses operate within a unidirectional backdrop of environmental change whereas others, like the variability selection hypothesis, focus on the degree of environmental variability hominin species have experienced irrespective of the specific environmental conditions. The more recent accumulated plasticity hypothesis states that a species that experiences increased temporal changes in their environment develop adaptive strategies that allow for dispersal into multiple habitats. Comparisons between paleoenvironmental and paleoanthropological records have proven difficult for numerous reasons. There are large uncertainties in the timing of specific events in hominin evolution as well as the difficulty in objectively determining what qualifies as a period of high or low paleoenvironmental variability. In this study, we attempt to address the latter of these two problems by developing new metrics of paleoenvironmental variability. The Standard Deviation

Variability Method (SDV) and the Accumulated Weighted Rate of Change Variability Method

(AWRCV) are easily applied to most paleoenvironmental time series records. These new 101 methods allow for direct comparisons of degrees of paleoenvironmental variability between different records. The application of the metrics demonstrated here offer insight into the inter- regional history of climate variability over the last 400 ka. Furthermore, the records span the

African continent as well western Asia and southern Europe. These approaches to the analysis of climate variability provide a detailed spatial understanding of how multiple regions relevant to the rise and expansion of Homo sapiens responded to global climatic events.

Our study shows that five separate climate regions experienced both periods of high and low climate variability, and that these periods are not synchronous across Africa, southern

Europe and western Asia. Higher latitude regions are more susceptible to global climate forcing as is tropical Africa. Eastern Africa has a dampened response to glacial/interglacial cycles except for the transition out of MIS 5 and into MIS 4 and 3, which is a period of high climate variability for the region. The first appearance of anatomically modern human morphological features as well as the earliest recognized Middle Stone Age stone tools correlate with a period of high climate variability in both North and Eastern Africa. Furthermore, times of inferred major human dispersals occur after periods of increased climate variability in the source region, with the prolonged period of repeated migrations during MIS 5 occurring during a period of increased variability everywhere but eastern Africa. However, inferred dispersal at ~60 ka occurred after a period of extreme variability in eastern Africa.

1. Introduction

There have been many proposed hypotheses about how environmental history could have influenced hominin evolution through time. The ways in which environmental history might elicit evolutionary responses in organisms revolve around three broad scenarios: selection in the face of stable environmental forcing, as a result of directional environmental change, or as a 102 consequence of increasing levels of environmental variability. The first scenario incorporates a fundamental aspect of natural selection, selective consistency, in which a stable environment allows for selection towards fitness in one specific habitat, i.e. open grasslands or closed woodlands (Potts, 1998). The second scenario of directional change suggests that consistent selection occurred over time as the environment changes from one optimum to another. In the case of early hominin evolution in Africa, this scenario is exemplified by the savannah hypothesis, in which hominin adaptations of bipedality and perhaps other attributes evolved as a consequence of a broad directional shift from closed woodland environments to open grasslands

(Washburn, 1960). The third set of hypotheses suggest that rather than adapting to any one particular environment, either to a stable, consistent environment or to a directional change from one environment to another, a species may develop a suite of adaptive strategies to deal with large and repeated variability in the environment (Potts, 1998). In this scenario, fluctuations in the environment create an inconsistency in the selection process which allows for habitat specific adaptations to be replaced by adaptive strategies selected to deal with a variety of novel environments (Potts, 1996, 1998). In hominin evolution this is has been referred to as the variability selection hypothesis (VSH) (Potts, 1996, 1998). The VSH has been widely discussed as a driver of human evolution as environmental records, specifically from eastern Africa, have indicated that the unidirectional move towards increased aridity and open landscapes is overlaid with a pattern of increased variability in the region since the Plio-Pleistocene (deMenocal, 1995;

Potts, 1996, 1998; Maslin and Trauth, 2009). In addition to the VSH, the accumulated plasticity hypothesis (APH) proposes that populations that developed adaptive strategies suitable for more variable environments would be out competed by those more suited for any one particular environment as the environment becomes more stable (Grove, 2014, 2015). APH suggests that 103 those with accumulated plasticity (i.e. a suite of characteristics and technology capable of dealing with a variety of environments) could disperse into novel regions as the environment in the source region stabilized (Grove, 2014, 2015).

Any discussion of the role of environmental variability as a driver of hominin evolution must consider how to define variability from the perspective of geological or palaeoclimatological records as well as how to determine a way to quantifying exactly what constitutes a more variable environment on a timescale relevant to human evolution. Previous attempts at solving this problem have implicitly assumed that environmental variability is registered in the range of variation expressed by orbitally-controlled insolation (Potts, 1998,

2013; Potts and Faith, 2015). This approach is appealing because past insolation is readily calculated and has shown significant explanatory power in understanding the tempo of hominin evolutionary events. However, although orbitally forced insolation certainly plays a major role in controlling environment, its history of increased or decreased rates of change has two significant limitations as a metric of environmental variability. First, it varies over time frames that are rather large. Second, and most importantly, the record of insolation by itself does not encapsulate the entire range of variability experienced by hominins on the landscape. Specifically, it does not register the heterogeneous responses of climate to insolation caused by topographic, oceanic, glacial and continental influences and feedbacks.

In this study, we develop new approaches to estimating past environmental variability directly from aggregated paleorecords. We introduce two new quantitative metrics in defining paleoenvironmental variability, which can be applied to any type of paleorecord, no matter what the initial indicator is or what the indicator represents (i.e. precipitation, temperature, etc.). We also introduce methods for stacking these variability metric time series to produce a master 104 record of the history of environmental variability at a regional scale. As an example of how this approach can be used, we applied these methods to a suite of 31 paleorecords from around Africa and in adjacent regions of , spanning the last 400-ka, to develop a picture of paleoenvironmental variability throughout this region and time period. We then sorted these records into regional groups to explore possible relationships between climate variability and known or inferred events related to hominin evolution and dispersals. Going forward, these methods promise an improved way to measure environmental variability over even longer timescales, to directly compare human evolutionary milestones to variability at any given site, and thus allow us to understand better how climate and environmental variability could have influenced the evolutionary history of hominins.

1.1 Regional Climate Variability Over the Study Area

The spatial extent of this study is extensive and encompasses a vast range of environments, climates, and seasonality. The mechanisms that control spatial and temporal patterns of climate variability across this region are complex, with internal and external forcing and feedbacks playing a role in regional responses to global climate events, such as variations in orbital forcing and glaciations (Blome et al., 2012). Precipitation patterns, both today and in the past, are controlled by a variety of mechanisms such as the position of the Westerlies, the migration of both the tropical rain belt and the Congo Air Boundary (CAB), the inland extent of the West African Monsoon (WAM), local insolation and sea surface temperatures (Nicholson,

2000; Nicholson, 2018). These climatological features respond to orbital forcing on the

Milankovitch cycles of precession (19 kyr -23 kyr), obliquity (41 kyr) and eccentricity (100 kyr), and are often coincident with Northern Hemisphere stadial/interstadial patterns. There are also topographical controls on local convection processes. In order to simplify the complexity created 105 by multiple climate mechanisms, as well as determine climate variability at a spatial and temporal scale experienced by early AMH, a regional approach is needed, which guides our analysis and the geographic partitioning of data used below.

1.1.1 Southern Europe and Southwest Asia

The region of southern Europe and southwestern Asia incorporated in this study extends across the eastern Mediterranean between 35⁰ N and 45⁰ N and between 20° E and 43° E. There is a climatic difference between the northern Levant and the southern Levant created by the placement of atmospheric currents and topography which generate a distinct N-S and E-W gradient in precipitation and vegetation (Belmaker and Hovers, 2011; Gasse et al., 2015). The region receives most of its rainfall during Boreal winter with the southward displacement of the

Westerlies. Therefore, the area experiences cold wet winters and hot dry summers (Rohling et al., 2013; Kwiecien et al., 2014). Over Pleistocene time scales, the climate in this region is strongly regulated by Northern Hemisphere glaciation events, with a southward displacement of vegetation and climate zones during glacial advances and increased moisture availability in the region during interglacials (Rohling et al., 2013; Kwiecien et al., 2014; Gasse et al., 2015).

Periods of intense aridity correlate to colder temperatures in Europe (Stockhecke et al., 2014).

Interglacials may also be marked by increased seasonality and spring/summer temperatures in the region (Francke et al., 2016a). There has also been an increase in climate variability in the region over the last 500 ka (Rohling et al., 2013). This variability represents switches from relatively arid to mesic habitats, with a precipitation swings from 550 mm/year to 1,700 mm/year

(Blumenthal et al., 2017). Sea surface temperatures (SSTs) in the Mediterranean experience fluctuations on millennial scales correlating with Dansgaard-Oescheger and Heinrich events, 106 which influences moisture availability and local advection processes in the region (Rohling et al.,

2013).

1.1.2 Northern Africa and the Southern Levant

Northern Africa and the southern Levant region is defined in the study as the area from

15⁰ N to 35⁰ N from the Atlantic Ocean to the eastern Mediterranean. The area is currently dominated by the extremely arid Sahara and is bounded on the south by the . The aridity is mainly controlled by the subduction of the northern arm of the Hadley cell, which prevents local convection processes and onshore moisture transport (Nicholson, 2000). In the past, however, there have been periods of humidity in the region during interglacials when more moisture has been documented on the landscape, specifically during the African Humid Period, approximately

15 ka – 5 ka, and during MIS 5, 135 ka -115 ka and 105 ka – 75 ka (Drake et al., 2010; Blome et al., 2012a; Larrasoaña et al., 2013; Grant et al., 2017; Tierney et al., 2017a; Quade et al., 2018).

These periods correlate to higher SST in the North Atlantic that allowed for an increased land/sea temperature and pressure gradient as well as higher moisture availability and development of a stronger monsoonal system (Grant et al., 2017). Vegetation on the landscape in this region varies on orbital scales, with Mediterranean forests expanding southward into Africa and grasslands expanding further north into the Sahara during interglacials (Dupont, 2011;

Larrasoaña et al., 2013).

1.1.3 Eastern Africa

Eastern Africa is here defined as the area of the continent between 15⁰ N and 5⁰ S and east of 20⁰ E. It is differentiated from tropical and northern Africa by the presence of several rift systems that topographically and climatically isolate the region (Maslin and Trauth, 2009;

Tierney et al., 2011). Most of the region currently has two rainy seasons controlled by the 107 migration of the tropical rain belt as it follows local insolation across the equator bringing moisture into the region from the Indian Ocean (Nicholson, 2000). Climate modeling coupled with studies of past lake level fluctuations indicate orbitally driven wet-dry climate cycles on the order of 800 kyr, 400 kyr, 41 kyr and 23 kyr (Kutzbach and Otto-Bliesner, 1982; Deino et al.,

2006; Kingston et al., 2007a; Campisano and Feibel, 2008; Maslin and Trauth, 2009; Maslin et al., 2015). Millennial, multidecadal and interannual scale variability has also been observed in some lake records (Shultz and Maslin, 2013; Wilson et al., 2014). The region has experienced increased aridity and climate variability since the onset of Northern Hemisphere glaciation

(deMenocal, 2004; Maslin and Trauth, 2009).

1.1.4 Tropical Africa

In this study, we define tropical Africa from 5⁰ N to 15⁰ S from the Atlantic Ocean on the west to Lake Malawi in the east. It is separated from eastern Africa geographically by the East

African Plateau and climatically by the meridional placement of the CAB (Nicholson, 2000,

2018). The northern extent of tropical Africa is the Sahel, which separates tropical Africa from the Sahara. The climate of tropical Africa is mainly controlled by fluctuations in the West

African Monsoon system (WAM), which is greatly influenced by sea surface temperatures in the equatorial East Atlantic as well as the intensification of solar radiation in the Northern

Hemisphere (Lezine, 1991; Weldeab et al., 2007; Tierney et al., 2008; Shanahan et al., 2015).

Dry intervals in tropical Africa are associated with periods of high-latitude cooling and a weakened Atlantic Meridional Overturning Circulation (AMOC) (Shanahan et al., 2015). The eastern extent of the region has also been shown to be influenced by SST in the Indian Ocean, with increased temperature assisting in local advection processes (Tierney et al., 2008). 108

There is a history of abrupt changes in precipitation in the region recorded in both offshore terrigenous input in the Atlantic and in onshore lake records (Adegbie et al., 2003;

Tierney et al., 2008). Biomes in the region fluctuate latitudinally with changes in the intensification of the trade winds and the inland extent of the WAM as recorded in offshore pollen records, with interglacial periods correlating with humid forests and increased river input while glacial periods have increased aeolian transport and a more open, grassland vegetation

(Lezine, 1991). Furthermore, vegetation in the region has been shown to correlate well with the interannual extent of the tropical rain belt, which is directly influenced by local insolation intensity (Kutzbach and Liu, 1997; Ivory et al., 2018). Much of this region, as well as parts of

Eastern Africa, experienced episodes of extremely arid conditions between 135 – 90 ka, during a period of relatively high climate variability (Scholz et al., 2011a, 2011b).

1.1.5 Southern Africa

Southern Africa is defined here as the region of the continent south of 15⁰ S. The climate is greatly influenced by the cool Benguela current on the west, created by cold water upwelling off the coast, and the warm Agulhas current to the east, which brings warm water south from the equatorial Indian Ocean. The western part of South Africa receives most rainfall during the

Austral winter, when the Westerlies move north over the continent, while the eastern part of southern Africa receives most of its rainfall during the summer months (Botai et al., 2018). Lake level records from Paleo-lake Makgadikgadi indicate that the region experienced punctuated humid periods over 140 ka, with arid periods occurring 145 ka – 140 ka, 125 ka – 122 ka, 110 –

90 ka, and 50 ka – 45 ka (Burrough and Thomas, 2009). Pollen assemblages from ocean drill cores taken of southeastern Africa indicate that during interglacials the vegetation is mainly evergreen and deciduous forests while mountainous scrubland dominate during glacials (Dupont, 109

2011). Montane forests and C3 vegetation expanded during the warm, humid periods of interglacials when higher local insolation and increased SST of the Agulhas current allowed for more precipitation (Dupont et al., 2011; Castañeda et al., 2016). The area also experiences strengthened trade winds during glacial periods (Ning and Dupont, 1997).

2. Methods

Since we are interested in quantifying and comparing patterns of variability within and between paleorecords, it is important to initially remove any long-term trends in a data set that may impact the calculation of short-term variance that is of interest. Determining an appropriate approach to detrending a data set for this application is not straightforward because the removal of higher order (polynomial) trends could inadvertently remove some of the very patterns of internal variability we are interested in quantifying. As a result, we take a conservative approach, and only apply a first-order linear detrend to our data sets (Figure 1b). Since our ultimate goal is to aggregate and compare data sets of many different types, and with many different numeric ranges of variability, it is necessary to normalize all data as z-scores around the detrended means for each data set (Figure 1c).

The next step is to determine appropriate time bin durations over which to examine the aggregated variability of the z-score data points. This decision must be based on both the question being asked and the available time series data density. This process may require some trial and error assessments of various bin durations in order to determine an adequate number of data points per bin. For the purposes of this study, we chose to parcel the z-scores into 10 kyr bins (Figure 1d-e). The selection of a 10 kyr bin allowed for the highest resolution to be achieved while retaining data adequacy in each bin. Smaller bin sizes often lacked data points in multiple bins in most of the time series preventing calculations of variability in those records for those 110 periods. Also, the VSH is a time-integrated selective process which requires environmental variability through multiple generations (Potts, 1998). Modeling of human population responses to climate variability suggests that maximum selection towards plasticity may be achieved within

500 generations, implying that 10 kyr time bins of climate variability may be an objectively useful duration for comparison to hominin evolutionary events assuming a hominin generation time of ~20yr (Grove, 2015). In order to maintain continuity of the record for each bin, values were linearly interpolated for the time series at the bin breaks. For example, for our second 10kyr bin which begins at time step 10 ka and ends at time step 20 ka, values were calculated for those two time steps if values were not present in the original paleoenvironmental record. This allows for beginning and ending points for the calculation of variability in each bin.

Variability embedded in a paleoenvironmental time series record could, in principle, be quantified in various ways. Here we develop two relatively simple approaches that are applicable to many data sets. Our first approach relies on the measurement of the variance of data within the defined bin by simply calculating the standard deviation of the data points for that bin. By comparing individual bins throughout the entire time-series, periods of high or low variability can be determined by the amount of variance within any given bin. To put it more simply, bins with higher standard deviations are assumed to represent periods of higher variability. In this method, it is the range of the data which defines the variability. We refer to this approach as the

Standard Deviation Variability Method (SDV) (Figure 1e). This approach has the great advantage of simplicity in calculation but could be subject to sampling artifacts when the data points are not evenly distributed throughout the time bin or if there are too few data points to have a statistically significant standard deviation. 111

Our second approach to determining climate variability relies on the change occurring between sequential data points. In this method, the absolute value of the rate of change, or slope, is calculated between all points within a given time bin. These slopes are then weighted based on the duration of the bin (i.e. if the timing between two data points is 1 kyr and the bin duration is

10 kyr, the slope is weighted 1/10 of the entire change over the duration of the bin). Weighting the slopes allows for appropriate representation of rates of change even in records with an uneven distribution of data. The calculated slopes, normalized for the bin duration, are then summed for each bin of the time series. We refer to this method as Accumulated Weighted Rate of Change Variability Method (AWRCV). In this method, it is the change between data which defines variability. It is assumed that bins with the highest values for accumulated rate of change are periods of higher environmental variability. This method is more useful than SDV when the data are unevenly distributed or if there are too few data points to accurately determine a standard deviation as a slope can be determined with less data and may represent the total variability across a bin.

The mean, standard deviation and an accumulated weighted rate-of-change were calculated for each of the 10 kyr bins. Interpolated data was not included in the calculation for the SDV method. The records were then divided into five regions, and the AWRCV metric was calculated for each region for each 10 kyr bin. Each 10 kyr bin of the regional records were bootstrapped 1000 times using MATLAB® ‘bootstrap’ function. The mean and the standard deviation of the bootstraps were used to create the upper and lower limits of the confidence bands. Periods of higher variability were defined as those above one standard deviation of the mean variability for each region in any one bin similar to the previous approach to estimating environmental variability by Potts and Faith (2015). 112

For an evenly dispersed time series, the SDV and the AWRCV will produce proportionally very similar results, as shown by applying the methods to an evenly sampled portion of an insolation curve for the equator (Figure 1). For the remainder of this application we will only use the latter approach, the AWRCV.

2.1 Methods Specific to This Study

Time series data sets for this study consisted of terrestrial records from caves, lakes and outcrops as well as offshore ocean drill cores, encompassing a variety of indicators of climate and environmental processes (Table 1). This study uses 31 different climate reconstructions from

23 locations (Figure 2). Because of the regional climate heterogeneity discussed previously, we wanted to explore our data subdivided geographically. Accordingly, each record was placed within one of five geographic regions: Southern Europe and Southwest Asia, Northern Africa and the Southern Levant, Eastern Africa, Tropical Africa and Southern Africa. The regions were determined following Blome et al. (2012), based on regionally consistent patterns of climate over extended periods (Table 1). Individual variability time series within a region were stacked by time bin by taking the means for each bin (Figure 3).

The records incorporated in this study were chosen because they were both relatively continuous over the 400-ka time duration selected for this study and of a high enough resolution for our methods to be applied. The quantifying measures were run on the original data set when possible. For some complex, multivariate data sets such as pollen analyses, it is necessary to initially create a simplified univariate metric for analysis. We did this using correspondence analysis (CA) on the fourteen pollen records incorporated in this study, using the CA first axis eigenvalues as our transformed time series. 113

Correlation analysis was performed between the averaged summed rate of change per 10 kyr of each region and the corresponding summed rate of changes for possible forcing mechanisms (sea surface temperatures, glacial/interglacial changes and insolation) using the

Pearson r in the PAST 3.x analytical software. Significance was calculated using a two tailed t test with 2 degrees of freedom.

3. Results

3.1 Variability in insolation

The “proof-of-concept” application of our metrics were initially analyzed using insolation from the Equator calculated at 1 kyr time steps over 400 ka (Figure 1). Initial detrending of the insolation time series removed a small directionality imposed by the arbitrary 400 – 0 ka cutoffs.

The mean z-score for each 10 kyr bin tracked the amplitude changes in the original record well, with higher z-scores correlating with higher W/m2 (Figure 1d). Both the SDV metric for the bins and the AWRCV metric follow similar patterns, with the more extreme changes of amplitude of the original time series being reflected with increased values in both calculations (Figure 1e-f).

The periods of 140 – 110 ka and 230 – 170 ka have the highest values for both standard deviation (1.17 – 1.07 and 1.19 – 1.44, respectively) and rate-of-change (0.293 – 0.359 and 0.288

– 0.402, respectively). During these periods, insolation is oscillating between its highest and lowest values.

3.2 Variability in regional records

The mean of the z-scores for each bin of the 31 paleorecords used in this study follow the changes in amplitude of the original data. As seen in the insolation record reviewed above, the

SDV and AWRCV metrics for each record are in concordance with each other, with increases in both occurring when there are major transitions in the amplitude of the original records. All of 114 the regions have periods of high variability, in which the averaged summed rate-of-change for a

10 kyr bin is above one standard deviation from the mean (Figure 3).

The range of AWRCV values for southern Europe and southwestern Asia is 0.108 –

0.487, with a mean of 0.274 and a standard deviation of 0.059. There are three periods of high climate variability in this region, from 90 – 80 ka, 240 – 230 ka and 400 – 380 ka, when

AWRCV values for the 10 kyr bins were above 0.333. North Africa and the southern Levant have a range of 0.131 – 0.705 with a mean of 0.464 and standard deviation of 0.179. There are four periods of high climate variability from 70 – 60 ka, 130 – 110 ka, 350 – 340 ka and 390 –

380 ka, when the AWRCV was above 0.642. Tropical Africa has a range of 0.268 – 0.910 with a mean of 0.399 and standard deviation of 0.137. There were four periods of high variability from

20 – 10 ka, 100 – 90 ka, 210 -200 ka, and 400 – 390 ka, when the AWRCV was above 0.536.

Southern Africa AWRCV valued range between 0.104 – 0.807, with a mean of 0.456 and a standard deviation of 0.175. There were five periods of high variability from 10 – 0 ka, 140 –

130 ka, 200 – 190 ka, 220 – 210 ka and 340 – 330 ka, when the AWRCV was higher than 0.632.

Eastern Africa has an AWRCV range of 0.002 – 1.072 with a mean of 0.175 and a standard deviation of 0.233. There were three periods of high variability in this region, from 10 – 0 ka, 90

– 60 ka, and 330 – 320 ka, when the AWRCV was higher than 0.408 (Table 2).

The confidence bands for the AWRCV for the regional records are broader in regions with fewer records. This is most evident for the eastern African data where only two available data sets met our completeness criteria. There is also a notable decrease in confidence in the northern Africa and southern Levant record prior to 150 ka, where 6 of the 8 records are missing data. Our calculations fall within the range for confidence in all regions and times except for southern Africa between 400 – 315 ka, when the AWRCV of the region is beginning to be 115 determined by less individual records as two records do not extend to the full 400 ka. When the

AWRCV is at the upper most limit of the confidence band, the average regional variability of the records is highly influenced by the extreme variability in one record.

3.3 Correlations between regional variability and potential forcing mechanisms

Correlation analysis indicates that there is a significant positive correlation, 0.3881, between high variability in southern Europe and southwest Asia and tropical Africa (Table 3).

The tropical African record also displays a significant positive correlation with both insolation at the equator (0.3173) and at 35° S (0.32273). The southern African record is significantly and positively correlated with insolation at 65°N (0.35722), 35° N (0.355), the equator (0.31839), and 35°S (0.33301).

4. Discussion

4.1 New Methods for Quantifying Climate Variability and Possible Applications

One of the key questions in evolutionary studies is whether species are constantly evolving in stable environments under the influence of biotic selective forcing such as competition or escalatory predator-prey coevolution (i.e. Red Queen theory) or if speciation is occurring during infrequent, drastic physical environmental changes due to abiotic forcing

(Benton, 2010). It has been suggested that the Red Queen theory of constant evolutionary advances during periods of environmental stability is valid on smaller timescales and in a localized spatial extent, but may lose significance on a larger geological timescale as it is overwhelmed by global climate changes and other large-scale events such as orogenesis and volcanic eruptions (Barnosky, 2006). Attempts have been made to model different evolutionary scenarios to determine if there is a steady state of change in species even during consistent environmental stability or if change only occurs during extreme physical reconstruction of the 116 environment. These hypothetical scenarios have been difficult to test in nature because of the practical difficulty in determining that there are no changes in the physical environment for a given time period and place (Stenseth and Smith, 2006; Benton, 2010). Specifically, it has been difficult to determine what constitutes extreme environmental variability in relation to a stable state.

The methods put forth in this study may offer a framework to help determine what selective process or combination of processes (abiotic or biotic) may be acting upon a species or ecosystem at any given time. Both the SDV and the AWRCV are helpful metrics in defining periods of environmental instability as well as periods of relative stability. The methods can be applied to local, regional and global paleoenvironmental records of any type of indicator allowing for full functionality when determining environmental variability. This can allow for direct comparisons of environmental variability to paleobiological and paleontological records in almost all settings. Although this study focuses specifically on the regional variability in and around Africa over the last 400 ka in the context of human evolution and dispersals, these methods can be modified and applied to other evolutionary questions regarding environmental variability on both smaller and larger spatial-temporal scales.

4.2 Regional Climate Variability

Our study indicates considerable change in the degree of environmental variability within and between regions around Africa over the past 400 ka. The periods of high variability are not synchronous across the area of study indicating regional differences in responses to external forcing factors. The pattern of regional differences indicates that changes in variability that hominins in the region experienced over this time period cannot be completely encompassed by variability in orbitally-induced insolation alone. Throughout all five regions, there are clear 117 responses to interglacial/glacial transitions, with the global climate shifts from one state to another creating increased variability in any given region (Figure 4). However, regional differentiation can be seen as periods of high variability rarely overlap, even in regions adjacent to each other. For example, eastern Africa experienced a long period of high environmental variability, 90 – 60 ka, during the transition from MIS 5 through MIS 4 (Lisiecki and Raymo,

2005). Southern Europe and southwestern Asia, 90 – 80 ka, as well as northern Africa and the southern Levant, 70 – 60 ka, also experienced high environmental variability during this transition; however, the durations of the periods were shorter and did not overlap with each other

(Lisiecki and Raymo, 2005). This suggests that although the overall pace of environmental variability in and around Africa is controlled by orbitally-influenced global climate events, localized feedbacks and responses cannot be ignored.

Southern Europe and southwestern Asia experienced three periods of unusually high environmental variability over our study interval, between 400 – 380 ka, 240 – 230 ka, and 90 –

80 ka. The first of these periods of high climate variability occurred as the global climate shifted from the warm interglacial MIS 11 into MIS 10, an interglacial termination occurring during a period of high obliquity (Huybers and Wunsch, 2005). This event coincides with the retreat of temperate forests around Lake Ohrid (Macedonia), Yammouneh (Lebanon) and the Tenaghi

Philippon (Greece) regions, as well as a trend towards increased aridity and cooling throughout the eastern Mediterranean (Tzedakis et al., 2006; Gasse et al., 2015; Francke et al., 2016; Kousis et al., 2018). The second period of high climate variability for this region, 240 – 230 ka, coincides with a drastic drop in global temperature during MIS 7 (Lisiecki and Raymo, 2005).

MIS 7, 246 – 186 ka, was a period of at least five warm-cool climate shifts in Europe, with the most pronounced being the transition from MIS7e into MIS 7d, ~230 ka – 215 ka, when again a 118 transition from temperate woodlands to steppe vegetation occurred (Roucoux et al., 2006; López-

García et al., 2014). During this period, the Mediterranean experienced much cooler temperatures and an increase in the intensification of the north-westerlies (Cortina et al., 2011).

The final period of high climate variability for Europe and southwest Asia occurred during the cold period MIS 5b, when again the region transitioned from a warm, humid, forested regime to cooler temperatures, more open environments, and increased aridity (Helmens, 2014).

Northern Africa and the southern Levant experienced four periods of high variability over our study interval. The first, from 390 – 380 ka, overlaps with a similar period in southern

Europe and southwestern Asia. Northern Africa was more humid than present during MIS 11, with warmer SSTs in the North Atlantic and a northward displacement of the tropical rain belt.

However, as the climate transitioned into MIS 10, North Africa became much more arid, as indicated by increased dust flux off the North African coast (deMenocal, 1995; Helmke et al.,

2008; Trauth et al., 2009; Kandiano et al., 2012). This aridity may also have been influenced by increased upwelling in the North Atlantic, which limits onshore transport of moisture and increases millennial scale variability (Kandiano et al., 2012; Rohling et al., 2013). The second period of high climate variability occurred between 350 – 340 ka, a period that coincides with the onset of MIS 9. This transition is marked in the North Atlantic by increasing SSTs, ice sheet collapses, and a decrease in North Atlantic Deep-Water (NADW) formation, which may have facilitated weaker NE trade winds off the coast of Africa in MIS 9 compared to MIS 10 creating climate instability during this period (Kostygov et al., 2010; Dupont, 2011).

The onset of MIS 5 ushered in an extended period of climate variability for northern

Africa and the southern Levant between 130 – 110 ka. The region experienced a substantial increase in precipitation at approximately 130 ka, as recorded in both the Pequiin and Soreq cave 119 records, followed by a drastic decrease at 115 ka (Bar-Matthews et al., 2010). Increased effective moisture and a greening of the Sahara also occurred during this time, when an increase in northern hemisphere insolation allowed for a more prominent and extensive WAM (Kutzbach and Liu, 1997; Ruddiman, 2006; Drake et al., 2010; Dupont, 2011). The last period of high climate variability for this region occurred 70 – 60 ka as the global climate became cooler transitioning from MIS 5 into MIS 4, and northern Africa and the southern Levant became much more arid and woodlands and tropical forests contracted southwards towards the equator (Bar-

Matthews et al., 2010; Dupont, 2011).

Eastern Africa experienced three periods of high climate variability over our study interval, 10 –0 ka, 90 – 60 ka and 330 – 320 ka. During MIS 9, eastern Africa experienced a transition from arid to more humid conditions, as insolation at the equator increased. This allowed for a strengthening of the East African monsoonal system, and this transition was also marked by high variability in the region between 330 – 320 ka (Szabo et al., 1995; Owen et al.,

2018). The transition from MIS 5 to 4/3 coincides with a second period of high variability in eastern Africa, 90 – 60 ka. This period saw multiple lake level fluctuations at Chew Bahir and

Lake Tana in Ethiopia as well as sharp decreases in SST and rainfall in the (Trauth et al., 2015; Tierney et al., 2017a; Viehberg et al., 2018). The termination of the African Humid

Period, when the region experienced a severe drop in precipitation between 8 ka and 5 ka, created a period of high variability in eastern Africa, 10 – 0 ka (Tierney et al., 2011; Costa et al.,

2014; Owen et al., 2018).

Tropical Africa experienced four periods of high climate variability during our study period, 20 – 10 ka, 100 – 90 ka, 210 – 200 ka, and 400 – 390 ka. The first period of high variability corresponds with the shift from interglacial maxima of MIS 11 into MIS 10 (Lisiecki 120 and Raymo, 2005). During this transition, tropical Africa shifted from a wet, warm landscape with lowland tropical forests into more open, arid wooded grasslands (Dupont, 2011; Ivory et al.,

2016). Similarly, the second period of high variability in tropical Africa is associated with the transition from interglacial to glacial. From 210 – 200 ka, during the transition from interglacial

MIS 7 into MIS 6, there was a rapid change from a relatively wet to a very dry environment at

Lake Malawi, with an accompanying decrease in lake level (Ivory et al., 2016, 2018). The west coast of Africa also saw an expansion of grasslands with a decrease in precipitation during this period (Dupont, 2011).

Between 130 – 90 ka, tropical Africa experienced several periods of extended aridity, megadroughts, during which lake level at Lake Malawi fell as much as 600 m (Cohen et al.,

2007; Scholz et al., 2007). The third period of high variability, 100 – 90 ka, represents the transition of this region out of the megadrought (Ivory et al., 2016). During this period, Lakes

Malawi and Tanganyika experienced increases in lake level, and there was an expansion of tropical forests in (Dupont et al., 2001; Tierney et al., 2008; Burnett et al.,

2011; Dupont, 2011; Scholz et al., 2011a; Stone et al., 2011; Lyons et al., 2015). Our study indicates that relatively high variability persisted throughout the period from 150 – 70 ka when compared to the rest of the record. However, it is this major transition from hyperaridity into wetter conditions (~100 – 90 ka) which represents the most extreme change and, therefore, the highest variability. The Holocene was also a period of high environmental variability in tropical

Africa. From 10 ka to present there have been many shifts between wet and arid climates, the most drastic of which is the rapid transition out of the African Humid Period (Tierney et al.,

2008; Shanahan et al., 2015). 121

There were five periods of high variability in southern Africa over the past 400 ka observed in this study: 340 – 330 ka, 220 – 210 ka, 200 – 190 ka, 140 – 130 ka, and 10 – 0 ka.

The first of these intervals corresponds with the transition from MIS 10 to MIS 9 and coincides with an increase of offshore upwelling, which appears to have increase millennial scale variability and habitat fragmentation (Rohling et al., 2013). Furthermore, offshore records indicate that as the ice sheets retreated, southern Africa was getting wetter creating a landscape dominated by evergreen and deciduous forests rather than the mountainous scrubland seen during more arid glacials (Dupont, 2011; Caley et al., 2018). The next period of high variability,

220 – 210 ka, corresponds with the transition from MIS 7d to 7c. The transition from MIS 7 into

MIS 6, 200 – 190 ka, was marked by a more variable climate with landscape and hydroclimate change related to the restriction of the tropical rainbelt towards the Equator during glacial expansion and more arid environments in southern Africa (Castañeda et al., 2016; Caley et al.,

2018). The transition into the last interglacial was a period of high variability in southern Africa,

140 – 130 ka. Similarly, the Holocene is a period of high variability in the region as well as southern Africa transitioned from relatively cooler temperatures into a warmer climate (Scott,

1999).

The new methods developed for this study allowed us to determine periods of high climate variability in each of the five regions in and around Africa. The periods of variability seen in this study also coincide with increased variability noted previously for individual paleoenvironmental records, which gives us greater confidence concerning our new metrics.

Furthermore, this study shows that regions responded differently to global climate shifts based on local feedback processes, with variability in the potential forcing mechanisms not necessarily correlating with regional variability. 122

4.3 Potential forcing mechanisms and regional responses

The lack of correlation of periods of highest variability between the regions, and thus single, explanatory and global forcing mechanism(s) is a noteworthy result of this study. The temporal changes in each region do align well to glacial/interglacial shifts, and many paleoenvironmental studies have shown that sea surface temperature, insolation, and ice sheet expansion have influences over environmental changes. As our study has been able to identify periods of high variability in both regional and global records, we need to consider plausible explanations for the lack of correlation between these widely investigated drivers of global climate and regional variability. One possibility lies with the duration of the time step bins used in this analysis

As noted previously, a 10 kyr time bin was selected in part to explore landscape-scale environmental changes on timescales relevant to realistic models of hominin selective and dispersal responses to the environment. From this perspective longer duration bins might be less desirable because they would potentially encompass multiple forcing events. However, whereas a 10 kyr time bin is capable of capturing the full amplitude of the higher frequency events that could have significant selective implications for hominin populations, it may simultaneously be too short to capture the coherent changes in variability that exist in global scale processes. In part, this results from that fact that there are variably delayed responses in regional climate to most global forcing mechanisms (deMenocal et al., 2000; Abe-Ouchi et al., 2013; Shanahan et al., 2015). For instance, if there is a 3 kyr delay between the initial variability in the forcing mechanism and the regional environmental response, the peak variability for each element may end up in different temporal bins. To illustrate this, we can compare the effects of varying bin duration on the correspondence between variability metrics for the benthic foraminifera δ18O 123 isotope stack (Lisiecki and Raymo, 2005) and the dust flux record from ODP 721/722, collected off the coast of East Africa (deMenocal, 1995) (Figure 5).

For the 10 kyr bin example from 10 – 0 ka, the dust flux record displays both high SDV and AWRCV values, whereas the Lisiecki and Raymo curve is relatively flat, with low SDV and

AWRCV values. However, when the bin duration is expanded to a 20 kyr window, the analysis indicates that both records were experiencing a period of rapid change and, subsequently, higher climate variability. The expansion of the bin duration in this example captured the interval when the full range of amplitude change of the benthic δ18O stack could be tied to the full range of the amplitude change in the dust record. The higher SDV and AWRCV for both indicate this shift.

It should be noted that expanding the window may dampen the signal of climate variability as drastic shifts tend to be diluted mathematically over time by steady states, as we see in the 50 kyr time bin. Here the SDV and AWRCV are lower than the values seen in the 20 kyr bin. However, even at this bin size, periods of high variability can be differentiated from those of low variability. Perhaps not surprisingly then, when correlations are calculated over longer bin sizes, there are more significant correlations between the local records and global forcing mechanisms (Cohen et al., 2018 and in prep). But conversely, a single long duration bin’s measured variability may mask the “churning” evident from a series of sequential, shorter bin, high variability episodes that could take place over the longer bin’s duration. The takeaway message is that there is no one correct bin duration for a variability analysis. The choice must be made as a tradeoff between the needs for greater insights into regional impacts of variability versus an ability to infer potential global drivers of the same.

124

4.4 Climate Variability and Implications for Hominin Evolution and Dispersal

Differences in environmental variability such as those quantified in this study suggest a landscape constantly changing, creating both pathways for hominin dispersals and interactions as well as environmental fractionation, which would isolate populations from each other for extended periods of time (Kingston, 2007b; Blome et al., 2012; Scerri et al., 2014, 2018; Potts and Faith, 2015). Multiple hypotheses have been put forward in order to understand how both climate variability and climate stability may have influenced the speciation, technological advancement and dispersal of early anatomically modern humans. The variability selection hypothesis (VSH) suggests that periods of high variability in the climate would provide the environmental background needed to spawn the biological and technological adaptations seen among adaptively flexible anatomically modern humans (AMHs) (Potts, 1996). Drawing correlations between hominin evolutionary events and climate fluctuations is made difficult by our still limited and often poorly dated hominin fossil and archaeological record, uncertainties in age constraints in the paleoenvironmental records and uncertainties concerning calibrations of genetic clocks. However, some comparisons can be made in order to develop a framework of how climate variability may have served as a backdrop to the development and expansion of

AMH.

Prior studies suggest that environmental corridors may have allowed AMHs to disperse across previously uninhabitable places (Drake et al., 2010; Blome et al., 2012b; Larrasoaña et al.,

2013; Scerri et al., 2018). The accumulated plasticity hypothesis (APH) suggests that increased environmental variability would have allowed for selection towards plasticity, defined as traits which would allow an individual the ability to adapt to changing environments rapidly (Grove,

2014, 2015; Grove and Burke, 2015). In contrast, a stable environment would select populations 125 towards specialization (Figure 6) (Grove, 2015). Once a region switches from a period of high climate instability to a relative steady state, theory suggests that individuals with plasticity would be outcompeted in the new environmental regime, encouraging their dispersal into new territories, where they would be able to thrive given their previously accumulated plasticity

(Grove, 2014, 2015).

As summarized above, our 400-ka record of climate variability indicates that the five regions in and around Africa experienced both periods of climate stability and instability, with many of these periods differing between regions for any given time period. If the APH and the

VSH were actually driving factors in human evolution and dispersal events over this timescale, then we would expect to see a relationship in time between these two types of data (e.g. Potts and

Faith, 2015).

Genetic studies indicate that AMH split from the Neanderthals approximately 500 ka, but the first fossil which shows morphologically modern facial features appears in Jebel Irhoud,

Morocco approximately 315 ka with the MSA stone tools at the site dating to 374 – 337 ka

(Hublin et al., 2017; Richter et al., 2017; Stringer and Galway-Witham, 2017). Northern Africa experienced two periods of high variability between 380 – 390 ka as well as 350 – 340 ka, which may suggest that both the evolution of AMH morphologies as well as the development of MSA technology documented at Jebel Irhoud at this time were in response to increased climate variability in the region. Similarly, the earliest appearance of MSA at Olorgesailie in eastern

Africa dates to approximately 320 ka, which falls into a period of high climate variability for the region, 330 – 320 ka (Deino et al., 2018; Potts et al., 2018).

By approximately 337 – 300 ka, Levallois technology was widespread across the African continent, following a period from 400 – 320 ka where all five regions experienced at least one 126 period of high climate variability (Richter et al., 2017). Furthermore, fossils with possibly transitional AMH-like features have been found at Kathu Pan, South Africa (~291 ka), Baringo,

Kenya (~284 ka), Gademotta, Ethiopia (~275 ka), Florisbad, South Africa (~260 ka) and

Maghreb (~254 ka) (Hublin et al., 2017; Richter et al., 2017; Stringer and Galway-Witham,

2017). Although these first appearance datum (FAD) times do not fall within the regional periods of variability, the relatively simultaneous appearance of morphological similarities across multiple regions suggests that AMH morphology evolved in response to increased environmental variability.

Genetic evidence suggests that ancient San people in South Africa separated from East

Africans between 260 – 350 ka (Schlebusch et al., 2017). 340 – 330 ka was a period of high variability for Southern Africa, after which the region was relatively stable until 220 ka. This genetic split of two populations may suggest a dispersal from southern Africa into eastern Africa, which was experiencing a period of high variability between 330 – 320 ka.

AMH were most likely established in eastern Africa around 195 ka, with fossil evidence at Herto, Ethiopia confirming full AMH morphology at 160 ka (White et al., 2003; Richter et al.,

2017). However, an archaic hominin fossil sharing affinities with AMH morphology from Lake

Eyasi, dates to 132 – 88 ka, which correlates with periods of high variability in eastern

Africa (90 – 60 ka) and tropical Africa (100 – 90 ka) (Domínguez-Rodrigo et al., 2008). This suggests that there may have been regional persistence archaic hominin populations after AMH was fully established, perhaps facilitated by heterogeneous climate variability and stability patterns across the continent. After the initial appearance of AMH morphology and MSA technology, most of the regions in this study experienced extended periods of relative climate stability. During this period, between 160 ka and 110 ka, there is evidence for the predecessors 127 of San populations diverging from most other African populations (Henn et al., 2018 and references therein).

The dispersal of AMH out of Africa is one of the more hotly debated topics in paleoanthropology. Genetic studies indicate an African origin for all extant populations outside of Africa, but the timing of the dispersals is still unclear (Reyes-Centeno, 2016). Genetic clocks indicate a relatively recent dispersal of the ancestors of modern AMH at approximately 65 ka; however, there are multiple lines of evidence suggesting much earlier occupations of the Levant, southern Europe, Arabia, India and Southeast Asia. The earliest AMH fossils found outside of

Africa are a fossilized human cranium from Apidima Cave, Greece dated to 210 ka and a maxilla from Mislia Cave, Israel dated to 194 – 177 ka (Hershkovitz et al., 2018; Harvati et al., 2019).

Both of these sites support genetic evidence for a dispersal out of Africa approximately 220 ka

(Posth et al., 2017). The timing of this dispersal correlates to a period of high variability in the southern Europe and southwestern Asia (240 – 230 ka) determined by this study, while both north Africa and the southern Levant and eastern Africa were experiencing relative climate stability.

Archaeological evidence can place AMH in Jebel Faya, UAE approximately 125 ka, and

Nubian complex sites also show that by 106 ka AMH were in Oman (Armitage et al., 2011; Rose et al., 2011). The AMH sites of Skhul and Qafzeh in Israel date from 115 – 71 ka, with aDNA overlap restraining the two sites to an occupation between 95 – 90 ka (Grove, 2015). The earlier expansion during MIS 5 was most likely facilitated by optimal climate conditions and lower sea level allowing for passage through the Sahara or along the southern coast of Arabia (Armitage et al., 2011; Larrasoaña et al., 2013). Our study indicates that northern Africa and the southern

Levant experienced an extended period of high climate variability between 130 – 110 ka 128 followed by a period of relatively stable conditions. This earlier dispersal out of Africa by AMH would then be consistent with both the VSH and the APH models, according to our data analysis.

Evidence is accumulating showing that the 65-ka dispersal was one of several such events that had occurred since the late-Middle Pleistocene dispersals. Both fossil and archaeological data support multiple dispersals out of Africa, with the later one being simultaneous through the Levant and Eurasia (Reyes-centeno, 2016). As both eastern Africa, 90

– 60 ka, and northern Africa and the southern Levant, 70 – 60 ka, experienced periods of high variability before this final push out of Africa, it may suggest that this later wave of dispersal was also influenced by the variability of climate in the population source region. Furthermore, archaeological evidence shows an increase in hominin occupations sites in eastern Africa during

MIS 4 and 3 (70 – 60 ka) (Will et al., 2019).

More recent human migrations such as the Arabian backflow into Africa between 12 ka –

4 ka and the Bantu expansion out of tropical Africa into southern and eastern Africa starting at 5 ka may have followed a similar pattern of climate variability driving human movement

(Fadhlaoui-Zid et al., 2013; Grollemund et al., 2015). As the continent of Africa unevenly transitioned into and out of the African Humid Period, variability increased and habitats became fragmented which may have encouraged populations to move (Shanahan et al., 2015). Our study indicates that tropical Africa experienced high climate variability between 20 – 10 ka, while southern and eastern Africa experienced higher variability between 10 – 0 ka. The expansion of the Bantu into the new environments fits well then in the framework of the APH.

Although direct comparisons between human evolutionary events and climate variability may be hampered by age model constraints in both fields allowing for diverse interpretations, our study shows that a more regional and temporally restrained assessment of climate variability 129 can offer a more nuanced framework for testing evolutionary hypotheses such as APH and VSH.

When our variability records are compared to known hominin speciation and technological advances, such as the onset of the MSA, we can see a pattern where increased climate variability may be influencing adaptive strategies. Furthermore, changes from higher to lower variable environments may have been an important driving factor for divergence and dispersals of AMH out of (and at least once back into) Africa.

5. Conclusions

We developed two new metrics for assessing environmental variability over long paleoenvironmental times series and applied these in an analysis of records covering the last 400 ka in and around Africa. The Standard Deviation Variability Method (SDV) uses the variance in data points in any given bin to define climate variability, with bins having a higher standard deviation representing periods of higher climate variability. This method is useful for paleoenvironmental records which have relatively high resolution and evenly spaced data, and is more easily applied as the calculations are relatively simple. The second method, the

Accumulated Weighted Rate of Change Variability Method (AWRCV), uses the accumulated change, or slope, between sequential data points to determine climate variability. The more change occurring within any given bin will give a higher AWRCV value and therefore point to higher climate variability. This method may be more useful in records which have unevenly spaced data or lack enough data points to determine statistically meaningful standard deviations.

In evenly spaced and high-resolution data sets, both the SDV and AWRCV produce proportionally similar results suggesting that both are valuable tools in the assessment of climate variability through time. 130

We applied these methods to analyzing 31 paleoenvironmental reconstructions from 23 separate locations. We further subdivided our analysis regionally (southern Europe and southwestern Asia, northern Africa and the southern Levant, eastern Africa, tropical Africa, and southern Africa) to explore patterns of regional heterogeneity in environmental variability. Our analysis indicates that the five regions of this study experienced periods of both high variability and low variability. Periods of high climate variability identified in this study correspond well to known vegetation, temperature and precipitation changes in each of the regions indicating that both the SDV and AWRCV metrics for set time bins are capable of identifying specific elements of variability at a regional or landscape level that cannot be encapsulated by a global index of variability such as solar insolation by itself. The climate of each region in the study has responded differently to known forcing mechanisms such as glacial expansion, insolation and sea surface temperature, which would have enhanced the habitat fragmentation experienced by early

AMH. When our record is compared to known evolutionary milestones such as the onset of the

MSA, it indicates that climate variability could have been a key driving factor behind technological advances and adaptive strategies allowing for new environments to be inhabited by

AMH and our near relatives. Furthermore, divergences in populations as well as the MIS 5 and

65 ka dispersals out of Africa seem to have occurred as the environment in the source region switched from high to low environmental variability conditions.

Acknowledgements

The authors would like to thank all of our fellow scientists who enthusiastically engaged us in conversation regarding our methods and applications, specifically James Russell, Christopher

Campisano, Henry Lamb, Jonathan Dean, Jose Joordens, Kevin Uno, Mona Stockhecke 131

Christopher Lepre, Christian Tryon, Jay Quade, Peter Reinthal and Kevin Anchukaitis. This is publication xx of the Hominin Sites and Paleolakes Drilling Project (HSPDP).

References

Abe-Ouchi, A., Saito, F., Kawamura, K., Raymo, M.E., Okuno, J., Takahashi, K., Blatter, H.,

2013. Insolation-driven 100,000-year glacial cycles and hysteresis of ice-sheet volume.

Nature. 500, 190-193.

Adegbie, A.T., Schneider, R.R., Röhl, U., Wefer, G., 2003. Glacial millennial-scale fluctuations

in central African precipitation recorded in terrigenous sediment supply and freshwater

signals offshore Cameroon. Palaeogeography, Palaeoclimatology, Palaeoecology. 197, 323-

333.

Armitage, S.J., Jasim, S.A., Marks, A.E., Parker, A.G., Usik, V.I., Uerpmann, H.P., 2011. The

southern route “out of Africa”: Evidence for an early expansion of modern humans into

Arabia. Science. 331, 453-456.

Bar-Matthews, M., Marean, C.W., Jacobs, Z., Karkanas, P., Fisher, E.C., Herries, A.I.R., Brown,

K., Williams, H.M., Bernatchez, J., Ayalon, A., Nilssen, P.J., 2010. A high resolution and

continuous isotopic speleothem record of paleoclimate and paleoenvironment from 90 to 53

ka from Pinnacle Point on the south coast of South Africa. Quaternary Science Reviews. 29,

2131–2145.

Barnosky, A.D., 2006. Distinguishing the effects of the Red queen and Court Jester on Miocene

mammal evolution in the northern Rocky Mountains. Journal of Vertebrate Paleontology.

21, 172-185.

Belmaker, M., Hovers, E., 2011. Ecological change and the extinction of the Levantine

Neanderthals: Implications from a diachronic study of micromammals from Amud Cave, 132

Israel. Quaternary Science Reviews. 30, 3196-3209.

Benton, M.J., 2010. New take on the Red Queen. Nature. 463, 306-307.

Blome, M.W., Cohen, A.S., Tryon, C.A., Brooks, A.S., Russell, J., 2012a. The environmental

context for the origins of modern human diversity: A synthesis of regional variability in

African climate 150,000-30,000 years ago. Journal of Human Evolution. 62, 563–592.

Blumenthal, S.A., Levin, N.E., Brown, F.H., Brugal, J.-P., Chritz, K.L., Harris, J.M., Jehle, G.E.,

Cerling, T.E., 2017. Aridity and hominin environments. Proceedings of the National

Academy of Sciences. 114, 7331-7336.

Botai, C.M., Botai, J.O., Adeola, A.M. 2018. Spatial distribution of temporal precipitation

conatrasts in South Africa. South African Journal of Science. 114, 1-9,

dx.doi.org/10.17159/sajs.2018/20170391.

Burnett, A.P., Soreghan, M.J., Scholz, C.A., Brown, E.T., 2011. Tropical East African climate

change and its relation to global climate: A record from , Tropical East

Africa, over the past 90+kyr. Palaeogeography, Palaeoclimatology, Palaeoecology. 303,

155–167.

Burrough, S.L., Thomas, D.S.G., 2009. Geomorphological contributions to palaeolimnology on

the African continent. Geomorphology. 103, 285-298.

Caley, T., Extier, T., Collins, J.A., Schefuß, E., Dupont, L., Malaizé, B., Rossignol, L., Souron,

A., McClymont, E.L., Jimenez-Espejo, F.J., García-Comas, C., Eynaud, F., Martinez, P.,

Roche, D.M., Jorry, S.J., Charlier, K., Wary, M., Gourves, P.Y., Billy, I., Giraudeau, J.,

2018. A two-million-year-long hydroclimatic context for hominin evolution in southeastern

Africa. Nature. 560, 76-79.

Campisano, C.J., Feibel, C.S., 2008. Depositional environments and stratigraphic summary of 133

the Pliocene Hadar Formation at Hadar, Afar Depression, Ethiopia. The Geology of Early

Humans in the . 252, 179–201.

Castañeda, I.S., Caley, T., Dupont, L., Kim, J.-H., Malaizé, B., Schouten, S., 2016. Middle to

Late Pleistocene vegetation and climate change in subtropical southern East Africa. Earth

and Planetary Science Letters. 450, 306–316.

Cohen, A.S., Stone, J.R., Beuning, K.R.M., Park, L.E., Reinthal, P.N., Dettman, D., Scholz,

C.A., Johnson, T.C., King, J.W., Talbot, M.R., Brown, E.T., Ivory, S.J., 2007. Ecological

consequences of early Late Pleistocene megadroughts in tropical Africa. Proceedings of the

National Academy of Sciences. 104, 16422-16427.

Cohen, A.S., Billingsley, A.L., Campisano, C., Lupien, R., Muiruri, V., Owen, R.B., Russell, J.,

Stockhecke, M., Yost, C. 2018. Records of Plio-Pleistocene environmental variablity from

the Hominin Sties and Paleolakes Drilling Project (HSPDP): Comparsions with other

African records and implications for hominin evolution. Abstract 349645 prsending at the

2018 Fall Meeting, AGU, Washington, D.C. Dec 10 -14.

Cortina, A., Sierro, F.J., González-Mora, B., Asioli, A., Flores, J.A., 2011. Impact of climate and

sea level changes on the ventilation of intermediate water and benthic foraminifer

assemblages in the Gulf of Lions, off South France, during MIS 6 and 7. Palaeogeography,

Palaeoclimatology, Palaeoecology. 309, 215–228.

Costa, K., Russell, J., Konecky, B., Lamb, H., 2014. Isotopic reconstruction of the African

Humid Period and Congo Air Boundary migration at Lake Tana, Ethiopia. Quaternary

Science Reviews. 83, 58–67.

Deino, A.L., Behrensmeyer, A.K., Brooks, A.S., Yellen, J.E., Sharp, W.D., Potts, R., 2018.

Chronology of the Acheulean to Middle Stone Age transition in eastern Africa. Science. 134

360, 95-98.

Deino, A.L., Kingston, J.D., Glen, J.M., Edgar, R.K., Hill, A., 2006. Precessional forcing of

lacustrine sedimentation in the late Cenozoic Chemeron Basin, Central Kenya Rift, and

calibration of the Gauss/Matuyama boundary. Earth and Planetary Science Letters. 247, 41–

60. deMenocal, P., 1995. Plo-Pleistocene African Climate. Science. 270, 53–59. deMenocal, P., Ortiz, J., Guilderson, T., Adkins, J., Sarnthein, M., Baker, L., Yarusinsky, M.,

2000. Abrupt onset and termination of the African Humid Period: Rapid climate responses

to gradual insolation forcing. Quaternary Science Reviews. 19, 347-361. deMenocal, P.B., 2004. African climate change and faunal evolution during the Pliocene-

Pleistocene. Earth and Planetary Science Letters. 220, 3–24.

Domínguez-Rodrigo, M., Mabulla, A., Luque, L., Thompson, J.W., Rink, J., Bushozi, P., Díez-

Martin, F., Alcala, L., 2008. A new archaic Homo sapiens fossil from , Tanzania.

Journal of Human Evolution. 54, 899–903.

Drake, N.A., Blench, R.M., Armitage, S.J., Bristow, C.S., White, K.H., 2011. Ancient

watercourses and biogeography of the Sahara explain the peopling of the desert.

Proceedings of the National Academy of Sciences. 108, 458-462.

Dupont, L., 2011. Orbital scale vegetation change in Africa. Quaternary Science Reviews. 30,

3589–3602.

Dupont, L.M., Donner, B., Schneider, R., Wefer, G., 2001. Mid-Pleistocene environmental

change in tropical Africa began as early as 1.05 Ma. Geology. 29, 195–198.

Dupont, L.M., Linder, H.P., Rommerskirchen, F., Schefuß, E., 2011. Climate-driven rampant

speciation of the Cape flora. Journal of Biogeography. 38, 1059-1068. 135

Fadhlaoui-Zid, K., Haber, M., Martínez-Cruz, B., Zalloua, P., Elgaaied, A.B., Comas, D., 2013.

Genome-wide and paternal diversity reveal a recent origin of human populations in north

africa. PLoS ONE. 8: e80293. doi:10.1371/journal.pone.0080293.

Francke, A., Wagner, B., Just, J., Leicher, N., Gromig, R., Baumgarten, H., Vogel, H., Lacey,

J.H., Sadori, L., Wonik, T., Leng, M.J., Zanchetta, G., Sulpizio, R., Giaccio, B., 2016a.

Sedimentological processes and environmental variability at Lake Ohrid (Macedonia,

Albania) between 637 ka and the present. Biogeosciences. 13, 1179–1196.

Gasse, F., Vidal, L., Van Campo, E., Demory, F., Develle, A.L., Tachikawa, K., Elias, A., Bard,

E., Garcia, M., Sonzogni, C., Thouveny, N., 2015. Hydroclimatic changes in northern

Levant over the past 400,000 years. Quaternary Science Reviews. 111, 1–8.

Grant, K.M., Rohling, E.J., Westerhold, T., Zabel, M., Heslop, D., Konijnendijk, T., Lourens, L.,

2017. A 3 million year index for North African humidity/aridity and the implication of

potential pan-African Humid periods. Quaternary Science Reviews. 171, 100-118.

Grollemund, R., Branford, S., Bostoen, K., Meade, A., Venditti, C., Pagel, M., 2015. Bantu

expansion shows that habitat alters the route and pace of human dispersals. Proceedings of

the National Academy of Sciences. 112, 13296-13301.

Grove, M., 2014. Evolution and dispersal under climatic instability: A simple evolutionary

algorithm. Adaptive Behavior. 22, 235-254.

Grove, M., 2015. Palaeoclimates , plasticity , and the early dispersal of Homo sapiens.

Quaternary International. 369, 17–37.

Grove, M., Burke, A., 2015. Environmental variability and hominin dispersal. Journal of Human

Evolution. 87, 1–4.

Harvati, K., Roding, C., Bosman, A.M., Karakostis, F.A., Rainer, G., Stringser, C., Karkanas, P., 136

Thompson, N.C., Koutoulidis, V., Moulopoulos, L.A., Gorgoulis, V.G., Kouloukoussa, M.,

2019. Apidima cave fossils provide earliest evidence of Homo sapiens in Eurasia. Nature.

571, 500 - 504. doi.org/1.1038/s41586-019-1376-z.

Helmens, K.F., 2014. The last interglacial-glacial cycle (MIS 5-2) re-examined based on long

proxy records from central and . Quaternary Science Reviews. 86, 115–143.

Helmke, J.P., Bauch, H.A., Röhl, U., Kandiano, E.S., 2008. Uniform climate development

between the subtropical and subpolar Northeast Atlantic across marine isotope stage 11.

Climate of the Past. 4, 181-190.

Henn, B.M., Steele, T.E., Weaver, T.D., 2018. Clarifying distinct models of modern human

origins in Africa. Current Opinion in Genetics and Development. 53, 148–156.

Hershkovitz, I., Weber, G.W., Quam, R., Duval, M., Grün, R., Kinsley, L., Ayalon, A., Bar-

Matthews, M., Valladas, H., Mercier, N., Arsuaga, J.L., Martinón-Torres, M., De Castro,

J.M.B., Fornai, C., Martín-Francés, L., Sarig, R., May, H., Krenn, V.A., Slon, V.,

Rodríguez, L., García, R., Lorenzo, C., Carretero, J.M., Frumkin, A., Shahack-Gross, R.,

Mayer, D.E.B.Y., Cui, Y., Wu, X., Peled, N., Groman-Yaroslavski, I., Weissbrod, L.,

Yeshurun, R., Tsatskin, A., Zaidner, Y., Weinstein-Evron, M., 2018. The earliest modern

humans outside Africa. Science.359, 456-459.

Hublin, J.J., Ben-Ncer, A., Bailey, S.E., Freidline, S.E., Neubauer, S., Skinner, M.M., Bergmann,

I., Le Cabec, A., Benazzi, S., Harvati, K., Gunz, P., 2017. New fossils from Jebel Irhoud,

Morocco and the pan-African origin of Homo sapiens. Nature. 546, 289-292.

Huybers, P., Wunsch, C., 2005. Obliquity pacing of the late Pleistocene glacial terminations.

Nature. 434, 491-494.

Ivory, S.J., Blome, M.W., King, J.W., McGlue, M.M., Cole, J.E., Cohen, A.S., 2016. 137

Environmental change explains cichlid adaptive radiation at Lake Malawi over the past 1.2

million years. Proceedings of the National Academy of Sciences. 113, 11895–11900.

Ivory, S.J., Lézine, A.M., Vincens, A., Cohen, A.S., 2018. Waxing and waning of forests: Late

Quaternary biogeography of southeast Africa. Global Change Biology. 24, 2939–2951.

Kandiano, E.S., Bauch, H.A., Fahl, K., Helmke, J.P., Röhl, U., Pérez-Folgado, M., Cacho, I.,

2012. The meridional temperature gradient in the eastern North Atlantic during MIS 11 and

its link to the ocean-atmosphere system. Palaeogeography, Palaeoclimatology,

Palaeoecology. 333–334, 24–39.

Kingston, J.D., Deino, A.L., Edgar, R.K., Hill, A., 2007a. Astronomically forced climate change

in the Kenyan Rift Valley 2.7-2.55 Ma: implications for the evolution of early hominin

ecosystems. Journal of Human Evolution. 53, 487–503.

Kingston, J.D., 2007b. Shifting adaptive landscapes: Progress and challenges in reconstructing

early hominid environments. Yearbook of Physical Anthropology. 134, 20–58.

Kostygov, S.A., Kandiano, E.S., Bauch, H.A., 2010. Reconstruction of deep-water conditions in

the North Atlantic during MIS 9 based on benthic foraminiferal assemblages. Oceanology.

50, 397–407.

Kousis, I., Koutsodendris, A., Peyron, O., Leicher, N., 2018. Centennial-scale vegetation

dynamics and climate variability in SE Europe during Marine Isotope Stage 11 based on a

pollen record from Lake Ohrid. 190, 20–38.

Kutzbach, J.E., Liu, Z., 1997. Response of the African monsoon to orbital forcing and ocean

feedbacks in the middle Holocene. Science. 278, 440-443.

Kutzbach, J.E., Otto-Bliesner, B.L. 1982. The sensitivity of hte African-Asian monsoonal

climate to orbital parameter changes for 9000 years B.P. in a low-resolution general 138

circulation model. Journal of the Atmospheric Sciences. 39, 1177-1188.

Kwiecien, O., Stockhecke, M., Pickarski, N., Heumann, G., Litt, T., Sturm, M., Anselmetti, F.,

Kipfer, R., Haug, G.H., 2014. Dynamics of the last four glacial terminations recorded in

Lake Van, Turkey. Quaternary Science Reviews. 104, 42–52.

Larrasoaña, J.C., Roberts, A.P., Rohling, E.J., 2013. Dynamics of Green Sahara Periods and

Their Role in Hominin Evolution. PLoS ONE. 8: e76514.doi:101371/journal.pone.0076514.

Lezine, A.M., 1991. West African paleoclimates during the last climatic cycle inferred from an

Atlantic deep-sea pollen record. Quaternary Research. 35, 456-463.

Lisiecki, L.E., Raymo, M.E., 2005. A Pliocene-Pleistocene stack of 57 globally distributed

benthic δ 18O records. Paleoceanography. 20, 1–17.

López-García, J.M., Blain, H.A., Julià, R., Maroto, J., 2014. Environment and climate during

MIS 7 and their implications for the late Middle Pleistocene hominins: The contribution of

Mollet cave, Serinyà, Girona, northeastern Iberian Peninsula. Quaternary International. 337,

4–10.

Lyons, R.P., Scholz, C.A., Cohen, A.S., King, J.W., Brown, E.T., Ivory, S.J., Johnson, T.C.,

Deino, A.L., Reinthal, P.N., McGlue, M.M., Blome, M.W., 2015. Continuous 1.3-million-

year record of East African hydroclimate, and implications for patterns of evolution and

. Proceedings of the National Academy of Sciences. 112, 15568-15573.

Maslin, M.A., Shultz, S., Trauth, M.H., 2015. A synthesis of the theories and concepts of early

human evolution. Philosophical Transactions of the Royal Society B: Biological Sciences.

370, 20140064–20140064.

Maslin, M.A., Trauth, M.H., 2009. Plio-Pleistocene East African Pulsed Climate Variability and

Its Influence on Early Human Evolution. In: The First Humans – Origin and Early 139

Evolution of the Genus Homo. pp. 151–158.

Nicholson, S.E., 2000. The nature of rainfall variability over Africa on time scales of decades to

millenia. Global and Planetary Change. 26, 137–158.

Nicholson, S.E., 2018. The ITCZ and the seasonal cycle over equatorial Africa. Bulletin of the

American Meteorological Society. 99, 337-348.

Ning, S., Dupont, L.M., 1997. Vegetation and climatic history of southwest Africa: A marine

palynological record of the last 300,000 years. Vegetation History and Archaeobotany. 6,

117-131.

Owen, R.B., Muiruri, V.M., Lowenstein, T.K., Renaut, R.W., Rabideaux, N., Luo, S., Deino,

A.L., Sier, M.J., Dupont-Nivet, G., McNulty, E.P., Leet, K., Cohen, A., Campisano, C.,

Deocampo, D., Shen, C.-C., Billingsley, A., Mbuthia, A., 2018. Progressive aridification in

East Africa over the last half million years and implications for human evolution.

Proceedings of the National Academy of Sciences. 115, 11174-11179.

Posth, C., Wißing, C., Kitagawa, K., Pagani, L., Van Holstein, L., Racimo, F., Wehrberger, K.,

Conard, N.J., Kind, C.J., Bocherens, H., Krause, J., 2017. Deeply divergent archaic

mitochondrial genome provides lower time boundary for African gene flow into

Neanderthals. Nature Communications. 8, 16046. doi:10.1038/ncomms16046.

Potts, R., 1996. Evolution and climate variability. Science. 273, 922-923.

Potts, R., 1998. Variability selection in hominid evolution. Evolutionary Anthropology.7, 81-96.

Potts, R., 2013. Hominin evolution in settings of strong environmental variability. Quaternary

Science Reviews. 73, 1-13.

Potts, R., Behrensmeyer, A.K., Tyler Faith, J., Tryon, C.A., Brooks, A.S., Yellen, J.E., Deino,

A.L., Kinyanjui, R., Clark, J.B., Haradon, C.M., Levin, N.E., Meijer, H.J.M., Veatch, E.G., 140

Bernhart Owen, R., Renaut, R.W., 2018. Environmental dynamics during the onset of the

Middle Stone Age in eastern Africa. Science. 360, 86-90.

Potts, R., Faith, J.T., 2015. Alternating high and low climate variability: The context of natural

selection and speciation in Plio-Pleistocene hominin evolution. Journal of Human

Evolution. 87, 5–20.

Quade, J., Dente, E., Armon, M., Ben Dor, Y., Morin, E., Adam, O., Enzel, Y., 2018. Megalakes

in the Sahara? A Review. Quaternary Research. 90, 253-275. doi:10.1017//qua.2018.46.

Reyes-Centeno, H., 2016. Out of Africa and into Asia : Fossil and genetic evidence on modern

human origins and dispersals. Quaternary International. 416, 249–262.

Richter, D., Grün, R., Joannes-Boyau, R., Steele, T.E., Amani, F., Rué, M., Fernandes, P.,

Raynal, J.P., Geraads, D., Ben-Ncer, A., Hublin, J.J., McPherron, S.P., 2017. The age of the

hominin fossils from Jebel Irhoud, Morocco, and the origins of the Middle Stone Age.

Nature. 546, 293-296.

Rohling, E.J., Grant, K.M., Roberts, A.P., Larrasoaña, J.-C., 2013. Paleoclimate Variability in

the Mediterranean and Red Sea Regions during the Last 500,000 Years. Current

Anthropology. 54, S183-S201.

Rose, J.I., Usik, V.I., Marks, A.E., Hilbert, Y.H., Galletti, C.S., Parton, A., Geiling, J.M., Černý,

V., Morley, M.W., Roberts, R.G., 2011. The Nubian complex of Dhofar, Oman: An African

Middle Stone Age industry in Southern Arabia. PLoS ONE. 6, e28239.

doi:10.1371/journal.pone.0028239.

Roucoux, K.H., Tzedakis, P.C., de Abreu, L., Shackleton, N.J., 2006. Climate and vegetation

changes 180,000 to 345,000 years ago recorded in a deep-sea core off Portugal. Earth and

Planetary Science Letters. 249, 307–325. 141

Ruddiman, W.F., 2006. What is the timing of orbital-scale monsoon changes? Quaternary

Science Reviews. 25, 657-658.

Scerri, E.M.L., Drake, N.A., Jennings, R., Groucutt, H.S., 2014. Earliest evidence for the

structure of Homo sapiens populations in Africa. Quaternary Science Reviews. 101, 207-

216.

Scerri, E.M.L., Thomas, M.G., Manica, A., Gunz, P., Stock, J.T., Stringer, C., Grove, M.,

Groucutt, H.S., Timmermann, A., Rightmire, G.P., d’Errico, F., Tryon, C.A., Drake, N.A.,

Brooks, A.S., Dennell, R.W., Durbin, R., Henn, B.M., Lee-Thorp, J., deMenocal, P.,

Petraglia, M.D., Thompson, J.C., Scally, A., Chikhi, L., 2018. Did Our Species Evolve in

Subdivided Populations across Africa, and Why Does It Matter? Trends in ecology &

evolution. 33, 582–594.

Schlebusch, C.M., Malmström, H., Günther, T., Sjödin, P., Coutinho, A., Edlund, H., Munters,

A.R., Vicente, M., Steyn, M., Soodyall, H., Lombard, M., Jakobsson, M., 2017. Southern

African ancient genomes estimate modern human divergence to 350,000 to 260,000 years

ago. Science. 358, 652-655.

Scholz, C.A., Johnson, T.C., Cohen, A.S., King, J.W., Peck, J.A., Overpeck, J.T., Talbot, M.R.,

Brown, E.T., Kalindekafe, L., Amoako, P.Y.O., Lyons, R.P., Shanahan, T.M., Castaneda,

I.S., Heil, C.W., Forman, S.L., McHargue, L.R., Beuning, K.R., Gomez, J., Pierson, J.,

2007. Eastern African megadroughts between 135 and 75 thousand years ago and bearing

on early-modern human origins. Proceedings of the National Academy of Sciences. 102,

16416–16421.

Scholz, C.A., Cohen, A.S., Johnson, T.C., King, J., Talbot, M.R., Brown, E.T., 2011a. Scientific

drilling in the Great Rift Valley: The 2005 Lake Malawi Scientific Drilling Project - An 142

overview of the past 145,000years of climate variability in Southern Hemisphere East

Africa. Palaeogeography, Palaeoclimatology, Palaeoecology. 303, 3–19.

Scholz, C.A., Talbot, M.R., Brown, E.T., Lyons, R.P., 2011b. Lithostratigraphy, physical

properties and organic matter variability in Lake Malawi Drillcore sediments over the past

145,000years. Palaeogeography, Palaeoclimatology, Palaeoecology. 303, 38-50.

Scott, L., 1999. Vegetation history and climate in the Savanna biome South Africa since 190,000

ka: A comparison of pollen data from the Tswaing Crater (the Pretoria Saltpan) and

Wonderkrater. Quaternary International. 57–58, 215–223.

Shanahan, T.M., Mckay, N.P., Hughen, K.A., Overpeck, J.T., Otto-Bliesner, B., Heil, C.W.,

King, J., Scholz, C.A., Peck, J., 2015. The time-transgressive termination of the African

humid period. Nature Geoscience. 8, 386-301.

Shultz, S., Maslin, M., 2013. Early Human Speciation, Brain Expansion and Dispersal

Influenced by African Climate Pulses. PLoS ONE. 8, e76750. doi:

10.1371/journal.pone.0076750.

Stenseth, N.C., Smith, J.M., 2006. Coevolution in Ecosystems: Red Queen Evolution or Stasis?

Evolution. 38, 870-880.

Stockhecke, M., Sturm, M., Brunner, I., Schmincke, H.U., Sumita, M., Kipfer, R., Cukur, D.,

Kwiecien, O., Anselmetti, F.S., 2014. Sedimentary evolution and environmental history of

Lake Van (Turkey) over the past 600 000 years. Sedimentology. 61, 1830-1861.

Stone, J.R., Westover, K.S., Cohen, A.S., 2011. Late Pleistocene paleohydrography and diatom

paleoecology of the central basin of Lake Malawi, Africa. Palaeogeography,

Palaeoclimatology, Palaeoecology. 303, 51-70.

Stringer, C., Galway-Witham, J., 2017. Palaeoanthropology: On the origin of our species. 143

Nature. 545, 212-214.

Szabo, B.J., Haynes, C. V., Maxwell, T.A., 1995. Ages of Quaternary pluvial episodes

determined by uranium-series and radiocarbon dating of lacustrine deposits of Eastern

Sahara. “Palaeogeography, Palaeoclimatology, Palaeoecology.” 113, 227–242.

Tierney, J.E., deMenocal, P.B., Zander, P.D., 2017a. A climatic context for the out-of-Africa

migration. Geology. 45, 1023-1026.

Tierney, J.E., Pausata, F.S.R., De Menocal, P.B., 2017b. Rainfall regimes of the Green Sahara.

Science Advances. 3: e1601503.

Tierney, J. E., Russell, J.M., Huang, Y., Damste, J.S.S., Hopmans, E.C., Cohen, A.S., 2008.

Northern Hemisphere Controls on Tropical Southeast African Climate During the Past

60,000 Years. Science. 322, 252-255.

Tierney, J.E., Russell, J.M., Sinninghe Damsté, J.S., Huang, Y., Verschuren, D., 2011. Late

Quaternary behavior of the East African monsoon and the importance of the Congo Air

Boundary. Quaternary Science Reviews. 30, 798-807.

Trauth, M.H., Bergner, A.G.N., Foerster, V., Junginger, A., Maslin, M.A., Schaebitz, F., 2015.

Episodes of environmental stability versus instability in Late Cenozoic lake records of

Eastern Africa. Journal of Human Evolution. 87, 21-31.

Trauth, M.H., Larrasoaña, J.C., Mudelsee, M., 2009. Trends, rhythms and events in Plio-

Pleistocene African climate. Quaternary Science Reviews. 28, 399–411.

Tzedakis, P.C., Hooghiemstra, H., Pälike, H., 2006. The last 1.35 million years at Tenaghi

Philippon: revised chronostratigraphy and long-term vegetation trends. Quaternary Science

Reviews. 25, 3416-3430.

Viehberg, F.A., Just, J., Dean, J.R., Wagner, B., Franz, S.O., Klasen, N., Kleinen, T., Ludwig, P., 144

Asrat, A., Lamb, H.F., Leng, M.J., Rethemeyer, J., Milodowski, A.E., Claussen, M.,

Schäbitz, F., 2018. Environmental change during MIS4 and MIS 3 opened corridors in the

Horn of Africa for Homo sapiens expansion. Quaternary Science Reviews. 202, 139-153.

Vrba, E.S., 1985. Ecological and Adaptive Changes Associated With Early Hominid Evolution.

In: Ancestors: The Hard Evidence. Paleobiology. 10, 123-164.

Washburn, S. 1960. Tools and Human Evolution. Scientific American. 3, 62-75.

Weldeab, S., Lea, D.W., Schneider, R.R., Andersen, N., 2007. 155,000 Years of West African

Monsoon and Ocean Thermal Evolution. Science. 316, 1303–1307.

White, T.D., Asfaw, B., DeGusta, D., Gilbert, H., Richards, G.D., Suwa, G., Howell, F.C., 2003.

Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature. 423, 742-747.

Will, M., Conard, N.J., Tryon, C.A., 2019. Timing and trajectory of cultural evolution on the

African continent 200,000-30,000 years ago. In: Modern Human Origins and Dispersal.

Wilson, K.E., Maslin, M.A., Leng, M.J., Kingston, J.D., Deino, A.L., Edgar, R.K., Mackay,

A.W., 2014. East African lake evidence for Pliocene millennial-scale climate variability.

Geology. 42, 955–958.

145

Figures

146

Figure 1: Equatorial insolation for the last 400ka plotted with the results of the analysis. Periods of high variability are represented by both increased SDV and AWRCV. Both metrics demonstrate a correspondence of these periods with the largest changes in insolation.

147

Figure 2: Locations of individual records used in this study with the type of indicator used for the analysis. Geographic subdivisions used for regional records are indicated. More information regarding each record is in Table 1. 148

Figure 3 The mean AWRCV for each region are plotted with confidence bands and 1 sigma above mean for each region. Periods where the AWRCV are outside of the 1 sigma boundary are considered periods of high climate variability.

149

Figure 4: 400 ka AWRCV record for potential forcing mechanisms with periods of high regional climate variability highlighted. Colors represent the periods of high environmental variability in each region determined by the AWRCV method. 150

Figure 5: Comparison between the original time series and variability metrics for δ18O foraminifera benthic stack (Lisiecki and Raymo, 2005) and the ODP721/722 dust flux

(deMenocal, 1995) for three bin durations (10, 20 and 50 kyr) indicating how bin size may influence calculations of variability and correspondence of variability among records. 151

Figure 6: Diagram showing where and how climate variability may influence selection processes on human populations. Periods highlighted in red represent an instable environment, when the environmental record is going through the most change. These periods would have the most potential for selection of plasticity. In contrast, the periods highlighted in yellow would have the least environmental change and therefore would most likely select towards specialization or fitness. 152

Figure 7: Map of fossil and archaeological sites mentioned in the text along with the timing of

AMH dispersals. 153

Tables

Table 1: List of all of the paleoenvironmental records used in this analysis.

154

Table 2: Results of the AWRCV analysis for each region.

Table 3 Correlations between records and potential forcing mechanisms. Red indicates significant correlation between regions, green significant correlation between region and forcing mechanism, blue significant correlations between forcing mechanisms.