1 A MULTIDISCIPLINARY APPROACH TO LATE QUATERNARY PALEOCLIMATOLOGY WITH AN EMPHASIS ON SUB-SAHARAN WEST AFRICA AND THE LAST INTERGLACIAL PERIOD by Nicholas Paul McKay ___________________________ 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 2012 3 STATEMENT BY AUTHOR This dissertation has been submitted in partial fulfillment of requirements for an advanced degree at the University of Arizona and is deposited in the University Library to be made available to borrowers under rules of the Library. Brief quotations from this dissertation are allowable without special permission, provided that accurate acknowledgment of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgment the proposed use of the material is in the interests of scholarship. In all other instances, however, permission must be obtained from the author. SIGNED: Nicholas P. McKay 4 ACKNOWLEDGEMENTS This research was supported by several funding sources, including National Science Foundation Grant for Earth Sciences grant 0602355, an Institute for the Study of Planet Earth Fellowship, an Institute of the Environment Dissertation Improvement Grant, Dr. H. Wesley Pierce Scholarship, and a Chevron-Texaco Summer Research Grant. The sediment drilling was supported by the International Continental Scientific Drilling Program, and the U.S., Austrian, and Canadian National Science Foundations as well as the Ghanaian Geological Survey. I am greatly indebted to the drilling team and Lake Bosumtwi project collaborators, John Peck, Jonathan Overpeck, Tim Shanahan, Chip Heil, John King and Chris Scholz, who designed and carried out the drilling program, and who have spent countless hours processing, and analyzing the sediment. Without their efforts none of my research on Lake Bosumtwi would have been possible. Furthermore, I am thankful for access to laboratory equipment, sample analysis and insight from Erik Brown, David Dettman, Bob Downs, and Heidi Barnett. Discussions with my advisor, Jonathan Overpeck, with my committee members Julia Cole, Andy Cohen, Erik Brown, and Joellen Russell, as well as my collaborators Bette Otto-Bliesner, John Peck, Tim Shanahan, Chip Heil, John King, Chris Scholz, David Dettman, and Bob Downs greatly improved my research. I am especially grateful for the support and contributions of past and present graduate student colleagues Cody Routson, Jeremy Weiss, Jessica Conroy, Sarah White, Sarah Truebe, Toby Ault, Diane Thompson and Suz Tolwinski-Ward. 5 DEDICATION I dedicate this dissertation to my wife, Amber, who has always supported me and my decision to pursue this research and degree, even during times when my obligations required her to manage our household and raise our children without much help from me. Without her selflessness I would have never completed this dissertation, and she doesn't get enough credit for all the things she does. I also dedicate this dissertation to my boys, Carter, Ethan and Landon, who throughout the past five years have helped me to remember what's important in life, and for the most part, to keep my priorities straight. 6 TABLE OF CONTENTS LIST OF FIGURES...........................................................................................................10 ABSTRACT.......................................................................................................................11 1. INTRODUCTION.........................................................................................................13 2. PRESENT STUDY........................................................................................................21 3. REFERENCES..............................................................................................................27 4. APPENDICES...............................................................................................................32 APPENDIX A: A HIGH-RESOLUTION PALEOHYDROLOGICAL RECORD FOR THE PAST 530,000 YEARS FROM LAKE BOSUMTWI, SOUTHERN GHANA........33 Abstract..............................................................................................................................35 1. Introduction...................................................................................................................36 1.1 Setting......................................................................................................................38 1.2 Regional Climatology..............................................................................................39 1.3 Past work on Lake Bosumtwi..................................................................................41 2. Methods.........................................................................................................................43 2.1 Lake Drilling and core processing...........................................................................43 2.2 Geochronology........................................................................................................44 2.3 Scanning X-Ray fluorescence analysis....................................................................45 2.4 Carbonate isotope geochemistry..............................................................................46 2.5 Turbidites.................................................................................................................48 2.6 Laminations ............................................................................................................49 3. Results...........................................................................................................................49 3.1 Geochronology........................................................................................................49 3.1.1 Comparisons with regional and global dust record..........................................51 3.1.2 Tuning to EDC timescale..................................................................................53 3.2 X-ray fluorescence...................................................................................................54 3.2.1 Correcting for biases in the XRF data..............................................................55 3.2.1.1 X-ray absorption and attenuation by water................................................56 3.2.1.2 Sediment compaction with depth..............................................................58 3.2.2 Principal component analysis...........................................................................59 3.2.2.1 Carbonate dilution of XRF-PC1................................................................60 3.3 Lamination indices..................................................................................................61 3.4 Turbidites.................................................................................................................62 7 4. Discussion......................................................................................................................63 4.1 Past 530 ka...............................................................................................................63 4.1.1 400-kyr eccentricity forcing.............................................................................63 4.1.2 100-kyr eccentricity forcing and glacial boundary conditions.........................67 4.1.3 West African contribution to global atmospheric methane concentrations......67 4.2 Latitudinal variability in West African climate over the past 150 ka......................71 4.3 Millenial-scale events during the past 130 ka..........................................................73 4.4 Intervals of extreme aridity......................................................................................75 5. Conclusion and summary..............................................................................................77 6. References.....................................................................................................................80 APPENDIX B: A 12,000-YEAR-LONG, ANNUALLY-RESOLVED VARVE RECORD SPANNING THE LAST INTERGLACIAL FROM LAKE BOSUMTWI, SOUTHERN GHANA...........................................................................................................................117 Abstract............................................................................................................................119 1. Introduction.................................................................................................................120 1.1 Site.........................................................................................................................124 1.2 Modern Climate.....................................................................................................125 2. Methods.......................................................................................................................127 2.1 Core Drilling and sampling...................................................................................127 2.2 Sampling and laminae analysis..............................................................................128 2.3 XRF analysis..........................................................................................................128 2.4 Development of the varve chronology..................................................................129 2.5 Geochronology......................................................................................................132 2.6 Carbonate isotope geochemistry............................................................................135 3. Results.........................................................................................................................136 3.1 Generalized Lithology and Laminae structure......................................................136 3.1.1 Unit 1a (60.3 – 60.7 RMCD) .........................................................................137 3.1.2 Unit 1b (60.2 – 60.3 RMCD)..........................................................................137 3.1.3 Unit 1c (59.8 – 60.2 RMCD) .........................................................................137 3.1.4 Unit 1d (59.6 – 59.8 RMCD)..........................................................................138 3.1.5 Unit 2a (59.1 – 59.6 RMCD)..........................................................................138 3.1.6 Unit 2b (58.7 – 59.1 RMCD)..........................................................................139 3.1.7 Unit 2c (58.6 - 58.7 RMCD)...........................................................................139 3.1.8 Unit 2d (58.4 - 58.6 RMCD)..........................................................................140 3.1.9 Unit 3a (58.1 - 58.4 RMCD)...........................................................................140 3.1.10 Unit 3b (57.4 - 58.1 RMCD)........................................................................140 8 3.2 Carbonate isotope geochemistry and mineralogy..................................................141 3.3 Varve thickness......................................................................................................145 3.4 XRF data................................................................................................................146 4. Discussion....................................................................................................................148 4.1 Hydrology during the interglacial..........................................................................148 4.1.1 Early LIG (Varve years 1-5600; 128.6 – 123 ka)...........................................149 4.1.2 Mid-LIG (Varve years 5600-9000; 123 – 119.6 ka).......................................150 4.1.3 Late LIG (Varve years 9000-12100; 119.6 – 116.5 ka)..................................151 4.2 Comparison between the LIG and the Holocene...................................................152 4.2.1 Relationship with solar insolation..................................................................154 4.3 Relationships with sea surface temperature records..............................................156 4.4 An multi-century aridity event 118 ka...................................................................158 4.5 Drought frequency during the LIG........................................................................159 4.5.1 Influence of solar variability...........................................................................163 5. Conclusions and summary...........................................................................................165 5.1 Primary findings....................................................................................................165 5.2 Implications for future hydrology in the region....................................................168 6. References...................................................................................................................169 APPENDIX C: THE ROLE OF OCEAN THERMAL EXPANSION IN LAST INTERGLACIAL SEA LEVEL RISE.............................................................................198 Abstract............................................................................................................................200 Introduction......................................................................................................................200 Methods............................................................................................................................202 Paleoclimate data.........................................................................................................202 Global Climate Model Simulations.............................................................................203 Paleoceanographic data synthesis....................................................................................204 Global Climate Model Simulations.................................................................................207 The thermosteric component of LIG sea level rise..........................................................210 Conclusion.......................................................................................................................213 Acknowledgements..........................................................................................................213 9 References........................................................................................................................214 Auxiliary Material............................................................................................................217 Error Analysis..............................................................................................................217 Bias Analysis................................................................................................................218 130 ka climate model simulation.................................................................................219 Auxiliary References....................................................................................................220 APPENDIX D: ON THE POTENTIAL OF RAMAN-SPECTROSCOPY-BASED CARBONATE MASS SPECTROMETRY.....................................................................234 Abstract............................................................................................................................236 Introduction......................................................................................................................236 Experimental Procedure...................................................................................................238 Calcite synthesis...........................................................................................................238 Raman Spectroscopy....................................................................................................239 Isotope notation............................................................................................................240 Water Mass Spectrometry............................................................................................241 Carbonate Mass Spectrometry.....................................................................................242 Results and Discussion....................................................................................................243 Raman-inferred isotope ratios (RRaman).........................................................................243 Potential sources of error and bias...............................................................................244 Potential for improving the accuracy and precision of RRaman estimates......................245 Carbon isotopes............................................................................................................246 Potential for quantifying isotope ratios in other molecules.........................................247 Conclusions......................................................................................................................248 References........................................................................................................................249 APPENDIX E: PERMISSIONS......................................................................................258 10 LIST OF FIGURES Figure 1, Map of Lake Bosumtwi......................................................................................26 11 ABSTRACT A primary goal of paleoclimatology is to extend the instrumental record to capture a wider range of natural variability, documenting the climate system's response to past changes that have no analog in the historical record. Sediment archives of the recent geologic past, both marine and lacustrine, offer the opportunity to study how climate responds to a range of forcings and changing boundary conditions on timescales ranging from years to millennia. In this dissertation I use lacustrine and marine sediment to investigate changes late Quaternary climate, with particular focus on the Last Interglacial period (LIG). First, I use multiple approaches to reconstruct long-term changes in the West African Monsoon by investigating centennial-scale hydrologic variability recorded in Lake Bosumtwi sediments over the past 530,000 years. Over this interval, hydrology in the region is driven by a complex interplay of orbital forcing and glacial-interglacial boundary conditions. Lake level was generally much lower between 50 and 300 ka, likely due to the redistribution of rainfall from the tropics to the subtropics, driven by eccentricity's amplification of precession. Consequently, the Holocene highstand at the lake was both larger and longer lived than the maximum highstand during the LIG. Annual layers were continuously deposited through the LIG in Lake Bosumtwi, and I also present a new, 12,100 year-long, varve record spanning the interval from 128.6 to 116.5 ka. Over the course of the LIG, lake level generally tracks sea surface temperatures (SST) in Gulf of Guinea, including an abrupt drop in lake level that lasted about 500 years ca. 118 ka, coincident with cool SSTs in the North Atlantic and severe aridity in Europe. I find that the despite the generally drier conditions, hydrology varied 12 on similar timescales as the late Holocene, with pronounced multidecadal to centennialscale variability with non-stationary periodicities. I also investigate the contribution of ocean thermal expansion to sea level rise during the LIG, using a synthesis of paleoceanographic data and a climate model simulation. Globally, LIG SSTs were similar to, or slightly cooler than late Holocene SSTs, with the exception of the North Atlantic, which was several degrees warmer. Consequently, thermal expansion was likely a minor component of sea level rise during the interval, explaining between -0.3 and 0.4 m. of the 6 to 8 m highstand. Lastly, I tested the potential of Raman spectroscopy as a new, non-destructive technique to rapidly measure oxygen isotopic ratios in carbonates at extremely high resolution. Analyses on a suite a synthetic calcites indicate that 18O/16O ratios can be measured directly from the Raman spectra and have a 1:1 correspondence with traditional mass-spectrometry measurements. At present, the technique does not have the precision necessary to record natural variability, although there is considerable potential for improving the precision of the technique. 13 1. INTRODUCTION One of the greatest challenges of the 21st century will be mitigating and adapting to global climate change. As concentrations of greenhouse gases continue to build up, ice caps continue melting, and atmospheric and ocean temperatures continue to rise, Earth's climate system will be subjected to forcings and boundary conditions outside the range of observed climate variability. A robust, physically-based understanding of the climate system is therefore a prerequisite for anticipating global, and more challengingly, regional climate impacts. The relative brevity of the instrumental record limits our capacity to understand some aspects of natural climate variability, primarily the nature of long-term (decadal-scale and longer) variability, and the system's response to changes in forcing that are outside the limited range of variability in recent decades to centuries. Herein lies a key role of paleoclimatology: to extend the instrumental record to capture a wider range of natural variability, and to document the climate system's response to past changes that have no modern analog. This Earth system forcing and response approach to paleoclimatology requires continuous, well-dated, and high-resolution records, especially when we want to know how dramatic changes affect climate on human timescales. The integration of paleoclimate records are critical for understanding long-term climate dynamics and providing insight into future change in sub-Saharan West Africa. The region is prone to abrupt and prolonged drought; the “Sahel Drought” of the late 20th century began unexpectedly in the late 1960's and lasted for decades (Nicholson, 2001). The severity of the drought was greater than what was previously thought possible; to 14 illustrate, annual precipitation totals for the region for every year from 1970 to 1997 were below the 100-yr mean (Nicholson and Webster, 2007). Subsequent research has linked the decadal- to multidecadal-scale variability in West African hydrology to slow-varying changes in the meridional structure of sea surface temperatures in tropical and subtropical Atlantic Ocean (Lamb, 1978; Fontaine and Jacinot, 1996; Janicot et al., 2001; Camberlin et al., 2001; Nicholson and Grist, 2007). The instrumental record in the region is far too short (< 100 yrs; Nicholson, 2001) to capture such long term variability, so using highresolution paleoclimate records to extend the instrumental record in region (e.g., Shanahan et al., 2009) is necessary to understand the dynamics of drought in the region. Considerable uncertainty remains in projections of future change for sub-Saharan West Africa. Precipitation variability in the region is controlled by the complex dynamics of the West African Monsoon (WAM). Unlike convective precipitation in other equatorial and monsoonal regions, the West African monsoon is not driven by the convergence of moist southwesterly monsoonal flow and the Intertropical Convergence Zone (ITCZ), but rather, by the interaction of upper-level circulation and jets with low level moisture transport (Nicholson, 2009; Pu and Cook, 2010; 2012). This complexity, and the importance of small-scale features not resolved in many climate models, such as the West African Westerly Jet, is part of the reason why most climate models do not simulate a realistic West African Monsoon, both over the 20th century and in the mid-Holocene (Cook and Vizy, 2006; Randall et al., 2007; Christensen, 2007; Pu and Cook, 2010, Rodriguez-Fonseca et al., 2011). Only 4 of the 18 climate models in the Fourth Assessment Report (AR4) of the Intergovernmental Panel on Climate Change (IPCC) 15 demonstrated reasonably realistic simulations of the WAM during the second half of the 20th century (Cook and Vizy, 2006). Furthermore, Rodriguez-Fonseca et al. (2011) concluded that, in general, the IPCC AR4 models did not accurately simulate the tropical SST teleconnections that drive both interannual and decadal variability of the WAM. Even among models that performed reasonably well in the 20th century, projections for the 21st century vary wildly (Cook and Vizy, 2006). Taken together, the short instrumental record in a region dominated by long-term variability and the uncertainty of future projections for the region highlights the potential of paleoclimatology to bolster our understanding of climate in the region. Unfortunately, continental West Africa is one of the most data-poor regions in the world, both over the past two thousand years (cf., Mann et al., 2008) and during the past glacial period (cf., Blome et al., 2012). The primary focus of this dissertation is using lake sediments from Lake Bosumtwi in southern Ghana to improve our understanding of climate variability in the region, and to gain insight into future change in sub-Saharan West Africa. Lake Bosumtwi occupies a 1.08 +/- 0.04 Ma (Koeberl et al., 1997; Jourdan et al., 2009) meteorite-impact crater located in southern Ghana (6°30' N, 1°25' W) approximately 150 km north of the Gulf of Guinea (Figure 1). The lake has no outlet and is isolated from the groundwater table (Turner et al., 1996). Furthermore, the lake surface occupies about 50% of surface area of the 11-km-diameter impact crater (Turner et al., 1996). These features make lake level extremely sensitive to changes precipitation and evaporation, and lake level responds to interannual and decadal-scale changes in precipitation (Shanahan et al., 2007). 16 Lake Bosumtwi has been recognized as a valuable archive of past hydrologic and environmental change in West Africa since the 1970's. Seminal work by Talbot and Delibrias (1977) revealed large changes in lake level during the past 12 kyr recorded by terraces and sediments exposed by river cuts. This led to a series of studies on sediment cores taken from the lake and the geomorphology of the lake basin. Pollen studies revealed that arboreal forests, which were abundant in the region during the Holocene, were scarce in the region during the end of the last glaciation (Maley and Livingstone, 1983; Talbot et al., 1984; Maley 1987, 1991). Paleoecological studies from the lake also indicated a decrease in the tropical forests around the lake in the late Holocene (Maley 1997; 2002). These changes in paleoecology are consistent with geochemical evidence (elemental C and N, as well as δ13C and δ15N), which suggested a much lower lake level between 30 and 12 ka, and an extremely wet interval between 12 and 3 ka, before dropping to intermediate conditions in the late Holocene (Talbot and Johannessen, 1992; Russell et al., 2003). The extremely wet early to mid Holocene period corresponded to a distinct interval in the sediment, termed the “sapropel” by Talbot et al. (1984). The interval is extremely rich in organic material (organic carbon 15-20%), and poor in clastic minerals, and contains an important component of the filamentous blue-green algae anabeana (Talbot et al., 1984). Terraces, highstand deposits and an overflow notch indicate that the lake was very high, and occasionally overflowing during the deposition of this layer (Talbot and Delibrias, 1977, 1980; Shanahan et al., 2006). Lastly, Talbot and Kelts (1986) investigated the mineralogy and geochemistry of carbonates formed in the water column or surface sediments of the lake. This work highlighted the extreme 17 variability in both mineralogy and the δ13C and δ18O of carbonates in the sediment sequence. Recently, a second phase of research has begun on the lake. One goal of this research was to conduct further analysis on sediment cores from the the lake, using new approaches to better understand recent variability. For example, recent studies have taken advantage of the annual laminations (varves) persistent through the late Holocene to reconstruct regional hydrology with unprecedented resolution. This culminated in a 2700yr-long varve record that revealed pronounced mulitdecadal to century-scale variability in regional hydrology associated with changes in Atlantic SSTs and the Atlantic Multidecadal Oscillation (Shanahan et al., 2009). Peck et al. (2004) examined the mineral magnetic characteristics of the sediments over a longer time period to reveal intervals with abundant iron sulfide minerals associated with drought. These intervals are associated with cold events in the North Atlantic (the Younger Dryas and Heinrich events 1 and 2), indicating that millennial-scale fluctuations in North Atlantic SSTs and glacial boundary conditions also influence hydrology at the lake. The other goal of this second phase of research was to use drill cores to extend the sedimentary record through the entire million-year history of the lake (Koeberl et al., 2005; Peck et al., 2005). Work towards this goal is the primary focus of this dissertation, and a high-resolution record from lake extending back 530,000 years is presented in Appendix A. A particularly useful feature of Lake Bosumtwi is that throughout most of its history, the lake preserved finely laminated varves. One varved interval of particular interest corresponds to the Last Interglacial period (LIG). The LIG (roughly equivalent to 18 the Eemian, or Marine Isotope Stage 5e), which occurred from about 128 to 118 ka, is an interesting period to study for several reasons; first, it's the most recent analog to the Holocene, with generally similar conditions, although differences in the orbital configuration resulted in altered insolation forcing, especially seasonally (Berger and Loutre, 1991). It is often considered a partial analog for future climate, because global mean temperatures were slightly warmer (0-2°C; Turney and Jones, 2010; McKay et al., 2011) and sea level was 6-8 m higher than preindustrial conditions (Hearty et al., 2007; Kopp et al., 2009), although global CO2 (and other GHG) concentrations were far below modern and projected future concentrations (270-290 ppm; Luthi et al., 2008). The largest difference in forcing relative to the Holocene during the LIG is the change in seasonal forcing due to the amplification of the precession cycle by eccentricity (Berger and Loutre, 1991). This resulted in a more than 30 W m-2 increase in Northern Hemisphere summer insolation, and a corresponding decrease in Southern Hemisphere winter insolation at the height of the interglacial, ca. 124 ka. This drove warmer temperatures in the Northern Hemisphere, especially on land during the summer (Turney and Jones, 2010). The northern Hemisphere is consistently projected to warm faster in the future than the southern Hemisphere in all seasons, and especially on land (Meehl et al., 2007). More regionally, Weldeab et al., (2007) show that the SST in the Gulf of Guinea was generally warmer during the LIG than the Holocene, which is also a robust feature of GCM projections for the 21st century (Cook and Vizy, 2006). The Lake Bosumtwi record offers the rare opportunity to investigate past climate variability with annual resolution during this particularly interesting interval; these results are presented in Appendix B. 19 Another aspect of the LIG that is relevant to future climate change is that it was the last interval with substantially higher-than-modern sea level (6-8 m above modern; Hearty et al, 2007; Kopp et al., 2009). Sea level rise is one of the major socio-economic hazards associated with global warming, however expected contributions of the different components of sea level rise (i.e., ocean thermal expansion, melting of glaciers, and wasting of the Greenland and Antarctic Ice Sheets) are poorly understood. For example, estimates and projections of the thermal expansion component of sea level rise during the 21st century vary from 20 to 70% of total sea level rise over the 21st century (Meehl et al., 2007; Pfeffer et al., 2008; Vermeer and Rahmstorf, 2009). The sea level highstand during the Last Interglacial offers one approach to improve understanding of the relative contributions of the different sources of sea level rise. Previous work has made it clear that the majority of the sea level rise originated from melting of the Greenland Ice Sheet (GIS) and the Antarctic Ice Sheets (Otto-Bliesner et al., 2006; Kopp et al., 2009; Clark and Huybers, 2009), but the role of thermal expansion has not been carefully examined. In Appendix C, we combine analyses of paleoceanographic records and global climate model simulations to better constrain the amount of thermal expansion that occurred during the LIG sea-level highstand. High-resolution paleoclimatology is a rapidly-developing field. A critical aspect of progress in the field is the development of new techniques that allow high-resolution and rapid acquisition of paleoclimate archive data. Variability in oxygen isotope ratios is among the most widely used indicators of environmental change, and is a primary paleoclimate proxy for ice cores, corals, speleothems, soils, and marine and lacustrine 20 sediments. For most of these archives, δ18O is measured on carbonate minerals, typically using a stable isotope mass spectrometer. Recently, laser-based optical approaches, such as wavelength-scanned cavity ring-down spectroscopy (WS-CRDS), are allowing for more rapid, and portable, isotopic analysis. Raman spectroscopy is a technique that relies on the interaction of light with a molecule to infer information about the vibrational modes of the molecule. The shifts in energy resulting from excitation of the vibrational modes result in a corresponding shift in the wavelength of the monochromatic light (usually a laser) used to excite the sample. This is termed a “Raman shift” and the resulting spectrum of Raman shifts and their intensities corresponds to the energies of the stretching modes in the molecule. In carbonates, the symmetric stretching mode of the oxygen atoms about the carbon atom forms a distinct Raman peak, and the Raman shift is different for carbonate molecules with one (or more) 18O atom, than for those with three 16 O atoms. It is therefore theoretically possible to quantify the relative abundance of the different molecule types, and calculate δ18O. If the technique works with sufficient precision and accuracy, it would be extremely useful in high-resolution paleoclimatology. Raman spectroscopy is rapid, non-destructive, and can be performed at a resolution on the order of 1 micron. Such a device could even be made portable. Investigation into the potential and utility of the technique is presented in Appendix D. 21 2. PRESENT STUDY The methods, results and conclusions of this study are presented as four papers formatted for publication in professional journals. All papers are coauthored, and I am senior author on all four manuscripts. The following is a brief summary of the most important findings in this dissertation. The majority of the research presented in this dissertation has stemmed from my work on the Lake Bosumtwi sediment cores, the primary focus of my work as a student at the University of Arizona. In the first study (Appendix A), we present the first continuous, high-resolution terrestrial record of sub-Saharan paleohydrological variability for the past 530,000 years; a sediment geochemical record from Lake Bosumtwi in southern Ghana (6°30' N, 1°25' W). In the study, we used the abundance of terrigenous elements in the sediment, as well as the mineralogy and isotope geochemistry of carbonates formed in the lake to infer past changes in lake level and hydrology in the region. Over the past 530 ka, moisture balance at the lake is driven by a complex interplay of orbital forcing and glacial-interglacial boundary conditions. The 400-kyr component of orbital eccentricity drove a more arid background state at the lake between 300 and 100 ka, likely due to the amplification of precession and the redistribution of tropical moisture to more subtropical regions. On glacial-interglacial timescales, lake level displays a positive relationship with the 100-kyr component of eccentricity, such that interglacial periods are generally less arid than glacial periods, likely due to the influence of warmer north Atlantic SSTs during interglacial periods. The influence of 22 Atlantic SSTs is apparent on millennial-timescales as well. Over the past 130 kyr, cold periods recorded in Greenland ice cores (NGRIP 2004; Wolff et al., 2010), especially those associated with to Heinrich Events, consistently correspond to low lake level at Lake Bosumtwi. Six multi-millennial scale intervals of extreme aridity were recorded in the sediments during the past 530 ka, all during the eccentricity-high interval from 30050 ka. These events generally occurred during times of similar orbital and glacial boundary conditions, that is, deep glacial conditions and local summer insolation minima. Two of the intervals represent abrupt drying from wetter-than-average conditions, and were immediately preceded and followed by intervals of high lake level. The transitions into and out of these two intervals occurred rapidly, on the order of decades to centuries, although the origin of these arid excursions remains unclear. In Appendix B we present a new 12,100-year-long varve record for the Last Interglacial (LIG) at Lake Bosumtwi, spanning the interval from 128.6 to 116.5 ka. We use the abundance of terrigenous elements in the sediment, varve thickness, and the isotope geochemistry and mineralogy of authigenic carbonates in the sediment to infer changes in lake level during the interval. The LIG peak highstand was both lower and shorter-lived than the prolonged highstand in the early Holocene, and unlike the Holocene, the evidence suggests the lake never overflowed during LIG. This result is consistent with those from Appendix A, which indicated generally drier conditions during intervals of high eccentricity (on the 400-kyr timescale). In other respects, the LIG and the Holocene were similar. Both interglacials had two distinct millennial-scale intervals of higher lake level that occurred during times of precession-driven insolation maxima in 23 July and October, corresponding to the two rainy seasons in the modern climatology. Over the course of the LIG, lake level generally tracks sea surface temperatures (SST) in the Gulf of Guinea (Weldeab et al., 2007), including an abrupt drop in lake level that lasted about 500 years ca. 118 ka, corresponding to cooling in the Gulf of Guinea and much of the North Atlantic (e.g., Lehman et al., 2002; Oppo et al., 2006) during the interval. The timing and duration of the event are comparable to the Late Eemian Aridity Pulse (LEAP) that is observed at several European sites (Sirocko et al., 2005; Seelos and Sirocko, 2007) and has been interpreted to result from abrupt cooling in the North Atlantic, and possibly a reduction of Atlantic meridional overturning circulation (AMOC). A reduction in AMOC, and a cooling of North Atlantic would likely result in aridity in West Africa (Street-Perrott and Perrot, 1990; Chang et al., 2008; Mulitza et al., 2008; Shanahan et al., 2009) so the aridity ca. 118 ka is the first indication that the LEAP occurred in Africa as well as Europe. Further analysis of the Last Interglacial period is presented in Appendix C. In this study we produced the first quantitative estimate of the thermosteric component of LIG sea level rise. To do this, we compiled paleoceanographic data and used a coupled atmosphere-ocean climate model simulation to examine global SSTs during the height of the LIG sea level highstand, which was 6 to 8 meter higher than present (Hearty et al., 2007; Kopp et al., 2009). The paleoceanographic data suggest a peak LIG global sea surface temperature (SST) warming of 0.7±0.6°C compared to the late Holocene. Our LIG climate model simulation suggests a slight cooling of global average SST relative to preindustrial conditions (ΔSST = -0.4°C), associated with a reduction in atmospheric 24 water vapor in the Southern Hemisphere driven by a northward shift of the Intertropical Convergence Zone, and substantially reduced seasonality in the Southern Hemisphere. Taken together, the data suggest a small contribution of thermal expansion during the interval; it is unlikely that thermosteric sea level rise exceeded 0.4±0.3 m during the LIG. This constraint, along with estimates of the sea level contributions from the Greenland Ice Sheet (Otto-Bliesner et al., 2006), glaciers and ice caps (Radić and Hock, 2010), implies that 4.1 to 5.8 m of sea level rise during the Last Interglacial period was derived from the Antarctic Ice Sheet. The final study, Appendix D, investigates the potential for using Raman spectroscopy to measure 18O-16O oxygen isotopic ratios in carbonates. To do this, we synthesized a suite of isotopically-enriched calcite crystals ranging in composition from natural (0.2% 18O) to six-times-natural abundance (1.2% 18O). We determined the Raman-inferred isotopic ratios (RRaman) of individual crystals by fitting pseudoVoigt curves to the ν1 symmetric stretching peak at 1086 cm-1 and the smaller satellite peak, associated with the ν1 stretching mode of singly-substituted carbonate groups (C16O218O) at 1065 cm-1. The ratio of the two numerically-integrated peak areas shows a 1:1 correspondence with the 18O/16O ratios measured by water and carbonate mass spectrometry, confirming that the relative intensities of the ν1 symmetric stretching peaks is a direct measure of the isotopic ratio in the carbonates. The 1-sigma uncertainties of the RRaman values of the individual crystals were 0.00079 (384‰ PDB), and 0.00043 (210‰ PDB) for the 4-crystal sample means. Although this level of uncertainty is too high to detect natural variability, there are multiple prospects for improving the precision of the 25 technique. Carbon isotope ratios in carbonates cannot be measured by this approach, however our results do highlight the potential of Raman-based mass spectrometry for C and other elements in minerals and organic compounds, such as calcium oxalates. 26 90 1°27' W 300 90 90 90 45 6°32' N 30 m 50 m 90 90 45 90 90 45 70 m 45 90 6°27' N 90 90 0 1 2 km Figure 1. Site map of the Lake Bosumtwi crater, modified from Shanahan (2006). Solid gray lines show bathymetry (10 m contour). Dashed line indicates crater rim. Inset: location of Lake Bosumtwi in southern Ghana, West Africa. 27 3. REFERENCES Berger, A., and M. F. Loutre (1991), Insolation values for the climate of the last 10 million years, Quaternary Science Reviews, 10(4), 297–317. Camberlin, P., S. Janicot, and I. Poccard (2001), Seasonality and atmospheric dynamics of the teleconnection between African rainfall and tropical sea-surface temperature: Atlantic vs. ENSO, International Journal of Climatology, 21(8), 973– 1005. Chang, P., R. Zhang, W. Hazeleger, C. Wen, X. Wan, L. Ji, R. J. Haarsma, W.-P. Breugem, and H. Seidel (2008), Oceanic link between abrupt changes in|[nbsp]| the North Atlantic Ocean and the|[nbsp]|African|[nbsp]|monsoon, Nature Geoscience, 1(7), 444–448, doi:10.1038/ngeo218. Christensen, J. H. et al. (2007), Regional climate projections, 2007: The Physical Science Basis. Contribution of Working group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, University Press, Cambridge, Chapter 11, 847–940. Clark, P. U., and P. Huybers (2009), Global change: Interglacial and future sea level, Nature, 462(7275), 856–857. Cook, K. H., and E. K. Vizy (2006), Coupled model simulations of the West African monsoon system: Twentieth-and twenty-first-century simulations, Journal of climate, 19(15), 3681–3703. Fontaine, B., and S. Janicot (1996), Sea surface temperature fields associated with West African rainfall anomaly types, Journal of climate, 9(11), 2935–2940. Hearty, P. J., J. T. Hollin, A. C. Neumann, M. J. O’Leary, and M. McCulloch (2007), Global sea-level fluctuations during the Last Interglaciation (MIS 5e), Quaternary Science Reviews, 26(17-18), 2090–2112. Janicot, S., S. Trzaska, and I. Poccard (2001), Summer Sahel-ENSO teleconnection and decadal time scale SST variations, Climate Dynamics, 18(3), 303–320. Jourdan, F., P. R. Renne, and W. U. Reimold (2009), An appraisal of the ages of terrestrial impact structures, Earth and Planetary Science Letters, 286(1), 1–13. Koeberl, C., R. Bottomley, B. P. Glass, and D. Storzer (1997), Geochemistry and age of Ivory Coast tektites and microtektites, Geochimica et Cosmochimica Acta, 61(8), 1745–1772. 28 Koeberl, C., J. Peck, J. King, B. Milkereit, J. Overpeck, and C. Scholz (2005), The ICDP Lake Bosumtwi Drilling Project: A First Report, Scientific Drilling, 1, 23–27. Kopp, R. E., F. J. Simons, J. X. Mitrovica, A. C. Maloof, and M. Oppenheimer (2009), Probabilistic assessment of sea level during the Last Interglacial stage, Nature, 462(7275), 863–867. Lamb, P. J. (1978), Large-scale tropical Atlantic surface circulation patterns associated with Subsaharan weather anomalies, Tellus, 30, 240–251. Lehman, S. J., J. P. Sachs, A. M. Crotwell, L. D. Keigwin, and E. A. Boyle (2002), Relation of subtropical Atlantic temperature, high-latitude ice rafting, deep water formation, and European climate 130,000-60,000 years ago, Quaternary Science Reviews, 21(18-19), 1917–1924. Lüthi, D. et al. (2008), High-resolution carbon dioxide concentration record 650,000– 800,000 years before present, Nature, 453(7193), 379–382. Maley, J. (1987), Fragmentation de la forêt dense humide africaine et extension des biotopes montagnards au Quaternaire récent: nouvelles données polliniques et chronologiques, implications paléoclimatiques et biogéographiques (1987). Palaeoecology of Africa and the surrounding islands. Maley, J. (1991), The African rain forest vegetation and palaeoenvironments during late Quaternary, Climatic change, 19(1), 79–98. Maley, J. (1997), Middle to Late Holocene changes in tropical Africa and other continents: paleomonsoon and sea surface temperature variations, Nato ASI Series I: Global Environmental Change, 49, 611–640. Maley, J. (2002), A catastrophic destruction of African forests about 2,500 years ago still exerts a major influence on present vegetation formations, IDS bulletin, 33(1), 13–30. Maley, J., and D. A. Livingstone (1983), Extension d’un élément montagnard dans le sud du Ghana (Afrique de l’Ouest) au Pléistocène supérieur et à l’Holocène inférieur: premières données polliniques, CR Acad. Sci. Paris, Sér, 2(196), 1287–1292. Mann, M. E., Z. Zhang, M. K. Hughes, R. S. Bradley, S. K. Miller, S. Rutherford, and F. Ni (2008), Proxy-based reconstructions of hemispheric and global surface temperature variations over the past two millennia, Proceedings of the National Academy of Sciences, 105(36), 13252. McKay, N. P., J. T. Overpeck, and B. L. Otto-Bliesner (2011), The role of ocean thermal 29 expansion in Last Interglacial sea level rise, Geophysical Research Letters, 38(14), L14605. Meehl, G. A. et al. (2007), Chapter 10: Global Climate Projections. IPCC Fourth Assessment Report. The Physical Science Basis. Contribution of Working group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, University Press, Cambridge, Chapter 10, 747-846. Mulitza, S., M. Prange, J. B. Stuut, M. Zabel, T. von Dobeneck, A. C. Itambi, J. Nizou, M. Schulz, and G. Wefer (2008), Sahel megadroughts triggered by glacial slowdowns of Atlantic meridional overturning, Paleoceanography, 23(4), PA4206. NGRIP et al. (2004), High-resolution record of Northern Hemisphere climate extending into the Last Interglacial period, Nature, 431(7005), 147–151. Nicholson, S. E. (2001), Climatic and environmental change in Africa during the last two centuries, Climate Research, 17(2), 123–144. Nicholson, S. E. (2009), A revised picture of the structure of the “monsoon” and land ITCZ over West Africa, Climate Dynamics, 29, doi:10.1007/s00382-008-0514-3. Nicholson, S. E., and J. P. Grist (2001), A conceptual model for understanding rainfall variability in the West African Sahel on interannual and interdecadal timescales, International Journal of Climatology, 21(14), 1733–1757. Nicholson, S. E., and P. J. Webster (2007), A physical basis for the interannual variability of rainfall in the Sahel, Quarterly Journal of the Royal Meteorological Society, 133(629). Oppo, D. W., J. F. McManus, and J. L. Cullen (2006), Evolution and demise of the Last Interglacial warmth in the subpolar North Atlantic, Quaternary Science Reviews, 25(23-24), 3268–3277. Otto-Bliesner, B. L., S. J. Marshall, J. T. Overpeck, G. H. Miller, and A. Hu (2006), Simulating Arctic climate warmth and icefield retreat in the last interglaciation, Science, 311, 1751–1753. Peck, J. A., R. R. Green, T. Shanahan, J. W. King, J. T. Overpeck, and C. A. Scholz (2004), A magnetic mineral record of Late Quaternary tropical climate variability from Lake Bosumtwi, Ghana, Palaeogeography, Palaeoclimatology, Palaeoecology, 215(1-2), 37–57. Pfeffer, W. T., J. T. Harper, and S. O’Neel (2008), Kinematic Constraints on Glacier Contributions to 21st-Century Sea-Level Rise, Science, 321(5894), 1340 –1343, 30 doi:10.1126/science.1159099. Pu, B., and K. H. Cook (2010), Dynamics of the West African westerly jet, Journal of Climate, 23(23), 6263–6276. Pu, B., and K. H. Cook (2012), Role of the West African Westerly Jet in Sahel Rainfall Variations, Journal of Climate, 25(8), 2880–2896, doi:10.1175/JCLI-D-1100394.1. Radić, V., and R. Hock (2010), Regional and global volumes of glaciers derived from statistical upscaling of glacier inventory data, Journal of Geophysical Research, 115, 10 PP., doi:201010.1029/2009JF001373. Randall, D. A. et al. (2007), Climate models and their evaluation: The Physical Science Basis. Contribution of Working group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, University Press, Cambridge, Chapter 8, 589-662. Rodríguez-Fonseca, B. et al. (2011), Interannual and decadal SST-forced responses of the West African monsoon, Atmospheric Science Letters, 12(1), 67–74. Russell, J., M. R. Talbot, and B. J. Haskell (2003), Mid-Holocene climate change in Lake Bosumtwi, Ghana, Quaternary Research, 60(2), 133–141. Seelos, K., and F. Sirocko (2007), 14. Abrupt cooling events at the very end of the Last Interglacial, Developments in Quaternary Sciences, 7, 207–229. Shanahan, T. M., J. T. Overpeck, W. E. Sharp, C. A. Scholz, and J. A. Arko (2007), Simulating the response of a closed-basin lake to recent climate changes in tropical West Africa (Lake Bosumtwi, Ghana), Hydrological Processes, 21(13). Shanahan, T. M., J. T. Overpeck, C. W. Wheeler, J. W. Beck, J. S. Pigati, M. R. Talbot, C. A. Scholz, J. Peck, and J. W. King (2006), Paleoclimatic variations in West Africa from a record of late Pleistocene and Holocene lake level stands of Lake Bosumtwi, Ghana, Palaeogeography, Palaeoclimatology, Palaeoecology, 242(3), 287–302. Shanahan, T., J. Overpeck, K. Anchukaitis, J. Beck, J. Cole, D. Dettman, J. Peck, C. Scholz, and J. King (2009), Atlantic forcing of persistent drought in West Africa, Science, 324(5925), 377. Sirocko, F., K. Seelos, K. Schaber, B. Rein, F. Dreher, M. Diehl, R. Lehne, K. Jäger, M. Krbetschek, and D. Degering (2005), A late Eemian aridity pulse in central Europe during the last glacial inception, Nature, 436(7052), 833–836. 31 Street-Perrott, F. A., and R. A. Perrott (1990), Abrupt climate fluctuations in the tropics: the influence of Atlantic Ocean circulation, Nature, 343(6259), 607–612, doi:10.1038/343607a0. Talbot, M. R., and G. Delibrias (1977), Holocene variations in the level of Lake Bosumtwi, Ghana, Nature, 268(5622), 722–724. Talbot, M. R., and T. Johannessen (1992), A high resolution palaeoclimatic record for the last 27,500 years in tropical West Africa from the carbon and nitrogen isotopic composition of lacustrine organic matter, Earth and Planetary Science Letters, 110(1-4), 23–37. Talbot, M. R., and K. Kelts (1986), Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana, Geology, 14(11), 912–916. Talbot, M. R., D. A. Livingstone, P. G. Palmer, J. Maley, J. M. Melack, G. Delibrias, and S. Gulliksen (1984), Preliminary results from sediment cores from Lake Bosumtwi, Ghana, Palaeoecology of Africa, 16, 173–192. Turner, B. F., L. R. Gardner, and W. E. Sharp (1996), The hydrology of Lake Bosumtwi, a climate-sensitive lake in Ghana, West Africa, Journal of Hydrology, 183(3-4), 243–261. Turney, C. S. ., and R. T. Jones (2010), Does the Agulhas Current amplify global temperatures during super-interglacials?, Journal of Quaternary Science, 25(6), 839–843. Vermeer, M., and S. Rahmstorf (2009), Global sea level linked to global temperature, Proceedings of the National Academy of Sciences, 106(51), 21527–21532. Weldeab, S., D. W. Lea, R. R. Schneider, and N. Andersen (2007), 155,000 years of West African monsoon and ocean thermal evolution, Science, 316(5829), 1303. Wolff, E. W., J. Chappellaz, T. Blunier, S. O. Rasmussen, and A. Svensson (2010), Millennial-scale variability during the last glacial: The ice core record, Quaternary Science Reviews, 29(21–22), 2828–2838, doi:10.1016/j.quascirev.2009.10.013. 32 4. APPENDICES 33 APPENDIX A A HIGH-RESOLUTION PALEOHYDROLOGICAL RECORD FOR THE PAST 530,000 YEARS FROM LAKE BOSUMTWI, SOUTHERN GHANA 34 A HIGH-RESOLUTION PALEOHYDROLOGICAL RECORD FOR THE PAST 530,000 YEARS FROM LAKE BOSUMTWI, SOUTHERN GHANA Nicholas P. McKay1, Jonathan T. Overpeck1,2,3, Timothy M. Shanahan4, Erik T. Brown5, John A. Peck6, Clifford W. Heil7, John W. King7, and Chris A. Scholz8 1 Department of Geosciences, University of Arizona Department of Atmpospheric Science, University of Arizona 3 Institute of the Environment, University of Arizona 4 Jackson School of Geosciences, University of Texas at Austin 5 Large Lakes Observatory, University of Minnesota Duluth 6 Department of Geology and Environmental Science, University of Akron 7 Graduate School of Oceanography, University of Rhode Island 8 Deptartment of Earth and Environmental Sciences, Syracuse University 2 35 Abstract The West African Monsoon (WAM) system is an important component of the global climate system. It is also the primary source of moisture for millions of people in sub-Saharan Africa. The region is prone to decadal to multi-decadal droughts and pluvials associated with slow changes in Atlantic sea surface temperatures (SSTs). The instrumental climate record is too short to understand natural variability in the WAM, let alone how it responds to large changes in global climate. Here we present the first, highresolution terrestrial record of sub-Saharan paleohydrological variability for the past 530,000 years; a sediment geochemical record from Lake Bosumtwi in southern Ghana (6°30' N, 1°25' W). The abundance of terrigenous elements in the sediment, as well as the mineralogy and isotope geochemistry of carbonates formed in the lake, are both sensitive to changes in lake level, and reveal a dynamic history of hydrologic variability over the past 530 kyr. Hydrology in the region is driven by a complex interplay of orbital forcing and glacial-interglacial boundary conditions. The 400-kyr component of eccentricity drove a more arid background state at the lake between 300 and 100 ka, likely due to the amplification of precession and the redistribution of tropical moisture to more subtropical regions. On glacial-interglacial timescales, lake level displays a positive relationship with the 100-kyr component of eccentricity, such that interglacial periods are generally less arid than glacial periods, likely due to the influence of warmer north Atlantic SSTs during interglacial periods. The competing roles of local insolation forcing and North Atlantic boundary conditions are apparent over the past 150 kyr on a latitudinal transect of records sensitive to West African hydrology. During that interval, more northerly sites in West 36 Africa are more sensitive to obliquity and North Atlantic temperatures, whereas more southerly sites respond more to local insolation and precession forcing. The extent of the competing impacts changes with the amplitude of local insolation forcing, as larger swings in precession before 70 ka had a more widespread impact on West African climate. Six multi-millennial scale intervals of extreme aridity were recorded in the sediments during the past 530 ka, all during the eccentricity-high interval from 300-50 ka. These events are generally associated with deep glacial conditions and local summer insolation minima, and in two cases, were immediately preceded and followed by intervals of high lake level. The transitions, both into and out of the arid intervals, appear to be rapid, occurring on the order of centuries or less, although the underlying climate dynamics of these transitions remains uncertain. 1. Introduction The tropics are the heat engine of global climate, and tropical monsoon systems play a critical role in distributing moisture and energy (Wallace and Hobbs, 2006; Chang et al., 2011). The tropics also initiate, modulate and interact with both natural and anthropogenic climate change (Peterson et al., 2000; Visser et al., 2003; Hoerling and Kumar, 2003; Wang et al., 2007; Lu et al., 2009). Much of observed variability in tropical monsoon systems has been associated with slow-varying changes in the tropical oceans (Lamb, 1978; Fontaine et al., 1998; Janicot et al., 2001; Cook and Vizy, 2001; Nicholson and Webster, 2007); however the terrestrial instrumental records are too short to understand natural variability in the monsoon systems, let alone their response to abrupt 37 changes. The paleoclimatic history of continental tropics are particularly poorly understood when compared to the oceans and higher latitudes, both in terms of lowfrequency climate variability and abrupt change. This is largely because of the lack of long, continuous records comparable to ice and ocean sediment cores, and because of the lack of well-constrained age models. Recently, records from the Cariaco Basin (Peterson et al., 2000; Yarinik et al., 2000; Hughen et al., 2004), Lakes Malawi (e.g., Cohen et al., 2007; Brown et al., 2007; Johnson et al., 2011; Stone et al., 2011) Challa (Verschuren et al., 2009; Moernaut et al., 2010) and Tanganyika (Tierney et al., 2008), and South American speleothems (Cruz et al., 2005; Cruz et al., 2009; Kanner et al., 2012) have broadened our perspective of how tropical climate has varied on long timescales, responds to (and possibly forces) abrupt change in the North Atlantic, and is modulated by orbital variability. New, long drill records from lakes in both East and West Africa, including initial results from Lake Bosumtwi, suggest that severe aridity was much more common on the continent between 135 and 75 ka than from 75 to 0 ka (Scholz et al., 2007). Nevertheless, the long-term climatic history of tropical West Africa, and the past response of the West African Monsoon (WAM) to extreme climate changes remain poorly understood, which is of particular concern given the region's vulnerability to drought, and considerable uncertainty about how continued climate change will affect the hydrology of sub-Saharan Africa (Cook and Vizy, 2006). The continuous, high-resolution sediment sequence from Lake Bosumtwi, offers the first opportunity to use a terrestrial record to investigate how West Africa responds to orbital forcing and glacial boundary conditions over the past 38 million years (Peck et al., in prep). 1.1 Setting Lake Bosumtwi is a 1.08 +/- 0.04 Ma (Koeberl et al., 1997; Jourdan et al., 2009) meteorite-impact crater lake located in southern Ghana (6°30' N, 1°25' W; Figure A1). The crater is the youngest and best preserved complex impact structure on Earth (Koeberl and Milkereit, 2007). The modern lake is about 8 km in diameter, and fills about half of crater area, so the majority (80%) of the annual water input to the lake comes from rainfall directly on the surface (Turner et al., 1996a). The steep sides of the crater and depth of the lake (76 m) promote semi-permanent stratification of the water column; the upper 35 m overturn seasonally, however the lake does not mix below 35 m depth (Almond and Hecky, 2000). Consequently, the bottom waters are anoxic, and sediments preserve finely laminated varves. Lake level is sensitive to changes precipitation and evaporation, and responds to intraannual to decadal-scale changes in precipitation, varying by more than a meter seasonally, and falling more than 2 m during the “Sahel Drought” from 1970-2000 (Shanahan et al., 2007). Sediment physical properties and the stable isotope geochemistry of carbonates precipitated in the lake are sensitive to lake level (Talbot and Delibrias, 1977; Talbot and Kelts, 1986; Peck et al., 2004; Shanahan, 2006; Scholz et al., 2007; Shanahan et al., 2009). This study focuses on the upper 160 meters of the longest sedimentary drill core from the lake. 39 1.2 Regional Climatology The annual migration of the tropical rainbelt over Guinea Coast drives a bimodal distribution of annual precipitation, with peak rainfall occurring in May-July and a smaller peak from September-October (Nicholson, 2009) (Figure A2). Wind speed and direction in the region is also seasonal, and are controlled by the position of the Intertropical Convergence Zone (ITCZ). In the boreal summer the ITCZ is positioned north of the lake over the Sahel and Sahara, and the monsoonal southwesterlies peak in August. The winds drive upwelling off the Guinea Coast, cooling the Gulf of Guinea and the Guinea Coast region to its annual minimum (Shanahan, 2006). This reduces evaporation and stabilizes the atmosphere, suppressing convection in the region (Moron, 1994, Opoku-Ankomah, 1994). During the boreal winter the ITCZ is positioned near the Guinea Coast, wind speed is reduced, and warm, aerosol-rich northeasterly 'Harmattan' winds are common. Temperatures are highest during this season (~28°C at Kumasi; Shanahan, 2006). Dust supplied to the region by these winds has been shown to be a substantial component of Ghanaian soils (He et al., 2007) and solute load in Lake Bosumtwi (Turner et al., 1996b). Precipitation variability in West Africa is strongly influenced by the distribution of sea surface temperatures (SSTs), particularly in the tropical Atlantic (Lamb, 1978; Janicot et al., 2001), but also in the eastern Mediterranean (Rowell, 2003), and tropical Pacific (Rowell, 2001; Janicot et al., 2001). Two modes of variability in the tropical Atlantic have been linked to West African precipitation. The equatorial zonal mode, termed “Atlantic Niños” (Zebiak, 1993) is primarily associated with interannual variability, and 40 the warm phase of the mode drives increased convection and precipitation over the Guinea Coast (Moron, 2001). The pronounced decadal variability in West African rainfall has been associated with the meridional mode, which controls the cross-equatorial SST and pressure gradient during the boreal summer (Fontaine and Jacinot, 1996; Camberlin et al., 2001; Nicholson and Grist, 2007). This sea-level pressure gradient influences the strength and position of the mid-level African Easterly Jet (AEJ), which along with the strength and position of the upper level Tropical Easterly Jet (TEJ), control the intensity and position of deep convection during the monsoon season (Nicholson and Webster, 2007; Nicholson, 2008; 2009a; Dezfuli and Nicholson, 2011). Upper atmospheric circulation modulates interannual variability in the convective conditions, whereas moisture is primarily derived from lower level in the atmosphere. There are two sources of low-level moisture for the WAM: the southerly monsoonal flow, and a distinct, lowlevel westerly jet originating off the coast of West Africa near the Atlantic ITCZ, which was identified by Grodsky et al. (2003). This low-level jet has been termed the West African Westerly Jet (WAWJ), and recent work suggests that it is critical in driving interannual and decadal variability in precipitation in the Sahel (Pu and Cook, 2010; 2012). Past hydrologic changes in the region have been related to interhemispheric Atlantic SST patterns as well. Shanahan et al., (2009) showed that lake level at Lake Bosumtwi was tied to the Atlantic Multidecadal Oscillation (AMO) over the past 2700 years. Cold intervals in the North Atlantic have been associated with extremely dry conditions in much of West Africa, most notably during Heinrich 1 (Stager et al., 2011); 41 although similar conditions have been observed in both terrestrial and marine proxies throughout the past 100 ka (Peck et al., 2004; Weldeab et al., 2007; Mulitza et al., 2008; Tjallingii et al., 2008; Castañeda et al., 2009; Itambi et al.; 2009). 1.3 Past work on Lake Bosumtwi Lake Bosumtwi has been recognized as a valuable archive of past hydrologic and environmental change in West Africa since the 1970's. Seminal work by Talbot and Delibrias (1977) revealed large changes in lake level during the past 12 kyr recorded by terraces and sediments exposed by river cuts. This led to a series of studies on sediment cores taken from the lake and the geomorphology of the lake basin. Pollen studies revealed that arboreal forests, which were abundant in the region during the Holocene, were scarce in the region during the end of the last glaciation (Maley and Livingstone, 1983; Talbot et al., 1984; Maley 1987, 1989, 1991). Further studies suggested that the tropical forests of the early Holocene abruptly declined ca. 3 ka (Maley, 1997; 2002). These changes in paleoecology are consistent with geochemical evidence (elemental C and N, as well as δ13C and δ15N), which suggested much lower lake levels between 30 and 12 ka, and an extremely wet interval between 12 and 3 ka, before dropping to intermediate conditions in the late Holocene (Talbot and Johannessen, 1992; Russell et al., 2003). The extremely wet early to mid Holocene period corresponded to a distinct interval in the sediment, termed the “sapropel” by Talbot et al. (1984). The interval is extremely rich in organic material (organic carbon 15-20%), and poor in clastic minerals, and contains an important component of the filamentous blue-green algae anabeana 42 (Talbot et al., 1984). Terraces, highstand deposits and an overflow notch indicate that the lake was very high, and occasionally overflowing during the deposition of this layer (Talbot and Delibrias, 1977, 1980; Shanahan et al., 2006). Lastly, Talbot and Kelts (1986) investigated the mineralogy and geochemistry of carbonates formed in the water column or surface sediments of the lake. This work highlighted the extreme variability in both mineralogy and the δ13C and δ18O of carbonates in the sediment sequence. A second phase of research on the lake began around the turn of the 21st Century, with two major emphases. One goal of this research was to conduct further analysis on sediment cores from the the lake, using new approaches to better understand recent variability. For example, taking advantage of the annual laminations (varves) persistent through the late Holocene to reconstruct regional hydrology with unprecedented resolution. This culminated in a 2700-yr-long varve record that revealed pronounced mulitdecadal to century-scale variability in regional hydrology associated with changes in Atlantic SSTs and the Atlantic Multidecadal Oscillation (AMO; Shanahan et al., 2009). Peck et al. (2004) examined the mineral magnetic characteristics of the lake to reveal intervals with abundant iron sulfide minerals associated with drought. These intervals are associated with cold events in the North Atlantic (the Younger Dryas and Heinrich events 1 and 2), indicating that millennial-scale fluctuations in North Atlantic SSTs and glacial boundary conditions also influence hydrology at the lake. High-resolution seismicreflection data from the lake revealed pronounced, dense layers corresponding to lake lowstands (Scholz et al., 2002; Brooks et al., 2005). The primary goal of this second phase of research was to use drill cores to extend the sedimentary record through the 43 entire million-year history of the lake (Koeberl et al., 2007). Initial results from the long drill cores identified several intervals of severe aridity between 75 and 135 ka, including the presence of a thick, dense unit at 32 m in the drill cores that corresponds to a lake wide seismic unconformity (Brooks et al., 2005), and likely resulted from nearly complete dessication of the lake (Scholz et al., 2007). In this study, we examine the geochemistry of the upper 160 m of the 296-m sediment sequence at cm-scale, extending the drill core record to 530 ka. 2. Methods 2.1 Lake Drilling and core processing In July and August of 2004, the global lake drilling system (GLAD-800) was used to recover 12 sediment cores from five core sites along a depth transect in the lake (Figure A3). This study focuses on the 296-m-long core from deep-water (76 m) site 5 (core BOS04-5B), which extends from the present-day lake floor to the brecciated bedrock, and includes a basal sediment layer of impact-glass bearing accretionary lapilli. This layer corresponds to the final fallback of material following the meteorite impact, indicating that the record spans the entire 1.08 Myr history of lacustrine sedimentation at the lake. After drilling, the cores were shipped to the University of Rhode Island, where they were split, and described, then imaged and analyzed using a Geotek® multi-sensor core logger for a suite of magnetic and physical parameters. The paleomagnetic parameters of the sediments were measured on U-channels taken central axis of the split core sections (Heil et al., in prep.). The cores were sampled at 4-cm-resolution for the 44 analysis of sediment magnetic hysteresis, x-ray diffraction mineralogy, total organic and inorganic carbon content, bulk organic carbon and nitrogen isotopes, and grain size. Core logging and physical sediment properties were used to integrate the depth scales for the three drill cores from site 5 (BOS04-5A,B&C) into a relative mean core depth (RMCD; pers. comm. C. Heil). 2.2 Geochronology The uppermost 22 m (RMCD) of the drill core have been dated by 155 radiocarbon and low-blank radiocarbon ages, as well as varve counts in both late Holocene and late Pleistocene intervals (Shanahan et al., submitted). The majority of these ages were measured on piston cores that have been tied into the RMCD scale (Shanahan et al., submitted). Beyond the limit of radiocarbon dating the age model is more uncertain; however, we have employed multiple, independent dating techniques to constrain the timing of sediment deposition in the lake. Optically stimulated luminescence (OSL) ages were determined in collaboration with S. Forman at the University of Illinois, Chicago (Shanahan et al., in prep). To avoid age underestimates that can be associated with anomalous fading of feldspars, we relied on the post-infrared, blue-light response that is mainly associated with quartz (Banerjee, 2001). Furthermore, the ages were also corrected for downcore variability in porosity, organic matter and biogenic silica content, which affect the dose rate (Forman et al., 2007). OSL ages between 0 and 50 ka were generally in agreement with radiocarbon age estimates (Figure A4). Below about 60 RMCD (130 ka), measured OSL samples were at or near saturation, 45 and provide only minimum ages, making this the effective limit of the technique for the drill core. U-Th dating of authigenic carbonates in the sediments provide additional age control; however the challenges associated with cleaning and mechanically separating the fine layers of carbonate to remove detrital Th limits both the precision and capacity of measuring sediments from deep in deep-water core 5B. Carbonate nodules found in shallow-water cores are easier to analyze, and more consistently yield usable ages. U-Th ages from core site 1 are tied into the site 5 chronology by correlating the down-hole natural gamma logs between the sites (Shanahan et al., in prep). Paleomagnetic properties of the sediments were used to identify the Brunhes/Matuyama and Upper Jaramillo geomagnetic reversal boundaries (Heil et al., in prep). The ages of these events are well known (781 +/- 3 and 988 +/- 3 ka; Horng et al., 2002), and constrain the chronology deep in the drill core. Finally, the timing of the meteor impact, and therefore the base of the sediment sequence, has been dated by 40Ar39 Ar and fission track techniques (Koeberl et al., 1997) to 1.08 +/- 0.04 Ma (Jourdan et al., 2009). 2.3 Scanning X-Ray fluorescence analysis The bulk elemental abundance of the upper 159 meters of core BOS04-5B was analyzed using the Itrax® scanning x-ray fluorescence (XRF) analyzer at the Large Lake Observatory at the University of Minnesota, Duluth. The archive halves of split sediment cores were analyzed at 1-cm-resolution, with 60 sec count times and a Mo x-ray source 46 operated at 30 kV and 20 mA. The Itrax was also used to generate digital x-radiographs with 200-μm-resolution in the downcore direction for each of the core segments. 2.4 Carbonate isotope geochemistry Carbonate mass spectrometry samples have not been measured systematically in core 5B, however, high-resolution samples in varved intervals between 60.7 - 57.6 (Appendix B) and 1.5 – 0 RMCD (Shanahan, 2006; Shanahan et al., 2009), as well as bulk sediment samples taken from carbonate-rich intervals, primarily above 50 RMCD have been analyzed for δ18O and δ13C isotope geochemistry. Dried sediments were heated under vacuum at 100°C for several hours to remove volatile organic compounds prior to measurement. The δ18O and δ13C of these carbonates were measured at the University of Arizona, using an automated carbonate preparation device (KIEL-III) coupled to a gas-ratio mass spectrometer (Finnigan MAT 252). Powdered samples were reacted with dehydrated phosphoric acid under vacuum at 70°C. The isotope ratio measurement is calibrated based on repeated measurements of NBS-19 and NBS-18 and precision is ±0.10‰ for δ18O and ±0.08‰ for δ13C (1 sigma). The δ18O of authigenic carbonates in Lake Bosumtwi is a function of the lake water δ18O (δ18Olw), the water temperature, and the carbonate mineralogy. Whereas changes in lake water temperature probably exert a minor influence, the extremely variable carbonate mineralogy in Lake Bosumtwi complicates the interpretation of past changes in the δ18Olw. At least six carbonate minerals are present in the sediments (calcite, high Mg-calcite, aragonite, manganosiderite, dolomite and ankerite), which all fractionate 47 oxygen differently during formation. Furthermore, in the cases of calcite and high Mgcalcite, and dolomite and ankerite, the composition of the carbonates exist on a continuum, varying between two end members, resulting in variable fractionation factors even within individual mineralogies. To account for this, we calculate δ18Olw for each sample based on its mineralogy, over a wide (10°C) range in temperatures. This results in a fair amount of uncertainty, however the variability in carbonate δ18O in Lake Bosumtwi is extremely large (>10‰), so large-scale trends and rhythms are still apparent. That said, there are still several problems with the approach. First, the mineralogy of each sample, much less its precise elemental composition, has not been determined. Furthermore, whereas fractionation for most of these carbonates have been studied empirically, only theoretical estimates are available for ankerite, and minerals along the ankerite-dolomite series (Chacko and Deines, 2008). For this study, we have adopted a simplified approach for adjusting for mineralogy. First, manganosiderite-rich intervals are readily identified by high Mn peaks in XRF data. No other mineral has been identified in the sediments that corresponds to Mn intensities higher than 1000 cps. Dolomite and ankerite are identified by their very high δ18O and δ13C values (Appendix B; Talbot and Kelts, 1986). No other carbonate mineral in the lake has been identified that has δ18O higher than 7‰ and δ13C higher than 15‰. Moreover, neither dolomite or ankerite has been found with δ18O or δ13C lower than those values. Samples which do not appear to be either manganosiderite or dolomite/ankerite are classified as calcite/Mg-calcite/aragonite. There are differences in the fractionation factors of these three Ca-carbonates, however they most likely result in a maximum offset 48 of ~1‰ (Tarutani et al., 1968; Kim and O'Neill, 1997). Because we cannot reliably distinguish the minerals in the samples that have been analyzed, we treat calcite, high Mg-calcite and aragonite as calcite. The differences in fractionation between dolomite and ankerite are larger, likely up to 4‰, but are extremely uncertain (Chacko and Deines, 2008). We treat all of these samples as dolomite, likely resulting in an underestimation of δ18Olw for the more ankeritic carbonates. The dolomites and ankerites are consistently formed in the most 18O-enriched lake waters, so this misrepresentation results in a more conservative estimate of δ18Olw during lowstands, and ankerite-rich intervals are interpreted cautiously. Ultimately, δ18Olw is estimated using the simplified mineralogies are used to account for the different phosphoric acid fractionations (Rosenbaum and Shephard, 1986; das Sharma, 2002), and water-carbonate fractionations (Carothers et al., 1988; Kim and O'Neill, 1997; Vasconcelos et al., 2005) for each mineral over a 10°C range in water temperature. 2.5 Turbidites The depths and thicknesses of turbidites greater than 1 cm in thickness were visually identified for the entire length of core BOS04-5B. For the upper 159 meters of the core, both core face images and high-resolution x-radiographs were used to identify turbidites, which are generally apparent as fining upward packages of homogenized sediment with a distinct, dense, light-gray clay cap (Figure A5). 49 2.6 Laminations Two approaches were used to estimate downcore variability in the presence and quality of laminations in the sediment. First, core face images were used to visually rank the laminations for consecutive 10-cm intervals throughout all of core BOS04-5B. Each interval was subjectively deemed well-laminated (lamination code {LC} = 3), moderately-laminated (LC = 2), poorly-laminated (LC=1), or unlaminated (LC=0). A second, objective approach used the x-radiographs that span the upper 159-m of core BOS04-5B to estimate changes in the presence of laminations by calculating the ratio of vertical to horizontal variance in consecutive 5-cm intervals of the x-radiograph (Figure A6). Well-laminated intervals have a lot of vertical variance in the x-radiograph, and very little horizontal variance, resulting in high vertical to horizontal variance (V:H) ratios. Alternatively, massive intervals have little variance in both directions, resulting in much lower V:H ratios (Figure A6). Gaps in the sediment cores can greatly affect this calculation, and were removed prior to calculation. 3. Results 3.1 Geochronology To model the age-depth relationship in the drill core and assign ages to depths over the million-year record we used the “CLassical Age Modelling” (CLAM) approach of Blaauw (2009). In this technique an ensemble of splines are fit to the ages in the record, with the the ages for each ensemble member selected stochastically from their probability distribution functions. For 14C ages, the full calibrated probability 50 distributions are used, for all other ages normal distributions are used with their analytical standard deviations (Figure A4). There is no evidence for reversals in core BOS04-5B, therefore we removed all of the ensemble members with age reversals. Because turbidites make up nearly 25% of core BOS04-5B, and they are deposited extremely rapidly (on the order of hours to days) when compared to deep-water sedimentation at the site, we removed turbidites from the depth scale before developing the age model. By fitting smooth splines through the ages, a built-in assumption with CLAM is that sedimentation rate changes gradually between age-control points. This approach certainly underestimates variability in sedimentation rate, but by removing turbidites from the depth scale, we avoid the biases that would result from the gradual decrease in turbidite abundance with depth. The ages and uncertainties of all the dates, as well as the ensemble envelope and median estimate from CLAM are shown in Figure A5. The density and precision of geochronometers decreases with depth in the core, and this is reflected in the spread of model ensembles. The age model is well-constrained for the upper 30 m (RMCD) of the core by radiocarbon and OSL ages, however below 30 m there are far fewer ages, with greater uncertainties. Below ~ 75 m (RMCD) the few U/Th ages in the interval have large error bars, and this uncertainty is reflected in a wide range of age model ensembles. Geomagnetic reversals and the impact age reduce this uncertainty at the base of the core, however, the range of age model ensembles exceeds 100 kyr for much of the middle of the drill core. This range of uncertainty most likely precludes the possibility of useful millennial-scale comparison between paleohydrological indicators at Lake Bosumtwi and 51 other regional and global records of climate variability during the late Pleistocene. Therefore, we examined the potential of “tuning” the poorly-dated interval of the record to another record, and ultimately used a proxy for dust content in the sediments to tune the record to glacial-interglacial changes in the EPICA Dome C dust record. To minimize overfitting or confirmation bias, we rely on the well-dated past 200 kyr to validate any potential matching between records. To be acceptable, any match must be mechanistically sound, and clearly covarying and persistent over the most recent 200 kyr. 3.1.1 Comparisons with regional and global dust record Dust export from the Sahara and Sahel varied dramatic during the Pleistocene, generally in phase with other records of dust variability throughout the world in response to large-scale changes in global climate (Tiedemann et al., 1994; deMenocal, 1995; Moreno et al., 2002; Larrasoaña et al., 2003; Moreno 2012). It is reasonable to assume that glacial-interglacial-scale changes in Saharan dust flux are recorded in the Lake Bosumtwi sediments as well, and provide a potential means of synchronizing the records. The ratio of coercivity (Hc) to the coercivity of remanence (Hcr) in the Lake Bosumtwi sediments is indicative of the magnetic grain size of the sediment (Peck et al., 2004). Intervals with high (Hc/Hcr) ratios are associated with hematite and goethite, which are delivered to the lake as aerosols from the Sahara and the Sahel (Peck et al., 2004). Over the past 200 ka, Hc/Hcr generally covaries with dust flux deposited in the Gulf of Guinea (ODP 663; Figure A7; deMenocal et al., 1993) and off the west coast of North Africa (ODP 659; deMenocal, 1995). However, it is difficult to establish robust tie points 52 between the records because of the low resolution of marine cores. Interestingly, dust flux recorded in the EPICA Dome C (EDC) Antarctic ice core matches fairly well with Lake Bosumtwi dust content, especially when comparing the general presence and absence of dust, rather than the amplitude of the peaks (Figure A7). This similarity continues over the entire 800 kyr of EDC core (Figure A8). Obviously, the source areas and winds that deliver dust to Antarctica and Lake Bosumtwi are in different hemispheres, and may respond differently to even global-scale climate changes. Glacial intervals, however, are characterized by much higher dust flux in both hemispheres (deMenocal et al., 1993, deMenocal, 1995, Lambert et al., 2008). In West Africa, the increase in dust flux is associated with drier conditions in North Africa, and an intensification of the winter trade winds, both of which occur during cold intervals in the North Atlantic (deMenocal and Rind, 1993; Moreno, 2012). At EDC, the dramatic increase in dust during glacial periods is believed to be associated with increased aridity in South American source regions, and a longer lifetime of aerosols in a drier atmosphere (Lambert et al., 2008). The southern westerlies were likely slower during glacial intervals, however they were also shifted northward, potentially enlarging the South American source areas (Toggweiler and Russell, 2008). Dust flux in both regions is sensitive to glacial-interglacial changes in Atlantic SSTs and probably underwent similar changes at similar times during the Pleistocene. This, combined with the good visual match between the high-resolution records allows us to “tune” the Lake Bosumtwi chronology to the commonly-used EDC chronology. 53 3.1.2 Tuning to EDC timescale To roughly synchronize Bosumtwi dust record (Peck et al., in prep) to the EDC dust record (Lambert et al., 2008), we picked 14 tie points at glacial-interglacial transitions over the past 800 kyr (Figure A8). Because the uppermost sediments are well dated, we did not allow any adjustment to the chronology between 0 and 70 ka. Given the tie points, there appears to be no drift or systematic bias in the Lake Bosumtwi chronology. Five of the tie points were younger in the Lake Bosumtwi chronology, and eight were older (the tie point at 70 ka is the same in both records). The age model was then adjusted by stretching, or condensing the curve to fit the tie points, no other tuning was performed. This approach preserves any relative changes in sedimentation rate inferred by the original model (Figure A4). The tuned age model falls within the bounds of the ensemble envelope until near the Brunhes/Matuyama reversal where the ensemble narrows, and the tuned model slightly exceeds the envelope (Figure A4). Generally, the tuning increased sedimentation rates during glacial intervals and reduced sedimentation rates during interglacials. This is consistent with the hypothesis that interglacials are generally associated with higher lake level at Lake Bosumtwi, since the basin area, and sedimentation rates are reduced when the lake is high (Shanahan et al., 2008; 2009). After tuning on glacial-interglacial transitions, the power spectra of lake level indicators show significant variability at orbital (i.e., eccentricity, obliquity, and precession) timescales (Figure A9). The presence of 100-kyr cycles could be a function of the tuning, however the occurrence of significant peaks at the independent obliquity and precession timescales provide support for the tuning procedure (cf., Imbrie et al., 1984; 54 EPICA, 2004). 3.2 X-ray fluorescence The relative abundance of Si, K, Ti, and to a certain extent, Fe (Figure A10) are interpreted as terrigenous elements, following Shanahan et al., (2009). The primary source of these elements in the lake sediments is terrigenous material that washes in from the drainage basin during the rainy season, and they comprise a major component of annual sedimentation in varves in both the late Holocene (Shanahan et al., 2008; 2009) and the Last Interglacial (Appendix B). Because the terrigenous elements are washed in from the lake basin, their relative abundance is strongly influenced by the size of the drainage basin and the proximity of the core site to the shoreline, both of which are primarily functions of lake level. Therefore, the terrigenous elements are expected to vary inversely with lake level; most abundant when the lake is low, and least when it is high. The abundance of terrigenous elements is a robust lake level indicator for the late Holocene, covarying with carbonate oxygen isotope values, geomorphic lake level indicators, and a variety of other forms evidence (Shanahan et al., 2009). Fe is also a terrigenous element, and is an important component of the sediment that washes into the lake. However, Fe is also present in minerals precipitated in the lake, or post-depositionally in the sediment, primarily manganosiderites and various species of iron-sulfide minerals. This complicates the interpretation of the Fe intensities, and we consider them separately from the other terrigenous elements. Mn in the sediment record is present in several minerals, but is predominately associated with manganosiderites. Mn 55 is also associated with iron-sulfide minerals and ankerite, but the intensity of the Mn peaks associated with these minerals is much lower (<100 cps) than with the manganosiderites (>1000 cps). Therefore, we consider Mn to be a robust indicator of the presence of manganosiderite. Ca XRF intensities primarily indicate the presence, and relative abundance of various species of calcium carbonates in the sediments. There is very little Ca in the bedrock of the drainage, which is predominantly schist with small outcrops of granite (Koeberl et al., 1998), and the low intensities of Ca in the terrigenous layers of the sediment are consistent with this observation. The Ca that is incorporated in authigenic carbonates in the sediments is most likely originally delivered to the lake as dust. Dust deposited by northerly Harmattan winds is an important component of Ghanaian soils (He et al., 2007), and is the likely origin of much of the dissolved Na in Lake Bosumtwi (Turner et al., 1996b). Finally, the relative abundance of S in the sediment shows pronounced variability, and is present in three distinct components of the sediment. For most of the sediment, S XRF intensities are low, and somewhat noisy, but generally track the abundance of organic matter in the sediment. Higher S intensities are associated with various the various iron-sulfide minerals (pyrrhotite), and less frequently, diagenetic gypsum. 3.2.1 Correcting for biases in the XRF data Scanning XRF analysis of wet, split sediment cores allows for rapid and highresolution determination of relative elemental abundance, and is a particularly useful tool 56 for long sediment sequences. However, XRF intensities can be affected by a number factors, especially when applied over highly variable lithologies, such as those present in the core BOS04-5B. One favored approach in dealing with these uncertainties is to examine elemental ratios; however, in Lake Bosumtwi sediments, the absolute abundance of terrigenous elements is closely related to lake level (Shanahan et al., 2009). In this section, we correct for changes in sediment water content and compaction of the core with depth. 3.2.1.1 X-ray absorption and attenuation by water One challenge associated with XRF scanning wet sediments is evaluating the effect of x-ray absorption by interstitial water and water condensed on the film used to reduce dessication of the cores during analysis (Kido et al., 2006, Tjallingii et al., 2007). Intervals with elevated water content have consistently lower XRF counts, and the relative change in counts varies between elements such that lighter elements are more greatly affected. Both empirical and theoretical corrections have been proposed and applied to this problem, in both cases the goal is to estimate the counts that would have been collected from dry sediment. Because we do not have the XRF measurements of packed, dry sediment necessary to calibrate the record, we use the theoretical correction of Kido et al. (2006) which uses the known mass attenuation constants of the different elements to calculate the impact of x-ray absorption and scattering by water. The correction requires knowledge of the bulk sediment composition and the water content. Both of these variables can calculated from the XRF data themselves, so we employ the 57 correction iteratively, first calibrating the data, then correcting the XRF intensities, then recalibrating with the adjusted intensities and reapplying the correction until both the calibrations and adjusted counts stabilize. The bulk sediment composition is estimated by simple 3-point calibrations with standard reference materials (SRMs) that are routinely measured on the Itrax core scanner at the University of Minnesota, Duluth. The resulting compositional estimates are probably not very accurate, however uncertainty in the composition has only a minor effect on the correction of Kido et al. (2006), because the differences in mass attenuation values for the primary sedimentary constituents are small. On the other hand, the determination of water content results in more uncertainty in the correction. The water content of core BOS04-5B has been calculated from the dry and wet bulk densities measured using on the Geotek Core Scanner, and by loss-in-ignition, respectively. The ratio of coherent (Rayleigh) to incoherent (Compton) scattering (coh/inc) is strongly related to bulk density, and is calibrated to sediment water content. Unfortunately, the water content values used for calibration were measured immediately after the core was split, whereas the XRF data were collected several years later. This means that the calibration likely overestimates the water content at the time of measurement, since the cores had dried somewhat during storage. We estimate the amount of drying as about 30% based on comparison of the cores at the time of XRF analysis with core photos taken after splitting, and therefore multiply the original water content values by 0.7. This choice is somewhat arbitrary, however allowing this value from 60 to 80% results in only subtle changes, especially with regard to relative variability within each estimate (Figure A11). 58 3.2.1.2 Sediment compaction with depth Unconsolidated sediment compacts as a function of the increasing weight of overlying material with depth. In marine and lacustrine sediments, this results in a downcore increase in density and decrease in porosity that changes most rapidly near the surface (Bahr et al., 2001). Because XRF intensities increase with density, it is worthwhile to account for the downcore compaction component of changes in wet bulk density (WBD) of the sediment. This is generally done empirically, by fitting an exponential (Bahr et al., 2001), logarithmic (Sheldon and Retallack, 2001), or low-order polynomial (Huybers and Wunch, 2004) to measured WBD or porosity data. These approaches work reasonably well for cores with relatively little downcore variability in WBD, such as most marine sediment cores. Core BOS04-5B, however, contains greatly variable lithologies and WBD throughout the 300 m interval (Pers. comm. J. King). Attempting to fit an exponential, logarithmic or polynomial to the WBD results in a compaction curve that either overfits the data, resulting in reversals in the compaction curve, or fits poorly in the upper 50 meters of the core, the section most strongly affected by differential compaction. Therefore we adopt a piecewise approach, and fit a piecewise 2nd and 1st order polynomial to the data. To do this, we use a least-squares fitting algorithm to fit the uppermost part of the curve with a quadratic polynomial, and then at a depth determined by the algorithm, the function becomes linear with the slope of the quadratic at the depth where the function switches. This approach has several advantages. First, it allows the uppermost part of the curve to be fit well, while maintaining the structure of a simple compaction curve with no negative slopes. Second, the algorithm is 59 non-parametric, in that the only user input was the type (2nd and 1st-order polynomials) of the function. Secondly, all of the parameters for the equations, including the depth at which they switch, are chosen by the algorithm to best fit the WBD data. The relationship between density and XRF counts is complex, varying among different elements and sediment compositions, however over the relatively small range in density associated with the core compaction curve, the relationship is probably nearly-linear. Therefore, we adjust the XRF intensities for downcore compaction by dividing the intensities by the fit compaction curve. 3.2.2 Principal components analysis Principal components analysis (PCA) is a technique that can be used to identify and consolidate the leading modes of variability shared between multiple time series into several, uncorrelated records (Ghil et al., 2002). The first principle component (PC) of late Holocene XRF data is strongly associated with the abundance of terrigenous material, and tracks lake level over the past 2700 years (Shanahan et al., 2009). We performed PCA on the water-content and core-compaction adjusted XRF intensities of Al, Si, K, S, Ca, Ti, Mn, Fe, Sr, and Rb. The first PC of the XRF data spanning the past 530 ka (XRF-PC1) explains 39% of the variance, and is strongly associated with the terrigenous elements Al, Si, K, Ti and Rb (Table 1; Figure A12). The other elements are primarily loaded on other PCs, except for Mn, which loads negatively on PC1. PC2 explains 28% of the variance, and is dominated by variability in carbonate mineralogy, with Ca and Sr associated with the various Ca-carbonates in the sediment, and Mn and Fe 60 associated with manganosiderite. The results of the PCA are very similar to those for the late Holocene varve record (Shanahan et al., 2009), and the whole Holocene XRF record (Shanahan, 2006) from Lake Bosumtwi. 3.2.2.1 Carbonate dilution of XRF-PC1 XRF-PC1, because it is so strongly driven by changes in the terrigenous content of core BOS04-5B, likely responds to changes in the availability of terrigenous sediment supply, and terrigenous drainage area, which at Lake Bosumtwi are primarily functions of lake level (Shanahan et al., 2008; 2009). Because we interpret XRF-PC1 primarily as a lake level indicator, it is important to identify and account for processes that affect XRFPC1 that are not necessarily related to changes in the influx of terrigenous sediment. Dilution of terrigenous material by carbonate is one such process. Figure A13 shows the correlation between XRF-PC1 and Ca, and Mn, both of which track the abundance of different carbonate minerals in the sediments. The precipitation of these carbonates can be related to changes in lake level, however, it is much more likely that anti-correlation of terrigenous material with Ca and Mn (Figure A13) is due to dilution, rather than a prolonged reduction of terrigenous influx into the lake. To avoid misinterpreting these decreases in XRF-PC1, we remove this dilution effect by fitting a line to the Ca-PC1 and Mn-PC1 trends where Ca and Mn are particularly abundant. The relationship between Ca and Mn and XRF-PC1 is not significant at low Ca and Mn abundances (i.e., where there is little or no carbonate). Consequently, we only adjusted XRF-PC1 above intensities where the anti-correlation is statistically significant (above 260 and 160 cps, respectively; 61 Figure A13). The corresponding XRF-PC1 data are then adjusted, following the fit line. The adjusted correlations are shown in Figure A13. This adjustment has a minor impact on most of the data points in the record. Forty-one percent of the XRF data are adjusted, however the adjustment is very small for most points, only 12% are adjusted by more than one unit, and 5% by more than two. The net effect of all of the corrections applied to the XRF data can be examined by comparing Figures A10a and b. The primary difference is the removal of the long-term decrease in counts towards the top of the core, the smaller-scale and deeper variability in each element is generally unaffected. For the most part, the corrections have little influence on the main conclusions. However, they do influence the interpretation of the inferred wetness of the past 75 kyr relative to the rest of record. This interval is most influenced by the lack of sediment compaction, and lake level appears artificially high in the uncorrected data. 3.3 Lamination indices The subjective lamination index and the logarithm of the V:H density variance ratio are generally consistent over the upper 160 m of core BOS04-5B (Figure A14). Taken together, the two indices reveal significant variability in the presence of well preserved laminations, which are present for most of the upper half of the sediment sequence. Laminations are formed when the lake floor is anoxic, limiting bioturbation. XRF Mn and Ca intensities, indicative of carbonate abundance in the sediment, generally track the lamination indices, and are more abundant when laminations are present. 62 Carbonates often form fine laminations in the sediment, so the two indicators are not independent. Much of the carbonate found in the sediments from the deep-water core sites is interpreted to have formed in anoxic water, at or near the sediment-water interface (Talbot and Kelts, 1986). Covariance between carbonate abundance and laminations indices supports the hypothesis that the indices record variability in the oxygenation state of the lake bottom waters. 3.4 Turbidites Turbidite abundance generally decreases upcore, from an average of 41 cm per meter over the bottom 50 m to an average of 3 cm per meter for the upper 50 m (Figure A15). This likely reflects lake basin evolution the past million years (Peck et al., 2006; in prep). Early in the lake's history, the crater sides were steeper, and the crater floor was 300 m deeper. As the margins eroded and filled the basin with lake sediment, the relief decreased, as did the probability of mass wasting events. Considerable shorter-term variability is overprinted on the gradual decreasing trend, suggesting that lake level and climate variability also influenced the turbidite abundance. Over the past 530 ka, this shorter-term variability in turbidite abundance shows a remarkable correspondence with Mn abundance (Figures A16 and A17), as well as organic matter content in the core (pers. comm C. Scholz). Mn indicates the presence of manganosiderite, in the interval, which along with organic carbon, is an indicator of high lake level. It is important to note that the manganosiderite does not occur in the turbidites, nor is it consistently deposited immediately before or after a turbidite. Therefore, we do not interpret this co-occurrence 63 as a direct cause and effect relationship. That is, the manganosiderite is not being deposited because a turbidite was deposited. A lag correlation analysis (Figure A17) suggests that on average, Mn deposition leads turbidite deposition by 1000-3000 years. We hypothesize that turbidites are associated with terminations of highstands. During a sustained highstand, deltas and terraces build up high on the crater walls, as lake level falls, these deposits become unstable and prone to mass wasting, leading to increased turbidite abundance. 4. Discussion 4.1 Past 530 ka 4.1.1. 400-kyr eccentricity forcing The magnetic hysteresis parameter Hc:Hcr, which is primarily a proxy for dust content in Lake Bosumtwi, is shown along with EDC dust concentration in Figure A16 (Lambert et al., 2008, Peck et al., in prep). The similarity between these records were used to “tune” the Bosumtwi chronology at glacial-interglacial boundaries, so leads or lags between the records cannot be evaluated. The two records often covary ten thousand year timescales, likely tracking changes in dust flux between stadials and interstadials. The greatest mismatch between the records is found between 250 and 150 ka, where Hc:Hcr is generally much higher than EDC dust. This suggests increased dust flux in West Africa, relative to the high latitudes of the southern Hemisphere, during most of Marine Isotope Stages (MIS) 6 and 8 relative to other glacial intervals during the past 530 ka. During particularly arid conditions greigite forms in Lake Bosumtwi sediments, 64 which also increases Hc:Hcr (Peck et al., 2004). Regardless, the mineral magnetic indicators suggest particularly arid conditions during this interval. The XRF-PC1 record, an indicator of terrigenous content and lake level at Lake Bosumtwi, also suggests generally dry conditions between 300 and 100 ka. This interval of lower lake levels, combined with wetter conditions from 530 to 300 ka, and since 100 ka suggests changes in background aridity in the region with a periodicity of ~400-kyr (Figure A16). To compare the XRF-PC1 records to the 400-kyr of orbital eccentricity, we isolate the 400- and 100-kyr components of eccentricity over the past 1.1 Myr using singular spectrum analysis (SSA), a non-parametric technique that extracts the primary modes of periodic or semi-periodic variability from a timeseries (Ghil et al., 2002). The 400- and 100-kyr components of eccentricity make up 41 and 54% of the variance in eccentricity over the past 1.1 Myr, and are shown in Figure A16. XRF-PC1 shows a close correspondence to 400-kyr component of orbital eccentricity; the dry interval corresponds to the interval of highest eccentricity over the past 530 kyr. Deeper in the core, variability in the presence of laminations (Figure A18) and organic carbon content (not shown; pers. comm C. Scholz) suggest a similar trend during the penultimate 400-kyr eccentricity cycle. Superimposed on the 400-kyr change in background state is shorter-term variability that corresponds to the 100-kyr component of eccentricity, but with the opposite sign (Figure A16). Whereas wet conditions are associated with low eccentricity on the 400-kyr timescale, they correspond with high-eccentricity on the 100-kyr timescale. This complex relationship likely results from the interplay of local orbital 65 forcing with glacial boundary conditions. We suggest that the inverse relationship between Lake Bosumtwi hydrology and 400-kyr-scale eccentricity is driven by eccentricity's modulation of precession, following the hypothesis of Scholz et al. (2007). During intervals of high eccentricity (e.g. 100-300 ka), the amplitude of seasonal insolation changes on precessional timescales is substantially increased. Both the West African and Asian Monsoon systems intensified and expanded northward during the high-eccentricity Last Interglacial (Miller et al., 1991; Szabo et al., 1995; Wunneman et al., 2007; Wang et al., 2007, Drake et al., 2011). An intensified and northward shifted monsoon would impact the hydrology at Lake Bosumtwi (and more generally, the Guinea Coast) in several ways. Lake Bosumtwi presently has a double peak in precipitation, associated with the northward passage of the rainbelt in May, June and July, and the southward passage during September and October (Figure A2). July and August are dry; the core of the rainbelt is situated to the north and intensified southeasterly trades increasing upwelling along the Guinea Coast, lowering SSTs in the Gulf of Guinea, and suppressing convection over the Guinea Coast (Moron, 1994; Opoku-Ankomah, 1994; Nicholson, 2009b). The WAM was probably most intensified and north-shifted when an eccentricity-amplified precession cycle drove insolation maxima during July and August. Under such conditions, the rainbelt would most likely spend more time north of the Guinea Coast, shifting the first rainy season earlier in the year and the second rainy season later in the year. This would move rainy seasons away from the peak monsoon season, and intensify the southeasterly trades during July and August, lengthening the summer dry season at Lake Bosumtwi. This longer dry interval occurring at the same 66 time as the season of highest solar insolation would also drive increased evaporation from the lake. Redistribution of tropical moisture from equatorial regions towards the subtropics on 400-kyr timescales is also observed in anti-phased dust records from the Atlantic and Mediterranean (Tiedemann et al., 1994; Larrasoaña et al., 2003), consistent with our observations from Lake Bosumtwi. The hypothesis described above could explain dry conditions during precession minima (JJA insolation maxima), but not the generally dry conditions throughout the rest of a precession cycle. High-resolution studies of the Holocene (Shanahan, 2006) and the Last Interglacial (Appendix B) suggest that the wettest conditions at the lake occur when the annual insolation maximum occurs during one of the two rainy seasons (i.e, June-July or September-October). However, if the mean position of the rainbelt moves north and south in response to precessional forcing, seasonality of precipitation would vary in counterpoint with seasonality of maximum insolation maximum (Figure A19). For example, when insolation peaks in midsummer, the rainbelt would cross the Guinea Coast earlier and later in the year, moving rainy seasons away (temporally) from the season of highest insolation. When insolation peaks in May or October, the northward extent of the WAM is probably reduced, and the rainbelt would cross the Guinea Coast later in the season, moving the rainy season closer to mid-summer, but once again, away from the season of highest insolation. Following this model, twice during each precession cycle the season of insolation maximum will coincide with one of the rainy season, consistent with what is observed in the Last Interglacial and the Holocene. However when eccentricity is higher, seasonal differences 67 in insolation are amplified, and the period when insolation maxima and the rainy seasons coincide may be much shorter (Figure A19). If true, this implies that for most of a precession cycle, the season of highest insolation occurs during a dry season, and maximum seasonal insolation is higher at higher eccentricities, potentially resulting in increased evaporation from the lake. 4.1.2 100-kyr eccentricity forcing and glacial boundary conditions The 100-kyr periodicity of Lake Bosumtwi lake level is most likely due to changes in glacial boundary conditions. The West African Monsoon is sensitive to the interhemispheric temperature gradient in the Atlantic Ocean, and sea surface temperatures (SSTs) in the North Atlantic on timescales ranging from annual to millennial (e.g., Mulitza et al., 2008; Shanahan et al., 2009; Stager et al., 2011). Sustained periods of cold SSTs in the North Atlantic, most notably Heinrich Events, have a dramatic influence on African hydrology, driving drought at Lake Bosumtwi (Peck et al., 2004) and many other sites in Africa (Collins et al., 2010; Stager et al., 2011). Periods of low-eccentricity on the 100-kyr timescale correspond with deep glacial conditions in the North Atlantic, and drought at Lake Bosumtwi, most likely through the influence of cold Atlantic SSTs. 4.1.3 West African contribution to global atmospheric methane concentrations Obliquity and precession-scale periods are also observed in the power spectra of XRF-PC1, Mn and Ca (Figure A9). The role of precession can be observed in the 68 similarity between atmospheric methane and Mn and Ca abundance in the sediments (Figure A16d). Tropical West Africa is an important methane-producing region (Lelieveld et al., 1998; Frankenberg et al., 2005, Singarayer et al., 2011), and variability in the West African Monsoon drives changes in the extent and moisture balance of African wetlands. Precession-scale variability in ice-core methane records is commonly attributed to changes in Northern Hemisphere monsoon systems (e.g., Ruddiman and Raymo 2003, Loulergue et al., 2008) and this hypothesis is consistent with records of the Asian Monsoon inferred from speleothems over the past 380 ka (Wang et al., 2008; Cheng et al., 2009). The Lake Bosumtwi record offers the first opportunity to evaluate the role of West Africa with a moisture-sensitive record form a terrestrial site over the past 530 kyr. XRF-PC1, and especially the abundance of Mn and Ca in the sediments show a close correspondence with atmospheric methane. Mn and Ca track carbonate mineralogy in core BOS04-5B, Mn is almost exclusively associated with manganosiderite, and Ca is associated with the calcium carbonate minerals calcite, aragonite and dolomite. Manganosiderite and the calcium carbonates are not deposited at the same time, rather, the ratio of dissolved Fe2+ and Ca in the lake water controls which type of carbonate is deposited (Berner, 1971). We interpret manganosiderite as a wet indicator for several reasons. To reach sufficiently high Fe/Ca ratios (i.e., > 0.05; Berner 1971), most likely requires a sustained anoxic conditions at the lake floor, supported in part by increased deposition of organic material, and a reduction of eolian Ca supply to the lake. These conditions are consistent with high lake level and a generally wetter climate. This is also supported by other evidence: during the Holocene manganosiderite was only deposited 69 when the lake was very high, and other lake level indicators, such as varve thickness (Appendix B), δ18O of carbonates (Figure A20) and δD leafwax (pers. comm. T. Shanahan) all confirm that manganosiderite is only deposited during relatively high lake level. Manganosiderite was not deposited when the lake was overflowing during the early Holocene, probably due to a reduction in bicarbonate driven by a freshening of the lake. The majority of carbonates in Lake Bosumtwi sediments are formed near or just below the sediment-water interface under anoxic conditions, their formation partially driven by methanogenesis (Talbot and Kelts, 1986). Consequently, carbonates are rarely found in unlaminated intervals when the bottom waters were at least seasonally oxygenated. Perhaps, during intervals of lower lake level, calcium carbonate deposition primarily occurs nearer the shoreline. This is consistent with the observation of abundant calcite nodules in intervals of the shallow-water drill cores (pers. comm. J. Peck). Therefore, the range of calcium carbonate minerals comprising the XRF Ca record are interpreted as indicating intermediate lake level. Calcite and aragonite are probably indicative of higher lake levels than Mg-calcite and dolomite, but these cannot be easily distinguished in the XRF data. Intervals with abundant Mn are consistently associated with high atmospheric methane concentrations, and peaks in Mn often occur during precessional-scale peaks in CH4 (Figure A16d). Typically, smaller peaks in CH4 correspond to peaks in Ca, although some of the high methane peaks during MIS 7 and 5 are associated with Ca, rather than Mn. This is part of the general trend of lower lake level during the high-eccentricity interval from 300-100 ka, where Mn is less common. The influence of the 400-kyr 70 eccentricity component is clear during low-CH4 intervals as well. Calcium carbonates were deposited throughout most of MIS 12, 10 and 2, however, there are prolonged intervals of unlaminated sediment with no carbonates during parts of MIS 8 and 6, suggesting very dry conditions at the lake. The close correspondence between the Mn and Ca and EDC CH4 over the past 530 ka suggests that the contribution of CH4 from sub-Saharan Africa is generally in phase with other Northern Hemisphere methane and sources. This conclusion could be biased by the fact the chronologies are not independent, however the two appear in phase during the well-dated and untuned most recent 50 ka, and it seems improbable that the close correspondence of high-frequency peaks is spurious, given that the dust records were only synchronized on glacial-interglacial boundaries. On longer timescales, however, tropical West Africa seems to have had a variable contribution to atmospheric methane, as the conditions at the lake were consistently drier between 300 and 100 ka than at other times. The 400-kyr eccentricity cycle is not an important component of atmospheric methane variability over the past 800 kyr (Lambert et al., 2008). This likely means one of two things. First, it is possible that this pronounced 400-kyr variability is restricted to the Guinea Coast region, which provides a relatively small contribution to the global methane budget; the modelling study of Singarayer et al. (2011) suggests that sub-Saharan North Africa is responsible for about 5-10% of global emissions. If true, low-frequency changes in regional methane production might be overwhelmed by the rest of the global production. However, an expansion of the rainbelt, and increased precipitation in the subtropics at the expense of the tropics would affect a much greater 71 proportion of the global methane producing regions. This possibility is supported by other long equatorial paleohydrological records that record drier conditions in past higheccentricity intervals (i.e., the Last Interglacial Period), in both Africa (Stone et al., 2011) and South America (Bush et al., 2004, Hanselman et al., 2005). In this case, the reduced contribution of tropical wetlands during high-eccentricity intervals may be compensated by increases in subtropical moisture. Despite the fact that precession-scale variability in moisture balance at Lake Bosumtwi appears to be in phase with global CH4 concentrations, the influence of eccentricity and glacial-interglacial boundary conditions on regional hydrology highlights the complexity of the influence of orbital precession on the WAM. Global CH4 emissions are partly driven by hydrological variability in disparate regions, that each have their own, often complex relationship with orbital forcing (Singarayer et al., 2011), and for more equatorial regions, these relationships remain largely unknown. Therefore, it seems premature to assume that an in-phase relationship between Northern Hemisphere monsoon systems and precession was present throughout the late Quaternary, or to use that assumption to adjust a geochronology (cf., Ruddiman and Raymo, 2003), especially since the late Holocene (6ka to preindustrial) increase in methane does not appear to have resulted from anthropogenic activities, invalidating the relationship between methane and Northern Hemisphere insolation in the Holocene (Singarayer et al., 2011). 4.2 Latitudinal variability in West African climate over the past 150 ka Over the past 150 kyr, moisture balance at Lake Bosumtwi, as recorded by XRF- 72 PC1 and δ18Olw, is generally consistent with ocean core records of terrestrial moisture variability from West Africa (Figure A20). A latitudinal profile of the available records that have been interpreted to record West African moisture variability over the past 150 kyr indicates a latitudinally varying response to North Atlantic boundary conditions (i.e., temperature and ice volume) and local insolation forcing. Strong similarities with the NGRIP δ18O record in the northern-most West African records gradually become more similar to precessional-scale changes in insolation at lower latitudes (Figure A20). The variable response of the records during two high-obliquity and high-precession (low NH insolation) intervals from about 100 to 85 ka and 55 to 40 ka are particularly interesting. During both intervals obliquity was high and Greenland temperatures were relatively warm, whereas summer insolation at 10°N was low. During the interval between ca. 50 and 40 ka, the most northerly records (> ~10°N) appear near-average in terms of moisture balance, whereas the Sahelian records suggest that effective moisture during this interval was high, in some cases comparable with the Holocene (deMenocal, 1995; Zabel et al., 2001; Castañeda et al., 2009). Records sensitive to moisture in more southerly areas, such as the Congo and Sanaga river basins (Schneider et al., 1997; Weldeab et al., 2007), show very dry conditions from ~50-40 ka, tracking low-latitude NH summer insolation. Lake Bosumtwi lies on the boundary between these two responses; XRF-PC1 suggests wet conditions from about 48-43 ka, right through the minimum in local summer insolation. This is, however superimposed drier conditions for the rest of the interval. A similar orbital configuration occurred ca. 90 ka, however the West African response was quite different (Figure A20). The influence of increased obliquity is evident 73 in several of the more northerly records, but, all of the records suggest below average or very dry conditions during the interval, especially centered on the low latitude Northern Hemisphere summer insolation minimum ca. 95 ka. Extreme aridity during this general interval (115 – 90 ka) is a common feature of paleohydrological records from both East and West Africa, to latitudes as far a south as the Northern Kalahari Desert (Scholz et al., 2007; Blome et al., 2012). AMOC during the interval appears to have been comparable to 50 ka (Castañeda et al., 2009; Lopes dos Santos et al., 2010), so the most likely cause of the differing responses is the difference in amplitude of the precessional forcing. The very low summer insolation (15-20 W m-2 less than 45 ka) between the two responses seems to have overwhelmed the influence of the North Atlantic between 100 and 85 ka (Figure A20). 4.3 Millenial-scale events during the past 130 ka In the North Atlantic, the last glacial period was characterized by abrupt, millennial-scale changes in temperature; most notably, Dansgaard-Oeschger (D-O) events (Severinghaus and Brook 1999; Masson-Delmotte et al., 2005). Many of the D-O warm events occurred in sequences of progressively cooler peaks, terminating in a Heinrich event, periods of iceberg discharge that were the among the coldest intervals the North Atlantic, and are believed to have resulted in a weakening of AMOC (McManus et al., 2004). D-O and Heinrich events have a global signature (Overpeck and Cole, 2006), and are often associated with past arid intervals in the tropics (Peterson et al., 2000; Lea et al., 2003) including West Africa (Peck et al., 2004; Mulitza et al., 2008; Collins et al., 2010; 74 Stager et al., 2011). The climate dynamics underlying this connection remain uncertain; in parts of the tropics it seems likely that the aridity was associated with a southward shift in the ITCZ, a hypothesis that is supported by climate models (Broccoli et al., 2006). Whereas there is strong evidence that ITCZ shifted south of northern Lake Malawi during Heinrich Events (Brown et al., 2007; Johnson et al., 2011), there is little evidence of wetter conditions in southern Africa associated with this shift (Stager et al., 2011). In fact, virtually all records from both hemispheres suggest drier conditions during Heinrich 1, suggesting a weakening or contraction, instead of, or at least in addition to, a shift in the rainbelt (Collins et al., 2010; Stager et al., 2011). Cold intervals in the North Atlantic consistently coincide with intervals of low lake at Lake Bosumtwi over the past 130 ka (Figure A21). Furthermore between 60 and 30 ka, each arid Heinrich interval is preceded by gradually decreasing millennial-scale wet peaks that may correspond with D-O cycles. This link between West African hydrology and North Atlantic temperatures is also found in records to the north (off the coast of Senegal; Mulitza et al., 2008) and south (Sanaga River discharge; Weldeab et al., 2007) for at least the past 60 kyr. This provides support for the hypothesis that for Heinrich Events 1 – 5, West Africa experienced a contraction, rather than a shift, of the rainbelt. The Younger Dryas (YD; or Heinrich Event 0), appears to have had a much more limited impact on the region. At Lake Bosumtwi, the interval was a relative lowstand, but lake level was still ~30 m higher than modern levels (Peck et al., 2004; Shanahan et al., 2006). XRF-PC1 suggests that the interval was particularly wet, although this result is likely biased by the presence of manganosiderite during the interval, and not 75 immediately after in the early Holocene, when the lake was overflowing. Nevertheless, the presence of manganosiderite, the δ18Olw and the low terrigenous content all agree that the Younger Dryas was wet in the context of the past 130 ka (Figure A21), however it was still drier than the preceding Bølling-Allerød period and the early Holocene (Peck et al., 2004; Shanahan et al., 2006). 4.4 Intervals of extreme aridity One pronounced feature of core BOS04-5B is the presence of several thick, dense, organic-poor clay intervals. These intervals are massive, and are almost exclusively comprised of fine-grained terrigenous material; organic carbon content drops below 2% in the intervals, and they are typically carbonate-free (Scholz et al., 2007; Peck et al., in prep.). Root casts, and filled mud cracks are also observed in some intervals (pers. comm. J. Peck). We interpret these deposits as evidence for extremely low lake-level. The lack of evaporite minerals (neither halites nor evaporitic carbonates are found in the sediments) suggest that the lake did not completely dry up, but the drill site may have been exposed seasonally. The upper most of these layers, Arid Interval 1 (AI-1), was described by Scholz et al., (2007) and is fairly well dated by nearby OSL and U/Th ages, suggesting that the layer was deposited between 74 to 70 ka. The duration of the event is not well constrained, as sedimentation rates may have been significantly higher than the mean during the interval. Given the observed relationship between North Atlantic temperatures and hydrology at Lake Bosumtwi, it seems most likely the event coincides with Heinrich 7. The OSL and U/Th ages, as well as this correlation with Heinrich 7 suggest that the 76 event also occurred at about the same time as the eruption of the Toba super-volcano, 74+/-2 ka (Oppenheimer, 2002). The Toba eruption is the largest known eruption in the Quaternary, and is thought to have had a severe impact on global climate for at least decades, and possibly accelerated global cooling into glacial conditions (Rampino and Self, 1992; Zielinksi et al., 1996; Jones et al., 2005). It remains unclear, however, the extent of global change resulting from the eruption (Oppenheimer, 2002). We cannot rule out the possibility that AI-1 resulted in part from the Toba eruption. However, the strongest argument against such a possibility is the presence of five other similar units deposited over the past 530 ka, which were, for the most part, deposited during intervals with similar orbital forcing and glacial boundary conditions (Figure A22). All six intervals of extreme aridity of the past 530 kyr occurred between 300 and 50 ka, when the 400-kyr component of orbital eccentricity was highest (Figure A16). Five of the six intervals also occurred during glacial conditions, and they all occurred when the global temperatures were cooler than average, and when global CH4 production was low (Figure A22). AI-1, 2, 3, and 6 occurred during precession-driven insolation minima; however that does not appear to be a prerequisite, since AI-4 and 5 were deposited near insolation maxima. That said, uncertainty in the age model prevents more certain analysis of the phasing between the events and precession. AI-2 was deposited during an interval when severe aridity was common throughout Africa (Scholz et al., 2007; Blome et al., 2012). The lower section of AI-2 contains unlaminated deposits of ankerite, and is the most ankerite-rich interval of core BOS04-5B. Ankerite is indicative of very dry conditions, and the carbonate geochemistry of those ankerites suggest a lake water δ18O 77 of at least 10-13‰ (SMOW; Figure A21), although the oxygen fractionation during ankerite precipitation is uncertain, and is probably a minimum value (Chacko and Deines, 2008). AI-3 and 4, deposited during MIS 6, are the thickest arid intervals, and correspond to the longest period of low organic and inorganic deposition in the past 530 kyr. AI-5 occurred during MIS 7, and is most similar to AI-1. AI-1 and 5 are particular interesting in that they both appear to have occurred soon after, and immediately preceding wet conditions at the lake. In addition to the presence of manganosiderite and organic rich sediment, low lake water δ18O values and the occurrence of rain forest pollen taxa (pers. comm. W. Gosling) further suggest that the the change from immediately before to AI-1 was dramatic, and probably occurred on the order of centuries. Further research into the nature and timing of the transitions into and out of these severe arid intervals is a subject of ongoing investigation. 5. Conclusion and summary In this study, we presented the first high-resolution, terrestrial record from West Africa spanning the last 530 kyr.  The chronology of the record is well constrained over the past 150 kyr by 14 C, OSL and U/Th ages, and is constrained deeper in the core by geomagnetic reversal ages and the age of meteor impact. However, the chronology of the middle of the sediment sequence is less wellconstrained. To allow comparisons with regional records, climatic forcings and boundary conditions we use the strong relationship between mineral- 78 magnetic dust proxies in Lake Bosumtwi and dust content in Antarctic ice cores to tune the record on glacial-interglacial transitions, similar to the approach of EPICA (2004).  XRF-inferred elemental abundance and carbonate geochemistry and mineralogy are indicative of lake level and hydrologic balance at the lake, and indicate an interplay of orbital forcing and Atlantic SSTs boundary conditions over the past 530 kyr.  Lake Bosumtwi hydrology tracked the 400-kyr component of orbital eccentricity, with substantially drier background conditions during the interval of high eccentricity between 300 and 100 ka.  On 100-kyr timescales, lake level had a positive relationship with eccentricity, with generally wetter conditions occurring during interglacials and eccentricity highs, although the relative wetness of each glacial and interglacial period varied substantially, modulated by the 400kyr component of eccentricity.  On shorter timescales, Lake Bosumtwi hydrology appears to have been a balance between high- and low-latitude forcing. Over the past 150 kyr, West African hydrology shows a latitudinally varied response to these two forcings; more northerly sites are more sensitive to obliquity and temperatures in the North Atlantic, whereas more southerly sites respond more to local insolation and precession forcing. Lake Bosumtwi typically lies between these two responses. Interestingly, the amplitude of the 79 forcing influences the latitudinal extent of the response: only more southerly sites responded to the local insolation low ca. 45 ka; whereas all of West Africa dried during the more severe decrease ca. 90 ka.  Abrupt, millennial-scale climate changes in the North Atlantic (i.e., Dansgaard-Oeschger and Heinrich Events) are recorded as abrupt changes in lake level in Lake Bosumtwi. As is common at several sites in both East and West Africa, Heinrich events consistently correspond to dry intervals at Lake Bosumtwi. There is some evidence to suggest that D-O cycles may correspond to progressively drier wet-intervals preceding each Heinrich Event, but uncertainty in the chronology precludes conclusive evidence of this coupling. Millennial-scale variability is greatly reduced during the penultimate glacial period, suggesting a variable influence of abrupt events in the North Atlantic, probably driven by changes in the amplitude of orbital forcing.  Six multi-millennial scale intervals of extreme aridity were recorded in the sediments during the past 530 ka, all during the eccentricity-high interval from 300-50 ka. These events share common features, typically occurring during deep-glacial conditions, and often coinciding with local summer insolation minima. Two of the events (AI-1 and 5), appear to have began and ended abruptly, with wetter-than-average conditions occurring immediately before and after the arid intervals. Sedimentation rates into and out of the transitions are uncertain, but the wet-to-dry shift most likely 80 occurred on the order of centuries or less. The amplitude, duration, and most importantly, the climate dynamics underlying these abrupt changes is an area of continuing investigation. 6. References Almond, M. J. R., and R. E. Hecky (2000), Water Column Properties of Lake Bosumtwi, Ghana, Proceedings of the 45 th Conference on Great Lakes Research., pp. 2–3. Bahr, D. B., E. W. . Hutton, J. P. . Syvitski, and L. F. Pratson (2001), Exponential approximations to compacted sediment porosity profiles, Computers & geosciences, 27(6), 691–700. Banerjee, D. (2001), Supralinearity and sensitivity changes in optically stimulated luminescence of annealed quartz, Radiation measurements, 33(1), 47–57. Berger, A., and M. F. Loutre (1991), Insolation values for the climate of the last 10 million years, Quaternary Science Reviews, 10(4), 297–317. Berner, R. A. (1971), Principles of Chemical Sedimentology, McGraw-Hill Inc.,US. Blaauw, M. (2010), Methods and code for “classical” age-modelling of radiocarbon sequences, Quaternary Geochronology, 5(5), 512–518. Blome, M., A. Cohen, C. Tryon, A. Brooks, and J. Russell (in press), 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. Broccoli, A. J., K. A. Dahl, and R. J. Stouffer (2006), Response of the ITCZ to Northern Hemisphere cooling, Geophys. Res. Lett, 33(10.1029). Brooks, K., C. A. Scholz, J. W. King, J. Peck, J. T. Overpeck, J. M. Russell, and P. Y. . Amoako (2005), Late-Quaternary lowstands of lake Bosumtwi, Ghana: evidence from high-resolution seismic-reflection and sediment-core data, Palaeogeography, Palaeoclimatology, Palaeoecology, 216(3), 235–249. Brown, E. T., T. C. Johnson, C. A. Scholz, A. S. Cohen, and J. W. King (2007), Abrupt change in tropical African climate linked to the bipolar seesaw over the past 55,000 years, Geophysical Research Letters, 34(20), L20702. 81 Bush, M. B., P. E. De Oliveira, P. A. Colinvaux, M. C. Miller, and J. E. Moreno (2004), Amazonian paleoecological histories: one hill, three watersheds, Palaeogeography, Palaeoclimatology, Palaeoecology, 214(4), 359–393. Camberlin, P., S. Janicot, and I. Poccard (2001), Seasonality and atmospheric dynamics of the teleconnection between African rainfall and tropical sea-surface temperature: Atlantic vs. ENSO, International Journal of Climatology, 21(8), 973– 1005. Carothers, W. W., L. H. Adami, and R. J. Rosenbauer (1988), Experimental oxygen isotope fractionation between siderite-water and phosphoric acid liberated CO2siderite, Geochimica et Cosmochimica Acta, 52(10), 2445–2450. Castañeda, I. S., S. Mulitza, E. Schefuss, R. A. . dos Santos, J. S. . Damsté, and S. Schouten (2009), Wet phases in the Sahara/Sahel region and human migration patterns in North Africa, Proceedings of the National Academy of Sciences, 106(48), 20159–20163. Chacko, T., and P. Deines (2008), Theoretical calculation of oxygen isotope fractionation factors in carbonate systems, Geochimica et Cosmochimica Acta, 72(15), 3642– 3660. Chang, C.-P., Y. Ding, and N.-C. Lau (2011), The Global Monsoon System: Research and Forecast, World Scientific. Cheng, H., R. L. Edwards, W. S. Broecker, G. H. Denton, X. Kong, Y. Wang, R. Zhang, and X. Wang (2009), Ice age terminations, Science, 326(5950), 248–252. Cohen, A. S., J. R. Stone, K. R. M. Beuning, L. E. Park, P. N. Reinthal, D. Dettman, C. A. Scholz, T. C. Johnson, J. W. King, and M. R. Talbot (2007), Ecological consequences of early Late Pleistocene megadroughts in tropical Africa, Proceedings of the National Academy of Sciences, 104(42), 16422. Collins, J. A. et al. (2010), Interhemispheric symmetry of the tropical African rainbelt over the past 23,000 years, Nature Geoscience, 4(1), 42–45. Cook, K. H., and E. K. Vizy (2006), Coupled model simulations of the West African monsoon system: Twentieth-and twenty-first-century simulations, Journal of climate, 19(15), 3681–3703. Cruz, F. W., S. J. Burns, I. Karmann, W. D. Sharp, M. Vuille, A. O. Cardoso, J. A. Ferrari, P. L. Silva Dias, and O. Viana (2005), Insolation-driven changes in atmospheric circulation over the past 116, 000 years in subtropical Brazil, Nature, 434(7029), 82 63–66. Cruz, F. W., M. Vuille, S. J. Burns, X. Wang, H. Cheng, M. Werner, R. Lawrence Edwards, I. Karmann, A. S. Auler, and H. Nguyen (2009), Orbitally driven east– west antiphasing of South American precipitation, Nature Geoscience, 2(3), 210– 214, doi:10.1038/ngeo444. deMenocal, P. B., and others (1995), Plio-pleistocene African climate, Science, 270(5233), 53–59. deMenocal, P. B., and D. Rind (1993), Sensitivity of Asian and African Climate to Variations in Seasonal Insolation, Glacial Ice Cover, Sea Surface Temperature,and Asian Orography, Journal of Geophysical Research, 98(D4), PP. 7265–7287, doi:199310.1029/92JD02924. deMenocal, P. B., W. F. Ruddiman, and E. M. Pokras (1993), Influences of high-and lowlatitude process on African terrestrial climate to variations in seasonal insolation, glacial ice cover, sea-surface temperature and Asian orography, Journal of Geophysical Research, 98, 7265–7287. Dezfuli, A. K., and S. E. Nicholson (2011), A note on long-term variations of the African easterly jet, International Journal of Climatology, 31(13), 2049–2054. Drake, N. A., R. M. Blench, S. J. Armitage, C. S. Bristow, and K. H. White (2011), Ancient watercourses and biogeography of the Sahara explain the peopling of the desert, Proceedings of the National Academy of Sciences, 108(2), 458–462. EPICA (2004), Eight glacial cycles from an Antarctic ice core, Nature, 429(6992), 623– 628. Fontaine, B., and S. Janicot (1996), Sea surface temperature fields associated with West African rainfall anomaly types, Journal of climate, 9(11), 2935–2940. Fontaine, B., S. Trzaska, and S. Janicot (1998), Evolution of the relationship between near global and Atlantic SST modes and the rainy season in West Africa: statistical analyses and sensitivity experiments, Climate dynamics, 14(5), 353– 368. Forman, S., J. Pierson, J. Gómez, J. Brigham-Grette, N. Nowaczyk, and M. Melles (2007), Luminescence geochronology for sediments from Lake El’gygytgyn, northeast Siberia, Russia: constraining the timing of paleoenvironmental events for the past 200 ka, Journal of Paleolimnology, 37(1), 77–88. Frankenberg, C., J. F. Meirink, M. van Weele, U. Platt, and T. Wagner (2005), Assessing 83 methane emissions from global space-borne observations, Science, 308, 1010– 1014. Ghil, M. et al. (2002), Advanced spectral methods for climatic time series, Rev. Geophys, 40(1), 1003. Grodsky, S. A., J. A. Carton, and S. Nigam (2003), Near surface westerly wind jet in the Atlantic ITCZ, Geophys. Res. Lett, 30(19), 2009. Hanselman, J. A., W. D. Gosling, G. M. Paduano, and M. B. Bush (2005), Contrasting pollen histories of MIS 5e and the Holocene from Lake Titicaca (Bolivia/Peru), Journal of Quaternary Science, 20(7-8), 663–670. He, C., H. Breuning-Madsen, and T. W. Awadzi (2007), Mineralogy of dust deposited during the Harmattan season in Ghana, Danish Journal of Geography, 107. Heil, C., J. King, J. Peck, T. Shanahan, N. McKay, J. Overpeck, and C. Scholz, (in prep.) Geochronology of the Lake Bosumtwi sediment sequence for the past million years, Quaternary Science Reviews. Hoerling, M., and A. Kumar (2003), The perfect ocean for drought, Nature, 299, 691– 694. Horng, C. S., M. Y. Lee, H. Palike, K. Y. Wei, W. T. Liang, Y. Iizuka, and M. Torii (2002), Astronomically calibrated ages for geomagnetic reversals within the Matuyama chron, Earth Planets and Space, 54(6), 679–690. Hughen, K. A., T. I. Eglinton, L. Xu, and M. Makou (2004), Abrupt tropical vegetation response to rapid climate changes, Science, 304, 1955–1959. Huybers, P., and C. Wunsch (2004), A depth-derived Pleistocene age model: Uncertainty estimates, sedimentation variability, and nonlinear climate change, Paleoceanography, 19(1), PA1028. Imbrie, J., J. D. Hays, D. G. Martinson, A. McIntyre, A. C. Mix, J. J. Morley, N. G. Pisias, W. L. Prell, and N. J. Shackleton (1984), The orbital theory of Pleistocene climate: Support from a revised chronology of the marine delta 18O record, Milankovitch and climate: Understanding the response to astronomical forcing, p. 269. Itambi, A. C., T. Von Dobeneck, S. Mulitza, T. Bickert, and D. Heslop (2009), Millennialscale northwest African droughts related to Heinrich events and DansgaardOeschger cycles: Evidence in marine sediments from offshore Senegal, Paleoceanography, 24(1), PA1205. 84 Janicot, S., S. Trzaska, and I. Poccard (2001a), Summer Sahel-ENSO teleconnection and decadal time scale SST variations, Climate Dynamics, 18(3), 303–320. Janicot, S., S. Trzaska, and I. Poccard (2001b), Summer Sahel-ENSO teleconnection and decadal time scale SST variations, Climate Dynamics, 18(3), 303–320. Johnson, T. C., E. T. Brown, and J. Shi (2011), Biogenic silica deposition in Lake Malawi, East Africa over the past 150,000 years, Palaeogeography, Palaeoclimatology, Palaeoecology, 303(1-4), 103–109. Jones, G. S., J. M. Gregory, P. A. Stott, S. F. . Tett, and R. B. Thorpe (2005), An AOGCM simulation of the climate response to a volcanic super-eruption, Climate Dynamics, 25(7), 725–738. Jourdan, F., P. R. Renne, and W. U. Reimold (2009), An appraisal of the ages of terrestrial impact structures, Earth and Planetary Science Letters, 286(1), 1–13. Kanner, L. C., S. J. Burns, H. Cheng, and R. L. Edwards (2012), High-Latitude Forcing of the South American Summer Monsoon During the Last Glacial, Science, 335(6068), 570–573. Kido, Y., T. Koshikawa, and R. Tada (2006), Rapid and quantitative major element analysis method for wet fine-grained sediments using an XRF microscanner, Marine geology, 229(3), 209–225. Kim, S., and J. Oneil (1997), Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates, Geochimica et Cosmochimica Acta, 61, 3461–3475. Koeberl, C. et al. (2007), An international and multidisciplinary drilling project into a young complex impact structure: The 2004 ICDP Bosumtwi Crater Drilling Project—An overview, Meteoritics & Planetary Science, 42(4-5), 483–511. Koeberl, C., R. Bottomley, B. P. Glass, and D. Storzer (1997), Geochemistry and age of Ivory Coast tektites and microtektites, Geochimica et Cosmochimica Acta, 61(8), 1745–1772. Koeberl, C., and B. Milkereit (2007), Continental Drilling and the Study of Impact Craters and Processes-an ICDP Perspective, Continental Scientific Drilling (eds. Harms, U., Koeberl, C., and Zoback, MD), Springer, Heidelberg, 95–161. Koeberl, C., W. U. Reimold, J. D. Blum, and C. P. Chamberlain (1998), Petrology and geochemistry of target rocks from the Bosumtwi impact structure, Ghana, and comparison with Ivory Coast tektites, Geochimica et Cosmochimica Acta, 62(12), 85 2179–2196. Lambert, F., B. Delmonte, J. R. Petit, M. Bigler, P. R. Kaufmann, M. A. Hutterli, T. F. Stocker, U. Ruth, J. P. Steffensen, and V. Maggi (2008), Dust-climate couplings over the past 800,000 years from the EPICA Dome C ice core, Nature, 452(7187), 616–619. Lamb, P. J. (1978), Large-scale tropical Atlantic surface circulation patterns associated with Subsaharan weather anomalies, Tellus, 30, 240–251. Larrasoaña, J. C., A. P. Roberts, E. J. Rohling, M. Winklhofer, and R. Wehausen (2003), Three million years of monsoon variability over the northern Sahara, Climate Dynamics, 21(7), 689–698. Lea, D. W., D. K. Pak, L. C. Peterson, and K. A. Hughen (2003), Synchroneity of tropical and high-latitude Atlantic temperatures over the last glacial termination, Science, 301(5638), 1361–1364. Lelieveld, J., P. J. Crutzen, and F. J. Dentener (1998), Changing concentration, lifetime and climate forcing of atmospheric methane, Tellus B, 50(2), 128–150. Lopes dos Santos, R. A., M. Prange, I. S. Castañeda, E. Schefuss, S. Mulitza, M. Schulz, E. M. Niedermeyer, J. S. Sinninghe Damsté, and S. Schouten (2010), Glacial– interglacial variability in Atlantic meridional overturning circulation and thermocline adjustments in the tropical North Atlantic, Earth and Planetary Science Letters, 300(3), 407–414. Loulergue, L., A. Schilt, R. Spahni, V. Masson-Delmotte, T. Blunier, B. Lemieux, J. M. Barnola, D. Raynaud, T. F. Stocker, and J. Chappellaz (2008), Orbital and millennial-scale features of atmospheric CH4 over the past 800,000 years, Nature, 453(7193), 383–386. Lu, J., C. Deser, and T. Reichler (2009), Cause of the widening of the tropical belt since 1958, Geophysical Research Letters, 36, 5 PP., doi:200910.1029/2008GL036076. Maley, J. (1987), Fragmentation de la forêt dense humide africaine et extension des biotopes montagnards au Quaternaire récent: nouvelles données polliniques et chronologiques, implications paléoclimatiques et biogéographiques (1987). Palaeoecology of Africa and the surrounding islands. Maley, J. (1989), Late Quaternary climatic changes in the African rain forest: forest refugia and the major role of sea surface temperature variations (1989). Paleoclimatology and paleometeorology: modern and past patterns of global atmospheric transport. 86 Maley, J. (1991), The African rain forest vegetation and palaeoenvironments during late Quaternary, Climatic change, 19(1), 79–98. Maley, J. (1997), Middle to Late Holocene changes in tropical Africa and other continents: paleomonsoon and sea surface temperature variations, Nato ASI Series I: Global Environmental Change, 49, 611–640. Maley, J. (2002), A catastrophic destruction of African forests about 2,500 years ago still exerts a major influence on present vegetation formations, IDS bulletin, 33(1), 13–30. Maley, J., and D. A. Livingstone (1983), Extension d’un élément montagnard dans le sud du Ghana (Afrique de l’Ouest) au Pléistocène supérieur et à l’Holocène inférieur: premières données polliniques, CR Acad. Sci. Paris, Sér, 2(196), 1287–1292. Masson-Delmotte, V., J. Jouzel, A. Landais, M. Stievenard, S. J. Johnsen, J. W. C. White, M. Werner, A. Sveinbjornsdottir, and K. Fuhrer (2005), GRIP deuterium excess reveals rapid and orbital-scale changes in Greenland moisture origin, Science, 309(5731), 118–121. McManus, J. F., R. Francois, J. M. Gherardi, L. D. Keigwin, and S. Brown-Leger (2004), Collapse and rapid resumption of Atlantic meridional circulation linked to deglacial climate changes, Nature, 428(6985), 834–837. Miller, G. H., F. Wendorf, R. Ernst, R. Schild, A. E. Close, I. Friedman, and H. P. Schwarcz (1991), Dating lacustrine episodes in the eastern Sahara by the epimerization of isoleucine in ostrich eggshells, Palaeogeography, Palaeoclimatology, Palaeoecology, 84(1-4), 175–189. Moernaut, J., D. Verschuren, F. Charlet, I. Kristen, M. Fagot, and M. De Batist (2010), The seismic-stratigraphic record of lake-level fluctuations in Lake Challa: Hydrological stability and change in equatorial East Africa over the last 140ákyr, Earth and Planetary Science Letters, 290(1-2), 214–223. Moreno, A. (2012), A Multiproxy Paleoclimate Reconstruction over the Last 250 kyr from Marine Sediments: The Northwest African Margin and the Western Mediterranean Sea, in: Hublin J.-J. and McPherron S.P. (Eds.), Modern Origins, Springer Netherlands, pp. 3–17. Moreno, A., I. Cacho, M. Canals, M. A. Prins, M.-F. Sánchez-Goñi, J. O. Grimalt, and G. J. Weltje (2002), Saharan Dust Transport and High-Latitude Glacial Climatic Variability: The Alboran Sea Record, Quaternary Research, 58(3), 318–328, doi:10.1006/qres.2002.2383. 87 Moron, V. (1994), Guinean and Sahelian rainfall anomaly indices at annual and monthly scales (1933–1990), International journal of climatology, 14(3), 325–341. Moron, V., A. Navarra, M. N. Ward, C. K. Folland, P. Friederichs, K. Maynard, and J. Polcher (2001), Analysing and combining atmospheric general circulation model simulations forced by prescribed SST: northern extratropical response, Annals of Geophysics, 44(4), doi:10.4401/ag-3565. Mulitza, S., M. Prange, J. B. Stuut, M. Zabel, T. von Dobeneck, A. C. Itambi, J. Nizou, M. Schulz, and G. Wefer (2008), Sahel megadroughts triggered by glacial slowdowns of Atlantic meridional overturning, Paleoceanography, 23(4), PA4206. Nicholson, S. E. (2001), Climatic and environmental change in Africa during the last two centuries, Climate Research, 17(2), 123–144. Nicholson, S. E. (2008), The intensity, location and structure of the tropical rainbelt over west Africa as factors in interannual variability, International Journal of Climatology, 28(13). Nicholson, S. E. (2009a), On the factors modulating the intensity of the tropical rainbelt over West Africa, International Journal of Climatology, 29(5), 673–689, doi:10.1002/joc.1702. Nicholson, S. E. (2009b), A revised picture of the structure of the “monsoon” and land ITCZ over West Africa, Climate Dynamics, 29. Nicholson, S. E., and J. P. Grist (2001), A conceptual model for understanding rainfall variability in the West African Sahel on interannual and interdecadal timescales, International Journal of Climatology, 21(14), 1733–1757. Nicholson, S. E., and P. J. Webster (2007), A physical basis for the interannual variability of rainfall in the Sahel, Quarterly Journal of the Royal Meteorological Society, 133(629). Opoku-Ankomah, Y., and I. Cordery (1994), Atlantic Sea Surface Temperatures and Rainfall Variability in Ghana., Journal of Climate, 7, 551–558. Oppenheimer, C. (2002), Limited global change due to the largest known Quaternary eruption, Toba ~74 kyr BP?, Quaternary Science Reviews, 21(14-15), 1593–1609. Overpeck, J. T., and J. E. Cole (2006), Abrupt change in Earth’s climate system, Annu. Rev. Environ. Resour., 31, 1–31. 88 Peck, J. A., R. R. Green, T. Shanahan, J. W. King, J. T. Overpeck, and C. A. Scholz (2004), A magnetic mineral record of Late Quaternary tropical climate variability from Lake Bosumtwi, Ghana, Palaeogeography, Palaeoclimatology, Palaeoecology, 215(1-2), 37–57. Peck, J., N. McKay, T. Shanahan, C. Heil, J. Overpeck, J. King, and C. Scholz (in prep.), A million-year history of the West African Monsoon, in preparation for Science. Peterson, L. C., G. H. Haug, K. A. Hughen, and U. Rohl (2000), Rapid changes in the hydrologic cycle of the tropical Atlantic during the last glacial, Science, 290, 1947–1951. Pu, B., and K. H. Cook (2010), Dynamics of the West African westerly jet, Journal of Climate, 23(23), 6263–6276. Pu, B., and K. H. Cook (2012), Role of the West African Westerly Jet in Sahel Rainfall Variations, Journal of Climate, 25(8), 2880–2896. Rampino, M. R., and S. Self (1992), Volcanic winter and accelerated glaciation following the Toba super-eruption, Nature, 359(6390), 50–52. Rosenbaum, J., and S. M. F. Sheppard (1986), An isotopic study of siderites, dolomites and ankerites at high temperatures, Geochimica et Cosmochimica Acta, 50(6), 1147–1150. Rowell, D. P. (2001), Teleconnections between the tropical Pacific and the Sahel, Quarterly Journal of the Royal Meteorological Society, 127(575). Rowell, D. P. (2003), The impact of Mediterranean SSTs on the Sahelian rainfall season, Journal of Climate, 16(5), 849–862. Ruddiman, W. F., and M. E. Raymo (2003), A methane-based time scale for Vostok ice, Quaternary Science Reviews, 22(2-4), 141–155. Russell, J., M. R. Talbot, and B. J. Haskell (2003), Mid-holocene climate change in Lake Bosumtwi, Ghana, Quaternary Research, 60(2), 133–141. Schneider, R. R., B. Price, P. J. Mtiller, D. Kroon, and I. Alexanderz (1997), Monsoon related variations in Zaire (Congo) sediment load and influence of fluvial silicate supply on marine productivity in the east equatorial Atlantic during the, Paleoceanography, 12(3), 463–481. Scholz, C. A., T. C. Johnson, A. S. Cohen, J. W. King, J. A. Peck, J. T. Overpeck, M. R. Talbot, E. T. Brown, L. Kalindekafe, and P. Y. O. Amoako (2007), East African 89 megadroughts between 135 and 75 thousand years ago and bearing on earlymodern human origins, Proceedings of the National Academy of Sciences, 104(42), 16416. Scholz, C. A., T. Karp, K. M. Brooks, B. Milkereit, P. Y. O. Amoako, and J. A. Arko (2002), Pronounced central uplift identified in the Bosumtwi impact structure, Ghana, using multichannel seismic reflection data, Geology, 30(10), 939–942. Severinghaus, J. P., and E. J. Brook (1999), Abrupt Climate Change at the End of the Last Glacial Period Inferred from Trapped Air in Polar Ice, Science, 286(5441), 930– 934, doi:10.1126/science.286.5441.930. Shanahan, T., J. Beck, J. Overpeck, N. McKay, and J. Pigati (Submitted), Late Quaternary sedimentological and climate changes at Lake Bosumtwi Ghana: new constraints from laminae analysis and radiocarbon age modeling, Quaternary Science Reviews. Shanahan, T. M. (2006), West African monsoon variability from a high-resolution paleolimnologic record (Lake Bosumtwi, Ghana), University of Arizona. Shanahan, T. M., J. T. Overpeck, J. W. Beck, C. W. Wheeler, J. A. Peck, J. W. King, and C. A. Scholz (2008), The formation of biogeochemical laminations in Lake Bosumtwi, Ghana, and their usefulness as indicators of past environmental changes, Journal of Paleolimnology, 40(1), 339–355. Shanahan, T. M., J. T. Overpeck, W. E. Sharp, C. A. Scholz, and J. A. Arko (2007), Simulating the response of a closed-basin lake to recent climate changes in tropical West Africa (Lake Bosumtwi, Ghana), Hydrological Processes, 21(13). Shanahan, T. M., J. T. Overpeck, C. W. Wheeler, J. W. Beck, J. S. Pigati, M. R. Talbot, C. A. Scholz, J. Peck, and J. W. King (2006), Paleoclimatic variations in West Africa from a record of late Pleistocene and Holocene lake level stands of Lake Bosumtwi, Ghana, Palaeogeography, Palaeoclimatology, Palaeoecology, 242(3), 287–302. Shanahan, T., J. Overpeck, K. Anchukaitis, J. Beck, J. Cole, D. Dettman, J. Peck, C. Scholz, and J. King (2009), Atlantic forcing of persistent drought in West Africa, Science, 324(5925), 377. Shanahan, T., J. Peck, S. Forman, C. Heil, N. McKay, J. Overpeck, J. King, and C. Scholz (in prep), Geochronology of the Lake Bosumtwi sediment sequence for the past 200,000 years, Quaternary Science Reviews. das Sharma, S., D. J. Patil, and K. Gopalan (2002), Temperature dependence of oxygen 90 isotope fractionation of CO2 from magnesite-phosphoric acid reaction, Geochimica et cosmochimica acta, 66(4), 589–593. Sheldon, N. D., and G. J. Retallack (2001), Equation for compaction of paleosols due to burial, Geology, 29(3), 247. Singarayer, J. S., P. J. Valdes, P. Friedlingstein, S. Nelson, and D. J. Beerling (2011), Late Holocene methane rise caused by orbitally controlled increase in tropical sources, Nature, 470(7332), 82–85. Stager, J. C., D. B. Ryves, B. M. Chase, and F. S. . Pausata (2011), Catastrophic drought in the Afro-Asian monsoon region during Heinrich event 1, Science, 331(6022), 1299. Stone, J. R., K. S. Westover, and A. S. Cohen (2011), Late Pleistocene paleohydrography and diatom paleoecology of the central basin of Lake Malawi, Africa, Palaeogeography, Palaeoclimatology, Palaeoecology, 303(1-4), 51–70. Szabo, B., C. Haynes, and T. A. Maxwell (1995), Ages of Quaternary pluvial episodes determined by uranium-series and radiocarbon dating of lacustrine deposits of Eastern Sahara, Palaeogeography, Palaeoclimatology, Palaeoecology, 113(2), 227–242. Talbot, M. R., and G. Delibrias (1977), Holocene variations in the level of Lake Bosumtwi, Ghana, Nature, 268(5622), 722–724. Talbot, M. R., and G. Delibrias (1980), A new late Pleistocene-Holocene water-level curve for Lake Bosumtwi, Ghana, Earth and Planetary Science Letters, 47(3), 336–344. Talbot, M. R., and T. Johannessen (1992), A high resolution palaeoclimatic record for the last 27,500 years in tropical West Africa from the carbon and nitrogen isotopic composition of lacustrine organic matter, Earth and Planetary Science Letters, 110(1-4), 23–37. Talbot, M. R., and K. Kelts (1986), Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana, Geology, 14(11), 912–916. Talbot, M. R., D. A. Livingstone, P. G. Palmer, J. Maley, J. M. Melack, G. Delibrias, and S. Gulliksen (1984), Preliminary results from sediment cores from Lake Bosumtwi, Ghana, Palaeoecology of Africa, 16, 173–192. Tarutani, T., R. N. Clayton, and T. K. Mayeda (1969), The effect of polymorphism and magnesium substitution on oxygen isotope fractionation between calcium 91 carbonate and water, Geochimica et Cosmochimica Acta, 33(8), 987–996, doi:10.1016/0016-7037(69)90108-2. Tiedemann, R., M. Sarnthein, and N. J. Shackleton (1994), Astronomic timescale for the Pliocene Atlantic δ18O and dust flux records of Ocean Drilling Program Site 659, Paleoceanography, 9(4), P. 619, doi:199410.1029/94PA00208. Tierney, J. E., J. M. Russell, Y. Huang, J. S. S. Damste, E. C. Hopmans, and A. S. Cohen (2008), Northern Hemisphere Controls on Tropical Southeast African Climate During the Past 60,000 Years, Science, 322(5899), 252. Tjallingii, R., M. Claussen, J. B. . Stuut, J. Fohlmeister, A. Jahn, T. Bickert, F. Lamy, and U. Röhl (2008), Coherent high-and low-latitude control of the northwest African hydrological balance, Nature Geoscience, 1(10), 670–675. Tjallingii, R., U. Röhl, M. Kölling, and T. Bickert (2007), Influence of the water content on X-ray fluorescence core-scanning measurements in soft marine sediments, Geochemistry, Geophysics, Geosystems, 8(2), 1–12. Toggweiler, J. R., and J. Russell (2008), Ocean circulation in a warming climate, Nature, 451(7176), 286–288, doi:10.1038/nature06590. Turner, B. F., L. R. Gardner, and W. E. Sharp (1996a), The hydrology of Lake Bosumtwi, a climate-sensitive lake in Ghana, West Africa, Journal of Hydrology, 183(3-4), 243–261. Turner, B. F., L. R. Gardner, W. E. Sharp, and E. R. Blood (1996b), The geochemistry of Lake Bosumtwi, a hydrologically closed basin in the humid zone of tropical Ghana, Limnology and Oceanography, 1415–1424. Vasconcelos, C., J. A. McKenzie, R. Warthmann, and S. M. Bernasconi (2005), Calibration of the δ18O paleothermometer for dolomite precipitated in microbial cultures and natural environments, Geology, 33(4), 317–320, doi:10.1130/G20992.1. Verschuren, D. et al. (2009), Half-precessional dynamics of monsoon rainfall near the East African Equator, Nature, 462(7273), 637–641. Visser, K., R. Thunell, L. Stott, and others (2003), Magnitude and timing of temperature change in the Indo-Pacific warm pool during deglaciation, Nature, 421(6919), 152–155. Wallace, J. M., and P. V. Hobbs (2006), Atmospheric science: an introductory survey, Academic press. 92 Wang, X., A. S. Auler, R. L. Edwards, H. Cheng, E. Ito, Y. Wang, X. Kong, and M. Solheid (2007), Millennial-scale precipitation changes in southern Brazil over the past 90,000 years, Geophysical Research Letters, 34(23), L23701. Wang, Y., H. Cheng, R. L. Edwards, X. Kong, X. Shao, S. Chen, J. Wu, X. Jiang, X. Wang, and Z. An (2008), Millennial-and orbital-scale changes in the East Asian monsoon over the past 224,000 years, Nature, 451(7182), 1090–1093. Weldeab, S., D. W. Lea, R. R. Schneider, and N. Andersen (2007), 155,000 years of West African monsoon and ocean thermal evolution, Science, 316(5829), 1303. Wolff, E. W., J. Chappellaz, T. Blunier, S. O. Rasmussen, and A. Svensson (2010), Millennial-scale variability during the last glacial: The ice core record, Quaternary Science Reviews, 29(21–22), 2828–2838, doi:10.1016/j.quascirev.2009.10.013. Wünnemann, B., K. Hartmann, N. Altmann, U. Hambach, H. J. Pachur, and H. Zhang (2007), 22. Interglacial and glacial fingerprints from lake deposits in the Gobi Desert, NW China, Developments in Quaternary Sciences, 7, 323–347. Yarincik, K. M., R. W. Murray, T. W. Lyons, L. C. Peterson, and G. H. Haug (2000), Oxygenation history of bottom waters in the Cariaco Basin, Venezuela, over the past 578,000 years: results from redox-sensitive metals (Mo, V, Mn, and Fe), Paleoceanography, 15(6), 593–604. Zabel, M., and H. D. Schulz (2001), Importance of submarine landslides for non-steady state conditions in pore water systems—Lower Zaire (Congo) deep-sea fan, Marine Geology, 176(1), 87–99. Zebiak, S. E. (1993), Air-sea interaction in the equatorial Atlantic region, Journal of Climate, 6, 1567–1567. Zielinski, G. A., P. A. Mayewski, L. D. Meeker, S. Whitlow, and M. S. Twickler (1996), A 110,000-yr record of explosive volcanism from the GISP2 (Greenland) ice core, Quaternary Research, 45(2), 109–118. 93 Table A1. Elemental loadings of the first two principal components of the XRF data Element Al Ti K Si Fe Mn Ca Sr Rb Si PC1 -0.39 -0.43 -0.44 -0.42 0.03 0.30 0.15 0.06 -0.40 0.12 PC2 0.03 0.11 0.00 0.03 0.51 0.35 -0.49 -0.50 -0.04 0.33 94 90 1°27' W 300 90 90 90 45 6°32' N 30 m 50 m 90 90 45 90 90 45 70 m 45 90 6°27' N 90 90 0 1 2 km Figure A1. Site map of the Lake Bosumtwi crater, modified from Shanahan (2006). Solid gray lines show bathymetry (10 m contour). Dashed line indicates crater rim. Inset: location of Lake Bosumtwi in southern Ghana, West Africa. 4.8 10 4.4 8 4.0 6 3.6 4 3.2 2 2.8 monthly rainfall (cm) Wind speed (m s-1) 95 0 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Figure A2. 1961-2000 climatology for Kumasi, Ghana, 30 km from Lake Bosumtwi. Grey bars indicate monthly precipitation totals (cm), dashed line shows wind speed (m s-1). Modified from Shanahan (2006). 96 Figure A3. Multichannel seismic reflection profile showing draped reflectors of the lacustrine fill and the brecciated bedrock of the central uplift (from Scholz et al., 2002). The 13 drill holes at the five drill sites are shown as black bars. Drill hole 5B is shown as a solid line, along with the down core median grain size profile at the site. 97 0 Age model ensemble envelope Age model median estimate EDC-dust-tuned Age Model 14C ages OSL ages U/Th ages Geomagnetic reversal ages Meteor impact age RMCD turbidite−adjusted 50 100 150 200 0 200 400 0 600 Age (ka) 800 1000 RMCD turbidite−adjusted 10 20 30 40 50 60 70 80 0 50 100 150 200 Age (ka) Figure A4. Upper: Age-depth relation and age models for all of core BOS04-5B. Error bars show 2-sigma uncertainty where it is larger than the size of the marker. Median age model and ensemble range determined by the Classical Age Modelling approach of Blaauw (2010). Dashed box shows depth-age range shown in lower plot. 260 285 310 335 360 385 410 435 460 485 510 535 98 BOS04−5B 10−1, 26 to 55 cm Mean downcore x-ray density X−ray intensity 500 1000 1500 2000 2500 560 585 610 635 66 X-radiograph 300 350 400 450 500 450 500 Core face image 300 350 400 depth (mm) Figure A5. An example of a turbidite in Lake Bosumtwi sediments, and the analyses and tools used to identify turbidites in the core. Upper: Mean x-radiograph intensitty of the sediment; higher x-ray intensities correspond to lower density material. Middle: x-radiograph. Bottom: core face image. Turbidite identified by red bar in all three plots. 600 650 600 650 99 A mean x−ray intensity Depth (mm) 2000 1190 1200 1200 1210 1210 1220 1220 1230 1230 Depth (mm) 1200 1190 mean x−ray intensity 2000 Vert. var. = 3.4 x 104 Hor. var. = 9.3 x 102 log V:H ratio = 3.6 1200 mean x−ray intensity 1120 Depth (mm) 480 1240 480 490 490 500 500 510 510 520 520 mean x−ray intensity 1240 Depth (mm) B Vert. var. = 6.8 x 102 Hor. var. = 3.4 x 102 log V:H ratio = 0.67 1120 Figure A6. Examples of the calculation of the vertical:horizontal density variance ratio for both A) well-laminated and B) unlaminated sections of core BOS04-5B. In both A) and B), both core face images and x-radiographs are shown for 5-cm sections core. To the right and below each x-radiograph, horizontal and vertical intensity means are shown, respectively. 100 200 180 160 140 120 100 80 60 40 20 0 A 0 100 B 50 Bosumtwi Hc:Hcr Dust/Moisture Proxy ODP 663 Dust Flux (g cm-2 kyr-1) 20 0.8 C 0.6 0.4 0.2 EDC Dust Concentration ng/g ODP 659 Dust Flux (g cm-2 kyr-1) Age (ka) 1000 D 200 180 160 140 120 100 80 Age (ka) 60 40 20 0 0 Figure A7. Records of dust flux over the past 200 kyr. A) Dust flux recorded in marine core ODP 659 off the coast of West Africa (deMenocal, 1995). B) Dust flux recorded in marine core ODP 663 in the Gulf of Guinea (deMenocal et al., 1993). C) Hc:Hcr a mineral-magnetic dust proxy for Lake Bosumtwi (Peck et al., in prep.). D) Dust concentration in the EPICA Dome C ice core (Lambert et al., 2008). Age (ka) 800 700 600 500 400 300 200 100 0 20 0 100 0.8 0.6 0.4 0.2 1000 800 700 600 500 400 Age (ka) 300 200 100 0 0 EDC dust (ng/g) Bosumtwi Hc:Hcr Dust/Moisture Proxy 50 Dust flux (gcm-2 kyr-1 ) Dust flux (gcm-2 kyr-1 ) 101 Figure A8. Records of dust flux over the past 800 kyr. A) Dust flux recorded in marine core ODP 659 off the coast of West Africa (deMenocal, 1995). B) Dust flux recorded in marine core ODP 663 in the Gulf of Guinea (deMenocal et al., 1993). C) Hc:Hcr a mineral-magnetic dust proxy for Lake Bosumtwi (Peck et al., in prep.). D) Dust concentration in the EPICA Dome C ice core (Lambert et al., 2008). Tie points used to tuned the Lake Bosumtwi to glacialinterglacial transitions on the EDC timescale are shown in red. Tie points are also estimated for A) and B), but are not used in adjusting the chronology. 102 XRF-PC1 Untuned 2 10 XRF-PC1 Dust-tuned 2 10 270 1 Power 10 270 1 25 17 42 23 16 11 9 9 8 0 0 10 −1 10 10 10 110-90 10 110-90 −1 0 0.1 0.2 0 0.1 Mn Untuned 4000 0.2 Mn Dust-tuned 90 110-90 Power 3000 2000 34 16 14 11 10 22 26 1000 0.1 0.2 0 0 8 6 0.1 Ca Untuned 6000 13 10 0 0 0.2 Ca Dust-tuned 180 67 180 Power 16 41 4000 68 19 2000 34 26 0 0 13 11 25 8 0.1 Frequency (1 kyr-1) 0.2 0 0 17 12 8 7-6 0.1 0.2 Frequency (1 kyr-1 ) Figure A9. Multitaper-method power spectra for XRF-PC1, Mn and Ca, both before and after tuning the age model to glacial-interglacial transitions in the EDC dust record (Lambert et al., 2008). 103 Age (ka) 500 450 400 350 300 250 200 150 100 50 0 100 200 20 0 40 100 60 Si (cps) 0 4 200 0.5 300 1 Mn (cps) 3000 1.5 Fe (cps) x 10 2000 2 1000 3000 0 2000 1000 S (cps) 60 0 40 20 500 450 400 350 300 250 Age (ka) 200 150 100 50 0 Figure A10a. Unadjusted XRF counts per second for Ti, Si, K, Fe, Mn, Ca, and S. Ca (cps) K (cps) Ti (cps) 0 104 Age (ka) Ti (cps) 0 500 450 400 350 300 250 200 150 100 50 0 50 100 0 20 40 200 5000 1500 10000 1000 15000 Fe (cps) 100 500 2000 0 1000 S (cps) 60 40 0 20 0 500 450 400 350 300 250 Age (ka) 200 150 100 50 0 Figure A10b. XRF counts per second for Ti, Si, K, Fe, Mn, Ca, and S after adjusting for water content, down-core compaction and turbidites. See text for details. Ca (cps) Mn (cps) K (cps) 0 Si (cps) 150 105 40% 30% 20% 50 Si 30 Water−contentadjusted Ti (cps) 10 Water−contentadjusted Si (cps) Estimated reduction in water content since core splitting Ti 200 150 100 0 20 40 60 80 100 Depth (RMCD) 120 140 Figure A11. The effect on Si and Ti XRF counts of uncertain drying of the cores on the adjustment for attenuation and absorption of x-rays by water. The median estimate of 30% was used in the study, see text for details. 106 0.6 Ca Sr 0.4 PC2 0.2 Rb K Al Si Ti 0 −0.2 −0.4 Mn S Fe −0.4 −0.2 0 PC1 0.2 0.4 0.6 Figure A12. Loadings of XRF data on XRF-PC1 and XRF-PC2. XRF-PC1 is dominated by the abundance of terrigenous elements, whereas PC2 tracks carbonate mineralogy. 107 A B C D Figure A13. Relation between XRF-PC1 and Mn (A) and Ca (C) showing the effect of dilution of terrigenous material by carbonates. Black lines show linear regressions used to adjust for dilution, and B) and D) show the relation after adjustment. 108 Depth (RMCD) 0 50 100 150 2.5 2 1.5 1 4 0.5 3 0 2 1 0 Mn and Ca (cps) 3000 −1 −2 log V:H density variance ratio Lamination index 3 2000 1000 0 0 50 100 Depth (RMCD) 150 Figure A14. Variations in the A) subjective and B) objective lamination indices for the upper 160 m of core BOS04-5B. The objective index is calculate as the ratio of vertical to horizontal density variance in the x-radiographs. C) Mn (red) and Ca (gray), indicative of carbonates in the sediment. Well-laminated sediments and most carbonates in the sediment are indicative of anoxic bottom waters (Talbot and Kelts, 1986). 109 Turbidites (cm m−1) 100 80 60 40 20 0 0 50 100 150 200 Depth (RMCD) 250 300 Figure A15. Abundance of turbidites in core BOS04-5B. Turbidite thicknesses summed over contiguous 1-meter increments. 110 550 0.8 500 450 400 350 300 250 200 150 100 50 0 A 0.6 0.4 0.2 1000 B 0 100-kyr (54% var) -4 (-) 0.01 0 0 (+) -0.01 XRF-PC1 Lake Level Eccentricity C 400-kyr (41% var) (Axis inv.) EDC Dust (ng/g) Bosumtwi Hc:Hcr Dust/Moisture Proxy Age (ka) D 5000 2000 600 Turbidite abundance (cm kyr-1) 400 2000 0 50 E 0 550 500 450 400 350 300 250 Age (ka) 200 150 100 50 Mn and Ca (cps) 800 4 EPICA CH (ppbv) 4 0 Figure A16. Lake Bosumtwi paleohydrology indicators compared with orbital forcing and Antarctic dust and global methane. A) Dust abundance in Lake Bosumtwi inferred from mineral magnetic properties (Hc:Hcr; Peck et al., in prep.). B) Dust content in the EPICA Dome C (EDC) Antarctic ice core (Lambert et al., 2008). C) The two primary frequency components of orbital eccentricity along with Lake Bosumtwi XRF-PC1 (This study). D) Atmospheric methane concentration (Loulergue et al., 2008) with Lake Bosumtwi Mn and Ca abundance (This study). Mn and Ca are indicative of manganosiderite and Ca-carbonates, respectively. E) Total thickness of turbidites summed over each 1000-yr interval in the record (This study). Interglacial periods shown in yellow. 1 A 2000 0.5 correlation coefficient (r) 0 1000 500 400 300 200 Age (ka) 0.3 B 0.3 C 0.2 0.2 0.1 0.1 −100 0 turbidite lag (cm) 100 −5000 100 0 turbidite lag (yrs) 0 0 5000 Figure A17. A) Turbidite abundance and Mn over the past 530 ka. The lagged correlations between turbidites and Mn with B) depth and C) time. On average, turbidites lag Mn by 20-30 cm, or 1000 to 3000 yrs. Mn cps Turbidites (m m-1) 111 112 1000 800 600 400 200 0 A 100 50 Lamination Index 0 3 2 1 0 B 1000 800 600 400 Age (ka) 200 0 Figure A18. A) Turbidite abundance and B) the presence and quality of laminations over the past 1.08 Ma. For the lamination index 3=well laminated, 2=moderately laminated, 1=poorly laminated, and 0=unlaminated. Turbidites (cm kyr-1) Age (ka) −2 10° JJA insolation (W m ) 10° JJA insolation (W m−2) 113 480 A Timing of rainy seasons Jun Sep May Oct Apr Nov 460 440 420 400 20 15 10 Age (ka) 5 0 480 B Timing of rainy seasons Jun Sep May Oct Apr Nov 460 440 420 400 225 220 215 Age (ka) 210 Figure A19. Boreal summer (JJA) insolation during intervals of A) low and B) high eccentricity. Conceptual changes in the in timing of rainy seasons on the Guinea Coast, in response to shifts in the mean, and estimated summer maximum positions are shown in red (qualitative scales). When insolation is highest the rainy seasons shift away from the boreal solstice. This effect is more dramatic when insolation is higher, limiting the duration of wet conditions at the lake. Age (ka) NGRIP δ18O 100 50 0 1.5 B 1 21°N 0 -1 C 12°N 2.5 3.5 -30 -28 -26 20 40 E 9°N 25 60 80 F 24 5 6.5°N H 0 5 10°N July Insolation (W m-2) 4 20 I 500 2-6°N 25 J 30 450 K 7 6 11°S-6°N 150 100 50 0 5 Gulf of Guinea SSS (p.s.u.) 0 6 4-14°N 10 Congo drainage weathering (Al:K) Lake Bosumtwi XRF-PC1 G 18 23 7 −4 Drier δ Olw Niger drainage Obliquity (‰) weathering (Al:K) (°) Terrigenous dust (%) D C3-C4 veg. balance (C29 n-alkane δ13C) −45 Terrestrial Humidity Index −40 10°N Terrestrial moisture 114 75°N A −35 Eolian vs Fluvial sediment (Ti:Al) Wetter 150 Age (ka) Figure A20. West African terrestrial moisture indicators and climatic forcing conditions over the past 150 kyr. A) NGRIP δ18O on the GICC05ext timescale (NGRIP 2004, Wolff et al., 2010). B) Terrestrial humidity index from core GeoB7920-2 (Tjallingii et al., 2008). C) Ti:Al ratio from core GeoB9527-5 (Itambi et al., 2009) D) C29 n-alkane δ13C from core GeoB9528-3 (Castañeda et al., 2009). E) Terrigenous dust content from core ODP 661 (Demenocal, 1995). F) Obliquity (Berger and Loutre, 1991). G) Al:K ratios from core GeoB4901-8 (Zabel et al., 2001). H) Lake Bosumtwi XRF-PC1 and δ18Olw (this study). I) Sea surface salinity inferred from Ba/Ca from core MD03-2707 (Weldeab et al., 2007). J) 10°N July insolation (Berger and Loutre, 1991). K) Al:K ratios from core GeoB1008-3 (Schneider et al., 1997). Vertical colored bars indicate intervals of high eccentricity and precession, and the latitudinally varied terrestrial moisture response in each interval. 115 Age (ka) −35 100 80 60 H7 A 40 H5.2 H6 H4 H5 20 H3 0 YD H2 H1 −6 −45 −4 B −2 O lake water (‰) 18 δ Lake Bosumtwi 0 2 0 4 C 5 10 29 D 28 27 26 25 Gulf of Guinea SSS (psu) XRF−PC1 (Terrigenous) −40 22 24 24 23 26 Gulf of Guinea SST (°C) 18 NGRIP δ O (‰) 120 E 28 30 32 120 100 80 60 Age (ka) 40 20 0 Figure A21. Lake Bosumtwi hydrology over the past 130 kyr, compared with regional records and hemispheric boundary conditions. A) δ18O of Greenland ice tracking North Atlantic temperature (NGRIP, 2004; Wolff et al., 2010). B) Lake Bosumtwi terrigenous input, an indicator of lake level (This study). C) δ18O of lake water inferred from authigenic carbonates in Lake Bosumtwi (this study). D) Sea surface temperature and E) salinity in the Gulf of Guinea (Weldeab et al., 2007). Millennial-scale cold events in Greenland shown in yellow, often coincident with labeled Heinrich events. 116 Age (ka) AI-6 700 600 250 200 150 AI-5 100 AI-3 AI-4 A 50 0 AI-1 AI-2 500 400 300 −4 B −2 July Insolation 15oN (W m-2) 500 −6 0 2 400 4000 4 C 2000 2000 0 −360 −380 D −400 −420 −440 EDC δD (per mil) Mn and Ca (cps) XRF−PC1 EDC CH4 (ppbv) 300 800 −460 300 250 200 150 Age (ka) 100 50 0 Figure A22. The timing of extreme arid intervals during the past 300 ka. A) EDC methane concentration (Lambert et al., 2008). B) Lake Bosumtwi XRF-PC1 lake level indicator (black; This study) with July insolation at 15oN (red; Berger and Loutre, 1991). C) Mn (red) and Ca (gray) intensities, indicating the presence of manganosiderite and calcium carbonates, respectively (This study). D) EDC δD (Jouzel et al., 2007), indicative of Antarctic temperatures. Aridity intervals (AI) shown in yellow bars. 117 APPENDIX B A 12,000-YEAR-LONG, ANNUALLY-RESOLVED VARVE RECORD SPANNING THE LAST INTERGLACIAL FROM LAKE BOSUMTWI, SOUTHERN GHANA 118 A 12,000-YEAR-LONG, ANNUALLY-RESOLVED VARVE RECORD SPANNING THE LAST INTERGLACIAL FROM LAKE BOSUMTWI, SOUTHERN GHANA Nicholas McKay1, Jonathan T. Overpeck1,2,3, Timothy M. Shanahan4, John A. Peck5, Clifford W. Heil6, John W. King6, and Chris A. Scholz7 1 Department of Geosciences, University of Arizona Department of Atmpospheric Science, University of Arizona 3 Institute of the Environment, University of Arizona 4 Jackson School of Geosciences, University of Texas at Austin 5 Department of Geology and Environmental Science, University of Akron 6 Graduate School of Oceanography, University of Rhode Island 7 Deptartment of Earth and Environmental Sciences, Syracuse University 2 119 Abstract The impact of continued global warming on the likelihood of severe drought in sub-Saharan West Africa remains uncertain, as climate models generally do not simulate realistic climate dynamics in the region and have inconsistent projections for the future. The Last Interglacial period (LIG), occurring between 128 and 116 thousand years ago, is a partial analog for future warming because at its peak, global temperatures were slightly higher, and this warming was accentuated in Northern Hemisphere terrestrial summer temperatures. Here we present a new, annually-resolved, 12,100-year-long varve record for the LIG from Lake Bosumtwi in southern Ghana (6.5°N, 1.4°W). The abundance of terrigenous elements in the sediment, varve thickness, and the isotope geochemistry and mineralogy of authigenic carbonates in the sediment are all sensitive to changes in lake level, and record a dynamic history of hydrologic variability in the region. The LIG lake highstand was lower and shorter-lived than the the prolonged highstand in the early Holocene, and unlike the Holocene, the lake never overflowed during LIG. The overall drier conditions during the Holocene are most likely driven by amplified precessional forcing during the interval, resulting in a northward shift in the rainbelt. The LIG, like the Holocene, had two distinct millennial-scale moist intervals, from 125 - 123 and 121 – 120 ka. In both the LIG and the Holocene, these peaks occurred during times of precessiondriven insolation maxima in July and October, corresponding to the two rainy seasons in the modern climatology. This suggests that, at least during interglacials, prolonged wet conditions occur at the lake when rainy season insolation is highest. Over the course of the LIG, lake level generally tracked sea surface temperatures (SST) in Gulf of Guinea, 120 including an abrupt drop in lake level that lasted about 500 years ca. 118 ka, corresponding to cooling in the Gulf of Guinea and much of the North Atlantic during the interval. The timing and duration of the event are comparable to the Late Eemian Aridity Pulse (LEAP) that is observed at several European sites, and has been interpreted to result from abrupt cooling in the North Atlantic, and possibly a reduction of Atlantic meridional overturning circulation. This scenario would likely result in drought in West Africa, so the aridity ca. 118 ka is the first indication that the LEAP occurred in Africa as well as Europe. The occurrence of quasiperiodic variability at multidecadal to centennial timescales is consistent with the hypothesis that slowly varying changes in Atlantic SST structure (e.g., the Atlantic Multidecadal Oscillation) drives long term hydrologic variability in the region. The periodicities vary over the course of the record, further implicating the role ocean circulation. 1. Introduction Sub-Saharan West Africa is prone to decadal to century-scale droughts, which often begin abruptly, and are difficult to predict. The recent “Sahel Drought” began suddenly and unexpectedly in the late 1960's, resulting in widespread famine, mass emigration, and political instability, and exemplifies the effects of abrupt climate change (Benson and Clay, 1998). Virtually no instrumental precipitation records in the region were collected before 1915 (Nicholson, 2001), and the prolonged drought spanning the end of the 20th Century is well outside the variability observed prior to 1968, when the drought began. For example, annual precipitation totals for the region for every year from 1970 to 1997 121 were below the 100-yr mean (Nicholson and Webster, 2007). The instrumental record is too short to understand the full range of hydrologic variability in the region, especially because of the occurrence of multidecadal droughts and pluvials in the pre-instrumental record (Shanahan et al., 2009). This motivates the investigation of paleoclimate archives. Unfortunately, terrestrial West Africa has among the poorest coverage of paleoclimate data in the world; however, recent work on Lake Bosumtwi has revealed extreme variability in hydrology over the past 2700 years, indicating that the “Sahel Drought” is well within the range of past variability, and that the region has experienced more severe and longer lasting droughts (Shanahan et al., 2009). The implications of this finding for the influence of continued global warming on moisture availability in the region remains unclear. Climate model projections for precipitation and evaporation in sub-Saharan west Africa vary widely from model to model (Cook and Vizy, 2006; Randall et al., 2007; Christensen et al., 2007; Rodriguez-Fonseca et al., 2011). This is partly because the region lies on the boundary between two zonal bands with well-understood responses to global warming. Climate models consistently simulate increases in water vapor with increases in temperature, following the Clausius-Clapeyron relation, which reduces the poleward redistribution of moisture in the tropics, resulting in increases in equatorial precipitation and decreases in subtropical moisture (Held and Soden, 2006). In addition, continued tropospheric warming and stratospheric cooling, driven by both increases in greenhouse gases and anthropogenic ozone depletion, results in a widening of the tropical belt (Lu et al., 2009) and a poleward shift in the westerlies (Shindell and Schmidt, 2004; 122 Toggweiler and Russell, 2008), further drying the subtropics. The Sahel and Guinea Coast regions are in the transition zone between increases in precipitation in the tropics and decreases in the subtropics. Projecting changes in hydrology in the region is further complicated by the dynamics of the West African Monsoon (WAM). Unlike convective precipitation in other equatorial and monsoonal regions, the West African monsoon is not driven by the convergence of moist southwesterly monsoonal flow and the ITCZ, but rather, by the complex interaction of upper-level circulation and jets with low level moisture transport (See section 1.2) for more detail; Nicholson, 2009; Pu and Cook, 2010; 2012). This complexity, and the importance of small-scale features, such as the West African Westerly Jet, is probably part of the reason why GCMs struggle to simulate the West African Monsoon, both over the 20th century and in the mid-Holocene (Cook and Vizy, 2006; Randall et al., 2007; Christensen, 2007; Pu and Cook; 2012). Only 4 of the 18 AR4 models analyzed by Cook and Vizy (2006) demonstrated reasonably realistic simulations of the WAM during the second half of the 20th century. The models that did well in the 20th century, each simulated substantially different responses for the 21st century. For the A2 scenario, GFDL CM2.0 and CM2.1 (Held et al., 2005) showed slight drying in the Guinea Coast region, and extreme drying in the Sahel, MIROC (medres model) simulated wetter conditions in the Sahel and drier on the Guinea Coast, and MRI simulated small increases in precipitation on the Guinea Coast, and a slight drying of the Sahel (Cook and Vizy, 2006). Given the uncertainty in climate model projections for the region, using 123 paleoclimate records to understand how hydrology in the region responds to changes in forcing may provide insight to the region will respond to continued warming. The Last Interglacial period (LIG; roughly equivalent to the Eemian, or Marine Isotope Stage 5e), which occurred from about 128 to 118 ka, is an interesting period to study for several reasons; first, it's the most recent analog to the Holocene, with generally similar conditions, although differences in the orbital configuration resulted in altered insolation forcing, especially seasonally (Berger and Loutre, 1991). It is often considered a partial analog for future climate, because global mean temperatures were slightly warmer (02°C; Turney and Jones, 2010; McKay et al., 2011) and sea level was 6-8 m higher than preindustrial conditions (Hearty et al., 2007; Kopp et al., 2009), although global CO2 (and other GHG) concentrations were far below modern and projected future concentrations (270-290 ppm), Luthi et al. (2008). The largest difference in forcing relative to the Holocene during the LIG is the change in seasonal forcing due to the amplification of the precession cycle by eccentricity (Berger and Loutre, 1991). This resulted in a >30 W m-2 increase in Northern Hemisphere insolation summer, and a corresponding decrease in Southern Hemisphere winter insolation at the height of the interglacial ca. 124 ka. This resulted in much warmer temperatures in the Northern Hemisphere, especially on land during the summer (Turney and Jones, 2010; McKay et al., 2011). The northern Hemisphere is consistently projected to warm faster in the future than the southern Hemisphere in all seasons, and especially on land. More regionally, Weldeab et al., (2007) show that the SST in the Gulf of Guinea was generally warmer during the LIG than the Holocene, which is also a consistent projection of GCMs for the 21st century 124 (Cook and Vizy, 2006). Altogether, reconstructing hydrologic variability in the region during the LIG has the potential to provide insight into future climate in the region. 1.1 Site Lake Bosumtwi is a 1.08 +/- Ma (Koeberl et al., 1997; Jourdan et al., 2009) meteoriteimpact crater lake located in southern Ghana (6°30' N, 1°25' W) approximately 150 km north of the Gulf of Guinea (Figure B1). The lake has no outlet, is isolated from the groundwater table, and the lake surface occupies about half of the basin, and approximately 80% of the annual water input from rainfall directly on the surface (Turner et al., 1996a). These features make lake level extremely sensitive to changes precipitation and evaporation, and lake level responds to interannual and decadal-scale changes in precipitation, rising more than 5 m during the wet interval from 1940-1970, and falling more than 2 m during the “Sahel Drought” from 1970-2000 (Shanahan et al., 2007). Sedimentation in the lake and the stable isotope geochemistry of carbonates precipitated in the lake are sensitive to lake level (Talbot and Delibrias, 1977; Talbot and Kelts, 1986; Peck et al., 2004; Shanahan, 2006; Shanahan et al., 2009). The lake is meromictic and typically anoxic below 15 m depth (Turner et al., 1996b), allowing for the preservation of finely laminated varves, which have accumulated over the past ca. 2700 yr (Shanahan et al., 2008a, 2009). Finely laminated, presumably varved, sediment is common in intervals throughout the 291-m sediment sequence deposited since the crater was formed, 1.08 Ma (Peck et al., in prep.; Appendix A). This study focuses on a continuously laminated interval between 57.4 and 60.7 meters below the present-day lake floor. 125 1.2 Modern Climate Climate in the region is strongly seasonal, and most of the variability can be attributed to the annual migration of the tropical rainbelt (Nicholson, 2009) (Figure B2). Annual precipitation at the lake has a bimodal distribution, with peak rainfall occurring in May-July and a smaller peak from September-October. Mean temperature at nearby Kumasi varies from about 24°C in the boreal summer to 28°C in the winter (Shanahan et al., 2008). The cooler temperatures in the summer are attributable to enhanced upwelling in the Gulf of Guinea, while the warm winters are indicative of the influence of the terrestrial air masses to the north. Wind in the region is also seasonal, and is controlled by the position of the Intertropical Convergence Zone (ITCZ). In the boreal summer the ITCZ is positioned north of the lake over the Sahel and Sahara, and the monsoonal southwesterlies peak in August. In the winter the ITCZ is positioned near the Guinea Coast, wind speed is reduced, and aerosol-rich northeasterly 'Harmattan' winds are common. Dust supplied to the region by these winds has been shown to be a substantial component of Ghanaian soils (He et al., 2007) and solute load in Lake Bosumtwi (Turner et al., 1996b). Interannual variability in precipitation in the region has two primary modes of variability: a monopole and dipole mode. The monopole mode describes years when the both the Guinea Coast and the Sahel are wetter or drier than average. This mode is associated with the amount of low-level moisture transport (Pu and Cook, 2012) and the strength of the Tropical Easterly Jet (TEJ; Nicholson 2008a,b), which, when stronger, drives upper level divergence and promotes deep convection through the rain belt. The 126 strength of the TEJ may in turn be associated with the strength of the Asian Monsoon and tropical Pacific SSTs (Chen and van Loon, 1987; Janicot et al., 2001; Nicholson, 2008b). The dipole mode describes rainfall anomalies in the Sahel region (north of 10° N) are of the opposite sign of those in the Guinea Coast region (south of 10° N) (Nicholson and Webster, 2007; Nicholson, 2008a). The mode of variability, the sign of that mode, and the amplitude of the anomalies are related to the strength, position and structure of the upper level TEJ (Nicholson 2008a,b) and the mid-level African Easterly Jet (AEJ; Nicholson and Webster, 2007; Nicholson, 2008a, Dezfuli and Nicholson, 2011). Whereas upper atmospheric circulation drives the convective conditions, moisture is primarily derived from low levels from the southerly monsoonal flow, and a distinct, low-level westerly jet originating off the coast of West Africa near the Atlantic ITCZ, which was identified by Grodsky et al. (2003). Recent work suggests that the West African Westerly Jet (WAWJ) is critical in driving interannual and decadal variability in precipitation in the Sahel (Pu and Cook, 2010; 2012). Variability in these atmospheric circulation patterns are related to a number of factors, including sea surface temperatures (SST) in the tropical Pacific (e.g., Rowell 2001; Janicot et al., 2001) and eastern Mediterranean (Rowell, 2003), as well as the strength of the Asian Monsoon (Chen and van Loon, 1987; Janicot et al., 2001; Nicholson, 2008b). However, SST variability and especially the interhemispheric sea level pressure (SLP) and SST gradients in the tropical Atlantic are most strongly linked to interannual and decadal scale variability in the regional precipitation (e.g., Lamb, 1978; Nicholson and Grist, 2001; Rowell, 2003; Nicholson and Webster, 2007; Nicholson, 2008a,b; Nicholson, 2009, Pu and Cook, 2012). 127 Interhemispheric SST structure has been implicated in past hydrologic changes in the region as well. Shanahan et al., (2009) showed that lake level at Lake Bosumtwi was tied to the Atlantic Multidecadal Oscillation (AMO) over the past 2700 years. Cold intervals in the North Atlantic have been associated with extremely dry conditions in much of West Africa, most notably during Heinrich 1 (Stager et al., 2011); although similar conditions have been observed in both terrestrial and marine proxies throughout the past 100 ka (Peck et al., 2004; Weldeab et al., 2007; Mulitza et al., 2008; Tjallingii et al., 2008; Castañeda et al., 2009; Itambi et al.; 2009). 2. Methods 2.1 Core Drilling and sampling In July and August of 2004, a total of 1.83 km of sediment was recovered from five drill sites using the global lake drilling system (GLAD-800; Peck et al., 2005; Koeberl et al., 2007). This study focuses on cores recovered from deep-water (74 m) site 5, drilled into the annular sediment moat around the impact-derived central uplift (Figure B3). Two cores from that site (5B and 5C) capture the entire 296-m-long lacustrine sediment sequence. After drilling, the cores were shipped to the University of Rhode Island, where they were split, described, imaged, and analyzed for a wide variety of physical parameters. U-channels were taken from the axis of each core to measure a suite of mineral magnetic parameters. Core logging and physical sediment properties were used to integrate the depth scales for the three drill cores from site 5 (BOS04-5A,B&C) into a relative mean core depth (RMCD; pers. comm. C. Heil). 128 2.2 Sampling and laminae analysis The U-channels from the core sections of cores 5B and 5C were subsampled, flash-frozen using liquid nitrogen, and freeze dried at the University of Arizona. Once dehydrated, the sediments were vacuum-embedded with epoxy resin (cf., Lotter and Lemcke, 1999). For optical analysis, the embedded sediments were cut and polished down into thin sections of 30-40 μm thickness, and then photographed at 25x magnification under crosspolarized light using a Canon EOS D20 SLR camera attached to an Olympus SZX9 binocular microscope. To create mosaic images without distortion, the photographs were stitched together using Canon PhotoStitch software version 3.1 in “Document” mode. The images were then spatially rectified in ArcGIS 10.0 (ESRI Software, Inc.) by aligning the images with the 25-μm-resolution μXRF data, or a 100-μm-resolution glass scale included in the photographs. Laminae were then counted and measured in ArcGIS 10.0. 2.3 XRF analysis The relative abundance of major and minor elements in the sediments was analyzed using two scanning X-ray fluorescence techniques. The embedded sediments from core 5B were analyzed at 25-μm resolution, under vacuum, using the EDAX Eagle III scanning XRF at the University of Arizona, using a Mo X-ray source and 16 sec count times. The short count times used on the embedded sediments were necessary to allow for high-resolution measurements of elemental variability within small lamina present in the the sediments, and to support laminae counting. After the varves were identified and 129 counted, two methods were used to calculate annual averages. In the first approach, the raw XRF spectra for each year were summed and reanalyzed in the EDAX software. This approach allows for better defined spectra (with effectively longer count times), but was too inefficient to use for the Monte Carlo ensembles (see section 2.4). For the ensemble members, the XRF counts for each point were simply averaged. The latter approach result is less precise than the former, however comparison of the two methods suggests that the large changes in XRF intensities throughout the record are well-captured by either approach. Nevertheless, we hope to improve the efficiency of the first approach and fully adopt this methodology in future research, as this two-step procedure would allow the XRF data to be collected and used to aid laminae counting at extremely high resolution, and then subsequently integrated into yearly averages from very well-defined spectra. Over the course of the analyses, a decrease in count rates were observed in all elements, possibly due to a weakening of the X-ray source. To account for this, Cl counts on pure resin at the beginning, end, or in gaps in the sediment in each puck were used to standardize the count rates. The standardized counts compare well with lower resolution XRF data collected using the Itrax scanning XRF at the University of Minnesota, Duluth, justifying the normalization (Figure B4; Appendix A). 2.4 Development of the varve chronology The fine laminations found throughout most of the Lake Bosumtwi record have been carefully studied in several intervals. The lamina comprising the top 1.5 m of the sediment sequence are annual layers, or varves, as shown by their correspondence with 130 210 Pb, and 14C “bomb-pulse” ages over recent decades and centuries, and 14C ages over the past 2700 yrs (Shanahan et al., 2007, 2009). A longer interval of laminations deposited between 12 and 29 ka closely correspond with the nearby 14C dates, confirming that the laminations deposited during this interval are also varves (Shanahan et al., in prep). The internal structure of the varves through these intervals varies considerably (Shanahan et al., 2007). Varves formed during the late Holocene are comprised of a clastic unit, which most likely washes in during the rainy season, and organic matter and calcite-rich unit which forms during the productive season (Shanahan et al., 2007). In intervals deeper in the core, the calcite is often absent or replaced by another carbonate mineral (i.e., dolomite, aragonite, siderite, or ankerite), and the organic layer typically appears in thin section as a thin, opaque layer. Nevertheless, the minerogenic (rainy season) component and the organic-rich (productive season) component, appear to be a robust indicator of the annual cycle across a wide range of environmental and climatic conditions. The laminations studied here also consist of couplets of alternating minerogenic and organic-rich units, among other components (including several carbonate minerals) which occur in sections of the study interval. The interval is well beyond the limits of radiocarbon dating, and there are no independent geochronological approaches that can be used to verify my interpretation of the annual nature of the laminations in this interval. Nevertheless, because the layers have a similar structure as the Holocene and late Pleistocene varves, and because their components are consistent with an annual cycle of sedimentation, it is most likely that they represent annual layers. In this study, the top of each varve was identified by the opaque, organic-rich unit deposited each productive 131 season. For some laminae, it is unclear whether or not they represent a year, or are subannual depositional events. For these layers, we followed the approach Rasmussen et al., (2006) used in layer counting the NGRIP ice core, and recorded these layers 0.5 +/0.5 years. These layers were later used to help quantify the uncertainty in the varve chronology, varve thickness, and XRF records. The entire LIG section (57.3 to 60.7 relative mean composite depth (RMCD)) from both cores 5B and 5C was thin sectioned, and the varves were counted and measured twice for each core, resulting in nearly complete replication of the varve sequence, and for most intervals, four estimates of varve counts and thickness. To allow for comparison and evaluation of counting uncertainty between cores, 379 marker layers, primarily turbidites, and units with distinct mineralogies (iron-sulfides and carbonates), were identified in cores 5B and 5C. The composite varve chronology was developed by creating an ensemble of varve counts and thicknesses between each set of marker layers, moving progressively up the sediment section. For most intervals, this resulted in four ensemble members for each interval. In sections where there were gaps in the sediment in one of the cores, only the two estimates from the continuous core were used. For each of the four varve count and thickness estimates, hundreds of additional ensemble members were generated stochastically by allowing the uncertain years identified during counting and measurement (i.e., 0.5 +/- 0.5 years) to be either a separate year, or part of the adjacent lamination. Following this approach, 1000 ensemble members for varve count and thickness were created for the gap between each set of marker layers. These ensembles allow quantification of the uncertainty associated with identifying, and 132 measuring the varves throughout the record and estimation of the chronological uncertainty between marker layers. By repeating analyses on individual ensemble members, we can investigate how counting and identification uncertainty influence the varve thickness and XRF records. It is important to note, however, that this approach represents only one component of the chronological uncertainty in the record. The other component, the veracity of my interpretation of annual cycles in the sediment cannot be independently determined, and is not included in the uncertainty measurements. Because we have identified synchronous marker layers in cores BOS04-5B and BOS04-5C, two different approaches for creating XRF and varve thickness ensembles are used. In the first approach, marker layers are forced to occur in the same year in each ensemble, so the the gap between marker layers also remains constant. This approach is preferable for examining the records in time space, since it enforces the observed stratigraphic order in the record, not allowing for varves below a marker layers in one core to be integrated with varves above the same layer in the other core. These ensembles, however, alter the serial nature and autocorrelation structure of the records. Therefore, a second set of ensembles, that disregard the marker layers, is used for spectral analysis of the records. 2.5 Geochronology Although the varve chronology for the interval is well constrained, the absolute age of any given varve is unknown; it is a “floating” chronology. For the purposes of comparing with other regional and global records, and understanding the intervals 133 response to orbital forcing, it is worthwhile to try to “pin” the “floating” chronology as accurately as possible. The uppermost sediments of the Lake Bosumtwi sediment sequence are well-dated by 14C ages and varve counting (Shanahan et al., 2009; Shanahan et al., in review), however, beyond the limits of radiocarbon dating the chronology becomes less certain. Deeper in the core, optically stimulated luminescence (OSL) ages, and U/Th ages on authigenic carbonates provide further age control, however, the analytical uncertainty and lack of agreement between ages make identification of MIS 5e in the sediment sequence ambiguous (Figure B5). In Appendix A, the Lake Bosumtwi sediment sequence was roughly synchronized with the EPICA Dome C (EDC) timescale by matching abrupt changes in the Lake Bosumtwi and EDC dust records (Lambert et al., 2008; Peck et al., in prep) that coincide with glacial-interglacial transitions. The two nearest tie points relevant to the LIG are the transition from MIS 6 to MIS 5e (133 ka on the EDC timescale) and the onset of abundant dust deposition associated with MIS 4 (70 ka on the EDC timescale). This tuned chronology includes its own assumptions, and does not allow for evaluation of leads and lags in the onset and termination of the period (Appendix A). However, it does increase our confidence that the varve interval was deposited during the LIG. Over the past 140 ka, the relative abundance of terrigenous sediment in the Lake Bosumtwi record generally covaries with nearby SST and SSS records (Weldeab et al., 2007), and to a lesser extent the NGRIP δ18O record (NGRIP, 2004; Figure B6). A distinct cold event in the NGRIP record, Greenland stadial 25 (GS25), followed by an abrupt shift into Greenland interstadial 24 (GI-24; ~110 ka) is clearly recognizable in SST and SSS records in the Gulf of Guinea, and likely corresponds to a 134 dry-to-wet transition in Lake Bosumtwi, apparent in the abundance of terrigenous material in the sediment at about 109 ka on the timescale from Appendix A. Assuming this excursion in Lake Bosumtwi hydrology is correlative with GS-25 and GI-24 implies a 1 kyr offset in the Lake Bosumtwi timescale (Appendix A; Shanahan et al., in prep) relative to the NGRIP GICC05ext timescale (Wolff et al., 2010). Therefore, we use this offset to pin the “floating” varve chronology to an absolute age; suggesting that the 12.1 kyr-long varve record spans the interval from 128.6 to 116.5 ka. It is very difficult to assess the uncertainty in these absolute ages. The assignment relies on the several assumptions, including that the large scale climate changes are synchronous between Greenland and West Africa. Although this hypothesis is well supported in the literature, there are few if any records for this time from West Africa that have chronologies that are completely independent from the NGRIP chronology. Another assumption is that the 1 kyr offset between the Lake Bosumtwi and NGRIP records apparent 110 kyr remains constant back to 116.5 ka. This assumption could be avoided by pinning the record to an event within the varve chronology, however, there are not any unambiguously correlative events in the varved interval. By tying the floating varve chronology to a single event we avoid potential biases associated with “wiggle-matching” the varve chronology to nearby high-resolution records. There are several other lines of evidence supporting my hypothesis that the study interval corresponds to the LIG. Dust proxies for Lake Bosumtwi generally share the main features of both nearby and high latitude dust records, particularly the difference between glacials and interglacial periods (Peck et al., in prep.; Appendix A). The study 135 interval comprises a period of low dust content immediately following high dust deposition at the end of MIS 6. The 3.4-m-long varved interval studied here has relatively high organic matter and manganosiderite content, both of which are indicative of relatively high lake level (Appendix A). Furthermore, the interval contains continuous, sub-mm laminations indicating the persistence of anoxic bottom waters, another indicator of wet or warm conditions at the lake. The overwhelming majority of terrestrial moisture sensitive records for West Africa indicate that interglacials are generally wet with respect to glacial periods (e.g., deMenocal et al., 1993; deMenocal, 1995; Weldeab et al., 2007; Mulitza et al., 2008; Tjallingii et al., 2008; Itambi et al.; 2009), and this was clearly the case at Lake Bosumtwi for the Holocene relative to the late Pleistocene (Talbot and Johannessen, 1992; Peck et al., 2004). Therefore, it is likely that lake level was higher during MIS 5e than MIS 6 or MIS 5d. Taken together, the general agreement of lake level indicators at Lake Bosumtwi and global-scale climate phenomena during MIS 5, and the occurrence of a high-moisture balance interval immediately following dust-rich MIS 6 makes the study interval (57.3 to 60.7 RMCD) the most likely candidate for the LIG at the lake. This is consistent with the nearby U/Th and OSL ages (Figure B5; Shanahan et al., in prep.). Finally, the Blake geomagnetic excursion, which occurred during the LIG (123 +/- 3 ka; Lund et al., 2006), has been identified in this interval (Heil et al., in prep.). 2.6 Carbonate isotope geochemistry Potential samples for carbonate mass spectrometry were identified from thin 136 section analysis and the XRF intensities of Ca (for calcite, dolomite and ankerite) or Mn (for manganosiderites). A 300 μm drill bit was then used to sample the carbonates from the resin-embedded sediments. Once cured, the epoxy resin does not react with the phosphoric acid used in isotopic analysis; powdered resin added to calcite standards NBS-19 and NBS-18 resulted in no offset from pure carbonates. The powdered carbonates were heated under vacuum at 100°C for several hours to remove volatile organic compounds prior to measurement. The δ18O and δ13C of these carbonates were measured using an automated carbonate preparation device (KIEL-III) coupled to a gasratio mass spectrometer (Finnigan MAT 252). Powdered samples were reacted with dehydrated phosphoric acid under vacuum at 70°C. The isotope ratio measurement is calibrated based on repeated measurements of NBS-19 and NBS-18 and precision is ± 0.10 ‰ for δ18O and ±0.08‰ for δ13C (1 sigma). 3. Results 3.1 Generalized Lithology and Laminae structure A generalized lithology is shown for the study interval in Figure B7. The studied interval is a 3.4-m-long (57.4 to 60.7 m RMCD), continuously laminated with variable lithologies. The interval is bounded above and below by unlaminated sediments, immediately below the study interval is 2-cm-thick carbonate-rich layer, underlain by massive, dense clays and silts with little internal structure. 137 3.1.1 Unit 1a (60.3 – 60.7 m RMCD) The base of the study interval is composed of dark-brown finely laminated (0.1 to 0.5 mm) muds and opaque organic layers with abundant well-rounded quartz grains, varying from 10 to 50 μm in diameter, presumably of eolian origin. Moving up in the section, the well-rounded grains gradually become finer and sparser, and are rare above 60.4 m RMCD. Calcite-rich sublaminae are common in this interval, but are not found within every lamination. 3.1.2 Unit 1b (60.2 – 60.3 m RMCD) In unit 1b, the laminae remain finely laminated (0.1 to 0.4 mm), but carbonaterich laminae are rare. The laminae are clearly defined as alternating tan, minerogenic layers and thinner, opaque, organic-rich layers. The minerogenic layers often have a distinctly reddish color, associated with fine-grained iron sulfide minerals. 3.1.3 Unit 1c (59.8 – 60.2 m RMCD) The base of unit 1c is composed of very finely laminated (0.05 to 0.15 mm) terrigenous-calcite-organic triplets that gradually thicken upwards and into thick (0.4 to 1.0 mm) terrigenous layers divided by poorly defined opaque organic units. Through this interval, distinct iron-sulfide-rich bands become common, and carbonate layers are rare in the thickest lamina. Above, 59.9 m RMCD, the laminations thin somewhat (0.3 to 0.6 mm), and thin calcite layers are common in most laminae, and thick, red iron sulfide layers are found occasionally. 138 3.1.4 Unit 1d (59.6 – 59.8 m RMCD) Here, finely laminated (0.1 to 0.4 mm) alternating minerogenic terrigenous layers alternate with opaque organic layers, with light tan calcite layers in each packet. Distinct, bright red discontinuous layers of coarse-grained iron sulfide occur sporadically, once every several layers. The coarseness of morphology of the layers suggest that these sulfide layers were formed post-depositionally, however, they remain confined to individual lamina and are traceable between cores, suggesting that they form in the shallow subsurface in response to bottom water geochemistry. 3.1.5 Unit 2a (59.1 – 59.6 m RMCD) Unit 1 grades into Unit 2, where the laminae become extremely fine (0.05 to 0.1 mm), and calcite layers are rare. Bright red iron-sulfide-rich layers are common, but are thinner and more continuous than their counterparts in Unit 1d. The laminae are still primarily formed of alternating minerogenic and organic units, but the minerogenic units become very thin, and the overall composition of the sediment is more organic-rich. Throughout the interval, there are distinctive, relatively thick (0.1 to 1 mm) orange manganosiderite units, and above 59.5 RMCD, thick (up to 7 cm) turbidites are common. The turbidites are primarily minerogenic, and coarsen upwards, terminating in lightcolored fine-grained caps. In places, the turbidites occur in packets, with no or few laminae between turbidites. Elsewhere in the unit, however, turbidites of various thicknesses are separated by hundreds of fine lamina. There is no evidence of erosive contacts at the base of any of the turbidites, although some of the thicker turbidites 139 include packets of laminae reworked from other, presumably shallower areas of the lake. 3.1.6 Unit 2b (58.7 – 59.1 m RMCD) Above the extremely fine laminations of unit 2a, the lamina thicken (0.1 to 0.4 mm) and the grain sizes of the terrigenous layers become visibly coarser, composed primarily of silt-size material. The unit is less organic-rich than unit 2a, but contains more dark-colored organic material than unit 1. The laminations in this interval often contain a thin reddish-brown layer in the middle of the clastic terrigenous layer. The layers are rich in Fe, but contain no sulfur, and may be associated with redox geochemistry accompanying an annual overturning of lake during this interval. Large, (1-3 mm) individually occurring gypsum crystals that cut across nearby lamina are found in this interval. The crystals are clearly diagenetic, and may be associated with increased amount of sedimentary organic-material in the interval, although the origin of the crystals and their significance remains uncertain. 3.1.7 Unit 2c (58.6 - 58.7 m RMCD) In unit 2c, the fine laminations (0.1 to 0.3 mm) are predominantly composed of a minerogenic clastic unit interbedded with a thin, opaque organic unit. The sediments here contain abundant, well-round silt-sized quartz, similar to the base of unit 1a, but somewhat finer grained and more densely packed. Dolomite, present in both distinct yellow bands, and mixed in with the terrigenous layer is also present in this interval. 140 3.1.8 Unit 2d (58.4 - 58.6 m RMCD) This unit is very finely laminated (0.05 to 0.15 mm), and the grain size of the terrigenous clastic unit is finer than in units 2c and 2b, with rare well-rounded quartz grains. The laminations are primarily composed of distinct minerogenic and opaque organic couplets, although reddish, iron-sulfide-rich layers are common. As in Unit 2b, large diagenetic (1-3 mm) gypsum crystals cut across laminations in places. 3.1.9 Unit 3a (58.1 - 58.4 m RMCD) Dark brown muds alternating with thin opaque organic layers form the laminations in this interval, which gradually thicken upwards (from about 0.2 mm to 0.3 mm). Well rounded, presumably eolian, quartz grains are abundant through this interval. Yellow-tan carbonates of variable compositions between dolomite and ankerite occur in places through the interval, as both fine grained crystals interbedded with terrigenous clastic sediments, and as larger (100 μm to 1 mm) cyrstals that form discontinuously across lamina and appear to have formed post-depositionally. 3.1.10 Unit 3b (57.4 - 58.1 m RMCD) Tan, terrigenous, minerogenic sediment dominates this interval, alternating with discontinuous thin opaque organic-rich layers to define laminations that thicken upward (from 0.4 mm near the base to about 1.5 mm near the top). Distinct red iron-sulfide-rich layers commonly form distinct bands that are traceable across cores. As in Unit 3a, dolomitic to ankeritic carbonates are found as both distinct yellow-tan units with large 141 crystals, and as fine-grained minerals within the terrigenous units. Near the top of the section the laminations continue to thicken and the opaque organic layer becomes less distinct until the laminations are no longer clearly recognizable, marking the top of the study interval. 3.2 Carbonate isotope geochemistry and mineralogy The δ18O and δ13C of the carbonates (δ18Ocarb, δ13Ccarb) show extreme variability across through the 12,000 yr record, varying from 0.7 to 12.5‰ and 3.0 to 23.5‰, respectively (Figure B8). Generally, δ18Ocarb and δ13Ccarb covary, consistent with previous studies at the lake (Figure B8; Talbot and Kelts, 1986). Furthermore, the mineralogy and mechanism of formation of the carbonates also have a strong control over their carbon and oxygen isotopic composition. Talbot and Kelts (1986) identified two primary modes of formation for the carbonates in Lake Bosumtwi, 1) near the lake surface driven by CO2 drawdown during the productive season, and 2) near or just below the sediment-water interface driven by methanogenesis. The latter mode of formation results in very high δ13C values (7 – 27‰) because the methane produced is extremely depleted (-60‰) in 13 C, causing the remaining carbon is isotopically enriched. Methanogenesis does not affect the oxygen isotopic composition of the carbonates, and the oxygen isotopes of carbonates formed by either mechanism should track that of the surface or deep water (Talbot and Kelts, 1986). Because the lake is small, has a short residence time (~30 yrs ; Turner et al., 1996a), and is relatively well-mixed (Turner et al., 1996b), the carbonates formed in the bottom waters probably represent a long-term average of the lake water 142 δ18O (δ18Olw), whereas those formed near the surface may reflect a shorter-term “snapshot” of δ18Olw. On decadal and longer timescales, however, the two are likely comparable. Changes in carbonate mineralogy complicate the δ18Ocarb record, because the water-carbonate oxygen isotope fractionation varies considerably between minerals (cf., Chacko and Deines, 2008). At least four species of carbonate are found in the study interval. The minerals were identified using several methods. First, the manganosiderites, found between 59.1 and 59.6 m RMCD, are readily apparent in thin section, because of their bright orange color, and in the XRF Mn intensities, which are very high (Figure B9). Calcite is identifiable in thin section, as distinct white layers, with sharp Ca XRF peaks between 59.6 and 60.6 m RMCD, and 58.8 and 59.1 m RMCD. The transition from calcite to dolomite at 59.1 m RMCD occurs abruptly, in less than 4 cm. Between 58.4 and 59.1 m RMCD, dolomite is visually distinct in thin section, forming thicker layers and is yellow in thin section. The shift in mineralogy across this transition was confirmed using Raman spectroscopy (Rutt and Nicola, 1974; Herman et al., 1987). It is likely that the calcite near the transition to dolomite is relatively rich in Mg, however the inclusion of small amounts Mg into the calcite structure has a minimal impact on the water-carbonate fractionation (~0.06‰/1% Mg at 25°C; Tauritani et al., 1969). Talbot and Kelts (1986) report up to 18% mol MgCO3 in Lake Bosumtwi calcites, that varied significantly, even within individual laminations. This is consistent with the results of Abebe (2010), who found that the molar % of MgCO3 varies from 7.9 to 19.2 throughout drill core 5B. This would correspond to a maximum offset of ~1‰, which would have a minor impact on 143 major trends in δ18Ocarb. Therefore, we ignore the amount of Mg in the calcites; however, ideally these data would be included in the analysis. Finally, both dolomite and ankerite [Ca(Fe,Mn,Mg)(CO3)2], occur between 57.4 and 58.4 m RMCD. Much of the carbonate in this section of the core is fine grained and intermixed with terrigenous sediment, and is only identifiable because of the weak Ca (and sometimes Mn) peak in the XRF data. It is difficult to obtain well-defined Raman spectra on the fine grained material, although ankerite is observed in X-ray diffraction analyses on samples from 57.70 and 57.84 m RMCD (Abebe, 2010). The occurrence of ankerite in the sediment complicates the interpretation of the δ18Ocarb through this interval for several reasons. First, the oxygen isotopic fractionation from water to ankerite has not been empirically determined, although a theoretical estimate suggests that it should be similar to calcite, and substantially lower than dolomite at relevant temperatures. This implies that the δ18O of an ankerite would be ~4‰ lower than a dolomite formed from the same water (Chacko and Deines, 2008). Secondly, the amount of Fe and Mn as opposed to Mg in the carbonates most likely falls on a continuum between the dolomite and ankerite end-member compositions. In fact, pure ankerite [CaFe(CO3)2] isn't found in nature, (Davidson et al., 1993). Again, the impact of substitution of Fe and Mn for Mg in the dolomite structure on oxygen isotopic fractionation has not been measured empirically, but theoretically, replacing Fe with Mg should increase the fractionation, while replacing Fe with Mn should decrease it (Chacko and Deines, 2008). The elemental composition of the dolomites and ankerites in this interval has not been carefully determined, however, comparison of the relative intensities of the Ca, Mn, and Fe XRF 144 peaks allows further investigation of the relationship between mineralogy and δ18Ocarb. Interestingly, Ca/Mn and Ca/Fe ratios generally correspond to δ18Ocarb measurements in this interval (Figure B10). Generally, lower δ18Ocarb values are associated with lower Ca/Mn and Ca/Fe ratios, probably corresponding to more ankeritic carbonates, whereas samples with high Ca/Mn and Ca/Fe ratios are more enriched in 18O. Accounting for the ~4‰ offset between dolomite and ankerite would reduce the variability in δ18O in this interval, and suggest that generally, the apparently more ankeritic samples would have been formed in lake water with higher δ18O. This is consistent with the observation that over the history of the lake, ankerite is only found in intervals that other indicators suggest are uncommonly dry (Abebe, 2010). To compensate for the effect of mineralogy on carbonate δ18O, for the ankeritefree interval between 58.4 and 60.7 m RMCD, we used the fractionation equations for calcite (Kim and O'Neill, 1997), siderite (Carothers et al., 1988), and dolomite (Vasconcelos et al., 2005), to calculate the δ18Olw at the time of formation. Because the water-carbonate oxygen isotope fractionation is temperature dependent, changes in lake water temperature could complicate the δ18Olw reconstruction. To account for this, we estimate changes in δ18Olw over a 6°C range (23°C to 29°C). This is similar to the modern seasonal range in temperature, and annual mean temperatures are unlikely to have varied by 6°C over the course of the record. Furthermore, a period of lake level decline (enriching δ18Olw) would most likely be associated with warmer temperatures, which would have the opposite effect on δ18Ocarb. Ultimately, the inferred variability in δ18Olw is substantially larger than even a large temperature effect can explain (Figure B11), 145 reinforcing the assumption of past workers that carbonate δ18O records the composition of the lake water (Talbot and Kelts, 1986; Shanahan et al., 2009). 3.3 Varve thickness Varve thickness varies over several orders of magnitude over the study interval, from 0.03 to 70 mm. The thickest layers correspond to turbidites, and although they occur rarely, they make up a substantial portion of the sediment sequence. The turbidites are typically found in very finely laminated intervals, and represent sedimentation rates several orders of magnitude greater than nearby years. The extreme variability in varve thickness illustrates the non-linear relationship between hydrology and sedimentation at the core site. Varve thickness is controlled by several factors. On short timescales (i.e., < 30 years), varve thickness is probably dominated by changes in precipitation. During the instrumental record, varve thickness tracks precipitation well; more sediment is deposited at the core site during rainier years (Shanahan et al, 2008). On long timescales (> 100 years), this relationship is dominated by other factors. First, because of the small size of the drainage basin relative to the lake, changes in lake level have a large impact on the area available for sediment supply. For example, the drop in lake level from its early Holocene maximum to modern level corresponds to a 160% increase in the size of the non-lake drainage area (Turner et al., 1996a, Shanahan et al., 2006). Additionally, a decrease in lake level means that there is a reduced lake floor area over which sediment is distributed, and the core site, which is near the center of the lake, is nearer the shore. Both of these factors would also increase sedimentation rates at the core site, even 146 without a change in the volume of sediment entering the lake. Finally, drops in lake level may be associated with changes in the vegetation structure of the lake basin, varying between rainforest during wet intervals, and savannah and grassland flora during drier intervals. This along, with lake previously-deposited lake silts could increase the supply of available sediment, although this feedback is less certain. Altogether, the morphology of the lake basin is expected to drive non-linear increases in sedimentation rates during lowstands, and greatly reduced sedimentation during highstands. 3.4 XRF data The relative abundance of Al, Si, K, and Ti, as indicated by changes in the intensities of their XRF peaks, generally tracks varve thickness (Figures B9 and B11). The primary source of these elements in the lake sediments is terrigenous material that washes in from the drainage basin during the rain season, and every varve contains a layer rich in these elements (Shanahan et al., 2008; 2009). Because they are associated with minerals washing into the lake, as opposed to forming in the lake, we refer to these elements as “terrigenous elements”. The terrigenous elements are washed in from the lake basin, and their relative abundance is strongly influenced by the size of the drainage basin and the proximity of the core site to the shoreline, both of which are primarily functions of lake level. Therefore, the terrigenous elements are expected to vary inversely with lake level; most abundant when the lake is low, and least when it is high. Changes in dust flux would amplify this effect, further increasing the abundance of terrigenous material when lake level is low. The abundance of terrigenous elements is a robust lake level indicator 147 for the late Holocene, covarying with carbonate oxygen isotope values, geomorphic lake level indicators, and a variety of other forms evidence (Shanahan et al., 2009). In this study, we focus on the intensities of Ti as an indicator of terrigenous content, because it forms relative low-noise peaks even at short count times (unlike Al and Si), and is not complicated by changes in weathering and dust input (unlike Si and K; see section). Fe is also a terrigenous element, and is an important component of the sediment that washes into the lake. However, Fe is also present in minerals precipitated in the lake, or post-depositionally in the sediment, primarily manganosiderites and various species of iron-sulfide minerals (Shanahan 2006; Shanahan et al., 2008; 2009). This complicates the interpretation of the Fe intensities, and we consider them separately from the other terrigenous elements. Mn in the sediment record is present in several minerals, but is predominately associated with manganosiderites. Mn is also associated with iron-sulfide minerals and ankerite, but the intensity of the Mn peaks associated with these minerals is much lower (<100 cps) than with the manganosiderites (>1000 cps). Therefore, we consider Mn to be a robust indicator of the presence of manganosiderite. Ca XRF intensities primarily indicate the presence, and relative abundance of various species of calcium carbonates in the sediments (Shanahan 2006; Shanahan et al., 2008; 2009). There is very little Ca in the bedrock of the drainage, which is predominantly schist with small outcrops of granite (Koeberl et al., 1998), and the low intensities of Ca in the terrigenous layers of the sediment are consistent with this observation. The Ca that is incorporated in authigenic carbonates in the sediments is most likely delivered to the lake as dust. Dust deposited by northerly Harmattan winds is an 148 important component of Ghanaian soils (He et al., 2007), and is the likely origin of much of the dissolved Na in Lake Bosumtwi (Turner et al., 1996b). During the early part of the LIG, Ca forms distinct peaks associated with the annual, or semi-annual precipitation of calcium carbonates in the lake, after about 7000 varve years, however, the peaks become broader and smaller, reflecting the change from distinct carbonate layers to diffuse fine grained carbonates that are partially intermixed with terrigenous sediment and probably formed post-depositionally. Finally, the relative abundance of S in the sediment shows pronounced variability, and is present in three distinct components of the sediment. For most of the sediment, S XRF intensities are fairly low, tracking the abundance of organic matter in the sediment. Higher S intensities are associated with various the various ironsulfide minerals (pyrrhotite), and less frequently, large (1-3 mm) crystals of diagenetic gypsum found in units 2b and 2d. 4. Discussion 4.1 Hydrology during the interglacial Varve thickness, the abundance of terrigenous sediment as indicated by Ti, and δ18Olw all show the same main patterns over the 12 kyr-long interval and indicate a consistent, dynamic history of hydrologic variability during the Last Interglacial (Figure B11). Whereas there are a few varve records for the interval (e.g., Sirocko et al., 2005; Brauer et al., 2007; Seelos and Sirocko, 2007), the overwhelming majority of the paleoclimatic and paleoenvironmental work done on the LIG is comprised of paleoecological records from Europe (e.g., Woillard 1978; Klotz et al., 2004; Sirocko et 149 al., 2005; Kühl et al., 2007; Brewer et al., 2008). This work has identified three primary phases of climate during the interval (c.f., Brewer et al., 2009), and although there is little a priori reason to assume that West African climate and hydrology experienced a similar history, both regions are strongly influenced by North Atlantic SST, and the framework provides a convenient way to examine climate as it evolved through the the interglacial. 4.1.1 Early LIG (Varve years 1-5600; 128.6 – 123 ka) The beginning of the early LIG interval shows gradually decreasing varve thickness and terrigenous sediment abundance, corresponding with decreasing δ18Olw, all of which suggest a gradual increase in lake level. This corresponds to a gradual decrease in the abundance of well-rounded eolian quartz grains in the sediment (Figure B7), suggesting that the increase in lake level was accompanied by some combination of decreasing local dust supply, driven by increasing moisture in the region, and less intense, or less frequent, northerly winds. Such a decrease in northerly winds may be associated with a northward shift in the boreal winter position of the ITCZ. All of this is consistent with a gradual strengthening of the West African Monsoon during the early part of the interglacial. By 127 to 126.3 ka, varve thicknesses were generally very thin, although the Ti and the δ18Olw data are more indicative of intermediate lake level. Beginning 126.2 ka, varve thicknesses dramatically increase from less than 0.1 mm to more than 1 mm by 125.8 ka, before decreasing to thin varves (<0.1 mm) by 125 ka. This feature is among the most distinctive throughout the LIG varve record, and is generally supported by a 150 corresponding increase in inferred δ18Olw. Interestingly, this feature is not apparent in the average abundance of terrigenous material, which remains near the long term average during this interval. This means that although the flux of terrigenous material into the lake certainly increased during this interval (since flux is proportional to the product of average terrigenous input and sedimentation rate), the relative proportions of terrigenous material did not increase, which is typical of lower lake levels (Figure B11). The significance of this is not yet clear. Following the return to thin varves, all three proxies suggest an extended period of high lake level from 125 to 123 ka. This interval is characterized by organic-rich sediment, with abundant manganosiderite layers and occasional, thick turbidites. The presence of turbidites during intervals of high lake level is consistent with observations over longer timescales; turbidites are frequent during relatively wet intervals throughout at least the past 530 kyr (Appendix A). Most likely, the turbidites formed during brief, and perhaps relatively small, decreases in lake level that exposed deltas formed during higher lake level. These landforms would be unsupported and poorly consolidated, and would be particularly susceptible to mass wasting events. Taken together, the combination of thin varves, low Ti content, low δ18Olw and the presence of abundant manganosiderite and turbidites identify this interval as the longest period of sustained lake level highstand in the LIG. 4.1.2 Mid-LIG (Varve years 5600-9000; 123 – 119.6 ka) Lake level began to decrease following the sustained highstand from 125-123 ka, 151 with Ti and varve thickness increasing to intermediate levels. δ18Olw increased to the highest early LIG values by 122.3 ka, after which varve thickness and δ18Olw both decrease suggesting a return to somewhat higher lake level by 121.8 ka. Following this varve thickness and δ18Olw rapidly increase to local maxima around 121.1 ka. This transition is accompanied by an abrupt change in carbonate mineralogy from calcite (probably Mg-rich calcite) to dolomite, and calcite is not found in the rest of the LIG interval. All three lake level indicators suggest a return to relatively higher lake level following 121 ka, with lake level remaining relatively high for the rest of the mid-LIG, until about 119.6 ka. 4.1.3 Late LIG (Varve years 9000-12100; 119.6 – 116.5 ka) The beginning of the late LIG interval corresponds to an abrupt increase in varve thickness, and the onset of ankeritic carbonates in the sediment sequence. The carbonates are not pure ankerite, which doesn’t exist in nature, but most likely a represent carbonates of variable composition between the dolomite and ankerite end members, where the Mg of the dolomite is progressively substituted with Fe or Mn. Water-ankerite oxygen isotope fractionation factors have not been measured empirically, and the effect of partial substitution of Fe or Mn for Mg remains uncertain, although theoretical arguments suggest that increasing Fe and Mn should decrease the fractionation factor. Interestingly, the oxygen isotopes of the carbonates in this interval closely track both varve thickness and Ti, with thicker, more terrigenous varves corresponding to lower δ18Ocarb values, opposite of what is observed throughout the rest of the study interval (Figure B12). This 152 is consistent with the hypothesis that δ18Ocarb in this interval tracks the relative amount of Fe and Mn in the carbonate, and that carbonates with lower δ18Ocarb values are more ankeritic, and deposited during intervals of very low lake level (Abebe, 2010). This interpretation is consistent with varve thickness and Ti during these intervals. Mineralogical complications prevent calculation of the δ18Olw during this interval, however the inference that more ankeritic samples have lower δ18Ocarb and dolomitic samples have higher δ18Ocarb, due to changes in water-carbonate fractionation factors, suggests that δ18Olw was on the order of 9 to 11 per mil during the interval, the highest values in the LIG varve record. This suggests a highly evaporated water mass, and is consistent with the other indicators of very low lake level during at this time (Figure B12). Varve thickness and Ti intensities both show gradual increases through this interval, suggesting a slow decrease in lake level, interrupted by two multicentury-scale lowstands from 118 to 117.5 ka and 117.2 to 116.9 ka (Figure B12). The end of the interval is defined by increasingly poorly-defined, terrigenous-rich varves with disparate opaque organic layers, probably indicative of seasonally oxic conditions at the lake floor, consistent with a continued drop in lake level, until the varves are no longer recognizable. 4.2 Comparison between the LIG and the Holocene The most pronounced difference between the LIG and Holocene intervals of the Lake Bosumtwi sediment sequence (Figure B13) is the lack of an equivalent to the thick, organic-rich unit deposited during the early Holocene (often referred to as the “sapropel”; 153 Talbot et al., 1984). Geomorphic evidence and and exposed shorelines indicate that the lake overflowed frequently during the early Holocene, from ca. 11 to 9 ka, and maintained very high lake levels until 3.2 ka (Shanahan et al., 2006). This interval of high lake level is distinct in the sediment sequence; characterized by the lack of varves, greatly reduced precipitation of carbonates, and a dramatic change in the aquatic biology, as indicated by δ15N (Talbot and Johannessen, 1992). There is no LIG equivalent of the Holocene highstand sediment (“sapropel”). The organic-rich interval from 59.7 to 59.1 m RMCD, corresponding to 124.9 to 122.8 ka (700-5800 varve years), is the most similar, but the laminations are still apparent, although very thin (often <50 μm). The presence of a very thin terrigenous component in these varves suggest a much smaller, but still active, terrestrial watershed, consistent with a high, but not overflowing lake. Manganosiderite is much more abundant in the LIG interval than in the Holocene highstand sediment, another indication that the lake did not overflow during the LIG. Dissolved Fe and Mn were likely abundant in the deep anoxic water masses during both intervals, however, closed-lake conditions throughout the LIG would have kept the alkalinity and concentration of HCO3 high, resulting in more carbonate precipitation. The final indication that the wettest part of the LIG was not overflowing is the inferred δ18Olw from the carbonates in the interval. There are no δ18Olw measurements from the early Holocene, but the lowest δ18Olw values in the LIG (2.5 to 3.5 ‰) are comparable to similar values derived from calcites precipitated between 2.7 and 1.7 ka, after the lake had fallen dramatically from the early Holocene highstand (Shanahan et al., 2009). 154 4.2.1 Relationship with solar insolation Unlike what has been observed in parts of the Asian (Wang et al., 2001; Wang et al., 2008) and South American (Cruz et al., 2005; Cruz et al., 2009) Monsoon regions, the West African Monsoon, at least in region of Lake Bosumtwi, does not vary in phase with changes in summer insolation (Appendix A; Shanahan, 2006; Peck et al., in prep). During the Holocene lake level remained very high for several thousand years after summer insolation began to decrease, not falling until insolation was nearing its minima 2.7 ka (Figure B13) . Lake level also appears to have lagged insolation during the LIG as well, reaching its maximum highstand 2 – 4 ka after the summer insolation maxima 127 ka (Figure B13). Lake level does appear to have decreased along with decreasing insolation between 123 and 121 ka, before recovering to higher lake level between 121 and 120 ka despite continued decrease in insolation. Around 120 ka, increases in varve thickness and the appearance of ankerite in the sediment suggest an abrupt decrease in lake level, although this is not well-recorded in Ti intensities (Figure B11). This occurred at about the same phase of the precession cycle (2-3 ka before the northern Hemisphere summer insolation minima) as the abrupt fall during the late Holocene. This may suggest a nonlinear response of the West African Monsoon to gradual decreases in summer insolation, although the uncertainty in the chronology during the LIG interval makes it impossible to compare the phasings precisely. Interestingly, both interglacials had two millennial-scale peaks in lake level, at 109 ka and 4-3 ka in the Holocene and from 124-123 and 121-120 ka in the LIG. These peaks both occurred at around the same phasing within the precession cycle during both 155 interglacials, the first peak corresponding to maximum 10N insolation in July, the second peak corresponding to maximum insolation in October (Figure B13). Sub-precessional periodicities in precipitation are predicted and observed for equatorial regions that experience double maximums in rainfall associated with the twice-yearly passing of the tropical rainbelt and ITCZ (Short et al., 1991; Trauth et al, 2003; Berger et al., 2006; Verschuren et al., 2009). Despite this, the climate dynamics associated with these subprecessional peaks and troughs remain somewhat uncertain, and most likely vary regionally. Trauth et al. (2003) suggested that lake level high stands at equatorial Lake Naivasha coincided with March and September local insolation maxima, corresponding to the timing of the two rainy seasons in the modern climatology. This would imply that higher insolation was driving enhanced convection and onshore moisture transport, and thus, an intensification of the East African Monsoon during whatever season has higher than normal solar insolation. Verschuren et al. (2009), came to a contradictory conclusion based on a record from nearby Lake Challa, where subprecessional lake level highstands correspond to insolation maxima near the solstices, and lowstands when insolation maxima occurred near the equinoxes. The authors suggested that onshore moisture transport was highest when either the Northern or the the Southern Hemispheric components of the East African Monsoon was strongest, and that the very low insolation during the other rainy season when March or September insolation may have contributed to aridity during those intervals. Lake Bosumtwi also experiences two distinct rainy seasons, centered on June and October (Figure B2). The two millennial-scale highstands in each interglacial correspond 156 to the periods of maximum insolation during the two rainy seasons (Figure B13). Furthermore, in both interglacials, the first of the two highstand intervals were longer and are associated with higher lake level than the second highstand. The same pattern is also observed in modern precipitation climatology; the first rainy season is both longer and provides more precipitation than the second rainy season (Figure B13). It is reasonable to assume that an intensification of a longer and wetter rainy season would result in a longer and wetter highstand, although how the relative amplitudes and durations of the rainy seasons may have varied in the past is uncertain. 4.3 Relationships with sea surface temperature records There are very few records of sea surface temperature in the region for the LIG, and only a Mg/Ca-based SST record from the Gulf of Guinea (Weldeab et al., 2007) is sufficiently well-resolved to allow comparison of the variability within the interval. The floating varve chronology is pinned to this record based on an abrupt event 110 ka, so although some uncertainty remains in the absolute timing of the records, relative changes between the two should be comparable. Generally speaking, varve thickness, Ti intensities and carbonate mineralogy and isotope geochemistry covary with SSTs, with warmer SSTs co-occurring with periods of higher lake level (Figure B11). The SST record shows a gradual cooling trend after 127 ka that is consistent with gradual drying in the Lake Bosumtwi region, as shown by increases in varve thickness and δ18Olw, but this is not reflected in the abundance of Ti, possibly due to dilution by authigenic minerals during the interval. A local minimum in SSTs occurred between 122 and 120 ka, 157 coincident with the drier conditions in between the two LIG highstands. This decrease in SST is consistent with a northward shifted maximum position of the ITCZ, which would drive increased upwelling off the Guinean Coast and potentially drier conditions at Lake Bosumtwi. During the late Holocene, variability in North Atlantic SSTs played a prominent role in driving hydrologic variability at Lake Bosumtwi (Shanahan et al., 2009). High resolution SST data for the LIG from the North Atlantic are also sparse, however, SST records from the western Mediterranean and Iberian Margin regions (Martrat et al., 2004; 2007) and from the eastern subpolar and western North Atlantic (Lehman et al., 2002; Oppo et al., 2006), provide some context for the interval (Figure B14). Both regions experienced rapid warming between 128 and 127 ka, followed by generally warm temperatures until about 122 ka before gradually cooling through the rest of the interval (Figure B14). The initial warming apparent in the SST data from all three regions corresponds to a period of low varve thickness, and perhaps wet conditions at Lake Bosumtwi, but gaps in the XRF and δ18Olw data during this interval prevent further conclusions. Interestingly, SSTs in both the Gulf of Guinea and Iberian Peninsula region suggest cooler conditions from about 127 to 125.5 ka, at about the same time as a pronounced, gradual increase and decrease in varve thickness, that is partly corroborated by δ18Olw data. The low resolution and apparent noise in the North Atlantic records make it impossible to evaluate potential linkages on shorter timescales. 158 4.4 An multi-century aridity event 118 ka Immediately after 118 ka, Ti intensity and varve thickness both rapidly increase in less than 100 years (Figure B12), corresponding with an abrupt decrease in SSTs in the Gulf of Guinea (Figure B11). Carbonate mineralogy and geochemistry also shows abrupt changes a this interval, with an abrupt decrease in δ18Ocarb most likely associated with a transition towards more ankeritic carbonate. The hypothesis that lower δ18Ocarb values is consistent with differences between dolomite and ankerite fractionation factors (Chacko and Deines, 2008), but has yet to be tested for this interval. Ankerite is present in the two XRD samples measured in the interval, corresponding to ca. 117.3 and 117.1 ka (Abebe, 2010). Throughout the 1.08 Myr record, ankerite is only rarely found in Lake Bosumtwi sediments, and is consistently associated with extreme aridity (Abebe, 2010; Peck et al., in prep.). The arid interval lasted about 400-500 years, before recovering somewhat around 117.5 ka (Figure B12). Afterward, varve thicknesses gradually increase towards the top of the LIG section, and Ti counts remain high. δ18Ocarb also increases, although whether this is due to continued drying of the lake driving an increase in δ18Ocarb, or a shift in mineralogy toward more dolomitic carbonates is unclear. The arid interval is consistent in timing and duration with the Late Eemian Aridity Pulse (LEAP), which has been observed in some European records, most notably those from the Eifel Maar complex (Figure B12; Sirocko et al., 2005; Seelos and Sirocko, 2007). In northern Germany, LEAP corresponds to about 500 years of greatly increased loess deposition, suggesting much colder, and more arid conditions in the region at the time, although the precise timing and duration of the event vary from site to site (Sirocko 159 and Seelos, 2007). LEAP most likely corresponds to the North Atlantic cooling event C26, which is observed in some, but not all SST records from the North Atlantic, and may be associated with a small peak in ice rafted debris at the time (Lehman et al., 2002; Sirocko et al., 2005; Oppo et al., 2006). Sirocko et al., (2005) hypothesize that LEAP, and C26, resulted from a slowdown of AMOC, or at least a southward shift of the northern limit of Atlantic heat transport. A reduction in AMOC, and a cooling of North Atlantic would likely result in aridity in West Africa (Street-Perrott and Perrot, 1990; Chang et al., 2008; Mulitza et al., 2008; Shanahan et al., 2009), so it is reasonable to assume that the aridity observed near the end of the LIG varve record may be the West African expression of LEAP, although uncertainty in the ages of the records make it impossible to link the two records conclusively. 4.5 Drought frequency during the LIG Both varve thickness and Ti counts (Figure B11), as well as other XRF counts from other terrigenous elements (Figure B9), show pronounced decadal to century scale variability during the LIG. To investigate the timescales of variability in the records, as well as the influence of uncertainty in the varve chronologies, spectral estimates were calculated using the multi-taper method (MTM; Mann and Lees, 1996) for all of the ensemble members of the varve thickness, and Ti count records (Figure B15). Confidence limits were calculated for each ensemble member using the “Robust” noise estimation method of Mann and Lees (1996), the median 90 and 95% confidence limits are shown (Figure B15), but there is little variability in the confidence levels between ensemble 160 members. The Monte Carlo MTM spectra (MC-MTM) for both Ti and varve thickness (Figure B15) show distinct changes in slope around frequencies of 1/100 years. This “kink” in the spectra is a common feature of paleotemperature records, and may reflect the midpoint between the predominant influence of the influence of the annual cycle at higher frequencies, and Milankovitch cycles at lower frequencies (Huybers and Curry, 2006). Although Huybers and Curry restricted their analysis to temperature records, it is not surprising to find a similar feature at Lake Bosumtwi, since hydrology at the lake is greatly influenced by annual and orbital cycles. The median MC-MTM spectra for both varve thickness and Ti do not have any significant peaks at frequencies higher than 1/50 yrs, although there are significant peaks in each of the individual ensemble members. On the other hand, whereas individual ensemble members may have a grouping of individual significant peaks, the median estimate shows broader bands of relatively higher variance, such as the peaks centered on periodicities of 19, 16, and 14 years in Ti, and the peaks at 28 and 20 years in varve thickness (Figure B15). The diffusion of high frequency peaks is a predictable result of incorporating age model uncertainty, both because of the relatively greater influence of small changes in the age model, and because each higher frequency spectral estimate corresponds to narrower window of time. The broad bands of power at multidecadal frequencies are not significant, but most likely reflect real modes variability in the record, that are either obscured by age model uncertainty, or are actually non-periodic and really do occur over a range of frequencies. To investigate the latter hypothesis, we calculated 161 evolutive MC-MTM spectra for Ti and varve thickness to analyze changes in periodicities over the course of the varve sequence. The evolutive MC-MTM (EvMC-MTM) spectra were calculated using the same method as the MC-MTM spectra, but over nine, partially overlapping, 2000-yr-windows for each series, with the exception of the first window for Ti, which ends at the start of the gap in XRF data coverage. Both series show concentrations of multidecadal to centennial-scale variance, including many that are significant at the 95% confidence level (Figures B16 and B17). In both EvMC-MTM spectra, however, the periodicities vary over the course of the interglacial. For example, significant peaks, or nearly significant peaks are found at periodicities between 37 and 33 years, and 77 and 67 years in Ti, and 32-27, 61-53, and 83-71 years in varve thickness, but the periodicities and relative amplitudes of the peaks vary greatly between windows. This suggests the presence of a non-periodic, but preferred range of timescales of hydrological variability at the lake. Alternatively, bias in the varve chronology could also result in such variability, distorting a truly periodic signal. However, the same phenomenon is observed at lower frequencies, which should be less susceptible to age bias. In the Ti EvMC-MTM spectra, peaks with periodicities from 120-95, and 220 to 170 years, and peaks from 180 to 130 years in the log varve thickness record show similar changes in frequency and relative amplitudes (Figures B16 and B17). Multichannel Singular Spectrum Analysis (M-SSA) was performed on the Ti and log varve thickness datasets, such that each input was a timeseries ensemble member (Ghil et al., 2002). This non-parametric approach identifies the primary oscillatory modes 162 of variability in the time series that are robust to uncertainty in the varve chronologies. Because we are particularly interested in multidecadal to centennial (M2C) variability, we detrended the series with a spline before performing the analysis to remove millennialscale variability, The M-SSA was performed with an embedding window of 400 yrs, and with 50 ensemble members. The full set of ensemble members could not be used because the procedure is extremely computationally expensive, even with high-performance computing capabilities. Generally, the results support the MC-MTM and EvMC-MTM analyses, indicating the predominance of M2C variability, especially periodicities around 110 and 275 yrs (Table B1). The M-SSA analyses on the ensembles without marker layers suggested more variance concentrated at high periodicities, likely due to increased variability between ensemble members in that set. For both types of ensembles, some of the predominance of lower frequencies likely reflects chronological uncertainty, which has a greater impact on higher frequency variability. That said, the Ti ensemble with marker layers has several important reconstructed components with multidecadal-scale variability, including those with periodicities of ~70, 55 and 24 years, consistent with the EvMC-MTM analysis for Ti (Figure B16). The finding that Ti abundance and log varve thickness show pronounced M2C variability, but over a range of frequencies, is consistent with the finding that changes in Atlantic sea surface temperature structure controlled by the AMO drove M2C drought variability during the late Holocene (Shanahan, 2006). The origin of the AMO remains a topic of debate; with hypotheses ranging from a global-scale series of ocean-atmospheresea ice teleconnections (Dima and Lohman, 2006), to internal variability in Atlantic 163 meridional overturning circulation (AMOC) (Knight et al., 2005), to the possibility that the AMO in the instrumental record is an artifact of the 20th century warming trend and two abrupt cooling events (Thompson et al., 2010), although this last hypothesis seems unlikely, given that there is evidence of AMO variability for at least the past several hundred years (Gray et al., 2004; Shanahan et al., 2009). The AMO is typically simulated in state-of-the-art climate models, but the underlying mechanism, as well as the timescales of variance vary from model to model (Danabasoglu, 2008). In CCSM4, AMO variability is associated with changes in AMOC and covers a broad low frequency spectrum (50-200 yr periodicities) (Danabasoglu et al., in press). The variability in CCSM4 is associated with deep-water formation in the Labrador Sea, and may be associated with the NAO, however the origin of the lower frequency time scale remains unclear (Danabasoglu et al., in press). Interestingly, in CCSM4, a externally-forced (with solar and volcanic variability) last millennium simulation shows substantially more multidecadal-scale variability in the AMO than the long control run (Landrum et al., submitted). The finding that Lake Bosumtwi hydrology varied with periods around 30, 50, 70 and 150 yrs during the LIG suggests that AMO, and AMOC variability during the Last Interglacial may have been similar to the late Holocene, despite the substantial differences in seasonal insolation forcing. That said, the lack of a well-defined timescale of variability and uncertainty in the underlying mechanism prevents further analysis. 4.5.1 Influence of solar variability The similarities between paleorecords of tropical and subtropical hydrology and 164 indicators of past solar variability (e.g., 14C, 10Be), have motivated the hypothesis that M2C variability in solar insolation drives changes in these regions (e.g., deMenocal, 2001; Hodell et al., 2001; Cook et al., 2004, Shindell et al., 2006, Meehl et al., 2009) although recent syntheses of global climate variability and cold events during the Holocene concluded that long-term solar-variability was not consistently associated with climate on global scale (Wanner et al., 2008; 2011). Modelling studies have suggested that certain regions may be more vulnerable than others to solar variability, especially the hydrology of the tropics, due both to direct heating of the land surface and changes the ozone chemistry response to altered ultraviolet radiation (Shindell et al., 1999; Shindell and Schmidt, 2004). Furthermore, tropical Atlantic SSTs and AMOC respond to simulated solar variability as well (Goosse and Renssen, 2004; Weber et al., 2004; Cubasch et al., 2005). Altogether, this indicates that Lake Bosumtwi should be particularly well situated to record the climatic influence of solar variability. The Holocene record from Lake Bosumtwi offered limited support for the hypothesis that solar variability modulates hydrology in West Africa. Shanahan (2006) found coherent variability between terrigenous sediment abundance and reconstructed atmospheric 14C concentrations at several frequencies, including inferred solar periodicities of 120 and 220 yrs. During the LIG, however, varve thickness and Ti abundance do not show significant variability at these periodicities. As discussed above, M2C variability during the LIG varied in periodicity and amplitude throughout the interglacial. It remains uncertain whether or not solar variability is periodic, or whether solar variability during the LIG occurred at the same periodicities as observed during the Holocene. It is clear, 165 however, that there is no noticeable concentration of spectral power at known modern solar periodicities, that is, the 11-year sunspot cycle, the 87-year Gleissberg cycle, or the 210-year DeVries cycle (Figures B16 and B17; Braun et al., 2005). Ultimately, the role of solar variability during the LIG remains unclear. 5. Conclusions and summary 5.1 Primary findings  In this study we presented a new, 12,100 year varve record from Lake Bosumtwi that spans an interval that represents the Last Interglacial period. Radiometric and cosmogenic age control points from within or near the study interval do not have sufficient precision to accurately “pin” the floating varve chronology, however, correlation between global and regional dust and SST records suggest that the varve section spans the interval from 128.6 to 116.5 ka. This is consistent nearby U/Th and OSL ages, as well the Blake geomagnetic excursion (123 +/- 3 ka; Lund et al., 2006) found within the varved interval (Heil et al., in prep).  A new approach was used to account for uncertainty in the varve count age model: Following the approach of Rasmussen et al., (2006) in layer counting the NGRIP ice core, ambiguous (possibly annual, possibly subannual) layers were recorded as 0.5 +/- 0.5 years. Varve count, thickness, and XRF ensembles were created by using the four separate varve counts (two on each core) and allowing the uncertain years to vary stochastically 166 between marker layers. For spectral analysis, a second set of ensembles were created that allowed the gaps between marker layers to vary as well. This approach allows quantification of the effect of counting uncertainty on the resulting record and analyses.  Varve thickness and XRF titanium intensities generally agree with the δ18O of lake water (δ18Olw), which was inferred from carbonate isotope geochemistry and mineralogy over a range of temperatures. All three of these parameters are sensitive lake level, and vary substantially throughout the interglacial.  The maximum LIG highstand was both smaller and shorter-lived than the the prolonged highstand in the early Holocene, and unlike the Holocene, the lake level indicators, including the continuous presence of varves through the interval, suggests that the lake never overflowed during LIG.  The LIG, like the Holocene, had two distinct millennial-scale moist intervals, from 125 - 123 and 121 – 120 ka. In both the LIG and the Holocene, these peaks occurred during times of precession-driven insolation maxima in July and October, corresponding to the two rainy seasons in the modern climatology. This suggests that, at least during interglacials, prolonged wet conditions occur at the lake when rainy season insolation is highest. Furthermore, in both interglacials, the larger highstand occurred during the July insolation maxima, which is the more substantial of the two rainy seasons. 167  Over the course of the LIG, lake level generally tracks sea surface temperatures (SST) in Gulf of Guinea, including an abrupt drop in lake level that lasted about 500 years ca. 118 ka, corresponding to cooling in the Gulf of Guinea and much of the North Atlantic during the C26 cooling event. The timing and duration of the event are comparable to the Late Eemian Aridity Pulse (LEAP; Sirocko et al.,, 2005) that is observed in several sites in Europe, most notably several lakes in the Eifel Maar complex (Sirocko et al., 2005; Seelos and Sirocko, 2007). The event is interpreted to have resulted from abrupt cooling in the North Atlantic, and possibly a reduction of Atlantic meridional overturning circulation. These two phenomena are known related to West African aridity over a range of timescales, so it seems likely that the 500-yr-long arid interval around 118 ka is responding to the same forcing.  The occurrence of quasiperiodic variability at multidecadal to centennial timescales is consistent with the hypothesis that slow-varying changes in Atlantic SST structure (e.g., the Atlantic Multidecadal Oscillation) drives long term hydrologic variability in the region. The periodicities vary over the course of the record, further implicating the role ocean circulation. This does not support the hypothesis that solar variability plays an important, direct role in regional hydrology, especially because there is no concentration of spectral power at well known periodicities of solar variability. 168 5.2 Implications for future hydrology in the region If the Last Interglacial holds as a partial analog for future warming in the region, the varve record for the interval has several important implications for the future in the region. One robust comparison between future projections and the LIG is the increase in Northern Hemisphere summer heating, especially on land, relative to the rest of the world. This asymmetric heating affects the hemispheric heat balance during the monsoon season and would likely effect the timing and intensity of monsoon rains throughout West Africa. When this occurred during the Last Interglacial, Lake Bosumtwi, and probably most of the Guinea Coast region, was substantially drier, on average, than during the Holocene. Most evidence from the Sahel (and Sahara) during the interval suggests wetterthan-Holocene conditions during the LIG (Drake et al., 2011). This scenario is precisely what is predicted for the 21st century by the AR4 MIROC model, one of the four AR4 models which simulated the WAM well. Whereas the mean state was drier in the Guinea Coast region, drought variability during the Eemian was similar to the late Holocene, dominated by multidecadal to centennial scale droughts and pluvials. Further extending the LIG analogy to the future would suggest similar variability throughout coming centuries superimposed on generally drier conditions on the Guinea Coast, and perhaps generally wetter conditions in the Sahel. That said, drought variability is most likely linked to variability in Atlantic SST and circulation, and the impact of future warming on the AMO and AMOC remains uncertain. The results presented here do imply that variability in the system was robust to the changes in seasonal forcing during the LIG. 169 Finally, the occurrence of an abrupt aridity event at the end of the interglacial, that lasted about 500 years, and may be correlative with widespread aridity in Europe, highlights the capacity of the system to rapidly fall into severe, prolonged drought. This drought appears to have been far more severe than the “Sahel Drought”, and occurred during an interglacial. The exact cause and the likelihood of such an event remain uncertain, the fact that such widespread drought can occur during conditions similar to today, justifies further investigation into the underlying mechanisms and the potential of such an event in the future. 6. References Abebe, N. T. (2010), Paleohydrology of West Africa Using Carbonate, Detrital and Diagenetic Minerals of Lake Bosumtwi, Ghana, M.S., University of Akron. Benson, C., and E. J. Clay (1998), The impact of drought on sub-Saharan African economies: a preliminary examination, World Bank Publications. Berger, A., and M. F. Loutre (1991), Insolation values for the climate of the last 10 million years, Quaternary Science Reviews, 10(4), 297–317. Bergner, A. G. N., M. H. Trauth, and B. Bookhagen (2003), Paleoprecipitation estimates for the Lake Naivasha basin (Kenya) during the last 175 ky using a lake-balance model, Global and Planetary Change, 36(1-2), 117–136. Brauer, A., J. R. . Allen, J. Mingram, P. Dulski, S. Wulf, and B. Huntley (2007), Evidence for Last Interglacial chronology and environmental change from Southern Europe, Proceedings of the National Academy of Sciences, 104(2), 450. Braun, H., M. Christl, S. Rahmstorf, A. Ganopolski, A. Mangini, C. Kubatzki, K. Roth, and B. Kromer (2005), Possible solar origin of the 1,470-year glacial climate cycle demonstrated in a coupled model, Nature, 438(7065), 208–211. Brewer, S., J. Guiot, M. F. Sánchez-Goņi, and S. Klotz (2008), The climate in Europe during the Eemian: a multi-method approach using pollen data, Quaternary 170 Science Reviews, 27(25-26), 2303–2315. Carothers, W. W., L. H. Adami, and R. J. Rosenbauer (1988), Experimental oxygen isotope fractionation between siderite-water and phosphoric acid liberated CO2siderite, Geochimica et Cosmochimica Acta, 52(10), 2445–2450. Castañeda, I. S., S. Mulitza, E. Schefuss, R. A. . dos Santos, J. S. . Damsté, and S. Schouten (2009), Wet phases in the Sahara/Sahel region and human migration patterns in North Africa, Proceedings of the National Academy of Sciences, 106(48), 20159–20163. Chacko, T., and P. Deines (2008), Theoretical calculation of oxygen isotope fractionation factors in carbonate systems, Geochimica et Cosmochimica Acta, 72(15), 3642– 3660. Chang, P., R. Zhang, W. Hazeleger, C. Wen, X. Wan, L. Ji, R. J. Haarsma, W.-P. Breugem, and H. Seidel (2008), Oceanic link between abrupt changes in the North Atlantic Ocean and the African monsoon, Nature Geoscience, 1(7), 444– 448, doi:10.1038/ngeo218. Chen, T. C., and H. van Loon (1987), Interannual variation of the tropical easterly jet, Monthly Weather Review, 115(8), 1739–1759. Christensen, J. H. et al. (2007), Regional climate projections, Climate Change, 2007: The Physical Science Basis. Contribution of Working group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, University Press, Cambridge, Chapter 11, 847–940. Cook, E. R., C. A. Woodhouse, C. M. Eakin, D. M. Meko, and D. W. Stahle (2004), Long-term aridity changes in the western United States, Science, 306(5698), 1015–1018. Cook, K. H., and E. K. Vizy (2006), Coupled model simulations of the West African monsoon system: Twentieth-and twenty-first-century simulations, Journal of climate, 19(15), 3681–3703. Cubasch, U., G. Bürger, I. Fast, T. Spangehl, and S. Wagner (2005), The direct solar influence on climate: modeling the lower atmosphere, Memorie della Societa Astronomica Italiana, 76, 810. Danabasoglu, G. (2008), On Multidecadal Variability of the Atlantic Meridional Overturning Circulation in the Community Climate System Model Version 3, Journal of Climate, 21(21), 5524–5544, doi:10.1175/2008JCLI2019.1. 171 Danabasoglu, G., Yeager, S.G., Kwon, Y.-O., Tribbia, J. J., Phillips, A. S., and Hurrell, J. W. (n.d.), Variability of the Atlantic Meridional Overturning Circulation in CCSM4., Journal of Climate. Davidson, P. M., G. H. Symmes, B. A. Cohen, R. J. Reeder, and D. H. Lindsley (1993), Synthesis of the new compound CaFe (CO3) 2 and experimental constraints on the (Ca, Fe) CO3 join, Geochimica et Cosmochimica Acta, 57(23-24), 5105– 5109. deMenocal, P. B. (2001), Cultural Responses to Climate Change During the Late Holocene, Science, 292(5517), 667–673, doi:10.1126/science.1059287. deMenocal, P. B., and others (1995), Plio-pleistocene African climate, Science, 270(5233), 53–59. deMenocal, P. B., W. F. Ruddiman, and E. M. Pokras (1993), Influences of high-and lowlatitude process on African terrestrial climate to variations in seasonal insolation, glacial ice cover, sea-surface temperature and Asian orography, Journal of Geophysical Research, 98, 7265–7287. Dezfuli, A. K., and S. E. Nicholson (2011), A note on long-term variations of the African easterly jet, International Journal of Climatology, 31(13), 2049–2054. Dima, M., and G. Lohmann (2007), A hemispheric mechanism for the Atlantic Multidecadal Oscillation, Journal of Climate, 20(11), 2706–2719. Drake, N. A., R. M. Blench, S. J. Armitage, C. S. Bristow, and K. H. White (2011), Ancient watercourses and biogeography of the Sahara explain the peopling of the desert, Proceedings of the National Academy of Sciences, 108(2), 458–462. Ghil, M. et al. (2002), Advanced spectral methods for climatic time series, Rev. Geophys, 40(1), 1003. Goosse, H., and H. Renssen (2004), Exciting natural modes of variability by solar and volcanic forcing: idealized and realistic experiments, Climate Dynamics, 23(2), 153–163. Gray, S. T., L. J. Graumlich, J. L. Betancourt, and G. T. Pederson (2004), A tree-ring based reconstruction of the Atlantic Multidecadal Oscillation since 1567 AD, Geophysical Research Letters, 31(12), L12205. Grodsky, S. A., J. A. Carton, and S. Nigam (2003), Near surface westerly wind jet in the Atlantic ITCZ, Geophys. Res. Lett, 30(19), 2009. 172 Hearty, P. J., J. T. Hollin, A. C. Neumann, M. J. O’Leary, and M. McCulloch (2007), Global sea-level fluctuations during the Last Interglaciation (MIS 5e), Quaternary Science Reviews, 26(17-18), 2090–2112. He, C., H. Breuning-Madsen, and T. W. Awadzi (2007), Mineralogy of dust deposited during the Harmattan season in Ghana, Danish Journal of Geography, 107. Heil, C., J. King, J. Peck, T. Shanahan, N. McKay, J. Overpeck, and C. Scholz, (in prep) Geochronology of the Lake Bosumtwi sediment sequence for the past million years, In preparation for Quaternary Science Reviews. Held, I., T. Delworth, J. Lu, K. Findell, and T. Knutson (2005), Simulation of Sahel drought in the 20th and 21st centuries, Proceedings of the National Academy of Sciences of the United States of America, 102(50), 17891. Held, I. M., and B. J. Soden (2006), Robust responses of the hydrological cycle to global warming, Journal of Climate, 19(21), 5686–5699. Herman, R. G., C. E. Bogdan, A. J. Sommer, and D. R. Simpson (1987), Discrimination among carbonate minerals by Raman spectroscopy using the laser microprobe, Applied spectroscopy, 41(3), 437–440. Hodell, D. A., M. Brenner, J. H. Curtis, and T. Guilderson (2001), Solar forcing of drought frequency in the Maya lowlands, Science, 292(5520), 1367–1370. Huybers, P., and W. Curry (2006), Links between annual, Milankovitch and continuum temperature variability, Nature, 441(7091), 329–332. Itambi, A. C., T. Von Dobeneck, S. Mulitza, T. Bickert, and D. Heslop (2009), Millennialscale northwest African droughts related to Heinrich events and DansgaardOeschger cycles: Evidence in marine sediments from offshore Senegal, Paleoceanography, 24(1), PA1205. Janicot, S., S. Trzaska, and I. Poccard (2001), Summer Sahel-ENSO teleconnection and decadal time scale SST variations, Climate Dynamics, 18(3), 303–320. Jourdan, F., P. R. Renne, and W. U. Reimold (2009), An appraisal of the ages of terrestrial impact structures, Earth and Planetary Science Letters, 286(1), 1–13. Kim, S., and J. Oneil (1997), Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates, Geochimica et Cosmochimica Acta, 61, 3461–3475. Klotz, S., U. Müller, V. Mosbrugger, J.-L. de Beaulieu, and M. Reille (2004), Eemian to early Würmian climate dynamics: history and pattern of changes in Central 173 Europe, Palaeogeography, Palaeoclimatology, Palaeoecology, 211(1–2), 107–126, doi:10.1016/j.palaeo.2004.04.009. Knight, J. R., R. J. Allan, C. K. Folland, M. Vellinga, and M. E. Mann (2005), A signature of persistent natural thermohaline circulation cycles in observed climate, Geophys. Res. Lett, 32(L20708). Koeberl, C. et al. (2007), An international and multidisciplinary drilling project into a young complex impact structure: The 2004 ICDP Bosumtwi Crater Drilling Project—An overview, Meteoritics & Planetary Science, 42(4-5), 483–511. Koeberl, C., R. Bottomley, B. P. Glass, and D. Storzer (1997), Geochemistry and age of Ivory Coast tektites and microtektites, Geochimica et Cosmochimica Acta, 61(8), 1745–1772. Koeberl, C., W. U. Reimold, J. D. Blum, and C. P. Chamberlain (1998), Petrology and geochemistry of target rocks from the Bosumtwi impact structure, Ghana, and comparison with Ivory Coast tektites, Geochimica et Cosmochimica Acta, 62(12), 2179–2196. Kopp, R. E., F. J. Simons, J. X. Mitrovica, A. C. Maloof, and M. Oppenheimer (2009), Probabilistic assessment of sea level during the Last Interglacial stage, Nature, 462(7275), 863–867. Kühl, N., T. Litt, C. Schölzel, and A. Hense (2007), Eemian and Early Weichselian temperature and precipitation variability in northern Germany, Quaternary Science Reviews, 26(25-28), 3311–3317. Lambert, F., B. Delmonte, J. R. Petit, M. Bigler, P. R. Kaufmann, M. A. Hutterli, T. F. Stocker, U. Ruth, J. P. Steffensen, and V. Maggi (2008), Dust-climate couplings over the past 800,000 years from the EPICA Dome C ice core, Nature, 452(7187), 616–619. Lamb, P. J. (1978), Large-scale tropical Atlantic surface circulation patterns associated with Subsaharan weather anomalies, Tellus, 30, 240–251. Landrum, L., Otto-Bliesner, B.L., Conley, A., Lawrence., P., Rosenbloom, N., and Teng, H. (in review), Last Millennium Climate and Its Variability in CCSM4., Journal of Climate. Lotter, A. F., and G. Lemcke (1999), Methods for preparing and counting biochemical varves, Boreas, 28(2), 243–252. Lu, J., C. Deser, and T. Reichler (2009), Cause of the widening of the tropical belt since 174 1958, Geophysical Research Letters, 36, 5 PP., doi:200910.1029/2008GL036076. Lund, S., J. S. Stoner, J. E. T. Channell, and G. Acton (2006), A summary of Brunhes paleomagnetic field variability recorded in Ocean Drilling Program cores, Physics of the earth and planetary interiors, 156(3-4), 194–204. Lüthi, D. et al. (2008), High-resolution carbon dioxide concentration record 650,000– 800,000 years before present, Nature, 453(7193), 379–382. Mann, M. E., and J. M. Lees (1996), Robust estimation of background noise and signal detection in climatic time series, Climatic Change, 33(3), 409–445. Martrat, B., J. O. Grimalt, C. Lopez-Martinez, I. Cacho, F. J. Sierro, J. A. Flores, R. Zahn, M. Canals, J. H. Curtis, and D. A. Hodell (2004), Abrupt temperature changes in the western Mediterranean over the past 250,000 years, Science, 306(5702), 1762. Martrat, B., J. O. Grimalt, N. J. Shackleton, L. de Abreu, M. A. Hutterli, and T. F. Stocker (2007), Four climate cycles of recurring deep and surface water destabilizations on the Iberian margin, Science, 317(5837), 502. McKay, N. P., J. T. Overpeck, and B. L. Otto-Bliesner (2011), The role of ocean thermal expansion in Last Interglacial sea level rise, Geophysical Research Letters, 38(14), L14605. Meehl, G. A., J. M. Arblaster, K. Matthes, F. Sassi, and H. van Loon (2009), Amplifying the Pacific climate system response to a small 11-year solar cycle forcing, Science, 325(5944), 1114–1118. Mulitza, S., M. Prange, J. B. Stuut, M. Zabel, T. von Dobeneck, A. C. Itambi, J. Nizou, M. Schulz, and G. Wefer (2008), Sahel megadroughts triggered by glacial slowdowns of Atlantic meridional overturning, Paleoceanography, 23(4), PA4206. Nicholson, S. E. (2001), Climatic and environmental change in Africa during the last two centuries, Climate Research, 17(2), 123–144. Nicholson, S. E. (2008), The intensity, location and structure of the tropical rainbelt over west Africa as factors in interannual variability, International Journal of Climatology, 28(13). Nicholson, S. E. (2009a), A revised picture of the structure of the “monsoon” and land ITCZ over West Africa, Climate Dynamics, 29, doi:10.1007/s00382-008-0514-3. Nicholson, S. E. (2009b), On the factors modulating the intensity of the tropical rainbelt over West Africa, International Journal of Climatology. 175 Nicholson, S. E., and J. P. Grist (2001), A conceptual model for understanding rainfall variability in the West African Sahel on interannual and interdecadal timescales, International Journal of Climatology, 21(14), 1733–1757. Nicholson, S. E., and P. J. Webster (2007), A physical basis for the interannual variability of rainfall in the Sahel, Quarterly Journal of the Royal Meteorological Society, 133(629). Oppo, D. W., J. F. McManus, and J. L. Cullen (2006), Evolution and demise of the Last Interglacial warmth in the subpolar North Atlantic, Quaternary Science Reviews, 25(23-24), 3268–3277. Peck, J. A., R. R. Green, T. Shanahan, J. W. King, J. T. Overpeck, and C. A. Scholz (2004), A magnetic mineral record of Late Quaternary tropical climate variability from Lake Bosumtwi, Ghana, Palaeogeography, Palaeoclimatology, Palaeoecology, 215(1-2), 37–57. Peck, J., C. Koeberl, J. King, B. Milkereit, J. Overpeck, and C. A. Scholz (2005), The Lake Bosumtwi drilling project: Initial Report, Geological Society of America, Limnogeology Division Newsletter, 2(2), 3–7. Peck, J., N. McKay, T. Shanahan, C. Heil, J. Overpeck, J. King, and C. Scholz (in prep), A million-year history of the West African Monsoon, in preparation for Science. Pu, B., and K. H. Cook (2010), Dynamics of the West African westerly jet, Journal of Climate, 23(23), 6263–6276. Pu, B., and K. H. Cook (2012), Role of the West African Westerly Jet in Sahel Rainfall Variations, Journal of Climate, 25(8), 2880–2896, doi:10.1175/JCLI-D-1100394.1. Randall, D. A. et al. (2007), Climate models and their evaluation, Climate Change, 2007: The Physical Science Basis. Contribution of Working group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, University Press, Cambridge, Chapter 8, 589-662. Rasmussen, S. O. et al. (2006), A new Greenland ice core chronology for the last glacial termination, Journal of geophysical research, 111(D6), D06102. Rodríguez-Fonseca, B. et al. (2011), Interannual and decadal SST-forced responses of the West African monsoon, Atmospheric Science Letters, 12(1), 67–74. Rowell, D. P. (2003), The impact of Mediterranean SSTs on the Sahelian rainfall season, 176 Journal of Climate, 16(5), 849–862. Rutt, H. N., and J. H. Nicola (1974), Raman spectra of carbonates of calcite structure, Journal of Physics C: Solid State Physics, 7, 4522. Seelos, K., and F. Sirocko (2007), 14. Abrupt cooling events at the very end of the Last Interglacial, Developments in Quaternary Sciences, 7, 207–229. Shanahan, T., J. Beck, J. Overpeck, N. McKay, and J. Pigati, (submitted) Late Quaternary sedimentological and climate changes at Lake Bosumtwi Ghana: new constraints from laminae analysis and radiocarbon age modeling, Quaternary Science Reviews. Shanahan, T. M. (2006), West African monsoon variability from a high-resolution paleolimnologic record (Lake Bosumtwi, Ghana), University of Arizona. Shanahan, T. M., J. T. Overpeck, J. W. Beck, C. W. Wheeler, J. A. Peck, J. W. King, and C. A. Scholz (2008a), The formation of biogeochemical laminations in Lake Bosumtwi, Ghana, and their usefulness as indicators of past environmental changes, Journal of Paleolimnology, 40(1), 339–355. Shanahan, T. M., J. T. Overpeck, J. B. Hubeny, J. King, F. S. Hu, K. Hughen, G. Miller, and J. Black (2008b), Scanning micro-X-ray fluorescence elemental mapping: A new tool for the study of laminated sediment records, Geochemistry, Geophysics, Geosystems, 9(2), Q02016. Shanahan, T. M., J. T. Overpeck, W. E. Sharp, C. A. Scholz, and J. A. Arko (2007), Simulating the response of a closed-basin lake to recent climate changes in tropical West Africa (Lake Bosumtwi, Ghana), Hydrological Processes, 21(13). Shanahan, T. M., J. T. Overpeck, C. W. Wheeler, J. W. Beck, J. S. Pigati, M. R. Talbot, C. A. Scholz, J. Peck, and J. W. King (2006), Paleoclimatic variations in West Africa from a record of late Pleistocene and Holocene lake level stands of Lake Bosumtwi, Ghana, Palaeogeography, Palaeoclimatology, Palaeoecology, 242(3), 287–302. Shanahan, T., J. Overpeck, K. Anchukaitis, J. Beck, J. Cole, D. Dettman, J. Peck, C. Scholz, and J. King (2009), Atlantic forcing of persistent drought in West Africa, Science, 324(5925), 377. Shanahan, T., J. Peck, S. Forman, C. Heil, N. McKay, J. Overpeck, J. King, and C. Scholz (in prep), Geochronology of the Lake Bosumtwi sediment sequence for the past 200,000 years, in preparation for Quaternary Science Reviews. 177 Shindell, D. T., G. Faluvegi, R. L. Miller, G. A. Schmidt, J. E. Hansen, and S. Sun (2006), Solar and anthropogenic forcing of tropical hydrology, Geophys. Res. Lett, 33(24), L24706. Shindell, D. T., R. L. Miller, G. A. Schmidt, and L. Pandolfo (1999), Simulation of recent northern winter climate trends by greenhouse-gas forcing, Nature, 399(6735), 452–455. Shindell, D. T., and G. A. Schmidt (2004a), Southern Hemisphere climate response to ozone changes and greenhouse gas increases, Geophysical Research Letters, 31, 4 PP., doi:200410.1029/2004GL020724. Shindell, D. T., and G. A. Schmidt (2004b), Southern Hemisphere climate response to ozone changes and greenhouse gas increases, Geophysical Research Letters, 31(18), L18209. Short, D. A., J. G. Mengel, T. J. Crowley, W. T. Hyde, and G. R. North (1991), Filtering of Milankovitch cycles by Earth’s geography, Quaternary Research, 35(2), 157– 173. Sirocko, F., K. Seelos, K. Schaber, B. Rein, F. Dreher, M. Diehl, R. Lehne, K. Jäger, M. Krbetschek, and D. Degering (2005), A late Eemian aridity pulse in central Europe during the last glacial inception, Nature, 436(7052), 833–836. Stager, J. C., D. B. Ryves, B. M. Chase, and F. S. . Pausata (2011), Catastrophic drought in the Afro-Asian monsoon region during Heinrich event 1, Science, 331(6022), 1299. Street-Perrott, F. A., and R. A. Perrott (1990), Abrupt climate fluctuations in the tropics: the influence of Atlantic Ocean circulation, Nature, 343(6259), 607–612, doi:10.1038/343607a0. Talbot, M. R., and G. Delibrias (1977), Holocene variations in the level of Lake Bosumtwi, Ghana, Nature, 268(5622), 722–724. Talbot, M. R., and T. Johannessen (1992), A high resolution palaeoclimatic record for the last 27,500 years in tropical West Africa from the carbon and nitrogen isotopic composition of lacustrine organic matter, Earth and Planetary Science Letters, 110(1-4), 23–37. Talbot, M. R., and K. Kelts (1986), Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana, Geology, 14(11), 912–916. Talbot, M. R., D. A. Livingstone, P. G. Palmer, J. Maley, J. M. Melack, G. Delibrias, and 178 S. Gulliksen (1984), Preliminary results from sediment cores from Lake Bosumtwi, Ghana, Palaeoecology of Africa, 16, 173–192. Tarutani, T., R. N. Clayton, and T. K. Mayeda (1969), The effect of polymorphism and magnesium substitution on oxygen isotope fractionation between calcium carbonate and water, Geochimica et Cosmochimica Acta, 33(8), 987–996, doi:10.1016/0016-7037(69)90108-2. Thompson, D. W. J., J. M. Wallace, J. J. Kennedy, and P. D. Jones (2010), An abrupt drop in Northern Hemisphere sea surface temperature around 1970, Nature, 467(7314), 444–447. Tjallingii, R., M. Claussen, J. B. . Stuut, J. Fohlmeister, A. Jahn, T. Bickert, F. Lamy, and U. Röhl (2008), Coherent high-and low-latitude control of the northwest African hydrological balance, Nature Geoscience, 1(10), 670–675. Toggweiler, J. R., and J. Russell (2008), Ocean circulation in a warming climate, Nature, 451(7176), 286–288, doi:10.1038/nature06590. Trauth, M. H., A. L. Deino, A. G. . Bergner, and M. R. Strecker (2003), East African climate change and orbital forcing during the last 175 kyr BP, Earth and Planetary Science Letters, 206(3), 297–313. Turner, B. F., L. R. Gardner, and W. E. Sharp (1996a), The hydrology of Lake Bosumtwi, a climate-sensitive lake in Ghana, West Africa, Journal of Hydrology, 183(3-4), 243–261. Turner, B. F., L. R. Gardner, W. E. Sharp, and E. R. Blood (1996b), The geochemistry of Lake Bosumtwi, a hydrologically closed basin in the humid zone of tropical Ghana, Limnology and Oceanography, 1415–1424. Turney, C. S. ., and R. T. Jones (2010), Does the Agulhas Current amplify global temperatures during super-interglacials?, Journal of Quaternary Science, 25(6), 839–843. Vasconcelos, C., J. A. McKenzie, R. Warthmann, and S. M. Bernasconi (2005), Calibration of the δ18O paleothermometer for dolomite precipitated in microbial cultures and natural environments, Geology, 33(4), 317–320, doi:10.1130/G20992.1. Verschuren, D. et al. (2009), Half-precessional dynamics of monsoon rainfall near the East African Equator, Nature, 462(7273), 637–641. Wanner, H. et al. (2008), Mid-to Late Holocene climate change: an overview, Quaternary 179 Science Reviews, 27(19-20), 1791–1828. Wanner, H., O. Solomina, M. Grosjean, S. P. Ritz, and M. Jetel (2011), Structure and origin of Holocene cold events, Quaternary Science Reviews, 30(21–22), 3109– 3123, doi:10.1016/j.quascirev.2011.07.010. Weber, S. L., T. J. Crowley, and G. der Schrier (2004), Solar irradiance forcing of centennial climate variability during the Holocene, Climate Dynamics, 22(5), 539–553. Weldeab, S., D. W. Lea, R. R. Schneider, and N. Andersen (2007), 155,000 years of West African monsoon and ocean thermal evolution, Science, 316(5829), 1303. Woillard, G. M. (1978), Grande Pile peat bog: a continuous pollen record for the last 140,000 years, Quaternary Research, 9(1), 1–21. Wolff, E. W., J. Chappellaz, T. Blunier, S. O. Rasmussen, and A. Svensson (2010), Millennial-scale variability during the last glacial: The ice core record, Quaternary Science Reviews, 29(21–22), 2828–2838, doi:10.1016/j.quascirev.2009.10.013. 180 Table B1. Multi-channel singular spectrum analysis of the detrended Ti and log varve thickness time series, for ensembles with and without marker layers (ML). Ti (ML-ensemble) Ti (noML-ensemble) Reconstructed Variance Reconstructed Variance Component Period (yrs) exlained (%) Component Period (yrs) exlained (%) 1+2 650-750 12 1+2 600-800 20 3+4 240-270 6 3+4 270-330 6 5+6 150-180 3 5+6 230-250 5 7 90 1 7 – 10 160-190 5 8 70 1 11+12 125-145 2 9+10 60 2 13+14 100-110 1.3 11+12 110-120 2 15+16 230-250 1.3 13+14 54-55 1.6 15+16 62-63 1.4 17+18 24 1.4 Log of varve thickness (ML-ensemble) Log of varve thickness (noML-ensemble) Reconstructed Variance Reconstructed Variance Component Period (yrs) exlained (%) Component Period (yrs) exlained (%) 1 1200 29 1 1300 38 2 660 14 2 750 15 3 340 5 3 400 4 4 260 3 4 570 3 5 190 2 5 430 3 6 160 1.5 6–8 250-280 4 7 130 1 9+10 230 2 8 112 0.9 11+12 155 1.5 9 100 0.8 13-17 190-220 3 10+11 80-90 1.2 12+13 55-60 1.1 14+15 40-55 1 181 90 1°27' W 300 90 90 90 45 6°32' N 30 m 50 m 90 90 45 90 90 45 70 m 45 90 6°27' N 90 90 0 1 2 km Figure B1. Site map of the Lake Bosumtwi crater, modified from Shanahan (2006). Solid gray lines show bathymetry (10 m contour). Dashed line indicates crater rim. Inset: location of Lake Bosumtwi in southern Ghana, West Africa. 4.8 10 4.4 8 4.0 6 3.6 4 3.2 2 2.8 monthly rainfall (cm) Wind speed (m s-1) 182 0 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Figure B2. 1961-2000 climatology for Kumasi, Ghana, 30 km from Lake Bosumtwi. Grey bars indicate monthly precipitation totals (cm), dashed line shows wind speed (m s-1). Modified from Shanahan (2006). 183 Figure B3. Multichannel seismic reflection profile showing draped reflectors of the lacustrine fill and the brecciated bedrock of the central uplift (from Scholz et al., 2002). The 13 drill holes at the five drill sites are shown as black bars. Drill hole 5B is shown as a solid line, along with the down core median grain size profile at the site. Ti intensity (cps) 184 200 Itrax XRF μXRF 150 100 50 0 2000 4000 6000 8000 Varve Year 10000 12000 Figure B4. Comparison between Itrax and μXRF data for the study interval. The very high values between varve year 4000 and 6000 correspond to turbidites that are averaged into one year in the varve record. 185 Age model ensemble envelope Age model median estimate EDC-dust-tuned Age Model Blake Excursion (Top and Bottom) OSL ages U/Th ages 30 Depth (RMCD turbidite−adjusted) 35 40 Interpreted LIG Interval (varve section) 45 50 55 60 65 70 50 100 150 Age (ka) Figure B5. Age control for Lake Bosumtwi core 5B between 30 and 75 RMCD. Error bars on OSL and U/Th ages show 2 sigma uncertainty ranges. The yellow band corresponds to the of the floating chronology after "pinning" to the NGRIP (Wolff et al., 2010) chronology, see text for details. 186 Age (ka) −35 A 100 80 60 H7 GS25 40 H5.2 H6 H4 H5 20 H3 0 YD H2 H1 −6 −45 1 kyr −4 B −2 O lake water (‰) 18 δ Lake Bosumtwi 0 2 0 4 C 5 10 29 D 28 27 26 25 Gulf of Guinea SSS (psu) XRF−PC1 (Terrigenous) −40 22 24 24 23 26 Gulf of Guinea SST (°C) 18 NGRIP δ O (‰) 120 E 28 30 32 120 100 80 60 Age (ka) 40 20 0 Figure B6. A) δ18O of Greenland ice tracking North Atlantic temperature (NGRIP, 2004; Wolff et al., 2010). B) Lake Bosumtwi terrigenous input, an indicator of lake level (Appendix A). C) δ 18O of lake water inferred from authigenic carbonates in Lake Bosumtwi (this study). D) Sea surface temperature and E) salinity in the Gulf of Guinea (Weldeab et al., 2007).The broad agreement between the series, and especially the abrupt event ca. 110 ka corresponding to GS25 and GI24, suggest that the MIS 5 section of the Bosumtwi chronology is too young by about 1 ka. This suggests that the end of the varved interval (red circle) occurred ca. 116.5 ka. 187 Calcite Siderite Dolomite Dolomite/Ankerite Unit Varve year 11000 58 3a 10000 2d 9000 8000 2c Depth (RMCD) Core Description Laminations thicken upwards, minerogenic muds with diffuse dolomite layers, diagenetic FeS layers, and occasional well-rounded quartz grains 3b 59 Fine-grained FeS 12000 57.5 58.5 Mineralogy Diagenetic FeS Diagenetic Gypsum 2b Dark brown muds with abundant well-rounded quartz grains and diagenetic dolomite 7000 Finely laminated organic-rich mud, with packets of dolomite and occasional large (1 - 3 mm) diagenetic gypsum crsytals Coarsely laminated mineral-rich muds with abundant medium to coarse silt-sized quartz grains 6000 Organic-rich silts with abundant reddish, FeS-rich layers, with occasionaly calcites and diagenetic gypsum 5000 2a Very finely-laminated dark brown, organic rich muds interbedded with large coarse-grained turbidites, and manganosiderite 59.5 4000 Well-laminated with occasional diagenetic FeS layers and well-defined calcite layers. 1d 3000 60 1c Thick, poorly defined minerogenic laminae. 1b Well-laminated tan terrigenous layers alternating with calcite and organic material, with more reddish, finegrained FeS-rich lamina at the base 2000 1000 60.5 1a Dark brown finely-laminated with occasional turbidites, and abundant well-rounded quartz grains at its base 1 Figure B7. Generalized lithology, mineralogy and lamination structure of the study interval. 188 25 20 δ13Cpdb 15 10 Calcite Diagenetic Calcite Manganosiderite Dolomite Dolomite/Ankerite 5 0 0 2 4 6 8 10 12 14 18 δ Opdb Figure B8. Relationship between δ18O, δ13C , and carbonate mineralogy. 189 Varve Years 4000 6000 8000 10000 12000 −4 −2 50 Ti (cps) 0 100 20 40 60 80 100 120 140 10 20 200 4 2 x 10 1.5 K (cps) 100 30 300 1 10000 0.5 Mn (cps) Fe (cps) Si (cps) 2 150 Al (cps) Log varve thickness 2000 5000 0 60 0 40 S (cps) Ca (cps) 2000 20 2000 4000 6000 Varve Years 8000 10000 12000 Figure B9. 10-yr means of varve thickness and XRF ensemble members for Ti, Si, Al, K, Fe, Mn, Ca and S. The 5 to 95% annual ensemble envelopes are shown in light gray. Note that the axes for log varve thickness, Ti, Si, Al, and K are inverted. 190 Ca/Mn ratio 30 12 25 20 10 15 8 10 6 10500 9500 10000 Varve year 10 0.4 8 9000 13 C 12 11 11 δ Ocarb 10 18 18 δ Ocarb 0.6 12 9 8 7 12 0.8 0.2 5 9000 13 14 B carb 1 δ18O 14 A Ca/Fe ratio 35 9500 10000 Varve year 6 10500 D 10 9 8 10 15 20 25 Ca/Mn ratio 30 35 7 0.2 0.4 0.6 0.8 Ca/Fe ratio 1 Figure B10. XRF Ca/Mn (A) and Ca/Fe (B) ratios plotted along with the corresponding δ 18O values in the ankeritic interval of the record. (C) and (D) show the same data as (A) and (B), respectively, plotted against each other. High Ca/(Mn or Fe) ratios likely corresponds to more dolomitic carbonate, which typically is more enriched in δ18O, consistent oxygen isotope fractionation theory. Samples with lower elemental ratios and δ18O values are probably more ankeritic. 191 128 126 124 122 120 118 116 50 100 10 150 -1 Gulf of Guinea SST (°C) 10 0 10 28 Turbidites Varve thickness (mm) Lake Level 5 Variable ankerite content, 18 likely range of lake δ O Low Ti intensity (cps) 0 18 Lake water δ O High Age (ka) 27 26 128 126 124 Age (ka) 122 120 118 116 Figure B11. Lake level indicators for Lake Bosumtwi last interglacial varve record, along with Gulf of Guinea sea surface temperatures (Weldeab et al., 2007). The blue band corresponds to the range of possible lake water δ18O, given the carbonate δ18O, mineralogy, and a range of temperature. The rectangular band corresponds to an interval of variably dolomitic to ankeritic carbonate, which cannot be properly accounted for. However, the band corresponds to the best estimate of mean lake water δ18O for the interval. See text for details. Sharp, single year peaks in varve thickness are turbidites. 192 Age (ka) 119 118 117 13 Dolomite 120 11 18 80 Ankerite 9 Lake Level δ OCarb Ti (cps) High 40 120 −1 Low 10 0 10 Varve thickness (mm) 7 Loess content (%) 0 20 40 HL2 Ei2 RS1 60 120 119 118 117 Age (ka) Figure B12. Lake Bosumtwi Ti abundance, varve thickness and carbonate δ18O between 120 and 116.5 ka, along with the loess content of lake sediments from the Eifel Maar Complex (Sirocko et al., 2005; Seelos and Sirocko, 2007. Eifel Maar lakes West of Hoer List (HL2), Eigelback (Ei2), and Rederstall (RS1), record abrupt increases in less during the Late Eemian Aridity Pulse (LEAP), ca. 118 ka. The yellow bar indicates a ~500 yr interval of extremely dry conditions at Lake Bosumtwi that is consistent in timing with the LEAP. 128 126 124 122 Age (ka) 120 400 118 Lake overflowing 470 460 40 450 80 440 430 120 10 8 6 Age (ka) 4 2 0 420 440 420 400 420 400 380 Figure B13. Comparison of Lake Bosumtwi lake level and 10N rainy season insolation for July (solid red) and October, for the last interglacial (top) and the Holocene (bottom). Inferred lake water δ18O, and Ti intensity as in figure B11. Geomorphic evidence indicates that the lake was overflowing between 11 and 9 ka, despite somewhat higher terrigenous content in the sediment (Shanahan, 2006; Shanahan et al., 2006). October insolation (W m−2) 420 150 10 Ti (cps) Low 440 440 October insolation (W m−2) 8 100 July insolation (W m−2) 6 460 460 July insolation (W m−2) 4 480 50 Ti (cps) Inferred water δ18O 2 Lake Level High lw 193 194 128 126 124 122 120 118 116 50 5 100 150 0 North Atlantic (ODP 980) JJA SST (oC) 10 10 5 25 0 20 23 10 21 28 19 27 26 128 126 124 122 Age (ka) 120 118 Varve thickness (mm) -1 10 Western Mediterranean and Iberian Margin SST (oC) 10 Variable ankerite content, 18 likely range of lake δ O Western North Atlantic (MD95-2036) JJA SST (oC) Ti intensity (cps) 0 Gulf of Guinea SST (oC) Lake water δ18O Age (ka) 116 Figure B14. Same as figure B11, but with additional high resolution SST records included, from the western Mediterranean and Iberian Margin regions (Martrat et al., 2004; 2007), the eastern subpolar North Atlantic (Oppo et al., 2006), the western North Atlantic (Lehman et al., 2002) as well as Gulf of Guinea SSTs (Weldeab et al., 2007). 195 A 4 10 269 112 73 3 10 Power 59 42 34 28 19 17 95% 90% 14 2 10 0 0.02 0.04 0.06 0.08 0.1 0.12 Frequency (1/yr) B 1 10 279 154 Power 0 10 104 78 59 46 27 95% 90% 20 −1 10 0 0.02 0.04 0.06 0.08 0.1 0.12 Frequency (1/yr) Figure B15. Monte carlo MTM spectra for (A) Ti intensities and (B) the log of varve thickness. For both, the black line indicates the median power spectrum for all ensembles, and the gray envelope shows the 5 to 95% range of the spectra. Red dashed lines show the median 90 and 95% confidence levels The periodicites of pronounced peaks are labeled, rounded to the nearest years. 196 Ti intensity (cps) 100 200 Frequency (1/yr) 0.02 0.03 0.04 0.01 400 120 12000 150 10000 110 44 77 7765 Varve Year 120 71 12000 23 10000 49 37 33 140 100 74 250 23 33 48 130 95 77 8000 36 32 0.05 40 50 33 34 26 23 28 8000 22 Varve Year 0 29 500 6000 95 74 220 140 91 69 4000 6000 53 56 38 33 30 26 24 43 4000 25 23 220 120 71 57 35 27 21 2000 2000 180 110 67 56 0 100 Ti intensity (cps) 200 0.01 43 0.02 34 28 0.03 26 0.04 23 0.05 Frequency (1/yr) Figure B16. Evolutive spectra for Ti intensity. Left panel: median (black) and 95% confidence envelope (gray) of full ensemble (no marker layers) used to calculate spectra. Right panel: MC-MTM spectra, as in Figure B15, for partially overlapping 2000 year windows. Vertical position corresponds to mid pointof each window. Dot-dashed vertical blue lines corresponds to well-known frequencies of solar variability. Periodicities of pronounced peaks are labeled on each spectrum. 197 log varve thickness −3 −2 −1 0 1 12000 Frequency (1/yr) 0.02 0.03 0.04 0.01 400 140 110 83 0.05 12000 59 47 42 180 10000 110 74 57 10000 32 330 170 120 250 330 130 47 33 28 25 8000 83 49 130 83 27 53 6000 23 27 21 6000 Varve Year Varve Year 8000 74 59 290 80 61 45 27 36 21 4000 4000 290 130 61 28 45 140 110 77 2000 50 32 20 28 2000 170 110 80 −3 −2 −1 0 1 log varve thickness 0.01 59 49 43 31 29 0.02 0.03 0.04 Frequency (1/yr) 23 0.05 Figure B17. Evolutive spectra for log varve thickness. Left panel: median (black) and 95% confidence envelope (gray) of full ensemble (no marker layers) used to calculate spectra. Right panel: MC-MTM spectra, as in Figure B15, for partially overlapping 2000 year windows. Vertical position corresponds to mid pointof each window. Dot-dashed vertical blue lines corresponds to well-known frequencies of solar variability. Periodicities of pronounced peaks are labeled on each spectrum. 198 APPENDIX C THE ROLE OF OCEAN THERMAL EXPANSION IN LAST INTERGLACIAL SEA LEVEL RISE Published in the professional journal: Geophysical Research Letters Copyright 2011 American Geophysical Union Reproduced by permission of American Geophysical Union. McKay, N. P., J. T. Overpeck, and B. L. Otto-Bliesner, The role of ocean thermal expansion in Last Interglacial sea level rise, Geophysical Research Letters, 38(14), L14605, 2011. Copyright 2011 American Geophysical Union. 199 The Role of Ocean Thermal Expansion in Last Interglacial Sea Level Rise 1* Nicholas P. McKay , Jonathan T. Overpeck 1,2,3 , Bette L. Otto-Bliesner 4 1. Department of Geosciences, University of Arizona, Tucson, AZ 85721, USA 2. Institute of the Environment, University of Arizona, 845 North Park Avenue, Suite 532, Tucson, AZ 85719, USA 3. Department of Atmospheric Science, University of Arizona, Tucson, AZ 85721, USA 4. National Center for Atmospheric Research, Post Office Box 3000, Boulder CO 80307 USA *Corresponding author: nmckay@email.arizona.edu 200 Abstract A compilation of paleoceanographic data and a coupled atmosphere-ocean climate model were used to examine global ocean surface temperatures of the Last Interglacial (LIG) period, and to produce the first quantitative estimate of the role that ocean thermal expansion likely played in driving sea level rise above present day during the LIG. Our analysis of the paleoclimatic data suggests a peak LIG global sea surface temperature (SST) warming of 0.7±0.6°C compared to the late Holocene. Our LIG climate model simulation suggests a slight cooling of global average SST relative to preindustrial conditions (ΔSST = -0.4°C), with a reduction in atmospheric water vapor in the Southern Hemisphere driven by a northward shift of the Intertropical Convergence Zone, and substantially reduced seasonality in the Southern Hemisphere. Taken together, the model and paleoceanographic data imply a minimal contribution of ocean thermal expansion to LIG sea level rise above present day. Uncertainty remains, but it seems unlikely that thermosteric sea level rise exceeded 0.4±0.3 m during the LIG. This constraint, along with estimates of the sea level contributions from the Greenland Ice Sheet, glaciers and ice caps, implies that 4.1 to 5.8 m of sea level rise during the Last Interglacial period was derived from the Antarctic Ice Sheet. These results reemphasize the concern that both the Antarctic and Greenland Ice Sheets may be more sensitive to temperature than widely thought. Introduction Sea-level rise is one of the major socio-economic hazards associated with global 201 warming, and a better understanding the mechanisms that underly sea-level rise is a prerequisite to accurate projections of global and regional sea-level rise. Despite this, variability in the different components of sea level rise (i.e., ocean thermal expansion, melting of glaciers, and wasting of the Greenland and Antarctic Ice Sheets) is poorly understood, especially with respect to the future. The Fourth Assessment Report of the Intergovernmental Panel on Climate Change (IPCC), which explicitly excluded rapid ice flow dynamics, projected that ocean thermal expansion would make up 55 to 70% of the sea level rise over the 21st century [Meehl et al., 2007], whereas the empirical model of Vermeer and Rahmstorf [2009] projects a much smaller proportion, between 20 and 30%, although this result is primarily driven by a larger contribution of ice melt. On longer timescales, the equilibrium response of ocean thermal expansion to warming has been estimated as 0.2 to 0.6 m °C-1 [Meehl et al. 2007], but the relative contributions of ice sheet melt and thermal expansion during a millennial-scale highstand remains unclear. One approach to address this uncertainty is to study past sea-level changes. The Last Interglacial period (LIG) is the most recent warm interval with substantially higher-thanmodern global sea level. During the LIG, from ca. 130 to 120 ka, sea level reached at least 6 m above present levels [Hearty et al., 2007; Kopp et al., 2009]. The majority of the sea level rise originated from melting of the Greenland Ice Sheet (GIS) and the Antarctic Ice Sheets [Otto-Bliesner et al., 2006; Overpeck et al., 2006; Kopp et al., 2009; Clark and Huybers, 2009], but the role of thermal expansion has not been carefully examined. Here, we compile the available paleoceanographic records and examine global climate model simulations to better constrain the amount of thermal expansion during the 202 LIG. Methods Paleoclimate data We compiled a dataset of 76 published sea surface temperature (SST) records that met several criteria. Only quantitative SST records that included both the LIG and the Holocene were included so that ΔSST values (warmest LIG – late Holocene) could be calculated internally for each record. We restricted our analyses to records that had an average temporal resolution of 3 kyr or better during both LIG and the Holocene. Records were obtained through the NOAA Paleoclimatology World Data Center (www.ngdc.noaa.gov/paleo/paleo.html), the Pangaea database (www.pangaea.de), and from individual site reports and papers (Supplemental Table 1). The sea surface temperatures (SSTs) were determined using Mg/Ca ratios in foraminifera, alkenone unsaturation ratios (i.e., Uk37), and faunal assemblage transfer functions (for radiolaria, foraminifera, diatoms and coccoliths), and were interpreted to reconstruct annual, austral summer, and boreal summer sea surface temperatures. Only records with published age models were included; however, there is substantial uncertainty between age estimates at different sites. For this study we chose to determine a maximum estimate of ocean warming during a sustained sea level highstand, so the average SST of a 5 kyr period centered on the warmest temperature between 135 and 118 ka was calculated for each record. ΔSST values were determined by subtracting the average SST of the late Holocene (5 to 0 ka) from the 5 kyr LIG average. The data set was supplemented by 94 203 LIG SST estimates from the CLIMAP project [CLIMAP Project Members, 1984]. For the CLIMAP data, ΔSST values were determined as the difference between LIG temperatures and core top temperature estimates at each site. Global mean SST anomalies (ΔSST) were calculated by averaging anomalies in 10°x10° boxes, then determining zonal averages, which were finally averaged after weighting each zonal average by the area of ocean for each latitudinal band. To complement our data synthesis, we performed the same analyses on the LIG SST dataset assembled by Turney and Jones [2010]. The Turney and Jones [2010] dataset differs from our synthesis in several regards: 1) only data that were interpreted to reconstruct annual mean temperatures were included, 2) the timing of LIG mean SST estimates was determined by corresponding marine δ18O records and 3) ΔSST values were calculated as the difference between LIG SSTs and instrumental SST climatology. Many of the same records went into both LIG SST syntheses, but analyzing both datasets allows us to evaluate the sensitivity of our results to markedly different data treatment approaches. Global Climate Model Simulations Climate simulations were conducted using a global, coupled ocean-atmosphereland-sea ice general circulation model (Community Climate System Model [CCSM], Version 3) [Collins et al., 2006]. The atmospheric model has ~1.4° latitude-longitude resolution (T85) with 26 levels, and the ocean model has ~1° resolution and 40 levels. The preindustrial 1870 AD control simulation includes the appropriate forcing conditions, 204 including trace gas concentrations (CO2: 289 ppmv; CH4 901 ppbv), solar constant (1365 W/m2), and orbital characteristics (obliquity: 23.44°, perihelion: 3 January, and eccentricity: 0.0167). The preindustrial control simulation was run for 550 model years, and climatologies were calculated using model years 530 to 549. The LIG simulation included forcing conditions appropriate for 125 ka; obliquity was 23.80°, perihelion was 23 July, and eccentricity was 0.0400 [Berger et al., 1991]. The trace gas concentrations were estimated from ice core data (CO2: 273 ppmv; CH4: 642 ppbv) [Petit et al., 1999]. The solar constant was set to the model present-day value of 1367 W/m2. Vegetation and land ice coverage were prescribed at their present-day distributions for both the preindustrial and LIG simulations. The LIG simulation was run for 200 model years, and climatologies were calculated using model years 180 to 199. CCSM3 is known to have regional SST biases, but is very well-suited for simulating global mean SST [Collins et al., 2006], which is the focus of this study. Global SST anomalies were calculated by zonal averaging and then calculating an area-weighted global mean. Paleoceanographic data synthesis LIG-Holocene SST anomalies varied regionally, and importantly, were not uniformly warmer during the LIG (Figure C1). Furthermore, our data synthesis shows the same primary patterns as the synthesis of Turney and Jones [2010], suggesting that the primary results results are robust to the choices of averaging and Holocene reference period. Records from the high latitudes of the Northern Hemisphere (>30°N) were consistently warmer during the LIG. This is consistent with the dramatic increase in 205 summer insolation (~12% above preindustrial), and extensive evidence for much warmer (4-5 °C) conditions in the Arctic during the interval [CAPE Project Members, 2006]. South of ~30°N, the anomalies are regionally variable (Figure C1). The Caribbean Sea and the tropical Atlantic Oceans appear to have been generally cooler during the LIG than the late Holocene (and late 20th century). The eastern equatorial Pacific Ocean shows both positive and negative anomalies, as does the rest of the Pacific ocean. The western Indian Ocean appears to have been slightly warmer, and the central Indian Ocean somewhat cooler, but the data coverage in both the Pacific and Indian Ocean is poor. The southeastern Atlantic Ocean, off the west coast of South Africa, was consistently warmer. The Southern Ocean changes are mixed, apparently cooler west of South America, somewhat warmer in the Atlantic sector and near New Zealand, and mixed in the Indian sector. The regional variability is interesting, and warrants further investigation, however interpreting the patterns in terms of modes of climatic and ocean variability is confounded by chronological errors, resolution differences and poor data coverage. Consequently, we chose to focus on the primary, global pattern: the warmer temperatures between 30°N and 70°N, and equivocal anomalies further south (Figure C1). The oceanarea-weighted global average SST anomaly is 0.7±0.6°C for our data synthesis, and 0.7°C for that of Turney and Jones [2010]. Interestingly, these global ΔSST estimates are lower than the global land and ocean temperature anomaly (1.5±0.1°C) calculated by Turney and Jones [2010]. This discrepancy may be due to the predominance of terrestrial records from the Northern Hemisphere that are particularly sensitive to summer 206 temperatures in their global synthesis. The error calculated for global ΔSST incorporates errors in the SST proxies, which typically range from 1 to 2°C, and the error associated with estimating global ΔSST from limited spatial coverage (Auxiliary Material). Nevertheless, this estimate does not capture all of uncertainty in global ΔSST. Because we calculated a maximum estimate for ΔSST, we excluded chronological errors, although differences in temporal resolution between sites contributes additional uncertainty. Furthermore, each of the SST proxies comes with its own set of errors and biases. A particular concern is that all three of the primary SST proxies in our database (faunal assemblages, Mg/Ca, Uk37) are known to be sensitive to changes in seasonality [Anand et al., 2003; Morey et al., 2005; Lorenz et al., 2006], and each proxy may exhibit different responses to changes in seasonality, even at the same location [e.g., Weldeab et al., 2007; Saher et al., 2009]. Given the extreme differences in seasonal insolation forcing during the LIG relative to the late Holocene, changes in the timing and distribution of the productive seasons likely biased the SST estimates. To evaluate some of the potential biases in our analysis, we subsampled our database by proxy type and seasonality (Auxiliary Material). Globally, ΔSST for the Uk37 and Mg/Ca proxies was about 1.5°C higher than the faunal assemblage proxies. Some of this offset is likely due to lower sample density and different spatial coverage of the Uk37 and Mg/Ca proxies, which are commonly located near coasts in upwelling regions. Regionally, ΔSST appears generally consistent between proxies, with some exceptions (Figure C1a). Subdividing seasonally, ΔSST in boreal summer (JJA) records was slightly 207 higher (0.2°C) than in austral summer (DJF) records, consistent with the change in insolation forcing. Global Climate Model Simulations Like the paleoceanographic data, the model simulations for 125 ka show substantial warming north of 40°N, and similar or slightly cooler SSTs south of 30°N (Figure C2a-b), similar to previously published simulations for this time period [Montoya et al., 2000; Kaspar and Cubasch, 2007]. The ocean-area-weighted global average ocean temperature difference between the 125 ka simulation and the preindustrial control is -0.4°C for both the surface temperatures and the top 200 m. The result of a cooler average ocean surface in the 125 ka simulation is surprising given that the annual insolation anomalies are positive globally (Figure C2c). This result merits a discussion of the climate dynamics simulated in the model that contribute to the cooling in the Southern Hemisphere. The most significant difference between the forcings for the LIG simulation and the preindustrial control are the different orbital parameters, and among those, the date of perihelion (or phase in the precession cycle) is most different. In the LIG simulation, perihelion occurs during the boreal summer, as opposed to the preindustrial control, when aphelion occurs during the boreal summer. The result is that relative to the preindustrial simulation, the Northern Hemisphere should experience much greater seasonality (warmer summers and colder winters), while the Southern Hemisphere should have colder summers and warmer winters (Figure C2c). 208 The Southern Hemisphere cooling in the model is associated with a decrease in longwave radiative forcing (Figure C2d), which is a function of decreased water vapor concentrations in the southern hemisphere (Figure C2e). Annually averaged, water vapor content was consistently lower in the LIG simulation than the preindustrial control throughout most of the Southern Hemisphere and over the Pacific Ocean, and substantially higher over the Northern Hemisphere monsoon regions, and the high Northern latitudes (Figure C2e). There appear to be two global scale mechanisms responsible for the hemispheric shift in water vapor. First, the large increase in summer insolation in the Northern Hemisphere results in a strengthening of the Asian, African and North American Monsoons in the model, along with a northward shift of the Intertropical Convergence Zone (ITCZ) (Figure C2e). The strengthened and northward-shifted monsoon systems pull more moisture further across the equator into the Northern Hemisphere, focusing precipitation in the monsoons while effectively drying the southern tropics. The effect of this northward shift on the Earth's energy budget is apparent in the changes in outgoing longwave radiation (OLR; Figure C2f), which is substantially reduced over the Northern Hemisphere monsoon regions, and increased over the Southern Hemisphere monsoon systems (e.g., South America, equatorial and southern Africa, Australia), effectively cooling the tropical Southern Hemisphere. The second mechanism is associated with the opposing changes in seasonality in each hemisphere. Due to the nonlinearity in the capacity of air to hold water vapor as a function of temperature (the Clausius-Clapeyron relation), the large decrease in insolation 209 during the Southern Hemisphere summer and fall is not compensated, in terms of specific humidity, by an equivalent increase in winter and spring insolation. This effect should be most important at higher latitudes, and the increase in Northern Hemisphere specific humidity is consistent with this hypothesis (Figure C2e). The impact is not immediately apparent in OLR (Figure C2f). At the high southern latitudes, both downwelling radiation at the surface (Figure C2d) and OLR (Figure C2f) are decreased. This is due to decreased absorption and attenuation of longwave radiation in the atmosphere, and is a function of both decreased specific humidity and cooler surface temperatures decreasing the amount of outgoing longwave radiation produced at the surface. The opposite scenario is apparent at the high northern latitudes. These two mechanisms provide a plausible explanation for the cooling over most of the world's ocean. The results do not appear to be specific to the CCSM3 model. Simulations with other climate models show cooler temperatures in the Southern Hemisphere, and near-zero or negative annual SST anomalies relative to preindustrial controls [e.g., Montoya et al., 2000; Kaspar and Cubasch, 2007]. Furthermore, an additional simulation using CCSM3 for the period 130 ka yields a similar cooling in the Southern Hemisphere, despite regional differences in SST, suggesting that our result is not specific to only this interval of the LIG (Auxiliary Material). This result calls into question the belief that the LIG was substantially warmer globally [e.g., LIGA Members, 1991; Clark and Huybers, 2009; Turney and Jones, 2010; Masson-Delmotte et al., 2010]. Much uncertainty remains in model simulations; but it is possible that the predominance of terrestrial, Northern Hemisphere, summer-sensitive temperature proxies may have 210 biased our understanding of global temperature anomalies during the interval. The thermosteric component of LIG sea level rise The amount of steric sea level rise can be determined by calculating the specific volume of the ocean, which requires integrating the temperature and salinity structure of the ocean. This is possible for the model simulations, but not for the paleoceanographic data, so other approaches must be utilized. A simple empirical approach is to estimate a thermal expansion sensitivity (i.e., cm/°C). This can be achieved with instrumental data; the IPCC [Bindoff et al., 2007] concluded that the top 700 m of the ocean warmed 0.1°C from 1961-2003, and that thermal expansion of the ocean was about 1.3 cm over the same interval, resulting in a sensitivity of ~13 cm/°C. To determine a maximum estimate, we assumed our average ΔSST of 0.7±0.6°C is representative of the top 700 m, resulting in 9±8 cm of thermosteric sea level rise. Alternatively, we estimate the thermal expansion using the Thermodynamic Equation of Seawater 2010 (TEOS-10) to calculate the change in the specific volume of the top 700 m of the ocean due to a 0.7±0.6°C warming, while holding the salinity constant, and neglecting changes in ocean area. This approach results in ~12±10 cm of thermosteric sea level rise. It is possible that sustained, warmer-than-modern conditions resulted in warming below 700 m in the oceans. If the average warming extended to 2000 m, the thermal expansion of the ocean would have been about 35±30 cm, consistent with the equilibrium ocean-temperature thermal expansion sensitivity observed in long climate model simulations (0.2 to 0.6 m °C-1) [Meehl et al. 2007] . 211 For the model simulation, the whole-ocean global average steric sea level change was -18 cm, primarily due to cooler ocean temperatures in the Southern Hemisphere. Because the model simulations are relatively short, the deep ocean was not equilibrated. This introduces additional uncertainty in our estimate of steric sea level change; however the volume-integrated ocean temperature trends are the same in both the LIG and preindustrial simulations (-0.12°C/century), suggesting that the steric sea level change would be comparable after equilibration. Altogether, it is clear that ocean thermal expansion during the LIG was a small component of the maximum LIG sea level highstand. A conservative estimate from the available paleoclimatic data is 0.4±0.3 m. The climate model simulations suggest that the thermosteric component may have been smaller or even negative 125 ka, near the time of the maximum highstand. This has several important implications. First the high-end estimate of sea level exceedance (33% probability that sea level exceeded 9.4 m during the LIG) by Kopp et al. [2009] is probably too high, because the stochastic thermosteric component in their model was unrealistically large (mean = 0 m, 1σ = 2 m). Using a more realistic thermosteric component should reduce the variance of the distribution of sea level histories, resulting in tighter error estimates and exceedance levels that are nearer to the median. Secondly, our results provide further constraints on the relative contributions to sea level rise during the Last Interglacial. The contribution from the GIS was likely 2.2 – 3.4 m [Otto-Bliesner et al., 2006], or even less [Oerlemans et al., 2006]. The maximum possible contribution from mountain glaciers and ice caps is 0.6±0.1 m [Radić and Hock, 212 2010], and our conservative estimate of maximum thermal expansion during the LIG (0.4±0.3 m). These data, combined with the median projection (50% exceedance) of maximum LIG sea level rise (8.5 m) [Kopp et al., 2009] imply that the Antarctic Ice Sheet (AIS), most likely the West Antarctic Ice Sheet (WAIS), contributed at least 4.1±0.3 m. Assuming a low-end contribution from the GIS (2.2 m), only glaciers and ice caps from the northern Hemisphere (0.4±0.1 m) and our low-end estimate for thermal expansion (0.1±0.1 m), the maximum contribution from Antarctica is 5.8±0.1 m. It remains unclear why so much more ice (4.1 to 5.8 m sea level equivalent) was lost from Antarctica during the LIG than the Holocene. Antarctic ice cores all suggest warmer-than-modern annual temperatures for East Antarctica [cf., Petit et al., 1999; EPICA, 2004; Kawamura et al., 2007], and recent evidence suggests that the warming anomaly may have been larger (~6°C warmer than the Holocene) than previously estimated [Sime et al., 2009]. This stands in contrast to the cooling simulated by our LIG simulation (Figure C2a-b), and is a consistent frustration of model-paleodata comparisons [Masson-Delmotte et al., 2010]. Furthermore, melt season solar insolation was substantially lower than present-day (Figure C2c). A recent study by Huybers and Denton [2008] suggested that Antarctic temperatures are primarily controlled by the duration of summer, which was very long during this interval, rather than the intensity of solar insolation (like the Northern Hemisphere), although this mechanism does not drive Antarctic temperatures in our LIG simulation. It has also been hypothesized that poorly simulated climatic feedbacks and changes in ocean circulation may be responsible for the mismatch [Overpeck et al., 2006; Masson-Delmotte et al., 2010], a hypothesis that 213 implies substantial vulnerability of the AIS in the future [Yin et al., 2011]. Finally, it is possible that much of the Antarctic contribution was derived during late deglaciation or early LIG, when melt-season insolation was much higher [Overpeck et al., 2006]. This possibility is consistent with the observation that substantial downwasting of the WAIS is necessary to simulate the high temperatures inferred from East Antarctic ice cores during the LIG in climate models [Holden et al., 2010]. Conclusion The available paleoceanographic records and our LIG GCM simulation suggest that global SSTs were not dramatically warmer than preindustrial conditions (paleodata = 0.7±0.6°C warmer, model = 0.4°C cooler). Taken together, the model and paleodata imply a minimal (-0.2 to 0.4 m) contribution of thermal expansion to LIG sea level rise. This constraint, along with estimates of the sea level contributions from the Greenland Ice Sheet, glaciers and ice caps, implies that 4.1 to 5.8 m of sea level rise during the LIG was derived from the Antarctic Ice Sheet. These results reemphasize the concern that the Greenland and especially the Antarctic Ice Sheets may be more sensitive to temperature than widely thought. Acknowledgements We thank Suz Tolwinski-Ward, Joellen Russell, John Fasullo, Toby Ault, and Sarah Truebe for insightful discussions on the work presented in this paper. Nan Rosenbloom provided the numbers for the model forcings. NCAR is funded by NSF. We also thank 214 Mark Siddall and Michael Oppenheimer for helpful comments in review of this paper. References Anand, P., H. Elderfield, and M. H. Conte (2003), Calibration of Mg/Ca thermometry in planktonic foraminifera from a sediment trap time series, Paleoceanography, 18, doi:200310.1029/2002PA000846. Berger, A., and M. F. Loutre (1991), Insolation values for the climate of the last 10 million years, Quaternary Science Reviews, 10, 297–317. Bindoff, N. et al. (2007), Observations: oceanic climate change and sea level. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, edited by S. Solomon and others, Cambridge University Press, Cambridge. CAPE Project Members (2006), Last interglacial Arctic warmth confirms polar amplification of climate change, Quaternary Science Reviews, 25, 1383–1400. Clark, P. U., and P. Huybers (2009), Interglacial and future sea level, Nature, 462, 856– 857. CLIMAP Project Members (1984), The Last Interglacial ocean, Quat. Res., 21, 123–224. Collins, W. D. et al. (2006), The Community Climate System Model Version 3 (CCSM3), J. Climate, 19, 2122-2143, doi:10.1175/JCLI3761.1. EPICA (2004), Eight glacial cycles from an Antarctic ice core, Nature, 429, 623–628. Hearty, P. J., J. T. Hollin, A. Neumann, M. O’Leary, and M. McCulloch (2007), Global sea-level fluctuations during the Last Interglaciation (MIS 5e), Quat. Sci. Rev., 26, 2090-2112. Holden, P. B. et al. (2010), Interhemispheric coupling, the West Antarctic Ice Sheet and warm Antarctic interglacials, Clim. Past, 6, 431-443. Huybers, P., and G. Denton (2008), Antarctic temperature at orbital timescales controlled by local summer duration, Nature Geoscience, 1, 787–792. Kaspar, F., and U. Cubasch (2007), Simulations of the Eemian interglacial and the subsequent glacial inception with a coupled ocean-atmosphere general circulation 215 model, in The Climate of Past Interglacials, edited by F. Sirocko, M. Claussen, M.F. Sánchez Goñi, and T. Litt, pp. 499-515, Elsevier. Kawamura, K. et al. (2007), Northern Hemisphere forcing of climatic cycles in Antarctica over the past 360,000 years, Nature, 448, 912–916. Kopp, R., F. J. Simons, J. X. Mitrovica, A. Maloof, and M. Oppenheimer (2009), Probabilistic assessment of sea level during the Last Interglacial stage, Nature, 462, 863–867. LIGA members (1991), Report of 1st discussion group: The Last Interglacial in high latitudes of the Northern Hemisphere: Terrestrial and marine evidence, Quat. Int., 10-12, 9-28. Lorenz, S. J., J. Kim, N. Rimbu, R. R. Schneider, and G. Lohmann (2006), Orbitally driven insolation forcing on Holocene climate trends: Evidence from alkenone data and climate modeling, Paleoceanography, 21, doi:200610.1029/2005PA001152. Masson-Delmotte, V. et al. (2010), EPICA Dome C record of glacial and interglacial intensities, Quaternary Science Reviews, 29, 113–128. Meehl, G. A. et al. (2007), Chapter 10: Global Climate Projections. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, edited by S. Solomon and others, Cambridge University Press, Cambridge. Montoya, M., H. von Storch, and T. J. Crowley (2000), Climate simulation for 125 kyr BP with a coupled ocean-atmosphere general circulation model, Journal of Climate, 13, 1057–1072. Morey, A. E., A. C. Mix, and N. G. Pisias (2005), Planktonic foraminiferal assemblages preserved in surface sediments correspond to multiple environment variables, Quaternary Science Reviews, 24, 925-950, doi:10.1016/j.quascirev.2003.09.011. Oerlemans, J., D. Dahl-Jensen, and V. Masson-Delmotte (2006), Ice Sheets and Sea Level, Science, 313, 1043 -1045, doi:10.1126/science.313.5790.1043c. Otto-Bliesner, B. L., S. Marshall, J. T. Overpeck, G. Miller, and A. Hu (2006), Simulating Arctic climate warmth and icefield retreat in the last interglaciation, Science, 311, 1751-1753. Overpeck, J. T., B. L. Otto-Bliesner, G. H. Miller, D. R. Muhs, R. B. Alley, and J. T. Kiehl (2006), Paleoclimatic evidence for future ice-sheet instability and rapid sea- 216 level rise, Science, 311, 1747-1750. Petit, J. R. et al. (1999), Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica, Nature, 399, 429–436. Radić, V., and R. Hock (2010), Regional and global volumes of glaciers derived from statistical upscaling of glacier inventory data, Journal of Geophysical Research, 115, doi:201010.1029/2009JF001373. Saher, M. H. et al. (2009), Western Arabian Sea SST during the penultimate interglacial: A comparison of U37K′ and Mg/Ca paleothermometry, Paleoceanography, 24, doi:10.1029/2007PA001557. Sime, L. C., E. W. Wolff, K. I. C. Oliver, and J. C. Tindall (2009), Evidence for warmer interglacials in East Antarctic ice cores, Nature, 462, 342–345. Turney, C. S., and R. T. Jones (2010), Does the Agulhas Current amplify global temperatures during super-interglacials?, Journal of Quaternary Science, 25, 839– 843. Vermeer, M., and S. Rahmstorf (2009), Global sea level linked to global temperature, Proceedings of the National Academy of Sciences, 106, 21527-21532. Weldeab, S., R. R. Schneider, and P. Müller (2007), Comparison of Mg/Ca-and alkenonebased sea surface temperature estimates in the fresh water–influenced Gulf of Guinea, eastern equatorial Atlantic, Geochemistry Geophysics Geosystems, 8, Q05P22, 10.1029/2006GC001360. Yin, J., J. T. Overpeck, S. T. Griffies, A. Hu, J. L. Russell, and R.J. Stouffer (2011), Different magnitudes of projected subsurface ocean warming around Greenland and Antarctica, Nature Geoscience, in press. 217 Auxiliary Material Error Analysis To quantify the error in our paleoceanographic global mean SST anomalies, we propagated the SST errors for each SST measurement though each of the averaging steps (temporal → grid cell → zonal → area-weighted global) in our ocean-area-weighted average. We used quoted error estimates for each study, except for those that only cited analytical errors, and not SST-calibration error. For these studies, and those which did not provide error estimates, we applied proxy-specific error estimates (UK37 = 1.5°C, Mg/Ca = 1.1°C, Faunal = 1.5°C), following Turney and Jones [2010] and Barrows et al. [2006]. To estimate the error associated with the limited spatial range of the oceanographic proxies, we calculated 1000 random 1-year global SST anomalies from the ERSST dataset over the period 1910 to 2009 AD (e.g., one iteration could be 1954 – 2003, another 1981 – 1939, and so on). We then calculate the global SST anomalies using all of the data, and a second estimate using just the grid cells where we have paleoceanographic data. We consider the difference between these estimates to be the spatial coverage error, and derive the statistics of the error from the distribution of the 1000 calculated distributions, and then incorporate this into our error analysis (Figure CS1). This is obviously not a perfect analog to our millennial scale analysis, but does diminish the effect of memory and climatic modes from annual estimates, and can have a similar magnitude as our global anomaly (0.7°C). We also evaluated the effect of grid-size on both sources of error, and the results are presented in Aux. Table 2. Calculated error decreases with larger grid cells, 218 presumably due to the fact the we do not incorporate the error associated with scaling up site-specific SST anomalies to grid-cell mean anomalies. Consistent with our conservative approach, the 10° by 10° we focus on in this study are on the high end of the range of errors. Bias Analysis To evaluate some of the potential biases in our database, we subsampled our database by proxy type and seasonality. The ocean-area-weighted global ΔSST for the Uk37 and Mg/Ca proxies are higher (1.6°C, n=28, and 1.4°C, n=10, respectively) than the average for faunal assemblage proxies (0.1°C, n=124). Some of the difference may be due to the different spatial coverage of the proxies; there are far fewer Uk37 and Mg/Ca records, and they tend to be located near coasts. Regionally, ΔSST appears generally consistent between proxies; however, in regions that appear to have been cooler during the LIG, the few available Mg/Ca- and Uk37-based reconstructions often suggest warmerthan-modern SSTs (Figure C1a). Subdividing seasonally, records interpreted to record annual SSTs had an oceanarea-weighted global average difference of 1.3°C (n=42). Records interpreted to record boreal summertime (JJA) SSTs had an average difference of 0.2°C (n=57), and 0.0°C (n=63) for boreal winter (DJF) temperature estimates. The higher ΔSST averages in annual records primarily reflects differences between proxies, since most Mg/Ca and Uk37 derived records are interpreted to reflect annual SST, whereas most of the JJA- and DJFsensitive records are derived from faunal assemblages. The lower ΔSST estimate for the 219 boreal winter is consistent with negative insolation anomalies during that interval. Finally, the ΔSST estimates from the CLIMAP project [CLIMAP Project Members, 1984], were slightly lower (0.1°C, n=94) than the other faunal assemblage estimates in the database (0.3°C, n=30). This small offset is may be due to the selection of the warmest 5 kyr interval for the LIG reference for the non-CLIMAP data. Finally, the impact of latitude-longitude grid size on global ΔSST was evaluated by calculating ΔSST for latitude and longitude grid sizes ranging from 5° to 60° (Auxiliary Table 2). Increasing longitude grid sizes had a minimal effect on ΔSST, but increasing latitude grid sizes resulted in lower global ΔSST estimates. To sample the dataset at higher resolution, and to determine a conservatively high estimate for ΔSST, we chose 10° by 10° boxes for the analyses in this study. 130 ka climate model simulation In addition to our model simulation for 125 ka, we have a second simulation with orbital parameters and greenhouse gas concentrations consistent with 130 ka (CO2: 300 ppm, CH4: 720 ppb). Comparing this simulation to our simulation for 125 ka allows us to evaluate whether our conclusions and the data-model mismatch are only specific to 125 ka, or are also characteristic of earlier parts the LIG. Figure C-S2 shows the difference in surface temperatures between the two simulations. There are regional differences between the two simulations, but the global mean temperatures are within 0.1°C, suggesting that our conclusions are not limited to the interval 125 ka. 220 Auxiliary References de Abreu, L., N. J. Shackleton, J. Schönfeld, M. Hall, and M. Chapman (2003), Millennial-scale oceanic climate variability off the Western Iberian margin during the last two glacial periods, Marine Geology, 196(1-2), 1–20. Bard, E., F. Rostek, and C. Sonzogni (1997), Interhemispheric synchrony of the last deglaciation inferred from alkenone palaeothermometry, Nature, 385(6618), 707– 710. Barrows, T. T., S. Juggins, P. De Decker , E. Calvo, and C. Pelejero (2007). Long-term sea surface temperature and climate change in the Australian–New Zealand region, Paleoceanography 22 doi:10.1029/2006PA001328. Becquey, S., and R. Gersonde (2003), A 0.55-Ma paleotemperature record from the Subantarctic zone: Implications for Antarctic Circumpolar Current development, Paleoceanography, 18(1), 1014. Budziak, D. (2001), Late Quaternary monsoonal climate and related variations in paleoproductivity and alkenone-derived sea-surface temperatures in the western Arabian Sea, Universität Bremen. Clemens, S. C., W. L. Prell, Y. Sun, Z. Liu, and G. Chen (2008), Southern Hemisphere forcing of Pliocene δ18O and the evolution of Indo-Asian monsoons, Paleoceanography, 23(4). CLIMAP Project Members (1984), The Last Interglacial ocean, Quaternary Research, 21, 123–224. Cortese, G., and A. Abelmann (2002), Radiolarian-based paleotemperatures during the last 160 kyr at ODP Site 1089 (Southern Ocean, Atlantic Sector), Palaeogeography, Palaeoclimatology, Palaeoecology, 182(3-4), 259–286. Cortese, G., A. Abelmann, and R. Gersonde (2007), The last five glacial-interglacial transitions: A high-resolution 450,000-year record from the subantarctic Atlantic, Paleoceanography, 22(4). Crosta, X., A. Sturm, L. Armand, and J. J. Pichon (2004), Late Quaternary sea ice history in the Indian sector of the Southern Ocean as recorded by diatom assemblages, Marine Micropaleontology, 50(3-4), 209–223. Crundwell, M., G. Scott, T. Naish, and L. Carter (2008), Glacial-interglacial ocean climate variability from planktonic foraminifera during the Mid-Pleistocene transition in the temperate Southwest Pacific, ODP Site 1123, Palaeogeography, Palaeoclimatology, Palaeoecology, 260(1-2), 202–229. Herbert, T. D., and J. D. Schuffert (2000), 16. Alkenone unsaturation estimates of seasurface temperatures at site 1002 over a full glacial cycle, in Proceedings of the Ocean Drilling Program, Scientific Results, vol. 165, pp. 239–247. Herbert, T. D., J. D. Schuffert, D. Andreasen, L. Heusser, M. Lyle, A. Mix, A. C. Ravelo, 221 L. D. Stott, and J. C. Herguera (2001), Collapse of the California Current during glacial maxima linked to climate change on land, Science, 293(5527), 71. Imbrie, J. et al. (1992), On the structure and origin of major glaciation cycles 1. Linear responses to Milankovitch forcing, Paleoceanography, 7(6), 701–738. Jasper, J. P., J. M. Hayes, A. C. Mix, and F. G. Prahl (1994), Photosynthetic fractionation of 13C and concentrations of dissolved CO2 in the central equatorial Pacific during the last 255,000 years, Paleoceanography, 9(6), 781–798. Kandiano, E. S., and H. A. Bauch (2003), Surface ocean temperatures in the north-east Atlantic during the last 500 000 years: evidence from foraminiferal census data, Terra Nova, 15(4), 265-271, doi:10.1046/j.1365-3121.2003.00488. Kiefer, T., and M. Sarnthein (1998), Distribution of planktic foraminifers and SST of sediment core GIK15637-1, , doi:10.1594/PANGAEA.202139. Kirst, G. J., R. R. Schneider, P. J. Müller, I. von Storch, and G. Wefer (1999), Late Quaternary temperature variability in the Benguela Current system derived from alkenones, Quaternary Research, 52(1), 92–103. Lawrence, K. T., and T. D. Herbert (2005), Late Quaternary sea-surface temperatures in the western Coral Sea: Implications for the growth of the Australian Great Barrier Reef, Geology, 33(8), 677. Lea, D. W., D. K. Pak, C. L. Belanger, H. J. Spero, M. A. Hall, and N. J. Shackleton (2006), Paleoclimate history of Galapagos surface waters over the last 135,000 yr, Quaternary Science Reviews, 25(11-12), 1152–1167. Lea, D. W., D. K. Pak, and H. J. Spero (2000), Climate impact of late Quaternary equatorial Pacific sea surface temperature variations, Science, 289(5485), 1719. Martinson, D. G., N. G. Pisias, J. D. Hays, J. Imbrie, T. C. Moore, and N. J. Shackleton (1987), Age dating and the orbital theory of the ice ages: Development of a highresolution 0 to 300,000-year chronostratigraphy, Quaternary Research, 27(1), 129, doi:10.1016/0033-5894. Martrat, B., J. O. Grimalt, C. Lopez-Martinez, I. Cacho, F. J. Sierro, J. A. Flores, R. Zahn, M. Canals, J. H. Curtis, and D. A. Hodell (2004), Abrupt temperature changes in the western Mediterranean over the past 250,000 years, Science, 306(5702), 1762. Martrat, B., J. O. Grimalt, N. J. Shackleton, L. de Abreu, M. A. Hutterli, and T. F. Stocker (2007), Four climate cycles of recurring deep and surface water destabilizations on the Iberian margin, Science, 317(5837), 502. Mashiotta, T. A., D. W. Lea, and H. J. Spero (1999), Glacial-interglacial changes in Subantarctic sea surface temperature and δ18O-water using foraminiferal Mg, Earth and Planetary Science Letters, 170(4), 417–432. Müller, P. J., R. Schneider, and G. Ruhland (1994), Late Quaternary pCO2 variations in the Angola Current: Evidence from organic carbon d13C and alkenone 222 temperatures, Carbon Cycling in the Glacial Ocean: Constraints on the Ocean’s Role in Global Change, 343–366. Nürnberg, D., A. Müller, and R. R. Schneider (2000), Paleo-sea surface temperature calculations in the equatorial east Atlantic from Mg/Ca ratios in planktic foraminifera: A comparison to sea surface temperature estimates from U37K′, oxygen isotopes, and foraminiferal transfer function, Paleoceanography, 15(1), 124–134. Nürnberg, D., M. Ziegler, C. Karas, R. Tiedemann, and M. W. Schmidt (2008), Interacting Loop Current variability and Mississippi River discharge over the past 400 kyr, Earth and Planetary Science Letters, 272(1-2), 278–289. Pahnke, K., and J. P. Sachs (2006), Sea surface temperatures of southern midlatitudes 0– 160 kyr BP, Paleoceanography, 21(2). Pahnke, K., R. Zahn, H. Elderfield, and M. Schulz (2003), 340,000-year centennial-scale marine record of Southern Hemisphere climatic oscillation, Science, 301(5635), 948. Pailler, D., and E. Bard (2002), High frequency palaeoceanographic changes during the past 140 000 yr recorded by the organic matter in sediments of the Iberian Margin, Palaeogeography, Palaeoclimatology, Palaeoecology, 181(4), 431–452. Pichon, J. J., L. D. Labeyrie, G. Bareille, M. Labracherie, J. Duprat, and J. Jouzel (1992), Surface water temperature changes in the high latitudes of the Southern Hemisphere over the last glacial-interglacial cycle, Paleoceanography, 7(3), 289– 318. Pisias, N. G., and A. C. Mix (1997), Spatial and temporal oceanographic variability of the eastern equatorial Pacific during the Late Pleistocene: Evidence from radiolaria microfossils, Paleoceanography, 12(3), PP. 381-393, doi:199710.1029/97PA00583. Rickaby, R. E. M., and H. Elderfield (1999), Planktonic foraminiferal Cd/Ca: Paleonutrients or paleotemperature?, Paleoceanography, 14(3), 293–303. Ruddiman, W. F., M. E. Raymo, D. G. Martinson, B. M. Clement, and J. Backman (1989), Pleistocene evolution: Northern hemisphere ice sheets and North Atlantic Ocean, Paleoceanography, 4(4), 353–412. Schaefer, G., J. S. Rodger, B. W. Hayward, J. P. Kennett, A. T. Sabaa, and G. H. Scott (2005), Planktic foraminiferal and sea surface temperature record during the last 1 Myr across the Subtropical Front, Southwest Pacific, Marine Micropaleontology, 54(3-4), 191–212. Schmidt, M. W., H. J. Spero, and D. W. Lea (2004), Links between salinity variation in the Caribbean and North Atlantic thermohaline circulation, Nature, 428(6979), 160–163. Schneider, R. R., P. J. Müller, and G. Ruhland (1995), Late Quaternary surface circulation 223 in the east equatorial South Atlantic: Evidence from alkenone sea surface temperatures, Paleoceanography, 10(2), 197–219. Sowers, T., M. Bender, L. Labeyrie, D. Martinson, J. Jouzel, D. Raynaud, J. J. Pichon, and Y. S. Korotkevich (1993), A 135,000-year Vostok-Specmap Common temporal framework, Paleoceanography, 8(6), 737, doi:10.1029/93PA02328. Stein, R., J. Hefter, J. Grützner, A. Voelker, and B. D. Naafs (2009), Variability of surface water characteristics and Heinrich-like events in the Pleistocene midlatitude North Atlantic Ocean: Biomarker and XRD records from IODP Site U1313 (MIS 16–9), Paleoceanography, 24(2). Weldeab, S., D. W. Lea, R. R. Schneider, and N. Andersen (2007), 155,000 years of West African monsoon and ocean thermal evolution, Science, 316(5829), 1303. Wolff, T., B. Grieger, W. Hale, A. Dürkoop, S. Mulitza, J. Pätzold, and G. Wefer (1999), On the reconstruction of paleosalinities, Use of Proxies in Paleoceanography: Examples From the South Atlantic, 315–344. Yamamoto, M., R. Suemune, and T. Oba (2005), Equatorward shift of the subarctic boundary in the northwestern Pacific during the last deglaciation, Geophysical Research Letters, 32(5), L05609. Yamamoto, M., M. Yamamuro, and Y. Tanaka (2007), The California current system during the last 136,000 years: response of the North Pacific High to precessional forcing, Quaternary Science Reviews, 26(3-4), 405–414. 224 Faunal Uk37 Mg/Ca Cd/Ca Ann DJF JJA Latitude A -5 -5 0 5 0 ΔSST (°C) 5 Δ (Warmest 5 kyr of LIG - Late Holocene) SST (°C) All proxies Latitude B -5 -9 0 Δ (LIG - modern) SST (°C) 9 0 5 ΔSST (°C) Figure C1. Maps of global ΔSST values in A) our database, where symbols indicate proxy type (see legend) and B) the synthesis of Turney and Jones [2010]. Note that in both maps, the locations of the symbols were adjusted slightly for visibility. To the right each map, ΔSST values are plotted by latitude. For our database (A), records interpreted to reflect annual, austral summer, and boreal summer temperatures are shown with different symbols. 225 Δ (LIG - Preindustrial) A B -3 -5 -4 -3 -2 -1 -.5 -.25.25 .5 1 2 3 4 5 Surface temperature (°C) 0 3 -5 -4 -3 -2 -1 -.5 -.25.25 .5 1 2 3 4 5 200 m potential ocean temperature (°C) -2 0 2 -10 0 10 C D 60N 30N 0 30S 60S J F M A M J J A S O N D J -5 0 5 -20 -15 -10 -6 -4 -2 -1 1 2 4 6 10 15 20 -30-20 -15 -10 -5 -3 -1 0 1 3 5 10 15 20 30 -2 Incoming solar radiation (W m ) -2 Downwelling longwave radiation at the surface (W m ) E F -30 -20 -15 -10 -6 -4 -2 2 4 6 10 15 20 30 -1 -2 Specific humidity (g kg x 10 ) -1 0 1 -15 -10 -7 -4 -2 -1 -.25 .25 1 2 4 7 10 15 -2 Outgoing longwave radiation (W m ) -10 0 Figure C2. LIG simulation - preindustrial control anomalies in our global climate model simulation parameters. The parameters include: A) surface air temperature, B) potential temperature averaged over the top 200 m of the ocean, C) incoming solar radiation, by latitude and month, D) downwelling longwave radiation at the surface, E) specific humidity, averaged over all layers of the atmosphere and F) outgoing longwave radiation at the top of the model. Zonal average anomalies are plotted to the right of each map. 10 226 Aux. Table 1. Site names, locations and reconstruction details for paleoceanographic synthesis. Core Site K-11 K-11 V28-56 V28-56 V27-86 V27-86 V28-14 V28-14 V27-20 V27-20 GIK23414-9 GIK23414-9 GIK23414-9 GIK23414-9 V23-82 V23-82 K708-1 K708-1 GIK15612-2 GIK15612-2 V29-179 V29-179 ODP1020 IODP-U1313B V30-97 V30-97 MD95-2040 MD95-2040 ODP607 ODP607 MD012443 MD95-2042 V21-146 V21-146 V32-128 V32-128 MD01-2421 V32-126 V32-126 ODP977A Lat 71.8 71.8 68.0 68.0 66.6 66.6 64.8 64.8 54.0 54.0 53.5 53.5 53.5 53.5 52.6 52.6 50.0 50.0 44.4 44.4 44.0 44.0 41.0 41.0 41.0 41.0 40.6 40.6 40.0 40.0 37.9 37.8 37.7 37.7 36.5 36.5 36.0 35.3 35.3 35.0 Long 1.6 1.6 -6.1 -6.1 1.1 1.1 -29.6 -29.6 -46.2 -46.2 -20.3 -20.3 -20.3 -20.3 -21.9 -21.9 -23.7 -23.7 -26.5 -26.5 -24.5 -24.5 -126.4 -33.0 -32.9 -32.9 -9.9 -9.9 -33.0 -33.0 -10.2 -10.2 163.0 163.0 117.2 117.2 141.8 117.9 117.9 -2.0 Proxy* Faunal Faunal Faunal Faunal Faunal Faunal Faunal Faunal Faunal Faunal Faunal (M) Faunal (M) Faunal (T) Faunal (T) Faunal Faunal Faunal Faunal Faunal (F) Faunal (F) Faunal Faunal Uk37 Uk37 Faunal (F) Faunal (F) Faunal (F) Uk37 Faunal (F) Faunal (F) Uk37 Uk37 Faunal Faunal Faunal Faunal Uk37 Faunal Faunal Uk37 Season ΔSST** Reference JJA 1.1 CLIMAP Project Members [1984] DJF 1.5 CLIMAP Project Members [1984] JJA -0.7 CLIMAP Project Members [1984] DJF -0.9 CLIMAP Project Members [1984] JJA -1.2 CLIMAP Project Members [1984] DJF -1.3 CLIMAP Project Members [1984] JJA 0.4 CLIMAP Project Members [1984] DJF -0.1 CLIMAP Project Members [1984] JJA -0.2 CLIMAP Project Members [1984] DJF 1.7 CLIMAP Project Members [1984] DJF 0.4 Kandiano and Bauch [2003] JJA 0.9 Kandiano and Bauch [2003] DJF 0.8 Kandiano and Bauch [2003] JJA 1.0 Kandiano and Bauch [2003] JJA 1.6 CLIMAP Project Members [1984] DJF 0.3 CLIMAP Project Members [1984] JJA 1.6 Imbrie et al. [1992] DJF 0.5 Imbrie et al. [1992] JJA 1.8 Kiefer and Sarnthein [1998] DJF 2.0 Kiefer and Sarnthein [1998] JJA 2.6 CLIMAP Project Members [1984] DJF 2.8 CLIMAP Project Members [1984] Annual 2.9 Herbert et al. [2001] Annual 0.4 Stein et al. [2009] JJA 3.4 Ruddiman et al. [1989] DJF 1.0 Ruddiman et al. [1989] JJA -2.5 de Abreu et al. [2003] Annual 2.4 Pailler and Bard [2002] DJF 0.2 Ruddiman et al. [1989] JJA 2.5 Ruddiman et al. [1989] Annual 2.8 Martrat et al. [2004] Annual 1.5 Pailler and Bard [2002] JJA 2.8 CLIMAP Project Members [1984] DJF 3.8 CLIMAP Project Members [1984] JJA 1.0 CLIMAP Project Members [1984] DJF -0.5 CLIMAP Project Members [1984] Annual 3.0 Yamamoto et al. [2005] JJA 1.8 CLIMAP Project Members [1984] DJF 1.2 CLIMAP Project Members [1984] Annual 3.3 Martrat et al. [2007] 227 Aux Table 1. (Continued) Core Site ODP1016C RC8-145 RC8-145 ODP1014A ODP1012 MD02-2575 LPAZ21P TR126-29 TR126-29 TR126-23 TR126-23 ODP1146 v12-122 v12-122 v34-88 v34-88 GeoB3007-1 TY93-929/P ODP999A V28-127 V28-127 ODP1002C RC12-339 RC12-339 V29-29 V29-29 GeoB1523-1 GeoB1523-1 MD85674 MD03-2707 TR163-19 V25-59 V25-59 V28-238 V28-238 W8402A-14 TR163-22 ODP806B A180-73 A180-73 Lat 34.5 33.6 33.6 32.8 32.3 29.0 23.0 21.4 21.4 20.5 20.5 19.5 17.0 17.0 16.5 16.5 16.2 13.7 12.8 11.7 11.7 10.7 9.1 9.1 5.1 5.1 3.8 3.8 3.2 2.5 2.3 1.4 1.4 1.0 1.0 1.0 0.5 0.3 0.2 0.2 Long -122.3 -62.4 -62.4 -118.9 -118.4 -87.1 -109.5 -94.0 -94.0 -95.6 -95.6 116.3 -74.4 -74.4 59.5 59.5 59.8 53.3 -78.7 -80.1 -80.1 -65.2 90.0 90.0 77.6 77.6 -41.6 -41.6 50.4 9.4 -91.0 -33.5 -33.5 160.5 160.5 -139.0 -92.4 159.4 -23.0 -23.0 Proxy* Uk37 Faunal Faunal Uk37 Uk37 Mg/Ca Uk37 Faunal Faunal Faunal Faunal Uk37 Faunal Faunal Faunal Faunal Uk37 Uk37 Mg/Ca Faunal Faunal Uk37 Faunal Faunal Faunal Faunal Faunal (F) Faunal (F) Uk37 Mg/Ca Mg/Ca Faunal Faunal Faunal Faunal Uk37 Mg/Ca Mg/Ca Faunal Faunal Season ΔSST** Reference Annual 1.5 Yamamoto et al. [2007] JJA 1.0 CLIMAP Project Members [1984] DJF 1.2 CLIMAP Project Members [1984] Annual 1.1 Yamamoto et al. [2007] Annual 2.7 Herbert et al. [2001] Annual 2.6 Nürnberg et al. [2008] Annual 1.7 Herbert et al. [2001] JJA -1.3 CLIMAP Project Members [1984] DJF -1.2 CLIMAP Project Members [1984] JJA -0.8 CLIMAP Project Members [1984] DJF -1.8 CLIMAP Project Members [1984] Annual 0.5 Clemens et al. [2008] JJA -0.8 CLIMAP Project Members [1984] DJF -2.3 CLIMAP Project Members [1984] JJA 0.6 CLIMAP Project Members [1984] DJF -0.1 CLIMAP Project Members [1984] Annual 1.2 Budziak [2001] Annual 1.5 Bard et al. [1997] Annual 0.9 Schmidt et al. [2004] JJA 0.4 CLIMAP Project Members [1984] DJF -1.3 CLIMAP Project Members [1984] Annual 1.4 Herbert and Schuffert [2000] JJA -1.5 CLIMAP Project Members [1984] DJF -1.0 CLIMAP Project Members [1984] JJA 1.1 CLIMAP Project Members [1984] DJF 0.2 CLIMAP Project Members [1984] JJA 0.3 Wolff et al. [1999] DJF 0.3 Wolff et al. [1999] Annual 0.9 Bard et al. [1997] Annual 1.1 Weldeab et al. [2007] Annual 2.1 Lea et al. [2000] JJA 0.0 CLIMAP Project Members [1984] DJF -2.4 CLIMAP Project Members [1984] JJA -0.1 CLIMAP Project Members [1984] DJF -3.6 CLIMAP Project Members [1984] Annual 0.3 Jasper et al. [1994] Annual 2.7 Lea et al. [2006] Annual 0.6 Lea et al. [2000] JJA -0.1 CLIMAP Project Members [1984] DJF -0.6 CLIMAP Project Members [1984] 228 Aux Table 1. (Continued) Core Site VNTR1-8 MD85668 V22-182 V22-182 V22-182 V22-182 RC13-115 GeoB1105 RC13-205 RC13-205 RC13-205 RC13-205 V19-29 V19-29 GeoB1112 GeoB10083 Y71_9_101 RC11-230 RC11-230 V22-38 V22-38 V22-38 V22-38 V22-174 V22-174 V22-174 V22-174 GeoB1016-3 GeoB1016-3 Y71-6-12 Y71-6-12 ODP820 v19-53 v19-53 V28-345 V28-345 GeoB10285 RC13-228 RC13-228 RC13-228 Lat 0.0 0.0 -0.6 -0.6 -0.6 -0.6 -1.4 -1.7 -2.3 -2.3 -2.3 -2.3 -3.6 -3.6 -5.8 -6.6 -6.6 -8.8 -8.8 -9.6 -9.6 -9.6 -9.6 -10.1 -10.1 -10.1 -10.1 -11.8 -11.8 -16.4 -16.4 -16.6 -17.0 -17.0 -17.7 -17.7 -20.1 -22.3 -22.3 -22.3 Long -110.5 46.0 -17.3 -17.3 -17.3 -17.3 -104.5 -12.4 5.2 5.2 5.2 5.2 -83.2 -83.2 -10.7 10.3 -107.0 -110.8 -110.8 -34.3 -34.3 -34.3 -34.3 -12.8 -12.8 -12.8 -12.8 11.7 11.7 -77.9 -77.9 146.3 -113.5 -113.5 118.0 118.0 9.2 11.2 11.2 11.2 Proxy* Faunal (R) Uk37 Faunal (C) Faunal (F) Faunal (C) Faunal (F) Faunal (R) Mg/Ca Faunal (F) Faunal (R) Faunal (F) Faunal (R) Faunal (R) Faunal (R) Mg/Ca Uk37 Faunal (R) Faunal Faunal Faunal (C) Faunal (F) Faunal (C) Faunal (F) Faunal (C) Faunal (F) Faunal (C) Faunal (F) Uk37 Uk37 Faunal Faunal Uk37 Faunal Faunal Faunal Faunal Uk37 Faunal (C) Faunal (F) Faunal (R) Season ΔSST** Reference Annual -0.8 Pisias and Mix [1997] Annual 1.3 Bard et al. [1997] JJA 1.8 CLIMAP Project Members [1984] JJA -2.4 CLIMAP Project Members [1984] DJF 0.6 CLIMAP Project Members [1984] DJF -3.0 CLIMAP Project Members [1984] Annual 0.0 Pisias and Mix [1997] JJA 2.1 Nürnberg et al. [2000] JJA 2.0 CLIMAP Project Members [1984] JJA 1.0 CLIMAP Project Members [1984] DJF 3.4 CLIMAP Project Members [1984] DJF 0.8 CLIMAP Project Members [1984] Annual -0.5 Pisias and Mix [1997] Annual -0.5 Pisias and Mix [1997] JJA 1.1 Nürnberg et al. [2000] Annual 1.2 Schneider et al. [1995] Annual -1.6 Pisias and Mix [1997] JJA -0.9 CLIMAP Project Members [1984] DJF -1.5 CLIMAP Project Members [1984] JJA 2.0 CLIMAP Project Members [1984] JJA 1.1 CLIMAP Project Members [1984] DJF 0.6 CLIMAP Project Members [1984] DJF 0.5 CLIMAP Project Members [1984] JJA -1.7 CLIMAP Project Members [1984] JJA -0.8 CLIMAP Project Members [1984] DJF -0.3 CLIMAP Project Members [1984] DJF -0.7 CLIMAP Project Members [1984] Annual 0.9 Müller et al. [1994] Annual 0.9 Schneider et al. [1995] JJA -4.9 CLIMAP Project Members [1984] DJF -5.3 CLIMAP Project Members [1984] Annual 0.7 Lawrence and Herbert [2005] JJA 3.1 CLIMAP Project Members [1984] DJF 1.0 CLIMAP Project Members [1984] JJA -0.9 CLIMAP Project Members [1984] DJF -0.8 CLIMAP Project Members [1984] Annual 2.7 Schneider et al. [1995] JJA 1.9 CLIMAP Project Members [1984] JJA -4.1 CLIMAP Project Members [1984] JJA 2.8 CLIMAP Project Members [1984] 229 Aux Table 1. (Continued) Core Site RC13-228 RC13-228 RC13-228 GeoB1712-4 GeoB1711-4 GeoB1710-3 RC13-229 RC13-229 RC13-229 RC13-229 RC11-86 RC11-86 RC11-86 RC11-86 RC12-294 RC12-294 RC12-294 RC12-294 MD97-2121 MD97-2121 ODP1089 ODP1089 ODP1123 PS2489-2 RC8-39 RC8-39 V22- 108 V22- 108 MD84-527 RC11-120 RC11-120 RC11-120 RC11-120 MD73025 MD73025 DSDP 90-594 MD97-2120 MD88-770 MD88-770 SO136-111 Lat -22.3 -22.3 -22.3 -23.3 -23.3 -23.4 -25.5 -25.5 -25.5 -25.5 -35.8 -35.8 -35.8 -35.8 -37.3 -37.3 -37.3 -37.3 -40.4 -40.4 -40.9 -40.9 -41.8 -42.9 -42.9 -42.9 -43.2 -43.2 -43.5 -43.5 -43.5 -43.5 -43.5 -43.8 -43.8 -45.5 -45.5 -46.0 -46.0 -50.7 Long Proxy* 11.2 Faunal (C) 11.2 Faunal (F) 11.2 Faunal (R) 12.8 Uk37 12.4 Uk37 11.7 Uk37 11.3 Faunal (F) 11.3 Faunal (R) 11.3 Faunal (F) 11.3 Faunal (R) 18.5 Faunal (C) 18.5 Faunal (F) 18.5 Faunal (C) 18.5 Faunal (F) -10.1 Faunal (C) -10.1 Faunal (F) -10.1 Faunal (C) -10.1 Faunal (F) 178.0 Uk37 178.0 Uk37 9.9 Faunal (R) 9.9 Faunal (R) -171.5 Faunal 9.0 Faunal (F) 42.4 Faunal (R) 42.4 Faunal (R) -3.3 Faunal -3.3 Faunal 51.2 Faunal (F) 79.9 Faunal (R) 79.9 Faunal (R) 79.9 Mg/Ca 79.9 Faunal (F) 51.3 Faunal 51.3 Faunal 174.9 Faunal (F) 174.9 Mg/Ca 96.5 Cd/Ca 96.5 Faunal (D) 160.2 Faunal (D) Season ΔSST** Reference DJF 2.1 CLIMAP Project Members [1984] DJF -3.3 CLIMAP Project Members [1984] DJF 2.1 CLIMAP Project Members [1984] Annual 1.6 Kirst et al. [1999] Annual 1.9 Kirst et al. [1999] Annual 1.1 Kirst et al. [1999] JJA -0.9 CLIMAP Project Members [1984] JJA 3.9 CLIMAP Project Members [1984] DJF -0.3 CLIMAP Project Members [1984] DJF 4.7 CLIMAP Project Members [1984] JJA 2.7 CLIMAP Project Members [1984] JJA -1.8 CLIMAP Project Members [1984] DJF 2.5 CLIMAP Project Members [1984] DJF -3.2 CLIMAP Project Members [1984] JJA 2.7 CLIMAP Project Members [1984] JJA 1.1 CLIMAP Project Members [1984] DJF 3.4 CLIMAP Project Members [1984] DJF 2.1 CLIMAP Project Members [1984] Annual 2.8 Pahnke et al. [2006] DJF 2.5 Pahnke et al. [2006] DJF 2.0 Cortese and Abelmann [2002] DJF 2.6 Cortese et al. [2007] Annual 2.0 Crundwell et al. [2008] DJF 1.3 Becquet and Gersonde [2003] JJA -0.8 CLIMAP Project Members [1984] DJF -0.7 CLIMAP Project Members [1984] JJA -0.3 CLIMAP Project Members [1984] DJF -0.7 CLIMAP Project Members [1984] DJF 0.6 Pichon et al. [1992] JJA -2.7 CLIMAP Project Members [1984] DJF -3.7 CLIMAP Project Members [1984] Annual -0.8 Mashiotta et al. [1999] DJF 0.8 Martinson et al. [1987] JJA 1.2 CLIMAP Project Members [1984] DJF 1.3 CLIMAP Project Members [1984] DJF 1.5 Schaefer et al. [2005] SON 2.1 Pahnke et al. [2003] Annual 3.0 Rickaby and Elderfield [1999] Annual 0.1 Sowers et al. [1993] DJF 0.2 Crosta et al. [2004] 230 Aux Table 1. (Continued) Core Site Lat Long Proxy* Season ΔSST** Reference V18-68 -54.6 -77.9 Faunal JJA -0.1 CLIMAP Project Members [1984] V18-68 -54.6 -77.9 Faunal DJF -0.3 CLIMAP Project Members [1984] MD84-551 -55.0 73.2 Faunal (F) DJF 0.4 Pichon et al. [1992] * In cases where more than one type of faunal proxy was used, further details are provided; C = Coccoliths, F = Foraminifera, R = Radiolaria, D = Diatoms. In Kandiano and Bauch [2003], M= Modern Analog Technique, T = Transfer Function Technique ** All ΔSST values were determined by subtracting late Holocene (5-0 ka) average SST from the average SST for the 5 kyr interval centered on the warmest value between 135 and 118 ka for each record, except for those from CLIMAP Project Members [1984]. CLIMAP ΔSST values were calculated as the difference between MIS 5e SST and core top SST estimates for each record. 231 Auxiliary Table 2. The effect of grid size on ocean-area weighted global mean ΔSST. 10° x 10° spacing was used in the study Longitude grid spacing (°) 5 Latitude grid spacing (°) 10 15 20 30 10 0.7 ± 0.7 0.7 ± 0.6 0.6 ± 0.5 0.5 ± 0.5 0.4 ± 0.4 20 0.7 ± 0.6 0.6 ± 0.5 0.5 ± 0.4 0.4 ± 0.4 0.4 ± 0.3 30 0.7 ± 0.6 0.7 ± 0.5 0.5 ± 0.4 0.4 ± 0.4 0.5 ± 0.3 45 0.7 ± 0.6 0.7 ± 0.5 0.5 ± 0.4 0.4 ± 0.4 0.5 ± 0.3 60 0.7 ± 0.6 0.6 ± 0.5 0.5 ± 0.4 0.5 ± 0.4 0.5 ± 0.3 Global SST data Coverage error, 1−yr means, 10 x 10 grids 140 1 120 St. Dev.= 0.08 100 # iterations Global weighted−mean SST anomaly calculated using data from grids in paleo analysis only (°C) 232 0 80 60 40 −1 −1 0 1 Global weighted−mean SST anomaly calculated using all data (°C) 20 0 −0.4 −0.2 0 0.2 Global SST Error 10 x 10 grid data coverage Figure C-S1. Data coverage error analysis. Top left: relationship between global SST calculated using full data coverage vs coverage limited to the sites of the paleooceanographic data. Red line shows 1:1 relationship. Top right: Histogram of the SST errors (residuals from top left). Bottom: Spatial coverage of the paleoceanographic data (green cells) using 10° by 10° grid cells, as used in the study. Figure C-S2. Surface temperature simulated for 130 ka (top left), 125 ka (top right), the difference between the two (bottom left) and the significances of the differences (bottom right). The area weighted global surface temperature difference between the two simulations is 0.09°C. 233 234 APPENDIX D ON THE POTENTIAL OF RAMAN-SPECTROSCOPY-BASED CARBONATE MASS SPECTROMETRY Submitted for publication in the professional journal: Journal of Raman Spectroscopy 235 ON THE POTENTIAL OF RAMAN-SPECTROSCOPY-BASED CARBONATE MASS SPECTROMETRY Nicholas P. McKay1, David L. Dettman1, Robert T. Downs1 and Jonathan T. Overpeck1,2,3 1 Department of Geosciences, University of Arizona Institute of the Environment, University of Arizona 3 Department of Atmospheric Science, University of Arizona 2 236 Abstract The potential for using Raman spectroscopy to measure 18O-16O oxygen isotopic ratios in carbonates is evaluated by measuring the Raman spectra and isotope ratios of a suite of sixty synthesized, isotopically-enriched calcite crystals ranging in composition from natural (0.2% 18O) to 1.2% 18O. We determined the Raman-inferred isotopic ratios (RRaman) by fitting curves to the ν1 symmetric stretching peak at 1086 cm-1 and the smaller satellite peak, associated with the ν1 stretching mode of singly-substituted carbonate groups (C16O218O) at 1065 cm-1. The ratio of the two peak areas shows a 1:1 correspondence with the 18O/16O ratios measured by the water and carbonate mass spectrometry, confirming that the relative intensities of the ν1 symmetric stretching peaks is a direct measure of the isotopic ratio in the carbonates. The 1-sigma uncertainties of the RRaman values of the individual crystals were 0.00079 (384‰ PDB), and 0.00043 (210‰ PDB) for the 4-crystal sample means. Although this level of uncertainty is too high to provide significant estimates of natural variability, there are multiple prospects for improving the accuracy and precision of the technique. Carbon isotope ratios in carbonates cannot be measured by our approach, however our results do highlight the potential of Raman-based mass spectrometry for C and other elements in minerals and organic compounds. Introduction Raman spectroscopy can potentially be used as a non-destructive, high-resolution mass spectrometer for a variety of substances, because differences in mass within a 237 molecule can be observed as shifts in the intensity or position of peaks in Raman spectra. Such an approach has been utilized for carbon isotopic composition in CO2 fluid inclusions in minerals (Arakawa et al., 2007), and pronounced shifts in the Raman spectra of organic molecules allow 13C to be used as a tracer of carbon uptake in individual microbes (Huang et al., 2007). Here, we investigate the potential of Raman spectroscopy as a method to directly measure O isotopic ratios in carbonates. The Raman spectra of carbonates are typically characterized by a well-defined ν1 symmetric stretching peak near 1100 cm-1 (1086 cm-1 in calcite) (Rutt and Nicola, 1974; White, 1974). Typically, a second, much smaller peak is also observed around 20 wavenumbers lower (1065 cm-1 in calcite). This satellite peak is the ν1 symmetric stretching peak associated with singlysubstituted carbonate groups (C16O218O) (Cloots et al., 1991; Gillet et al., 1996). Furthermore, the intensity of the C16O218O peak relative to the C16O3 peak is higher in synthetic carbonates grown in waters enriched in 18O (Gillet et al., 1996). For brevity, in this paper, we will refer to the ν1 symmetric stretching peak of the unsubstituted carbonate groups (C16O3) as the ν116O peak, and the ν1 symmetric stretching peak of the singlysubstituted carbonate groups (C16O218O) as the ν118O peak. The relative amount of 18O in carbonate is a widely used indicator of geologic and environmental variability, both in modern and past systems. The relative amount is typically expressed as the per mil deviation of the 18O/16O ratio from a standard (typically the Pee Dee Belemnite), or δ18OPDB. The δ18O of carbonates is a primary means of inferring past climatic and hydrologic variability from a wide variety of materials and environments, including cave deposits (speleothems), corals, carbonate shells from 238 marine and freshwater organisms, and inorganic and organic carbonates deposited in lake sediments and soils. A Raman-spectroscopy-based carbonate mass spectrometer could allow for rapid, non-destructive analysis at extremely high resolution (micron-scale), that would have abundant applications in environmental and geologic research. This study takes the first step towards this goal by determining the relationship between the intensities of the two Raman peaks and calcite 18O/16O, the precision with which this ratio can be determined from Raman spectra, and the potential application of this Raman mass spectrometer to natural systems. Experimental Procedure Calcite synthesis To create a set of a calcites with a wide range of oxygen isotopic ratios we synthesized 15 calcite samples in waters with oxygen isotope ratios (18O/16O) ranging from 0.002 (natural) to 0.012. To generate the waters, we initially diluted 3 ml of 97% 18O water from Sigma-Aldrich with 172 ml of distilled water. After precipitation of the mostenriched calcite, the water was further diluted to produce the remaining samples. To precipitate calcite from the enriched waters, 1-2 grams of laboratory-grade calcite were dissolved in the waters while they were bubbled with CO2 for several hours. The waters were filtered, and then capped and allowed to sit for two days to allow complete equilibration of the oxygen isotopes with HCO3- in the water. Next, the water was bubbled with air for two days to drive off dissolved CO2, driving the precipitation of calcite. Finally, the waters were sonicated and filtered, and the filter was dried, while the 239 remaining water was preserved for isotopic analysis, and further dilution and calcite synthesis. The synthesized calcites were generally euhedral crystals ranging from 25 to 50 μm in size. Raman Spectroscopy The Raman spectra of the synthesized calcites were collected on a Thermo Almega microRaman system, using a solid-state laser with an excitation wavelength of 780 nm at 100% power, and a thermoelectrically cooled CCD detector. The spectra were collected under 10x magnification, with 0.5 cm-1 resolution and a 1-μm spot size. For each sample, four individual calcite crystals were analyzed; Raman spectra were collected for fifteen minutes on each crystal. To increase the signal to noise ratio, only crystals with one face oriented orthogonal to the laser were analyzed, beyond this constraint, crystal orientation was random. Raman peak positions and intensities were determined by fitting pseudoVoigt curves to the ν118O and ν116O peaks centered at near 1065 and 1086 cm-1 respectively, corresponding to the ν1 symmetric stretching modes of the unsubstituted and singly-substituted carbonate groups (Cloots et al., 1991; Gillet et al., 1996). The ν116O peak, as measured by the Thermo Almega microRaman system is asymmetric, and required three pseudoVoigt curves centered on ~1086, ~1084 and ~1076 cm-1, to properly fit (Figure D1). This asymmetry is not observed in carbonates measured on other Raman systems (e.g., Cloots et al., 1991; Gillet et al., 1996), and is likely due to incomplete polarization of the laser interacting with the orientation of the crystal. The much smaller ν118O peak, centered near 1065 cm-1, can be fit with a single pseudoVoigt 240 curve. Before fitting peaks to the spectra, the background was removed from each spectrum using the software package CrystalSleuth (Laetsch and Downs, 2006). All four peaks in the ν1 stretching region were then fit simultaneously using a weighted nonlinear least squares fitting algorithm in Matlab 7.12 (The Mathworks, Inc.; Figure D1). The Raman-inferred 18O/16O ratio (RRaman) is then calculated as 1150 ∫ R Raman = 1150 1000 ∫∑ 1000 I ( ν 18 1 O) ∗ I ( ν 16 1 O 1,2,3 ) 1 3 (1) Where I(ν118O) are the intensities of the ν118O peak, and the intensities of the three ν116O peaks (ν116O1, ν116O2, ν116O3) are summed I(ν116O1,2,3) before numerically integrating using Matlab (Mathworks, Inc.). RRaman is one third of the integrated intensity ratio because only one of the three atoms in C16O218O is 18O. To investigate the potential of using Raman spectra to infer changes in calcite 13C/12C ratios, Raman spectra were also collected from a 13C-enriched (97% 13C) calcite and a natural-abundance reagent calcite (1.1% 13C) calcite from J.T. Baker Chemical. Isotope notation We use several notations common to isotopic studies in this paper, which we define here. R is the relative abundance of two isotopes, in this case 18O/16O. Isotopic fractionation, α, is the fractional difference between the R of two substances (A and B), defined as: 241 α A−B = RA RB (2) Finally, delta notation is used to express per mil (‰) differences from a standard, such that for oxygen: δ 18 O= ( R sample −R std ∗1000 R std ) (3) where Rstd is the R of Vienna Standard Mean Ocean Water (VSMOW) for waters and Vienna Pee Dee Belemnite (VPDB) for carbonates. Water Mass Spectrometry To measure the oxygen isotope ratios in the waters used for calcite synthesis, we diluted the enriched waters with distilled water (-8.9‰ VSMOW) in proportions resulting in a mixture with a predicted δ18O value less than +50‰. The diluted waters were then analyzed for δ18O using a dual inlet mass spectrometer (Delta-S, Thermo-Finnegan, Bremen, Germany) using an automated CO2-H2O equilibration unit. Standardization is based on internal standards referenced to VSMOW and VSLAP. Precision is better than ±0.08‰ for δ18O. The resulting δ18O values along with the measured masses of the two waters were used to determine the oxygen isotope ratios of the waters used to synthesize calcite. Because the mass spectrometer responds linearly between the calibration points of -55.5 (VSLAP) and 0‰ (VSMOW), we assume that the system maintains a linear response up to +50‰ VSMOW. 242 Carbonate Mass Spectrometry The fractionation of oxygen isotopes during the precipitation of calcite is relatively well known (α=1.0288 at 25°C; Friedman and O'Neill, 1977; or α=1.0285 at 25°C. Kim and O'Neill, 1997), and uncertainties are small compared to the uncertainty associated with calculating the oxygen isotope ratio from Raman spectra. Given the low precision of the Raman-spectra-derived ratios, calculating the oxygen isotope composition of the precipitated calcites from the δ18O value of the waters should be sufficiently precise. To verify this assumption, we measured the δ18O of five of the synthesized carbonate samples. Samples HC14 and HC15 had oxygen isotope ratios approaching natural abundances and could be measured without dilution. The δ18O and δ13C of these carbonates were measured using an automated carbonate preparation device (KIEL-III) coupled to a gas-ratio mass spectrometer (Finnigan MAT 252). Powdered samples were reacted with dehydrated phosphoric acid under vacuum at 70°C. The isotope ratio measurement is calibrated based on repeated measurements of NBS-19 and NBS-18 and precision is ±0.10‰ for δ18O and ±0.08‰ for δ13C (1 sigma). Samples HC2, HC5 and HC8 were too enriched in 18O to be measured directly, and were diluted with J.T. Baker calcite (δ18Opdb=-15.65‰) such that the resulting mixture would have a mean δ18Opdb of ~20‰. To have precise dilutions, 10 mg of total calcite was reacted overnight in sealed glass tubes with dehydrated phosphoric acid at 25°C. The evolved CO2 gas was cleaned cryogenically and measured on a Finnigan Delta-S gas-ratio mass spectrometer. The oxygen isotope fractionation between calcite and acid-liberated CO2 was taken from Swart et al. (1983). 243 Results and Discussion The Raman-inferred, and mass-spectrometer-measured water and calcite R and δ18O values for the 15 samples are presented Table D1. The five calcite samples measured on the mass spectrometer generally agree with the values predicted from the waters, within the uncertainty associated with the mass dilution analysis procedure. In samples HC14 and HC15, the δ18O values predicted from the waters underestimate those directly measured from the calcites by 2.4 and 1.9‰, respectively. These offsets may be associated with the extrapolation of the water and carbonate calibrations towards more enriched values, but may also be due to enrichment in the calcites associated with high dissolved [Ca+] during calcite synthesis (cf., Kim and O'Neill, 1997). Overall, the δ18O and R values inferred from the water and carbonate mass spectrometry appear to be sufficiently consistent for comparison with the less-certain Raman-inferred values. For consistency, in the rest of this paper, we compare the Raman-inferred isotope ratios (RRaman) with the calcite isotope ratios inferred from the water measurements (Rcalc-inf). Raman-inferred isotope ratios (RRaman) The RRaman values as calculated by equation 1 are reasonable for both the individual crystals and sample means (Table 1). The variability RRaman between crystals of the same batch displays a range in standard deviations from 0.00012 to 0.00166. Remarkably, the RRaman values demonstrate a clear 1:1 relationship with the Rcalc-inf values, especially when comparing the 4-crystal mean RRaman values (Figure D2). The residuals between the RRaman and Rcalc-inf values are normally distributed, with a standard deviation 244 of 0.00079 (corresponding to 384‰) for the individual crystals, and 0.00043 (210‰) for the 4-crystal means values. This range of uncertainty is about an order of magnitude outside the range of natural variability; therefore, this approach is not yet sufficiently resolved for most geologic and environmental applications. However, it should be noted that this is an exploratory study, and a number of existing-technology refinements could be applied that would significantly improve the accuracy and precision of our approach. These will be discussed below. Potential sources of error and bias Because the ν118O peak is small, particularly at lower, closer-to-natural R values, it would be expected that the signal to noise ratio (SNR) in the ν118O peak would be a primary factor driving the accuracy and precision of the RRaman values. However, this does not appear to be the case in these samples. There are no significant relationships between the residuals and the ν118O peak to background ratio, the ν118O:ν116O ratios, or even the root-mean squared error (RMSE) of the fit. Additionally, the size and breadth of the third ν116O peak centered at 1076 cm-1 (ν116O3) controls the separation between the ν116O and ν118O peaks, and in some cases, such as for spectra from HC6, which have the highest residuals of any samples, clearly affects the quality of the ν118O peak fit. Despite this, various metrics of the relative size and influence of the ν116O3 peak show no consistent relationship with the accuracy of RRaman, either for sample means or individual crystals. The only metric of SNR that shows a clear relationship with the residuals, for the samples with small relative ν118O peaks (R < 0.005), is the ratio between ν116O height and 245 background (the height of the nearby baseline, between 1020 and 1040 cm-1, before adjustment by CrystalSleuth; Laetsch and Downs, 2006) (Figure D3). This result suggests that the overall signal strength of the ν1 symmetric stretching mode is a primary control on the precision and accuracy of the RRaman estimates. Potential for improving the accuracy and precision of RRaman estimates There is considerable potential for reducing the uncertainty of the RRaman estimates, and increasing the utility of Raman-spectroscopic carbonate mass spectroscopy, even with existing techniques and technology. First, it should be emphasized that the Raman spectra collected here were preliminary, and that virtually no effort was made to screen crystals based on orientation or signal strength. Still, the fact that the Raman intensity ratios show a 1:1 relationship with mass spectrometer values, with no apparent biases, is extremely encouraging for future applications of the technique. There are several simple ways to potentially reduce uncertainty in the estimates. The incomplete polarization of the microAlmega Raman system causes the ν116O peak to be split into three smaller peaks, including a small peak near 1076 cm-1 that can influence the ν118O peak. Using a fully polarized spectrometer may reduce the overlap between the peaks resulting in a better fit. Second, the result that the ν116O peak height to background ratio affects the accuracy indicates that efforts to increase the overall signal to noise ratio of the Raman spectra, even screening crystals by signal strength, would likely reduce uncertainty in RRaman. Increasing the spectral resolution by focusing the CCD on a smaller area of the spectrum should increase the quality of the peak fit, and eliminate aliasing driven by the 246 precise position of the peak (cf. Arakawa et al., 2007). Another approach is resonance Raman (RR) scattering, which occurs when the wavelength of the exciting laser is chosen such that its energy corresponds to the electronic transition of specific atomic bonds within a molecule (Ferraro et al., 2003). RR spectroscopy can greatly increase (by a factor of 103 to 105) the signal strength of a Raman spectrum (Ferraro et al., 2003), and could potentially allow for much more precise determination of area of the ν118O peak. A simpler approach is immersing the sample in liquid nitrogen, which increases the separation of nearby peaks, and increases the sharpness of the peaks (Figure D5). The accuracy of RRaman would likely improve with frequent comparison of Raman-inferred values to an internal standard during sample collection. The use of relative measurements in isotope ratio mass spectrometry has allowed greatly improved precisions to be attained (Criss, 1999). Relative deviations from a standard can be measured far more accurately than the determination of the absolute 18O/16O ratio, as we did here. Carbon isotopes The displacement of the carbon atom in the ν1 symmetric stretching mode is not very large, so there is no significant shift of frequency in the 1050-1100 cm-1 region between the normal and 13C-enriched calcites. However, the substituted C should be apparent in stretching modes where the C atom does undergo large motions, such as the ν13 active translational mode near 281 cm-1. Here, the 13C-enriched calcite appears shifted to lower wavenumbers (centered near 279 cm-1) (Figure D4). This shift is consistent with theoretical expectation that the fractional change in position is a function of the square 247 root of the changes in mass (Gillet et al., 1996). In this case for the ν13 mode of Ca13CO3: ν 13 / ν 13= √( m 13 C 16 O / m 12 C 16 O ) =0 . 9918 (4) corresponding to a theoretical ν13 peak at 279 cm-1. Because the change in mass is small, and the peak is at low wavenumbers, small changes in the size of the Ca13CO3 ν13 peak would be completely obscured by the Ca12CO3 peak. This does not mean that the 13C/12C ratios could not be measured in other minerals or molecules. For example, the C-C stretching mode in the calcium oxalate weddellite (CaC2O4·2H2O) has a satellite peak at ~870 cm-1 (Figure D6). The position (following equation 2) and intensity of the peak suggest that the satellite peak is associated with a 13C-12C bond. Potential for quantifying isotope ratios in other molecules Following the discussion of the previous section on the inability to quantify changes in 13C/12C ratios in calcite Raman spectra, it makes sense to generalize the conditions necessary to directly quantify isotope ratios from Raman spectra. There are three primary criteria necessary to effectively fit and quantify peaks in Raman spectra, and to relate those peaks to isotopic ratios in the compound of interest. First, to avoid the complication of the effect of sample orientation on relative intensities, the two peaks must be associated with the same vibrational mode. This constraint drives the second criterion: the shift in the wavenumber associated with the change in mass must be large enough so that the secondary peak is distinct and identifiable from the primary peak. 248 Following equation 4, this is a function of both the relative change in mass in the molecule, and the position of the stretching peak, where larger changes in mass and vibrational modes at higher wavenumbers result in larger shifts. Finally, the rare isotope of interest must be sufficiently abundant that the associated peak is above the background intensity. Despite these constraints, it is likely that isotopic ratios in a wide variety of organic and inorganic compounds could be quantified from Raman spectra, with variable precision. Conclusions We have synthesized a suite of 18O-enriched calcites, and measured their 18O/16O ratios with both traditional mass spectrometry and through their Raman spectra. The Raman-inferred ratios show a clear 1:1 relationship with the ratio calculated from traditional mass spectrometry measurements, suggesting that oxygen isotope ratios can be measured directly in Raman spectra. The measurements of individual crystals were precise to within a ratio of 0.00079 (384‰) and the 4-crystal means were precise to within 0.00043 (210‰). These uncertainties are too large to be useful in natural systems, however this is a preliminary study, and the accuracy and precision could be greatly improved by increasing the signal to noise ratio of the Raman spectra, using a spectrometer that maintains the polarization of the laser, and refining our curve-fitting techniques. Carbon isotope ratios in carbonates cannot be measured by our approach, however our results do highlight the potential of Raman-based mass spectrometry for C and other elements in minerals and organic compounds. 249 References Arakawa, M., J. Yamamoto, and H. Kagi (2007), Developing micro-Raman mass spectrometry for measuring carbon isotopic composition of carbon dioxide, Applied spectroscopy, 61(7), 701–705. Cloots, R. (1991), Raman spectrum of carbonates MIICO3 in the 1100-1000 cm−1 region: observation of the ν1, mode of the isotopic (C16O182O)2 ion, Spectrochimica Acta Part A: Molecular Spectroscopy, 47(12), 1745–1750, doi:10.1016/05848539(91)80012-8. Criss, R. E. (1999), Principles of stable isotope distribution, Oxford University Press, USA. Ferraro, J. R. (2002), Introductory Raman Spectroscopy, Second Edition, 2nd ed., Academic Press. Friedman, I., and J. R. O’Neil (1977), Compilation of stable isotope fractionation factors of geochemical interest, USGPO. Gillet, P., P. McMillan, J. Schott, J. Badro, and A. Grzechnik (1996), Thermodynamic properties and isotopic fractionation of calcite from vibrational spectroscopy of 18 O-substituted calcite, Geochimica et cosmochimica acta, 60(18), 3471–3485. Huang, W. E., K. Stoecker, R. Griffiths, L. Newbold, H. Daims, A. S. Whiteley, and M. Wagner (2007), Raman-FISH: combining stable-isotope Raman spectroscopy and fluorescence in situ hybridization for the single cell analysis of identity and function, Environmental microbiology, 9(8), 1878–1889. Kim, S., and J. Oneil (1997), Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates, Geochimica et Cosmochimica Acta, 61, 3461–3475, doi:10.1016/S0016-7037(97)00169-5. Laetsch, T., and R. Downs (2006), Software for identification and refinement of cell parameters from powder diffraction data of minerals using the RRUFF project and American Mineralogist Crystal Structure Databases, 19th General Meeting of the International Mineralogical Association, Kobe, Japan, pp. 23–28. Rutt, H. N., and J. H. Nicola (1974), Raman spectra of carbonates of calcite structure, Journal of Physics C: Solid State Physics, 7, 4522. Swart, P. K., S. J. Burns, and J. J. Leder (1991), Fractionation of the stable isotopes of oxygen and carbon in carbon dioxide during the reaction of calcite with phosphoric acid as a function of temperature and technique, Chemical Geology: 250 Isotope Geoscience section, 86(2), 89–96. White, W. B. (1974), The carbonate minerals, The infrared spectra of minerals, 4, 227– 284. 251 Table D1. Water and Calcite oxygen isotope data, measured by traditional mass spectrometry and Raman spectroscopy. Water and calcite mass spectrometry Raman-based mass spectrometry 18 Sample Rwaters Rcalc-inf δ18OPDB Rcalc-meas* δ OPDB Rraman X1 Rraman X2 Rraman X3 Rraman X4 Rraman Mean HC01 0.0121 0.0125 5042.4 0.0143 0.0128 0.0135 0.0127 0.0133 HC02 0.0086 0.0088 3258.2 0.0083 3014.9 0.0082 0.0092 0.0093 0.0093 0.0090 HC03 0.0054 0.0055 1672.8 0.0055 0.0053 0.0060 0.0056 0.0056 HC04 0.0048 0.0050 1406.1 0.0045 0.0053 0.0052 0.0049 0.0050 HC05 0.0038 0.0039 868.2 0.0036 721.8 0.0036 0.0022 0.0044 0.0040 0.0036 HC06 0.0033 0.0034 623.7 0.0054 0.0065 0.0036 0.0029 0.0046 HC07 0.0028 0.0029 386.9 0.0029 0.0012 0.0028 0.0029 0.0025 HC08 0.0027 0.0028 329.2 0.0027 310.5 0.0031 0.0015 0.0036 0.0036 0.0029 HC09 0.0024 0.0025 206.3 0.0030 0.0021 0.0029 0.0028 0.0027 HC10 0.0024 0.0024 169.6 0.0038 0.0023 0.0033 0.0035 0.0032 HC11 0.0022 0.0023 110.2 0.0027 0.0021 0.0021 0.0027 0.0024 HC12 0.0021 0.0022 66.4 0.0023 0.0024 0.0024 0.0022 0.0023 HC13 0.0021 0.0022 53.2 0.0033 0.0026 0.0011 0.0030 0.0025 HC14 0.0021 0.0021 29.6 0.0021 32.0 0.0015 0.0029 0.0018 0.0016 0.0020 HC15 0.0020 0.0021 4.7 0.0021 6.6 0.0021 0.0026 0.0030 0.0030 0.0027 note: Rcalc-inf is the 18O/16O ratio of the calcite inferred from water composition (Kim and O'Neill, 1997), Rcalc-meas is the 18O/16O ratio of the calcite measured by mass spectrometry. RRaman values are shown for each of the four individual crystals measured, and their mean ratio. *Italicized calcite values were measured offline using mass dilution, and are less certain 252 −1 Wavenumber (cm ) 1090 1080 1070 1060 Rraman=0.0143 1070 1060 Intensity (normalized) Rraman=0.0045 1070 1060 Rraman=0.0027 1070 1090 1080 1070 −1 Wavenumber (cm ) 1060 1060 Figure D1. Examples of three measured Raman spectra in the ν1 symmetric stretching mode region, and the associated pseudoVoigt curve fits for each (top: HC1, middle: HC4, bottom: HC11). For all panels: Raman measurements are shown as black circles, the green, black and cyan curves show the fits to the three ν116O peaks, ν116O1, ν116O2, ν116O3, respectively; the dark blue curve shows the ν118O fit, and the red curve shows the sum of all four pseudoVoigt peaks. The Raman intensities for each sample are normalized to the maximum height of the ν116O peaks. The inset in each panel shows the ν118O region and associated curves with the Raman intensity scale amplified by a factor of 10 to better illustrate changes in ν118O peak area. The RRaman values calculated for each spectrum are shown in each panel. 253 Four−crystal mean 18O/16O Individual crystal mean 0.005 0 0 O/16O 0.01 RRaman RRaman 0.01 18 0.005 0.01 Rwaters 0.005 0 0 0.005 0.01 Rwaters Figure D2. Relation between Raman-inferred 18O/16O ratios (RRaman) and the R values predicted from the 18O/16O measured in the waters (Rcalc-inf), for both the 4-crystal mean of each sample (left) and measurements of the individual crystals (right). For each plot, a 1:1 relationship is shown in red. 254 16 Ov1 peak signal to noise ratio 70 60 50 40 30 20 10 0 0 0.001 0.002 0.003 Residual (RRaman − Rwaters) Figure D3. Relation between 18O/16O residuals (RRaman – Rcalc-inf) and the ν116O peak height to background ratio in the Raman spectra, for samples with R values < 0.005. 255 C−enriched (97% 13C) calcite 13 Standard (1.1% C) calcite Raman Intensity 13 300 290 280 270 −1 Wavenumber (cm ) 260 Figure D4. Raman spectra of 13C-enriched and standard calcite in the ν13 active translational mode region. 256 Intensity (normalized) Room temperature LN−cooled 1100 1095 1090 1085 1080 Wavenumber (cm−1) 1075 1070 Figure D5. The υ1 symmetric stretching mode region of Raman spectra of optical calcite before and after immersion in liquid nitrogen (LN). Intensity (normalized) 257 940 920 900 880 −1 Wavenumber (cm ) 860 Figure D6. Raman spectrum of the C-C stretching mode region in the calcium oxalate weddelite (CaC2O4·2H2O). Based on its position (following eq. 4) and relative intensity, the satellite peak at ~870 cm-1 likely corresponds to the 13 C-12C bond. 258 APPENDIX E PERMISSIONS 259 APPENDIX C: THE ROLE OF OCEAN THERMAL EXPANSION IN LAST INTERGLACIAL SEA LEVEL RISE is reprinted with permission of the American Geophysical Union. McKay, N. P., J. T. Overpeck, and B. L. Otto-Bliesner, The role of ocean thermal expansion in Last Interglacial sea level rise, Geophysical Research Letters, 38(14), L14605, 2011. Copyright 2011 American Geophysical Union.