Louisiana State University LSU Digital Commons

LSU Doctoral Dissertations Graduate School

2007 Paleoceanography of the Gulf of Papua using multiple geophysical and micropaleontological proxies Lawrence A. Febo Louisiana State University and Agricultural and Mechanical College, [email protected]

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_dissertations Part of the Earth Sciences Commons

Recommended Citation Febo, Lawrence A., "Paleoceanography of the Gulf of Papua using multiple geophysical and micropaleontological proxies" (2007). LSU Doctoral Dissertations. 3357. https://digitalcommons.lsu.edu/gradschool_dissertations/3357

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please [email protected]. PALEOCEANOGRAPHY OF THE GULF OF PAPUA USING MULTIPLE GEOPHYSICAL AND MICROPALEONTOLOGICAL PROXIES

A Dissertation Submitted to the Graduate Faculty of the Louisiana State University and Agricultural and Mechanical College in partial fulfillment of the requirements for the degree of Doctor of Philosophy

in

The Department of Geology and Geophysics

by Lawrence A. Febo B.S., Ohio State University, 1999 M.S., Ohio State University, 2003 December 2007

For my late advisor, Dr. John Wrenn

ii ACKNOWLEDGEMENTS

What you are about to read, and I certainly hope enjoy, is the result of a

significant effort not just on my part, but also from so many people who I would like to

thank. First, I thank my late advisor Dr. John Wrenn. John was an excellent mentor to

me and instrumental in shaping who I have now become as a scientist and teacher. He

was also a very good friend, never ceasing to surprise me with his vast scientific knowledge, genuine willingness to help others, and his ability to brighten my days with a

casual swipe of his witty humor. Being his student was a rare privilege and it is he to

who I dedicate this volume.

After some recovery time, I was very fortunate to be adopted by Brooks Ellwood

and I have benefited greatly from his scientific expertise. In fact, many of the tools and

ideas brought together in this dissertation are a result of his guidance. I also extend many

thanks to Dr. Barun Sen Gupta for all the time he took discussing data and ideas with me.

The stable isotope data in this dissertation would not have been possible without the help

of Dr. Brian Fry, thank you for teaching me so much. I am also grateful to Sam Bentley

for inviting me to be a part of the MARGINS Source-to-Sink research initiative and also

a part of his team over the years. I came to LSU to do marine geology and

paleoceanography, and he delivered. Sam provided countless opportunities unobtainable

anywhere else, circumnavigating Australasia certainly ranking high on the list, but none

as memorable as crossing the equator, facing King Neptune’s court, and walking the

plank all in the same day!

Research results presented in this dissertation was partly supported by NSF grant

OCE #0305373 awarded to Samuel Bentley. Results presented in Chapter 4 were funded

iii by LSU and a U.S. NSF -Egypt Joint Science and Technology Board grant to Dr.

Ellwood and Dr. Aziz M. Kafafy, Grant #OTH6-008-002. I would also like to thank BP

America for providing generous financial assistance to complete my research and also

gainful future employment. I also thank the Geological Society of America, the

American Association of Stratigraphic Palynologists, and the Department of Geology and

Geophysics at Louisiana State University. The Department has supported me with

generous assistantships and scholarships, even after my term limit.

Additionally, I thank Bob Olson of Baseline Resolution for discussions about

TOC and Rock-Eval Pyrolysis data. I also received thoughtful comments from Andre

Droxler, Gerry Dickens, Chuck Nittrouer, and two anonymous reviewers on an early

version of chapter 2.

Finally, I would like to thank my parents, family, and close friends for understanding my hawkish desire to do the Ph.D. I especially thank my shipmates Jason

Francis, Luke Patterson and Zahid Muhammad, and my “paleo” officemates Rebecca

Tedford, Jeff Agnew, and Juan Chow. Thanks for listening to me squawk the last five years.

iv TABLE OF CONTENTS

ACKNOWLEDGEMENTS...... iii

ABSTRACT……………………………………………………………………………...vi

CHAPTER 1. INTRODUCTION…………………………………………………………1

CHAPTER 2. LATE PLEISTOCENE AND HOLOCENE SEDIMENTATION, ORGANIC-CARBON DELIVERY, AND PALEOCLIMATIC INFERENCES ON THE CONTINENTAL SLOPE OF THE NORTHERN PANDORA TROUGH, GULF OF PAPUA...... 9

CHAPTER 3. SURFACE SEDIMENTS AND DISPERSAL PATHWAYS ON THE OUTER SHELF-EDGE AND BASIN IN THE GULF OF PAPUA: MAGNETIC SUSCEPTIBILITY, CALCIUM CARBONATE, AND ORGANIC GEOCHEMICAL MEASUREMENTS...... 67

CHAPTER 4. DETRITAL CONTROLS ON MAGNETIC SUSCEPTIBILITY AND CYCLOSTRATIGRAPHY RECORDS …...... 100

CHAPTER 5. CONCLUSIONS...... 121

APPENDIX A: PUBLISHED AND SUBMITTED CHAPTER ABSTRACTS...... 124

APPENDIX B: SUPPLEMENTAL CHAPTER 2 DATA TABLES NOT IN TEXT….128

APPENDIX C: CORE IMAGES OF MV-51JPC………………………………………137

APPENDIX D: PERMISSION REQUEST…………………………………………….138

APPENDIX E: COMPREHENSIVE REFERENCE LIST...... 140

VITA...... 156

v ABSTRACT

Recent marine and late Pleistocene sediments examined from the Gulf of Papua

(GoP), investigate the flux and fate of detrital sediments and organic

carbon over the last glacial-interglacial cycle. Based on surface sediment magnetic

susceptibility (MS) and calcium carbonate concentrations, recent marine sediment is

exported off the narrow shelf and into deeper regions via the Kerema Canyon of the

northern Pandora Trough. Detrital clastic sediment is then dispersed deeper into the

central and southern Pandora Trough. Except for pelagic deposition, very little material

reaches the Ashmore Trough and Eastern Plateau adjacent to the Great Barrier Reef.

Rock-Eval pyrolysis data and organic petrography indicate that late Pleistocene and recent organic matter were strongly degraded prior to burial.

Late Pleistocene depositional records come from two, 12 m piston cores retrieved

from the slope of the northern Pandora Trough. MV-54 was taken at the mid-slope of the

central Pandora Trough (923 m) and MV-51 was collected from a bathymetric high in the

northeastern Pandora Trough (804 m). Core sediments were analyzed with MS, calcium

carbonate, organic geochemistry, and benthic foraminiferal assemblages. Both cores

show two periods of rapid sediment accumulation. High accumulation rates characterize

15,800-17,700 Cal. yrs B.P. in MV-54 and correspond to the early transgression when

rivers delivered sediments closer to the shelf-edge. Benthic foraminiferal assemblages in

MV-51 indicate a seasonally variable flux of organic carbon during late LGM (~18,400-

20,400 Cal. yrs B.P.), suggesting enhanced contrast between monsoon seasons.

The oldest section, >32,000 14C yrs B.P., contains the highest mass accumulation

rates and TOC fluxes, with >50% of the organic carbon derived from C3 vascular plant

vi matter. MS and benthic foraminiferal accumulation rates are orders of magnitude higher

during this interval than any younger time indicating a greater influence of detrital

minerals and labile organic carbon. Because mineralogy and detrital input are shown to

be the main controls on MS variability, the MS data in this interval suggest more direct

dispersal pathways from central and eastern PNG Rivers to the core site when sea level was lower and dispersal gradients were higher.

vii CHAPTER 1

INTRODUCTION

Continental margins are one of the ultimate storage sites for sediments shed from

continental settings. Warm, wet tropical environments such as the Amazon Basin, west

central Africa, and Indonesia supply significant quantities of sediment and terrestrial

organic carbon to marine depositional systems [Meybeck, 1988; Milliman, 1995;

Nittrouer et al., 1995; Milliman et al., 1999; Schlünz and Schneider, 2000; Syvitski et al.,

2005]. At present, a minimum of 50% of the annual sediment discharged into the global ocean originates from tropical drainage basins [Meybeck, 1988] and accounts for

approximately 60% of the total organic carbon budget [Schlünz and Schneider, 2000]. In

the tropical western Pacific, Indonesia and Australasia are estimated to be responsible for

as much as 25% of the global sediment flux [Milliman et al., 1999] and ~40% of the

terrestrial organic carbon buried in coastal and shelf deposits [Schlünz and Schneider,

2000].

In addition to the importance of global sediment and organic carbon budgets, the

tropical western Pacific region is the location of dynamic climatic and oceanographic

features, making it an area of critical paleoceanographic importance for understanding

long-term sediment storage and organic carbon cycling [Bjerknes, 1969; Meyers et al.,

1986; Beaufort et al., 2001; Clemens et al., 2003; Wang et al., 2003]. The western

Pacific warm pool, or WPWP, is a large region north of Papua New Guinea (Figure 1.1)

and contains warm water with sea surface temperatures >29°C and an atmospheric low

pressure zone (Figures 1.2, 1.3). Fluctuations in the position of the Inter Tropical

1 Convergence Zone (ITCZ) causes the position of the WPWP to move southward during

the austral summer monsoon (January – March) and deliver large quantities of rain to the

young, mountainous terrain of Papua New Guinea and other Indonesian islands (Figure

1.1). However, the position shifts north of the equator with the onset of the austral winter

(July – September) causing relatively drier conditions [An, 2000; Clemens et al., 2003].

Marine productivity, as inferred from chlorophyll a concentrations, does not appear to

change significantly between seasons with the exception of a slight decrease in the seas

between the Indonesian islands when warmer sea surface temperatures occur during the

summer monsoon (Figure 1.2).

Surprisingly, the paleoceanography of the western Pacific is just now beginning to

be investigated in greater detail despite its potential to remove atmospheric carbon over

glacial-interglacial timescales [e.g. Kawahata, 1999; Müller and Opdyke, 2000; Beaufort et al., 2001; Kilbourne and Quinn, 2004]. Because the western tropical Pacific is the location of very active depositional systems, lush tropical forests, and a prominent climatic system, it is an excellent location to examine glacial - interglacial changes in sediment flux and organic carbon burial. In particular, the Gulf of Papua (GoP) is a large

tropical mixed siliciclastic-carbonate depositional system that is located between

northeastern and Papua New Guinea. The GoP is the recipient of between 300

and 400 x 106 tonnes per year (Mt y-1) in a basin one fourth the size of the Gulf of

Mexico [Milliman et al., 1999]. The contribution of sediment discharge from small, mountainous rivers for Papua New Guinea is shown in Figure 1.1 [see also Milliman et al., 1999].

2 In this dissertation, I address the issues of sediment accumulation rates, organic carbon burial, and inferred marine productivity over the past glacial-interglacial cycle

(LGM to recent). In chapter 2, I use two radiocarbon-dated piston cores to discuss changes in late Pleistocene to Holocene sediment accumulation rates, organic carbon burial, and the sources of the components preserved in the organic carbon pool. This chapter provides detailed information about how sedimentation rates and productivity responded with lower sea level and a cooler, drier climate system. Chapter 3 provides an examination of more recent sediment compositions and dispersal pathways in the Gulf of

Papua by examining surface sediments on the outer shelf-edge and basin in the Gulf of

Papua. Detrital and biogenic compositions are constrained by using magnetic susceptibility (MS), calcium carbonate concentrations, and organic geochemical measurements. Specifically, the quantity, type, and preservation of organic carbon in 68 core-top sediment samples are examined for the modern depositional system and contrasted with the late Pleistocene record.

Finally, in chapter 4, I examine the factors that control MS values and variations within marine cyclostratigraphy records. For example, MS is commonly used as a conservative tracer of terrigenous components in sediments [e.g. deMenocal et al., 1991;

Ellwood et al., 2000; Ellwood et al., 2006], but sometimes it is also argued to be a proxy for marine productivity [Meade and Tauxe, 1986; Warner and Domack, 2002]. Chapter 4 demonstrates that marine productivity alone cannot cause significant variability in MS records but instead that it is controlled by detrital fractions and mineralogy.

1.1 References

An, Z. (2000), The history and variability of the East Asian paleomonsoon climate, Quaternary Science Reviews, 19, 171-187.

3

Beaufort, L., T. de Garidel-Thoron, A.C. Mix, and N.G. Pisias (2001), ENSO-like forcing on oceanic primary production during the late Pleistocene, Science, 293, 2440-2444.

Figure 1.1. Map of Australasia and part of Indonesia showing major features of the western Pacific including seasonal weather patterns, surface circulation along the northeastern margin of Australia, and annual sediment flux directions and mass estimates for Papua New Guinea [map modified from Kershaw et al., 2003; flux estimates are in 106 tonnes per year and taken from Milliman et al., 1999]. The box denotes the Gulf of Papua study area, which is expanded in the lower right corner showing major rivers, bathymetry, and piston core sites. S.E.C. refers to the South Equatorial Current, E.A.C. is the East Australian Current, and C.S.C.C. is the Coral Sea Coastal Current. The position of the Inter Tropical Convergence Zone is shown for the austral summer (Jan. – Mar.) when it shifts below the equator. During austral winter (July - Sept.) it is positioned above the equator and out of the map area.

4

Figure 1.2. Summer Monsoon conditions. Sea surface temperatures, salinities, and chlorophyll a concentrations for part of the Indonesian Archipelago and Australasia. Temperature and salinity data are from World Ocean Atlas 2005 [Locarnini et al., 2005] and made with Ocean Data View software [Schlitzer, 2007].

5

Figure 1.3. Winter Monsoon conditions. Sea surface temperatures, salinities, and chlorophyll a concentrations for part of the Indonesian Archipelago and Australasia. Temperature and salinity data are from World Ocean Atlas 2005 [Locarnini et al., 2005] and made with Ocean Data View software [Schlitzer, 2007].

6

Bjerknes, J. (1969), Atmospheric teleconnections from the equatorial Pacific, Monthly Weather Review, 97, 163-172.

Clemens, S., P. Wang, and W. Prell (2003), Monsoons and global linkages on Milankovitch and sub-Milankovitch time scales, Marine Geology, 201, 1-3.

deMenocal, P., J. Bloemendal, and J. King (1991), A rock-magnetic record of monsoonal dust deposit ionto the Arabian Sea: Evidence for a shift in the mode of deposition at 2.4 Ma, in Proceedings ODP, Scientific Results 117, edited by W. Prell and L. Niitsuma, pp. 389-407.

Ellwood, B.B., W.L. Balsam, and H.H. Roberts (2006), Gulf of Mexico sediment sources and sediment transport trends from Magnetic Susceptibility measurements of surface samples, Marine Geology, 230, 237-248.

Ellwood, B.B., R.E. Crick, A. Hassani, S.L. Benoist, and R.H. Young (2000), Magnetosusceptibility event and cyclostratigraphy method applied to marine rocks: detrital input versus carbonate productivity, Geology, 28(12), 1135-1138.

Kawahata, H. (1999), Fluctuations in the ocean environment within the western Pacific warm pool during late Pleistocene, Paleoceanography, 14, 639-652.

Kilbourne, K.H. and T. M. Quinn (2004), A fossil coral perspective on western tropical Pacific climate ~350 ka, Paleoceanography, 19, PA1019, doi:10.1029/2003PA000944.

Locarnini, R.A., A.V. Mishonov, J.I. Antonov, T.P. Boyer, and H.E. Garcia (2005), World Ocean Atlas 2005, http://odv.awi.de/data/ocean/world-ocean-atlas-2005.html

Mead, G.A. and L. Tauxe (1986), Oligocene paleoceanography of the South Atlantic: Paleoclimatic implications of sediment accumulation rates and magnetic susceptibility measurements, Paleoceanography, 1(3), 273-284.

Meybeck, M. (1988), How to establish and use world budgets of riverine materials, in Physical and Chemical Weathering in Geochemical Cycles, edited by A. Lerman and M. Meybeck, The Hague, Kluwer Academic, pp. 242-272.

Milliman, J.D. (1995), Sediment discharge to the ocean from small mountainous rivers: the New Guinea example, Geo-Marine Letters, 15, 127-133.

Milliman, J.D. , K.L. Farnswerth, and C.S. Albertin (1999), Flux and fate of fluvial sediments leaving large islands in the East Indies, Journal of Sea Research, 41, 97-107.

Müller, A. and B.N. Opdyke (2000), Glacial-interglacial changes in nutrient utilization and paleoproductivity in the Indonesian Throughflow sensitive Timor Trough, eastern Indian Ocean, Paleoceanography, 15, 85-94.

7

Nittrouer, C.A., G.J. Brunskill, and A.G. Figueiredo (1995), Importance of tropical coastal environments, Geo-Marine Letters, 15, 121-126.

Schlitzer, R. (2007), Ocean Data View, http://odv.awi.de.

Schlünz, B. and R.R. Schneider (2000), Transport of terrestrial organic carbon to the oceans by rivers: Re-estimating flux- and burial rates, International Journal of Earth Sciences, 88, 599-606.

Syvitski, J.P.M., C.J. Vorosmarty, A.J. Kettner, and P. Green (2005), Impact of humans on the flux of terrestrial sediment to the global coastal ocean, Science, 308, 376-380.

Wang, B., S.C. Clemens, and P. Liu (2003), Contrasting the Indian and East Asian monsoons: implications on geologic timescales, Marine Geology, 201, 5-21.

Warner, N.R., and E.W. Domack,( 2002), Millennial- to decadal-scale paleoenvironmental change during the Holocene in the Palmer Deep, Antarctica, as recorded by particle size analysis, Paleoceanography, 17, Art. No. 8004, doi:10.1029/2000PA000602.

8 CHAPTER 2 LATE PLEISTOCENE AND HOLOCENE SEDIMENTATION, ORGANIC- CARBON DELIVERY, AND PALEOCLIMATIC INFERENCES ON THE CONTINENTAL SLOPE OF THE NORTHERN PANDORA TROUGH, GULF OF PAPUA*

2.1. Introduction

Rivers deliver large amounts of organic carbon (~0.4x1015 g C yr-1) from land to

continental margins, with approximately 30% of it originating from tropical rainforests alone [Schlesinger and Melack, 1981; Hedges et al., 1997]. Understanding the ultimate fate of this important carbon pool is crucial not only for understanding climatic changes, but also for accurately modeling the broader global carbon cycle [Schlünz and Schneider,

2000; Berner, 2003].

Within this context, the Gulf of Papua (GoP) (Figure 2.1) is a particularly

intriguing region to examine the flux and fate of terrestrial organic carbon over glacial to

interglacial timescales. The GoP is a marine depocenter for large and very active river

basins. In the GoP, three large rivers, the Fly, Kikori and Purari, currently deliver an

estimated 300-400 megatonnes yr-1 (Mt) of terrigenous siliciclastic material (e.g., clay, quartz, feldspars) from a young, mountainous and tropical region to the inner shelf each

year [Milliman, 1995]. This is an annual flux greater than that of the Mississippi River,

and comes from drainage basins with combined areas ~3% the size of the Mississippi

drainage basin [Milliman and Syvitski, 1992; Milliman, 1995]. With this immense siliciclastic load comes approximately 4 Mt yr-1 of terrestrial particulate organic carbon

[Bird et al., 1995]. Today, most of the fluvial discharge accumulates on the inner shelf of

______

* Submitted and accepted for publication in Journal of Geophysical Research – Earth Surface, Reproduced by permission of American Geophysical Union, see also Appendix D.

9 the GoP and in an extensive mangrove-covered deltaic system along the northern coast

(Figure 2.1); very little of the siliciclastic material and terrestrial organic carbon escapes

the shelf to deeper water [Bird et al., 1995; Walsh and Nittrouer, 2003].

The modern shelf edge of the GoP lies about 125 m below sea level (Figure 2.1).

However, during late Quaternary sea-level lowstands, water depths were ~60 to 125 m below those at present. Consequently, river mouths were at or near the shelf edge, and

large amounts of siliciclastic material and terrestrial organic carbon presumably

accumulated in deep water settings, as is known for other large river systems [e.g., Flood

and Piper, 1997, and references therein]. However, there is no information to evaluate whether massive terrestrial input indeed occurred in GoP slope and basin environments during glacial times, and, if so, what was the fate of that material with respect to burial, transformation, or remineralization.

This study aims to evaluate sediment accumulation in the northeastern GoP since

the last glacially induced sea-level lowstand. Specifically, I investigated: 1) the sediment

accumulation rates at two representative locations on the middle slope, 2) the late

Pleistocene to Holocene organic carbon contents and fluxes at these locations, and 3) the

proportions of terrestrial and marine organic carbon.

2.2. Gulf of Papua

The GoP is a tropical mixed siliciclastic-carbonate sedimentary system that is

presently affected by a warm, wet monsoon-dominated tropical climate and a young,

mountainous continental terrain. The dry season occurs during April to November when

SE trade winds dominate, setting up moderate wave fields (~1.3 m significant wave

heights) and invigorating the clockwise Coral Sea Coastal Current (Figure 2.1). In

10 contrast, the wet season occurs during December to March and is characterized by milder

NW monsoonal winds that bring significant amounts of rain and lower (~0.3 m)

significant wave heights [Thom and Wright, 1983].

Figure 2.1. General bathymetric map of the Gulf of Papua study area. Bathymetric contour interval is 100 m down to the 500-m contour line, and then every 500 m in deeper water. The 60-m contour line on the middle shelf indicates the extent of the modern shelf clinoform. The dashed arrow indicates the general direction of sediment movement from the Fly River. The bathymetric profiles A and B are plotted below the map and show the gradient from the coast to each core station - note the difference in horizontal scale. Bathymetric data set is from Daniell [2007].

11 The Papua receive 10-13 m of annual rainfall that ultimately drains into the GoP [Wolanski et al., 1984]. Coarse sediments discharged by

the Fly, Kikori, and Purari Rivers initially accumulate in deltaic sand bodies, whereas

suspended sediments are mostly advected clockwise to the northeast by geostrophic flow

[Wolanski et al., 1995; Wolanski and Alongi, 1995] and stored along the inner shelf in

waters generally shallower than 60 m [Bird et al., 1995; Brunskill et al., 1995; Harris

1996; Walsh et al., 2004; Keen et al., 2006]. Further remobilization may occur during the

wave-dominated trade-wind season [Hemer et al., 2004].

2.2.1. Pandora Trough

The Pandora Trough is a broad basin beyond the shelf edge covering >8000 km2

and is thought to receive a limited modern terrigenous sediment flux [Milliman et al.,

1999; Walsh and Nittrouer, 2003]. The Pandora Trough is rimmed by the GoP shelf that changes from broad and low-gradient (1:1000) in the northwest to a narrow and higher gradient (1:133) in the northeast (Figure 2.1). Sediments escape into the northern

Pandora Trough by nepheloid-layer transport down slope [Walsh and Nittrouer, 2003] and episodic turbidity currents in some locations [Bentley et al., 2006]. An estimated 2-

3% of the sediments entering the GoP each year may be transported off the shelf and deposited in the Pandora Trough [Walsh and Nittrouer, 2003; Muhammad et al., 2007].

2.3. Materials and Methods

2.3.1. PANASH Cruise and Sediment Cores

The Pandora and Ashmore Troughs are two large sediment sinks that ultimately

store sediments shed from the PNG highlands and coastal plain. The PANASH cruise,

which took place from March to April 2004, was part of the NSF MARGINS Source-to-

12 Sink Program to sample and quantitatively study this important sedimentary system.

During the 2004 field season aboard the R/V Melville, we collected a total of 30 multi- cores and 33 jumbo piston cores.

Piston core MV-54 was collected from the slope seaward of the broad shelf (~150 km wide) bordering the central Pandora Trough; piston core MV-51 was collected from the narrow shelf (~20 km wide) along the northern Pandora Trough (Figure 2.1). We selected cores on the basis of achieving comparable geologic setting, core length, and water depths to compare and contrast sediment accumulation in different regions of the

Pandora Trough. Cores MV-54 and MV-51 are both ~12 m in length and were obtained from 923 and 804 meters water depth, respectively, in open slope, non-channelized locations that are not likely to be inundated directly by turbidity currents (based on our multibeam surveys) [Francis et al., 2007] (Figure 2.1 and Table 2.1). MV-54 is located on the central Pandora mid-slope region (Figure 2.1) just NE of slump scars and between unfilled, relict channels as noted in the work of Francis et al., [2007, Figure 8]. MV-51 was collected on top of a NW-SE trending tectonic ridge that forms a local topographic high in the mid-slope [Francis et al., 2007, Figure 9]. The numerous tectonic ridges in the northeastern Pandora Trough slope also have troughs between them that serve to trap or divert sediments moving downslope. The ridges are probably sufficiently elevated above troughs to avoid direct inundation by gravity flows (~100 m) (Figure 1, bathymetric profile B), yet still receive suspended sediment transported off-shelf or down-slope. Consequently, MV-51 potentially holds a high-quality, high-resolution time-stratigraphic record.

13 Interpretations presented in this chapter apply specifically to the core locations

under investigation. However, the geologic settings of these cores suggest that the cores

should contain records of regional processes 1) because of the placement of MV-54 away

from surrounding channels and slump scars, 2) because MV-51 is positioned on top of a

tectonic ridge, and 3) because both are relatively close to river-influenced shelf edges.

Table 2.1. Geographic locations for sediment cores examined in this study.

Sediment Core Latitude Longitude Water Depth (m) Length (m) MV25-0403-54 8.936° S 145.389° E 923 12.16 MV 25-0403-51 8.767° S 146.018° E 804 12.28

2.3.2. Shipboard Analyses

After core retrieval, shipboard measurements were made on piston cores for bulk

density and magnetic susceptibility using a GeoTek Multi Sensor Core Logger (MSCL).

Bulk density was calculated from the attenuation of gamma rays emitted from a 10 milli-

Curie 137Cs source through sediment cores. GeoTek MSCL-derived bulk density units are g cm-3 and referred to as gamma-ray density (GRD) to separate them from traditional

wet and dry bulk density measurements [Weber et al., 1997]. Magnetic susceptibility

(MS) was measured using a Bartington MS2C Loop Sensor attached to the MSCL. MS

values were corrected with respect to volume and units are reported in 10-5 SI [Thompson

and Oldfield, 1986]. Both MS and GRD were measured on the MSCL at 1-cm intervals.

• X-radiographic Methods

Selected sections of MV-51 were imaged via X-radiography. Core subsections

were sliced into 2-cm-thick axial slabs, and imaged onboard using a Thales Flashscan 35

digital X-ray detector panel, illuminated with a Medison Acoma PX15HF X-ray

14 generator. Images were archived as 14-bit grayscale images with 127-micron pixel resolution.

2.3.3. Radiocarbon Analyses

Radiocarbon dates are based primarily on woody organic matter because of insufficient absolute concentrations of planktonic foraminifera. Nine wood samples from cores MV-54 (n=4) and MV-51 (n=5) were analyzed for their radiocarbon content at three labs (Table 2.2) using accelerator mass spectrometery (AMS). Bulk sediment samples were washed on a 63 µm sieve with distilled water and then at least 10 mg of woody organic matter were hand picked from each for analysis. Only large wood particles (>250 µm) that clearly demonstrated fibrous wood texture were selected.

Because wood and other organic matter may be reworked, particularly during periods of lower sea level, radiocarbon dates taken from wood may represent maximum ages for sediments.

Benthic foraminifera from two samples in MV-51 (Table 2.2) also were taken for radiocarbon analysis. These ages were reservoir corrected by subtracting the average age difference between planktonic and benthic foraminifera in the western Pacific Ocean

[1483 yrs; Broecker et al., 2004]. Benthic foraminifera were analyzed because commonly used planktonic foramnifera (e.g., Globigerinoides ruber, G. sacculifer) were not sufficiently abundant.

For six samples, reported radiocarbon ages were corrected for past variations in cosmogenic production using OxCal v 3.1 with the IntCal98 correction curve for dates

≤20,000 14C yrs B.P. [Stuiver et al., 1998]. Five of seven dates from MV-51 remain

“uncorrected” because internationally ratified calibration curves do not extend beyond

15 20,000 14C yrs B.P. [Stuiver et al., 1998]. Calendar ages for these samples may be as

much as 4000 to 6000 yrs older between 20,000-35,000 14C yrs B.P as a result of increased 14C production during changes in the Earth’s magnetic field intensity [Hughen et al., 2004; Fairbanks et al., 2005].

2.3.4. Magnetic Susceptibility

MS is a measure of the strength of induced magnetism experienced by mineral

grains in sediments when placed in an applied field. Low-field MS was independently

measured on piston-core samples at 5-20 cm intervals using a susceptibility bridge in a

magnetically shielded room at the LSU Rock Magnetism Laboratory. The MS Bridge

uses a balanced coil induction system and is sensitive to 1 (±0.2) x 10-9 m3 kg-1 [Ellwood

et al., 1996]. This method permits an independent cross check on data collected from the

MSCL.

MS values are excellent indicators of the total iron-containing compounds in

sediments [Nagata, 1961] and therefore MS is primarily a function of hinterland

mineralogy and climate. Terrigenous magnetic components originate from multiple

mineralogies including: 1) ferrimagnetic (such as magnetite and maghemite), 2)

paramagnetic sources such as clays (e.g., chlorite and illite), iron sulfides (e.g., pyrite), and 3) ferromagnesian silicates (such as biotite, amphibole, and pyroxene). Weakly negative diamagnetic components in sediments include calcite, quartz, and organic matter. Diamagnetic contributions may reduce the overall sediment MS because of a dilution effect, although this is generally relatively small, given the susceptibility of even small amounts of detrital components [Ellwood et al., 2000].

16 2.3.5. Calcium Carbonate

Calcium-carbonate content (Figures 2.3, 2.4) was analyzed in MV-51 and 54 using the vacuum-gasometric technique [Jones and Kaiteris, 1983] at the LSU Rock

Magnetism Laboratory. Dried sediment samples were ground to a fine powder and then a

0.25 g subsample was reacted under a vacuum in a reaction vessel with phosphoric acid for one hour. After complete reaction, the pressure difference was then read from a pressure gauge and used to calculate the sample carbonate concentration using a regression equation. Triplicate carbonate measurements on several samples resulted in a precision of +/- 0.10%. Tests of reagent grade (99%) carbonate and a mixed sample standard (95% feldspar 5% carbonate) gave accurate results to +/- 1%.

2.3.6. Sedimentary Organic Matter

The quantity and type of sedimentary organic carbon offer major evidence for

reconstructing paleoclimatic and paleoceanographic change. Organic matter in core

sediments was measured using three independent analyses. Total organic carbon (TOC)

was measured on core sediments using a LECO Carbon analyzer at 20 cm sample

intervals for both MV-51 and MV-54. Rock-Eval Pyrolysis (REP) was measured on

sediments from MV-51, because this second core contained the longest time-stratigraphic

record. TOC and REP measurements were made at Baseline Resolution Analytical Labs,

Houston Texas, with analytical errors of ± 0.05 wt. % and ± 0.02 mgHC/gTOC, respectively. TOC data quantify weight percent of organic carbon [Jarvie, 1991] and may serve as an initial proxy for productivity [Berger and Herguera, 1992; Rühlemann et

al., 1999]. Mass accumulation rates (MAR) for TOC were then calculated using the

accumulation rates in both piston cores:

17

= × × LSRDBDTOCMARTOC g g cm g (1) MARTOC =××= g cm3 kyr 2 ⋅kyrcm

where: C org g org C .%)( wtTOC .%)( ×= 100 sed g g sed DBD= Dry Bulk (gDensity 3- )cm LSR = Linear Sedimentat Rateion (cm 1- ).kyr

REP helps to determine the principle sources of the TOC by examining the amount of free (S1) and pyrolytic (S2) hydrocarbons preserved in sediments. REP was developed as a screening method to determine the hydrocarbon potential of petroleum source rocks [Tissot and Welte, 1984; Rullkötter, 2000], but it is becoming widely used in paleoceanography [Lallier-Vergès et al., 1993; Meyers, 1997; Bouloubassi et al., 1999;

Rullkötter, 2000; Wagner, 2000]. Two REP parameters are used in this study, the hydrogen index (HI) and the thermal maximum (Tmax). The hydrogen index is the mass of hydrocarbons per gram of TOC and is determined from heating kerogen in a sample until the maximum amounts of hydrocarbons are evolved [Espitalié and Bordenave,

1993]. Equation 2 was used to calculate the hydrogen index. Tmax is the temperature during pyrolysis at which the most hydrocarbons are evolved.

× S2100 mg 100g mgHC HI = =×= (2) TOC g g gTOC

where : HI = Hydrogen Index ⎛ mgHC ⎞ S2 = ⎜ ⎟ ⎝ g sediment ⎠

18 Hydrogen index values are generally low (≤150 mgHC gTOC-1) for organic

matter composed of vascular-land-plant carbon in recent deep-sea sediments, and are

characterized as Type 3 and Type 4 kerogen [Tissot and Welte, 1984; Meyers, 1997].

Algal-rich marine organic matter values range from 200 to 400 mgHC gTOC-1 for recent

marine sediments and are characterized as Type 2a kerogen [Rullkötter, 2000]. Tmax

values of modern “immature” organic carbon are below 430°C for kerogen types 3 and 4,

whereas “overmature” organic carbon from reworked fossil organic matter are >430°C

[Delvaux et al., 1990].

More precise determination of sedimentary organic-carbon types was achieved by

analyzing the C/N and δ13C ratios on selected samples from MV-51. Isotopic analyses

were performed on a Thermo Finnigan Delta Plus XP Isotope Ratio Mass Spectrometer

with an online elemental analyzer at the Coastal Ecology Institute, LSU. Dry sediment

samples were first acidified with 10% HCl until the reaction with calcium carbonate

ceased, or approximately 10 minutes. Sediments were then acidified with 50 ml

concentrated HF for eight hours to remove minerogenic matter and isolate the kerogen.

Residues were centrifuged and rinsed with distilled water several times until completely neutralized, then dried at 50°C and powdered for analysis. We also made slides from this isolated kerogen for later optical palynofacies analysis. Standard deviations of duplicate measurements show analytical precision of δ13C to be ±~0.15‰ and C/N ratios to ±~0.20

molar percent.

The terrigenous component of TOC was determined from δ13C values using a

two-source mixing model in which the marine (-20.5‰ ±0.5‰) and terrigenous (-26.5‰

±0.5‰) isotope values are used as end members [Bird et al., 1995; Aller and Blair,

19 2004]. Equations 3-5 are adapted from Amo and Minagawa [2003] and Bird et al.,

[1995].

13 13 (δ Cmeasured −δ Cmarine) f sterrigenou = (3) 13 13 (δ C lterrestria −δ Cmarine)

13 13 ⎡ δ Cmeasured −δ Cmarine ⎤ % = fTOClTerrestria TOC =× ( )× × .%100 TOCwt (4) sterrigenou ⎢ 13 13 ⎥ ⎣⎢()δ C lterrestria −δ Cmarine ⎦⎥

2.3.7. Benthic Foraminifera

Benthic foraminiferal assemblages are very useful paleoproductivity proxies because their abundances and assemblage compositions are linked to the flux of organic carbon arriving at the seabed. Several studies have demonstrated that foraminiferal densities are positively correlated with increases in nutrient fluxes, surface-water productivity, and the subsequent flux of that material to the seabed [e.g. Gooday, 1988;

Jorissen et al., 1992; Sjoerdsma and van der Zwann, 1992; Loubere, 1996]. Certain assemblages of benthic foraminifera are adapted to particular types of organic-carbon flux, such as highly variable or “seasonal” flux that is charactersistic of upwelling zones compared to relatively constant flux [Loubere and Farridudin, 1999; Smart and Gooday,

1997; Licari and Mackensen, 2005]. We use benthic foraminifera as an independent metric of organic-carbon flux, as well as the stability of flux through time.

In this study, eight to ten grams of 36 freeze-dried samples from core MV-51

were gently washed with distilled water over a 63 µm sieve and dried at 50°C. The >63

µm fraction was then weighed and further split into 150-250 µm and >250 µm fractions.

At least 300 benthic foraminifers were then picked from each fraction >150 µm, so that

20 ~600 specimens were counted from each sample. Specimens were placed on assemblage

slides, sorted, and then enumerated. Assemblage data are routinely collected from the

>150 µm fraction in ecological and paleoceanographic studies [e.g. Pérez et al., 2001]

because ecologically important taxa may be missed by examining only the >250 µm

fraction [Sen Gupta et al., 1987]. Examination of these fractions also permits comparison

with other studies. The >63 µm fraction was inspected for all sieved samples in MV-54

to determine presence or absence of foraminifera.

Benthic foraminiferal densities (BF# = total shells or specimens g-1) and benthic

foraminiferal accumulation rates (BFAR; # cm-2 kyr-1) were calculated from absolute foraminiferal densities [Herguera, 1992; Herguera and Berger, 1994]. Groupings of diagnostic species were examined with R-mode cluster analysis (Minitab v. 14.20) using a 1-Pearson correlation coefficient distance metric and an average linkage method.

Cluster analysis is based on the absolute abundance data (BF#, total specimens g-1) in 36 samples for the most abundant, recurrent, and diagnostic taxa, totaling 12 species. Q- mode cluster analysis was used to examine the stratigraphic clustering of samples based on BFAR data for the same 12 taxa. Taken together, these quantitative data are used to infer relative paleoproductivity during the past ~33,000 years in the northern Pandora

Trough.

2.3.8. Thermomagnetic Measurements

Thermomagnetic susceptibility measurements were performed on a number of

samples from MV-51 using the KLY 3S Kappa Bridge manufactured by AGICO, Czech

Republic. Results are given in SI and are dimensionless. This measurement involves

21 heating the sample from room temperature to 700oC while measuring the MS as temperatures rise.

Paramagnetic minerals show a parabolic-shaped MS decay during this process,

because the MS in these samples is inversely proportional to temperature of measurement

[Hrouda, 1994]. Based on measured standards, ferrimagnetic minerals usually show an

increase in MS up to a point where the MS decays toward the Curie temperature of the

minerals responsible for the MS. For magnetite, this temperature is ~580o C and for

maghemite it is ~610o C for most samples. Above 640oC, maghemite converts to

hematite with a Curie point of ~680oC. During measurement, some minerals that contain

iron and exhibit paramagnetic behavior at low temperatures, are unstable at higher

temperatures and chemically change during heating. This often produces diagnostic peaks

that result from the formation of a ferrimagnetic phase during this change [Ellwood et al.,

2007].

2.4. Results

2.4.1. Core Sediments

Sediments in MV-51 and MV-54 are composed of homogenous marine mud, and

contain few discernible sedimentary structures. MV-51 contains four ~5-cm thick

intervals at 1.6, 2.4, and 4.8 meters below sea floor (mbsf) that contain well preserved

biogenic sedimentary structures associated with diffuse ash layers. Another discrete,

partially bioturbated ash layer occurs at 6.9 mbsf. X-radiographs of MV-51 over the 3-

4.5 mbsf interval revealed homogeneous sediments, with no obvious bedding. MV-54

contains two coarse-grained, fining-upward, organic rich intervals at 1.5-1.9 mbsf and

22 6.1-6.3 mbsf. Sediments in the interval 6.1-6.3 mbsf are inclined by ~45°, suggesting

deposition by mass flow.

2.4.2. Geochronology and Accumulation Rates

Eleven radiocarbon dates illustrate a late transgression to Holocene section <2 m

thick, very rapid linear sedimentation rates (LSR), and high MAR for both cores during times older than ~15,000 Cal. yrs B.P. (Figure 2.2). Though we do not have radiocarbon dates or 210Pb data in the uppermost sediment of MV-51 and 54, we assume a recent age

for the core tops. Radiocarbon dates for MV-54 are generally all within the range of

error for each other, particularly when corrected to calendar years (Figures 2.2 and 2.3;

Table 2.2). The date at 1.9 mbsf in MV-54 may be an exception. We considered the dates of 17,000 Cal. yrs B.P. at 4.75 mbsf and 8.3 mbsf to be indistinguishable.

Figure 2.2. Time-Depth plot of radiocarbon ages. Arrows point to radiocarbon ages determined from mixed in situ benthic foraminifera. Numbers along profiles are sediment accumulation rates in m kyr-1. Ages and error estimates are listed in Table 2.2.

23 Two sections in MV-51, 1.40-3.50 and 10-12 mbsf, show high LSR and MAR

(Figure 2.3). The first section (~18,400-20,400 Cal. yrs B.P.) exhibits LSR of ~1.09 m

kyr-1 and ~81.2 g cm-2 kyr-1, and the LSR in the second section (>32,000 14C yrs B.P.) is

twice as much at ~2.70 m kyr-1 and ~244 g cm-2 kyr-1. Two sets of radiocarbon ages are

within the range of error for each other, 1.20 and 1.34 mbsf, and 10 and 12 mbsf (Figure

2.2, Table 2.2).

Table 2.2. Radiocarbon ages reported in this paper. Lab #: 1 Keck CCAMS Facility, UC Irvine, 2 Beta Analytic, 3 National Ocean Sciences AMS at WHOI. *Radiocarbon ages are calibrated to calendar ages (2 sigma 95% probability) using OxCal v. 3.10 [Bronk Ramsey, 1995] and the IntCal 98 calibration curve [Stuiver et al., 1998]. ** Corrected by subtracting age difference between planktic and benthic foraminifera in the Western Pacific [Average age difference: 1483 yrs; Broecker, 2004]. Depth (m) Lab Material 14C Age Error (yrs) Median Calendar Age, Range (± yrs) B.P.* MV-51 1.20 1 wood 15 565 ± 50 18 600 650 MV-51 1.34 2 wood 15 390 ± 100 18 405 505 MV-51 3.51 1 wood 20 400 ± 90 20 405 185 MV-51 7.40 3 Benthic Foraminifera 29 500 ± 220 28 017** 220 MV-51 7.94 1 wood 29 940 ± 180 out of range 180 3 Benthic 34 100 ± 250 250 MV-51 10.00 32 617** Foraminifera MV-51 12.00 1 wood 33 360 ± 320 out of range 320 MV-54 1.90 1 wood 13 320 ± 25 15 850 700 MV-54 4.75 1 wood 14 160 ± 35 17 000 500 MV-54 8.30 1 wood 14 175 ± 40 17 000 500 MV-54 9.94 2 wood 14 810 ± 80 17 740 440

The time intervals outside of high accumulation rate sections in MV-51 show 1) a

very thin late transgression to Holocene section (~1 m) and 2) a thick or expanded section

(~4 m) between ~20,400 and 29,000 14C yrs B.P., during the earliest to middle part of the

Marine-Isotope-Stage-2 glaciation that includes the LGM time of 23,000–19,000 Cal. yrs

B.P.. Sediments were deposited with very low LSR (0.07 m kyr-1) and MAR (~5 g cm-2 kyr-1) since late lowstand time (<~18,400 Cal. yrs B.P.). Intermediate LSR of 0.51 m

24 kyr-1 and moderate MAR of 94 g cm-2 kyr-1 occurs between 20,400 and 32,000 14C yrs. A short duration decrease in LSR (0.28 m kyr-1) and MAR (24 g cm-2 kyr-1) occurs between

28,000-29,900 14C yrs (Figures 2.2 and 2.3).

Two distinct intervals of LSR and MAR occur in MV-54 (Figure 2.4). The first

interval is from 0-1.9 mbsf (0-15,850 Cal. yrs B.P.) and is characterized by LSR of 0.11

m kyr-1 and MAR of 9.7 g cm-2 kyr-1. The second interval is from 1.9-9.4 mbsf and

contains the highest LSR of 3.96 m kyr-1 and MAR of 428 g cm-2 kyr-1.

2.4.3. Magnetic Susceptibility

Trends in data collected from the MS bridge at Louisiana State University are generally in good agreement with MSCL Bartington Loop measurements (R=0.81, p<<0.005; n=213; Figures 2.3 and 2.4). Values from the MSCL are relatively low for both cores (<40x10-5 SI) while MS bridge values are higher and show averages between

1.20 x 10-7 m3 kg-1 (MV-51) and 1.98 x 10-7 m3 kg-1 (MV-54). High frequency MS

fluctuations are evident in MV-51, particularly between 5.5-7.5 mbsf that are not present

in the Bartington Loop data alone (Figure 2.3). This is due to the diminished sensitivity

of the Bartington Loop Sensor at low MS values. Results presented below are limited to

detailing the more sensitive MS bridge data.

In MV-51, sediments less than 10 mbsf, or younger than 32,000 14C yrs B.P., have

low MS values, with some notable exceptions. MS maxima occur at 2.40, 4.80, and 6.94

mbsf (mean 3.56 x 10-7 m3 kg-1, n=3; Figure 2.3) and are nearly four times higher than

background levels over the interval from 0-10 mbsf (mean 9.35 x 10-7 m3 kg-1, n=51).

Higher MS values are present in MV-51 between 1.15 and 1.55 mbsf, depths

corresponding to approximately 18,400 Cal. yrs B.P.

25 A major shift to higher MS values by two to three orders of magnitude in MV-51

occurs below 10 m (>~32,000 14C yrs B.P.) and indicates enhanced input of detrital

minerals during that time. The magnetic detrital fraction in these sediments is restricted

to mud-sized particles because the sand-sized fraction (>63 µm) comprises <1.5% of sediments below 10 mbsf (Figure 2.3), and those coarse fractions are composed entirely

of foraminifera.

MS values are much more variable in MV-54 (Figure 2.4). Sediments from 0 to

1.9 and 9 to 12 mbsf have high MS values (mean 2.7 x 10-7 m3 kg-1, n=31), or

approximately <15,800 and 17,700 Cal. yrs B.P., respectively. A more variable interval

occurs from 2 to 9 mbsf (mean 1.33 x 10-7 m3 kg-1, n=35) and contains two maxima at

3.40 and 6.20 mbsf with values more similar to the upper and lower parts of the core.

The high values from ~1.5 – 1.9 mbsf are associated with dark, muddy sand and at 6.2

mbsf with dark, sandy, laminated and contorted sediment. The sand-sized fractions (>63

µm) comprise as much as 85% of the sediment from 1.5 – 1.9 mbsf and 25% at 6.20 mbsf

(Figure 2.4). Sediments are generally coarser in MV-54 (mean 15% >63 µm) than in

MV-51 (mean 2%).

2.4.4. Calcium Carbonate and Volcanic Ash Layers

Siliciclastic particles dominate sediments from cores MV-51 and MV-54.

Samples in MV-51 have very low calcium carbonate percentages (mean 3.8%, n=68)

whereas values are slightly lower in MV-54 (mean 3.0%, n=37). Calcium carbonate in

MV-54 occurs only in the upper 3.6 m and is completely absent below 4.0 mbsf (Figure

2.4).

26

27

28 Several volcanic ash layers are present in MV-51. MS maxima at 2.4, 4.8, and

6.9 mbsf are a result of elevated levels of paramagnetic grains such as biotite,

hornblende, and amphibole that were identified from sieved samples (>63 µm) using a

binocular microscope. These grains are accompanied by abundant pumice and glass

shards and result in sharp increases in the sand-sized fraction (>63 µm, Figures 2.3, 2.7),

suggesting a sudden input by volcanic origin. TOC and calcium-carbonate values decline

at the same levels, probably due to particle dilution, but they may also reflect a decline in biological productivity during deposition of ashfall. Another pumice-rich layer is present at 1.0 mbsf, but it does not correspond to an MS maximum because it lacks abundant biotite and other iron-rich paramagnetic minerals. Additionally, the pumice-rich layer at

1.0 mbsf does not show a significant decrease in either TOC or calcium carbonate contents.

2.4.4. Sedimentary Organic Matter

Organic carbon content in MV-51 and MV-54 is relatively low overall and at site

MV-51, appears largely to have a source from land-plant debris. TOC in core MV-51 ranges from 0.41 to 0.97 wt. % (mean 0.75), generally lower than values in core MV-54, which range from 0.63 to 2.29 wt. % (mean 0.92). In MV-51, TOC is highest from 1.4 to

1.8 mbsf and >7.4 mbsf, or during 18,400-20,400 Cal. yrs B.P. and >28,000 14C yrs B.P.,

respectively. MAR of TOC is also highest during these time intervals (Figure 2.3).

MV-54 shows nearly uniform values in TOC over most of the core length. Four

prominent subsurface maxima in excess of 1.3 wt. % occur at 1.8, 3.2, 6.2, and 7.0 mbsf

and correspond to maxima in both MS and the sand-sized fraction (Figure 2.4).

Sediments from 1.5-1.9 mbsf and at 6.2 mbsf are composed of a dark, sandy material,

29 with abundant and well preserved pteropods, foraminifera, and woody material. Organic

particles are common in all sieved samples from MV-54, many of which are unpyritized

and can be identified as plant cuticle and woody cortex under binocular microscopy. One

sample from 7.0 mbsf contained tree resin (amber) and a complete leaf (~2mm long) with

well preserved venation.

Hydrogen index values of sedimentary organic matter from MV-51 are very low

(7 to 152.6 mgHC gTOC-1; average 74) and contain very low S2 pyrolytic yields (Figure

2.5a). These data reflect hydrogen-poor organic carbon and therefore may indicate either

type 3 organic matter or a mixture of type 3 and oxidized type 2 marine organic matter.

Figures 5a and 5b show the data plotted in two different diagrams, both generally plotting

below the Type 3 terrestrial organic carbon thresholds. However, low S2 values, below 1

mgHC g sediment-1, may be a consequence of mineral matrix dilution of hydrogen-poor

hydrocarbons which makes it difficult to accurately determine the organic matter type

[Espitalié et al., 1980; Katz, 1983; Langford and Blanc-Valleron, 1990]. No correlation

exists between hydrogen index and TOC values, as expected given the very low hydrogen

index values and narrow range in TOC values [Bouloubassi et al., 1999].

Tmax measurements generally indicate that MV-51 sediments contain immature organic carbon, as expected for recent marine sediments. However, six samples from 3.0,

3.2, 4.6, 5.2, 8.8, 9.8 mbsf (Figure 2.5b) indicate overmature organic carbon, which probably had a source from reworked sedimentary rocks inland (e.g., coal). There is no stratigraphic significance to the timing of reworked organic matter in MV-51.

30

Figure 2.5. Rock-Eval 2 data from MV-51 plotted in an S2 versus TOC diagram (5a, after Langford and Blanc-Valleron, 1990) and a Van Krevelen diagram (5b, after Wagner, 2000). Boundaries in the Van Krevelen diagram between kerogen type 2, 3, and 4 are plotted with dashed lines and derived from reference values reported in Delvaux et al. [1990]. Hydrogen index values for MV-51 plot in the vascular land plant type 3 to type 4 kerogen window, with six samples representing reworked fossil organic matter (3.0, 3.2, 4.6, 5.2, 8.8, and 9.8 meters below sea floor, labeled on plot).

31

32 The percentages of terrestrial components calculated from stable isotopic data

exhibit nearly parallel changes with TOC, with maximum values (>40%) between 1.4-1.8

mbsf and deeper than 7.4 mbsf (Figure 2.6). Three terrestrial input maxima occur in

MV-51, one at 1.80 mbsf (62%), another at 8.08 mbsf (55%), and the third at 10 mbsf

(57%). A strong positive correlation exists between TOC concentrations and percent terrigenous TOC (R=0.85, p<0.001; n=24), indicating that increases in TOC concentrations result from enhanced input of continental sources and not from marine primary productivity.

Figure 2.7. Organic geochemical data from MV-51 are plotted in an elemental (atomic C/N) and stable-isotopic data cross plot. Data (empty circles) generally plot along a theoretical mixing line (dashed line) between two sources, modern marine organic matter and C3 land plants. OM refers to organic matter and SOM refers to soil organic matter. Plot after Meyers [1997] and sources mapped from values reported in Meyers [1994] and Goni et al., [2006].

Elemental and stable-isotope data provide further constraint on organic-carbon sources. A plot of elemental (atomic C/N) and stable isotopic (δ13C, ‰) values from

33 MV-51 illustrates two groupings of data (Figure 2.7). The first group (0.4-6.8 mbsf)

plots closer to modern marine organic matter values than does the second group (core-

top, and >7.4 mbsf). This second group (core-top, and >7.4 mbsf) clusters nearest to the

region of C3 land plants. Atomic C/N and stable-isotopic data generally fall on a mixing

line between the two sources, except for one point from an ash layer at 6.9 mbsf that rests

well below the mixing line (Figure 2.7). The ash layer forms an outlier, possible as a

result of a different organic matter source such as woody debris transported in pumice

rafts.

2.4.5. Benthic Foraminifera

Benthic foraminifera are abundant in MV-51 sediments. Absolute densities and

BFAR (Figure 2.8) generally give the same trends but with some differences in the lower

seven meters. Lowest densities of foraminifera (mean ~98 specimens g-1, n=22) occur

shallower than 6.8 mbsf, whereas densities between 6.8 and 12 mbsf are nearly two times

higher (mean ~174 specimens g-1, n=14; Figure 2.8). A sample at 6.8 mbsf contains the

highest density of foraminifera in MV-51 (~283 specimens g-1). In contrast, lower BFAR

(mean ~5800 cm-2 kyr-1, n=31) characterize core depths less than 10 mbsf, except for the

maximum at 6.8 mbsf (12,600 cm-2 kyr-1). Deeper than 10 mbsf, benthic foraminiferal densities are six times higher than in samples less than 10 mbsf (mean ~34,800 cm-2 kyr-1, n=5; Figure 2.8). The significance of benthic foraminiferal densities, BFAR, and organic carbon flux at site MV-51 (Figure 2.8) was examined with multiple linear regression analysis.

34

35 A statistically significant association exists between total benthic foraminiferal

densities >150 µm and TOC concentrations (R=0.46; p=0.003; n=35). However, the low

correlation coefficient suggests that while TOC plays a role in the densities of

foraminifera, other factors are important as well. If the mass accumulation rates are

incorporated into the raw data so that TOC MAR and BFAR are calculated, a much

stronger correlation of analytical data exists (R=0.95; p<0.001; n=35). These statistics

emphasize the importance of incorporating sediment accumulation rates with raw

foraminiferal data [Herguera, 1992]. Together, they show that the mass accumulation of

organic-carbon and foraminifera are intrinsically linked, and therefore that BFAR is a

reflection of organic-carbon flux.

Two periods of high BFAR and high TOC MAR occur in MV-51 (Figure 2.8).

The first interval is 1.4-3.4 mbsf and corresponds to ~18,400-20,400 Cal. yrs B.P., or middle to late lowstand time. The second interval occurs below 10 mbsf in sediments greater than ~32,000 14C yrs B.P., or approximately the late Stage 3 interstadial. The

latter period contains the highest BFAR and TOC MAR in the entire core. In addition to

total BFAR, important paleoecological information is contained in the accumulation rates

for groups of benthic foraminiferal species.

R-mode cluster analysis divided benthic foraminiferal species into two distinct

groups (Figure 2.9). The first cluster is characterized by low-abundance taxa including

Oridorsalis tener and Melonis barleeanus. Species of this cluster rarely exceed 1-5

specimens per gram or 150 specimens cm-2 kyr-1. The second cluster is composed of ten

taxa (Figure 2.9) with three distinct subclusters being the most abundant and

paleoceanographically significant. The first subcluster (A) is composed of Uvigerina

36 peregrina/hispida, Cibicidoides pachyderma, and Bolivina robusta (Figure 2.9). The

second subcluster (B) is composed of Bulimina aculeata, Sphaeroidina bulloides, and

Bolivinita quadrilatera as a companion taxon. The more distant third subcluster (C) contains Gavelinopsis translucens and Globocassidulina subglobosa. Species distributions in Figure 2.9 are arranged in order of the clusters described above.

Figure 2.9. Cluster analysis of absolute densities (specimens g-1) for the twelve most abundant species in MV-51 sediments. Analysis used is a 1-Pearson distance metric and average-linkage clustering, n=36 samples. Major clusters are numbered and smaller subclusters within cluster 2 are identified by alphabetic letters.

Three distinct species BFAR intervals are present in MV-51 (Figure 2.8, shaded

areas). The first interval (10-12 mbsf; >32,000-33,000 14C yrs B.P.) is characterized by

elevated abundances of nearly all foraminiferal species with the exception of M.

barleeanus (Figure 2.8). The most abundant taxa are, in order of decreasing abundance,

B. robusta, U. peregrina/hispida, S. bulloides, Bul. aculeata, C. pachyderma, and O.

37 tener. The second interval is composed of a single sample at 6.8 mbsf (<28,000 14C yrs

B.P.) and is composed chiefly of B. robusta, U. peregrina/hispida, and C. pachyderma.

The third abundance interval occurs from 1.4-3.4 mbsf (~18,400-20,400 Cal. yrs B.P.) and is composed mostly of B. robusta, G. translucens, Gc. subglobosa, O. tener, and M.

barleeanus. Overall, the most abundant species throughout MV-51 is B. robusta.

Large sections of MV-51 contain fewer benthic foraminifera and stand in contrast to previously mentioned high BFAR and abundance intervals. In particular, two intervals

0-1.4 and 3.4-6.8 mbsf contain the lowest BFAR values (Figure 2.8). The interval 0-2.4 mbsf (<18,400 Cal. yrs B.P.) is characterized by B. robusta, U. peregrina/hispida, Bul. aculeata, and Gc. subglobosa, in order of decreasing average BFAR values. In contrast, the second interval 3.4-6.8 mbsf (>20,400 Cal. yrs B.P. to <28,00014C yrs B.P.) is

composed mostly of B. robusta, Gc. subglobosa, Bul. aculeata, and G. translucens. The

interval 7-10 mbsf (>28,000 to <32,000 14C yrs B.P.) shows intermediate BFAR values

and is composed of the same dominant species as for 10-12 mbsf.

The Q-mode dendrogram shows the clustering of samples based on the BFAR

values for 12 species at each depth interval (Figure 2.10). In particular, the stratigraphic

record is grouped according to BFAR and the species assemblages discussed above.

2.4.6. Thermomagnetic Results

Samples measured from MV-51 over the entire range of the core (Figure 2.11, n=8)

exhibit a ferrimagnetic behavior at low temperatures but alter at higher temperatures to a

secondary, stronger ferrimagnetic phase. Such phases are typical of iron-containing

detrital minerals. The character of all of the samples measured is very similar, suggesting

that no major change in magnetic mineralogy is present in the core.

38

Figure 2.10. Q-mode cluster dendrogram of benthic foraminiferal samples from MV-51. Analysis used is a 1-Pearson distance metric and average-linkage clustering, n=36 samples.

We interpret this result to indicate that high MS values observed in the base of MV-51

(Figure 2.3) are the result of an increase in the detrital component at the site, with a corresponding decrease in percent calcium carbonate.

2.5. Discussion

2.5.1. Magnetic Susceptibility and Climate

Variations in MS signals primarily reflect the processes governing terrestrial erosion and transport into the marine environment, largely the result of climate change and eustacy

[Tite and Linington, 1975; Verosub et al., 1993; Banerjee, 1996; Ellwood et al., 1996;

39 Ellwood et al., 2000] and biological productivity [Ellwood and Ledbetter, 1977; Mead et al., 1986]. Fluvial and airborne delivery mechanisms account for most detrital input and are themselves primarily functions of precipitation and aridity, respectively. Volcanic ashfall may provide ferri- and paramagnetic compounds to marine sediments. Also affecting the MS, albeit to a lesser extent, are fluxes of diamagnetic calcium carbonate, biogenic silica, and organic matter to the seabed, which are functions of primary productivity in the surface water and therefore of nutrient supply.

Figure 2.11. Thermomagnetic results for eight samples from core MV-51JPC. Numbers adjacent to each curve represent the core depth (cm).

The effects of climate change and provenance on MS records are well documented for terrestrial and marine sediments [e.g. de Menocal et al., 1991; Heller and

40 Evans, 1995; Ellwood et al., 1999; Weedon et al., 1999; Balsam et al., 2005].

Precipitation is one of the principal drivers of chemical and physical weathering,

particularly in the monsoon-dominated tropics. Periods of increased precipitation are

generally associated with enhanced rates of runoff and erosion of weathered detrital

particles. Likewise, drier conditions are less favorable for erosion and transport of

weathered material, particularly during glacial ice advances, because they remove water

vapor available for erosion. During eustatic sea-level lowstands, river mouths may

prograde across exposed shelves to create coastlines that are much closer to shelf breaks,

and therefore may deliver greater amounts of detrital sediment to marine basins.

MS is not strongly influenced by diamagnetic minerals and therefore reflects a

dominantly detrital signal. Linear regression analysis was used to test MS values against

other variables such as TOC and calcium carbonate in cores MV-51 and 54. No association exists between TOC and MS in either MV-51 (R=0.04, p=0.29, n=70) or MV-

54 (r2=0.06, p=0.26, n=62). There is no correlation between calcium carbonate and MS

in either MV-51 (R=0.16, p=0.195, n=68) or MV-54 (R=0.17, p=0.311, n=37).

Additionally, no correlation exists between TOC and calcium carbonate content (MV-51:

R=.14, p=0.255, n=68; MV-54: R=.1, p=0.548, n=37).

Eroded detrital minerals dominate the MS signal, and their production, fate and

transport are controlled by mineral provenance and climate. Changes in river source(s)

and/or climate may explain MS patterns in MV-51. The most notable pattern in the MS

data for MV-51 is the major change at 10 mbsf (Figure 2.6). We propose some

possibilities to explain the higher MS values based on the surrounding PNG geology.

41 • PNG Geology

Papua New Guinea is composed of a complex, heterogeneous geological terrain

consisting of predominantly folded sedimentary rocks in western PNG and volcanogenic

rocks in the central-eastern PNG region [Rickwood, 1968; Davies and Smith, 1971]. The

Fly-Strickland River drainage basin in western PNG erodes predominantly limestone,

siltstone, and sandstone lithologies [Rickwood, 1968, Brunskill, 2004] and as a result discharges relatively mature sediments with high quartz/feldspar ratios (Figure 2.12, fluvial source 1) [Slingerland et al., 2007]. In contrast, central and eastern PNG are dominated by metamorphic and volcanic rocks that result in immature sediment lithologies draining from the other PNG rivers such as the Turama-Kikori-Purari Rivers

(Figure 2.12, fluvial source 2) [Davies and Smith, 1971; Slingerland et al., 2007].

In eastern PNG, the Owen Stanley Range is a large metamorphic region that

forms the prominent mountainous spine of the PNG highlands (Figure 2.12).

Metasedimentary rocks from the Owen Stanley Range consist of greenschist, slate, and

phyllite, and are interrupted by numerous igneous intrusions [Davies and Smith, 1971;

Blake, 1976]. Basaltic, volcanic rocks intrude the Owen Stanley Range and compose a

1000-km belt of Cretaceous to Pliocene rocks that are potassium rich and silica

undersaturated [Mackenzie, 1976; Smith, 1982]. Although the precise mineralogies of these metamorphic and igneous rocks are not well documented, a number of studies suggest that these rocks are mineralogically distinct from the dominantly sedimentary and intrusive andesitic volcanic rocks in western PNG [Rickwood, 1968; Mckenzie, 1976].

Therefore, we hypothesize that in the past, river sediments from central and eastern PNG

42 (Figure 2.12, fluvial sources 2-3) may have been more important than the Fly-Strickland as a source of iron-bearing clastics.

During the Last Glacial Maximum (LGM), lowered sea-level in the GoP enabled

river dispersal systems to incise across the exposed shelf to the shelf-edge [Harris et al.,

1996]. Given the proximity of MV-51 to central and eastern PNG and the higher

gradient, we suggest that the Turama-Kikori-Purari river system may have incised a more

direct pathway to MV-51 so that high MS values below 10 mbsf result from a more direct

input of volcaniclastic material from central and eastern PNG (Figure 2.12, fluvial sources 2-3). The Turama-Kikori-Purari river system would have had far less distance to travel to the northeastern GoP than the Fly River. This hypothesis is testable in the future by collecting quantitative mineralogical data from MV-51 and comparing them to different PNG river sediments. Another test could include age dating amphiboles from above and below 10 m. For example, Pleistocene age dates would indicate that amphiboles came from younger volcaniclastic sources of western PNG while older ages

(e.g. Cretaceous-Pliocene) would indicate a source from the Papuan Peninsula.

• Volcanic ash layers

The source(s) of MV-51 ash layers remain uncertain because few ash layers have

been reported in and around PNG that date to this time interval. Some ash deposits (e.g.

Tomba Tephra) have been studied around Mt. Hagen in western PNG and dated to

~28,000-50,000 yrs B.P. [Pain and Blong, 1976; Chartres and Pain, 1984]. Other ash

deposits, such as the Sagamasi and Natanga Tephras, have been examined around Mt.

Lamington in eastern PNG and dated to 15,000 ± 500-20,100 ±600 Cal. yrs B.P. [Ruxton,

1966].

43 Our radiocarbon dates provide some constraint on the ash layers present in MV-

51. The ash layer at 2.4 mbsf is well constrained between ~18,400 and ~20,400 Cal. yrs

B.P. The ashes at 4.8 and 6.9 mbsf are not as well constrained but were deposited

sometime between ~20,400 Cal. yrs B.P. and ~28,000 14C yrs B.P. Mt. Lamington may

have been a source for the ash layer at 2.4 mbsf but there are not enough dated ash layers

of the correct time frame on PNG to hypothesize the source for the older ash layers.

Therefore, they may originate from any of the active late Quaternary volcanic peaks

shown in Figure 2.12. One exception is Madilogo north of Port Moresby because it is

thought to have formed during the past ~1000 yrs [Blake, 1976]. These ash layers may

serve as correlation points in future GoP studies over this time interval.

2.5.2. Organic Geochemical Data

Organic geochemical proxies suggest that there were significant variations over

time in the paleoflux of organic carbon and the types of sedimentary organic matter

buried in the northeastern Pandora Trough. These data illustrate two important points.

First, TOC and hydrogen index values of sediments are consistently low. Secondly,

stable isotopic and elemental data indicate a complex mixture of marine and C3 vascular

plant matter in the organic carbon, with a potential contribution of other terrestrial plant

matter.

Terrestrial plant matter is common in the modern GoP and therefore it is not surprising that hydrogen index values are indicative of refractory types 3 and 4 organic matter. Small, mountainous rivers are important to the GoP [Milliman, 1995] and on average, 65% of river-borne terrestrial organic carbon is refractory [Ittekkot, 1988]. We frequently observed floating megadetritus (branches to whole tree trunks) in the

44 northeastern GoP during the cruise. In addition, Robertson and Alongi [1995] estimated

that as much as 9 Mt C yr-1 of floating macro and megadetritus occur in the surface water

of the Fly Delta, and that significant quantities are also found on the shelf and in deep-sea

troughs of the GoP. In addition, sediments with hydrogen poor organic matter (i.e.

terrestrial plant matter) are susceptible to much lower hydrogen indices because the

hydrocarbons are retained on mineral matrix during pyrolysis [Katz, 1983; Langford and

Blanc-Valleron, 1990].

Low hydrogen index values in MV-51 are consistent with recent studies of

fluvial-marine dispersal systems [Aller et al., 2004; Goni et al., 2006]. These systems

can promote efficient remineralization of labile organic compounds, particularly during

sediment reworking on the inner shelf, and leave only refractory organic matter.

Therefore, it is possible that a greater amount of labile organic carbon was originally

delivered to MV-51, but that it was later removed from the burial record by oxidation.

We note that trends in pyrolytic hydrogen index data do not correlate consistently

with TOC and stable-isotope-derived percentage of terrestrial TOC (Figure 2.7). One

explanation is that the hydrogen-index data have been biased by hydrogen-rich (algal) labile organic matter in sediments and their oxidation to lower hydrogen index values to reflect type 3 organic matter [Meyers, 1997]. Palynomorph observations indicate common pollen, spores, leaf and grass cuticles, but rare marine palynomorphs such as resistant dinoflagellate cysts. However, total kerogen observations also indicated an abundance of amorphous organic matter, which is derived from degraded marine organic matter and supports the presence of oxidized labile compounds [Lewan, 1986].

45 Stable-carbon isotopic data indicate that TOC contains >50% marine organic carbon throughout much of MV-51. However, the mixing model only takes into consideration two sources, marine and C3 plants. There is also the possibility of an influence from vascular C4 plant matter, soil organic matter upon which C3 and C4 plants grew, or also freshwater algae from rivers and estuaries (Figure 2.7). In addition, it is also possible that the values for the marine end member used in the mixing equation

(-20.5‰ ±0.5‰) changed during the recent geologic past to a more enriched value (e.g. -

19.5‰), which would cause the percentage to change toward more (>50%) terrestrial fractions. At this time, we are unsure about a dominance of marine organic matter as suggested by the stable isotope data and recognize that the estimates of terrestrial sedimentary organic carbon may be minimums.

While keeping in mind the assumptions made in the simple, two-end-member

mixing model, the stable-isotopic and elemental data illustrate a mixture of marine and

predominantly C3 vascular plant matter (Figure 2.7). C3 plants are arboreal/herbaceous

angiosperms and gymnosperms that occur in a wide range of settings where precipitation

is abundant, such as warm and wet tropical rainforests [e.g. mangroves; Tyson, 1995].

C4 plants are fast growing, herbaceous angiosperms (e.g. grasses) that prefer warm and

arid tropical environments such as savannah grasslands [Tyson, 1995; Wagner, 2000;

Wagner et al., 2003]. C3 plants are geochemically distinct from C4 plants in that they

are ~10-15‰ depleted in δ13C (-27‰) compared to C4 plants (-14‰) [Tyson, 1995;

Meyers, 1997; Wagner et al., 2003]. Because of these geochemical distinctions, we suggest that a large proportion, though not all, of the vascular plant matter is from C3 plants.

46 Many global climate records indicate that LGM was both a cooler and drier time

than present; even in the warm tropics [Thompson et al., 1995; Broecker, 1996; Lea et al.,

2000]. Therefore, we had hypothesized that during LGM, a grassland savannah may

have covered much of the PNG coastal plain as it did in northeastern Queensland

[Kershaw, 1978; 1988]. However, none of the samples measured from MV-51 exhibit an

obvious C4 contribution during LGM. Instead, values show a C3 signature that seems to

indicate that wet conditions and C3 vegetation persisted in PNG for the past ~33,000 14C

yrs B.P.

Terrestrial climate records from the GoP are mainly from the Papuan Highlands

[Bowler et al., 1976; Hope, 1976; Hope and Tulip, 1994], although some records exist from the coastal plains [Garrett-Jones, 1979]. Pollen records from the PNG highlands indicate an atmospheric cooling of 6-10°C from 40,000 - 30,000 14C yrs B.P. and during

LGM from 25,000 - 15,000 Cal. yrs B.P. [Bowler et al., 1976; Webster and Streten,

1978]. During the past ~10,000 Cal. yrs B.P. the coastal plain in northern PNG, adjacent to the Solomon Sea, was similar to today in that it was covered by a Lower Montane

Rainforest with some savannah grasses [Garrett-Jones, 1979]. Very little is known about

the coastal plain surrounding the GoP during the past ~10,000 Cal. yrs or preceding time

intervals. Pollen records from Lynch’s Crater, northeastern Queensland (~1000 km SW)

indicate dry conditions and Eucalyptus woodlands dominated during 26,000 – 8,000 Cal.

yrs B.P. [Kershaw, 1978; 1988].

We tentatively suggest that the C3 signal preserved in MV-51 sediments is

indicative of wet tropical vegetation covering the PNG coastal plain and mangroves

rimming the coast similar to today [Harrison and Dodson, 1993; Robertson and Alongi,

47 1995]. Nevertheless, we are unable to rule out the possibility of C4 grasslands in the

vicinity of the GoP without pollen data. Therefore, we recommend a future examination

of pollen spectra from GoP sediment cores to further test this interpretation.

TOC values are slightly higher in MV-54 than they are in MV-51. One

explanation for higher TOC values may be that the high accumulation rates found in MV-

54 would have limited exposure time to organic carbon oxidation more than at site MV-

51 [Hartnett et al., 1998; Hedges et al., 1999]. However, wood-rich organic matter is

abundant throughout MV-54 in the >63µm grain-size fraction and that probably best

explains the overall higher TOC values compared to MV-51. MV-54 also contains a

higher sand-sized fraction of terrigenous siliciclastic particles (mean 15%) compared to

the sand-poor sediments of MV-51 (mean 2%) (Figures 2.3, 2.4). Even without carbon

isotope data, it is apparent from microscopic observations of sand-sized particles, and the limited biogenic calcium carbonate, that MV-54 is dominated by vascular plant matter.

2.5.3. Benthic Foraminifera and Organic Carbon Fluxes

Accumulations of fossil benthic foraminifers are byproducts of multiple

competing biological and taphonomic factors, but overall their abundance and taxonomic composition can indicate relative abundance of labile organic carbon at the time of

deposition. In general, benthic foraminifera live in the upper few cm of sediment, and

their microhabitat selection and maintenance primarily are controlled by food and oxygen, and secondarily by competition for those resources [Jorissen et al., 1995; Van

der Zwaan et al., 1999]. Shell accumulation results from growth and reproduction within

their microhabitat and increases as nutrient supplies increase [Nees and Struck, 1999; Van

der Zwaan et al., 1999]. Postmortem shell abandonment is followed by mixing via

48 bioturbation, transport, and possible removal by dissolution. Shells that eventually pass

into the historical record are a reflection of all of those processes [Loubere, 1989;

Loubere et al., 1993].

In core MV-51, BFAR and assemblages during >32,000 14C yrs B.P. and 18,400-

20,400 Cal. yrs B.P. indicate substantially different oceanographic conditions compared

to modern. The older interval >32,000 14C yrs B.P. contains the highest BFAR of B.

robusta, U. peregrina/hispida, S. bulloides, Bul. aculeata, C. pachyderma, and O. tener.

Species such as U. peregrina/hispida, C. pachyderma, and S. bulloides are characteristic

of high productivity surface waters and higher flux rates of organic matter under low

seasonality [Altenbach and Sarnthein, 1989; Wahyudi and Minagawa, 1997; Loubere and

Farridudin, 1999]. Cibicidoides pachyderma is an epifaunal taxon (0-1 cm) and adapted

to uptake of labile organic matter [Rathburn et al., 1996; Jorissen et al., 1998].

Collectively, these taxa are epifauna to shallow infauna (upper 2 cm), meaning they are

adapted to abundant food and oxygen present in the near surface [Jorissen et al., 1995;

Jorissen et al., 1998].

It is surprising that the lower part of MV-51, with highest MAR and terrestrial

organic-matter content also contains the highest densities of benthic foraminifera.

Normally, high sediment accumulation rates dilute benthic foraminiferal concentrations, and productivity taxa are not adapted to refractory organic carbon. However, the higher accumulation rates and burial of refractory, terrigenous dominated TOC suggests that

there was a much greater river influence at the time. We account for these discrepancies

by suggesting that enhanced river discharge may have supplied nutrients to increase productivity. This provided labile organic carbon for foraminifera, but the labile carbon

49 was not preserved. Highest BFAR before ~32,000 14C yrs B.P. suggests that this was the

time of highest marine productivity compared to any other time interval in MV-51.

The benthic foraminiferal assemblage during ~18,400-20,400 Cal. yrs B.P. indicates a much higher seasonally variable flux of organic carbon to the seabed compared with assemblages from >32,000 14C yrs B.P. Bolivina robusta, G. translucens,

Gc. subglobosa, O. tener, and M. barleeanus comprise the assemblage during this time

interval. Globocassidulina subglobosa is a shallow infaunal (0-2 cm) and opportunistic

species that is able to respond rapidly to phytodetrital pulses from the photic zone [Linke

and Lutze, 1993; Rathburn et al., 1996; Smart and Gooday, 1997]. Gavelinopsis

translucens is typically a shallow infaunal taxon and is found associated with fresh, labile

organic carbon [Jorissen et al., 1998]. In the eastern South Atlantic, an association

between G. translucens and E. exigua is correlated to a higher seasonality of export

production resulting from seasons of maximal upwelling [Licari and Mackensen, 2005].

Melonis barleeanus is an intermediate infaunal taxon (1-4 cm) adapted to degraded

organic matter in areas of high upwelling seasonality [Corliss, 1991; De Stigter et al.,

1998; Jorissen et al., 1998; Loubere and Fariduddin, 1999].

Oceanographic settings with high seasonality in the flux of organic carbon to the

seafloor can be found in many locations from the high latitudes to the tropics. In the

northeast Atlantic Rockall Trough, primary productivity and the flux of phytodetrital

material to the seafloor is restricted to warmer, ice-free summer months and as a result,

specific benthic foraminifera rapidly increase in numbers with the arrival of labile

material at the seafloor [Gooday and Hughes, 2002]. In the Gulf of Guinea, coastal

upwelling is perennial, but it also has strong seasonal upwelling events during May-

50 September, and December which are in part also related to maximal discharge from the

Congo River [Licari and Mackensen, 2005]. Unlike the Gulf of Guinea, the GoP lacks a

western boundary current to promote upwelling, therefore we cannot invoke an upwelling

seasonality in the past. However, changes in the monsoons are seasonal and have

significant effects on the ocean. The enhanced seasonality of organic carbon flux

inferred from benthic foraminifera may be a reflection of past monsoons.

Enhanced seasonality in the GoP may have occurred in the past by increasing the contrast between the present day monsoonal climate so that the southeast trades were stronger and the northwest trades were weaker. Marine productivity, which is dependent on nutrient availability in the surface waters, would have increased when ocean currents were more invigorated from stronger southeast blowing trades, and then reduced during calmer surface waters from lower energy northwest trades. This type of seasonal contrast has been demonstrated in the present day Sulu Sea where primary productivity is highest when wind speeds and upper ocean nutrient mixing are maximal during the East Asian winter monsoon (January – March) [de Garidel-Thoron et al., 2001].

A third interval at 6.8 mbsf is inhabited by B. robusta, U. peregrina/hispida, and

C. pachyderma (Figure 2.8). This assemblage is indicative of high productivity, but it is

a very short-lived event and interrupts an interval of otherwise low BFAR. We suggest

that this sample may be a recovery fauna after deposition of ash at 6.9 mbsf. The

presence of epifaunal species such as C. pachyderma is consistent with post-ashfall

recovery faunas [Hess et al., 2001].

Intervening lower BFAR intervals experienced low organic carbon flux. Bolivina

robusta is consistently abundant in MV-51 and probably is a generalist species, though

51 little is known about its ecology (Figure 2.8, unshaded areas). Dissolution may be one

cause for the low numbers of foraminifera in MV-51; however, the pattern is similar to

TOC MAR. Additionally, there is no correlation between the percent calcium-carbonate

content and benthic foraminiferal densities (R=-0.14; p=0.59; n=32) as one would expect

if dissolution controlled the carbonate content of sediments.

2.5.4. Mass Fluxes, Sea Level, and Sediment Sources

Two periods of high sediment and TOC accumulation are recorded in MV-54 and

MV-51 sediments. The first period (32,000-33,000 14C yrs B.P., late regression)

corresponds to a time when sea-level fall slowed and an extensive mid-shelf clinoform

was aggrading on the mid-shelf [Slingerland et al., 2007]. High accumulation rates are

also observed at other locations in the GoP during this time and may have contributed to

mass transport deposits in the Pandora and Moresby Troughs [Francis et al., 2007]. For

example, sediment cores from the northern Ashmore Trough and southern Pandora

Trough are characterized by turbidites during late regression [Jorry et al., 2007] and a similar “regressive package,” characterized by high terrigenous sediment flux, but deposited during MIS stage 4, is also observed on the slopes of the Ashmore Trough

[Francis et al., 2006; Dickens et al., 2006]. The source for the Ashmore Trough regressive unit may have been sediment resuspended by waves and tides from the prograding mid-shelf clinoform [Dickens et al., 2006] when sea level was ~60 m lower

[Chappell and Shackleton, 1986; Lambeck and Chappell, 2001]. However, MV-51 rests on a bathymetric high (not strongly influenced by turbidity currents), and so sediment must have been delivered through the water column via nepheloid-layer advection or similar means.

52 The MV-51 MS and thermomagnetic results suggest a source shift from sediments with high concentrations (early) to lower concentrations (later) of iron-rich paramagnetic minerals around 32,000 14C yrs B.P. Additionally, benthic foraminiferal

species indicate oxygenated bottom water and greater organic-carbon flux before that

time than after. Organic geochemical proxies suggest greater terrestrial input before

32,000 14C yrs B.P. than after. These observations implicate greater fluvial discharge for

these sediments before 32,000 14C yrs B.P, possibly originating from the Papuan

Peninsula (Figure 2.12, fluvial source 3).

Sediment delivery was very high during ~15,000 – 18,000 yrs B.P. at MV-54 and

likely corresponds to a stillstand time after the first major transgression (10-15 m) associated with a meltwater pulse at 19,000 yrs B.P. [Clark et al., 2004]. During this

time, sea level was ~110 m lower and likely permitted the Fly, Turama, Kikori, and

Purari rivers to empty much closer to the shelf edge [Figure 2.12; Lambeck and

Chappell, 2001; Clark et al., 2004; Milliman et al., 2004]. For example, the chronology of MV-54 indicates extremely rapid emplacement of sediments, on the order of six meters in <1000 yrs (Figures 2.2, 2.4). Mass failure from upslope may be one explanation, because some coarse, inclined beds were observed in that interval. Fluxes at

MV-51 during 18,400-20,400 yrs B.P. were also high and likely resulted from direct delivery of river sediment when sea level was lower (Figure 2.12, fluvial source 2).

MV-54 and MV-51 show significantly lowered accumulation rates from the late

transgression to early Holocene time (<15,000 yrs B.P.). This partly resulted from inland

sediment storage as sea level flooded the shelf and increased accommodation space

[Harris et al., 1996].

53

Figure 2.12. Map of the GOP and hinterland topography and geology showing major sources of river input. The geological map was constructed based on information from various sources including Davies and Smith [1971], Mackenzie [1976], Smith [1982] and Steinshouer et al., [1999]. Fluvial source 1 is the Fly-Strickland-Bamu river system, 2 is the Turama-Kikori-Purari rivers, and 3 is a proposed input from the Papuan Peninsula to explain data in MV-51. Two bathymetric lines at 60 and 120 m are shown to reconstruct probable position of the stage 3 (>30,000 yrs B.P.) and the Last Glacial Maximum (23- 19,000 yrs B.P.) coastlines, respectively. Bathymetric profiles A and B illustrate decreased shelf space during LGM relative to recent. The coastlines shown were subsampled from data provided by Daniell [2007].

A coralgal ridge developed at the shelf-edge during sea-level rise from 14,500-12,000 yrs

B.P. suggesting less suspended sediment in the water column; this ridge may also have

54 served as a barricade to sediment escape in the central and northeastern Pandora Trough

[Droxler et al., 2006]. These reefs started growing on the shelf-edge up dip of MV-51 at

~19,000 yrs B.P. with additional growth during the meltwater pulse 1a (14,500–12,500

yrs B.P.) [Droxler et al., 2006].

2.6. Conclusions

Sediment cores examined in this study span the time interval from ~15,000 to

33,000 yrs B.P. and provide a unique window into depositional events at sites MV-54 and

MV-51 in the northeastern GoP. Geochemical and paleontological evidence indicate

different oceanographic conditions in the GoP >32,000 14C yrs B.P. compared to any younger time First, higher TOC MAR and total BFAR prior to 32,000 14C yrs B.P suggest that greater amounts of organic carbon were being delivered to the seabed.

Increases in the densities of productivity taxa (U.peregrina/hispida-C. pachyderma-S. bulloides) suggest some fraction of this organic carbon was in labile form. However, low hydrogen index values may suggest that if any labile compounds were originally present, they were mostly oxidized and not buried. During this time, increased percentages of terrestrial organic carbon over marine organic carbon suggest an abundant input of C3 and possibly other vascular plant matter. A major clastic sediment source change at

~32,000 14C yrs B.P. is suggested by MS, but thermomagnetic results indicate similar

magnetic mineralogy throughout the core. Higher MS values during this interval are

probably the result of increased clastic flux from the central and eastern PNG rivers

rather than significant changes in mineralogy.

The time from 15,000-20,400 yrs B.P. was a period of high sediment

accumulation rates. MV-54 experienced the greatest accumulation rates during late

55 transgression (~15,000-18,000 yrs B.P.) when rivers were situated on the exposed shelf

and delivered sediments directly to the slopes, although we do not know what

accumulation rates were prior to 18,000 yrs B.P. at this site. Stable-carbon isotopes from

MV-51 point to enhanced terrestrial organic carbon delivery, while the G. translucens,

Gc. subglobosa, and M. barleeanus benthic foraminiferal assemblage indicates higher

seasonality of organic carbon flux during this time. The significantly lower accumulation

rates at MV-54 and MV-51 during late transgression to Holocene time are evidence of

inland and coastal sediment storage and possible isolation of sediment supply from the

GoP slope by a shelf-edge coralgal complex upslope from MV-51.

2.7. References

Aller, R.C. and N.E. Blair (2004), Early diagenetic remineralization of sedimentary organic C in the Gulf of Papua deltaic complex (Papua New Guinea): Net loss of terrestrial C and diagenetic fractionation of C isotopes, Geochimica et Cosmochimica Acta, 68(8), 1815-1825.

Altenbach, A.V. and M. Sarnthein (1989), Productivity record in benthic foraminifera, in Productivity of the Ocean: Present and Past, edited by W.H. Berger, V.S. Smetacek, and G. Wefer, pp. 255-269, John Wiley & Sons Limited, New York, New York.

Amo, M. and M. Minagawa (2003), Sedimentary record of marine and terrigenous organic matter delivery to the Shatsky Rise, western North Pacific, over the last 130 kyrs, Organic Geochemistry, 34, 1299-1312.

Balsam, W., B.B. Ellwood, and J. Ji (2005), Direct correlation of the marine oxygen isotope record with Loess Plateau iron oxide and magnetic susceptibility records, Palaeogeography, Palaeoclimatology, Palaeoecology, 221, 141-152.

Banerjee, S.K. (1996), Sediment reveals early Holocene climate change in China, Eos Transactions AGU, 77(1), 3, 5.

Bentley, S.J., Z. Muhammad, L.Patterson, J. Dickens, B.N. Opdyke, L. Peterson, and A.W. Droxler (2006), Modern and Pleistocene turbidite sedimentation in the Gulf of Papua S2S study area: Implications for modulation of sediment sources by oceanic and climatic processes. AAPG Annual Meeting, Houston Texas.

56 Berger, W.H. and J.C. Herguera (1992), Reading the sedimentary record of the ocean’s productivity. in Primary Productivity and Biogeochemical Cycles in the Sea, edited by P.F. Falkowski and A.D. Woodhead, pp. 455-486, Plenum Press, New York, New York.

Berner, R.A. (2003), The long-term carbon cycle, fossil fuels and atmospheric composition, Nature, 426: 323-326.

Bird, M.I., G.J. Brunskill, and A.R. Chivas (1995), Carbon-isotope composition of sediments from the Gulf of Papua, Geo-Marine Letters, 15, 153-159.

Blake, D.H. (1976), Madilogo, a late Quaternary near Port Moresby, Papua New Guinea, in Volcanism in Australasia, edited by R.W. Johnson, pp. 253-258, Elsevier Scientific Publishing Company, Amsterdam, The Netherlands.

Bouloubassi, I, J. Rullkötter, and P.A. Meyers (1999), Origin and transformation of organic matter in Pliocene-Pleistocene Mediterranean sapropels: Organic geochemical evidence reviewed, Marine Geology, 153, 177-197.

Bowler, J.M., G.S. Hope, J.N. Jennings, G. Singh, and D. Walker (1976), Late Quaternary climates of Australia and New Guinea, Quaternary Research, 6, 359-394.

Broecker, W. (1996), Glacial climate in the tropics, Science, 272, 1902-1904.

Broecker, W., S. Barker, E. Clark, I. Hajdas, G. Bonani, and L. Stott (2004), Ventilation of the glacial deep Pacific Ocean, Science, 306, 1169-1172.

Bronk Ramsey, C. (1995), Radiocarbon calibration and analysis of stratigraphy: The OxCal Program, Radiocarbon, 37(2), 425-430.

Brunskill, G.J., K.J. Woolfe, and I. Zagorskis (1995), Distribution of riverine sediment chemistry on the shelf, slope and rise of the Gulf of Papua, Geo-Marine Letters, 15 (3-4), 160-165.

Chappell, J. and N.J. Shackleton (1986), Oxygen isotopes and sea level, Nature, 321, 137-140.

Chartres, C.J. and C.F. Pain (1984), A climosequence of soils on Late Quaternary volcanic ash in highland Papua New Guinea, Geoderma, 32, 131-155.

Clark, P.U., A.M. McCabe, A.C. Mix, and A.J. Weaver (2004), Rapid rise of sea level 19,000 years ago and its global implications, Science, 304, 1141-1144.

Corliss, B.H. (1985), Microhabitats of benthic foraminifera within deep-sea sediments, Nature, 314, 435-438.

57 Corliss, B.H. (1991), Morphology and microhabitat preferences of benthic foraminifera from the northwest Atlantic Ocean, Marine Micropaleontology, 17, 195-236.

Daniell, J., In Press, Development of a 100-m bathymetry grid for the Gulf of Papua, Journal of Geophysical Research - Earth Surface

Davies, H.L. and I.E. Smith (1971), Geology of Eastern Papua, Geological Society of America Bulletin, 82, 3299-3312.

Delvaux, D., H. Martin, P. Leplat, and J. Paulet (1990), Geochemical characterization of sedimentary organic matter by means of pyrolysis kinetic parameters, Organic Geochemistry, 16(1-3), 175-187.

de Garidel-Thoron, T., L. Beaufort, B.K. Linsley, and S. Dannenmann (2001), Millenial- scale dynamics of the East Asian winter monsoon during the last 200,000 years, Paleoceanography, 16, 1-12.

deMenocal, P., J. Bloemendal, and J. King (1991), A rock-magnetic record of monsoonal dust deposit ionto the Arabian Sea: Evidence for a shift in the mode of deposition at 2.4 Ma, in Proceedings ODP, Scientific Results 117, edited by W. Prell and L. Niitsuma, pp. 389-407.

De Stigter, H.C., F.J. Jorissen, and G.J. van der Zwann (1998), Bathymetric distribution and microhabitat partitioning of live (rose bengal stained) benthic foraminifera along a shelf to bathyal transect in the southern Adriatic Sea, Journal for Foraminiferal Research, 28(1), 40-65.

Dickens, G.R., A.W. Droxler, S.J. Bentley, L.C. Peterson, B.N. Opdyke, L. Beaufort, J. Daniell, L.A. Febo, J. Francis, P.T. Harris, S. Jorry, G. Mallarino, M. McFadden, Z. Muhammad, B. Carson, L. Patterson, E. Tcherepanov, and C.A. Zarickian (2006), A preliminary late quaternary view of sedimentary sinks in the Papua New Guinea Source- to-Sink focus area, MARGINS Newsletter, 16, 1-5.

Droxler, A., G. Mallarino, J.M. Francis, J. Dickens, L. Beaufort, S. Bentley, L. Peterson, B. Opdyke (2006), Early Part of Last Deglaciation: A Short Interval Favorable for the Building of Coralgal Edifices on the Edges of Modern Siliciclastic Shelves (Gulf of Papua and Gulf of Mexico), AAPG Annual Meeting.

Ellwood, B.B., C.E. Brett, and W.D. MacDonald (2007), Magnetosusceptibility Stratigraphy of the Upper Ordovician Kope Formation, Northern Kentucky, Palaeogeography, Palaeoclimatology, Palaeoecology, 243, 42-54.

Ellwood, B.B., R.E. Crick, and A. Hassani (1999), The magneto-susceptibility event and cyclostratigraphy (MSEC) method used in geological correlation of Devonian rocks of the Anti-Atlas Morocco, AAPG Bulletin, 83(7), 1119-1134.

58 Ellwood, B.B., R.E. Crick, A. Hassani, S.L. Benoist, and R.H. Young (2000), Magnetosusceptibility event and cyclostratigraphy method applied to marine rocks: detrital input versus carbonate productivity, Geology, 28(12), 1135-1138.

Ellwood, B.B. and M.T. Ledbetter (1977), Antarctic bottom water fluctuation in the Vema Channel: Effects of velocity changes on particle alignment and size, Earth Planetary Science Letters, 35, 189-198.

Ellwood, B.B. K.M. Petruso, and F.B. Harrold (1996), The utility of magnetic susceptibility for detecting paleoclimatic trends and as a stratigraphic tool: an example from Konispol Cave sediments, SW Albania, Journal of Field Archaeology, 23, 263-271.

Espitalié, J., M. Madec, and B. Tissot (1980), Role of mineral matrix in kerogen pyrolysis: Influence of petroleum generation and migration, AAPG Bulletin, 64, 59-66.

Espitalié, J. and M.L. Bordenave (1993), Rock-Eval pyrolysis, in Applied Petroleum Geochemistry, edited by M.L. Bordenave, pp. 237-261, Editions Technip, Paris, France.

Fairbanks, R.G., R.A. Mortlock, T.C. Chiu, L. Cao, A. Kaplan, T.P. Guilderson, T.W. Fairbanks, A.L. Bloom, P.M. Grootes, M.J. Nadeau (2005), Radiocarbon calibration curves spanning 0 to 50,000 years BP based on paired 230Th/234U/238U and 14C dates on pristine corals, Quaternary Science Reviews, 24, 1781-1796.

Francis, J. M., J. Daniell, A. W. Droxler, G. R. Dickens, S. J. Bentley, L. C. Peterson, B. Opdyke, and L. Beaufort (2007), Deepwater geomorphology and sediment pathways of the mixed siliciclastic/carbonate system, Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Francis, J.M., B. Olson, G.R. Dickens, A.W. Droxler, L. Beaufort, T. De Garidel-Thoron, S. Bentley, L.C. Peterson, B. Opdyke (2006), High Siliciclastic Flux during the Last Sea Level Transgression, Ashmore Trough, Gulf of Papua, AAPG Annual Meeting.

Flood, R., and D. Piper (1997), Amazon Fan sedimentation: the relationship to equatorial climate change, continental denudation, and sea-level fluctuations, in Proceedings ODP, Scientific Results 155, edited by R.D. Flood, D.J.W. Piper, A. Klaus, and L.C. Peterson, pp. 653-675.

Garrett-Jones, S.E. (1979), Holocene vegetation and lake sedimentation in the Markham Valley, Papua New Guinea, Ph.D. Dissertation, 420 pp., Australia National University, Canberra, Australia.

Gooday, A.J. (1988), A response by benthic Foraminifera to the deposition of phytodetritus in the deep sea, Nature, 332, 70-73.

Gooday, A.J. and J.A. Hughes (2002), Foraminifera associated with phytodetritus deposits at a bathyal site in the northern Rockall Trough (NE Atlantic): seasonal contrasts

59 and a comparison of stained and dead assemblages, Marine Micropaleontology, 46, 83- 110.

Goni, M.A., N. Monacci, R. Gisewhite, A. Ogston, J. Crockett, and C. Nittrouer (2006), Distribution and sources of particulate organic matter in the water column and sediments of the Fly River Delta, Gulf of Papua (Papua New Guinea), Estuarine Coastal and Shelf Science, 69, 225-245.

Harris, P.T., E.K. Baker, A.R. Cole, and S.A. Short (1993), A preliminary study of sedimentation in the tidally dominated Fly River Delta, Gulf of Papua, Continental Shelf Research, 13, 441-472.

Harris, P.T., C.B. Pattiaratchi, J.B. Keene, R.W. Dalrymple, J.V. Gardner, E.K. Baker, A.R. Cole, D. Mitchell, P. Gibbs, and W.W. Schroeder (1996), Late Quaternary deltaic and carbonate sedimentation in the Gulf of Papua foreland basin: Response to sea-level change, Journal Sedimentary Research, 66(4), 801-819.

Harrison, S.P. and J. Dodson (1993), Climates of Australia and New Guinea since 18,000 yr B.P., in Global Climates Since the Last Glacial Maximum, edited by H.E. Wright, J.E. Kutzbach, T. Webb, W.F. Ruddiman, F.A. Street-Perrott, and P.J. Bartlein, pp. 265-293, University of Minnesota Press, Minneapolis, Minnesota.

Hartnett, H.E., R.G. Keil, J.I. Hedges, and A.H. Devol (1998), Influence of oxygen exposure time on organic carbon preservation in continental margin sediments, Nature, 391, 572-574.

Hedges, J.I., R.G. Keil, and R. Benner (1997), What happens to terrestrial organic matter in the ocean?, Organic Geochemistry, 27: 195-212.

Hedges, J.I., F.S. Hu, A.H. Devol, H.E. Hartnett, E. Tsamakis, and R.G. Keil (1999), Sedimentary organic matter preservation: A test for selective degradation under oxic conditions, American Journal of Science, 299, 529-555.

Heller, F. and M.E. Evans (1995), Loess magnetism, Review of Geophysics, 33, 211-240.

Hemer, M.A., P.T. Harris, R. Coleman, and J. Hunter (2004), Sediment mobility due to currents and waves in the Torres Strait - Gulf of Papua region, Continental Shelf Research, 24, 2297-2316.

Herguera, J.C. (1992), Deep-sea benthic foraminifera and biogenic opal: Glacial to postglacial productivity changes in the western equatorial Pacific, Marine Micropaleontology, 19(1/2), 79-98.

Herguera, J.C. and W.H. Berger (1994), Glacial to post glacial drop on productivity in the western equatorial Pacific: mixing rate vs. nutrient concentrations, Geology, 22, 629-632.

60 Hess, S., W. Kuhnt, S. Hill, M.A. Kaminski, A. Holbourn, and M. de Leon (2001), Monitoring the recolonization of the Mt Pinatubo 1991 ash layer by benthic foraminifera, Marine Micropaleontology, 43, 119-142.

Hope, G.S. (1976), The vegetational history of Mt. Wilhelm, Papua New Guinea, Ecology, 64, 627-664.

Hope, G. and J. Tulip (1994), A long vegetation history from lowland Irian Jaya, Indonesia, Palaeogeography, Palaeoclimatology, Palaeoecology, 109, 385-398.

Hrouda, F. (1994), A technique for the measurement of thermal changes of magnetic susceptibility of weakly magnetic rocks by the CS-2 apparatus and the KLY-2 Kappabridge, Geophysical Journal International, 118, 604-612.

Hughen, K., S. Lehman, J. Southon, J. Overpeck, O. Marchal, C. Herring, and J. Turnbull (2004), 14C activity and global carbon cycle changes over the past 50,000 years, Science, 303, 202-207.

Ittekkot, V. (1988), Global trends in the nature of organic matter in river suspensions, Nature, 332, 436-438.

Jarzie, D.M. (1991), Total organic carbon (TOC) analysis, in Source and Migration processes and evaluation techniques. Treatise of Petroleum Geology, Handbook of Petroleum Geology, edited by R.K. Merrill, American Association of Petroleum Geologists, Tulsa, pp. 113-118.

Jones, G.A. and P. Kaiteris (1983), A vacuum-gasometric technique for rapid and precise analysis of calcium carbonate in sediments and soils, Journal of Sedimentary Petrology, 53, 655-659.

Jorissen, F.J., D.M. Barmawidjaja, S. Puskaric, and G.J. van der Zwann (1992), Vertical distribution of foraminifera in the northern Adriatic Sea: The relation with the organic flux, Marine Micropaleontology, 1(1/2), 131-146.

Jorissen, F.J., H.C. De Stigter, and J.G.V. Widmark (1995), A conceptual model explaining benthic foraminiferal microhabitats, Marine Micropaleontology, 26, 3-15.

Jorissen, F.J., I. Wittling, J.P. Peypouquet, C. Rabouille, and J.C. Relexans (1998), Live benthic foraminiferal faunas off Cape Blanc, MW-Africa: Community structure and microhabitats, Deep Sea Research Pt. 1, 45, 2157-2188.

Jorry, S.J., A.W. Droxler, G. Mallarino, G.R. Dickens, S.J. Bentley, L. Beaufort, and L.C. Peterson (2007), Bundled turbidite deposition in the Central Pandora Trough (Gulf of Papua) since Last Glacial Maximum: Linking sediment nature and accumulation to sea- level fluctuations at millennial time scale, Journal of Geophysical Research - Earth Surface.

61

Katz, B. (1983), Limitations of ‘Rock-Eval’ pyrolysis for typing organic matter, Organic Geochemistry, 4, 195-199.

Keen, T.R., D.S. Ko, R.L. Slingerland, S. Reidlinger, and P. Flynn (2006), Potential transport pathways of terrestrial material in the Gulf of Papua, Geophysical Research Letters, 33, 4, L04608, doi:10.1029/2005GL025416

Kershaw, A.P. (1978), Record of the last interglacial-glacial cycle from northeastern Queensland, Nature, 212, 156-161.

Kershaw, A.P. (1988), Australasia, in Vegetation History, edited by B. Huntley and T. Webb, pp. 237-306, Kluwer Academic Publishers, Dordrecht, The Netherlands.

Lallier-Vergès, E., P. Bertrand, and A. Desprairies (1993), Organic matter composition and sulfate reduction intensity in Oman Margin sediments, Marine Geology, 112, 57-69.

Lambeck, K. and J. Chappell (2001), Sea level change through the last glacial cycle, Science, 292, 679-686.

Langford, F.F. and M.M. Blanc-Valleron (1990), Interpreting Rock-Eval Pyrolysis data using graphs of pyrolizable hydrocarbons vs. total organic carbon, AAPG Bulletin, 74, 799-804.

Lea, D.W., D.K. Pak, and H.J. Spero (2000), Climate impact of Late Quaternary Equatorial Pacific sea surface temperature variations, Science, 289, 1719-1724.

Lewan, M.D. (1986), Stable carbon isotopes of amorphous kerogens from Phanerozoic sedimentary rocks, Geochimica et Cosmochimica Acta, 50, 1583-1591.

Licari, L. and A. Mackensen (2005), Benthic foraminifera off West Africa (1°N to 32°): Do live assemblages from the topmost sediment reliably record environmental variability?, Marine Micropaleontology, 55, 205-233.

Linke, P. and G.F. Lutze (1993), Microhabitat preferences of benthic foraminifera- a static concept or a dynamic adaptation to optimize food acquisition? Marine Micropaleontology, 20, 215-234.

Loubere, P. (1989), Bioturbation and sedimentation rate control of benthic microfossil taxon abundances in surface sediments: A theoretical approach to the analysis of species microhabitats, Marine Micropaleontology, 14, 317-325.

Loubere, P., A. Gary, and M. Lagoe (1993), Generation of the benthic foraminiferal assemblage: Theory and preliminary data, Marine Micropaleontology, 20, 165-181.

Loubere, P. (1996), The surface ocean productivity and bottom water oxygen signals in

62 deep water benthic foraminiferal assemblages, Marine Micropaleontology, 28, 247-261.

Loubere, P. and M. Farridudin (1999), Benthic foraminifera and the flux of organic carbon to the seabed, in Modern Foraminifera, edited by B.K. Sen Gupta, pp. 181-199, Kluwer Academic Publishers, Dordrecht, The Netherlands.

Mackenzie, D.E. (1976), Nature and origin of Late Cainozoic volcanoes in western Papua New Guinea, in Volcanism in Australasia, edited by R.W. Johnson, pp. 221-238, Elsevier Scientific Publishing Company, Amsterdam, The Netherlands.

Mead, G.A. and L. Tauxe (1986), Oligocene paleoceanography of the South Atlantic: Paleoclimatic implications of sediment accumulation rates and magnetic susceptibility measurements, Paleoceanography, 1(3), 273-284.

Meyers, P.A. (1994), Preservation of elemental and isotopic source identification of sedimentary organic matter, Chemical Geology, 144, 289-302.

Meyers, P.A. (1997), Organic geochemical proxies of paleoceanographic, paleolimnologic, and paleoclimatic processes, Organic Geochemistry, 27(5/6), 213-250.

Milliman, J.D. and J.P.M. Syvitski (1992), Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers, Journal of Geology, 100, 525-544.

Milliman, J.D. (1995), Sediment discharge to the ocean from small mountainous rivers: the New Guinea example, Geo-Marine Letters, 15, 127-133.

Milliman, J.D. , K.L. Farnswerth, and C.S. Albertin (1999), Flux and fate of fluvial sediments leaving large islands in the East Indies, Journal of Sea Research, 41, 97-107.

Milliman, J. D., N. W. Driscoll, R. Slingerland, J. Babcock, and J. P. Walsh (2004), Isotopic Stage 3 Deposition and Stage 2 Erosion of a Clinoform in the Gulf of Papua: Regional Tectonics Versus Eustatic Sea-Level Change, Eos Trans. AGU, 85(47), Fall Meet. Suppl., OS44A-05.

Muhammad, Z., S.J. Bentley, L.A. Febo, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke (2007), Excess 210Pb inventories and fluxes along the continental slope and basins of the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Nagata, T. (1961), Rock Magnetism, Tokyo, Maruzen, 350pp.

Nees, S. and U. Struck (1999), Benthic foraminiferal response to major paleoceanographic changes: A view from the deep-sea restaurant menu, in Reconstructing Ocean History: A Window into the Future, edited by F. Abrantes, and A. Mix, pp. 195-216, Plenum Publishers, New York.

63 Pain, C.F. and R.J. Blong (1976), Late Quaternary tephras around Mount Hagen and , Papua New Guinea, in Volcanism in Australasia, edited by R.W. Johnson, pp. 239-252, Elsevier Scientific Publishing Company, Amsterdam, The Netherlands.

Patterson, L.J., S.J. Bentley, D.J. Henry, A.W. Droxler, G.R. Dickens, B.N. Opdyke and L.C. Peterson (2006), Multiple-source areas for turbidite sands reaching the deep basin in the Gulf of Papua, AAPG Annual Meeting.

Pérez, M.E., C.D. Charles, and W.H. Berger (2001), Late Quaternary productivity fluctuations off Angola: Evidence from benthic foraminifers, site 1079, in Proceedings ODP, Scientific Results 175, edited by G. Wefer, W.H. Berger, and C. Richter, pp. 1-19.

Rathburn, A.E., B.H. Corliss, K.D. Tappa, and K.C. Lohmann (1996), Comparisons of the ecology and stable isotopic compositions of living (stained) benthic foraminifera from the Sulu and South China seas, Deep Sea Research Pt. 1, 43 (10), 1617-1646.

Rickwood, F.K. (1968), The geology of Western Papua, The APEA Journal, 8 (2), 51-61.

Robertson, A.I. and D.M. Alongi (1995), Role of riverine mangrove forests in organic carbon export to the tropical coastal ocean: A preliminary mass balance for the Fly Delta (Papua New Guinea), Geo-Marine Letters, 15, 134-139.

Rühlemann, C., P.J. Müller, and R.R. Schneider (1999), Organic carbon and carbonate as paleoproductivity proxies: Examples from high and low productivity areas of the Tropical Atlantic, in Use of Proxies in Paleoceanography: Examples from the South Atlantic, edited by G. Fischer and G. Wefer, pp. 315-344, Springer-Verlag, Berlin, Germany.

Rullkötter, J. (2000), Organic matter: The driving force for early diagenesis, in Marine Geochemistry, edited by H. D. Schultz and M. Zabel, pp. 129-172, Springer-Verlag, Berlin, Germany.

Ruxton, B.P. (1966), Correlation and stratigraphy of dacitic ash-fall layers in northeastern Papua, Journal of the Geological Society of Australia, 13, 41-67.

Schlesinger, W.H. and J.M. Melack (1981), Transport of organic carbon in the world’s rivers, Tellus, 33, 172-187.

Schlünz, B. and R.R. Schneider (2000), Transport of terrestrial organic carbon to the oceans by rivers: Re-estimating flux- and burial rates, International Journal of Earth Sciences, 88, 599-606.

Sen Gupta, B.K., I.C. Shin, and S.T. Wendler (1987), Relevance of specimen size in distribution studies of deep-sea benthic foraminifera, Palaios, 2, 332-338.

64 Sjoerdsma, P.G. and G.J. van der Zwaan (1992), Simulating the effect of changing organic flux and oxygen content on the distribution of benthic foraminifers, Marine Micropaleontology, 19(1/2), 163-180.

Slingerland, R., N.W. Driscoll, J.D. Milliman, S.R. Miller, and E.A. Johnstone In Press, Anatomy and growth of a Holocene clinothem in the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Smart, C.W. and A.J. Gooday (1997), Recent benthic foraminifera in the abyssal northeast Atlantic Ocean: Relation to phytodetrital inputs, Journal of Foraminiferal Research, 27 (2), 85-92.

Smith, I.E.M. (1982), Volcanic evolution in Eastern Papua, Tectonophysics, 87, 315-333.

Steinshouer, D.W., J. Qiang, P.J. McCabe, and R.T. Ryder (1999), Maps showing geology, oil and gas fields, and geologic provinces of the Asia Pacific Region. U.S.G.S. Open File Report 97-470F.

Stuiver, M., P.J. Reimer, E. Bard, J.W. Beck, G.S. Burr, K.A. Hughen, B. Kromer, G. McCormac, J. van der Plicht, and M. Spurk (1998), INTCAL98 Radiocarbon Age Calibration, 24000-0 cal BP, Radiocarbon, 40(3), 1041-1083.

Thom, B.G. and L.D. Wright (1983), Geomorphology of the Purari Delta, in The Purari: Tropical Environment of a High Rainfall River Basin, edited by T. Petr, pp. 47-65, Dr. W. Junk Publishers, The Hague.

Thompson, L.G., E. Mosley Thompson, M.E. Davis, P.N. Lin, K.A. Henderson, J. Coledai, J.F. Bolzan, and K.B. Liu (1995), Late-glacial stage and Holocene tropical ice core records from Huascaron, Peru, Science, 269, 46-50.

Thompson, R. and F. Oldfield (1986), Environmental Magnetism, Allen and Unwin, London, England.

Tissot, B.P. and D.H. Welte (1984), Petroleum Formation and Occurrence, 699 pp., Springer-Verlag, Berlin, Germany.

Tite, M.S. and R.E. Linington (1975), Effect of climate on the magnetic susceptibility of soils, Nature, 256, 565-566.

Tyson, R.V. (1995), Sedimentary Organic Matter: Organic Facies and Palynofacies, 615 pp., Chapman and Hall, New York.

Van der Zwaan, G.J., I.A.P. Duijnstee, M. den Dulk, S.R. Ernst, N.T. Jannink, and T.J. Kouwenhoven (1999), Benthic foraminifers: Proxies or problems? A review of paleoecological concepts, Earth Science Reviews, 46, 213-236.

65 Wagner, T. (2000), Control of organic carbon accumulation in the late Quaternary equatorial Atlantic (Ocean Drilling Program sites 664 and 663): Productivity versus terrigenous supply, Paleoceanography, 15(2), 181-199.

Wagner, T., M. Zabel, L. Dupont, J. Holtvoeth, and C.J. Schubert (2004), Terrigenous signals in sediments of the low latitude Atlantic - Indications to environmental variations during the Late Quaternary: Part 1: Organic Carbon, in The South Atlantic in the Late Quaternary: Reconstructions of Material Budgets and Current Systems, edited by G. Wefer, S. Mulitza, and V. Ratmeyer, pp. 295-322, Springer-Verlag, Berlin, Germany.

Wahyudi and M. Minagawa (1997), Response of benthic foraminifera to organic carbon accumulation rates in the Okinawa Trough, Journal of Oceanography, 53, 411-420.

Walsh, J.P. and C.A. Nittrouer (2003), Contrasting styles of off-shelf sediment accumulation in New Guinea, Marine Geology, 196, 105-125.

Walsh, J.P., C.A. Nittrouer, C.M. Palinkas, A.S. Ogston, R.W. Sternberg, and G.J. Brunskill (2004), Clinoform mechanics in the Gulf of Papua, New Guinea, Continental Shelf Research, 24, 2487-2510.

Weber, M.E., F. Niessen, G. Kuhn, and M. Wiedicke (1997), Calibration and application of marine sedimentary physical properties using a multi-sensor core logger, Marine Geology, 136, 151-172.

Webster, P.J. and N.A. Streten (1978), Late Quaternary ice age climates of tropical Australasia: Interpretations and reconstructions, Quaternary Research, 10, 279-309.

Weedon, G.P., H.C. Jenkyns, A.L. Coe, and S.P. Hesselbo (1999), Astronomical calibration of the Jurassic time-scale from cyclostratigraphy in British mudrock formations, Philosophical Transactions of the Royal Society of London, A, 357, 1787- 1813.

Wolanski E., G. L. Pickard, and D. L. B. Jupp (1984), River plumes, coral reefs and mixing in the Gulf of Papua and the northern Great Barrier Reef, Estuarine, Coastal and Shelf Science, 18, 291-314.

Wolanski, E. and D.M. Alongi (1995), A hypothesis for the formation of a mud bank in the Gulf of Papua, Geo-Marine Letters, 15, 166-171.

Wolanski, E., A. Norro, and B. King (1995), Water circulation in the Gulf of Papua, Continental Shelf Research, 15(2-3), 185-212.

66 CHAPTER 3 SURFACE SEDIMENTS AND DISPERSAL PATHWAYS ON THE OUTER SHELF-EDGE AND BASIN IN THE GULF OF PAPUA: MAGNETIC SUSCEPTIBILITY, CALCIUM CARBONATE, AND ORGANIC GEOCHEMICAL MEASUREMENTS

3.1. Introduction

Deep marine sedimentary basins beyond the shelf-edge are one of the ultimate storage sites for siliciclastic and biogenic particles over long-term geological timescales.

During times of sea-level lowstand, siliciclastic and organic particles are transported directly into deep-water environments and form mass transport deposits usually in the form of submarine fans [Posamentier and Vail, 1988; Flood and Piper; 1997; Maslin and

Mikkelsen, 1997]. However, detrital and organic particles are stored in clinoform shelf deposits during times of higher sea-level (e.g. highstands) and are rarely transported farther offshore. For example, in mixed siliciclastic-carbonate depositional systems such as the Queensland Margin [Francis et al., 2007] and the Gulf of Papua (GoP), modern siliciclastic sediments discharged from rivers are stored in near-shore inner-shelf sediment wedges or clinoforms [Walsh et al., 2004; Palinkas et al., 2006; Slingerland et al., 2007]. Nevertheless, certain amounts of recent sediment are episodically re- deposited onto the slopes and troughs through gravity flows and dense nepheloid layer plumes. Therefore, understanding the composition and distribution of sediments blanketing the seafloor in deep-marine environments can potentially help to discern the overall net transport pathways of sediments [Nittrouer and Wright, 1994; Ellwood et al.,

2006]. It is also important to accurately model sediment dispersal systems from a source- to-sink perspective as well as to improve our understanding of global biogeochemical cycles.

67 In this study, we examine sediments blanketing the upper slope and basin of the

GoP, which is one of the National Science Foundation’s MARGINS Source-to-Sink

(S2S) focus areas. Our main goal is to determine the nature of recent deep-sea sediments

and the origin of their formation. In particular, we examine the detrital and biogenic

composition of outer shelf-edge and basinal sediments using magnetic susceptibility

(MS), calcium carbonate content, organic geochemical measurements, and organic

petrography on surficial sediments, and then map the distribution of sediment types. On

the basis of these maps, the overall transport pathways are then determined.

3.2. Gulf of Papua (GoP)

3.2.1. Recent Marine Sedimentation

The GoP (Figure 3.1) is a tropical mixed siliciclastic-carbonate depositional

setting between the northeastern Great Barrier Reef (GBR) platform and Papua, New

Guinea, that receives a very large supply of terrigenous sediment from numerous river

systems. Between 300 and 400 megatonnes (Mt) yr-1 of terrigenous clastic sediment is

supplied to the GoP by the Fly-Kikori-Purari rivers, a sediment load in size exceeding

that from the Mississippi River [Milliman and Syvitski, 1992; Milliman, 1995; Wolanski

and Alongi, 1995; Milliman et al., 1999]. However, this load is deposited in a basin just

one fourth the size of the Mississippi River basin. Recent studies indicate that a majority

(up to 75%) of the sediment exiting these rivers is deposited within river floodplains and

mangrove intertidal zones [Walsh et al., 2004]. The remainder of this sediment is deposited in water depths shallower than 60 m, resulting in an inner shelf clinoform that is rapidly prograding across older, late Holocene deposits [Harris et al., 1996; Walsh and

Nittrouer, 2003; Walsh et al., 2004].

68 Inner-shelf sediment deposits account for approximately 20% of the modern Fly-

Kikori-Purari river sediment load [Walsh et al., 2004]. The maximum short-term accumulation of sediment (up to 4 cm yr-1) is found on clinoform foresets in water depths between 20 and 45 m, whereas very little sediment accumulates seaward of the 60m isobath [Walsh et al., 2004; Palinkas et al., 2006]. Sediment deposited on the inner shelf clinoform forms a prograding clastic wedge that is slowly moving toward the northeast as a result of seasonal changes in currents driven by spring tides, wind, and thermohaline currents [Wolanski and Alongi, 1995; Walsh and Nittrouer, 2003; Walsh et al., 2004;

Slingerland et al., 2007]. In contrast, siliciclastic sediment accumulation rates are minimal on the middle and outer shelf as evidenced by the presence of relict carbonate sands and Halimeda banks [Alongi and Robertson, 1995; Brunskill et al., 1995; Harris et al., 1996].

Approximately 3.3% of inner-shelf sediment, or ~1x107 tons yr-1, is transported to the outer shelf-edge where it may episodically spill onto the upper slope as the shelf accommodation space narrows significantly in the vicinity of the Kerema Canyon

[Figure 3.1; Brunskill, 2004; Walsh et al., 2004]. A combination of hemipelagic depositional and nepheloid layer transport processes are thought to be the dominant sediment transport mechanisms [Walsh and Nittrouer, 2003; Muhammad et al., 2007].

Siliciclastic sediment that is liberated from the shelf-edge is mixed with calcium carbonate rich sediment that has been exported off carbonate platforms. Added to this is a calcareous neritic marine plankton sedimentary component that has settled into deep- water environments [Francis et al., 2007]. Unlike the relatively well-studied shallow

69 water depositional environments in the GoP, very few studies have examined the

composition of sediment deposited in deep-water or mapped its distribution.

Figure 3.1. General Map of the Gulf of Papua (GoP) with labeled major physiographic features, bathymetry, and sample locations. The green shaded region highlights the area of the inner shelf clinoform (shallower than 60m) where most recent sediment is deposited. The contour interval is 20 m to -100 m, and then 500 deeper than 1000 m. Contour lines between the -100 and -1000 m isobaths are the 200 and 500 m isobath, respectively. Bathymetry from Daniell, [2007].

3.2.2. Seafloor Physiography

The GoP (Figure 3.1) is a complex basin composed of a variable gradient shelf and a series of troughs and carbonate pinnacles [see Francis et al., 2007]. In general, the shelf changes from a broad and low-gradient shelf (1:1000) seaward of the Turama River

70 to a narrower and higher gradient shelf (1:133) up dip of the Kerema Canyon. The

Pandora Trough is a ~2000 m deep basin exhibiting a NE-SW trend beyond the central and northeastern shelf edge of the GoP. Unfilled paleo channels extend from the shelf- edge and onto the slope as evidenced by irregular inlets in the 100 m isobath [Harris et al., 1996; Francis et al., 2007; Slingerland et al., 2007]. To the west and adjacent to the

GBR is the Ashmore Trough, a ~1000 m deep basin separated from the Pandora Trough by the Ashmore Reef. The Portlock Reef extends outward from the modern shelf-edge and is perched between the Ashmore and Pandora Trough. The GoP deepens to >2000 m in the Moresby Trough that ultimately leads into the deeper Coral Sea Basin.

3.2.3 Regional Geology

Sediment lithologies discharged from the Fly-Kikori-Purari rivers generally reflect the hinterland geology of their drainage basins. Mature siliciclastic sediments with a high illite:smectite ratio and detrital carbonate content (~60-70%) characterize the sediment load of the Fly and Kikori rivers [Brunskill, 2004; Slingerland et al., In Press].

However, much less of the Purari River sediment load (~20%) is composed of carbonate, whereas 30-35% of sediments from the Kikori and Purari rivers are volcaniclastic. These differences reflect the carbonate-dominated folded sedimentary rocks in the Fly drainage basin and an increasing amount of extrusive igneous rocks in parts of the Kikori and

Purari river basin [see Chapter 2; Rickwood, 1968; Davies and Smith, 1971].

3.3. Materials and Methods

3.3.1. PANASH and PECTEN Cruises

Data presented in this study are based on 68 measurements of surficial sediments taken from selected cores collected during the PANASH and PECTEN cruises of the R/V

71 Melville and R/V Marion Dufresne (Table 3.1). The PANASH cruise (Pandora-

Ashmore) took place in the GoP from March to April, 2004, during which time three

gravity cores, two box cores, 30 multi-cores, and 33 jumbo piston cores were collected

primarily in outer shelf edge and trough settings. The PECTEN cruise (Past Equatorial

Climate: Tracking El Niño) took place during June to July, 2005 on board the R/V

Marion Dufresne and collected 16 jumbo CALYPSO cores, six CASQ cores, and one

gravity core within the GoP study area.

3.3.2. Surficial Sediments

A set of 68 undisturbed core-tops (0-2cm) were selected for analyses from the

PANASH and PECTEN core collections. Multi-core, gravity core, box core, and CASQ

core-tops contain the least disturbance in the upper two cm, whereas jumbo piston core

tops may potentially have several centimeters (cm) of loss due to over-penetration. We

excluded piston core tops where they overlapped with multi-core locations, and when

possible, trigger core-tops were used. Samples collected from piston core tops were

carefully inspected for sediment loss by ensuring that they contained characteristics of

surface sediments. The selection criteria we used were 1) the sediment should be water

charged, and 2) sediment color should be lighter than downcore segments to indicate

surface oxidation.

3.3.3. Magnetic Susceptibility (MS) Measurements

MS measurements are a simple, quick, and nondestructive way to determine the

relative abundance of iron-bearing detrital minerals within a sediment or rock sample.

MS values record the strength of induced magnetism experienced by mineral grains when placed in an applied magnetic field. We measured the bulk low-field MS on dried

72 sediment samples using a susceptibility bridge in a magnetically shielded room at the

LSU Rock Magnetism Laboratory [Ellwood et al., 2006]. The LSU Bridge uses a balanced coil induction system that is sensitive to <1 (±0.2) x 10-9 m3 kg-1 [Ellwood et al.,

1996] and calibrated using values for standard salts reported by Swartzendruber [1992].

We report the mean of three measurements for each sample (Table 3.1).

3.3.4. Calcium Carbonate

Calcium carbonate content was analyzed in surficial sediment samples using both wet chemical HCl acid digestion and the gasometric technique at the LSU Rock

Magnetism Laboratory. Multi-core samples were sampled for 0.18 g dry weight and then treated with 50 ml of HCl. The weight loss after 1 hr of digestion was used to calculate total percent carbonate.

Other sediment samples were treated with the gasometric technique [Jones and

Kaiteris, 1983]. This method uses 5 ml of phosphoric acid and 0.25 g of dry, powdered sediment. Triplicate carbonate measurements on several samples resulted in a precision of +/- 0.10%. Tests of reagent grade (99%) carbonate and a mixed sample standard (95% feldspar 5% carbonate) gave accurate results to +/- 1%.

3.3.5. Sedimentary Organic Matter

Organic matter in core sediments was measured using two independent analyses.

Total organic carbon (TOC) was measured on sediment samples using a LECO Carbon analyzer, while Rock-Eval Pyrolysis (REP) was measured on samples using the Rock-

Eval II system. Both measurements were made at Baseline Resolution Analytical Labs, in Houston Texas. Analytical errors for TOC are ±0.05 wt. % and for REP measurements they are ± 0.02 mg of HC per gram of TOC.

73 TOC is a basic parameter measured to quantify the amount of organically-bound

carbon in a sample, whether it is of marine or terrestrial plant origin [Tissot and Welte,

1984; Jarzie, 1991]. The origin of sedimentary organic matter is examined through two

REP parameters, the amount of free (S1) and pyrolytic (S2) hydrocarbon (HC) preserved in sediment samples [Meyers, 1997; Rullkötter, 2000; Wagner, 2000]. In particular, the

hydrogen index (HI in Equation 1) is calculated using TOC and S2 numbers and is given as:

HI = 100xS2/TOC (1)

HI is the mass of S2 hydrocarbons per gram of TOC and is determined by heating

kerogen in a sample until the maximum amount of nonvolatile and volatile hydrocarbons

is evolved [Espitalié and Bordenave, 1993].

Hydrogen index values range from near 0 to >400 and may help to distinguish if

sediments are dominated by vascular land plant matter (≤150 mg of HC per gram of

TOC), or dominated by algal-rich marine organic matter (200 to 400 mg of HC per gram

of TOC). Sediment with hydrogen index values ≤150 mg of HC per gram of TOC are

characterized as Types 3 and 4 kerogen [Tissot and Welte, 1984; Meyers, 1997]. Organic

matter rich in marine plankton are Type 2a kerogen and are in the range of 200-400 mg

of HC per gram of TOC [Rullkötter, 2000].

Tmax is the temperature above which a sample must be heated before most of the

hydrocarbons are evolved. Tmax values of recent or “immature” organic matter are

<430°C, whereas organic matter is “overmature” (>430°C) if derived from reworked

fossil organic matter such as coal [Delvaux et al., 1990].

74 3.3.6. Organic Matter Petrography

Acid insoluble organic matter was optically examined to determine the major

types of kerogen particles contained in selected surface sediment samples and to constrain geochemical measurements [Teichmüller, 1986; Tyson, 1995; Batten, 1996;

Meyers, 1997; Omura and Hoyanagi, 2004; Sebag et al., 2006]. Standard total kerogen processing techniques were utilized in sample preparation [e.g. Tyson, 1995; Oboh-

Ikuenobe and Villiers, 2003]. Approximately 5 g of dry sediment from ten samples were acidified in 10% HCl until reaction ceased, and then acidified in concentrated HF overnight. After complete acidification, samples were then neutralized with distilled

water and reacidified with concentrated HCl to remove any calcium fluoride crystals that

formed during reaction with HF. The remaining residue was diluted with distilled water,

centrifuged, and this process repeated until a neutral pH was achieved. Three permanent

kerogen slides were made for each sample using glycerine jelly. Kerogen residues were

not sieved or oxidized prior to slide making because those steps can remove most of the amorphous organic matter and bias the total kerogen assemblage [Oboh-Ikuenobe and de

Villiers, 2003].

Organic particles were counted using a Leitz Dialux-20 micropscope with a 40X

objective. The identification criteria for counting follows Hart [1986] and total kerogen

categories such as % amorphous organic matter, % phytoclasts, and % marine

palynomorphs are reported [e.g. Tyson, 1989; Tyson, 1993; Table 3.2]. A minimum of

300 particles were counted over several traverses and two slides per sample were

examined to minimize within-sample heterogeneity.

75 Kerogen preservation state was examined using fluorescence microscopy [Van

Gizel, 1979; Tyson, 1995]. After point counts were completed, total kerogen slides were

subjected to incident blue-light excitation and examined using a Leitz Ortholux-II

fluorescing microscope and a 40x NPL Fluotar objective.

3.3.6.1. Phytoclast Group

Phytoclasts are organic particles that are of terrestrial plant origin such as leaf cuticles and tree cortex. Seven subcategories are differentiated in this main category and

include fungal hyphae, well and poorly preserved phytoclasts, amorphous structured and

nonstructured phytoclasts, and opaque particles (Figure 3.2, 1a-h). Amorphous

structured and nonstructured phytoclasts correspond to reddish-brown, highly degraded

or “gelified” plant particles [Hardy, 2000; Sebag et al., 2006]. In reality, amorphous

structured and nonstructured are better described as degraded vs. destroyed, respectively

[Hardy, 2000]. Opaque phytoclasts are dark brown to black and represent extensively

oxidized wood or coal particles [Tyson, 1995; Hardy, 2000; Oboh-Ikuenobe and de

Villiers, 2003; Sebag et al., 2006].

3.3.6.2. Palynomorph Group

Palynomorphs are complete organic-walled microfossils derived from pollen and

spores (sporomorphs), phytoplankton such as dinoflagellate cysts, and are the organic

resistant remains of zooplankton, foraminifera, and annelid worms. Zooclasts are the

recognizable remains of planktonic and benthic marine metazoans (Figures 3.2, 2a-f).

Zooclasts identified in this study include foraminiferal organic linings, annelid worm jaw

parts (scolecodonts), and zooplankton egg-cases [see also Van Waveren, 1992]. Because

76 zooclasts are strictly of marine origin and dinoflagellate resting cysts are exceedingly rare

to absent, zooclasts are combined in the marine “paly” category for data analysis.

Figure 3.2. Acid insoluble organic particles identified in kerogen slides. 1, Phytoclast group: (1a) fungal hyphae surrounded by amorphous organic matter (AOM) (1b) well- preserved phytoclast (1c) poorly-preserved phytoclast with partial gelification (1d) amorphous structured phytoclast (1e-g) nonstructured amorphous phytoclast (1h) opaque phytoclast. 2, Palynomorph group: (2a) pollen grain (2b) trilete spore (2c) copepod egg case (2d) foraminiferal organic lining (2e) fecal pellet containing and surrounded by AOM (2f) scolecodont. 3, AOM surrounding a foraminiferal organic lining. All images to same scale, scale bars = 125µm.

77 3.3.6.3. Amorphous Organic Matter Group

Amorphous organic matter (AOM; Figure 3.2, 3) is a structureless organic constituent that is usually yellow-brown to grey in color and has a thin, flaky to microparticulate and globular morphology. AOM, or “sapropel”, is the end product of accumulating sinking particulate organic matter or “marine snow” that has undergone extensive bacterial degradation [Alldredge, 1979; Alldredge and Silver, 1988; Tyson,

1995]. Carbon isotope analysis of AOM indicates that it is derived from autochthonous marine sources [Lewan, 1986].

3.3.6.4. Absolute Densities

Absolute kerogen abundance was determined by relating the percentage of

kerogen types to the TOC wt. % of each sample [e.g. Hart, 1986; Hardy, 2000].

Equation 2 was used to convert the percent of AOM, phytoclasts, and marine palynomorphs (including zooclasts) to grams of TOC for each category, where

×TOCX X oc = (2) 100

and Xoc = the wt. % of organic kerogen Type X and X = relative percentage of organic

kerogen type.

3.3.7. Map Bathymetry

Several maps are presented in this paper to illustrate the concentrations and

distributions of MS, calcium carbonate, and TOC concentrations. All maps were

constructed using ArcGIS 9.1 and then sediment properties were hand-contoured on base

maps using Corel Draw 9.0. The bathymetry contours for the GoP are from a gridded

78 dataset that utilizes data from numerous sources including multibean sonar coverage collected during the PANASH and PECTEN cruises [see Daniell, 2007].

3.4. Results

3.4.1. Magnetic Susceptibility Measurements from Surface Sediments

Surficial sedimentary MS values are shown in Figure 3.3. In general, MS values

exhibit a gradient from highest values (6-10 x 10-7 m3kg-1) on the upper slope of the

central Pandora Trough to lowest values in the Ashmore Trough and south of Eastern

Field’s Reef. A zone of moderately high values rims the upper slope from the central to

the northern Pandora Trough and also one station in the Moresby Trough (29JPC).

Deeper and seaward in the GoP, moderate values (1-4 x 10-7 m3kg-1) extend from and the

northern to the southern Pandora Trough, into the Moresby Trough and around to the

Eastern Plateau.

3.4.2. Calcium Carbonate Concentrations in Surface Sediments

Calcium carbonate concentrations display a seaward increasing gradient

from low to high concentrations (Figure 3.4) and are approximately the inverse

distribution of the MS values. Lowest concentrations (<30%) are found on the upper

slopes of the Pandora Trough, which also wrap around the margin of the Papuan

Peninsula and into the eastern portion of the Moresby Trough. A transition or inter-

fingering zone characterizes the central and southern Pandora Trough from moderately

low (30-49%) to moderately high calcium carbonate concentrations. Highest

concentrations are found in the basin to upper slopes of the Ashmore Trough.

79

Figure 3.3. Distribution of magnetic susceptibility (MS) values in surface sediment from the GoP.

Figure 3.4. Distribution of calcium carbonate concentrations in surface sediment from the GoP.

80 Two exceptions to this general pattern are found on the shelf-edge (<100 mwd) of

the eastern margin of the Papuan Peninsula (Figure 3.4). Higher calcium carbonate concentrations are found at 43MC (54%) and slightly higher at 44MC (33%) compared to other adjacent deeper cores that have <30% calcite.

3.4.3. TOC Concentrations from Surface Sediment

TOC concentrations are low (≤1 wt. %) throughout the GoP, but they still display

a high (landward) to low (seaward) distribution similar to MS and inverse of calcium

carbonate. Highest concentrations (>0.8 wt. %) are observed in core-tops taken from the

upper slope of the northern Pandora Trough and form a bi-lobed distribution extending

south and southwestward to deeper water depths (Figure 3.5).

Figure 3.5. Distribution of total organic carbon (TOC) concentrations in surface sediment from the GoP.

81 Surrounding this and deeper into the Pandora and Moresby Troughs is an area of

low to moderate TOC concentrations (0.5-0.79 wt. %). A region of low organic carbon

content rims the upper slope of the central Pandora Trough and then extends into the

southern Pandora Trough and throughout the Ashmore Trough. Lowest values are also

found in a single core-top from the Moresby Trough (29JPC) and also the outer shelf-

edge of the eastern Papuan margin (43MC).

3.4.4. RE Pyrolysis Measurements from Surface Samples

HI values of sedimentary organic matter in core-tops are generally low (28 to 346

mg of HC per gram of TOC; mean 125, n=68) and contain low S2 pyrolytic yields

(Figure 3.6a). HI and S2 values indicate hydrogen-poor organic carbon that is typical of terrestrial organic carbon such as Type 3 kerogen or a mixture of Type 3 and oxidized

Type 2 marine organic matter (Figure 3.6b). Low S2 yields are typical of all kerogen regardless of the TOC content (Figure 3.6a); however, highest HI values are

characterized by moderate to low TOC concentrations (Figure 3.6b).

The elevated HI values correspond to TOC depleted sediments found in the

carbonate-rich southern Pandora and Ashmore Troughs (Figure 3.5, Table 3.1). This is consistent with a reduced detrital input and greater marine influence, and to higher sedimentary organic matter levels in this region (i.e. low MS, Figure 3.3). In addition, recent marine HI values are generally shifted higher compared to late Pleistocene values observed in MV-51JPC (Figure 3.6b, see also Chapter 2). However, it is difficult to accurately determine the organic matter type when S2 yields are < 1 mg of HC per gram of TOC, because they may be the result of mineral matrix dilution of hydrogen-poor hydrocarbons [Espitalié et al., 1980; Katz, 1983; Langford and Blanc-Valleron, 1990].

82 Tmax measurements for core-top sediments contain immature organic carbon, as

expected for recent marine sediment. All core-tops are <430°C and range from 296 to

428°C (mean = 390, n=68, Table 3.1.). In contrast, late Pleistocene sediment contains

six samples with higher Tmax values, indicative of older, more mature carbon exported to the basin during lower sea level stands (Figure 3.6b, see also Chapter 2).

Figure 3.6. Rock-Eval 2 data from core-tops, plotted in an S2 versus TOC diagram (6a, after Langford and Blanc-Valleron, 1990) and a Van Krevelen diagram (6b, after Wagner, 2000). Boundaries in the Van Krevelen diagram between kerogen Types 2, 3, and 4 are indicated with dashed lines and are derived from reference values reported in Delvaux et al. [1990]. Hydrogen index (HI) values for MV-51 (Chapter 2) are plotted in Figure 6b for comparison of recent marine sediments with late Pleistocene samples.

83

Table 3.1. Data for all stations presented in this paper. Geographic coordinates are given in decimal degrees and water depths are listed in meters below sea level. Magnetic susceptibility 3 (MS) data are reported in m /kg and both CaCO3 and TOC concentrations are given as weight percentages of dry sediment. Rock-Eval II pyrolysis data are presented for the hydrogen index (HI, mgHC per g TOC), thermal maximum (Tmax, °C), and hydrocarbon sources S1 and S2 (mg HC per g sediment) and S3 (mg CO2 per g sediment). BC=box core, MC=multi-core, JPC=jumbo piston core, TC=trigger core, C2=Casq core (R/V Marion Dufresne).

Coretop Long Lat Depth CaCO3 MS TOC S1 S2 S3 Tmax HI 1 BC 144.00 -10.31 342 94.457 5.97E-09 0.18 0.05 0.39 1.09 418 216.67 2 BC 144.16 -10.31 661 88.919 4.09E-08 0.31 0.09 0.54 1.58 408 174.19 10 MC 144.48 -9.87 584 80.554 2.92E-08 0.37 0.15 0.84 1.81 418 227.03 12 MC 144.46 -9.81 372 90.5 4.65E-08 0.44 0.16 1.27 1.60 417 288.64 14 MC 144.66 -9.91 760 80.032 6.72E-08 0.42 0.15 0.83 2.21 405 197.62 16 MC 144.74 -9.74 686 74.657 1.36E-07 0.40 0.09 0.51 2.12 414 127.50 18 MC 145.51 -8.51 267 39.209 5.00E-07 0.42 0.08 0.33 1.72 366 78.57 19 MC 145.51 -8.52 229 24.115 5.42E-07 0.21 0.05 0.10 0.97 314 47.62 20 MC 145.02 -9.29 661 57.667 1.99E-07 0.35 0.08 0.32 1.92 379 91.43 21 MC 145.06 -9.36 1001 65.242 1.47E-07 0.43 0.13 0.52 2.41 379 120.93 24 MC 146.25 -9.79 2102 41.131 1.99E-07 0.51 0.14 0.43 2.75 354 84.31 26 MC 146.38 -10.01 2232 38.748 2.24E-07 0.51 0.15 0.40 3.07 372 78.43 28 MC 147.13 -10.17 2426 34.498 2.47E-07 0.55 0.15 0.46 3.06 349 83.64 30 MC 146.17 -9.67 2023 44.279 2.23E-07 0.55 0.16 0.56 2.83 374 101.82 32 MC 145.35 -9.55 1620 56.431 1.76E-07 0.51 0.16 0.73 2.68 378 143.14 38 MC 145.35 -9.01 961 47.859 2.23E-07 0.56 0.16 0.58 3.01 372 103.57 39 MC 145.61 -8.61 690 42.519 2.06E-07 0.74 0.19 0.79 3.15 384 106.76 42 MC 145.84 -8.37 91 27.75 4.77E-07 1.14 0.20 1.11 2.94 402 97.37 43 MC 146.40 -8.91 62 54.586 2.77E-07 0.46 0.09 0.36 1.82 368 78.26 44 MC 146.19 -8.61 99 33.701 5.35E-07 0.61 0.12 0.50 2.21 382 81.97 47 MC 145.83 -8.43 399 31.408 2.19E-07 1.01 0.18 1.01 2.90 415 100.00 50 MC 146.05 -8.59 795 27.634 2.76E-07 1.00 0.24 1.09 4.18 372 109.00 53 MC 145.94 -8.73 1111 31.504 2.61E-07 0.80 0.21 0.85 3.38 355 106.25 56 MC 145.23 -8.94 450 19.779 1.15E-06 0.25 0.02 0.07 1.13 353 28.00 59 MC 144.92 -9.39 577 40.765 6.82E-07 0.27 0.03 0.15 1.41 383 55.56 60 MC 145.81 -9.24 1671 47.006 2.17E-07 0.52 0.15 0.62 2.65 373 119.23 65 MC 145.35 -10.47 1012 76.033 6.52E-08 0.24 0.07 0.27 1.94 296 112.50 67 MC 144.90 -9.58 1147 71.981 1.55E-07 0.38 0.11 0.51 2.36 384 134.21 69 MC 145.87 -9.98 1638 69.746 1.10E-07 0.50 0.15 0.81 2.51 386 162.00 70 MC 145.15 -9.78 1758 61.711 1.47E-07 0.46 0.13 0.56 2.53 371 121.74 72 MC 144.69 -10.25 1320 80.936 6.91E-08 0.33 0.09 0.46 2.00 406 139.39 76 MC 144.86 -10.44 1551 67.287 1.14E-07 0.34 0.10 0.46 2.30 375 135.29 7 JPC 144.21 -10.44 773 82.31 3.58E-08 0.38 0.29 0.75 2.31 424 197.37 9 JPC 144.48 -9.87 574 82.26 3.29E-08 0.36 0.25 0.66 2.25 424 183.33

84

Table 3.1. Continued Coretop Long Lat Depth CaCO3 MS TOC S1 S2 S3 Tmax HI 13 JPC 144.46 -9.81 376 85.84 4.46E-08 0.35 0.21 0.74 1.86 423 211.43 22 JPC 146.47 -9.77 2063 20.96 2.51E-07 0.65 0.31 0.39 3.78 369 60.00 23 JPC 146.00 -9.88 2073 32.97 2.28E-07 0.51 0.31 0.37 3.70 386 72.55 25 JPC 146.56 -9.93 2199 21.95 1.76E-07 0.57 0.34 0.41 3.63 378 71.93 27 JPC 146.68 -10.17 2077 27.17 1.40E-07 0.51 0.32 0.47 3.63 386 92.16 29 JPC 146.63 -10.03 2295 14.91 4.58E-07 0.38 0.20 0.28 1.82 387 73.68 31 JPC 146.07 -9.56 1964 21.77 2.01E-07 0.54 0.29 0.35 3.16 368 64.81 33 JPC 145.60 -9.87 1795 44.13 1.39E-07 0.49 0.31 0.44 3.19 394 89.80 37 JPC 145.19 -9.24 1011 45.62 2.01E-07 0.51 0.37 0.61 2.96 412 119.61 48 JPC 145.55 -8.56 494 24.61 5.72E-07 0.47 0.23 0.46 2.12 395 97.87 49 TC 145.83 -8.58 859 20.49 1.96E-07 0.70 0.37 0.61 2.85 391 87.14 51 JPC 146.02 -8.77 804 5.1437 7.93E-08 0.73 0.42 0.56 2.51 393 76.71 51TC 146.02 -8.77 804 17.047 2.21E-07 0.85 0.52 0.82 3.41 405 96.02 52 JPC 145.94 -8.73 1110 15.98 1.83E-07 0.88 0.42 0.70 3.47 390 79.55 54 JPC 145.39 -8.94 923 20.94 1.73E-07 0.64 0.33 0.47 3.32 389 73.44 54TC 145.39 -8.94 923 31.149 1.71E-07 0.60 0.34 0.46 3.55 398 76.79 55 JPC 145.24 -8.94 405 8.53 8.04E-07 0.52 0.25 0.35 2.05 395 67.31 57 JPC 145.02 -9.29 654 49.17 2.51E-07 0.31 0.24 0.39 2.19 392 125.81 58 JPC 144.91 -9.41 415 29.73 6.14E-07 0.22 0.12 0.12 1.43 410 54.55 61 JPC 145.97 -9.00 1616 19.76 2.47E-07 0.66 0.38 0.62 3.34 399 93.94 63 TC 146.25 -10.62 1668 1.87E-07 0.57 0.67 1.00 3.13 410 175.44 64 JPC 145.51 -10.37 994 69.61 8.31E-08 0.25 0.24 0.29 2.27 390 116.00 66 JPC 145.15 -10.00 1793 50.74 1.12E-07 0.43 0.31 0.44 2.62 372 102.33 68 JPC 144.91 -9.58 1142 63.27 7.51E-08 0.45 0.22 0.40 2.51 398 88.89 71 JPC 145.02 -9.69 1593 52.66 9.11E-08 0.49 0.30 0.53 3.12 388 108.16 73 JPC 144.42 -9.78 137 87.85 3.62E-08 0.06 0.33 1.04 2.05 418 236.36 74 JPC 144.08 -10.73 684 87.59 1.40E-08 0.44 0.03 0.02 0.88 402 33.33 75 JPC 144.02 -10.71 530 88.59 1.58E-08 0.41 0.50 1.42 2.08 420 346.34 6 GC 144.21 -10.44 775 81.88 2.96E-08 0.33 0.32 0.80 2.07 427 242.42 8 GC 144.48 -9.87 594 83.66 3.07E-08 0.34 0.39 0.89 2.08 420 261.76 11 GC 144.46 -9.81 377 88.09 4.53E-08 0.47 0.46 1.41 2.09 428 300.00 2932 C2 146.01 -8.74 837 18.484 3.04E-07 0.82 0.63 1.05 3.19 400 128.05 2933 C2 145.88 -8.33 71 5.7973 3.81E-07 0.78 0.56 0.92 2.84 412 117.95 2951 C2 144.66 -9.90 765 70.224 9.13E-08 0.53 0.43 1.10 2.84 413 207.55

Statistical tests of TOC, calcium carbonate, and HI values demonstrate strong and significant correlations among variables. TOC and calcium carbonate concentrations are inversely correlated (R=-0.58, n=68, P<0.001) whereas HI values and calcium carbonate are positively correlated (R=0.74, n=68, P<0.001; Figure 3.7).

85

Figure 3.7. Cross-plot of core-top TOC vs. calcium carbonate (R=0.58, n=68, significant at P<0.001) and hydrogen index (HI) vs. calcium carbonate R=0.74, n=68, significant at P<0.001).

3.4.5. Organic Petrography

All sediment samples examined are dominated by AOM kerogen and contain

fewer amounts of phytoclasts and palynomorphs (Tables 3.2, 3.3). AOM ranged from

62-87% (mean 76%) regardless of water depth or location. A strong and significant

correlation exists for AOM abundance and TOC (Figure 3.8a, Table 3.4). Phytoclasts

ranged from 10-34% (mean 21%) in kerogen residues and are also significantly

correlated with TOC content. Phytoplankton were the least abundant kerogen

components ranging from 0.6 to 8%, the majority of which are zooclasts. Dinocysts and

other identifiable algal remains are essentially absent from surface sediments.

Terrestrialy-derived palynomorphs, such as pollen and spores, comprise <3% of organic

components.

Phytoclasts are dominated by degraded amorphous nonstructured constituents and

recalcitrant opaque phytoclasts. Fewer than 3% were well to poorly preserved

86 phytoclasts, which is indicative of limited terrigenous plant input. On a ternary diagram, all kerogen data plot near the % AOM apex (Figure 3.8b). Data plotting at the AOM apex, or in region 4, are indicative of basinal palynofacies dominated by marine organic matter under oxic to dysoxic conditions [Tyson, 1993].

Table 3.2. Raw data for kerogen counts in core-top samples. Kerogen categories are explained in section 3.6 and illustrated in Figure 3.2.

Palynomorph Group Phytoclast Group AOM Core-top Phytoplankton Zoomorphs Spores & pollen fungal hyphae Well- preserved Poorly- preserved Amorphous structured Amorphous Nonstructured Opaque Phytoclasts AOM Total Non AOM Total kerogen

43 MC 13 1 4 1 0 0 5 26 30 226 80 306 33 C2 4 4 10 2 0 1 4 15 38 229 78 307 42 MC 3 2 2 1 0 1 1 27 35 251 72 323 42 MC 2 2 2 0 1 3 4 18 53 218 85 303 44 MC 2 1 2 3 0 0 2 7 22 261 39 300 47 MC 8 7 11 5 0 0 6 47 94 300 178 478 39 MC 2 8 3 0 0 1 3 21 17 253 55 308 51 C2 19 7 2 0 0 2 2 17 26 233 75 308 51 C2 13 6 4 3 1 3 8 50 123 550 211 761 51 JPC 1 2 13 3 0 0 7 60 69 302 155 457 32 C2 6 5 2 1 0 2 0 3 24 257 43 300 32 C2 4 2 5 2 0 0 1 17 32 239 63 302 63 TC 1 1 2 1 0 0 1 17 40 242 63 305

Table 3.3. Percentage data and counting error for kerogen groups. The counting error is 1 standard deviation of duplicate counts on two separate slides.

Core Marine Pollen Structured top Paly error +/- AOM error +/- Terrestrial error +/- &Spores Phytos Opaques 43 MC 4.58 73.86 21.57 1.31 10.13 9.80 33 C2 2.61 74.59 22.80 3.26 6.51 12.38 42 MC 1.55 0.16 77.71 4.07 20.74 4.24 0.62 8.98 10.84 42 MC 1.32 71.95 26.73 0.66 8.58 17.49 44 MC 1.00 87.00 12.00 0.67 3.00 7.33 47 MC 3.14 62.76 34.10 2.30 11.09 19.67 39 MC 3.25 82.14 14.61 0.97 8.12 5.52 51 C2 8.44 75.65 15.91 0.65 6.82 8.44 51 C2 2.50 4.20 72.27 2.39 25.23 6.59 0.53 8.15 16.16 51 JPC 0.66 66.08 33.26 2.84 14.66 15.10 32 C2 3.67 85.67 10.67 0.67 1.67 8.00 32 C2 1.99 1.19 79.14 4.62 18.87 5.80 1.66 5.96 10.60 63 TC 0.66 79.34 20.00 0.66 5.90 13.11

87

Table 3.4. Kerogen absolute abundance data. Percent data from Table 3.3 and TOC from Table 3.1 were used in Equation 2. Errors were calculated using the 1 standard deviation from Table 3.3.

Core-top g Marine Paly error +/- g AOM error +/- g Phytoclasts error +/- 43 MC 0.02 0.34 0.10 33 C2 0.02 0.58 0.18 42 MC 0.02 0.00 0.89 0.03 0.24 0.03 42 MC 0.02 0.82 0.30 44 MC 0.01 0.53 0.07 47 MC 0.03 0.63 0.34 39 MC 0.02 0.61 0.11 51 C2 0.04 0.02 0.40 0.01 0.08 0.02 51 C2 0.01 0.38 0.13 51 JPC 0.00 0.48 0.24 32 C2 0.03 0.01 0.70 0.03 0.09 0.03 32 C2 0.02 0.65 0.15 63 TC 0.00 0.45 0.11

Figure 3.8. Kerogen abundance data. (a) Absolute kerogen data vs. sample TOC. Vertical error bars are based on two sample count replicates and the horizontal error bar is the analytical error for TOC measurements (Table 3.4). The AOM abundance is positively correlated with whole sample TOC wt. % (R=0.94, n=10, P<0.001). The phytoclast kerogen component is also correlated with whole sample TOC wt. % (R=0.77, n=10, P=0.02). (b) Ternary diagram of major kerogen categories [adapted from Tyson, 1993]. Sub-fields on the palynofacies diagram correspond to: (1) source proximal environment such as river delta or inner-shelf; (2) proximal oxic shelf; (3) distal oxic shelf; (4) distal basinal facies.

88 AOM particles are microparticulate and vary from thin, discrete flakes to large

grumose pieces (Figure 3.2). Surprisingly, almost all of the kerogen is nonfluorescent

with the exception of rare autofluorescing particles within AOM, which would not be

expected if marine material dominated the organic matter [Batten, 1996; Omura and

Hoyanagi, 2004]. However, strong oxidation of labile organic compounds in kerogen

can destroy the fluorescent characteristics of marine-derived constituents, such as AOM

[Tyson, 1995]. The fecal pellet (Figure 3.2, 2f) is microparticulate and similar in color to other typical AOM, suggesting that some AOM particles may be components liberated

from fecal excreta. Unfortunately, it is difficult to determine whether the provenance of

the fecal pellets is from zooplankton or from detritus-feeding benthic invertebrates.

3.5. Discussion

Based on the distribution of MS, calcium carbonate, and TOC, there is a clear

particle gradient from the shelf-edge and upper slope near the Kerema Canyon to the

Ashmore Trough. MS values are a function of the mineralogy and/or abundance of

detrital siliciclastic components [deMenocal et al., 1991; Ellwood et al., 2000; Ellwood et

al., 2007; see also Chapter 4]. The MS distribution pattern clearly demonstrates a change

from high to low values on the source proximal shelf-edge to the Ashmore Trough. In

contrast, calcium carbonate concentrations are lowest in regions of high MS and become

much greater in the Ashmore Trough where clastic input is least and pelagic deposition

dominates. The origin of part of the calcium carbonate in core-top sediments is from

neritic settling of coccolithophores, planktonic foraminifera, and pteropods. However,

some fraction may also come from coralgal debris reworked by wave erosion off the

89 Portlock and Ashmore pinnacle reefs as it did during the early Holocene [e.g. Carson et al., 2007; Francis et al., 2007].

TOC concentrations demonstrate the clearest trend from high to low concentrations, beginning on the upper slope of the Kerema Canyon and decreasing to the

Ashmore Trough. The distribution of TOC is consistent with previous studies of geochemical distributions, transport models, and accumulation rates [Bird et al., 1995;

Brunskill et al., 1995; Brunskill et al., 2003; Walsh and Nittrouer, 2003; Keen et al.,

2006; Muhammad et al., 2007]. However, part of the distribution may also be a consequence of enhanced preservation of total organic carbon contents in regions with higher accumulation rates. For example, Muhammad et al., [2007] also observed a similar pattern in which highest accumulation rates were found on the upper slope of the

Kerema Canyon and then decreasing seaward. Burial of organic carbon has been demonstrated to be strongly correlated with the flux of organic carbon to the sediments

[Müller and Seuss, 1979] and sedimentation rates [Heath et al., 1977; Emerson and

Hedges, 1988].

Regardless of total preservation, degradation of labile organic compounds is likely the result of oxidation within the zone of extensive bioturbation [Stein, 1990;

Hedges and Keil, 1995; Hartnett et al., 1998; Hedges et al., 1999]. Figure 3.9 shows

TOC profiles in the upper 12 cm of three representative multi-cores ranging in water depth of 91 m to 2023 m. Except for the core-top of 42MC, data within each profile do vary significantly and are within the range of analytical error, which is indicative of effective oxidation and sediment mixing.

90 Rock Eval (RE) pyrolysis data and kerogen palynofacies data seem at odds with one another. For example, the RE data indicate that organic residues in the sediments are primarily Type 3, or derived from a botanical provenance. However, kerogen residues are dominated by AOM (algal or bacterial in origin) and much fewer amounts of highly degraded phytoclasts. Two things can explain the apparent differences between the geochemical and optical data. First, almost all of the kerogen constituents lack autofluorescence, a sign of severe oxic degradation that can lead to hydrogen index data

<100 mg of HC per gram of TOC [Tuweni and Tyson, 1994]. Second, HI data can be

lowered by oxidation of organic matter, principally because hydrogen atoms are lost from

the hydrocarbons, leaving behind hydrogen-poor residual carbon (see also chapter 2).

TOC, wt. % 0.45 0.55 0.65 0.75 0.85 0.95 1.05 1.15 0 1 2 3 4 5 6 7 Depth, cm 8 9 30 MC 10 42 MC 11 47 MC 12

Figure 3.9. Distribution of TOC in the upper 12 cm of three multi-cores from the GoP (Table 3.1). Selected multi-cores (Figure 2.1) range in water depth from 91 m (42MC, Kerema Canyon), 399 m (47MC, Kerema Canyon), and 2023 m (30MC, Pandora Trough). Only 42MC demonstrates a decrease in TOC from the surface layer and deeper. All mult-core profiles exhibit values within the range error, which is indicative of efficient sediment mixing via bioturbation.

91 A marine dominance in offshore GoP kerogen is supported by numerous other isotope studies. A carbon isotope study from Bird et al. [1995] suggests that TOC in sediments beyond the outer shelf are of >90% marine origin. Similarly, Goni et al.

[2006] found that terrestrial organic matter is greatest in the Fly delta and inner-shelf sediments, but that marine organic matter dominates farther into the GoP. Burns et al.

[2004] found that organic detritus collected in sediment traps from the Pandora Trough contained few terrestrial plant biomarkers from fluvial sources.

It is surprising that more terrestrial phytoclasts are not observed in the study area, given the proximity of large and very active tropical rivers. However, this likely attests to the retention of particles near shore in the mangrove-rimmed coastline and in the inner- shelf clinoform. It also reflects effective recycling of allochthonous phytoclasts in oxygenated water within the outer topset and upper foreset of the inner-shelf clinoform where tidally-driven physical reworking is strongest [Aller, 1998; Aller and Blair, 2004].

3.6. Conclusions

In this manuscript, the detrital and biogenic components of outer shelf-edge and basinal sediment in the Gulf of Papua (GoP) are examined using magnetic susceptibility

(MS), calcium carbonate content, organic geochemical measurements, and organic petrography. A detrital gradient is observed in the MS in which highest values are found on the outer shelf-edge and upper slopes riming the Pandora Trough. Moderate MS values extend into the central and southern Pandora Trough and then decrease to lowest values in the Ashmore Trough. Calcium carbonate displays an inverse trend relative to

MS where highest concentrations are observed in the Ashmore Trough adjacent to the

Great Barrier Reef. Moderate calcium carbonate concentrations interfinger with detrital

92 surface sediments in the southern Pandora Trough, and then reach lowest concentrations

on the upper slopes in the vicinity of the Kerema Canyon. Together, the distribution of

these two surface sediment parameters demonstrates a mixing of carbonate and clastic

depositional systems in the central to southern Pandora Trough.

Total organic carbon concentrations are highest (>0.8 wt. %) on the outer shelf-

edge and upper slopes in the Kerema Canyon and then decrease into the Pandora and

Ashmore Trough. Rock-Eval Pyrolysis data indicate that the sedimentary organic matter is terrestrial Type 3 or 4 but kerogen petrography indicates <~30% to be of botanical provenance. Instead, kerogen is dominated by plankton or bacterially-derived amorphous organic matter, nearly all of which is nonfluorescent. Enhanced preservation of organic particles in the vicinity of the Kerema Canyon is probably favored by higher sediment accumulation rates found in that region [Muhammad et al., 2007] whereas surface oxidation and intense bioturbation of organic matter probably explains the severe degradation or organic particles that are preserved.

3.7. References

Alldredge, A.L. (1979), The chemical composition of macroscopic aggregates in two neritic seas, Limnology and Oceanography, 24, 855-866.

Alldredge, A.L. and M.W. Silver (1988), Characteristics, dynamics and significance of marine snow, Progress in Oceanography, 20, 41-82.

Aller, R.C. (1998), Mobile deltaic and continental shelf muds as suboxic, fluidized bed reactors, Marine Chemistry, 61, 143-155.

Aller, R.C. and N.E. Blair (2004), Early diagenetic remineralization of sedimentary organic C in the Gulf of Papua deltaic complex (Papua New Guinea): net loss of terrestrial C and diagenetic fractionation of C isotopes, Geochimica et Cosmochimica Acta, 68, 1815-1825.

Alongi, D.M. and A.I. Robertson (1995), Factors regulating benthic food chains in tropical river deltas and adjacent shelf areas, Geo-Marine Letters, 15, 145-152.

93

Batten, D.J. (1996), Palynofacies and palaeoenvironmental interpretation, in Palynology: Principles and Applications, edited by D.C. McGregor, American Association of Stratigraphic Palynologists Foundation, pp. 1011–1064.

Bird, M.I., G.J. Brunskill, and A.R. Chivas (1995), Carbon-isotope composition of sediments from the Gulf of Papua, Geo-Marine Letters, 15, 153-159.

Brunskill, G.J. (2004), New Guinea and its coastal seas, a testable model of wet tropical coastal processes: an introduction to Project TROPICS, Continental Shelf Research, 24, 2273-2295.

Brunskill, G.J., K.J. Woolfe, and I. Zagorskis (1995), Distribution of riverine sediment chemistry on the shelf, slope and rise of the Gulf of Papua, Geo-Marine Letters, 15 (3-4), 160-165.

Brunskill, G.J., I. Zagorskis, and J. Pfitzner (2003), Geochemical mass balance for lithium, boron, and strontium in the Gulf of Papua, Papua New Guinea (Project TROPICS), Geochimica et Cosmochimica Acta, 67(18), 3365-3383.

Burns, K.A., P. Greenwood, R. Benner, D. Brinkman, G. Brunskill, S. Codia, and I. Zagorskis (2004), Organic biomarkers for tracing carbon cycling in the Gulf of Papua (Papua New Guinea), Continental Shelf Research, 24, 2373-2394.

Carson, B.E., R.M. Leckie, J.M. Francis, A.W. Droxler, G.R. Dickens, S.J. Bentley, L.C. Peterson, and B.N. Opdyke (2007), Foraminiferal Populations of the Southern Ashmore Trough (Gulf of Papua) Mixed Siliciclastic-Carbonate System over the past 75 kyrs, Journal of Geophysical Research - Earth Surface.

Daniell, J., (2007), Development of a 100-m bathymetry grid for the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Davies, H.L. and I.E. Smith (1971), Geology of Eastern Papua, Geological Society of America Bulletic, 82, 3299-3312.

Delvaux, D., H. Martin, P. Leplat, and J. Paulet (1990), Geochemical characterization of sedimentary organic matter by means of pyrolysis kinetic parameters, Organic Geochemistry, 16(1-3), 175-187.

deMenocal, P., J. Bloemendal, and J. King (1991), A rock-magnetic record of monsoonal dust deposit ionto the Arabian Sea: Evidence for a shift in the mode of deposition at 2.4 Ma, in Proceedings ODP, Scientific Results 117, edited by W. Prell and L. Niitsuma, pp. 389-407.

94 Ellwood, B.B., W.L. Balsam, and H.H. Roberts (2006), Gulf of Mexico sediment sources and sediment transport trends from Magnetic Susceptibility measurements of surface samples, Marine Geology, 230, 237-248.

Ellwood, B.B., R.E. Crick, A. El Hassani, S. Benoist, and R. Young (2000), MagnetoSusceptibility Event and Cyclostratigraphy (MSEC) in marine rocks and the question of detrital input versus carbonate productivity, Geology, 28, 1135 - 1138.

Ellwood, B.B., K.M. Petruso, and F.B. Harrold (1996), The utility of magnetic susceptibility for detecting paleoclimatic trends and as a stratigraphic tool: an example from Konispol Cave sediments, SW Albania, Journal of Field Archaeology, 23, 263-271.

Ellwood, B.B., J.H. Tomkin, K.T. Ratcliffe, M. Wright, and A.M. Kafafy (2007), Magnetic Susceptibility and geochemistry for the Cenomanian/Turonian boundary GSSP with correlation to time equivalent core. GSA Annual Meeting, Denver.

Emerson, S. and J.I. Hedges (1988), Processes controlling the organic carbon content of open ocean sediments, Paleoceanography, 3(5), 621-634.

Espitalié, J. and M.L. Bordenave (1993), Rock-Eval pyrolysis, in Applied Petroleum Geochemistry, edited by M.L. Bordenave, pp. 237-261, Editions Technip, Paris, France.

Espitalié, J., M. Madec, and B. Tissot (1980), Role of mineral matrix in kerogen pyrolysis: Influence of petroleum generation and migration, AAPG Bulletin, 64, 59-66.

Flood, R., and D. Piper (1997), Amazon Fan sedimentation: the relationship to equatorial climate change, continental denudation, and sea-level fluctuations, in Proc. ODP, Sci. Results 155, edited by R.D. Flood, D.J.W. Piper, A. Klaus, and L.C. Peterson, pp. 653- 675.

Francis, J. M., J. Daniell, A. W. Droxler, G. R. Dickens, S. J. Bentley, L. C. Peterson, B. Opdyke, and L. Beaufort (2007), Deepwater geomorphology and sediment pathways of the mixed siliciclastic/carbonate system, Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Francis, J. M., G.B. Dunbar, G.R. Dickens, I.A. Sutherland, and A.W. Droxler (2007), Siliciclastic sediment across the north queensland margin (Australia): A Holocene perspective on reciprocal versus coeval deposition in tropical mixed siliciclastic- carbonate systems, Journal of Sedimentary Research, 77, 572-586.

Goni, M.A., N. Monacci, R. Gisewhite, A. Ogston, J. Crockett, and C. Nittrouer (2006), Distribution and sources of particulate organic matter in the water column and sediments of the Fly River Delta, Gulf of Papua (Papua New Guinea), Estuarine, Coastal and Shelf Science, 69, 225-245.

95 Hardy, M. J. (2000), Origin, distribution, and degradation of sedimentary organic matter in a modern tropical deltaic system (Mahakam Delta, Borneo, Indonesia), unpublished Dissertation, Louisiana State University, 368 p.

Harris, P.T., C.B. Pattiaratchi, J.B. Keene, R.W. Dalrymple, J.V. Gardner, E.K. Baker, A.R. Cole, D. Mitchell, P. Gibbs, and W.W. Schroeder (1996), Late Quaternary deltaic and carbonate sedimentation in the Gulf of Papua foreland basin: Response to sea-level change, Journal of Sedimentary Research, 66(4), 801-819.

Hart, G.F. (1986), Origin and classification of organic matter in clastic systems, Palynology, 10, 1-23.

Hartnett, H.E., R.G. Keil, J.I. Hedges, and A.H. Devol (1998), Influence of oxygen exposure time on organic carbon preservation in continental margin sediments, Nature, 391, 572-574.

Heath, G.R., T.C. Moore, and J.P. Dauphin (1977), Organic carbon in deep-sea sediments, in The Fate of Fossil Fuel CO2 in the Oceans, edited by N.R. Anderson and A. Malahoff, Plenum, New York, pp. 605-626.

Hedges, J.I., F.S. Hu, A.H. Devol, H.E. Hartnett, E. Tsamakis, and R.G. Keil (1999), Sedimentary organic matter preservation: A test for selective degradation under oxic conditions, American Journal of Science, 299, 529-555.

Hedges, J.I. and R.G. Keil (1995), Sedimentary organic matter preservation: an assessment and speculative assessment, Marine Chemistry, 49, 81-115.

Jarzie, D.M. (1991), Total organic carbon (TOC) analysis, in Source and Migration processes and evaluation techniques. Treatise of Petroleum Geology, Handbook of Petroleum Geology, edited by R.K. Merrill, American Association of Petroleum Geologists, Tulsa, pp. 113-118.

Jones, G.A. and P. Kaiteris (1983), A vacuum-gasometric technique for rapid and precise analysis of calcium carbonate in sediments and soils, Journal of Sedimentary Petrology, 53, 655-659.

Katz, B. (1983), Limitations of ‘Rock-Eval’ pyrolysis for typing organic matter, Organic Geochemistry, 4, 195-199.

Keen, T.R., D.S. Ko, R.L. Slingerland, S. Reidlinger, and P. Flynn (2006), Potential transport pathways of terrestrial material in the Gulf of Papua, Geophysical Research Letters, 33, 4, L04608, doi:10.1029/2005GL025416.

Langford, F.F. and M.M. Blanc-Valleron (1990), Interpreting Rock-Eval Pyrolysis data using graphs of pyrolizable hydrocarbons vs. total organic carbon, AAPG Bulletin, 74, 799-804.

96

Lewan, M.D. (1986), Stable carbon isotopes of amorphous kerogens from Phanerozoic sedimentary rocks, Geochimica et Cosmochimica Acta, 50, 1583-1591.

Maslin, M. and N. Mikkelsen (1997), Amazon fan mass transport deposits and underlying interglacial deposits: age estimates and fan dynamics, in Proceedings ODP, Scientific Results 155, edited by R.D. Flood, D.J.W. Piper, A. Klaus, and L.C. Peterson, pp. 353- 365.

Meyers, P.A. (1997), Organic geochemical proxies of paleoceanographic, paleolimnologic, and paleoclimatic processes, Organic Geochemistry, 27(5/6), 213-250.

Milliman, J.D. , K.L. Farnswerth, and C.S. Albertin (1999), Flux and fate of fluvial sediments leaving large islands in the East Indies, Journal of Sea Research, 41, 97-107.

Milliman, J.D. and J.P.M. Syvitski (1992), Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers, Journal of Geology, 100, 525-544.

Milliman, J.D. (1995), Sediment discharge to the ocean from small mountainous rivers: the New Guinea example, Geo-Marine Letters, 15, 127-133.

Muhammad, Z., S.J. Bentley, L.A. Febo, A.W. Droxler, G.R. Dickens, B.N. Opdyke, and L.C. Peterson (2007), Excess 210Pb inventories and fluxes along the continental slope and basin of the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Müller, P.J. and E. Seuss (1979), Productivity, sedimentation rate, and sedimentary organic matter in the oceans – I. Organic carbon preservation, Deep-Sea Research, 26A, 1347-1362.

Nittrouer, C.A. and L.D. Wright (1994), Transport of particles across continental shelves Review of Geophysics, 32(1), 85-113.

Oboh-Ikuenobe, F.E. and S.E. de Villiers (2003), Dispersed organic matter in samples from the western continental shelf of Southern Africa: palynofacies assemblages and depositional environments of Late Cretaceous and younger sediments, Palaeogeography, Palaeoclimatology, Palaeoecology, 201, 67-88.

Omura, K. and K. Hoyanagi (2004), Relationships between composition of organic matter, depositional environments, and sea-level changes in backarc basins, central Japan, Journal of Sedimentary Research, 74(5), 620-630.

Palinkas, C.M., C.A. Nittrouer, and J.P. Walsh (2006), Inner-shelf sedimentation in the Gulf of Papua, New Guinea: A mud-rich, shallow shelf setting, Journal of Coastal Research, 22(4), 760-772.

97 Posamentier, H.W., and P.R. Vail (1988), Eustatic controls on clastic deposition II - sequence and systems tract models, in Sea-level Changes: an Integrated Approach, SEPM Special Publication, vol. 42, edited by C.K. Wilgus, B.S. Hastings, C.G.S.C. Kendall, H.W. Posamentier, C.A. Ross, and J.C. Van Wagoner, SEPM, Tulsa.

Rickwood, F.K. (1968), The geology of Western Papua, The APEA Journal, 8 (2), 51-61.

Rullkötter, J. (2000), Organic matter: The driving force for early diagenesis, in Marine Geochemistry, edited by H. D. Schultz and M. Zabel, pp. 129-172, Springer-Verlag, Berlin, Germany.

Sebag, D., C. Di Giovanni, S. Ogier, V. Mesnage, F. Laggoun-Défarge, A. Durand (2006), Inventory of sedimentary organic matter in modern wetland (Marais Vernier, Normandy, France) as source-indicative tools to study Holocene alluvial deposits (Lower Seine Valley, France), International Journal of Coal Geology 67, 1-16.

Slingerland, R., N.W. Driscoll, J.D. Milliman, S.R. Miller, and E.A. Johnstone (2007), Anatomy and growth of a Holocene clinothem in the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Stein, R. (1990), Organic carbon content/sedimentation rate relationship and its paleoenvironmental significance for marine sediments, Geo-Marine Letters, 10, 37-44.

Swartzendruber L.J. (1992), Properties, units and constants in magnetism, Journal of Magnetism and Magnetic Materials, 100, 573–575.

Teichmüller, M. (1986), Organic petrology of source rocks, history and state of the art, Organic Geochemistry, 10, 581-590.

Tissot, B.P. and D.H. Welte (1984), Petroleum Formation and Occurrence, 699 pp., Springer-Verlag, Berlin, Germany.

Tuweni, A.O. and R.V. Tyson (1994), Organic facies variations in the Westbury formation (Rhaetic, Bristol Channel, S.W. England), Organic Geochemistry, 200: 1001- 1014.

Tyson, R.V. (1989), Late Jurassic palynofacies trends, Piper and Kimmeridge Clay Formations, UK onshore and offshore, in Northwest European Micropaleontology and Palynology, edited by D.J. Batten and M.C. Keen, British Micropalaeontological Society Series, pp. 135-172.

Tyson, R.V. (1993), Palynofacies analysis, in Applied Micropalaeontology, edited by D.G. Jenkins, Kluwer Academic Publishers, pp. 153-183.

Tyson, R.V. (1995), Sedimentary Organic Matter: Organic Facies and Palynofacies, 615 pp., Chapman and Hall, New York.

98

Van Gizel, P. (1979), Manual of the techniques and some geological applications of fluorescence microscopy, American Association of Stratigraphic Palynologists Workshop, Dallas, Oct. 29-Nov. 2, 55p.

Van Waveren, I. (1992), Morphology of probable planktonic crustacean eggs from the Holocene of the Banda Sea (Indonesia), in Neogene and Quaternary Dinoflagellate Cysts and Acritarchs, edited by M.J. Head and J.H. Wrenn, American Association of Stratigraphic Palynologists Foundation, pp. 89-120.

Wagner, T. (2000), Control of organic carbon accumulation in the late Quaternary equatorial Atlantic (Ocean Drilling Program sites 664 and 663): Productivity versus terrigenous supply, Paleoceanography, 15(2), 181-199.

Walsh, J.P. and C.A. Nittrouer (2003), Contrasting styles of off-shelf sediment accumulation in New Guinea, Marine Geology, 196, 105-125.

Walsh, J.P., C.A. Nittrouer, C.M. Palinkas, A.S. Ogston, R.W. Sternberg, and G.J. Brunskill (2004), Clinoform mechanics in the Gulf of Papua, New Guinea, Continental Shelf Research, 24, 2487-2510.

Wolanski, E. and D.M. Alongi (1995), A hypothesis for the formation of a mud bank in the Gulf of Papua, Geo-Marine Letters, 15, 166-171.

99 CHAPTER 4 DETRITAL CONTROLS ON MAGNETIC SUSCEPTIBILITY AND CYCLOSTRATIGRAPHY RECORDS

4.1. Introduction

All materials are susceptible to becoming magnetized in an inducing magnetic

field, and magnetic susceptibility (MS) is a rapid, simple, and nondestructive indicator of

that magnetization. Because many marine sedimentary records demonstrate Milankovitch

climatic cycles, MS is used for the astronomical tuning, in part to better define geological

time scales [deMenocal et al., 1991; Hellar and Evans, 1995; Shackleton et al., 1999;

Weedon et al., 1999], and for correlation [i.e. Ellwood et al., 2007]. In addition, core and coretop MS datasets from marine settings have been used to assess sediment transport pathways and provenance [Ellwood and Ledbetter, 1977b; Bulfinch et al., 1982; Sachs and Ellwood, 1988; Andrews and Stravers, 1993; Ellwood et al., 2006]. To correctly interpret long-term cyclostratigraphic MS records and geological processes from cored marine sediments, however, it is important to understand the factors influencing MS

values.

Of particular concern here is to understand the proximate cause(s) responsible for

variability within marine MS datasets. The simplest explanation is that MS variations are proportional to changes in terrigenous concentrations and/or mineralogical composition

[deMenocal et al., 1991; Ellwood et al., 1997; Ellwood et al., 2000]. However, because an inverse relationship between calcium carbonate and MS was demonstrated in surficial marine sediment samples [i.e., Ellwood and Ledbetter, 1977b], it is the perception of

many in the geological community that calcium carbonate variability, and therefore

productivity, is an important proximate control on MS variability [Mead and Tauxe,

100 1986; Andrews and Stravers, 1993]. Along these lines, others have inferred that in the absence of magnetite, siliceous biological productivity may overwhelm and then lower weaker MS values [Warner and Domack, 2002].

Therefore, the two models for MS variability in marine sediments are that MS is controlled 1) by changes in detrital fraction flux or 2) by variable marine productivity.

This means that, depending on the model adopted, the ultimate cause (climate) of MS variations could be interpreted very differently. With the first model in mind,

Milankovitch climate cyclicity will influence MS by changing the rates of detrital sediment supply. For example, weathering processes that erode and transport terrigenous particles to marine basins will operate faster and/or more broadly during times of wetter climates. In contrast, if the second model is correct, then Milankovitch climate cyclicity must cause variations in marine biological productivity, perhaps through pulses of nutrient availability, and thus cause changes in the flux and accumulation of organic carbon, coccolithophores, foraminifera, and/or biosiliceous microfossils at the seabed. If the second model is correct, then it should be possible to discern productivity variations in the past from MS cyclostratigraphic records and aid in interpreting the global carbon cycle proposed by Berner [2003].

Here we discuss some theoretical aspects of MS values and then experimentally examine the two models for proximate causes of MS variability. We further test the two models by examining MS from recent marine sediments collected in the Gulf of Papua

(GoP) and two marine sediment outcrop records, one from the Paleocene of Spain, and the other from Upper Cretaceous Niobrara Chalk exposures in Kansas.

101 4.1. Magnetic Susceptibility (MS) of Marine Sediments

Marine sediments are composed of a complex mixture of abiotic and biotic ingredients that influence MS records. Terrigenous components in marine sediments are typically composed of a range of mineralogies that have a range of magnetic susceptibilities. For example, ferrimagnetic iron-bearing minerals such as magnetite, titanomagnetite, or maghemite commonly occur in marine sediments, resulting in high

MS values. When ferrimagnetic mineral content is low or absent, MS values may still be high because of the presence of relatively large amounts of weakly magnetic, paramagnetic minerals, mainly ferromagnesian silicates and clay minerals [Ellwood et al., 2000]. Paramagnetic compounds, such as the clay mineral illite, can dominate the detrital MS signal in marine sedimentary rocks [Ellwood et al., 2007].

The most common minerals in marine sediments are calcite, quartz, feldspar, and chert, and they are diamagnetic, exhibiting a negative MS. The question we are testing here is whether variations in these abundant diamagnetic minerals can significantly lower overall sediment MS in the presence of ferrimagnetic and/or paramagnetic compounds.

4.2. Materials and Methods

4.2.1. MS Instrumentation

Bulk low-field MS was measured on experimental, core, and outcrop samples using a susceptibility bridge in a magnetically shielded room in the LSU Rock

Magnetism Laboratory. The MS Bridge is sensitive to <1 (±0.2) x 10-9 m3 kg-1 [Ellwood et al., 1996], calibrated using values for standard salts reported by Swartzendruber

[1992], and we report the mean of three replicate measurements of each sample. Because many MS data sets in the literature are measured using a MS2 Bartington instrument,

102 commonly the MS2C Loop Sensor configured on a GeoTek Multi Sensor Core Logger

(MSCL), we examine how measurements made from a Bartington MS2C compare with

the LSU MS bridge measurements. Bartington units are reported in the volume-corrected

units of 10-5 SI, while values from the LSU bridge are reported in m3 kg-1 (mass corrected

units).

4.2.2. Experimental Samples

Three mineral samples were prepared to simulate the range of MS values in

marine sediments (Table 4.1). The first sample is composed of ~1-2 g of powdered

Iceland basalt [from samples collected by Ellwood and Ledbetter, 1977b] and represents

a ferrimagnetic end-member sample. The second sample contains ~1 g of an illite

standard [Ellwood et al., 2007] and represents a paramagnetic clay mineral found in

marine sediments. The third sample is composed of a ~3 g 50/50 mixture of the illite

standard and diamagnetic quartz sand. This third sample simulates the value of a marine

sediment sample containing both a paramagnetic clay mineral and diamagnetic

siliciclastic quartz sand.

Prepared samples were placed in separate small sample bags, weighed, and then

measured for initial MS. To simulate the effect of marine productivity on MS, samples

were diluted with powdered reagent-grade calcium carbonate (calcite) in 0.1 g increments

up to 0.5 g and then in 0.25 g increments. After enough calcite was added to the samples

to represent 100% dilution, samples were then diluted with an additional 1 g of calcite in

succeeding increments until at least one order of magnitude shift (e.g. 1x10-7 → 1x10-8) was achieved. MS was measured after each dilution increment.

103 We chose calcite as the diluting mineral because it is common in the marine

environment, is diamagnetic (Table 4.1), and because its concentration is often reported

in the literature in conjunction with MS datasets [Ellwood and Ledbetter, 1977a; Mead

and Tauxe, 1986, Andrews and Stravers, 1993]. We chose not to use pure powdered

silica to simulate biosiliceous productivity because of its unusual electromagnetic

properties in an applied field, and its MS value is very similar to calcite (Table 4.1).

TOC is also diamagnetic but not considered significant here because TOC concentrations in sediments (except for coal) are generally very low, typically <2% [Tissot and Welte,

1984].

Table 4.1. Minerals, initial mass, and initial MS values for samples used in the experiments conducted in this research. Sample # Mineral(s) Initial mass, g Initial MS, m3 kg-1 1 Iceland Basalt 1.8 7.58x10-6 2 Illite Standard 1.0 1.20x10-7 3 Illite/Quartz sand 2.7 4.53x10-8 Diamagnetic calcite -3.00 x10-9 Quartz sand -3.25x10-9

4.2.3. Core-top, Piston-core, and Outcrop Samples

A variety of published and unpublished datasets are used in this manuscript to test

for controls on MS [e.g., Febo et al., 2007; Ellwood et al., In Review]. Piston core samples include two radiocarbon-dated jumbo piston cores collected from the GoP [Febo

et al., 2007, see also Chapter 2] and contain sediment archives deposited in >800 meters

water depth (mwd) for the past 38,000 Cal. yrs. B.P. A core-top dataset from the GoP

includes 42 surface samples (0-2 cm) from multi-cores, box-cores, and trigger-cores

104 [carbonate reported in Febo et al., 2006, see also Chapter 3]. The GoP is a modern

example of a mixed siliciclastic-carbonate depositional setting situated between the Great

Barrier Reef and Papua New Guinea.

Outcrop samples were collected from two localities within the Upper Cretaceous

and the Paleocene. The Upper Cretaceous dataset includes 1,840 samples collected at 10

cm increments from a ~200 m composite sequence of the Niobrara Chalk Formation

(Fort Hays and Smokey Hill Chalk members), Kansas. This section was selected because the entire sequence is composed of chalk and therefore provides an excellent test of carbonate controls on MS variations. The Paleocene section is a 20 sample subset we measured for MS and calcite out of a larger collection (n=175) across the proposed

Danian-Selandian Global Stratotype Section and Point (GSSP) at Zumaia, Spain

[Ellwood et al., In Review a]. The Zumaia sequence represents a change from carbonate- dominated sediments during the Danian to siliciclastic-dominated sediments during the

Selandian and contains a cyclostratigraphic record controlled by Milankovitch climate cycles [Ellwood et al., In Review].

4.2.4. Calcium Carbonate Measurements

Calcium-carbonate concentrations were analyzed using the vacuum-gasometric

technique of Jones and Kaiteris [1983] at the LSU Rock Magnetism Laboratory. A ~2 g specimen from each sample was ground to a fine powder, and then dried for four hours

on a hotplate at 40°C. A 0.25 g sub-sample was then taken and reacted with phosphoric

acid for one hour under a vacuum in a glass reaction vessel. After complete reaction, the

pressure difference was then read from a pressure gauge and used to calculate the sample

carbonate concentration from a calibration equation. Replicate measurements on several

105 samples resulted in a precision of +/- 0.1%. Tests of reagent grade (99.9%) carbonate and a mixed sample standard (95% feldspar 5% carbonate) gave accurate results to +/-

1% and precision to <0.5%.

4.2.5. Numerical Methods: Un-mixing MS Values

Because MS values are a function of many minerals, they can potentially be decomposed into mineral fractions if one or more fractions are known and if sediment samples represent relatively simple mixtures [Potter et al., 2004]. A mass balance equation (equation 1) illustrates that the sample MS is a function of numerous mineral fractions and their mass-dependent MS values.

MS χ t )( = χ + χ 2211 + ...FFF χ nn

where χ t = total sample MS (1)

F1 = component fraction

This equation makes it possible to examine only the terrigenous MS of a sample if the non-terrigenous fractions, such as calcium carbonate, are known (equation 2).

χ − F χ cct χ nc = Fnc

where χ nc = noncarbona MSte (2) Fc = fraction carbonate

If we separate just the calcite fraction, there is still an unresolved fraction of diamagnetic siliceous material in the new MS value, both biological and detrital in origin. Therefore,

106 the new MS is conservatively labeled the non-carbonate MS. Then it is possible to

consider the importance of the ferrimagnetic and paramagnetic components. In most

-9 cases (MS > 1x10 ), χc is so small that it can be ignored and therefore equation 2

becomes:

χ t χ nc = (3) Fnc

4.3. Results

4.3.1. Instrumental Comparisons

In general, there is a very good agreement between MS measured using the LSU

bridge and a Bartington MS2C loop sensor (Figure 4.1). However, the LSU bridge

detected significant variability in MS in core MV-51JPC from 3-9 m that was not

detected using the MS2C sensor (Figure 4.1). This may be attributed to low sensitivity

in the MS2C sensor and therefore limited instrumental detection in the very narrow range

below ~5x10-5 SI.

4.3.2. Experimental Results

All three mixing experiments reveal that very high amounts of calcite must be

added to offset the initial paramagnetic MS of a sample. The basalt sample required

~700% of additional calcite to be added before the value shifted by one order of

magnitude (Figure 4.2). The illite sample required slightly more dilution by calcite

(~860%) to change MS by one order of magnitude. The mixed illite-sand sample

required ~880% dilution by calcite to reduce the MS by one order of magnitude. A

power curve fits all experimental data ( y=axb, regression coefficient R=0.99).

107

Figure 4.1. Comparison between Bartington MS2C loop sensor and the LSU bridge measurements for two piston cores from the Gulf of Papua. The R value is from a logarithm regression of the combined core datasets presented in the central cross-plot. Individual core profiles on the left (MV-51JPC) and right (MV-54JPC) illustrate similar trends in data. Gaps in the Bartington MS are deleted erroneous values measured at core section breaks during logging. Bartington data below ~ 5x10-5 SI do not detect the wide range in MS values observed by the LSU bridge because of reduced instrumental sensitivity in that range.

The results shown in Figure 4.2 consistently demonstrate a non-linear, inverse

correlation between calcite concentrations and MS. In addition, MS is not easily reduced

by diamagnetic calcite. Figure 4.2 illustrates the difficulty in producing diamagnetic values by dilution of illite.

108

Figure 4.2. (A) Results of experimental dilution with calcite of ferrimagnetic (basalt), paramagnetic (illite), and a mixed paramagnetic/diamagnetic sample. The illite curve shows that >850% dilution by calcite is required to lower the MS by one order of magnitude, representative of typical variations observed in MS records. (B) A schematic representation illustrating the extremely large difference between diamagnetic calcite and paramagnetic illite.

4.3.3. Geological Samples

MS and detrital non-carbonate (100-% carbonate) data are shown in Figure 2 from two contrasting locations and geological ages. The Spanish Danian-Selandian samples and the GoP core-tops both show highly significant and positive correlations (Figure

4.3). The Danian-Selandian dataset is linearly correlated whereas the GoP dataset is best

109 fit with a non-linear regression line. However, piston-core data are not correlated

because the range in calcite is very low (<10%) [Febo et al., 2007].

Figure 4.3. MS and non-carbonate percent (100-% carbonate) data for geological samples. The Danian-Selandian data show a positive and highly significant linear regression (R=0.84) whereas core-tops from the Gulf of Papua show a strong nonlinear b correlation, also highly significant (power curve, y=y0+ax , R=0.80). The two piston cores show little variation because they contain extremely low amounts of calcite.

The Niobrara Chalk dataset is complex and is best broken down into subsets for

analysis. Overall, there is a slight positive correlation (Section A, R=0.13) between non-

carbonate and MS (Figure 4.4). However, there are portions of the sequence that, while appearing to be inversely correlated, actually show no sample to sample correlations

(Section B, R=0). In addition, some intervals show a statistically significant sample-to-

sample positive correlation, (e.g. Section C, R=0.50).

110

Figure 4.4. MS and non-carbonate percent data for the Niobrara Chalk Formation. Overall (A), there is a slight positive correlation (R=0.13), however data subsets demonstrate variations in the correlation (see text). Section B shows an apparent inverse correlation while section C shows a positive correlation. In contrast, the section between 40 and 80 m is characterized by an increase in detrital component while the MS is relatively non-varying. These data illustrate the complex relationships between terrigenous fractions and MS.

4.3.4. Numerical Un-mixing

MS values for all geological samples were recalculated using equation 3 to determine the MS for only the non-carbonate fraction. Data in Figure 4.5 demonstrate statistically significant positive correlations for all data sets with both piston-cores demonstrate the strongest linear correlation (R=0.99). The GoP core-tops are best fit

111 with a non-linear regression line and also exhibit a strong correlation (R=0.81). Most

important to note here is the highly significant, strong linear correlation found between

these parameters for the whole Niobrara Chalk dataset (R=0.61) that was not evident in

the parameters plotted in Figure 4 (R=0.13). Correlations would be even stronger if it

were possible to extract and remove the diamagnetic siliceous component in these

samples.

Figure 4.5. Non-carbonate MS versus the whole sample MS for the geological data sets examined. Non-carbonate MS values were recalculated from the initial measured sample MS using equation 3. Piston-core samples form an approximate 1:1 line whereas lower slopes formed by other data sets reveal variable contribution from other noncarbonate diamagnetic fractions.

4.4. Discussion

4.4.1. Non-Carbonate Controls

The experimental data demonstrate that calcite can change MS values, but only at

very high levels of dilution (Figure 4.2). For example, a marine sample with a small

amount of basalt or illite (as in our experimental samples) would require thousands to

112 tens of thousands of percent dilution to cause a significant reduction in MS (Figure 4.2).

It is clear from these data that detrital (non-carbonate) fluxes added to marine sediments

control the observed MS. The main reason for this is that diamagnetic, siliceous, and

carbonate minerals exhibit such a low MS that slight additions of detrital paramagnetic

and ferrimagnetic minerals to the sample overwhelm the diamagnetic components

(Figure 4.2B).

The rock record of productivity indicates that natural variations in calcite, opaline

silica, or organic carbon rarely operate at the scale of thousands of percent dilution,

except in extreme cases. Chalk and diatomite are two rock types that form thick

accumulations (up to a few kilometers) in ancient basins that were both detrital sediment-

starved and contained favorable oceanographic conditions for marine productivity

[Hattin, 1982; White et al., 1992]. The Niobrara Chalk from the Western Interior Basin

is one example of enhanced Upper Cretaceous calcite productivity and reduced clastic

input. Here, even though productivity is high the overall correlation between detrital

component and MS is positive.

In addition to the work reported here for carbonate-dominated samples, several

samples of the well-known Oamaru Diatomite (Bain’s Farm site) from the South Island of New Zealand were measured with the LSU MS bridge. Diatom and radiolarian

(siliceous, diamagnetic) microfossils in these samples record extensive productivity

during the Late Eocene in a sediment-starved warm tropical to subtropical basin in water

depths of ~75-150 m [O’Connor, 1999]. Samples from this diatomite, just like the chalk,

are in middle range of marine sediment MS (4.59x10-8 m3 kg-1) and not diamagnetic as

might be predicted. This is because the diatomite is only composed of ~60% biosiliceous

113 microfossils, leaving the other 40% as detrital components such as clay and fine silt

[O’Connor, 1999].

One of the best examples of detrital controls on MS is from the GoP. A good correlation occurs (Figure 4.3) because there is a range in calcite concentrations in the

core-tops that is the result of cores taken from the “carbonate-dominated” western region of the Ashmore Trough adjacent to the Great Barrier Reef (GBR) platform and the clastic dominated Pandora and Moresby Trough. Today, sediments draining from rivers on the

Queensland margin are deposited on the inner-shelf, allowing the mid to outer-shelf GBR communities to thrive [Scoffin and Tudhope, 1985; Belperio and Searle, 1988; Niel et al.,

2002; Larcombe and Carter, 2004]. However, sediments from the Fly and adjacent rivers inundate the GoP inner-shelf and are then transported into the Pandora Trough via hemipelagic settling and episodic nepheloid layer transport, thus diluting carbonate in the sediments [Walsh et al., 2004; Muhammad et al., 2007]. Sediments deposited during the late Pleistocene, recovered in piston-cores from the central and northern Pandora Trough, show a very limited range in calcite and therefore there is no correlation between MS and non-carbonate fraction.

The Danian-Selandian data set is also an excellent example of non-carbonate controls on MS. The Danian stage sequence contains inter-bedded limestone and marl that becomes dominantly marl in the overlying Selandian stage. Here, a range of calcite concentrations are present in sediments because sediment accumulation rates changed across the stage boundary, lower in the Danian and higher in the Selandian [Ellwood et al., In Review].

114 Perhaps the best evidence of detrital controls on MS is shown in Figure 4.5.

Non-carbonate MS is easy to recalculate once the whole sample MS and the calcite

concentrations are known and this can be used to tease apart relationships not observed

with either the calcite or the non-carbonate fractions (Figure 4.5). While all correlations

are significant, not all are perfectly correlated (e.g. Niobrara Chalk) because of

diamagnetic compounds (i.e. quartz and feldspar) comprising an unknown portion of the

non-carbonate (inferred detrital) sediments.

4.4.2. Interpreting MS and Cyclostratigraphic Records

Comparing MS and calcite can understandably lead to misinterpretations because

sometimes the datasets are correlated and sometimes they are not (Figures 4.3 and 4.4).

As was discussed earlier, it is common to interpret sections with high MS values and

those with low values differently. For example, intervals of high MS values are often

interpreted as being strongly controlled by terrigenous input or provenance, while

intervals of low MS are inferred to be controlled by biological productivity [e.g. Andrews

and Stravers, 1993; Warner and Domack, 2002]. This then leads some to conclude that

MS may be used as an indication of past biogenic productivity, particularly for

cyclostratigraphic MS interpretation.

The proximate causes of MS variations are changes in the detrital concentrations

and mineralogy. The ultimate cause is the process(s) that govern changes to detrital input and compositions, such as climate change, provenance, topography, tectonics, and volcanism. MS variations producing cyclostratigraphic records with orbital frequencies

are best attributed to climatic changes that mitigate the detrital flux to marine basins.

Specifically, variations in wet-dry and warm-cool climates cause changes in weathering

115 rates and river discharge that transport detrital particles to ocean basins [Milliman and

Meade, 1983; Milliman and Syvitski, 1992; Hovius, 1998]. It is much more difficult to argue that long-term climate cyclicity is due to variation in productivity than to argue that this cyclicity is responding to detrital fluxes. This is particularly true when very large productivity rates are required to produce the range in MS values.

4.5. Conclusions

We demonstrate experimentally that dilution of detrital minerals with diamagnetic calcite cannot easily reproduce MS values within the natural range of variation of calcite compositions in marine sediments. Three experimental sediment sample sets were prepared to cover the range of marine sediment MS intensities, including ferrimagnetic

Icelandic basalt, paramagnetic illite standard, and a 50/50 mixed diamagnetic quartz sand and illite. All samples require a calcite dilution of at least 700% to shift MS values by one order of magnitude. In addition, we tested non-carbonate (100%-% calcite) concentrations and MS for several natural marine data sets, covering the range from detrital-dominated (GoP) to calcite-dominated (Niobrara Chalk) samples. Strong correlations were found in data sets where a range of non-carbonate values occurred, but no correlations in data sets where non-carbonate values were >95% (GoP piston cores) were observed. In this context, MS cyclostratigraphic records cannot be interpreted to represent changes in marine productivity but rather changes in climate that act to increase detrital sediment flux.

4.6. References

Andrews, J.T. and J.A. Stravers (1993), Magnetic susceptibility of late Quaternary marine sediments, Frobisher Bay, N.W.T.: An indicator of changes in provenance and processes, Quaternary Science Reviews, 12, 157-167.

116 Belperio, A.P., and Searle, D.E. (1988), Terrigenous and Carbonate Sedimentation in the Great Barrier Reef Province, in Carbonate - Clastic Transitions, edited by L.J. Doyle and H.H. Roberts, Elsevier, pp. 143-174.

Berner, R.A. (2003), The long-term carbon cycle, fossil fuels and atmospheric composition, Nature, 426, 323-326.

Bulfinch, D.L., M.T. Ledbetter, B.B. Ellwood, and W.L. Balsam (1982), The high velocity core of the Western Boundary Undercurrent at the base of the U.S. Continental Rise, Science, 215, 970-973.

deMenocal, P., J. Bloemendal, and J. King (1991), A rock-magnetic record of monsoonal dust deposit ionto the Arabian Sea: Evidence for a shift in the mode of deposition at 2.4 Ma, in Proceedings ODP, Scientific Results 117, edited by W. Prell and L. Niitsuma, pp. 389-407.

Ellwood, B.B., W.L. Balsam, and H.H. Roberts (2006), Gulf of Mexico sediment sources and sediment transport trends from Magnetic Susceptibility measurements of surface samples, Marine Geology, 230, 237-248.

Ellwood, B.B., C.E. Brett, and W.D. MacDonald (2007), Magnetosusceptibility Stratigraphy of the Upper Ordovician Kope Formation, Northern Kentucky, Palaeogeography, Palaeoclimatology, Palaeoecology, 243, 42-54.

Ellwood, B.B., R.E. Crick, A. Hassani, S.L. Benoist, and R.H. Young (2000), Magnetosusceptibility event and cyclostratigraphy method applied to marine rocks: detrital input versus carbonate productivity, Geology, 28(12), 1135-1138.

Ellwood, B.B. and M.R. Fisk (1977a), Anisotropy of magnetic susceptibility variations in a single Icelandic columnar basalt, Earth and Planetary Science Letters, 35, 116-122.

Ellwood, B.B., and M.T. Ledbetter (1977b), Antarctic bottom water fluctuation in the Vema Channel: Effects of velocity changes on particle alignment and size, Earth and Planetary Science Letters, 35, 189-198.

Ellwood, B.B., W.D. MacDonald, C. Wheeler, and S.L. Benoist (2003), The K-T boundary in Oman: identifed using magnetic susceptibility field measurements with geochemical confirmation, Earth and Planetary Science Letters, 206, 529-540.

Ellwood, B.B. K.M. Petruso, and F.B. Harrold (1996), The utility of magnetic susceptibility for detecting paleoclimatic trends and as a stratigraphic tool: an example from Konispol Cave sediments, SW Albania, Journal of Field Archaeology, 23, 263-271.

Ellwood, B.B., J.H. Tomkin, L.A. Febo, and C.N. Stuart, Jr. In Review, Time Series Analysis of Magnetic Susceptibility Variations in Deep Marine Sediments: A Test Using

117 Upper Danian-Lower Selandian Proposed GSSP, Spain, Palaeogeography, Palaeoclimatology, Palaeoecology

Ellwood, B.B., J.H. Tomkin, K.T. Ratcliffe, M. Wright, and A.M. Kafafy (2007), Magnetic Susceptibility and geochemistry for the Cenomanian/Turonian boundary GSSP with correlation to time equivalent core. GSA Annual Meeting, Denver.

Febo, L.A., S.J. Bentley, J.H. Wrenn, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke (2007), Late Pleistocene and Holocene Sedimentation, Organic-Carbon Delivery, and Paleoclimatic Inferences on the Continental Slope of the Northern Pandora Trough, Gulf of Papua. Journal of Geophysical Research - Earth Surface.

Febo, L.A. S.J. Bentley, J.H. Wrenn, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke (2006), Recent to Late Pleistocene Sedimentary Organic Matter in the Gulf of Papua, Papua-New Guinea. AAPG Annual Meeting Abstracts.

Hattin, D.E. (1982), Stratigraphy and Depositional Environments of Smoky Hill Chalk Member, Niobrara Chalk (Upper Cretaceous) of the Type Area, Western Kansas, Kansas Geological Survey, Bulletin 225, 108p.

Heller, F. and M.E. Evans (1995), Loess magnetism, Review of Geophysics, 33, 211-240.

Hovius, N. (1998), Controls on sediment supply by large rivers, in Relative Role of Eustacy, Climate, and Tectonism in Continental Rocks, edited by K.W. Shanley and P.J. McCabe, pp. 3-16.

Jones, G.A. and P. Kaiteris (1983), A vacuum-gasometric technique for rapid and precise analysis of calcium carbonate in sediments and soils, Journal of Sedimentary Petrology, 53, 655-659.

Larcombe, P., and Carter, R.M. (2004), Cyclone Pumping, Sediment Partitioning and the Development of the Great Barrier Reef Shelf System: A Review, Quaternary Science Reviews, 23, 107-135.

Mead, G.A. and L. Tauxe (1986), Oligocene paleoceanography of the South Atlantic: Paleoclimatic implications of sediment accumulation rates and magnetic susceptibility measurements, Paleoceanography, 1(3), 273-284.

Milliman, J.D. and R.H. Meade (1983), World-wide delivery of river sediment to the oceans, The journal of Geology, 91(1), 1-21.

Milliman, J.D. and J.P.M. Syvitski (1992), Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers, Journal of Geology, 100, 525-544.

118 Muhammad, Z., S.J. Bentley, L.A. Febo, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke (2007), Excess 210Pb inventories and fluxes along the continental slope and basins of the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Niel, D.T., A.R. Orpin, P.V. Ridd, and B. Yu (2002), Sediment yield and impacts from river catchments to the Great Barrier lagoon, Marine and Freshwater Research, 53, 733- 752.

O’Connor, B (1999), Radiolaria from the Late Eocene Oamaru Diatomite, South Island, New Zealand, Micropaleontology, 45(1), 1-55.

Potter, D.K., P.W.M. Corbett, S.A. Barclay, and R.S. Haszeldine (2004), Quantification of illite content in sedimentary rocks using magnetic susceptibility – A rapid complement or alternative to X-ray diffraction, Journal of Sedimentary Research, 74(5), 730-735.

Sachs, S. D., and B. B. Ellwood (1988), Controls on magnetic grain-size variations and concentration in the Argentine Basin, South Atlantic Ocean, Deep-Sea Research, 35, 929–942.

Scoffin, T.P., and A.W. Tudhope (1985), Sedimentary environments of the Central Region of the Great Barrier Reef of Australia, Coral Reefs, 4, 81-93.

Shackleton, N.J., S.J. Crowhurst, G.P. Weedon, and J. Laskar (1999), Astronomical calibration of Oligocene-Miocene time, Philosophical Transactions of the Royal Society of London A, 357, 1907-1929.

Tissot, B.P. and D.H. Welte (1984), Petroleum Formation and Occurrence, 699 pp., Springer-Verlag, Berlin, Germany.

Swartzendruber, L. J. (1992), Properties, units and constants in magnetism: Journal of Magnetic Materials, 100, 573-575.

Walsh, J.P., C.A. Nittrouer, C.M. Palinkas, A.S. Ogston, R.W. Sternberg, and G.J. Brunskill (2004), Clinoform mechanics in the Gulf of Papua, New Guinea, Continental Shelf Research, 24, 2487-2510.

Warner, N.R., and E.W. Domack, (2002), Millennial- to decadal-scale paleoenvironmental change during the Holocene in the Palmer Deep, Antarctica, as recorded by particle size analysis, Paleoceanography, 17, Art. No. 8004, doi:10.1029/2000PA000602.

Weedon, G.P., H.C. Jenkyns, A.L. Coe, and S.P. Hesselbo (1999), Astronomical calibration of the Jurassic time-scale from cyclostratigraphy in British mudrock formations, Philosophical Transactions of the Royal Society of London, A, 357, 1787- 1813.

119 White, L.D., R.E. Garrison, and J.A. Barron (1992), Miocene intensification of upwelling along the California margin as recorded in siliceous facies of the Monterey Formation and offshore DSDP sites, in Upwelling Systems: Evolution Since the Early Miocene, edited by C.P. Summerhayes, W.L. Prell, and K.C. Emeis, Geological Society Special Publication, 64, pp. 429-442.

120 CHAPTER 5

CONCLUSIONS

Surface sediment and sediment core samples examined in this study span the time

interval from recent to ~15,000 - 33,000 yrs B.P. and offer a unique window into

depositional events in the northeastern GoP. At present, a detrital gradient is observed in

the MS in which highest values are found on the outer shelf-edge and upper slopes riming

the Pandora Trough and lowest values are observed in the Ashmore Trough. Calcium

carbonate displays an inverse trend relative to MS where highest concentrations are

observed in the Ashmore Trough adjacent to the Great Barrier Reef and lowest values are found on the upper slopes of the Pandora Trough in the vicinity of the Kerema Canyon.

The distribution of these two surface sediment parameters demonstrates a classic mixing between carbonate and clastic depositional systems in the central to southern Pandora

Trough.

Total organic carbon concentrations are generally low in surface sediments from

the GoP. A gradient of organic carbon is observed from highest values on the outer

shelf-edge and upper slopes in the Kerema Canyon to lower concentrations in the

Ashmore Trough. Organic geochemical data indicate a terrestrial Type 3 or 4 organic matter, but kerogen petrography indicates <~30% to be of botanical origin. Kerogen is dominated by marine derived amorphous organic matter, nearly all of which is highly degraded. Sea-bed oxidation and intense bioturbation of organic matter probably explains the severe degradation of organic particles that are preserved.

121 In core records, geophysical, geochemical, and micropaleontological evidence indicate different oceanographic conditions in the GoP >32,000 14C yrs B.P. The

following are significant new findings for the late Pleistocene to recent depositional

history in the GoP.

1) Higher total organic carbon mass accumulation rates (TOC MAR) and benthic

foraminiferal accumulation rates (BFAR) prior to 32,000 14C yrs B.P. suggest that

greater amounts of organic carbon were being delivered to the seabed compared

to today. Increases in the densities of productivity taxa such as

U.peregrina/hispida-C. pachyderma-S. bulloides suggest some fraction of this

organic carbon was in a labile form.

2) Low hydrogen index values suggest highly oxidized organic matter of terrestrial

origin. If any labile compounds were originally present in sediments older than

32,000 14C yrs B.P., then they were mostly oxidized and not buried. Carbon

isotope data indicate that during this time, terrestrial organic carbon was more

abundant in the organic carbon pool than marine organic carbon and also suggest

that a significant fraction of the input was from C3 vascular plant matter.

3) A major change in the clastic sediment source is suggested by a shift magnetic

susceptibility values (MS) at ~32,000 14C yrs B.P. In contrast, an increase in

detrital flux rather than a change in mineralogy is implicated by thermomagnetic

results. During this time, other sediment properties (e.g. BFAR and carbon

isotopes) in core MV-51 are very different from modern sediments suggesting

that they were derived from a different source. The change in MS may be a

response to a different sediment provenance, such as a river(s) source draining the

122 Papuan Peninsula, or more direct dispersal pathways of central Papua New

Guinea rivers at lower sea level.

4) The time from ~32,000 14C yrs B.P. to 20,400 yrs B.P. corresponds to the

transition into the Last Glacial Maximum (LGM) and is marked by relatively

lower sediment and TOC accumulation rates but higher percentages of marine

organic carbon. This contrasts with many global records indicating enhanced

organic carbon burial during the LGM.

5) During the time from 15,000 to 20,400 yrs B.P., a period of high sediment

accumulation rates returned. Greatest accumulation rates are experienced in MV-

54 during late transgression (~15,000-18,000 yrs B.P.) when rivers were situated

on the exposed shelf and delivered sediments directly to the slopes. Stable-carbon

isotopes from MV-51 indicate enhanced terrestrial organic carbon delivery, but

the G. translucens, Gc. subglobosa, and M. barleeanus benthic foraminiferal

assemblage indicates higher seasonality of organic carbon flux during this time.

6) The significantly lower accumulation rates at MV-54 and MV-51 during the late

transgression to Holocene time are evidence in support of inland and coastal

sediment storage at higher sea level.

123 APPENDIX A: PUBLISHED AND SUBMITTED CHAPTER ABSTRACTS

CHAPTER 2

Lawrence A. Febo1*, Samuel J. Bentley2, John H. Wrenn**, André W. Droxler3, Gerald R. Dickens3, Larry C. Peterson4, and Bradley N. Opdyke5

1Department of Geology and Geophysics, Louisiana State University, E-235 Howe- Russell Geosciences Complex, Baton Rouge, LA 70803

2Earth Sciences Department, Memorial University of Newfoundland, St. John’s, Newfoundland, Canada A1B 3X5

3Department of Earth Science, Rice University, 6100 Main Street, Houston, TX 77251- 1892

4Rosenstial School of Marine and Atmospheric Science, University of Miami, Miami, FL 33149

5 Department of Earth and Marine Sciences, The Australian National University, Canberra, ACT 020, Australia.

* Corresponding Author, Lawrence A. Febo, [email protected]

**Deceased

Status: In Press to JGR Earth Surface, print scheduled for March 2008

Abstract

We investigated sediment and organic-carbon accumulation rates in two jumbo

piston cores (MV-54, MV-51) retrieved from the mid slope of the northeastern Pandora

Trough in the Gulf of Papua, Papua New Guinea. Our data provide a first assessment of

mass fluxes over the past ~33,000 14C yrs B.P. and variations in organic-carbon sources.

Core sediments were analyzed using a suite of physical properties, organic geochemistry,

and micropaleontological measurements. MV-54 and MV-51 show two periods of rapid

sediment accumulation. The first interval is from ~15,000-20,400 Cal. yrs B.P. (MV-51:

124 ~1.09 m kyr-1 and ~81.2 g cm-2 kyr-1) and the second occurs >32,000 14C yrs B.P. (~2.70

m kyr-1 and ~244 g cm-2 kyr-1). Extremely high accumulation rates (3.96 m kyr-1; ~428 g

cm-2 kyr-1) characterize 15,800-17,700 Cal. yrs B.P. in MV-54 and likely correspond to

early transgression when rivers delivered sediments much closer to the shelf-edge. A

benthic foraminiferal assemblage in MV-51 from ~18,400-20,400 Cal. yrs B.P. indicates

a seasonally variable flux of organic carbon, possibly resulting from enhanced contrast

between monsoon seasons. The oldest sediments, >32,000 14C yrs B.P., contain TOC fluxes >200 g cm2 kyr-1, with >50% of it derived from C3 vascular plant matter.

Magnetic susceptibility values are two to three times higher and benthic foraminiferal

accumulation rates are six times higher during this interval than any younger time

indicating a greater influence of detrital minerals and labile organic carbon. The MS data

suggest more direct dispersal pathways from central and eastern PNG Rivers to the core

site.

125 CHAPTER 4

Lawrence A. Febo1, Brooks B. Ellwood1, and David K. Watkins2

1Department of Geology and Geophysics, Louisiana State University, E-235 Howe- Russell Geosciences Complex, Baton Rouge, LA 70803

2Department of Geosciences, University of Nebraska, 214 Bessey Hall, Lincoln, NE 68588

Status: Will be submitted to Geology

Abstract

When marine sediments are analyzed, it is often argued that magnetic susceptibility (MS) values exhibit an inverse relationship with calcium carbonate concentrations. As a result, MS records are often interpreted as being controlled by calcium carbonate, and therefore, that MS may be a reliable indication of carbonate

productivity. However, because carbonate MS values are extremely small, other studies

have argued (and shown) that varying terrigenous detrital fluxes control MS values. To

test these two models, we present new experimental results and several natural large data

sets in support of the second model, that changes in mineralogy and/or detrital input are the main controls on MS variability, not carbonate productivity.

To model the effect of variations in biological productivity on marine sediments,

we prepared three separate samples of iron-rich minerals found as detrital components in

marine sediments. These were then incrementally diluted with powdered reagent grade

calcium carbonate and the MS measured. Prepared samples were designed to cover the

126 common range of MS observed for lithified marine sediments. In order of decreasing

initial MS, the samples are: 1) ~2 g of powdered Iceland basalt; 2) ~2 g of a paramagnetic

illite standard; and 3) ~3 g 50/50 mixture of illite and quartz sand. The calcium

carbonate dilution required to reduce initial sample susceptibilities by one order of

magnitude is >700% for all samples, far exceeding natural variations in carbonate

productivity. In addition, we observe that MS and calcium carbonate data for a 200 m

section of the Upper Cretaceous Niobrara Chalk, Kansas, show no significant correlation

(R=0.14, n=1840) when the whole data set is considered.

127 APPENDIX B: SUPPLEMENTAL CHAPTER 2 DATA TABLES NOT IN TEXT

Table 2.3. Sediment and organic carbon accumulation rate date for MV-54. TOC mass accumulation rate data were calculated using equation 1 in the text. MAR and TOC MAR are not reported where either dry bulk density or TOC data are absent.

Depth Age DBD LSR MAR, TOC TOC MAR m kyrs g cm-1 cm kyr-1 g cm-2 kyr-1 g cm-2 kyr-1

0 0.694 11.99 8.32 0.64 5.33 0.2 0.698 11.99 8.37 0.65 5.44 0.4 0.753 11.99 9.03 0.71 6.41 0.6 0.680 11.99 8.15 0.68 5.54 0.8 0.761 11.99 9.13 0.71 6.48 1 0.787 11.99 9.43 0.63 5.94 1.2 0.794 11.99 9.52 0.67 6.38 1.4 0.839 11.99 10.06 0.76 7.64 1.5 0.755 11.99 9.05 1.6 0.906 11.99 10.86 0.76 8.25 1.8 1.147 11.99 13.75 2.30 31.56 1.9 15.85 1.239 11.99 14.86 2 0.781 396.83 309.81 0.69 213.77 2.2 0.745 396.83 295.53 0.68 200.96 2.4 0.741 396.83 294.22 0.71 208.89 2.6 0.753 396.83 298.93 0.76 227.19 2.8 0.786 396.83 311.74 0.72 224.46 3 0.809 396.83 320.86 0.73 234.23 3.2 1.103 396.83 437.58 1.59 693.57 3.4 0.927 396.83 368.02 0.87 320.18 3.6 0.862 396.83 341.94 0.75 256.45 3.8 1.088 396.83 431.57 0.58 250.31 4 1.061 396.83 421.17 0.75 315.88 4.2 0.969 396.83 384.52 0.92 353.76 4.4 1.134 396.83 449.95 0.90 404.95 4.6 1.137 396.83 451.17 0.92 415.07 4.75 17 396.83 4.8 1.154 396.83 458.11 0.90 412.30 5 1.091 396.83 432.85 0.95 411.21

128

Table 2.3b. Continued

Depth Age DBD LSR MAR, TOC TOC MAR m kyrs g cm-1 cm kyr-1 g cm-2 kyr-1 g cm-2 kyr-1

5.4 1.152 396.83 457.25 0.81 370.38 5.2 1.105 396.83 438.57 1.00 438.57 5.6 1.112 396.83 441.27 0.98 432.44 5.8 1.098 396.83 435.90 0.94 409.75 6 1.143 396.83 453.43 0.93 421.69 6.2 1.215 396.83 481.98 2.25 1084.46 6.4 1.155 396.83 458.33 0.98 449.16 6.6 1.161 396.83 460.89 1.03 474.71 6.8 1.166 396.83 462.83 0.95 439.69 7 1.095 396.83 434.66 0.79 341.21 7.2 1.195 396.83 474.29 0.95 450.57 7.4 1.220 396.83 484.13 0.91 440.56 7.6 1.188 396.83 471.25 0.91 428.83 7.8 1.203 396.83 477.42 0.94 448.77 8 1.201 396.83 476.53 0.97 462.23 8.2 1.198 396.83 475.55 0.93 442.26 8.3 17 396.83 8.4 1.173 396.83 465.33 1.01 469.98 8.6 1.192 396.83 473.02 0.90 425.72 8.8 1.207 396.83 479.07 0.94 450.33 9 1.221 396.83 484.38 0.92 445.63 9.2 1.176 396.83 466.69 0.99 462.02 9.4 17.74 1.228 396.83 487.29 0.87 423.94

129

Table 2.4. Sediment and organic carbon accumulation rate date for MV-51. TOC mass accumulation rate data were calculated using equation 1 in the text. MAR and TOC MAR are not reported where either dry bulk density or TOC data are absent. Depth Age DBD TOC LSR MAR, TOC MAR m kyrs g cm-1 cm kyr-1 g cm-2 kyr-1 g cm-2 kyr-1 0.01 0.669 0.73 7.28 4.87 3.56 0.21 0.694 0.76 7.28 5.05 3.84 0.41 0.691 0.70 7.28 5.03 3.52 0.61 0.718 0.74 7.28 5.23 3.87 0.81 0.698 0.71 7.28 5.08 3.61 1.01 0.712 0.64 7.28 5.19 3.32 1.21 18.6 0.723 0.79 7.28 5.27 4.16 1.34 18.405 7.28 1.41 0.703 0.90 108.77 76.42 68.78 1.61 0.701 0.84 108.77 76.26 63.68 1.81 0.745 0.91 108.77 81.07 73.77 2.01 0.730 0.79 108.77 79.43 62.75 2.21 0.790 0.69 108.77 85.94 59.30 2.41 0.801 0.41 108.77 87.18 35.74 2.61 0.741 0.63 108.77 80.61 50.78 2.81 0.754 0.70 108.77 82.07 57.45 3.01 0.744 0.68 108.77 80.94 55.04 3.21 0.742 0.67 108.77 80.73 54.09 3.41 0.752 0.61 108.77 81.84 49.92 3.51 20.4 0.764 108.77 83.08 3.61 0.709 0.57 51.20 36.28 20.50 3.81 0.761 0.64 51.20 38.96 24.93 4.01 0.773 0.58 51.20 39.56 22.94 4.21 0.822 0.61 51.20 42.11 25.48 4.41 0.775 0.67 51.20 39.67 26.58 4.61 0.773 0.68 51.20 39.57 26.91 4.81 0.957 0.44 51.20 49.00 21.56 5.01 0.796 0.69 51.20 40.75 28.12 5.21 0.811 0.67 51.20 41.53 27.83

130

Table 2.4b. Continued Depth Age DBD TOC LSR MAR, TOC MAR m kyrs g cm-1 cm kyr-1 g cm-2 kyr-1 g cm-2 kyr-1 5.41 0.840 0.70 51.20 43.03 30.12 5.61 0.849 0.74 51.20 43.48 32.17 5.81 0.888 0.72 51.20 45.46 32.73 6.01 0.852 0.68 51.20 43.60 29.65 6.21 0.874 0.64 51.20 44.75 28.64 6.41 0.871 0.67 51.20 44.57 29.87 6.61 0.872 0.70 51.20 44.64 31.25 6.81 0.874 0.69 51.20 44.74 30.87 6.91 0.877 51.20 44.91 6.935 1.058 0.42 51.20 54.17 22.75 7.01 0.872 0.64 51.20 44.65 28.58 7.135 0.873 0.75 51.20 44.70 33.53 7.21 0.883 0.68 51.20 45.20 30.74 7.41 28.017 0.904 0.82 51.20 46.28 37.95 7.61 0.868 0.76 27.56 23.91 18.17 7.81 0.884 0.76 27.56 24.37 18.52 7.94 29.94 27.56 8.01 0.879 0.75 56.52 49.69 37.26 8.085 0.854 0.96 56.52 48.29 46.36 8.21 0.904 0.80 56.52 51.07 40.86 8.285 0.967 0.87 56.52 54.67 47.56 8.41 0.962 0.76 56.52 54.38 41.33 8.485 0.982 0.84 56.52 55.53 46.65 8.61 0.895 0.73 56.52 50.58 36.92 8.685 0.896 0.71 56.52 50.62 35.94 8.81 0.903 0.85 56.52 51.04 43.38 8.895 0.891 0.78 56.52 50.35 39.27 9.01 0.882 0.82 56.52 49.86 40.88 9.085 0.905 0.84 56.52 51.14 42.96 9.21 0.893 0.86 56.52 50.46 43.39 9.285 0.882 0.78 56.52 49.86 38.89 9.61 0.782 0.81 56.52 44.19 35.79 9.81 0.823 0.83 56.52 46.52 38.38 10.01 32.617 0.880 0.84 56.52 49.72 41.76 10.21 0.904 0.82 269.18 243.25 199.47 10.41 0.902 0.87 269.18 242.86 211.29 10.61 0.917 0.85 269.18 246.86 209.83 10.81 0.897 0.97 269.18 241.39 234.15 11.01 0.918 0.87 269.18 246.98 214.87 11.21 0.911 0.91 269.18 245.10 223.04 11.41 0.914 0.91 269.18 246.13 223.98 11.61 0.926 0.91 269.18 249.24 226.81 11.81 0.918 0.86 269.18 247.19 212.58 12.01 33.36 0.908 0.86 269.18 244.49 210.26 12.21 0.869 0.94 269.18 233.90 219.87

131

Table 2.5. MS bridge data and other physical properties measured from MV-51 sediments. - - - 3 3 3 kg kg kg 3 3 3 1 1 1 depth, m m depth, MS, m %>63µm %CaCO m depth, MS, m %>63µm %CaCO m depth, MS, m %>63µm %CaCO

0.01 7.93E-08 1.60 5.44 5.21 9.85E-08 1.04 3.63 8.81 8.56E-08 1.34 3.69 0.21 1.05E-07 1.04 5.35 5.41 8.09E-08 0.85 2.86 8.895 8.93E-08 0.41 8.35E-08 5.52 3.40 5.61 8.76E-08 0.68 3.79 9.01 9.38E-08 0.79 3.18 0.61 7.12E-08 1.14 5.42 5.81 1.03E-07 1.13 3.07 9.085 6.31E-08 0.81 1.04E-07 2.15 4.36 6.01 9.36E-08 .83 3.39 9.21 8.74E-08 1.52 3.01 1.01 9.97E-08 1.31 4.16 6.21 9.40E-08 0.62 3.73 9.285 9.10E-08 3.528 1.21 1.25E-07 8.26 4.23 6.41 9.55E-08 3.47 9.61 1.19E-07 1.50 4.24 1.41 7.86E-08 0.77 3.30 6.61 9.81E-08 0.82 2.60 9.81 1.52E-07 3.64 1.61 2.19E-07 3.32 4.31 6.81 9.00E-08 0.87 3.33 10.01 2.52E-07 0.51 3.23 1.81 1.03E-07 0.84 3.57 6.86 6.40E-08 0.98 3.04 10.21 2.31E-07 2.95 2.01 9.92E-08 0.93 5.12 6.91 7.26E-08 1.86 2.65 10.41 2.53E-07 1.00 3.20 2.21 8.97E-08 1.87 8.81 6.935 1.46E-07 9.80 1.21 10.61 1.68E-07 3.14 2.41 7.13E-07 6.30 4.29 7.01 8.85E-08 0.69 4.49 10.81 2.17E-07 3.21 2.61 7.31E-08 1.44 7.44 7.135 7.74E-08 3.63 11.01 1.64E-07 0.88 3.57 2.81 6.87E-08 1.00 7.90 7.21 9.04E-08 3.96 11.21 2.37E-07 2.94 3.01 7.04E-08 1.60 6.94 7.41 9.19E-08 0.74 2.75 11.41 2.42E-07 0.70 3.50 3.21 8.09E-08 0.92 5.04 7.61 7.51E-08 3.61 11.61 2.23E-07 3.17 3.41 1.02E-07 2.28 2.96 7.81 9.95E-08 1.29 4.60 11.81 2.61E-07 3.40 3.61 7.09E-08 0.77 4.47 8.01 8.28E-08 3.26 12.01 2.91E-07 1.31 3.44 3.81 8.89E-08 0.60 4.29 8.085 8.30E-08 1.01 2.95 12.21 1.53E-07 3.15 4.01 9.97E-08 0.56 3.58 8.21 8.88E-08 3.27 4.21 8.29E-08 4.10 8.285 9.95E-08 3.3 4.41 9.22E-08 1.19 4.91 8.41 8.59E-08 4.01 4.61 8.82E-08 0.62 3.55 8.485 8.27E-08 1.28 3.70 4.81 2.10E-07 20.44 1.54 8.61 1.10E-07 3.57 5.01 8.54E-08 0.56 4.35 8.685 7.59E-08

132

Table 2.6. MS bridge data for MV-54, other physical properties such as dry bulk density and TOC for this core are reported in Table 3. depth, m MS, m3 kg-1 depth, m MS, m3 kg-1 depth, m MS, m3 kg-1 0.01 1.92E-07 4.61 1.80E-07 10.21 2.91E-07 0.11 1.88E-07 4.81 1.23E-07 10.41 2.84E-07 0.21 2.25E-07 5.01 1.28E-07 10.61 2.49E-07 0.31 1.49E-07 5.21 1.36E-07 10.81 2.18E-07 0.41 2.65E-07 5.41 1.30E-07 11.01 1.76E-07 0.51 1.82E-07 5.61 1.46E-07 11.21 3.43E-07 0.61 2.81E-07 5.81 1.56E-07 11.41 3.88E-07 0.81 3.45E-07 6.01 1.82E-07 11.61 3.17E-07 1.01 2.17E-07 6.21 2.67E-07 11.81 2.01E-07 1.21 1.85E-07 6.41 1.34E-07 12.01 2.23E-07 1.41 2.41E-07 6.61 1.81E-07 12.11 1.94E-07 1.51 3.76E-07 6.81 1.64E-07 1.61 3.78E-07 7.01 1.04E-07 1.81 3.00E-07 7.21 1.32E-07 1.91 7.92E-07 7.41 1.18E-07 2.01 1.22E-07 7.61 1.11E-07 2.21 1.04E-07 7.81 9.77E-08 2.41 1.12E-07 8.01 8.25E-08 2.61 9.07E-08 8.21 9.32E-08 2.81 6.07E-08 8.41 9.56E-08 3.01 1.33E-07 8.61 8.86E-08 3.21 1.13E-07 8.81 1.14E-07 3.41 2.56E-07 9.01 1.33E-07 3.61 1.09E-07 9.21 1.96E-07 3.81 1.41E-07 9.41 2.37E-07 4.01 1.60E-07 9.61 3.11E-07 4.21 1.48E-07 9.81 2.85E-07 4.41 1.54E-07 10.01 2.52E-07

133

Table 2.7. Rock-Eval pyrolysis data measured from MV-51 sediments. S1, S2, and S3 refer to “source” data generated during pyrolysis and are in mg/g units. Tmax is the thermal maximum in degrees Celsius. HI is the hydrogen index calculated using the S2 and TOC values (Table 4) with equation 2, in mgHC gTOC-1 units. depth, m S1 S2 S3 Tmax HI depth, m S1 S2 S3 Tmax HI 0.01 0.42 0.56 2.51 393 76.71 6.41 0.75 0.90 2.00 414 134.33 0.21 0.52 0.63 2.84 383 82.89 6.61 0.44 0.65 2.04 414 92.86 0.41 0.55 0.69 2.50 394 98.57 6.81 0.18 0.30 1.99 388 43.48 0.61 0.55 0.75 2.52 392 101.35 7.01 0.50 0.66 2.34 408 103.13 0.81 0.55 0.74 2.33 398 104.23 7.21 0.38 0.54 2.19 396 79.41 1.01 0.46 0.56 2.16 392 87.50 7.61 1.64 1.16 1.65 399 152.63 1.21 0.55 0.77 2.22 394 97.47 7.81 0.98 0.86 2.15 399 113.16 1.41 0.44 0.60 2.97 399 66.67 8.01 0.68 0.61 2.11 403 81.33 1.61 0.59 0.80 2.55 397 95.81 8.21 0.91 0.86 1.96 403 107.50 1.81 0.56 0.74 2.62 394 81.32 8.41 0.90 0.73 1.68 396 96.05 2.01 0.54 0.59 2.95 392 74.68 8.61 0.80 0.52 2.48 387 71.23 2.21 0.56 0.56 2.84 402 81.16 8.81 0.02 0.06 0.45 492 7.06 2.61 0.26 0.30 2.74 391 47.62 9.01 0.74 0.59 2.01 395 71.95 2.81 0.41 0.48 2.56 398 68.57 9.21 0.19 0.31 1.88 400 36.05 3.21 0.17 0.35 2.32 507 52.24 9.61 0.41 0.46 2.30 400 56.79 3.41 0.14 0.35 1.95 557 57.38 9.81 0.54 0.65 2.01 555 78.79 3.61 0.12 0.17 3.09 384 30.09 10.01 0.62 0.64 2.03 410 76.19 3.81 0.41 0.52 2.41 397 81.25 10.21 0.40 0.52 2.12 398 63.41 4.01 0.13 0.15 2.81 418 25.86 10.41 0.72 0.70 1.99 399 80.46 4.21 0.19 0.33 2.31 406 54.55 10.61 0.73 0.76 2.01 401 89.41 4.41 0.34 0.45 2.21 403 67.16 10.81 0.71 0.58 1.91 384 59.79 4.61 0.21 0.32 2.21 462 47.06 11.01 1.01 0.84 2.21 402 96.55 5.01 0.13 0.12 2.24 393 17.39 11.21 0.18 0.30 2.17 401 32.97 5.21 0.53 0.72 2.38 407 107.46 11.41 1.10 0.73 2.14 416 80.22 5.41 0.29 0.54 1.98 451 77.14 11.61 1.36 0.91 1.75 415 100.00 5.61 0.17 0.27 2.03 392 36.49 11.81 1.43 0.86 1.84 412 100.00 5.81 0.33 0.52 1.93 415 72.22 12.01 0.22 0.35 2.05 380 40.70 6.01 0.26 0.34 2.17 394 50.37 12.21 1.06 0.83 1.91 418 88.30 6.21 0.20 0.29 2.15 390 45.31

134

Table 2.8. Elemental and stable isotopic data measured from MV-51 sediments. Equations 3-5 were used to calculate the fterrigenous and % terrigenous TOC values. % terrigenous 13 depth, m δ C C/N fterrigenous TOC 0.01 -23.91 20.38 0.57 41.50 0.41 -22.90 10.49 0.40 28.04 0.61 -23.74 11.30 0.54 40.01 1.01 -24.24 10.27 0.62 39.89 1.21 -23.85 13.12 0.56 44.12 1.41 -23.43 11.88 0.49 43.93 1.61 -23.60 11.26 0.52 43.19 1.81 -24.59 11.27 0.68 62.10 2.01 -23.35 11.41 0.47 37.48 2.41 -23.63 9.74 0.52 21.36 3.01 -22.64 11.07 0.36 24.26 3.41 -22.84 11.81 0.39 23.78 4.01 -22.01 12.83 0.25 14.62 4.81 -23.35 11.95 0.48 20.93 5.01 -23.19 12.24 0.45 30.91 5.61 -23.57 13.29 0.51 37.89 6.21 -23.30 13.65 0.47 29.90 6.81 -23.20 9.66 0.45 31.00 6.93 -28.93 17.51 7.41 -23.58 18.51 0.51 42.14 8.09 -23.98 19.31 0.58 55.60 8.49 -23.66 18.49 0.53 44.28 9.01 -23.82 16.94 0.55 45.39 10.01 -24.58 19.24 0.68 57.09 11.01 -23.73 19.55 0.54 46.77

135 Table 2.9. Benthic foraminiferal densities (BF#, # g sediment-1) from MV-51 sediments for the 12 most abundant species.

depth, m Total BF# peregrina/ U. hispida Bolivina robusta Bulimina aculeata Cibicidoides pachyderma Sphaeroidina bulloides Gavelinopsis translucens Globocassidulin a subglobosa Oridorsalis. spp. O. tener Melonis barleeanus Bolivinita quadrilatera Pullenia spp.

0.01 142.07 21.38 34.25 20.00 5.06 4.83 13.33 9.89 1.84 1.61 1.15 1.15 3.45 0.21 137.11 16.94 32.15 14.40 10.25 10.14 4.26 14.52 1.96 0.23 0.69 4.15 0.23 0.41 69.61 8.43 17.33 6.44 4.45 8.13 1.84 8.59 0.77 1.38 0.15 1.23 0.31 0.61 56.64 4.30 16.55 7.16 0.64 3.34 3.66 10.02 0.95 0.80 0.32 0.32 0.16 1.01 150.14 10.10 64.90 11.71 0.40 6.60 12.93 13.87 2.29 5.25 1.48 1.75 0.81 1.41 83.24 3.75 44.61 7.40 0.71 2.74 3.14 7.50 1.32 2.03 0.20 0.81 0.41 1.61 172.34 17.48 81.14 15.73 1.31 15.15 5.24 12.82 3.64 4.08 0.87 1.60 0.29 1.81 98.85 7.84 59.64 6.85 0.20 1.39 2.18 8.14 0.89 1.79 0.60 0.00 0.30 2.01 80.64 6.57 36.50 4.73 1.37 1.83 3.97 9.77 1.53 1.53 0.15 0.46 0.46 2.21 130.63 15.75 63.84 12.23 1.69 1.69 6.33 9.84 1.69 1.69 0.70 0.00 0.00 2.41 84.28 2.80 34.73 5.93 0.99 1.48 9.05 8.72 1.48 1.81 0.66 0.99 0.16 3.01 140.46 2.23 26.67 4.89 0.56 3.35 2.93 3.77 3.35 3.07 0.84 0.84 0.70 3.21 81.47 3.12 35.32 9.69 1.15 4.60 2.30 12.16 1.64 0.99 0.66 2.30 0.00 3.41 41.35 0.56 16.96 4.35 0.98 1.12 1.68 4.35 2.38 0.84 1.12 1.12 0.14 4.01 91.41 1.28 38.83 9.75 2.55 1.18 7.84 13.49 2.01 2.46 0.27 2.55 0.73 4.41 121.13 2.48 69.18 12.40 0.00 1.74 6.08 7.44 1.86 3.97 0.50 0.99 0.12 4.61 49.84 1.12 26.90 6.82 0.00 1.24 2.11 1.74 1.49 1.36 0.12 0.12 0.25 4.81 34.50 0.44 23.23 1.32 0.09 0.79 2.38 2.55 0.44 0.53 0.00 0.09 0.26 5.01 67.55 3.08 34.37 6.64 0.12 1.78 6.04 3.91 2.73 1.07 0.00 0.36 0.47 5.61 70.71 0.23 39.37 2.94 0.45 5.32 3.39 3.85 2.15 2.04 0.45 3.96 0.79 6.21 91.35 6.16 53.56 4.65 0.20 1.11 4.75 4.14 0.81 2.43 0.40 0.91 0.30 6.61 151.95 16.11 79.82 10.70 24.55 0.00 2.64 5.54 0.38 1.26 0.13 0.38 0.88 6.81 282.70 61.20 145.62 11.60 39.82 0.11 6.60 6.60 1.93 1.25 0.00 2.16 0.23 6.91 128.19 33.26 53.81 8.93 15.93 0.43 1.61 3.77 0.32 1.83 0.00 0.00 1.18 6.935 67.37 7.30 33.22 2.64 6.83 2.33 3.42 3.57 0.78 0.93 0.00 0.00 0.93 7.01 134.78 26.90 64.27 13.45 5.84 3.40 4.08 4.62 2.31 0.27 0.00 0.54 0.82 7.41 144.70 34.21 62.85 10.68 17.77 11.81 3.12 2.65 3.21 0.47 0.00 1.13 0.57 8.085 188.93 30.62 97.92 9.17 16.45 10.96 5.86 5.77 0.95 0.85 0.00 1.70 2.65 8.485 241.12 19.80 135.18 10.13 18.09 21.16 9.79 10.13 1.25 0.57 0.68 0.23 0.23 9.01 218.79 31.00 101.67 7.83 16.08 26.31 10.75 6.89 2.51 1.36 0.00 1.36 0.42 9.61 160.10 29.44 51.59 9.23 10.97 23.90 10.64 5.32 3.15 1.96 0.00 1.52 0.43 10.01 199.75 36.22 46.34 27.43 10.40 31.31 9.93 10.31 3.50 1.61 0.00 2.27 1.13 10.41 154.32 34.14 27.95 20.96 7.09 27.45 8.38 4.99 1.50 0.60 0.00 4.29 1.50 11.01 167.91 36.34 35.55 20.96 12.20 34.48 4.51 7.69 1.59 2.03 0.00 1.59 0.53 11.41 137.77 23.32 40.07 16.95 8.32 23.53 4.11 4.01 1.75 1.23 0.31 3.39 1.34 12.01 209.15 40.50 67.09 25.83 8.18 24.45 9.25 10.16 1.83 2.06 0.00 3.21 0.92

136 APPENDIX C: CORE IMAGES OF MV-51JPC

137 APPENDIX D

PERMISSION REQUEST

1. Original Email Request

Subject: Permission Request (Urgent) From: "Lawrence Febo" Date: Mon, 12 Nov 2007 12:13:00 -0600 To: "Eric Moore"

Hi Eric,

I recently defended my dissertation and I am preparing my final copy for the graduate school. Do I need permission to publish the manuscript I submitted to JGR-Earth Surface as a modified chapter in my doctoral thesis? Please let me know as soon as possible as I need to submit it to the graduate school this Wednesday.

Thank you,

Lawrence

2. Email Response

Subject: permission From: “Michael Connolly” Date: Tuesday, November 13, 2007 8:17:16 AM To: [email protected]

We are pleased to grant permission for the use of the material requested for inclusion in your thesis. The following non-exclusive rights are granted to AGU authors:

• All proprietary rights other than copyright (such as patent rights). • The right to present the material orally. • The right to reproduce figures, tables, and extracts, appropriately cited. • The right to make hard paper copies of all or part of the paper for classroom use. • The right to deny subsequent commercial use of the paper.

Further reproduction or distribution is not permitted beyond that stipulated. The copyright credit line should appear on the first page of the article or book chapter. The following must also be included, “Reproduced by permission of American Geophysical Union.” To ensure that credit is given to the original source(s) and that authors receive full credit through appropriate citation to their papers, we recommend that the full bibliographic reference be cited in the reference list. The standard credit line for journal

138 articles is: "Author(s), title of work, publication title, volume number, issue number, page number(s), year. Copyright [year] American Geophysical Union."

If an article was placed in the public domain, in which case the words “Not subject to U.S. copyright” appear on the bottom of the first page or screen of the article, please substitute “published” for the word “copyright” in the credit line mentioned above.

Copyright information is provided on the inside cover of our journals. For permission for any other use, please contact the AGU Publications Office at AGU, 2000 Florida Ave., N.W., Washington, DC 20009.

Michael Connolly Journals Publications Specialist

139 APPENDIX E

COMPREHENSIVE REFERENCE LIST

Alldredge, A.L. (1979), The chemical composition of macroscopic aggregates in two neritic seas, Limnology and Oceanography, 24, 855-866.

Alldredge, A.L. and M.W. Silver (1988), Characteristics, dynamics and significance of marine snow, Progress in Oceanography, 20, 41-82.

Aller, R.C. (1998), Mobile deltaic and continental shelf muds as suboxic, fluidized bed reactors, Marine Chemistry, 61, 143-155.

Aller, R.C. and N.E. Blair (2004), Early diagenetic remineralization of sedimentary organic C in the Gulf of Papua deltaic complex (Papua New Guinea): Net loss of terrestrial C and diagenetic fractionation of C isotopes, Geochimica et Cosmochimica Acta, 68(8), 1815-1825.

Alongi, D.M. and A.I. Robertson (1995), Factors regulating benthic food chains in tropical river deltas and adjacent shelf areas, Geo-Marine Letters, 15, 145-152.

Altenbach, A.V. and M. Sarnthein (1989), Productivity record in benthic foraminifera, in Productivity of the Ocean: Present and Past, edited by W.H. Berger, V.S. Smetacek, and G. Wefer, pp. 255-269, John Wiley & Sons Limited, New York, New York.

Amo, M. and M. Minagawa (2003), Sedimentary record of marine and terrigenous organic matter delivery to the Shatsky Rise, western North Pacific, over the last 130 kyrs, Organic Geochemistry, 34, 1299-1312.

An, Z. (2000), The history and variability of the East Asian paleomonsoon climate, Quaternary Science Reviews, 19, 171-187.

Andrews, J.T. and J.A. Stravers (1993), Magnetic susceptibility of late Quaternary marine sediments, Frobisher Bay, N.W.T.: An indicator of changes in provenance and processes, Quaternary Science Reviews, 12, 157-167.

Balsam, W., B.B. Ellwood, and J. Ji (2005), Direct correlation of the marine oxygen isotope record with Loess Plateau iron oxide and magnetic susceptibility records, Palaeogeography, Palaeoclimatology, Palaeoecology, 221, 141-152.

Banerjee, S.K. (1996), Sediment reveals early Holocene climate change in China, Eos Transactions AGU, 77(1), 3, 5.

140 Batten, D.J. (1996), Palynofacies and palaeoenvironmental interpretation, in Palynology: Principles and Applications, edited by D.C. McGregor, American Association of Stratigraphic Palynologists Foundation, pp. 1011–1064.

Beaufort, L., T. de Garidel-Thoron, A.C. Mix, and N.G. Pisias (2001), ENSO-like Forcing on Oceanic Primary Production During the Late Pleistocene, Science, 293, 2440- 2444.

Belperio, A.P., and Searle, D.E. (1988), Terrigenous and Carbonate Sedimentation in the Great Barrier Reef Province, in Carbonate - Clastic Transitions, edited by L.J. Doyle and H.H. Roberts, Elsevier, pp. 143-174.

Bentley, S.J., Z. Muhammad, L.Patterson, J. Dickens, B.N. Opdyke, L. Peterson, and A.W. Droxler (2006), Modern and Pleistocene turbidite sedimentation in the Gulf of Papua S2S study area: Implications for modulation of sediment sources by oceanic and climatic processes. AAPG Annual Meeting, Houston Texas.

Berger, W.H. and J.C. Herguera (1992), Reading the sedimentary record of the ocean’s productivity. in Primary Productivity and Biogeochemical Cycles in the Sea, edited by P.F. Falkowski and A.D. Woodhead, pp. 455-486, Plenum Press, New York, New York.

Berner, R.A. (2003), The long-term carbon cycle, fossil fuels and atmospheric composition, Nature, 426: 323-326.

Bird, M.I., G.J. Brunskill, and A.R. Chivas (1995), Carbon-isotope composition of sediments from the Gulf of Papua, Geo-Marine Letters, 15, 153-159.

Bjerknes, J. (1969), Atmospheric teleconnections from the equatorial Pacific, Monthly Weather Review, 97, 163-172.

Blake, D.H. (1976), Madilogo, a late Quaternary volcano near Port Moresby, Papua New Guinea, in Volcanism in Australasia, edited by R.W. Johnson, pp. 253-258, Elsevier Scientific Publishing Company, Amsterdam, The Netherlands.

Bouloubassi, I, J. Rullkötter, and P.A. Meyers (1999), Origin and transformation of organic matter in Pliocene-Pleistocene Mediterranean sapropels: Organic geochemical evidence reviewed, Marine Geology, 153, 177-197.

Bowler, J.M., G.S. Hope, J.N. Jennings, G. Singh, and D. Walker (1976), Late Quaternary climates of Australia and New Guinea, Quaternary Research, 6, 359-394.

Broecker, W. (1996), Glacial climate in the tropics, Science, 272, 1902-1904.

Broecker, W., S. Barker, E. Clark, I. Hajdas, G. Bonani, and L. Stott (2004), Ventilation of the glacial deep Pacific Ocean, Science, 306, 1169-1172.

141 Bronk Ramsey, C. (1995), Radiocarbon calibration and analysis of stratigraphy: The OxCal Program, Radiocarbon, 37(2), 425-430.

Brunskill, G.J. (2004), New Guinea and its coastal seas, a testable model of wet tropical coastal processes: an introduction to Project TROPICS, Continental Shelf Research, 24, 2273-2295.

Brunskill, G.J., K.J. Woolfe, and I. Zagorskis (1995), Distribution of riverine sediment chemistry on the shelf, slope and rise of the Gulf of Papua, Geo-Marine Letters, 15 (3-4), 160-165.

Brunskill, G.J., I. Zagorskis, and J. Pfitzner (2003), Geochemical mass balance for lithium, boron, and strontium in the Gulf of Papua, Papua New Guinea (Project TROPICS), Geochimica et Cosmochimica Acta, 67(18), 3365-3383.

Bulfinch, D.L., M.T. Ledbetter, B.B. Ellwood, and W.L. Balsam (1982), The high velocity core of the Western Boundary Undercurrent at the base of the U.S. Continental Rise, Science, 215, 970-973.

Burns, K.A., P. Greenwood, R. Benner, D. Brinkman, G. Brunskill, S. Codia, and I. Zagorskis (2004), Organic biomarkers for tracing carbon cycling in the Gulf of Papua (Papua New Guinea), Continental Shelf Research, 24, 2373-2394.

Carson, B.E., R.M. Leckie, J.M. Francis, A.W. Droxler, G.R. Dickens, S.J. Bentley, L.C. Peterson, and B.N. Opdyke (2007), Foraminiferal Populations of the Southern Ashmore Trough (Gulf of Papua) Mixed Siliciclastic-Carbonate System over the past 75 kyrs, Journal of Geophysical Research - Earth Surface.

Chappell, J. and N.J. Shackleton (1986), Oxygen isotopes and sea level, Nature, 321, 137-140.

Chartres, C.J. and C.F. Pain (1984), A climosequence of soils on Late Quaternary volcanic ash in highland Papua New Guinea, Geoderma, 32, 131-155.

Clark, P.U., A.M. McCabe, A.C. Mix, and A.J. Weaver (2004), Rapid rise of sea level 19,000 years ago and its global implications, Science, 304, 1141-1144.

Clemens, S., P. Wang, and W. Prell (2003), Monsoons and global linkages on Milankovitch and sub-Milankovitch time scales, Marine Geology, 201, 1-3.

Corliss, B.H. (1985), Microhabitats of benthic foraminifera within deep-sea sediments, Nature, 314, 435-438.

Corliss, B.H. (1991), Morphology and microhabitat preferences of benthic foraminifera from the northwest Atlantic Ocean, Marine Micropaleontology, 17, 195-236.

142 Daniell, J. (2007), Development of a 100-m bathymetry grid for the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Davies, H.L. and I.E. Smith (1971), Geology of Eastern Papua, Geological Society of America Bulletin, 82, 3299-3312.

Delvaux, D., H. Martin, P. Leplat, and J. Paulet (1990), Geochemical characterization of sedimentary organic matter by means of pyrolysis kinetic parameters, Organic Geochemistry, 16(1-3), 175-187.

de Garidel-Thoron, T., L. Beaufort, B.K. Linsley, and S. Dannenmann (2001), Millenial- scale dynamics of the East Asian winter monsoon during the last 200,000 years, Paleoceanography, 16, 1-12.

deMenocal, P., J. Bloemendal, and J. King (1991), A rock-magnetic record of monsoonal dust deposit ionto the Arabian Sea: Evidence for a shift in the mode of deposition at 2.4 Ma, in Proceedings ODP, Scientific Results 117, edited by W. Prell and L. Niitsuma, pp. 389-407.

De Stigter, H.C., F.J. Jorissen, and G.J. van der Zwann (1998), Bathymetric distribution and microhabitat partitioning of live (rose bengal stained) benthic foraminifera along a shelf to bathyal transect in the southern Adriatic Sea, Journal for Foraminiferal Research, 28(1), 40-65.

Dickens, G.R., A.W. Droxler, S.J. Bentley, L.C. Peterson, B.N. Opdyke, L. Beaufort, J. Daniell, L.A. Febo, J. Francis, P.T. Harris, S. Jorry, G. Mallarino, M. McFadden, Z. Muhammad, B. Carson, L. Patterson, E. Tcherepanov, and C.A. Zarickian (2006), A preliminary late quaternary view of sedimentary sinks in the Papua New Guinea Source- to-Sink focus area, MARGINS Newsletter, 16, 1-5.

Droxler, A., G. Mallarino, J.M. Francis, J. Dickens, L. Beaufort, S. Bentley, L. Peterson, B. Opdyke (2006), Early Part of Last Deglaciation: A Short Interval Favorable for the Building of Coralgal Edifices on the Edges of Modern Siliciclastic Shelves (Gulf of Papua and Gulf of Mexico), AAPG Annual Meeting.

Ellwood, B.B., W.L. Balsam, and H.H. Roberts (2006), Gulf of Mexico sediment sources and sediment transport trends from Magnetic Susceptibility measurements of surface samples, Marine Geology, 230, 237-248.

Ellwood, B.B., C.E. Brett, and W.D. MacDonald (2007), Magnetosusceptibility Stratigraphy of the Upper Ordovician Kope Formation, Northern Kentucky, Palaeogeography, Palaeoclimatology, Palaeoecology, 243, 42-54.

Ellwood, B.B., R.E. Crick, and A. Hassani (1999), The magneto-susceptibility event and cyclostratigraphy (MSEC) method used in geological correlation of Devonian rocks of the Anti-Atlas Morocco, AAPG Bulletin, 83(7), 1119-1134.

143

Ellwood, B.B., R.E. Crick, A. Hassani, S.L. Benoist, and R.H. Young (2000), Magnetosusceptibility event and cyclostratigraphy method applied to marine rocks: detrital input versus carbonate productivity, Geology, 28(12), 1135-1138.

Ellwood, B.B. and M.R. Fisk (1977a), Anisotropy of magnetic susceptibility variations in a single Icelandic columnar basalt, Earth and Planetary Science Letters, 35, 116-122.

Ellwood, B.B. and M.T. Ledbetter (1977b), Antarctic bottom water fluctuation in the Vema Channel: Effects of velocity changes on particle alignment and size, Earth Planetary Science Letters, 35, 189-198.

Ellwood, B.B., W.D. MacDonald, C. Wheeler, and S.L. Benoist (2003), The K-T boundary in Oman: identifed using magnetic susceptibility field measurements with geochemical confirmation, Earth and Planetary Science Letters, 206, 529-540.

Ellwood, B.B. K.M. Petruso, and F.B. Harrold (1996), The utility of magnetic susceptibility for detecting paleoclimatic trends and as a stratigraphic tool: an example from Konispol Cave sediments, SW Albania, Journal of Field Archaeology, 23, 263-271.

Ellwood, B.B., J.H. Tomkin, K.T. Ratcliffe, M. Wright, and A.M. Kafafy (2007), Magnetic Susceptibility and geochemistry for the Cenomanian/Turonian boundary GSSP with correlation to time equivalent core. GSA Annual Meeting, Denver.

Ellwood, B.B., J.H. Tomkin, L.A. Febo, and C.N. Stuart, Jr. In Review, Time Series Analysis of Magnetic Susceptibility Variations in Deep Marine Sediments: A Test Using Upper Danian-Lower Selandian Proposed GSSP, Spain, Palaeogeography, Palaeoclimatology, Palaeoecology.

Emerson, S. and J.I. Hedges (1988), Processes controlling the organic carbon content of open ocean sediments, Paleoceanography, 3(5), 621-634.

Espitalié, J. and M.L. Bordenave (1993), Rock-Eval pyrolysis, in Applied Petroleum Geochemistry, edited by M.L. Bordenave, pp. 237-261, Editions Technip, Paris, France.

Espitalié, J., M. Madec, and B. Tissot (1980), Role of mineral matrix in kerogen pyrolysis: Influence of petroleum generation and migration, AAPG Bulletin, 64, 59-66.

Fairbanks, R.G., R.A. Mortlock, T.C. Chiu, L. Cao, A. Kaplan, T.P. Guilderson, T.W. Fairbanks, A.L. Bloom, P.M. Grootes, M.J. Nadeau (2005), Radiocarbon calibration curves spanning 0 to 50,000 years BP based on paired 230Th/234U/238U and 14C dates on pristine corals, Quaternary Science Reviews, 24, 1781-1796.

Febo, L.A., S.J. Bentley, J.H. Wrenn, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke 2006. Recent to Late Pleistocene Sedimentary Organic Matter in the Gulf of Papua, Papua-New Guinea. AAPG Annual Meeting Abstracts.

144

Febo, L.A., S.J. Bentley, J.H. Wrenn, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke (2007), Late Pleistocene and Holocene Sedimentation, Organic-Carbon Delivery, and Paleoclimatic Inferences on the Continental Slope of the Northern Pandora Trough, Gulf of Papua. Journal of Geophysical Research - Earth Surface.

Francis, J. M., J. Daniell, A. W. Droxler, G. R. Dickens, S. J. Bentley, L. C. Peterson, B. Opdyke, and L. Beaufort (2007), Deepwater geomorphology and sediment pathways of the mixed siliciclastic/carbonate system, Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Francis, J. M., G.B. Dunbar, G.R. Dickens, I.A. Sutherland, and A.W. Droxler (2007), Siliciclastic sediment across the north queensland margin (Australia): A Holocene perspective on reciprocal versus coeval deposition in tropical mixed siliciclastic- carbonate systems, Journal of Sedimentary Research, 77, 572-586.

Francis, J.M., B. Olson, G.R. Dickens, A.W. Droxler, L. Beaufort, T. De Garidel-Thoron, S. Bentley, L.C. Peterson, B. Opdyke (2006), High Siliciclastic Flux during the Last Sea Level Transgression, Ashmore Trough, Gulf of Papua, AAPG Annual Meeting.

Flood, R., and D. Piper (1997), Amazon Fan sedimentation: the relationship to equatorial climate change, continental denudation, and sea-level fluctuations, in Proceedings ODP, Scientific Results 155, edited by R.D. Flood, D.J.W. Piper, A. Klaus, and L.C. Peterson, pp. 653-675.

Garrett-Jones, S.E. (1979), Holocene vegetation and lake sedimentation in the Markham Valley, Papua New Guinea, Ph.D. Dissertation, 420 pp., Australia National University, Canberra, Australia.

Gooday, A.J. (1988), A response by benthic Foraminifera to the deposition of phytodetritus in the deep sea, Nature, 332, 70-73.

Gooday, A.J. and J.A. Hughes (2002), Foraminifera associated with phytodetritus deposits at a bathyal site in the northern Rockall Trough (NE Atlantic): seasonal contrasts and a comparison of stained and dead assemblages, Marine Micropaleontology, 46, 83- 110.

Goni, M.A., N. Monacci, R. Gisewhite, A. Ogston, J. Crockett, and C. Nittrouer (2006), Distribution and sources of particulate organic matter in the water column and sediments of the Fly River Delta, Gulf of Papua (Papua New Guinea), Estuarine Coastal and Shelf Science, 69, 225-245.

Hardy, M. J. (2000), Origin, distribution, and degradation of sedimentary organic matter in a modern tropical deltaic system (Mahakam Delta, Borneo, Indonesia), unpublished Dissertation, Louisiana State University, 368 p.

145 Harris, P.T., E.K. Baker, A.R. Cole, and S.A. Short (1993), A preliminary study of sedimentation in the tidally dominated Fly River Delta, Gulf of Papua, Continental Shelf Research, 13, 441-472.

Harris, P.T., C.B. Pattiaratchi, J.B. Keene, R.W. Dalrymple, J.V. Gardner, E.K. Baker, A.R. Cole, D. Mitchell, P. Gibbs, and W.W. Schroeder (1996), Late Quaternary deltaic and carbonate sedimentation in the Gulf of Papua foreland basin: Response to sea-level change, Journal Sedimentary Research, 66(4), 801-819.

Harrison, S.P. and J. Dodson (1993), Climates of Australia and New Guinea since 18,000 yr B.P., in Global Climates Since the Last Glacial Maximum, edited by H.E. Wright, J.E. Kutzbach, T. Webb, W.F. Ruddiman, F.A. Street-Perrott, and P.J. Bartlein, pp. 265-293, University of Minnesota Press, Minneapolis, Minnesota.

Hart, G.F. (1986), Origin and classification of organic matter in clastic systems, Palynology, 10, 1-23.

Hartnett, H.E., R.G. Keil, J.I. Hedges, and A.H. Devol (1998), Influence of oxygen exposure time on organic carbon preservation in continental margin sediments, Nature, 391, 572-574.

Hattin, D.E. (1982), Stratigraphy and Depositional Environments of Smoky Hill Chalk Member, Niobrara Chalk (Upper Cretaceous) of the Type Area, Western Kansas, Kansas Geological Survey, Bulletin 225, 108p.

Heath, G.R., T.C. Moore, and J.P. Dauphin (1977), Organic carbon in deep-sea sediments, in The Fate of Fossil Fuel CO2 in the Oceans, edited by N.R. Anderson and A. Malahoff, Plenum, New York, pp. 605-626.

Hedges, J.I. and R.G. Keil (1995), Sedimentary organic matter preservation: an assessment and speculative assessment, Marine Chemistry, 49, 81-115.

Hedges, J.I., R.G. Keil, and R. Benner (1997), What happens to terrestrial organic matter in the ocean?, Organic Geochemistry, 27: 195-212.

Hedges, J.I., F.S. Hu, A.H. Devol, H.E. Hartnett, E. Tsamakis, and R.G. Keil (1999), Sedimentary organic matter preservation: A test for selective degradation under oxic conditions, American Journal of Science, 299, 529-555.

Heller, F. and M.E. Evans (1995), Loess magnetism, Review of Geophysics, 33, 211-240.

Hemer, M.A., P.T. Harris, R. Coleman, and J. Hunter (2004), Sediment mobility due to currents and waves in the Torres Strait - Gulf of Papua region, Continental Shelf Research, 24, 2297-2316.

Herguera, J.C. (1992), Deep-sea benthic foraminifera and biogenic opal: Glacial to

146 postglacial productivity changes in the western equatorial Pacific, Marine Micropaleontology, 19(1/2), 79-98.

Herguera, J.C. and W.H. Berger (1994), Glacial to post glacial drop on productivity in the western equatorial Pacific: mixing rate vs. nutrient concentrations, Geology, 22, 629-632.

Hess, S., W. Kuhnt, S. Hill, M.A. Kaminski, A. Holbourn, and M. de Leon (2001), Monitoring the recolonization of the Mt Pinatubo 1991 ash layer by benthic foraminifera, Marine Micropaleontology, 43, 119-142.

Hope, G.S. (1976), The vegetational history of Mt. Wilhelm, Papua New Guinea, Ecology, 64, 627-664.

Hope, G. and J. Tulip (1994), A long vegetation history from lowland Irian Jaya, Indonesia, Palaeogeography, Palaeoclimatology, Palaeoecology, 109, 385-398.

Hovius, N. (1998), Controls on sediment supply by large rivers, in Relative Role of Eustacy, Climate, and Tectonism in Continental Rocks, edited by K.W. Shanley and P.J. McCabe, pp. 3-16.

Hrouda, F. (1994), A technique for the measurement of thermal changes of magnetic susceptibility of weakly magnetic rocks by the CS-2 apparatus and the KLY-2 Kappabridge, Geophysical Journal International, 118, 604-612.

Hughen, K., S. Lehman, J. Southon, J. Overpeck, O. Marchal, C. Herring, and J. Turnbull (2004), 14C activity and global carbon cycle changes over the past 50,000 years, Science, 303, 202-207.

Ittekkot, V. (1988), Global trends in the nature of organic matter in river suspensions, Nature, 332, 436-438.

Jarzie, D.M. (1991), Total organic carbon (TOC) analysis, in Source and Migration processes and evaluation techniques. Treatise of Petroleum Geology, Handbook of Petroleum Geology, edited by R.K. Merrill, American Association of Petroleum Geologists, Tulsa, pp. 113-118.

Jones, G.A. and P. Kaiteris (1983), A vacuum-gasometric technique for rapid and precise analysis of calcium carbonate in sediments and soils, Journal of Sedimentary Petrology, 53, 655-659.

Jorissen, F.J., D.M. Barmawidjaja, S. Puskaric, and G.J. van der Zwann (1992), Vertical distribution of foraminifera in the northern Adriatic Sea: The relation with the organic flux, Marine Micropaleontology, 1(1/2), 131-146.

Jorissen, F.J., H.C. De Stigter, and J.G.V. Widmark (1995), A conceptual model explaining benthic foraminiferal microhabitats, Marine Micropaleontology, 26, 3-15.

147

Jorissen, F.J., I. Wittling, J.P. Peypouquet, C. Rabouille, and J.C. Relexans (1998), Live benthic foraminiferal faunas off Cape Blanc, MW-Africa: Community structure and microhabitats, Deep Sea Research Pt. 1, 45, 2157-2188.

Jorry, S.J., A.W. Droxler, G. Mallarino, G.R. Dickens, S.J. Bentley, L. Beaufort, and L.C. Peterson (2007), Bundled turbidite deposition in the Central Pandora Trough (Gulf of Papua) since Last Glacial Maximum: Linking sediment nature and accumulation to sea- level fluctuations at millennial time scale, Journal of Geophysical Research - Earth Surface.

Katz, B. (1983), Limitations of ‘Rock-Eval’ pyrolysis for typing organic matter, Organic Geochemistry, 4, 195-199.

Kawahata,H. (1999), Fluctuations in the ocean environment within the western Pacific warm pool during late Pleistocene, Paleoceanography, 14, 639-652.

Keen, T.R., D.S. Ko, R.L. Slingerland, S. Reidlinger, and P. Flynn (2006), Potential transport pathways of terrestrial material in the Gulf of Papua, Geophysical Research Letters, 33, 4, L04608, doi:10.1029/2005GL025416.

Kershaw, A.P. (1978), Record of the last interglacial-glacial cycle from northeastern Queensland, Nature, 212, 156-161.

Kershaw, A.P. (1988), Australasia, in Vegetation History, edited by B. Huntley and T. Webb, pp. 237-306, Kluwer Academic Publishers, Dordrecht, The Netherlands.

Kilbourne, K.H. and T. M. Quinn (2004), A fossil coral perspective on western tropical Pacific climate ~350 ka, Paleoceanography, 19, PA1019, doi:10.1029/2003PA000944.

Lallier-Vergès, E., P. Bertrand, and A. Desprairies (1993), Organic matter composition and sulfate reduction intensity in Oman Margin sediments, Marine Geology, 112, 57-69.

Lambeck, K. and J. Chappell (2001), Sea level change through the last glacial cycle, Science, 292, 679-686.

Langford, F.F. and M.M. Blanc-Valleron (1990), Interpreting Rock-Eval Pyrolysis data using graphs of pyrolizable hydrocarbons vs. total organic carbon, AAPG Bulletin, 74, 799-804.

Larcombe, P., and Carter, R.M. (2004), Cyclone Pumping, Sediment Partitioning and the Development of the Great Barrier Reef Shelf System: A Review, Quaternary Science Reviews, 23, 107-135.

Lea, D.W., D.K. Pak, and H.J. Spero (2000), Climate impact of Late Quaternary Equatorial Pacific sea surface temperature variations, Science, 289, 1719-1724.

148

Lewan, M.D. (1986), Stable carbon isotopes of amorphous kerogens from Phanerozoic sedimentary rocks, Geochimica et Cosmochimica Acta, 50, 1583-1591.

Licari, L. and A. Mackensen (2005), Benthic foraminifera off West Africa (1°N to 32°): Do live assemblages from the topmost sediment reliably record environmental variability?, Marine Micropaleontology, 55, 205-233.

Linke, P. and G.F. Lutze (1993), Microhabitat preferences of benthic foraminifera- a static concept or a dynamic adaptation to optimize food acquisition? Marine Micropaleontology, 20, 215-234.

Locarnini, R.A., A.V. Mishonov, J.I. Antonov, T.P. Boyer, and H.E. Garcia, (2005), World Ocean Atlas 2005, http://odv.awi.de/data/ocean/world-ocean-atlas-2005.html

Loubere, P. (1989), Bioturbation and sedimentation rate control of benthic microfossil taxon abundances in surface sediments: A theoretical approach to the analysis of species microhabitats, Marine Micropaleontology, 14, 317-325.

Loubere, P., A. Gary, and M. Lagoe (1993), Generation of the benthic foraminiferal assemblage: Theory and preliminary data, Marine Micropaleontology, 20, 165-181.

Loubere, P. (1996), The surface ocean productivity and bottom water oxygen signals in deep water benthic foraminiferal assemblages, Marine Micropaleontology, 28, 247-261.

Loubere, P. and M. Farridudin (1999), Benthic foraminifera and the flux of organic carbon to the seabed, in Modern Foraminifera, edited by B.K. Sen Gupta, pp. 181-199, Kluwer Academic Publishers, Dordrecht, The Netherlands.

Mackenzie, D.E. (1976), Nature and origin of Late Cainozoic volcanoes in western Papua New Guinea, in Volcanism in Australasia, edited by R.W. Johnson, pp. 221-238, Elsevier Scientific Publishing Company, Amsterdam, The Netherlands.

Maslin, M. and N. Mikkelsen (1997), Amazon fan mass transport deposits and underlying interglacial deposits: age estimates and fan dynamics, in Proceedings ODP, Scientific Results 155, edited by R.D. Flood, D.J.W. Piper, A. Klaus, and L.C. Peterson, pp. 353- 365.

Mead, G.A. and L. Tauxe (1986), Oligocene paleoceanography of the South Atlantic: Paleoclimatic implications of sediment accumulation rates and magnetic susceptibility measurements, Paleoceanography, 1(3), 273-284.

Meybeck, M. (1988), How to establish and use world budgets of riverine materials, in Physical and Chemical Weathering in Geochemical Cycles, edited by A. Lerman and M. Meybeck, The Hague, Kluwer Academic, pp. 242-272.

149 Meyers, P.A. (1994), Preservation of elemental and isotopic source identification of sedimentary organic matter, Chemical Geology, 144, 289-302.

Meyers, P.A. (1997), Organic geochemical proxies of paleoceanographic, paleolimnologic, and paleoclimatic processes, Organic Geochemistry, 27(5/6), 213-250.

Milliman, J.D. and J.P.M. Syvitski (1992), Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers, Journal of Geology, 100, 525-544.

Milliman, J.D. (1995), Sediment discharge to the ocean from small mountainous rivers: the New Guinea example, Geo-Marine Letters, 15, 127-133.

Milliman, J.D. , K.L. Farnswerth, and C.S. Albertin (1999), Flux and fate of fluvial sediments leaving large islands in the East Indies, Journal of Sea Research, 41, 97-107.

Milliman, J. D., N. W. Driscoll, R. Slingerland, J. Babcock, and J. P. Walsh (2004), Isotopic Stage 3 Deposition and Stage 2 Erosion of a Clinoform in the Gulf of Papua: Regional Tectonics Versus Eustatic Sea-Level Change, Eos Trans. AGU, 85(47), Fall Meet. Suppl., OS44A-05.

Milliman, J.D. and R.H. Meade (1983), World-wide delivery of river sediment to the oceans, The journal of Geology, 91(1), 1-21.

Müller, A. and B.N. Opdyke, (2000), Glacial-interglacial changes in nutrient utilization and paleoproductivity in the Indonesian Throughflow sensitive Timor Trough, eastern Indian Ocean, Paleoceanography, 15, 85-94.

Müller, P.J. and E. Seuss (1979), Productivity, sedimentation rate, and sedimentary organic matter in the oceans – I. Organic carbon preservation, Deep-Sea Research, 26A, 1347-1362.

Muhammad, Z., S.J. Bentley, L.A. Febo, A.W. Droxler, G.R. Dickens, L.C. Peterson, and B.N. Opdyke In Press, Excess 210Pb inventories and fluxes along the continental slope and basins of the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Nagata, T. (1961), Rock Magnetism, Tokyo, Maruzen, 350pp.

Nees, S. and U. Struck (1999), Benthic foraminiferal response to major paleoceanographic changes: A view from the deep-sea restaurant menu, in Reconstructing Ocean History: A Window into the Future, edited by F. Abrantes, and A. Mix, pp. 195-216, Plenum Publishers, New York.

Niel, D.T., A.R. Orpin, P.V. Ridd, and B. Yu (2002), Sediment yield and impacts from river catchments to the Great Barrier lagoon, Marine and Freshwater Research, 53, 733- 752.

150

Nittrouer, C.A., G.J. Brunskill, and A.G. Figueiredo (1995), Importance of tropical coastal environments, Geo-Marine Letters, 15, 121-126.

Nittrouer, C.A. and L.D. Wright (1994), Transport of particles across continental shelves Review of Geophysics, 32(1), 85-113.

Oboh-Ikuenobe, F.E. and S.E. de Villiers (2003), Dispersed organic matter in samples from the western continental shelf of Southern Africa: palynofacies assemblages and depositional environments of Late Cretaceous and younger sediments, Palaeogeography, Palaeoclimatology, Palaeoecology, 201, 67-88.

O’Connor, B (1999), Radiolaria from the Late Eocene Oamaru Diatomite, South Island, New Zealand, Micropaleontology, 45(1), 1-55.

Omura, K. and K. Hoyanagi (2004), Relationships between composition of organic matter, depositional environments, and sea-level changes in backarc basins, central Japan, Journal of Sedimentary Research, 74(5), 620-630.

Pain, C.F. and R.J. Blong (1976), Late Quaternary tephras around Mount Hagen and Mount Giluwe, Papua New Guinea, in Volcanism in Australasia, edited by R.W. Johnson, pp. 239-252, Elsevier Scientific Publishing Company, Amsterdam, The Netherlands.

Palinkas, C.M., C.A. Nittrouer, and J.P. Walsh (2006), Inner-shelf sedimentation in the Gulf of Papua, New Guinea: A mud-rich, shallow shelf setting, Journal of Coastal Research, 22(4), 760-772.

Patterson, L.J., S.J. Bentley, D.J. Henry, A.W. Droxler, G.R. Dickens, B.N. Opdyke and L.C. Peterson (2006), Multiple-source areas for turbidite sands reaching the deep basin in the Gulf of Papua, AAPG Annual Meeting.

Pérez, M.E., C.D. Charles, and W.H. Berger (2001), Late Quaternary productivity fluctuations off Angola: Evidence from benthic foraminifers, site 1079, in Proceedings ODP, Scientific Results 175, edited by G. Wefer, W.H. Berger, and C. Richter, pp. 1-19.

Posamentier, H.W., and P.R. Vail (1988), Eustatic controls on clastic deposition II - sequence and systems tract models, in Sea-level Changes: an Integrated Approach, SEPM Special Publication, vol. 42, edited by C.K. Wilgus, B.S. Hastings, C.G.S.C. Kendall, H.W. Posamentier, C.A. Ross, and J.C. Van Wagoner, SEPM, Tulsa.

Potter, D.K., P.W.M. Corbett, S.A. Barclay, and R.S. Haszeldine (2004), Quantification of illite content in sedimentary rocks using magnetic susceptibility – A rapid complement or alternative to X-ray diffraction, Journal of Sedimentary Research, 74(5), 730-735.

Rathburn, A.E., B.H. Corliss, K.D. Tappa, and K.C. Lohmann (1996), Comparisons of

151 the ecology and stable isotopic compositions of living (stained) benthic foraminifera from the Sulu and South China seas, Deep Sea Research Pt. 1, 43 (10), 1617-1646.

Rickwood, F.K. (1968), The geology of Western Papua, The APEA Journal, 8 (2), 51-61.

Robertson, A.I. and D.M. Alongi (1995), Role of riverine mangrove forests in organic carbon export to the tropical coastal ocean: A preliminary mass balance for the Fly Delta (Papua New Guinea), Geo-Marine Letters, 15, 134-139.

Rühlemann, C., P.J. Müller, and R.R. Schneider (1999), Organic carbon and carbonate as paleoproductivity proxies: Examples from high and low productivity areas of the Tropical Atlantic, in Use of Proxies in Paleoceanography: Examples from the South Atlantic, edited by G. Fischer and G. Wefer, pp. 315-344, Springer-Verlag, Berlin, Germany.

Rullkötter, J. (2000), Organic matter: The driving force for early diagenesis, in Marine Geochemistry, edited by H. D. Schultz and M. Zabel, pp. 129-172, Springer-Verlag, Berlin, Germany.

Ruxton, B.P. (1966), Correlation and stratigraphy of dacitic ash-fall layers in northeastern Papua, Journal of the Geological Society of Australia, 13, 41-67.

Sachs, S. D., and B. B. Ellwood (1988), Controls on magnetic grain-size variations and concentration in the Argentine Basin, South Atlantic Ocean, Deep-Sea Research, 35, 929–942.

Schlesinger, W.H. and J.M. Melack (1981), Transport of organic carbon in the world’s rivers, Tellus, 33, 172-187.

Schlitzer, R. (2007), Ocean Data View, http://odv.awi.de.

Schlünz, B. and R.R. Schneider (2000), Transport of terrestrial organic carbon to the oceans by rivers: Re-estimating flux- and burial rates, International Journal of Earth Sciences, 88, 599-606.

Scoffin, T.P., and A.W. Tudhope (1985), Sedimentary environments of the Central Region of the Great Barrier Reef of Australia, Coral Reefs, 4, 81-93.

Sebag, D., C. Di Giovanni, S. Ogier, V. Mesnage, F. Laggoun-Défarge, A. Durand (2006), Inventory of sedimentary organic matter in modern wetland (Marais Vernier, Normandy, France) as source-indicative tools to study Holocene alluvial deposits (Lower Seine Valley, France), International Journal of Coal Geology 67, 1-16.

Sen Gupta, B.K., I.C. Shin, and S.T. Wendler (1987), Relevance of specimen size in distribution studies of deep-sea benthic foraminifera, Palaios, 2, 332-338.

152 Shackleton, N.J., S.J. Crowhurst, G.P. Weedon, and J. Laskar (1999), Astronomical calibration of Oligocene-Miocene time, Philosophical Transactions of the Royal Society of London A, 357, 1907-1929.

Sjoerdsma, P.G. and G.J. van der Zwaan (1992), Simulating the effect of changing organic flux and oxygen content on the distribution of benthic foraminifers, Marine Micropaleontology, 19(1/2), 163-180.

Slingerland, R., N.W. Driscoll, J.D. Milliman, S.R. Miller, and E.A. Johnstone (2007), Anatomy and growth of a Holocene clinothem in the Gulf of Papua, Journal of Geophysical Research - Earth Surface.

Smart, C.W. and A.J. Gooday (1997), Recent benthic foraminifera in the abyssal northeast Atlantic Ocean: Relation to phytodetrital inputs, Journal of Foraminiferal Research, 27 (2), 85-92.

Smith, I.E.M. (1982), Volcanic evolution in Eastern Papua, Tectonophysics, 87, 315-333.

Stein, R. (1990), Organic carbon content/sedimentation rate relationship and its paleoenvironmental significance for marine sediments, Geo-Marine Letters, 10, 37-44.

Steinshouer, D.W., J. Qiang, P.J. McCabe, and R.T. Ryder (1999), Maps showing geology, oil and gas fields, and geologic provinces of the Asia Pacific Region. U.S.G.S. Open File Report 97-470F.

Stuiver, M., P.J. Reimer, E. Bard, J.W. Beck, G.S. Burr, K.A. Hughen, B. Kromer, G. McCormac, J. van der Plicht, and M. Spurk (1998), INTCAL98 Radiocarbon Age Calibration, 24000-0 cal BP, Radiocarbon, 40(3), 1041-1083.

Swartzendruber L.J. (1992), Properties, units and constants in magnetism, Journal of Magnetism and Magnetic Materials, 100, 573–575.

Syvitski, J.P.M., C.J. Vorosmarty, A.J. Kettner, and P. Green (2005), Impact of humans on the flux of terrestrial sediment to the global coastal ocean, Science, 308, 376-380.

Teichmüller, M. (1986), Organic petrology of source rocks, history and state of the art, Organic Geochemistry, 10, 581-590.

Thom, B.G. and L.D. Wright (1983), Geomorphology of the Purari Delta, in The Purari: Tropical Environment of a High Rainfall River Basin, edited by T. Petr, pp. 47-65, Dr. W. Junk Publishers, The Hague.

Thompson, L.G., E. Mosley Thompson, M.E. Davis, P.N. Lin, K.A. Henderson, J. Coledai, J.F. Bolzan, and K.B. Liu (1995), Late-glacial stage and Holocene tropical ice core records from Huascaron, Peru, Science, 269, 46-50.

153 Thompson, R. and F. Oldfield (1986), Environmental Magnetism, Allen and Unwin, London, England.

Tissot, B.P. and D.H. Welte (1984), Petroleum Formation and Occurrence, 699 pp., Springer-Verlag, Berlin, Germany.

Tite, M.S. and R.E. Linington (1975), Effect of climate on the magnetic susceptibility of soils, Nature, 256, 565-566.

Tuweni, A.O. and R.V. Tyson (1994), Organic facies variations in the Westbury formation (Rhaetic, Bristol Channel, S.W. England), Organic Geochemistry, 200: 1001- 1014.

Tyson, R.V. (1989), Late Jurassic palynofacies trends, Piper and Kimmeridge Clay Formations, UK onshore and offshore, in Northwest European Micropaleontology and Palynology, edited by D.J. Batten and M.C. Keen, British Micropalaeontological Society Series, pp. 135-172.

Tyson, R.V. (1993), Palynofacies analysis, in Applied Micropalaeontology, edited by D.G. Jenkins, Kluwer Academic Publishers, pp. 153-183.

Tyson, R.V. (1995), Sedimentary Organic Matter: Organic Facies and Palynofacies, 615 pp., Chapman and Hall, New York.

Van der Zwaan, G.J., I.A.P. Duijnstee, M. den Dulk, S.R. Ernst, N.T. Jannink, and T.J. Kouwenhoven (1999), Benthic foraminifers: Proxies or problems? A review of paleoecological concepts, Earth Science Reviews, 46, 213-236.

Van Gizel, P. (1979), Manual of the techniques and some geological applications of fluorescence microscopy, American Association of Stratigraphic Palynologists Workshop, Dallas, Oct. 29-Nov. 2, 55p.

Van Waveren, I. (1992), Morphology of probable planktonic crustacean eggs from the Holocene of the Banda Sea (Indonesia), in Neogene and Quaternary Dinoflagellate Cysts and Acritarchs, edited by M.J. Head and J.H. Wrenn, American Association of Stratigraphic Palynologists Foundation, pp. 89-120.

Wagner, T. (2000), Control of organic carbon accumulation in the late Quaternary equatorial Atlantic (Ocean Drilling Program sites 664 and 663): Productivity versus terrigenous supply, Paleoceanography, 15(2), 181-199.

Wagner, T., M. Zabel, L. Dupont, J. Holtvoeth, and C.J. Schubert (2004), Terrigenous signals in sediments of the low latitude Atlantic - Indications to environmental variations during the Late Quaternary: Part 1: Organic Carbon, in The South Atlantic in the Late Quaternary: Reconstructions of Material Budgets and Current Systems, edited by G. Wefer, S. Mulitza, and V. Ratmeyer, pp. 295-322, Springer-Verlag, Berlin, Germany.

154

Wahyudi and M. Minagawa (1997), Response of benthic foraminifera to organic carbon accumulation rates in the Okinawa Trough, Journal of Oceanography, 53, 411-420.

Walsh, J.P. and C.A. Nittrouer (2003), Contrasting styles of off-shelf sediment accumulation in New Guinea, Marine Geology, 196, 105-125.

Walsh, J.P., C.A. Nittrouer, C.M. Palinkas, A.S. Ogston, R.W. Sternberg, and G.J. Brunskill (2004), Clinoform mechanics in the Gulf of Papua, New Guinea, Continental Shelf Research, 24, 2487-2510.

Wang, B., S.C. Clemens, and P. Liu (2003), Contrasting the Indian and East Asian monsoons: implications on geologic timescales, Marine Geology, 201, 5-21.

Warner, N.R., and E.W. Domack, ( 2002), Millennial- to decadal-scale paleoenvironmental change during the Holocene in the Palmer Deep, Antarctica, as recorded by particle size analysis, Paleoceanography, 17, Art. No. 8004, doi:10.1029/2000PA000602.

Weber, M.E., F. Niessen, G. Kuhn, and M. Wiedicke (1997), Calibration and application of marine sedimentary physical properties using a multi-sensor core logger, Marine Geology, 136, 151-172.

Webster, P.J. and N.A. Streten (1978), Late Quaternary ice age climates of tropical Australasia: Interpretations and reconstructions, Quaternary Research, 10, 279-309.

Weedon, G.P., H.C. Jenkyns, A.L. Coe, and S.P. Hesselbo (1999), Astronomical calibration of the Jurassic time-scale from cyclostratigraphy in British mudrock formations, Philosophical Transactions of the Royal Society of London, A, 357, 1787- 1813.

White, L.D., R.E. Garrison, and J.A. Barron (1992), Miocene intensification of upwelling along the California margin as recorded in siliceous facies of the Monterey Formation and offshore DSDP sites, in Upwelling Systems: Evolution Since the Early Miocene, edited by C.P. Summerhayes, W.L. Prell, and K.C. Emeis, Geological Society Special Publication, 64, pp. 429-442.

Wolanski E., G. L. Pickard, and D. L. B. Jupp (1984), River plumes, coral reefs and mixing in the Gulf of Papua and the northern Great Barrier Reef, Estuarine, Coastal and Shelf Science, 18, 291-314.

Wolanski, E. and D.M. Alongi (1995), A hypothesis for the formation of a mud bank in the Gulf of Papua, Geo-Marine Letters, 15, 166-171.

Wolanski, E., A. Norro, and B. King (1995), Water circulation in the Gulf of Papua, Continental Shelf Research, 15(2-3), 185-212.

155 VITA

Lawrence Febo was born in Live Oak, Florida, in 1977, and grew up in

Washington Court House, Ohio. Lawrence graduated from The Ohio State University in

1999 with a Bachelor of Science degree in geological sciences and then with a master’s degree in 2003, also from Ohio State. His master’s thesis was a benthic and planktonic foraminiferal paleoceanographic reconstruction of the Chukchi Borderland, Arctic Ocean, over the past 50,000 years.

Lawrence has an unusual passion for rocks, fossils, and the sea. In 2003, he was awarded a biostratigraphy internship at BP America, which led him to a full-time offer, which he accepted. During Lawrence’s research at Louisiana State University, he traveled to the Gulf of Papua in 2004 and again in 2005 on a second, more extensive cruise beginning in Bitung, northern Sulawesi, and ending in Darwin, Australia.

Lawrence will graduate with a Doctor of Philosophy degree in geology, in

December 2007.

156