fluvial-lacustrine processes shaping the landforms …

162
University of Kentucky University of Kentucky UKnowledge UKnowledge Theses and Dissertations--Earth and Environmental Sciences Earth and Environmental Sciences 2017 FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS OF THE DISTAL PARAGUAY FLUVIAL MEGAFAN OF THE DISTAL PARAGUAY FLUVIAL MEGAFAN Edward Limin Lo University of Kentucky, [email protected] Author ORCID Identifier: https://orcid.org/0000-0003-3299-4495 Digital Object Identifier: https://doi.org/10.13023/ETD.2017.415 Right click to open a feedback form in a new tab to let us know how this document benefits you. Right click to open a feedback form in a new tab to let us know how this document benefits you. Recommended Citation Recommended Citation Lo, Edward Limin, "FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS OF THE DISTAL PARAGUAY FLUVIAL MEGAFAN" (2017). Theses and Dissertations--Earth and Environmental Sciences. 54. https://uknowledge.uky.edu/ees_etds/54 This Master's Thesis is brought to you for free and open access by the Earth and Environmental Sciences at UKnowledge. It has been accepted for inclusion in Theses and Dissertations--Earth and Environmental Sciences by an authorized administrator of UKnowledge. For more information, please contact [email protected].

Upload: others

Post on 02-Jun-2022

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

University of Kentucky University of Kentucky

UKnowledge UKnowledge

Theses and Dissertations--Earth and Environmental Sciences Earth and Environmental Sciences

2017

FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS

OF THE DISTAL PARAGUAY FLUVIAL MEGAFAN OF THE DISTAL PARAGUAY FLUVIAL MEGAFAN

Edward Limin Lo University of Kentucky, [email protected] Author ORCID Identifier:

https://orcid.org/0000-0003-3299-4495 Digital Object Identifier: https://doi.org/10.13023/ETD.2017.415

Right click to open a feedback form in a new tab to let us know how this document benefits you. Right click to open a feedback form in a new tab to let us know how this document benefits you.

Recommended Citation Recommended Citation Lo, Edward Limin, "FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS OF THE DISTAL PARAGUAY FLUVIAL MEGAFAN" (2017). Theses and Dissertations--Earth and Environmental Sciences. 54. https://uknowledge.uky.edu/ees_etds/54

This Master's Thesis is brought to you for free and open access by the Earth and Environmental Sciences at UKnowledge. It has been accepted for inclusion in Theses and Dissertations--Earth and Environmental Sciences by an authorized administrator of UKnowledge. For more information, please contact [email protected].

Page 2: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

STUDENT AGREEMENT: STUDENT AGREEMENT:

I represent that my thesis or dissertation and abstract are my original work. Proper attribution

has been given to all outside sources. I understand that I am solely responsible for obtaining

any needed copyright permissions. I have obtained needed written permission statement(s)

from the owner(s) of each third-party copyrighted matter to be included in my work, allowing

electronic distribution (if such use is not permitted by the fair use doctrine) which will be

submitted to UKnowledge as Additional File.

I hereby grant to The University of Kentucky and its agents the irrevocable, non-exclusive, and

royalty-free license to archive and make accessible my work in whole or in part in all forms of

media, now or hereafter known. I agree that the document mentioned above may be made

available immediately for worldwide access unless an embargo applies.

I retain all other ownership rights to the copyright of my work. I also retain the right to use in

future works (such as articles or books) all or part of my work. I understand that I am free to

register the copyright to my work.

REVIEW, APPROVAL AND ACCEPTANCE REVIEW, APPROVAL AND ACCEPTANCE

The document mentioned above has been reviewed and accepted by the student’s advisor, on

behalf of the advisory committee, and by the Director of Graduate Studies (DGS), on behalf of

the program; we verify that this is the final, approved version of the student’s thesis including all

changes required by the advisory committee. The undersigned agree to abide by the statements

above.

Edward Limin Lo, Student

Dr. Michael M. McGlue, Major Professor

Dr. Edward W. Woolery, Director of Graduate Studies

Page 3: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS OF THE

DISTAL PARAGUAY FLUVIAL MEGAFAN

Thesis

A thesis submitted in partial fulfillment of the requirements for the degree of Master of

Science in the College of Arts and Sciences at the University of Kentucky

By

Edward Limin Lo

Lexington, KY

Co-Directors

Dr. Michael M. McGlue, Assistant Professor, Earth and Environmental Sciences

Dr. Kevin M. Yeager, Associate Professor, Earth and Environmental Sciences

2017

Copyright © Edward Limin Lo 2017

Page 4: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

ABSTRACT OF THESIS

FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS OF THE

DISTAL PARAGUAY FLUVIAL MEGAFAN

Tropical wetlands such as the Pantanal help regulate global biogeochemical

cycles, but climate change is modifying these environments. Controls on environmental

changes can potentially be assessed from ancient, well-dated lacustrine sedimentary

records. An integrated field and laboratory approach was undertaken to study the

limnogeology of Lake Uberaba in the northern Pantanal, and test whether the lake has

preserved a reliable record of environmental change in its strata. This study was designed

to understand how the basin accumulates sediment and to assess its sensitivity to

hydroclimatic variability. The data showed that modern Lake Uberaba is a highly

dynamic, freshwater fluvial-lacustrine basin. Modern lake floor sediments are largely

siliciclastic silts, with limited organic matter content and abundant sponge spicules. This

sedimentary composition reflects the lake’s open hydrology and well-mixed water

column. Limited data from sediment cores indicates that Lake Uberaba may have formed

~1760 CE, following an abrupt transgression over an oxidized floodplain depositional

environment. The stratal contact between lacustrine and floodplain deposits suggests the

presence of an erosional unconformity, the timing and duration of which remains

unknown. The environmental change favoring lake formation appears to be linked to

increased effective precipitation provided by the Intertropical Convergence Zone (ITCZ)

in the northern Pantanal.

KEYWORDS: Distributive Fluvial Systems, Lake Uberaba, Limnogeology, Pantanal

Wetlands, Stratigraphy

Edward Limin Lo

October 10, 2017

Page 5: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS OF THE

DISTAL PARAGUAY FLUVIAL MEGAFAN

By

Edward Limin Lo

Michael M. McGlue

Thesis Co-Director

Kevin M. Yeager

Thesis Co-Director

Edward W. Woolery

Director of Graduate Studies

October 10, 2017

Date

Page 6: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

To everyone

Working to understand and manage tropical wetlands,

Improving the livelihoods of local and global communities

Page 7: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

iii

ACKNOWLEDGMENTS

The completion of this thesis was accomplished with the steadfast support of my

thesis co-directors Dr. Michael McGlue and Dr. Kevin Yeager. Dr. McGlue consistently

returned comments on my writing in a timely manner and supported my endeavors on

many levels. Dr. Yeager provided the resources for me to conduct grain size analyses and

learn to prepare samples for radionuclide measurements. No words can describe my

appreciation for their mentoring and tolerance of my imbroglios. My full committee

includes physical geographers Drs. Jonathan Phillips and Aguinaldo Silva. I thank Dr.

Phillips for sharing his assessment of the biogeomorphology of the study site. Special

thanks for Dr. Silva’s mentorship as my host professor during my Fulbright fellowship in

Brazil and for securing authorizations for fieldwork. Fieldwork was made possible by

Jocemir “Jaburu” Antunes, André Siqueira (ECOA), and the 7th Brazilian Army Border

Battalion—Porto Índio (Corumbá).

Lab work was accomplished at the Kentucky Geological Survey and with Dr.

Ivan Bergier at Embrapa-Pantanal. I benefited from the help of lab colleagues, namely

Meredith Swallom, Joseph Lucas, Bailee Hodelka, Wei Ji, Jeremy Eddy, Kim Schindler,

and Sean Swick. Hudson Macedo, Eder Merino, Leandro Luz, Giliane Rasbold, Luciana

Pereira and Edson Rodrigo provided me good company while I worked in Brazil. I am

grateful for their help with my project through processing samples, exchanging ideas, and

allowing me to vent my frustrations. Finally, I extend my gratitude to my family and

friends, namely Patrick Whalen, Wisam Muttashar, Paúl Rodriguez, and Scott Burke for

listening to my ramblings and musings on grad school life.

This project was funded by a National Science Foundation grant (#1541247) and

National Geographic Society grant to Drs. McGlue and Yeager. A Fulbright fellowship,

GSA Student Research Award, and LacCore Travel Grant to myself were supplemental

support. Thanks to the UK McNair fellowship, and the Department of Earth and

Environmental Sciences (Brown-McFarlan travel award, the Pirtle summer support, and

teaching and research assistantships). Brazilian funding and support originated from

Embrapa-Pantanal, the Federal University of Mato Grosso do Sul—Pantanal, UNESP-

Rio Claro on FAPESP project 2014/06889-2, FUNDECT (project 083/2016), and CNPq

Page 8: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

iv

(processes 447402/2014-5 and 448923/2014-9). A research grant (PQ2) awarded to Dr.

Silva (312.386/2014-1) and Pioneer professorship endowed funds are also appreciated.

Page 9: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

v

TABLE OF CONTENTS

ACKNOWLEDGMENTS ................................................................................................. iii

LIST OF TABLES ............................................................................................................ vii

LIST OF FIGURES ......................................................................................................... viii

CHAPTER ONE: INTRODUCTION ................................................................................. 1

1.1 Background ............................................................................................................... 1

1.2 Regional Setting ........................................................................................................ 6

1.3 Study Site .................................................................................................................. 9

CHAPTER TWO: METHODS ......................................................................................... 11

2.1 Satellite image analysis ........................................................................................... 11

2.2 Fieldwork ................................................................................................................ 13

2.2.1 Bathymetry and Stratabox™ imaging .............................................................. 13

2.2.2 Water chemistry and suspended sediments ...................................................... 13

2.2.3 Sediments and depositional environments ....................................................... 14

2.3 Limnological Data Analysis .................................................................................... 15

2.3.1 Bathymetry and Water Column ........................................................................ 15

2.4 Lake Floor Sediment and Core Analysis................................................................. 18

2.4.1 Bioavailable elements ....................................................................................... 18

2.4.2 Core Stratigraphy and Multi-Sensor Scanning ................................................. 18

2.4.3 Grain Size ......................................................................................................... 19

2.4.4 Indicators of Biological Paleoproductivity ....................................................... 20

2.4.5 Geochronology ................................................................................................. 21

CHAPTER THREE: RESULTS ....................................................................................... 26

3.1 Morphometrics and Limnology ............................................................................... 26

3.2 Lake floor sediments ............................................................................................... 33

3.3 Seismic Profile ........................................................................................................ 38

3.4 Lithostratigraphy ..................................................................................................... 44

3.5 Geochronology ........................................................................................................ 46

CHAPTER FOUR: DISCUSSION ................................................................................... 56

4.1 Modern Limnogeology of Lake Uberaba ................................................................ 56

4.2 Core Stratigraphy .................................................................................................... 70

4.3 Distributive Fluvial Systems (DFS) and their distal environments ........................ 75

CHAPTER FIVE: CONCLUSIONS ................................................................................ 79

APPENDICES .................................................................................................................. 82

Appendix A: Water Column Measurements ................................................................. 82

Appendix B: Lake Floor Sediment Analyses ................................................................ 85

Appendix C: Sample Statistical Analysis of Surface Sediment .................................... 97

Page 10: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

vi

Appendix D: Sediment Core Data ................................................................................. 98

REFERENCES ............................................................................................................... 126

VITA............................................................................................................................... 150

Page 11: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

vii

LIST OF TABLES

Table 2.1, Core Locations and Settings ............................................................................ 15

Table 2.2, Sediment Pretreatment for Optically Stimulated Luminescence (OSL) .......... 23

Table 2.3, Single Aliquot Regenerative-OSL Protocol to Estimate Equivalent Dose ...... 23

Table 3.1, Quantitative Results of Land Surface Thematic Classification ....................... 30

Table 3.2, Results of Ion Chromatography ....................................................................... 31

Table 3.3, Results of Water Chemistry ............................................................................. 32

Table 3.4, Post-Modern Radiocarbon Ages ...................................................................... 47

Table 3.5, Pre-Modern Radiocarbon Ages ....................................................................... 48

Table 3.6, Characteristics of the Optically Stimulated Luminescence Age Date ............. 49

Table 4.1, Summary of Distributive Fluvial System Termination Style .......................... 77

Page 12: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

viii

LIST OF FIGURES

Figure 1.1, Pantanal Basin Location Relative to the Intertropical Convergence Zone ..... 2

Figure 1.2, Pantanal Regional Overview .......................................................................... 3

Figure 1.3, Monthly precipitation and the Intertropical Convergence Zone ................... 8

Figure 1.4, Paraguay Megafan and Associated Lobes ...................................................... 10

Figure 2.1, Lake Uberaba Shoreline, Hydrology, and Survey and Sampling Maps ......... 12

Figure 2.2, Lake Uberaba Bathymetric Map .................................................................... 16

Figure 2.3, Pato Island Shoreline Substrates .................................................................... 17

Figure 3.1, LANDSAT 5 Classification Maps of Open Water and Macrophytes ............ 27

Figure 3.2, Comparison of Lake Uberaba, 1968 and 2015 ............................................... 28

Figure 3.3, Lake Uberaba Shore Vegetation ..................................................................... 29

Figure 3.4, Contour Maps of Grain Size, Total Organic Carbon, Biogenic Silica ........... 33

Figure 3.5, Total Suspended Sediment ............................................................................. 34

Figure 3.6, Contour Maps of X-ray Fluorescence Elemental Abundances ...................... 36

Figure 3.7, Contour Maps of Bioavailable Elements ........................................................ 37

Figure 3.8, Interpreted Subsurface Seismic Profile .......................................................... 39

Figure 3.9, Integrated Data for Core 3 .............................................................................. 40

Figure 3.10, Integrated Data for Core 4 ............................................................................ 41

Figure 3.11, Integrated Data for Core 2 ............................................................................ 42

Figure 3.12, Integrated Data for Core 1 ............................................................................ 43

Figure 3.13, Total 210Pbxs for Core 3 ................................................................................. 51

Figure 3.14, Total 210Pbxs for Core 2 ................................................................................. 52

Figure 3.15, Total 210Pbxs for Core 1 ................................................................................. 53

Figure 3.16, Sediment Mass and Linear Accumulation Rates of Core 3 .......................... 54

Figure 3.17, Sediment Mass and Linear Accumulation Rates of Core 2 .......................... 54

Figure 3.18, Sediment Mass and Linear Accumulation Rates of Core 1 .......................... 55

Figure 4.1, Depositional Environments near Lake Uberaba ............................................. 57

Figure 4.2, Contour Maps of Clay, Silt, and Sand ............................................................ 58

Figure 4.3, Classified Seismic Character .......................................................................... 60

Figure 4.4, Serra do Amolar Photo ................................................................................... 65

Figure 4.5, LANDSAT 8 False-Color (RGB-654) Image of Lake Uberaba .................... 69

Figure 4.6, Stratigraphic Correlations and Integration ..................................................... 74

Page 13: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

1

CHAPTER ONE: INTRODUCTION

1.1 Background

The sum of every type of wetland comprises just six percent of Earth’s land

surface, yet these environments are key sources of trace elements, carbon dioxide (CO2),

and methane (CH4), with the latter two strongly influencing atmospheric warming

(Barbier, 1994; Shindell et al., 2004; Vriens et al., 2014). Wetlands contribute

substantially to global climate and biogeochemical cycles, especially the carbon cycle. As

much as 30% of the global wetlands inventory is found in the tropics (Mitsch et al.,

2010). In contrast with their temperate counterparts, tropical wetlands can store 80%

more organic carbon (Mitsch et al., 2010). Compounded by the dearth of observational

and modeling data from these aquatic ecosystems (e.g., Sjögersten et al., 2014; Zhu et al.,

2017; Gumbricht et al., 2017), the response of tropical wetland systems to global change

remains unclear. This knowledge gap is concerning because tropical wetlands have been

shown to significantly influence atmospheric CO2 and CH4 (Cao et al., 1998; Shindell et

al., 2004; Fischer et al., 2008; Konijnendijk et al., 2011; Bousquet et al., 2011). One key

regulator of global atmospheric CH4 flux is believed to be the Pantanal of tropical South

America (e.g., Marani and Alvalá, 2007; Bastviken et al., 2010).

Occupying 1.38 x 105 km2, the Pantanal is the world’s largest freshwater tropical

wetland (Seidl and Moraes, 2000; McGlue et al., 2011; Ioris et al., 2014). The Pantanal is

a savanna-floodplain type wetland, located in the Upper Paraguay River basin of the Río

de la Plata watershed (Fig. 1.1). The Paraguay River is the primary river controlling the

Pantanal water supply (Fig. 1.2) (Tricart, 1982; Hamilton, 1999). The diverse wildlife of

the Pantanal has attracted many biological studies, and ecologists are working to

understand how environmental change may impact the fauna and flora there (e.g.,

Quigley and Crawshaw, 1992; Suárez et al., 2004; Harris et al., 2005; Junk et al., 2006).

Far fewer studies have examined the history of environmental change and associated

spatial complexities archived in recent geological materials within the basin.

Compounding this issue is that hydroclimate changes are expected to increase variability

of the tropical water cycle in the coming decades (Milly et al., 2005; Seager et al., 2012;

McGlue et al., 2015). To determine how the Pantanal will respond to a more variable

Page 14: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

2

water cycle in the near future, we can assess how it responded to past environmental

changes using sensitive sedimentary archives.

Figure 1.1: Outline of the Intertropical Convergence Zone (ITCZ) path during the

northern hemisphere summer (June-August; JJA) and the austral summer (December-

February; DJF). The southerly ITCZ passes through parts of the Pantanal (yellow) within

the greater Río de la Plata basin (dark grey). Modified from McGlue et al. (2012), based

on data from Liu and Battisti (2015) and Boers et al. (2016).

Page 15: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

3

Figure 1.2: Adapted from Assine et al. (2015a), this enlarged diagram of the upper

Paraguay River basin (yellow area in Fig. 1.1) features the diverse geomorphic and

hydrologic features of the Pantanal system. The rectangular inset box corresponds to the

location of Fig. 1.4.

Page 16: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

4

Some of what is presently known about the response of the Pantanal wetlands to

environmental change in the late Quaternary come from studies of floodplain and lake

sediments (e.g., Bezerra and Mozeto, 2008; McGlue et al., 2012; Kuerten et al., 2012;

Metcalfe et al., 2014; Assine et al., 2014; Fornace et al., 2016; Pupim et al., 2017).

However, most of these studies were focused on the central and southern regions of the

wetlands, thus presenting a knowledge gap for the north. Given the north-south gradient

in hydrological processes discussed by a number of authors (e.g., McGlue et al., 2011),

environmental changes may very well impact the Pantanal with spatial asymmetry. As a

result, paleoenvironmental information derived from geological archives in the northern

Pantanal is a pressing need to fully understand the region’s sensitivity to global change.

The Pantanal landscape and its numerous lakes are most commonly shaped by

fluvial hydrosedimentary processes (Assine and Soares, 2004). Fluvial erosion and

competitive aggradation of fluvial fans are two processes by which accommodation for

lakes may form (e.g., Wetzel, 2001; Ashworth and Lewin, 2012; Cohen et al., 2015).

Aggradation and variations with local topography can create (a) blocked-valley lakes (3-5

m depth; can become desiccated), (b) dish lakes (length:width > 5; calm currents during

floods), and (c) channel lakes (length:width < 5; active currents) (Hamilton and Lewis,

1990; Wetzel, 2001; Rowland et al., 2009). These examples are but a few of the immense

variety of floodplain lakes. In the Pantanal, regional processes include tectonic

subsidence, earthquakes, and faulting whereas the climate system drives precipitation,

evaporation, and vegetation dynamics. Altogether, these processes interact with the river

systems and influence lake morphometry and hydrology.

Fundamentally, lakes are depressions that can hold water and sediment on the

landscape (Kelts, 1988; Cohen, 2003). The floodplain lakes in the Pantanal are incredibly

diverse. For example, salt pans (salinas) are often recharged by groundwater, are highly

alkaline, 500-1000 m in diameter, 2-3 m deep, and enclosed by sand hills (Barbiero et al.,

2008; McGlue et al., 2017). Freshwater lakes (baías), particularly in the Nhecolândia

region (Fig. 1.2), are seasonal, no deeper than 2 m, maintain perennial floating

vegetation, and are recharged by surface water during floods (Furquim et al., 2010;

Almeida et al., 2011). Elsewhere in the central Pantanal, floods from major rivers sustain

Page 17: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

5

larger perennial freshwater lakes that are directly connected to those rivers (McGlue et

al., 2011; 2012).

Several factors may allow lake sediments to be good archives of environmental

history, once the hydroclimate filter has been considered. The hydroclimate filter consists

of a lake’s physical, chemical, and biological characteristics, which altogether modify

how the indicators of ancient environments are recorded in the sediment (Cohen, 2003).

First, lake sediment accumulates considerably faster (0.3-6.0 mm yr-1) than in many

marine environments (Appleby et al., 1998; McGlue et al., 2011). A higher sedimentation

rate helps discern indicators that are more sensitive to environmental change at decadal or

centennial scales. Second, floodplain lakes can persist at the scale of thousands of years,

thereby providing a potentially long record of change (Cohen, 2003; Schillereff et al.,

2014). Third, the strata of these varied floodplain lakes are well-suited to studying

environmental change and paleohydrology, due to preservation of microfossils (e.g.,

pollen, diatoms, sponge spicules), organic matter (OM), and grain size variability that in

certain instances reflects hydrodynamic energy and proximal fluvial activity (McGlue et

al., 2011; Whitney et al., 2014).

A number of large lakes in the Pantanal are spatially associated with distributive

fluvial systems (DFS) (Weissmann et al., 2010; Hartley et al., 2010). The Pantanal’s DFS

formed in conditions of seasonal discharge, as large rivers flow out of confined channels

in the Brazilian plateau highlands (planalto) to unconfined conditions in the lowlands,

where channels can migrate laterally (Assine and Soares, 2004; Assine, 2005; Leier et al.,

2005). These rivers are frequently affected by climate change, and the geomorphology of

the Pantanal reflects historic climate patterns (Bezerra and Mozeto, 2008). Historic

climate change has been recorded in the Nhecolândia region in the form of deflation

ponds, indicating drier conditions (McGlue et al., 2017). Drier conditions were also

identified as shifts in channel planform in the São Lourenço (Fig. 1.2) megafan (Pupim et

al., 2017).

How does the northern Pantanal respond to late Quaternary environmental

change? The overarching goal of this study is to add to a growing database of

environmental change studies of the Pantanal wetlands. We chose an unmapped lake

system found along the distal fringe of the Paraguay fluvial fan (a “megafan” following

Page 18: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

6

the terminology of Leier et al., 2005) in the northern Pantanal. Lake Uberaba (Fig. 1.2)

has never been comprehensively studied before, and is particularly valuable as a possible

archive of fluvial-lacustrine environmental history due to its association with the

Paraguay megafan. Comparatively few studies in the northern Pantanal (e.g., Horton and

DeCelles, 1997; Assine and Soares, 2004; Assine et al., 2015b) have been completed on

the region’s geological features. Our first step was to determine how Lake Uberaba

functions as a depositional system. Understanding how the lake functions is important to

probe its potential to record signals of Holocene environmental change. Because of its

connection to the Paraguay megafan, one of the most spectacular DFS in the Pantanal, the

sediments of Lake Uberaba provide an opportunity to reveal new details of the

sedimentary and hydrogeomorphic history of this fluvial landscape. Improved records of

geomorphological evolution of the DFS are valuable for understanding regional climate

sensitivities and also for advancing modern analog facies models for the rock record.

This study was designed to test one hypothesis examining whether lake sediment

in the north Pantanal can record environmental signals, modified by the lake’s

hydroclimate filter. This question is important to determine if lake sediments can be used

to reliably interpret Holocene environmental conditions in the Pantanal. Widely varying

interpretations of Holocene conditions in the Pantanal have been presented in existing

studies of the central and southern Pantanal (Whitney et al., 2011; McGlue et al., 2012;

Metcalfe et al., 2014; McGlue et al., 2015).

1.2 Regional Setting

The Pantanal lies in a tectonically active, late Cenozoic (~2.5 Ma) sedimentary

basin, as evidenced by basement faults, fault-controlled landforms, and occasional

earthquakes (Ussami et al., 1999; Assine and Soares, 2004; Assine et al., 2015a; Dias et

al., 2016). This basin is most likely a shallow back-bulge that formed by flexure during

the Andean orogeny (which began ~200 Ma), though some evidence of extensional

deformation exists as well (Ussami et al., 1999; Chase et al., 2009). The Pantanal basin is

bounded to the east by the Brazilian planalto, and to the west by the Upper Paraguay

River (Assine, 2005). Although topography is minimal, resistant inselbergs in the basin

rise up to ~900 m above mean sea level (amsl), and are composed of Neoproterozoic

Page 19: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

7

conglomerates and siltites (Lacerda Filho et al., 2004). Gravity anomaly data suggest that

the maximum stratal thickness in the basin center is on the order of ~500 m (Horton and

DeCelles, 1997). Most of the deposits in this basin are believed to have formed due to the

development of several large megafans in the late Pleistocene (Fig. 1.1) (Tricart, 1982;

Assine and Soares, 2004). Fluvial megafans are large DFS (> 103 km2) characterized by

high seasonal discharge (Leier et al., 2005; Weissmann et al., 2010; Latrubesse, 2015).

Megafan deposits of Pleistocene age almost entirely fill the available accommodation

space in this basin, producing a very low land surface gradient of ~0.015 m km-1 from

north to south (Assine and Silva, 2009). This low regional gradient allows wetlands to

form during seasonal flooding of the Upper Paraguay River, which is undammed in the

Pantanal and modulates local base level (McGlue et al., 2012).

The Pantanal floodplains would not form vast wetlands without the Upper

Paraguay River’s annual flood cycle (Silva and Girard, 2004). Seasonal flooding of the

Upper Paraguay River, which sustains valuable aquatic ecosystem services (Seidl and

Moraes, 2000; McGlue et al., 2011), is governed by the arrival of the South American

Summer Monsoon (SASM). During austral summer rains, the Upper Paraguay River

overflows its banks beginning in the northern Pantanal, and flood waters reach the

southern Pantanal over a period of several months (October to March). Flood waters

modulate the height of the water table, thereby changing the soil and vegetation

seasonally in many of the low-lying areas mapped by Franco and Pinheiro (1982).

Major climatic patterns, particularly monsoon-like precipitation, and migration of

the Intertropical Convergence Zone (ITCZ; Fig. 1.3), influence the Pantanal (Fig. 1.1).

The ITCZ results from low air pressure and is an equatorial convergence of the northern

and southern hemisphere trade winds. It is characterized by high sea surface temperatures

(SSTs), low atmospheric mixed-layer depths, heavy clouds, and precipitation (Garreaud

et al., 2009; Wade, 2014). Atlantic SSTs, which are affected by the El Niño Southern

Oscillation (ENSO), control the ITCZ’s latitudinal position (Rodríguez et al., 2011;

Viana et al., 2014). When the North Atlantic Ocean is relatively cool, the ITCZ occupies

a more southerly position, bringing more rainfall to the Pantanal (Fig. 1.1). The path of

the ITCZ has controlled the SASM for the last 2,000 years (Vuille et al., 2012), and the

SASM today provides water that sustains > 350 million people, floral and faunal

Page 20: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

8

biodiversity, and hydroelectric power in South America (Wade, 2014). The South

American Convergence Zone (SACZ) also affects rainfall distribution, depending on

SSTs in the North Atlantic Ocean and the magnitude of solar irradiance (Novello et al.,

2012; Bernal et al., 2016). The SACZ is an elongate convective band of concentrated

moisture (~400 mm mo-1) stretching from western Amazonia to southeastern Brazil

during the austral summer and is a major component of the SASM (Kodama, 1992;

Carvalho et al., 2004).

Figure 1.3: Monthly average precipitation (mm) in areas receiving monsoonal rains of

the ITCZ. When the ITCZ occupies a southern position in January, much of South

America experiences precipitation. When the ITCZ shifts north in July, only northern

South America receives substantial rain. Image adapted from Syvitski et al. (2014).

Page 21: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

9

1.3 Study Site

The Paraguay megafan is characterized by floodable regions with varying degrees

of soil moisture (Franco and Pinheiro, 1982). The formation of this megafan has been

linked to distributary processes (downstream channel bifurcation, narrowing, and

shoaling; Horton and DeCelles, 2001; Leier et al., 2005; Nichols and Fisher, 2007) south

of the Precambrian Serrana Province (Fig. 1.1), in the Quaternary fluvial plains described

by Assine and Silva (2009). According to those authors, the evolution of this megafan

resulted from lobe switching during periods of high rainfall. During the wet season

(January to April), the Paraguay River can attain discharge of up to 2000 m3 s-1, which

promotes sediment transport (estimates from < 1.1 to 22.5 million tons yr-1 deposited) to

the floodplain and lobes (Assine et al., 2015a). Three depositional lobes have been

identified on this megafan using satellite images, and numerous ephemeral streams

dispersed over the fan have been observed (Fig. 1.4A) (Assine and Silva, 2009). Most

ephemeral streams of the two relict lobes drain into the Corixo Grande River, which

flows along the toe of the megafan from west to east (Fig. 1.4).

Lake Uberaba is located on the distal, southeastern fringe of the Paraguay

megafan, on the border between Brazil and Bolivia (Fig. 1.2). The lake occupies a

position in a shallow depression between the toe of the megafan and the high relief

Precambrian plateaus found along the Bolivian border. The regional slope of the

Paraguay megafan is tilted gently towards the south, with elevations declining slightly

towards its subaqueous toe (Assine and Silva, 2009). Lake Uberaba receives inflows from

the Corixo Grande River to the west, ephemeral streams of the relict depositional lobes to

the north, and the Canzi River to the east (Fig. 1.4B) (Lo et al., 2017). The primary

outflow of the lake is to the south, through Canal Dom Pedro II into Lake Gaíva (Fig.

1.4). Secondary lake outflow can spill into the Paraguay River, through a bay-like

connection on the southeastern margin of the lake. The Paraguay River channel does not

directly connect to Lake Uberaba, but that river’s flooding can have a small effect on lake

levels.

Page 22: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

10

Figure 1.4: A) Lansat-8 satellite image from September 15, 2015 with adjusted

brightness color bands (RGB-654), emphasizing woody vegetation (bright green), aquatic

macrophytes (darker green), exposed savannah (pink), and water (blue-black). Labeled

lakes are a) Orion, b) Piranhas, c) Uberaba, and d) Gaíva. Image courtesy of the U.S.

Geological Survey (https://glovis.usgs.gov). B) Geomorphic map adapted from Assine

and Silva (2009), which interprets the various fluvial fan lobes (blue) and alluvial fans

(pink). This figure emphasizes that distinct vegetation composition corresponds with

distinct depositional environments. The lines in the Serrana Province are outlines of the

ridges and are not associated with elevation contours.

Page 23: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

11

CHAPTER TWO: METHODS

Three expeditions were organized to Lake Uberaba in 2015 and 2016, taking

advantage of relatively low water levels during the dry season to help facilitate sample

collection (Fig. 2.1). Prior to these surveying expeditions, virtually nothing was known

about the limnogeology of this basin. Initial fieldwork in September (2015) was focused

on measurements of water column chemical characteristics. Characterizing the water

column is needed to constrain hydroclimate variables such as pH and dissolved oxygen

(DO) that influence sediment composition and benthic life. Bathymetric mapping and

surface sediment sampling were important to gain an understanding of modern depth

distribution, arrangement of depositional environments, and patterns of sediment

accumulation. Sediment accumulation patterns can be determined based on interpolated

maps of grain size and sediment composition. Initial insights from this expedition

informed our strategy for collecting short (< 1 m) sediment cores (November 2015 and

June 2016) focused on areas comprised of fine-grained sediments on the lake floor.

2.1 Satellite image analysis

Satellite image analysis allows for rapid identification and comparison of various

landforms (e.g., Weissmann et al., 2010), as well as land surface change (e.g., Byrne et

al., 1980; Hansen and Loveland, 2012; Lo et al., 2017). In this study, lake morphology

was determined based on a Landsat-8 image (RGB-654) taken on September 15, 2015,

with 30 m resolution. A combination of adjusted cumulative cut stretch, local histogram

stretch, and brightness was applied for image enhancement in QGIS. This date was

selected to provide an image that most closely approximated the initial field season

(September 3-8, 2015). The georeferenced image was downloaded from the U.S.

Geological Survey’s (USGS) Global Visualization (GloVis) (https://glovis.usgs.gov)

database. The primary software for creating the shapefiles used in this study was QGIS

2.14.7, but lake surface area was computed in (INPE) SPRING 5.4.3. To compare with

the Landsat image analysis of Lo et al. (2017), reconnaissance satellite images were also

purchased from USGS GloVis. Images were assessed from June 30, 1968 taken by the

KH-4A satellite on mission number 1047-2, with 7-micron (3,600 dpi) resolution.

Page 24: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

12

Figure 2.1: The modern conditions of Lake Uberaba, and survey maps A) Landsat 8

image (September 15, 2015) illustrating the extent of the lake and its surrounding

landscape. False color image (RGB-654) shows water as blue, aquatic vegetation as dark

green, permanent vegetation as bright green, and exposed soil as pink. B) Hydrology of

the lake outlined with solid lines representing perennial flows, and dotted lines

representing ephemeral and seasonal channels. C) Map of water sample stations,

collected in September 2015 (blue) and November 2015 (orange). Yellow markers

indicate sites where the YSI multi-parameter sonde was deployed. D) Map of the

bathymetric survey route (green lines), SyQwest Stratabox™ transect (blue dotted line),

surface sediment sample stations (red), and core sites (black triangles).

Page 25: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

13

2.2 Fieldwork

2.2.1 Bathymetry and Stratabox™ imaging

Bathymetry relates information on available sediment accommodation space,

water column stratification, and water chemistry (Wetzel, 2001; Cohen, 2003), thus

making it a key variable for understanding the limnogeology of Lake Uberaba. We

surveyed the bathymetry of Lake Uberaba in September 2015 using a Furuno model GP-

1650WF 200 kHz echo sounder, sampling every 2 m at a boat speed of ~10 km hr-1.

Transect lines (~1.5 km spacing) were oriented parallel to each other trending NE-SW, to

maximize spatial coverage and allow for accurate contouring. A total of ~253 line-km of

depth measurements were recorded and integrated with sub-meter accuracy GPS

navigation data (Fig. 2.1D). Information derived from echo-sounding was complimented

by acoustic imaging of the shallow subsurface, which can provide a perspective on facies

architecture and paleo-depositional surfaces (e.g., Hatushika et al., 2007; Shen et al.,

2011; Zocatelli et al., 2012; Muttashar et al., 2012) (Fig. 2.1D). A 10 kHz SyQwest

Stratabox™ sub-bottom profiler was used to record a single NE-SW oriented seismic

profile, which provided several meters of penetration in the best case.

2.2.2 Water chemistry and suspended sediments

Recognizing seasonal changes in water chemistry helps identify factors affecting

modern lake floor sediments that may not have been accounted for in a single field visit.

Basic in situ water chemistry (temperature, pH, dissolved oxygen, salinity) was measured

with a YSI EXO1 multi-parameter sonde, both at the surface and just above the lake floor

in September 2015 and June 2016. Measurements were taken at the surface and at depth

to determine if differences existed vertically in the water column. Additionally, discrete

water samples were collected to evaluate alkalinity, suspended sediment concentrations,

and cations (e.g., Na+, K+, Mg2+, Ca2+) and anions (e.g., F-, Cl-, NO31-, SO4

2-). Dissolved

ions and alkalinity are important for interpreting brine chemistry, potential mineralogy,

and local biochemical cycling that affect sediment composition. Water chemistry

controls sedimentation indirectly by influencing ionic concentrations (mineral

precipitation) and organisms (OM and biogenic sediments) (Cohen, 2003). Flora and

fauna are also affected by suspended sediment, which contribute to sediment

Page 26: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

14

accumulation rates (e.g., Newcombe and MacDonald, 1991; Welch et al., 1996; Lo et al.,

2014). To characterize the water chemistry, seven sites were sampled using 300-ml

plastic bottles in September 2015. An additional 28 water samples were collected ~50 cm

below the water surface in 1-L plastic bottles exclusively for suspended sediment

analyses in November 2015 (Fig. 2.1C).

2.2.3 Sediments and depositional environments

Sixty-nine surface sediment samples (~0-2cm below the lake floor, b.l.f.) (Fig.

2.1D) were collected to provide a thorough characterization of modern sediments in Lake

Uberaba. Samples were collected in a 2 km x 2 km grid to produce an evenly distributed

data set spanning the entire lake. A Wildco™ Ponar grab sampler was deployed to collect

fine-grained sediments, whereas a heavier Van Veen grab sampler was utilized in denser,

sandy substrates. Sediments were carefully scraped with a spatula from the upper 1-2 cm

of the samplers, and transferred to 50 ml glass jars rinsed with lake water for storage

prior to transport and analysis.

Lake sediment cores can provide information about the history of changes to

climate (Wolfe et al., 2005; Polissar et al., 2013; Ghinassi et al., 2015; Lintern et al.,

2016), hydrology (Magee et al., 1995; Loverde-Oliveira et al., 2009), ecology (Larsen

and MacDonald, 1993), and depositional environments (Jacob et al., 2004; Kuerten et al.,

2012). Key parameters that are typically measured from lake sediment cores include: (a)

types of lithofacies; (b) bedding thickness and contacts; (c) microfossil assemblages; and

(d) major elemental abundances, which taken together provide important insights on

paleo-environmental conditions. During the November 2015 field season, four short

sediment cores were collected at Lake Uberaba. Three cores were obtained by hammer

coring using aluminum core barrels, and one was obtained using a Bolivia-style piston

coring system with a polycarbonate sleeve (Cores 1-4; Table 2.1) (Wright, 1967).

Additional cores were collected in 2016, but those data are not reported here. Core

locations are shown on Figure 2.1D.

Page 27: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

15

Table 2.1: Locations and characteristics of sediment cores recovered from Lake Uberaba.

Cores 1-4 were collected in November 2015, and cores 5-8 were collected in June 2016.

Water depth was not recorded for two cores (n/a). *Core has not been split, and was

measured based on core sleeve dimensions.

Name (sleeve) Latitude Longitude Water Depth

(m)

Core Length

(m)

Core 1 (Al) -17.491 -57.853 1.47 0.71

Core 2

(polycarb)

-17.436 -57.748 0.65 0.55

Core 3 (Al) -17.546 -57.759 1.87 0.98

Core 4 (Al) -17.546 -57.797 2.17 0.86

Core 5 (Al) -17.506 -57.722 2.74 *0.55

Core 6 (Al) -17.448 -57.793 n/a 0.73

Core 7 (Al) -17.452 -57.826 n/a *0.95

Core 8 (Al) -17.530 -57.826 3.21 *0.99

2.3 Limnological Data Analysis

2.3.1 Bathymetry and Water Column

Digital sounding data (latitude, longitude, depth) were processed in ArcMap along

with a vectorized shapefile of the modern Lake Uberaba shoreline based on the Landsat-8

image from September 15, 2015. The shoreline was set as a zero contour, in order to

generate the best possible interpolation. The TopoToRaster (kriging specific to

topography) function in ArcMap’s spatial analyst toolkit was applied to contour

bathymetry. The most probable source of error in the contouring statistics was the

distance between transects, which we attempted to obviate by importing the contours into

Adobe Illustrator CS6, and smoothing them manually. Bathymetric data are shown in

Figure 2.2, though this map best captures the distribution of isobaths during the dry

season. Indications of higher water levels (beach berms, trash lines) observed on the

lake’s islands (Fig. 2.3A, B) suggest that water levels can be 2-3 m higher during the wet

season.

Page 28: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

16

Figure 2.2: Interpolated bathymetric map illustrating shoaling near the north and

northwest lake margins, as well as at the southeast outlet. Shoaling is much less

pronounced towards the southwest and eastern margins. Red lines indicate areas where

bathymetric data were collected.

Page 29: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

17

Figure 2.3: The substrate on the western banks of Pato Island (see Fig. 3.2) is composed

of metasedimentary cobbles, forming beach berms traced in A and B. Berms are

numbered from closest (1) to farthest (4) from shore. The arrow points towards a person

for scale. Varying sizes of metasedimentary rocks are visible in C and D. The edge of a

boot for scale appears at the bottom of D.

Page 30: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

18

Using laboratory facilities at the Brazilian Agricultural Research Corporation

(EMBRAPA), 100 ml of each water sample was filtered through a 45 µm cellulose

membrane filter prior to analysis using a Dionex ICS-1100 ion chromatograph (Thermo

Scientific, 2012). This analysis measured concentrations of major cations (Li+, Na+,

NH4+, K+, Mg+2, Ca+2) and anions (F-, Cl-, Br-, NO2

-, NO3-, PO4

3-, SO42-) in the lake

water. Each water sample was also titrated from 4.0 pH to 3.2 pH to calculate alkalinity

(Gran, 1952). Glass fiber filters (45 µm) were used to collect suspended solids, and were

dried (105°C) and combusted (500°C) to determine the organic fraction (Guy, 1969).

This analysis yielded the percent of suspended sediment, as well as its OM content.

2.4 Lake Floor Sediment and Core Analysis

2.4.1 Bioavailable elements

Surface sediment samples were measured for bioavailable major (Ca, Mg, Al, Na,

K, P) and trace (Fe, Mn, Cu, Zn) elements at EMBRAPA. The bioavailable element

composition in surface samples was determined by atomic absorption spectrophotometry

(exchangeable Ca, Mg), flame photometer (exchangeable K, Na), and colorimetry

(available P), following the methodology presented in Silva (2009). The 10 ml samples

were oven dried at 45°C for four days. A 2.5 ml aliquot was extracted and measured for

major elements, and another 2.5 ml for minor elements. Major elements were extracted

by KCl, and minor elements by HCl and H2SO4.

2.4.2 Core Stratigraphy and Multi-Sensor Scanning

Initial core descriptions were conducted at the National Lacustrine Core Facility

(LacCore) at the University of Minnesota using standard procedures (Schnurrenberger et

al., 2003). Each core was scanned for physical properties using a GeoTek whole core

logger; measurements included P-wave velocity, magnetic susceptibility, and GRAPE

(gamma ray attenuation porosity evaluator) density. High resolution line-scan

photographs and color spectrophotometery were also completed using standard

techniques. Cores were evaluated both macroscopically and microscopically to determine

lithofacies characteristics. We produced smear slides every 2 cm in OM-rich sediment

and every 4 cm in OM-poor sediment along each core. This assessment provided

Page 31: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

19

information about the biogeochemical composition of the sediments. For cores 1 and 3,

each label corresponds to 1 cm above the label name (e.g., 3 cm = 2-3 cm depth). For

cores 2 and 4, each label corresponds to 0.5 cm of material above and below the labeled

depth (e.g., 3 cm = 2.5-3.5 cm depth).

2.4.3 Grain Size

Siliciclastic grain sizes in cores and surface samples provide information related

to the hydrology and environmental energy of depositional settings, and their change over

time. A series of chemical digestions was necessary to isolate the siliciclastic grains,

using 30% H2O2 to oxidize OM, NaOH (1 N) to eliminate biogenic silica, and HCl (1 N)

to remove carbonates (Konert and Vandenberghe, 1997; Mudie et al., 2006; Johnson and

McCave, 2008). Each digestion was completed in a 50 ml centrifuge tube for ~24 hours,

or until samples were no longer reacting. Samples were rinsed with DI water and

centrifuged until all suspended particulates dropped out of suspension. Magnesium

chloride (0.5M) was added as needed to help flocculate and settle the clay size fraction.

Three DI rinses were completed between each digestion. After the final digestion and

rinse, sodium hexametaphosphate (15%) was added to samples to disperse clay minerals,

and centrifuge tubes were placed on a wrist action shaker for ~30 min. We employed a

Malvern Mastersizer S2000 laser-diffraction particle size analyzer to determine volume

percentages of sand, silt, and clay (as defined by the Wentworth, 1922 scale), as well as

median (sometimes referred to as D50) detrital grain sizes. Each sample was thoroughly

mixed, and transferred by disposable pipette to the Malvern unit.

All sediment samples were subjected to nondestructive X-ray fluorescence (XRF)

to determine relative concentrations of major elements (mainly Si, Al, Fe, Ti, Ca, and K)

(Löwemark et al., 2011). Each sample was freeze dried and ground using an agate mortar

and pestle prior to analysis. Following calibration with the SARM 41 standard (Ring,

1993), we used a Bruker Tracer IV handheld ED-XRF to measure elemental

concentrations. This instrument ionizes the sample and records the absorption in counts

per second. Samples were measured for 90 seconds, and Bruker’s empirical conversion

was applied to quantify relative elemental abundances (Rowe et al., 2012). We focused

on Si (correlated with quartz and clay minerals), Al (clay minerals), and Ti

Page 32: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

20

(allochthonous inputs) (e.g., López Laseras et al., 2006; Liu et al., 2013; Hahn et al.,

2014) for our analysis.

2.4.4 Indicators of Biological Paleoproductivity

Total carbon (TC) was determined by combustion using a LECO model SC-

144DR apparatus at the Kentucky Geological Survey (LECO, 2008). The LECO was

primed three times with 0.15-0.20 g of pure coal powder, then calibrated using three

standards: 0.25-0.30 g of SARM 41, 0.20-0.30 g of 502-030 (LECO, 2015), and 0.10-

0.15 g of 502-630 (LECO, 2015). All primes and standards were run for 180 sec. The

502-321 combustion catalyst standard was added to each sample, and run for 60-70 sec.

(LECO, 2014). Standards were re-run at the end of each day. Total carbon percent values

were calculated as the sample mass after combustion, compared to the mass prior to

combustion.

Carbonate content in a sample size of ~0.10-0.12 g was measured with the UIC,

Inc. CM5014 CO2 coulometer, coupled to a UIC CM5130 acidification unit (UIC, 2017).

Phosphoric acid (1:10) was used to dissolve the inorganic carbon fraction, thereby

generating CO2 gas. Carbonate (12%) standards of 10.5-11.5 mg were run between every

ten samples. The CO2 gas was measured as µg of carbonate in the coulometer cell, and

converted to weight percent inorganic carbon. Carbonate was calculated with the

equation below (Liu et al., 2013):

%carbonate = (μg carbonate- μg blank value)/(μg sample mass) x (100/12) x 100 (1)

Total organic carbon (TOC) was computed as the mass difference between LECO

total carbon and coulometry-derived total inorganic carbon values (e.g., Formolo and

Lyons, 2007; Liu et al., 2013). Values of TOC serve as an important indicator of

biological productivity (Al-Ghadban et al., 1994; Dean and Gorham, 1998; Kigoshi et al.,

2014).

Biogenic silica (BiSi) concentration data provides additional information on

productivity, particularly related to the diatom flora living in the lake and inflowing

rivers. The BiSi concentration values also reflect all other biological siliceous

Page 33: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

21

contributions, which may be from phytoliths and sponge spicules. Freeze dried and

ground samples (~0.5 g) were shipped to the Sedimentary Records of Environmental

Change Laboratory at Northern Arizona University (NAU) for BiSi analysis, which

followed the hot base (Na2CO3) leaching method outlined in Mortlock and Froelich

(1989). The BiSi was reported as a weight percentage of the entire sample.

2.4.5 Geochronology

Age control was attempted using complimentary methods including 14C, optically

stimulated luminescence (OSL), and fallout radionuclides (137Cs and 210Pb). Radiocarbon

(14C) dating was conducted on charcoal, bulk sedimentary OM, and mollusk (clam) shells

in sediment cores 2, 3, and 4 (Fig. 2.1D). Core 2 was selected to examine the effects of

aquatic macrophyte growth on sediment deposition. Cores 3 and 4 were chosen for their

location in the deepest part of the lake, which notionally limits the effects of subaerial

exposure during periods of drought. How much a drought affects lake levels depends

mainly on lake volume, which itself is a limnological filter influencing sediment

deposition.

For radiocarbon dating, sediment samples were sonicated in a water bath, filtered

through 150 and 63 μm sieves, and charcoal was hand-picked under a dissecting

microscope using cleaned tweezers and paint brushes. Alternatively, sediment samples

were sieved as a freeze-dried powder, and charcoal was picked with tweezers. Fresh-

looking (e.g., angular, vitreous) charcoal fragments were collected from the 63-150 μm

size fraction. Because charcoal fragments were typically very small, an aggregate of ~10

mg was collected. For horizons of interest where charcoal was absent, fine-fraction bulk

OM was measured. For mollusks, we dated shells of living individuals to assess the

potential for an old carbon reservoir effect at Lake Uberaba (Riggs, 1984; Yu et al., 2007;

Zhang et al., 2016; Olsen et al., 2017). All samples were sent to the Center for Atomic

Mass Spectrometry (CAMS) at the Lawrence Livermore National Laboratory (LLNL).

Samples were pre-treated at LLNL using the standard acid-base-acid (ABA) treatment

with two base washes (ABBA) to remove contaminants. Most samples (bulk sediment

and charcoal) were treated with 1 N HCl, followed by as many cycles of 1 N NaOH as

needed. The HCl treatment was repeated, and the samples were washed three times in

Page 34: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

22

Milli-Q water, dried overnight on a heat block, and combusted. The samples were

graphitized according to Vogel et al. (1987). The mollusk shells were acidified under

vacuum with phosphoric acid and graphitized. Results were reported according to Stuiver

and Polach (1977), and converted to calendar ages with CalPal (http://www.calpal-

online.de/) for charcoal and OM dates, and Calibomb (http://calib.org/CALIBomb/) for

the post-bomb dates.

In a section that was dominantly sandy at the bottom of core 1, one sample was

dated using OSL. This core was carefully split in a dark room at LacCore, and samples

were delivered to the Luminescence and Gamma Spectrometry Laboratory (LEGaL) at

the University of São Paulo for pretreatment and analysis (Table 2.2). To isolate the

quartz fraction, the sediment sample was wet sieved to 63-125 μm, and treated with

27%H2O2 and 10%HCl to remove organic components and carbonates, respectively. The

quartz fraction was isolated by heavy liquid separation, and treated with 38%HF (Aitken,

1998). This procedure has been used successfully in several case studies (e.g., Sawakuchi

et al., 2016; Bertassoli et al., 2017; Pupim et al., 2017).

The OSL dating of quartz aliquots was based on the single aliquot regenerative

(SAR) dose protocol (Murray and Wintle, 2000; 2003; Table 2.3), and functions by

treating quartz grains as dosimeters that record low amounts of ionizing, ambient

radiation (Aitken, 1998). A Risø TL/OSL DA-20 with 90Sr/90Y beta radiation source

(dose rate of 0.0868 Gy/s), blue LEDs for stimulation, and Hoya U-340 filter for light

detection in the ultraviolet band was used to determine equivalent dose based on the

Central Age Model (Galbraith et al., 1999). Equivalent dose calculations were made only

on aliquots with a recycling ratio of 0.9-1.1 and recuperation < 5%. Infrared stimulation

was used to test for feldspar contamination in the quartz aliquots. Only aliquots with

negligible infrared stimulation signal was used for equivalent dose calculation. For the

dose recovery test, six sample aliquots were bleached for 3-5 hours, preheated to 200°C,

and subjected to a 25 Gy given dose. The average calculated-to-given dose ratio was 0.96

± 0.01.

To calculate dose rate (1.34 ± 0.09 Gy ka-1), radionuclide (U, Th, and K)

concentrations were determined from high-resolution gamma spectrometry using a

Canberra high purity germanium (HPGe) detector and background shield. The sample

Page 35: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

23

was sealed and given 28 days to reach secular equilibrium with parent radionuclides

(Aitken, 1998). The U, Th, and K concentrations were converted to dose rate using

conversion factors outlined by Guérin et al. (2011). Cosmic dose rates were calculated

according to Prescott and Hutton (1994). OSL age was calculated by (Dose)/(Dose Rate).

Table 2.2: Sediment pretreatment procedure as outlined for the LEGaL lab.

1. Wet sieve to extract the 63-125 μm fraction.

2. Oxidize OM with 27%H2O2.

3. Dissolve carbonate minerals with 10%HCl.

4. Isolate quartz fraction with heavy liquid (lithium metatungstate solution,

LMT = 2.85 g/cm3)

5. Eliminate remaining feldspar grains and remove ~10 μm outer layer of

quartz grains with 38%HF for 40 minutes

Table 2.3: Outlined SAR-OSL protocol to estimate equivalent dose (Pupim et al., 2017)

1. Dose (Di)

2. Preheat at 200°C for 10 sec

3. OSL at 125°C for 40 sec (Li)

4. Test dose

5. Heat at 160°C

6. OSL at 125°C for 40 sec (Ti)

7. Blue LED bleach at 280°C for 40 sec

8. Return to step 1

Activities of the fallout radionuclides 137Cs and 210Pb were measured to help

constrain geochronology and determine sediment accumulation rates at each core

location, using sequential samples (1 cm intervals) that were oven dried at 70°C, and

ground by Retsch mortar miller or mortar-and-pestle. This rate is known as the sediment

mass accumulation rate (SMAR = Δm/Δt), where m is the mass depth (g cm-2) (Appleby

and Oldfield, 1978; Yeager et al., 2012). Application of fallout radionuclides to lakes has

been useful, particularly when paired with other age dating methods (Eakins and

Page 36: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

24

Morrison, 1978; Oldfield and Appleby, 1984; Hercman et al., 2014; Baskaran et al.,

2014; Andersen, 2017). Lead-210 (half-life of 22.26 yrs.) accumulates in sediment from

constant atmospheric fallout (Appleby and Oldfield, 1978). This fallout originates from

238U decay, which produces 226Ra that adsorbs onto fine sediment particles. Radium-226

decays into 222Rn, a noble gas, which re-enters the atmosphere and decays into 210Pb

(Noller, 2000). Atmospheric 210Pb is provided to the Earth’s surface by both wet and dry

deposition, will rapidly adsorb onto fine sediment, and is known as unsupported, or

“excess” 210Pb (210Pbxs) activity. Supported 210Pb activity is produced from in situ decay

of 238U. Because 210Pb is not an alpha particle emitter, 210Po is counted as a proxy,

assuming that secular equilibrium between 210Pb and 210Po has been reached.

210Pb is measured by alpha spectrometry counts of 210Po. Counting the 210Po

activity is a major advantage due to the small sample size required and no 226Ra

background activity from mineral lattices (Aalto and Nittrouer, 2012). The radioactive

tracer 209Po (Eckert and Zeigler Isotope Products) was added in aliquots of 0.5 ml to each

1.0-1.1 g sample to quantify the 210Pbxs activity along each core to a depth of ~30 cm

b.l.f. If supported 210Pb activity was not reached by 30 cm b.l.f., alpha counting was

extended to ~50 cm b.l.f. A series of acid digestions with concentrated HCl, HNO3, and

HF and continuous heating at 125-130°C dissolved the sedimentary matrix. Next, the

samples were treated with ascorbic acid to remove ferric iron (Yeager et al., 2012) and

the 210Po was plated onto 99.9% silver discs (Yeager et al., 2012). The disks were

transferred to a Canberra model 7404 integrated alpha spectrometer and model 8224

multi-channel analyzer, and counted until 209Po activity uncertainties were < 5% in fine

grain sediments, and < 10% in coarser grain samples. The desired SMAR rate was

determined assuming constant flux (Robbins, 1978):

210Pbxs = F210 e(-λt/SMAR) (2)

where F210 represents 210Pb flux, λ is the 210Pb decay constant (0.031 yr-1), and t = time

(years). The constant flux model (also known as Constant Rate of Supply) requires that

supported 210Pb is known for each core, that activity declines exponentially with depth,

and that 210Pbxs has been in constant flux (Appleby and Oldfield, 1978; Binford, 1990;

Page 37: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

25

Sanchez-Cabeza and Ruiz-Fernández, 2012). This model was selected because it is a

proven realistic model that works where the SMAR has changed over time due to

meandering fluvial channels.

Cesium-137 is a fallout radionuclide with a half-life of 30.2 years. It is primarily

derived from surface detonation of nuclear weapons beginning in ~1952, and peaking in

1963, just prior to the enforcement of the Partial Nuclear Test Ban Treaty (Livingston

and Bowen, 1979; Huntley et al., 1995). Cesium-137 was measured in Canberra high

purity germanium (HPGe) detectors coupled with model DSA-1000 multi-channel

analyzers to determine mass accumulation rates (MAR) with the following equation:

S = (Dpk/t) (3)

The S represents MAR (g cm-2 yr-1), Dpk is the cumulative mass depth (g cm-2) at which

the 1963 137Cs peak occurs, and t is time (years) since the 1963 peak in 137Cs fallout.

For gamma counting, 2g of sediment was homogenized and packed into clear

plastic 10-ml test tubes to achieve a volume-to-mass ratio of 2 ml to 2 g. If the 1:1 ratio

was not attainable, silica gel was added as an inert filler. If insufficient sample material

was available, the sample was prepared at 1 ml to 1 g. The samples were sealed with

epoxy and given 21 days for the isotopes to reach secular equilibrium (226Ra and 222Rn),

prior to placing them in the detectors. We assumed that 137Cs would be present in

detectable concentrations, particularly during the peak of above-ground nuclear bomb

testing in 1963.

Page 38: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

26

CHAPTER THREE: RESULTS

3.1 Morphometrics and Limnology

Measurements made on satellite images from 2015 reveal that the lake covers an

area of ~230 km2 (Fig. 2.1A). Over the last ~30 years, the surface area of the lake has

declined (Fig. 3.1), interpreted to be the result of sediment progradation aided by

extensive macrophyte colonization along the northern margin (Table 3.1) (Lo et al.,

2017). The northern margin had low total suspended solids (TSS) concentrations (Table

3.2), which were elevated along the southern margins of the lake in November 2015. In

terms of its hydrology, the lake does not receive direct input from the Paraguay River, so

sediment and water delivery primarily occurs through ephemeral streams (also known as

vazantes or stream traces) on the Paraguay megafan, the Corixo Grande River, and the

Canzi River (Resende, 2000; Assine and Silva, 2009). During markedly dry interannual

periods, Lake Uberaba becomes a partially desiccated lake (~36 km2) fed only by the

Corixo Grande River, and draining through Canal Dom Pedro II (Fig. 3.2). Lake Uberaba

is shallow, reaching a maximum depth of ~3.5 m, with a mean depth of 2.2 m. Lake

Uberaba is deepest (~3.5 m) in the south and becomes shallower along the northern

margins (~1 m). The southwest and eastern lake margins are steeper, reaching ~2.1-2.8 m

deep a short distance from the shoreline. By contrast, the northern margin exhibits a

much gentler bathymetric gradient, with a broad littoral plain that is < 2.1 m deep. One

shallow channel-like feature is observed extending from the eastern margins of the lake

(Fig. 2.2). Distinct environments exist around the lake margins (Fig. 3.3), particularly on

the banks of Precambrian metasandstone inselbergs (Fig. 2.3) (Cole, 1960; Pupim et al.,

2015; Warren et al., 2015). Water hyacinths (Eichhornia crassipes) and sedges (Salvinia

sp.) have colonized the northern lake margins, whereas pioneer woodland species (e.g.,

Vochysia divergens, Curatella americana) occupy the western margin (Fig. 3.3, sub-

images 2, 3, 4) (Pott and Silva, 2015).

Page 39: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

27

Figure 3.1: Results of landscape classification and mapping from SPRING, based on

Landsat 5 and 7 images. Water (blue) and surrounding terrestrial vegetation (green) of

Lake Uberaba. The open water area decreased from 463 km2 (1985) to 364 km2 (2011), a

21.4% difference. Major progradation of aquatic macrophytes occurred during this period

on the northern margins of the lake, and the growth of aquatic plants that were spectrally

classified as terrestrial (green) or surroundings (all other types of non-flooded surfaces;

blank) (Lo et al., 2017). All images were taken from July to October to map the extent of

the lake and vegetation during the dry season. More information on lake surface and

macrophyte coverage are presented in Lo et al. (2017).

Page 40: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

28

Figure 3.2: The reconnaissance satellite image from 1968 shows Lake Uberaba occupying only an area of ~36 km2 during an acute,

multi-year drought (Bergier and Resende, 2010). Major fluvial inputs and outputs are traced in yellow in the 1968 image. The

Landsat-8 satellite image from 2015 shows the current configuration of Lake Uberaba.

Page 41: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

29

Figure 3.3: Photos documenting the diversity of lakeshore vegetation around Lake Uberaba. Woody vegetation was more prevalent

along the southern margins, though the short lifecycle and rapid proliferation of non-woody species on the northern margin

contributed to relatively higher (> 12%) TOC concentrations in sediment here and the accumulation of peaty silt and sand. 1) Marsh

with dead brushy vegetation and crabs, 2) Woody vegetation with large trees stabilizing the shoreline, 3) Fields of water hyacinth and

sedges (see Souza et al., 2011), 4) Water hyacinth and sedges with pioneer woody species, 5) Well-developed trees growing on

Malvinas Island, with camalotes of water hyacinth in the foreground, 6) Fields of sedges thriving with little competition from water

hyacinths, and 7) Ecological succession illustrated from water hyacinths in the foreground to sedges with woody species to mature

trees on Pato Island (left).

Page 42: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

30

Table 3.1: Numerical results of the land surface thematic classification are reproduced

from Lo et al. (2017), which organized land cover into open water, aquatic macrophytes,

and surroundings (Fig. 3.1). Open water area has generally declined, accompanied by

increased macrophyte surface area. The Paraguay River stage indirectly modulates these

values, particularly evident in 2001 when stage decreased < 500 cm.

Year Open water area (km2) Macrophyte surface area

(km2)

Paraguay River stage (cm)

1985 463 567 543

1991 504 572 565

1996 463 602 506

2001 343 738 463

2006 396 643 538

2011 364 644 540

Water chemistry was distinctly different between September 2015 and June 2016.

In September, the mean water temperature was ~5.5°C warmer (28.0°C versus 22.5°C),

dissolved oxygen was lower by ~10.2%, total dissolved solids were lower (0.04 mg l-1),

and electrical conductivity was higher by ~7.7 µs cm-1. Samples from both periods

exhibited similarly neutral pH (average 7.17) and low salinity (0.02-0.03 ppt) (Table

3.3). Thus, Lake Uberaba can be classified as a shallow, freshwater, well-oxygenated (>

70% dissolved oxygen) lake. Lake Uberaba is slightly basic with an average alkalinity of

582 μeq (l HCO3-)-1. Dissolved ion analysis showed that Ca2+ was the most readily

available dissolved ion, averaging 7.4 mg l-1. The second most dominant cation was Na+,

whereas the dominant anion was HCO3- (Table 3.2).

Page 43: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

31

Table 3.2: Ion chromatography data for water samples collected from Lake Uberaba in

September 2015 (Fig. 2.1C). Abbreviations are as follows: total suspended sediment

(TSS mg l-1), inorganic suspended sediment (ISS mg l-1), and organic suspended sediment

(OSS mg l-1).

A B C D E F G

Latitude -17.455 -17.4553 -17.51 -17.4919 -17.5464 -17.4919 -17.4417

Longitude -57.9097 -57.8528 -57.8531 -57.8153 -57.7592 -57.7592 -57.7425

Alkalinity

(μeq l-1

HCO3-)

540.9 ±

21.4 N/A

729.7 ±

11.6

495.8 ±

13.9

577.2 ±

10

589.4 ±

11.0

557.6 ±

17.1

TSS 6.0606 6.7084 5.8244 24.2224 5.6655 8.1081 0.4440

ISS 4.4966 5.0089 4.1219 21.5834 4.4065 6.5766 0.1776

OSS 1.5640 1.6995 1.7025 2.6390 1.2590 1.5315 0.2664

F-

(mg l-1) 0.1083 0.0602 0.0965 0.0690 0.0780 0.0497 0.091

Cl-

(mg l-1) 0.3321 0.4162 0.5567 2.4809 0.2783 0.1452 0.6390

NO31-

(mg l-1) 0.1061 0.0575 0.0974 0.0904 0.0612 0.0819 0.0275

SO42-

(mg l-1) 0.0298 0.0208 0.0274 0.0657 0.0625 0.0498 0.0185

Na

(mg l-1) 3.2994 2.9153 4.9469 2.7496 2.9133 2.3433 2.9923

K

(mg l-1) 2.3567 2.2933 3.6749 1.7878 1.9834 1.2973 2.6319

Mg

(mg l-1) 2.3155 2.2283 2.5664 2.5412 2.4114 2.4604 2.4728

Ca

(mg l-1) 7.1238 6.859 7.5617 7.4363 7.4222 7.5166 7.9996

Page 44: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

32

Table 3.3: List of water chemistry characteristics obtained from the YSI multi-parameter

sonde. Samples I-V were measured in June 2016 (Fig. 2.1C), and samples VI-IX were

collected in September 2015. Abbreviations are as follow: temperature (Temp. °C),

dissolved oxygen (DO %), salinity (Salt. PSU), total dissolved solids (TDS g l-1), and

conductivity (Cond. μs cm-1). Conditions and timing allowed for duplicate measurements

at sites I, V, and VII. Location data were not recorded for sites VII, VIII, and IX.

Temp. pH DO Salinity TDS Cond. Depth (m)

I 22.4 7.13 73.2 0.02 31.13 45.5 0.15

21.6 6.88 72.4 0.02 33.95 48.8 2.74

II 23.6 7.5 100.5 0.03 37.71 56.5 0.19

III 23.4 7.23 89.7 0.03 43.92 66.1 0.24

IV 22.9 7.48 96.7 0.03 44.54 65.7 0.39

V 22.3 7.66 100 0.03 39.21 67.3 0.28

21.2 7.55 94.9 0.03 39.05 55.7 3.21

VI 28.01 7.05 73.5 0.03 0.04 65 n/a

VII 28.2 7.04 90.4 0.03 0.043 70 n/a

27.6 6.84 88.4 0.03 0.044 70 n/a

VIII 28.5 6.64 71.8 0.03 0.037 61 n/a

IX 27.7 7.02 73.3 0.03 0.038 62 n/a

Page 45: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

33

3.2 Lake floor sediments

The D50 (median) grain size of each lake floor sediment sample ranges from 10.3-

148.6 µm (Fig. 3.4). Silt is the dominant grain size (mean = 54.4%, range = 3.2-81.6%),

followed by sand (mean = 35.7%, range = 4.4-95.8%). Clay was the least abundant grain

size (mean = 9.8%, range = 1.0-37.1%) along the basin’s central axis (N to S), but on the

eastern and western lake margins, the clay size fraction was the second most abundant

component. Most coarse grained sediment was found in an elliptically shaped area near to

the basin center, located between the two islands on the eastern side of the lake (Fig. 3.4).

When suspended sediments were measured on November 14, 2015, the suspended

sediments were concentrated along the southwestern margins of Uberaba (Fig. 3.5).

Figure 3.4: Contour maps of different compositional characteristics of Lake Uberaba

surface sediments, and suspended sediment concentrations. A) D50 particle size. B) Total

organic carbon (TOC). C) Biogenic silica. Biogenic silica and TOC concentrations were

higher in the fine grained sediments of the eastern and western margins. Sampling sites

are indicated by plus (+) symbols.

Page 46: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

34

Figure 3.5: Suspended sediment concentrations were greater near to the southern lake

margins in November 2015 due to hydrologic flow and resuspension. Sampling sites are

indicated by plus (+) symbols.

Page 47: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

35

Carbon content in surface sediments was dominated by the organic fraction. Trace

amounts (on average < 110 ppm or 0.011%) of TIC were measured in our coulometric

analysis. The TOC (average of ~1.7%) effectively constituted total carbon (TC) (Fig.

3.4). The TOC concentrations are between 0 and 2 weight percent (wt%) in most of the

lake from the northern margins to the primary outflow Canal Dom Pedro II at the

southern margin. The notable exceptions are the fluvial inputs along the eastern and

western lake margins, where TOC is 2-6 wt%. One sampling site along the eastern

margin has up to ~12 wt% TOC, and is located near to the suspected primary inflow of

the Canzi River. The location of this sampling site away from the macrophyte growth of

the northern margins was surprising. This relationship suggests that TOC is primarily

concentrated at fluvial point sources and may be comprised of allochthonous (terrestrial

plant) material.

Biogenic silica (Fig. 3.4) was concentrated in areas with greater TOC, i.e.,

principally along the eastern and western lake margins. The BiSi was lowest (~1-2%)

where sand is concentrated in the center of the lake. Much of the lake is dominated by

BiSi values in the range of 2-3% and mostly composed of sponge spicules, primarily

megascleres seen in the smear slides. The perennial fluvial inputs along the eastern and

western lake margins contain greater concentrations of BiSi (3-6%). The same sampling

site where TOC was ~12% (-17.5103, -57.7214) also has the highest BiSi value (~6%).

Surface samples demonstrated a pattern of elemental concentrations (bioavailable

and XRF) spatially linked to the grain size distribution (p-value < 0.0001 for 30-80 μm)

based on t-test analysis between pairs of elements (see Appendix C). The XRF elemental

counts (Fig. 3.6) showed that areas with coarser sediment presented higher values of Si.

However, the elements Al, K, Ca, Ti, and Fe were present in greater percentages in areas

comprised of fine-grained sediment. Similarly, concentrations of bioavailable K, Mg, and

Ca (Fig. 3.7) were closely coupled (p < 0.0001) spatially with the clay- and silt-sized (5-

80 μm) sediment fraction. Iron was strongly concentrated in southeastern Lake Uberaba.

Bioavailable elements, especially P, can limit vegetation growth, which affects water

chemistry and indirectly impacts sedimentation (e.g., Costa and Henry, 2010).

Concentrations of bioavailable P were lower along the northern margins, possibly

Page 48: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

36

suggesting that more intense macrophyte growth is depleting P relative to other areas of

the lake.

Figure 3.6: Plot of XRF elemental abundances measured from surface sediment samples

showed that XRF elemental abundance varied as a function of grain size, with lower

abundances in areas of coarse deposits. Plus (+) symbols represent sampling sites.

Page 49: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

37

Figure 3.7: Like the elemental weight abundance distribution (Fig. 3.6), these

bioavailable element distributions visually correspond well with grain size distribution

(Fig. 3.4A). Concentrations were elevated in finer sediments. Iron was an exception and

did not appear to be influenced by grain size distribution, as it has been eroded from

Precambrian rocks in the watershed upstream and entered from the Canzi River

(extension of active lobe). Sampling sites are indicated by plus (+) symbols.

Page 50: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

38

3.3 Seismic Profile

In one type of seismic character, multiples were rare, and distorted channels were

identified where the features have steep sides. One discontinuous but traceable subsurface

reflector allowed for the identification of paleochannels (Fig. 3.8A, B). Filled

paleochannels appeared as U-shaped cross sections in the transect. These channels

became more common towards the southern part of the transect line. However,

determining the orientation of the paleochannels was not possible with only one seismic

transect line.

Some limited areas towards the southern part of the lake contain ~2-5 continuous

multiples practically devoid of sedimentary features (Fig. 3.8C). This acoustic facies has

a thin layer of sedimentation (~0.3-0.4 m) capping the continuous high impedance

reflector (~0.3 m thick) along some sections of the profile. Over a depth of ~2 m, the high

impedance reflector attenuates rapidly. Acoustic blanking interrupted the strata in several

locations.

The shallow acoustic facies (Fig. 3.8D) shows thick (> 1 m) sections with faint

reflectors up to 3 m below the lake floor. The acoustic signal is characterized by a thin

(~0.2 m) sharp return representing the lake floor, usually ~2-3 m below the water surface.

The lake floor reflection has an undulating appearance in some areas, created by wave

action during the survey. Immediately below the lake floor, the signal is dampened by

soft sedimentary deposits, with some areas presenting acoustic blanking where

subsurface stratigraphy is absent for a short (< 100 m) distance. Some discontinuous

sections of the profile have a stronger signal, reflecting beds with denser deposits.

Isolated locations along the profile return maximum signal strength at the lake floor

accompanied by 2-3 multiples, indicating the potential for a dense surface or very limited

penetration. This facies was most common in the northern part of the lake.

Page 51: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

39

Figure 3.8: A) Interpreted and B) uninterpreted subsurface profile with a suspected relict floodplain surface (yellow line) 4 m below

the water surface. Numbers mark depth from water surface. Sedimentary fill below the yellow line likely constitutes Unit 2, whereas

the material above the yellow line likely constitutes Unit 3. C) Subsurface profile with continuously occurring 2-5 multiples and

absent sedimentary structures. D) Subsurface profile showing faint sedimentary structures, with rare occurrences of ~2-3 multiples.

Page 52: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

40

Figure 3.9: Core 3 integrated stratigraphy, geochronology, geochemistry, and physical properties. A) Lithology and geochronology,

B) grain size, TOC, BiSi, and XRF (Al, Si, Fe, K, Ti), and C) smear slide sample images. Interpretation of the lithofacies and

environmental conditions are listed in the far right column. Core 3 preserved the transition from subaerial floodplain to subaqueous

lake, including sediment mixing (10-22 cm). Silty sand deposits were derived from the active lobe of the Paraguay fluvial megafan.

Page 53: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

41

Figure 3.10: Core 4 integrated stratigraphy, geochronology, geochemistry, and physical properties. A) Lithology and geochronology,

B) grain size, TOC, BiSi, and XRF (Al, Si, Fe, K, Ti), and C) smear slide sample images. Core 4 yielded a slightly different

composition from nearby core 3. Although sediment mixing evident in core 3 was absent in core 4, the sandy silt represents the time

period when Lake Uberaba formed. The uppermost unit was thicker due to its proximity to Canal Dom Pedro II.

Page 54: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

42

Figure 3.11: Core 2 integrated stratigraphy, geochronology, geochemistry, and physical properties. A) Lithology and geochronology,

B) grain size, TOC, BiSi, XRF (Al, Si, Fe, K, Ti, Ca), and magnetic susceptibility, and C) smear slide sample images. The lithologic

character of Core 2 illustrates the immediate influence of the Paraguay fluvial megafan. On the toe of the relict lobes, the colonization

of aquatic vegetation enhanced TOC concentrations, until sand was directly deposited over the core site.

Page 55: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

43

Figure 3.12: Core 1 integrated stratigraphy, geochronology, geochemistry, and physical properties. A) Lithology and geochronology,

B) grain size, TOC, BiSi, and XRF (Al, Si, Fe, K, Ti), and C) smear slide sample images. An optically stimulated luminescence (OSL)

age date was obtained at 21,155 ± 1445 years, so this core recovered sediments from the Last Glacial Maximum (LGM). An extended

period of massively bedded clayey silt was deposited between the LGM and late Holocene.

Page 56: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

44

3.4 Lithostratigraphy

The longest cores in our data set (Cores 3 and 4) showed an abrupt change from a

basal, tan, massively bedded, clayey lithostratigraphic unit (Unit 2) to an overlying

brown, massively bedded, relatively organic-rich, sandy lithostratigraphic unit (Unit 3).

This stratigraphic transition is illustrated in Figure 3.9. Core 3 contained a sharp change

at 23 cm b.l.f., between Units 2 and 3, whereas this transition occurred at ~35 cm b.l.f. in

Core 4 (Fig. 3.10). Core 2 contained two lithostratigraphic units; the uppermost unit was

comprised of two sub-layers (Fig. 3.11). The base of Core 2 consists of a dense,

massively bedded, organic-rich silty lithostratigraphic unit (Unit I), which transitions (41-

46 cm b.l.f.) into a porous, peaty silt lithostratigraphic unit (Unit II) lacking a noticeable

grain size change from the underlying sediments. Overlying the peaty unit is a dark tan,

massively bedded, sandy silt sub-unit (Unit III). Core 1 consists mostly of the massively

bedded Unit 2A (2-62 cm b.l.f.) (Fig. 3.12). This core is capped by a sandy layer,

representing contemporary lacustrine deposition.

Core 3 (Fig. 3.9) is the longest (~1 m) core available from Lake Uberaba. In terms

of its grain size characteristics, Core 3 transitions from a basal clayey silt (51-98 cm), to

sandy silt (12-51 cm), to an uppermost silty sand (0-12 cm). Sand-filled desiccation

cracks were identified in the basal lithofacies. The increase in sand-sized particles up-

section is one of the most striking physical properties transitions in core 3. No grain size

change occurs at the abrupt lithostratigraphic transition at ~23 cm b.l.f. from Unit 2 to

Unit 3, but sand does markedly increase at ~11 cm b.l.f. The BiSi content ranges from

1.0-2.5%, with a whole-core average of ~1.7%. From the perspective of opal

chemostratigraphy, a slight decline is apparent at the ~23 cm b.l.f. lithostratigraphic

transition, as Unit 2 BiSi values are generally between 1.0-1.5%. Organic matter is

virtually absent in Unit 2. The TOC concentrations begin to increase at ~23 cm, reaching

a maximum (~0.6%) near the top of the core in Unit 3. Within Core 3, the %Ti and %K

concentrations peak in the lowermost clayey silt (Unit 2) and decline above the

lithostratigraphic transition, where silty-sands dominate (Fig. 3.9). The Al values in Unit

2 are ~6%, and remain relatively constant until the lithostratigraphic transition at 23 cm

b.l.f., where values decline to 3-4% (Fig. 3.9). Opposite of the %Al data, %Si values

Page 57: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

45

increase to a maximum of ~28.8% in Unit 3. Iron is notably lower (0.65%) in Unit 3 than

in Unit 2, where values are 1.5-3%.

Core 4 (Fig. 3.10) provides an important comparison with Core 3, due to similar

coring sites in southern Lake Uberaba. Grain size consists of clayey silt (35-86 cm),

transitioning to silty sand (27-35 cm), and then to an uppermost sandy silt (0-27 cm).

Unlike the transition in Core 3, the abrupt change at ~35 cm from light (Unit 2) to dark-

colored (Unit 3) sediment is accompanied by significant grain size change from clayey

silt to silty sand. The BiSi values range between 1.1 and 1.9%, with an average of 3.6%

(40-84 cm); BiSi content declines significantly at 35 cm b.l.f. Organic matter is virtually

absent in Unit 2. TOC concentrations first exceed detection limits at 35 cm b.l.f., and

increases to 1% towards the top of the core. Percentages of Ti and K (0.67% and 0.61%,

respectively) from 36 to 84 cm b.l.f. are, on average, higher in Unit 2, and are lower, on

average, after the facies change at 35 cm b.l.f. The percentage of Al is 5.8% (36-84 cm)

in Unit 2, and drops to 3% (2-34 cm) in Unit 3. The %Si averages 24.2% in Unit 2, and

29.5% in Unit 3. The %Fe values were stable (~2.5%) in Unit 2, and decreased beginning

at 20 cm (~1.6%) in the upper Unit 3.

Due to its distinct composition, Core 2 (Fig. 3.11) lithologies were given units

with Roman numerals to consider them separately from the other three cores. The core

contains clayey silt (18-54 cm, Units I and II), followed by an abrupt change to sandy silt

(0-18 cm, Unit III). Silt is the dominant grain size fraction. Volume magnetic

susceptibility (MS) levels, measurable only in cores collected in polycarbonate sleeves,

peak at the bottom of the core, but gradually decreases to as low as 1.1 SI x 10-5 at 18 cm

in the peaty silt (18-38 cm). Concentrations of TOC peak (~15%) in the peat-like silt (18-

38 cm), and then taper towards the ends of the core. BiSi values range from 1.6% to

6.2%, averaging 3.8%. Both TOC and BiSi are elevated in the middle lithofacies, where

macroscopic plant OM is abundant. The K elemental abundance average 0.25% (7-53

cm), and %Ti decreases gradually up section. The %Ca peaks at ~0.49% in the peaty

unit, and decreases in the upper and lower core. Iron averages ~1.5% in the peaty and

clayey section, but rapidly decreases to 0.3% (5-15 cm).

Core 1 (Fig. 3.12) exhibits a limited thickness (~2 cm interval) of Unit 3, but

provides a good preservation of Unit 2 (2-62 cm). The grain size of the core changes

Page 58: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

46

from a basal (66-72 cm, Unit 1) sandy composition, to clayey silt (2-62 cm, Unit 2), and

is capped by sandy silt (0-2 cm, Unit 3). BiSi values range from 1.5% to 2.9%, averaging

2.4% for the whole core. TOC concentrations are very low in the basal deposits (> 40

cm), and increase to 1.4% at the core top (1 cm). The %K values peak at 0.73%, and

average 0.68% (37-65 cm), and 0.61% (3-33 cm). The %Ti values range from 0.35-

0.76%, and increase towards the core top. The %Al is ~7% in sections where clay size

sediments are the secondary component of grain size, but decreases when sand becomes

the secondary component. The values of %Al and %Si are inversely related. The %Fe

values are higher in the clayey silt unit (3.7%, 25-61 cm b.l.f.), and generally decreases

towards the upper part of the core.

3.5 Geochronology

The radiocarbon ages (Table 3.4, Table 3.5) for bulk sediment differed

significantly from charcoal ages from the same stratigraphic horizons by as much as

several hundred years. For example, at 19 cm b.l.f. in Core 3, charcoal returned a

calibrated age of 1,750 ± 80 cal yr BP, whereas bulk sedimentary OM was dated to 3,460

± 90 cal yr BP. However, bulk ages in the uppermost lithostratigraphic unit were precise.

The bivalve buried ~10 cm deep in core 3 returned a post-modern (> 1950 CE) age of

1958 CE (1958.25-1958.48 CE), which shows bioturbation. Pre-modern ages (prior to

1950 CE) were rounded to the nearest tenth year because rounding to the nearest year is

too precise with the models that were used (Weninger and Jöris, 2008; Santos et al.,

2015).

Core 4 contained silty sand at 27.5-35 cm b.l.f., from which three charcoal ages

(210 ± 120, 140 ± 100, and 230 ± 70 cal yr BP) were determined, with dates within 2-

sigma error of each other. This lithologic boundary was visually correlated with a

similarly abrupt change at 23 cm b.l.f. in Core 3.

At the bottom of Core 1, OSL gave an age of 21,155 ± 1,445 years computed

using the Central Age Model (Galbraith et al., 1999) (Table 3.6). All 24 quartz aliquots

met the conditions (recycling ratio, recuperation, and feldspar test) to be included in the

equivalent dose calculations. This age is consistent with the sandy facies deposited prior

to the onset of the floodplain lithofacies.

Page 59: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

47

Table 3.4: List of post-bomb radiocarbon ages determined at Lawrence Livermore

National Laboratory’s Center for Atomic Mass Spectrometry. Most post-bomb dates

were ~1958 AD. Note that all ages are converted to calendar age (AD) by CALIBomb.

CAMS

Number

Sample

Name

Core

Depth

(cm)

Material δ13C Fraction

Modern Error

Calibrated

age (AD)

2-σ range

(AD)

175544 Live

mollusk 0

live

shell -12 1.0638 0.0033

1958.18 ±

0.06

1958.07-

1958.30

175543 3A-15

mollusk 10 shell -12 1.0758 0.0033

1958.36 ±

0.06

1958.25-

1958.48

176925 2A-1 1 charcoal -25 1.1237 0.0098 1996.04 ±

0.91

1993.28-

1996.95

176926 2A-21 21 charcoal -25 1.0805 0.0077 1958.43 ±

0.12

1958.19-

1958.67

Page 60: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

48

Table 3.5: List of pre-bomb radiocarbon ages determined at Lawrence Livermore

National Laboratory’s Center for Atomic Mass Spectrometry. Bulk sediment ages were

generally much older than charcoal ages. Note that all ages are converted with CalPal and

reported in calibrated years before present (cal yr BP).

CAMS

Number

Sample

Name

Core

Depth

(cm)

Material δ13C

Age

(14C yr

BP)

Error

Calibrated

age (cal yr

BP)

2-σ range

(cal yr BP)

175313 3A-7c 2 charcoal -25 410 100 430 ± 90 250-610

175289 3A-14b 9 bulk sed -25 3260 80 3,500 ± 90 3,330-3,680

175314 3A-24c 19 charcoal -25 1820 60 1,750 ± 80 1,590-1,910

175295 3A-24b 19 bulk sed -25 3210 80 3,460 ± 90 3,290-3,630

175290 3A-32b 27 bulk sed -25 5780 150 6,600 ± 170 6,270-6,940

175296 4A-1b 1 bulk sed -25 545 35 580 ± 40 490-670

175297 4A-5b 5 bulk sed -25 470 40 520 ± 20 490-550

175298 4A-10b 10 bulk sed -25 620 45 610 ± 40 530-690

175299 4A-20b 20 bulk sed -25 670 45 620 ± 50 530-710

175300 4A-28c 28 charcoal -25 220 60 210 ± 120 0-450

175315 4A-28b 28 bulk sed -25 1575 30 1,470 ± 40 1,390-1,550

175316 4A-34c 34 charcoal -25 135 30 140 ± 100 0-340

175291 4A-34b 34 bulk sed -25 2005 40 1,960 ± 40 1,870-2,050

175317 4A-37c 37 charcoal -25 225 35 230 ± 70 90-370

175292 4A-40b 40 bulk sed -25 18450 2280 22,210 ±

2,670

16,870-

27,540

176927 2A-39 39 charcoal -25 2880 60 3,030 ± 90 2,840-3,220

176928 2A-45 45 charcoal -25 3325 35 3,560 ± 50 3,450-3,660

Page 61: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

49

Table 3.6: All characteristics of the single OSL-dated sample are listed in this table. Dose

rate is displayed as Gy ka-1. OD = overdispersion. N = number of aliquots.

LEGaL code L0761 Latitude -17.491 Longitude -57.853

Elevation 93 m Depth (cm) 65.6-70 N 24

U (ppm) 1.97 ±

0.10

Th (ppm) 6.31 ± 0.36 K (%) 0.50 ± 0.03

Cosmic dose

rate

0.153±

0.015

Total dose

rate

1.34 ± 0.09 Equi. Dose

(Gy)

28.4 ± 0.3

Water (%) 0.167 OD (%) 2.6 OSL age (ka) 21.155 ± 1.4

Page 62: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

50

The 210Pb dating technique showed increasing rates of sedimentation over time in

most cores. In Core 3 (Fig. 3.13), the exponentially declining curve fit the 210Pbxs closely

(r2 = 0.7004). Core 4 data is currently being processed. Core 2 (Fig. 3.14) reached

supported 210Pb at 7 cm depth, and the exponential curve fit reasonably well (r2 =

0.5588). Core 1 (Fig. 3.15) had a poorly-fit exponential trend (r2 = 0.2382) due to

frequent fluctuations in the 210Pbxs activity. Sediment mass and linear accumulation rates

were calculated based on 210Pbxs. Core 3 SMAR (Fig. 3.16) was 0.2 g cm-2 yr-1 at 5.03

years before sampling. The SMAR then declined exponentially to approximately 0.07 g

cm-2 yr-1 at ~100 years before sampling. Linear accumulation rate (LAR) exhibits a

similar trend, from 0.13 cm yr-1 at 5.03 years before sampling to 0.05 cm yr-1 at ~100

years before sampling. Since 210Pbxs is presently being measured for Core 4,

accumulation rates have not been determined yet. Core 2 (Fig. 3.17) SMAR was initially

0.26 g cm-2 yr-1 at 5.14 years before sampling and exponentially decreases to 0.07 g cm-2

yr-1 at ~88 years before sampling. However, between 34 and 42 years before sampling,

SMAR was particularly elevated to some rates above 1 g cm-2 yr-1, corresponding to

peaty silts that comprise Unit II. This elevated accumulation rate was also reflected in the

LAR graph. Core 1 (Fig. 3.18) exhibited accumulation rates closely fitting the trend line

(r2 = 0.8271). The SMAR was initially 0.03 g cm-2 yr-1 at 105 years before sampling and

increased to 1.09 g cm-2 yr-1 at 2.18 years before sampling.

Page 63: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

51

Figure 3.13: Core 3 total 210Pbxs is plotted and best described as an exponentially

declining curve (p-value = 0.58 by student’s t-test). The 210Pbxs did not reach supported

210Pb, but based on the trendline, 210Pbxs can be expected to reach zero by ~17 cm depth.

y = 16.304e-0.163x

R² = 0.7419

0

2

4

6

8

10

12

14

16

18

20

0 5 10 15 20

De

pth

(c

m)

210Pbxs (Bq kg-1)

Page 64: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

52

Figure 3.14: 210Pbxs profile for Core 2. The 210Pbxs reaches supported levels by ~7 cm

depth (p-value = 0.006 by student’s t-test), but the peak between 15 and 20 cm depth

reaches up to 12.12 Bq/kg. Supported levels of 210Pb occur above and below this peak in

210Pbxs activity. This graph was fitted with an exponential curve, because the constant

flux model that was used assumes that 210Pbxs activity exponentially decreases over ~100

years.

y = 19.242e-0.127x

R² = 0.5588

0

5

10

15

20

25

30

0 5 10 15 20 25

Dep

th (

cm)

210Pbxs (Bq kg-1)

Page 65: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

53

Figure 3.15: The 210Pbxs profile for Core 1 (p-value = 0.19 by student’s t-test) exhibits

fluctuating activity levels within an exponentially decreasing trend towards supported

levels (not reached). This pattern is consistent with active sediment homogenization

primarily from hydraulic processes, but bioturbation also plays a role in sediment

remobilization.

y = 34.098e-0.126x

R² = 0.2382

0

5

10

15

20

25

30

0 5 10 15 20 25 30

Dep

th (

cm)

210Pbxs (Bq kg-1)

Page 66: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

54

Figure 3.16: Mass accumulation (p-value = 0.0012) and linear accumulation (p-value =

0.0012) rates plotted over time for Core 3. The overall trend is greater SMAR and LAR

in more recent years prior to sampling, followed by exponential decrease in increasingly

older sediments.

Figure 3.17: Mass (p-value = 3.2 x 10-8) and linear accumulation (p-value = 3.1 x 10-8)

rates plotted over time for Core 2. Although SMAR and LAR are slightly higher in the

last 30 years than in older sediments (> 50 years), the drastically higher accumulation 34-

42 years ago indicates a major environmental shift. These data were fitted with

exponential functions because the constant flux model assumes that accumulation rates

change exponentially over ~100 years in undisturbed cores.

y = 0.2094e-0.014x

R² = 0.7907

0.0

0.1

0.2

0.3

0 20 40 60 80 100 120MA

R g

cm

-2 y

r-1

Years before sampling

y = 0.1374e-0.014x

R² = 0.7907

0.0

0.1

0.2

0.3

0 20 40 60 80 100 120

LA

R c

m y

r-1

Years before sampling

y = 0.9302e-0.025x

R² = 0.1974

0.0

0.5

1.0

1.5

2.0

0 20 40 60 80 100 120

MA

R g

cm

-2 y

r-1

Years before sampling

y = 0.8497e-0.025x

R² = 0.1974

0.0

0.5

1.0

1.5

2.0

0 20 40 60 80 100 120

LA

R c

m y

r-1

Years before sampling

Page 67: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

55

Figure 3.18: Mass (p-value = 4.7 x 10-8) and linear accumulation (p-value = 3.5 x 10-8)

rates plotted over time for Core 1. The trendline fits the SMAR and LAR datasets well.

y = 1.7348e-0.033x

R² = 0.8271

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

0 20 40 60 80 100 120

MA

R g

cm

-2 y

r-1

Years before sampling

y = 0.9691e-0.033x

R² = 0.8271

0.0

0.5

1.0

1.5

2.0

2.5

0 20 40 60 80 100 120

LA

R c

m y

r-1

Years before sampling

Page 68: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

56

CHAPTER FOUR: DISCUSSION

4.1 Modern Limnogeology of Lake Uberaba

In this study, we developed a deeper understanding of Lake Uberaba, which at

present is the Pantanal’s largest lake by surface area (~230 km2). Lake Uberaba is

shallow (austral winter Zmax = 3.5 m), with a well-oxygenated sediment-water interface

due to active mixing by wind-driven waves in the austral summer and fall; water

chemistry, suspended sediment concentrations, and surface sediment geochemical data

strongly suggest a continuously warm polymictic lake (too shallow to develop thermal

stratification; Lewis, 1983). Moreover, the lake is neutral (mean pH = 7.17), and

characterized by a dilute Ca2+-Na+-HCO3- hydrochemistry that is under-saturated relative

to CaCO3. The water chemistry reflects a short residence time, on the scale of days to

months, which functionally inhibits chemical sedimentation (Cohen, 2003). The majority

of our data attests to the important influence of axial and lateral rivers on the

limnogeology of Lake Uberaba (Fig. 4.1). Lake Uberaba is hydrologically open, flowing

out southwards through Canal Dom Pedro II, and sedimentary processes and patterns of

deposition within the lake are conditioned by numerous, and mostly ephemeral, inflowing

rivers associated with the Paraguay River megafan (Assine and Silva, 2009). This

hydrologically open and shallow configuration implies that sediment remobilization on

the lake floor is common, but mineral precipitates are rare.

The influence of megafan rivers on Lake Uberaba is clear in bathymetric data

sets. Our bathymetric survey shows that accommodation is limited in Lake Uberaba.

Water depth, which we interpret as a proxy for accommodation, increases towards the

southern shoreline, and isobaths show the strong influence of lateral fluvial channels. A

NE-SW oriented channel emanating from Lake Uberaba’s eastern margin is particularly

well-defined by the bathymetric contours (Fig. 2.2). The active, easternmost lobe of the

Paraguay megafan is presently being constructed by deposition associated with the

anabranching Canzi River (e.g., Assine and Silva, 2009; Assine et al., 2015a), and we

posit that Lake Uberaba’s eastern margin is also shaped by fluvial processes from the

Canzi system. The bathymetry data and siliciclastic particle size maps (Fig. 4.2) appear to

indicate that episodic flooding from the Canzi River has impacted the eastern side of

Lake Uberaba, attesting to coupling between the evolution of the active lobe of the

Page 69: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

57

megafan and the lake. Satellite images reveal large changes in water levels in Lake

Uberaba over the past several decades. Such fluctuations are common in shallow lakes

formed by fluvial processes in low-gradient foreland basins (e.g., Van Wagoner, 1995;

Sáez et al., 2007; McGlue et al., 2012). We interpret that as water levels fall in Lake

Uberaba, incision by Canzi River-related distributary channels alter the morphology of

the lake floor.

Figure 4.1: Modern depositional environments immediately surrounding Lake Uberaba.

The areas in green contain actively meandering rivers and channelized flow, but prograde

minimally. The areas in tan contain perennially active channels. The northern margin in

yellow has ephemerally active channels that rework and transport sediments, which

aquatic macrophytes help to stabilize and further accelerate progradation.

Page 70: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

58

Figure 4.2: Plots of clay, silt, and sand size sediment distributions in Lake Uberaba. A)

Clays are distributed near to fluvial inputs, since these confined channels transport most

clays into the lake. B) Silts are the most common grain size accumulated in areas draining

from relict lobes, or from distal fan drainage due to the hydrodynamics of those channels.

C) Sand was largely concentrated near the eastern lake margin due to input from the

active depositional lobe. Plus (+) symbols represent sampling sites.

Page 71: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

59

The ~10 kHz seismic profile shows distinct differences in shallow acoustic facies

(Fig. 4.3), with evidence for at least one decameter-wide, filled trough in an area that is

today submerged beneath the lake (Fig. 2.2). Two seismic stratigraphic units were

identified in these acoustic data. One unit is the sandy silt lacustrine facies, represented

by the reflector along the lake floor. The second unit is the dense clayey silt facies

composing the relict basin floor prior to lake formation. The depth of penetration of the

seismic profile was limited by the frequency of the source, the water column depth (e.g.,

Alaamer, 2015), and substrate type. However, the acoustic facies variability suggests

river channels incised into the basin floor when Lake Uberaba was at lowstand.

Page 72: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

60

Figure 4.3: Classification of seismic characteristics resulted in three categories. Sediment fill shows no multiples, or subsurface

reflectors. Paleo-channel (outlined in Fig. 3.8) character shows clear subsurface reflectors but no multiples. Dense sediment fill

exhibits no visible sedimentary structures and has ~2-5 multiples (shown in Fig. 3.8). The southern transect contained more frequent

paleo-channels than in the northern transect. Red arrows indicate the lake floor.

Page 73: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

61

Taken together, the bathymetric and seismic data help to demonstrate that a recent

change in depositional environment, from fluvial to lacustrine, has occurred along the

eastern toe of the Paraguay megafan (Fig. 4.1). Along Lake Uberaba’s northern margin,

numerous small, distributary channels from the Paraguay megafan form a vast, swampy,

sedimentary mosaic, where fine sands are being deposited along with peaty muds. Peat-

rich sediments form in interdistributary environments on the distal fan when vegetation

growth and senescence rates are high. In a recent study, Lo et al. (2017) found

exceptionally high rates of aquatic macrophyte growth (based on satellite and field

images) on the northern Uberaba shore. The decay of these macrophytes fundamentally

influences the size and shape of Lake Uberaba, as successive generations of vegetation

grow over decaying macrophytes. The available evidence suggests that this lake has been

shrinking over the past few years due to these processes. Together, these insights show

that Lake Uberaba functions much differently than many of the other large Pantanal

floodplain lakes, which experience seasonal cycles of direct connectivity with the

Paraguay River (McGlue et al., 2011). Lakes that generate geological records (i.e., those

that produce lacustrine deposits) are frequently characterized based on the interaction of

water level with the basin outlet (Carroll and Bohacs, 1999; Bohacs et al., 2000). The

Carroll and Bohacs (1999) conceptual model shows that lake types can be placed into

three categories, based on the interactions among available accommodation, water inputs,

and sediment inputs: (i) balanced fill, (ii) underfilled, and (iii) overfilled. Briefly,

balanced fill lakes occur where hydrological inputs are approximately equal to

accommodation (i.e., the lake is weakly open), underfilled lakes occur where

accommodation greatly exceeds water and sediment input (typically associated with

heavy evaporation in a hydrologically closed basin), and overfilled lakes are persistently

hydrologically open, where hydro-sedimentary inputs greatly exceed basin

accommodation. Importantly, the nature of the basin type has implications for predicting

the facies that are produced from each lake depositional system. Although the complexity

of modern environments can make it a challenge to adequately characterize extant lakes

in this tripartite system (e.g., Bohacs et al., 2003), we interpret Lake Uberaba to be an

excellent example of a shallow, overfilled lake basin. Carroll and Bohacs (1999)

demonstrate that overfilled lake basins throughout geologic time accumulate fluvial-

Page 74: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

62

lacustrine facies associations. The composition of modern facies accumulating in Lake

Uberaba strongly reflects the shallow, hydrologically open lake. For example, lake floor

sediments are dominantly allochthonous quartz-rich silt with low BiSi and TOC

concentrations, particularly offshore. Lake floor sediments with elevated (> 2%) clay size

content are restricted to some areas of the shoreline, where dense carpets of aquatic

vegetation are interpreted to help trap fine-grained particles, and disperse wave energy.

The %clay size contour map (Fig. 4.2) illustrates this relationship, whereas the %silt and

%sand contour maps show the overall dominance of coarser sediments in the center of

the lake. We interpret that the texture and composition of these deposits are strongly

conditioned by inflowing megafan rivers.

Modern Lake Uberaba sediments are brown and massively bedded, which are

characteristics consistent with lake floor oxygen levels capable of supporting benthic

organisms that rework the sediment, such as the thumbnail Unionidae clams (Cohen,

2003) recovered in our survey. Shallow sediment mixing by wave energy is common at

Lake Uberaba, shown by suspended sediment concentrations up to ~43.2 mg l-1 in some

locations in the basin, as well as poor preservation of OM, reflected in low TOC

concentrations. Notably, sites with higher TOC concentrations were routinely found in

nearshore environments, where aquatic macrophyte densities are high (Lo et al., 2017).

Evidence for sediment mixing was also observed in the radionuclides with large increases

in SMAR and LAR in Core 2 (Fig. 3.17). Resuspension also inhibits sedimentary

structure preservation, which is manifest in the absence of laminations and other fine

structures in the cores. Sponge spicules and diatoms are the primary sources of biogenic

silica in modern sediments at Lake Uberaba. Lake Uberaba’s offshore sediments contrast

greatly with those of Lake Gaíva (in some publications known as Laguna la Gaiba;

Whitney et al., 2011; Metcalfe et al., 2014). Lake Gaíva is connected to the Paraguay

River by a tie channel (Rowland et al., 2009) and undergoes a seasonal water level cycle

associated with passage of the Paraguay River flood pulse (McGlue et al., 2011). Lake

Gaíva is important because it is connected to Lake Uberaba downstream, through Canal

Dom Pedro II. Although also a shallow (Zmax = 6 m), exorheic lake, the offshore

sediments of Lake Gaíva are green, OM-rich, silty clays with abundant biogenic silica.

We suggest that margin-coincident topography plays an important role as a barrier to

Page 75: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

63

strong fluvial coupling with the shoreline at Lake Gaíva, allowing fine detrital and

biochemical sediments to settle on the basin floor. Suspension settling is interpreted to be

far less common in Lake Uberaba, due to the enhanced influence of fluvial discharge,

high channel density, and wave reworking of lake floor sediments.

Overfilled lakes may be shallow or deep, and their persistence on a landscape

through time can vary, depending in part on how the lakes were formed. A high rate of

accommodation space generation may favor a lake with a long lifespan. For example,

deep, overfilled lakes with high relative volumes of accommodation space tend to be

long-lived features with tectonic origins, such as in broken (thick-skinned) foreland, or

rift settings. In the western continental U.S., the well-studied Eocene Green River Fm.

contains foreland basin overfilled lake deposits, represented by the Laney and Tipton

Shale Members of the Greater Green River and Washakie Basins (Wyoming, U.S.). The

open lakes that produced these fluvial-lacustrine strata lasted a few million years, and

were marked by relatively stable shorelines (Johnson, 2003; Pietras et al., 2003; Murphy

et al., 2014). In the East Africa rift valley, Lake Edward (Uganda/Democratic Republic of

the Congo) is a deep (Zmax ~120 m), overfilled lake basin that is hydrologically open,

draining into Lake Albert through the Semliki River (Russell et al., 2003; McGlue et al.,

2006). Lakes may transition from one basin type to another due to variability in sediment

supply, water availability (often from climate change), catchment evolution, and

subsidence or uplift. For example, stratigraphic evidence from both Lakes Albert and

Malawi (East Africa) indicate that those basins were overfilled at different times in the

late Cenozoic (Karp et al., 2012; Ivory et al., 2016). Shallow overfilled lakes are more

common in settings where accommodation space is limited or rapidly filled with

sediment, such as in some fluvial or glacial environments (Carrivick and Tweed, 2013).

Lake Gaíva, just to the south of Lake Uberaba, is hydrologically connected to the

Paraguay River, and is directly influenced by its inflows and outflows (McGlue et al.,

2011). Allochthonous sediment transport to Lake Gaíva has been more limited than in

Lake Uberaba, but limited accommodation and continuous inflows maintain overfilled

conditions there.

Why is Lake Uberaba an overfilled lake? Such lakes form where sediment loads

plus water supply greatly exceed the available accommodation, forcing water to exit the

Page 76: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

64

system through an outlet channel. Compared with the drier southern Pantanal (e.g.,

Almeida et al., 2011), the northern Pantanal experiences less evaporation and greater

hydrologic connectivity. An average monthly rainfall of 157.5 mm yr-1 in the austral

summer (October-March) is the primary control on water availability to the Paraguay

River megafan (Bergier and Resende, 2010; Novello et al., 2016). In addition, the flows

between Lake Uberaba and the Paraguay River are dominantly outflows to the Paraguay

River. Lake Uberaba sits in a dynamic location along the toe of the Paraguay megafan,

north of high relative topography produced by the Neoproterozoic Serra do Amolar

mountain range (Fig. 4.4). The best explanation for limited accommodation in the region

of Lake Uberaba is active fluvial sedimentation associated with the progradation and

lateral migration of the distal Paraguay megafan. This megafan initially formed a

depositional lobe towards the western margins of the Pantanal River basin. As sediments

prograded into and filled available accommodation, the Paraguay River shifted and

progressively built new depositional lobes eastward, resulting in the present teardrop-

shaped megafan landform (Assine and Silva, 2009). Note that a chain of swampy lakes

and wetlands exists along the toe of the Paraguay megafan, and includes (from west to

east): Lakes Orion, Piranhas, and Uberaba (Fig. 1.4A) (McGlue et al., 2011). Our

analysis shows that surface area may fluctuate dramatically in these lakes, but that Lake

Uberaba maintains the largest open water area along the toe of the megafan, signifying

the area with greatest sediment accommodation.

Page 77: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

65

Figure 4.4: The Neoproterozoic Serra do Amolar mountain range rises majestically south

of Lake Uberaba, along the western Pantanal margins. The contrasting relief illustrates

the vast accommodation available in the Pantanal tectonic depression.

Page 78: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

66

Thus, we interpret that the geometry of Lake Uberaba was directly shaped by

sedimentation, reducing accommodation from west to east, and from north to south.

Accommodation can also be produced vertically through subsidence or erosive processes.

Another plausible mechanism that may modify available accommodation at the toe of the

megafan is neotectonic deformation. The Pantanal is a tectonically active basin

characterized by subsiding blocks, fault movements, and even localized earthquakes (e.g.,

Ussami et al., 1999; Warren et al., 2015). For example, satellite images show that a

number of river courses in the Pantanal basin have been deflected by nearly 90˚, or had

channel planforms altered by recent faulting (Assine et al., 2015a). In addition,

bathymetric patterns, such as deepening adjacent to lake margins with pronounced relief

in Lakes Gaíva and Mandioré, are consistent with a role for tectonic processes on

limnology (McGlue et al., 2011). We cannot discount that a combination of relief,

neotectonics, and erosion-resistant lithologies (e.g., metaconglomerates of the Cerradinho

and Caiuéus Formations in the Corumbá Group) (Assine, 2003; Warren et al., 2015) may

have helped to shape the available accommodation within which Lake Uberaba formed.

These factors are also likely responsible for the chain of swampy lakes along the

Paraguay megafan fringe, abutting against the Neoproterozoic mountain range. Based on

our observations of turbidity, bathymetry, and nearshore environments, the age of Lake

Uberaba is considerably < 1,000 years old, which is further constrained by the

geochronology of the sediment cores (see below). High turbidity suggests rapid

depositional rates, which reduces the lifespan of the lake by quickly filling available

accommodation (e.g., Tye and Coleman, 1989; Bohacs et al., 2003). Typically, floodplain

lakes persist on landscapes for a snapshot of geologic time (thousands to hundreds of

thousands of years) (Cohen, 2003). Overall, we interpret that Lake Uberaba maintains

overfilled conditions because of channel density along its margins; numerous inlets make

hydrological closure difficult, as sedimentary inputs fill available accommodation and

produce a low gradient lake floor. Ínsua Island (Fig. 3.2), which presents a partial buffer

to the transport of Paraguay River sediments to Lake Uberaba, also contribute to a

productive, hydrologically open water body (Wetzel, 2001). Good hydrologic

connectivity is maintained when sedimentation is reduced, and the lake boundary is in

contact with many outlets.

Page 79: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

67

Satellite images from 1968 (Fig. 3.2) contrast greatly with much more recent

Landsat images, which shows that lake levels and basin floor morphology are dynamic at

Lake Uberaba. In the 1968 satellite image, lowstand Lake Uberaba remains

hydrologically open, with a primary inflow from the Corixo Grande River and outflow

through Canal Dom Pedro II. At that time, however, the Canzi River mostly bypasses the

lake and flows directly into the Paraguay River. During this drier period, the lake

contracted, resulting in subaerial exposure of the shallowest parts of the basin along the

northern and western margins. This environment was likely responsible for producing the

short-echo acoustic facies observed along several areas of the 10 kHz seismic line (Fig.

4.3). The 1968 lake was irregularly shaped and embayed, which contributed to the

formation of this acoustic facies. By 1979, lake level was high and allowed for multiple

lake inlets and outlets.

Adjacent to Lake Uberaba, several different depositional environments influence

nearshore and offshore sedimentation, including the relict depositional fan lobe along the

northwestern shoreline, the active fan lobe along the eastern shoreline, a meandering

channel belt (Canal Dom Pedro II), and the toe-of-fan drainage that enters the lake from

the west, the Corixo Grande River (Fig. 4.1). The relict depositional fan is reworked by

rainfall annually, and it provides nutrients and sediments in ephemeral channels that

sustain high rates of macrophyte productivity on Uberaba’s northern shoreline. Satellite

image analysis of the lake from 1985 to 2015 showed that macrophyte growth helped

reduce the lake surface area by ~21% (Lo et al., 2017). The active fan lobe provides

direct fluvial input from the Canzi River. Nutrient delivery dynamics and sedimentary

environmental conditions, such as rapid flows on the eastern lake margins, seem less

amenable to sustaining dense carpets of floating macrophytes. The bioavailable element

distribution (Fig. 3.7) suggests that most bioavailable elements are concentrated at the

lateral river inputs on the eastern and western lake margins. Given the high rates of

deposition (0.28 cm yr-1) in north Lake Uberaba, we suspect that dilution, in addition to

the growth of macrophytes, reduces elemental abundances along the northern lake

margins. The macrophyte density and diluted elemental weight percent indicate that

nutrient availability in shallow zones is greatly modulated by vegetation growth. Shallow

areas with sufficient light penetration are needed for macrophytes to colonize (Middelboe

Page 80: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

68

and Markager, 1997). The active meandering channel belt receives lake outflows, and in

turn transports water and sediment downstream. The Corixo Grande River, which drains

the toe of the megafan, appears to supply mainly silt and clay to western Lake Uberaba,

most likely due to up-dip storage of coarser sediments in and more variable flow to Lakes

Orion and Piranhas. Although nutrients are abundant on the western lake margins, the

reduced sediment input there hinders macrophyte progradation, because depth becomes

the main limiting factor. Similarly, rivers of higher gradient can experience hydraulic

eddies, which can lead to upstream sediment trapping (Mueller et al., 2014). In the supra-

littoral environment, highly productive marshes were encountered, especially along the

northern lakeshore. Silt and sand contour maps (Fig. 4.2) show that some distributaries

from the Paraguay megafan impact sedimentation in Lake Uberaba, and form small,

subaqueous fan deltas in the nearshore environment. An open lacustrine environment lies

farther offshore, where deposits are siliciclastic, oxidized, contain benthos (clams), and

lack bedding due to resuspension from water column mixing.

Rivers entering and exiting Lake Uberaba have an important influence over

patterns of biogenic sediment accumulation. The highest TOC concentrations in Lake

Uberaba were found along the lake margins, where the Corixo Grande and the Canzi

Rivers enter the basin from the west and east, respectively. We interpret that preservation

of autochthonous and potentially allochthonous particulate OM in these settings is higher,

due to reduced wave action. We expected TOC concentrations to be high on the northern

shoreline, where macrophytes are most dense, but the low values we measured suggest

that either preservation or dilution dynamics strongly influence organic facies

development here. It is clear that OM preservation is limited in this lake, due to

polymixis and abundant lake floor oxygen, while numerous point-sources of detrital

sediments are an effective dilutant (e.g., Rowland et al., 2009; Citterio and Piégay, 2009;

Vega et al., 2014). The distribution of TOC shows that the Canzi River is most likely the

main source of allochthonous OM to the lake. Organic matter transport is frequently

associated with sediment load (e.g., Boix-Fayos et al., 2015; Shynu et al., 2015). Despite

high OM and sediment delivery from the Canzi River, the eastern lake margins are too

energetic for macrophyte growth. Autochthonous OM production may be limited by in

situ turbulence, which increases turbidity and reduces light penetration, limiting

Page 81: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

69

photosynthesis. The concentration of suspended sediments in the 2015 satellite image

(Fig. 4.5) attests to this possibility (e.g., Min et al., 2012).

Figure 4.5: In this Landsat-8 satellite image (RGB-654) from September 15, 2015,

lighter-colored suspended sediment plumes are observed in Lake Uberaba. Plumes are

stronger in the northwest and southeast, which shows that the aquatic macrophytes help

trap particles (Zhu et al., 2015), and increase sediment stability.

Page 82: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

70

4.2 Core Stratigraphy

Sediment cores from Lake Uberaba contain two dominant lithofacies that capture

evidence of a significant aquatic transition in the northern Pantanal during the latest

Quaternary. In several of our longest cores, the basal lithofacies was a massively bedded,

tan clayey silt that we interpret to reflect deposition on the floodplain of a meandering

river. Massive bedding, filled desiccation cracks (Appendix D), and oxidized grains point

towards a depositional environment that was at least periodically exposed. This

interpretation is corroborated by satellite images, which show that the lake’s morphology

changes considerably during lowstands, exposing broad areas that are today submerged.

Similar facies have been described from former floodplains in western Colorado, where

sand-filled mud cracks were preserved in a siltstone matrix (Dubiel et al., 1992). The

upper lithofacies, typified by relatively organic-rich, massive, brown sandy silt reflects

the recent lake level highstand following transgression of a highly variable lake across

the floodplain. The contact between these two facies is sharp, irregular, and interpreted to

be unconformable. A number of the Pantanal’s large floodplain lakes have incomplete

stratigraphic records, due to drought-forced regressions when SASM strength waned in

the early Holocene (e.g., Bird et al., 2011; McGlue et al., 2012; Novello et al., 2017). For

example, the southern end of Lake Mandioré was desiccated in the early-middle

Holocene, leaving a clay pan where the Paraguay River formerly entered the lake. The

basal stratigraphic unit at Lake Uberaba shares some similarities with the desiccated

deposit at Lake Mandioré (McGlue et al., 2012). Both exhibit a massive, tan clayey

lithology, and both were influenced by an abrupt aquatic transition. The lithologies in

Lake Uberaba cores can be related to environment of deposition based on grain size. The

dominantly sandy silt in Lake Uberaba currently represents neither slow sedimentation

(clayey silt) nor direct channel deposition (silty sand).

We interpret that the expression of shallow stratigraphy, which in part mirrors

patterns of modern sedimentation in Lake Uberaba, depends on shoreline proximity and

local depositional processes. General evidence indicates that the character of fine-grained

lake deposits varies widely depending on physical, biological, and chemical limnological

processes, the geological surroundings of the basin, and hydroclimate (e.g., Gustavson,

1991; Rodbell et al., 1999; Cohen, 2003). For example, some lakes preserve delicate,

Page 83: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

71

mm-scale biochemical laminations in the sediments of their sub-littoral and profundal

environments. These types of deposits are most frequently produced in deep, strongly

stratified lake basins with anoxic bottoms, or those characterized by rapid sediment

accumulation rates. Such deposits are prized for the high-resolution environmental

signals they record, and they have been used to great effect in the highlands of tropical

South America (Abbott et al., 1997; Fritz et al., 2001; Sylvestre, 2002; Hillyer et al.,

2009; Fornace et al., 2014). In contrast, massively bedded muds are more common in

shallow, well-mixed lakes that experience cycles of deposition, subaerial exposure, and

reworking (e.g., Gustavson, 1991; Oviatt et al., 1999). These types of deposits can

contain useful environmental signals, albeit with lower temporal resolution due to time-

averaging. Many of South America’s lowland floodplain lakes accumulate muds of this

kind (e.g., Stevaux, 2000; Cohen et al., 2015). Our most complete stratal records, from

cores 3 and 4, bear evidence of this massive texture, and are consistent with sediment

mixing, as well as possible reworking and oxidation during subaerial exposure (Larsen

and MacDonald, 1993; Shanley and McCabe, 1994). As lake levels fell and rose in the

recent past, some areas of the lake floor were subjected to one or more cycles of

desiccation and rehydration. This process helps to explain the very low TOC

concentrations in offshore deposits, the fragmentation observed in siliceous microfossils

(diatoms and spicules) on smear slides, and the “hard” acoustic character of certain

reaches found along the ~10 kHz seismic line. Both cores 3 and 4 contain a bimodal,

shallow stratigraphy with a massive, oxidized basal unit we interpret as the more ancient

floodplain environment overlain by deposits signifying a transgression into a shallow,

well-mixed lake environment, which is significantly influenced by fluvial processes.

Facies models for meandering rivers are typified by thick floodplain mud complexes

overlying laterally accreted channel sands, whereas braided streams are usually defined

stratigraphically by a paucity of floodplain facies and coarse, central channel bars and

lags. The amount of fine-grained sediment (silt + clay) in this basal lithofacies suggests a

meandering river floodplain, which would have been susceptible to reworking from

channel migration (Bridge, 1993). Thus, a highly-resolved depositional history derived

from radiocarbon dating is probably not achievable from the floodplain facies. The cause

of the transgression that led to the formation of Lake Uberaba is likely related to a

Page 84: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

72

transition to a wetter climate in the late Holocene, which is known to have influenced

tropical Brazil (e.g., Garcia et al., 2004; Enters et al., 2010; Novello et al., 2012). Grain

size analysis of the lacustrine facies in both cores 3 and 4 shows that the open lake

deposits contain coarse silt and fine sand, which suggest the potential for high relative

stream power in the lateral (E-W) rivers filling the lake (Chen et al., 2004; Peng et al.,

2005; Macedo et al., 2017). Today, deposition of the coarsest siliciclastic sediments

occurs on the eastern half of the lake, which indicates that the Canzi River has become an

important influence on the composition and texture of the preserved lacustrine facies.

Our best estimate of lake formation centers on the late 18th century. We used 14C

dates derived from fresh-looking charcoal in Core 3 to provide age constraints (e.g., Geyh

et al., 1997); these dates are stratigraphically out of order, but fall within error of one

another (228 ± 72, 139 ± 101, 206 ± 120 cal yr BP). During this period, stalagmite

records show a 1792-1833 BCE wet event in the Brazilian Nordeste (Northeast) region

(Novello et al., 2012). One scenario explaining this hydroclimatic change is that the ITCZ

increased in amplitude southwards, and consequently, brought more precipitation to the

Pantanal. This interpretation is consistent with SASM rainfall being directly related to the

ITCZ (Vuille et al., 2003). The surge in humidity likely increased vegetation growth,

which may have helped to stabilize shoreline sediments. The encroaching Paraguay

megafan lobe, coupled with the topographic barrier produced by the eastern Bolivian

landscape helped create the shallow lake basin. The coarse sediment at the lithofacies

contact in Core 4 suggests that the lateral input from the Corixo Grande River provided

the largest input, enabling lake transgression to occur.

Although Core 3 exhibits an erosional unconformity at 23 cm b.l.f. (Fig. 3.9), the

lithofacies contact suggests that the transition from subaerial floodplain to lake increased

OM input without delivering coarser sediment. As lake levels rose, clay-sized detritus (as

well as its chemical proxy, %Al), and BiSi concentrations declined, as sand content and

OM increased. We interpret this pattern as another effect of siliciclastic dilution.

Although particulate OM and lithogenous sediment provenance data are presently

unavailable, we predict that a combination of shoreline macrophytes and allochthonous

OM is the most likely source of the higher TOC values. At 10 cm b.l.f. in Core 3, we

Page 85: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

73

interpret a change in sediment delivery from floodplain sediments to Canzi River

deposits, when sand surpassed silt as the dominant (70-80%) grain size.

Stratigraphic completeness varies across the lake (see Fig. 4.6). In core 1, the

presence of primarily floodplain facies indicates that little modern sediments are

accumulating in northwestern Lake Uberaba. A thin, fluvially deposited basal sand bed

(Unit 3) was deposited prior to Unit 2 deposition. Until the lake formed, the subaerial

floodplain persisted, with occasional fluvial sand input since at least 21,155 ± 1,445 years

based on the OSL measurement near the bottom of Core 1. In contrast, Core 2 contains

the most continuous record of modern lacustrine nearshore sedimentation. The aggressive

floral growth and faunal activity in this nearshore environment indicate that ecological

agents act to colonize submerged areas. The basal, clayey silts below 38 cm b.l.f. were

likely part of a greater wetland region with high aquatic macrophyte productivity. Soft

peaty silts occupy the 18-38 cm b.l.f. section, which were probably deposited after the

lake formed. The aquatic vegetation that could not adapt to the increased water depth

perished, and rapidly accumulated in a thick peaty silt (Bergier et al., 2012). This peat-

rich layer is now capped by distributary sands delivered from the encroaching Paraguay

River megafan. Stratigraphic completeness in the last century is indicated by

exponentially increasing SMARs displayed in Core 3, consistent with the recent increase

in sedimentation after the lake level increased in the 1970s. The increase in SMAR can

likely be attributed to the encroaching Paraguay megafan in northern Lake Uberaba as

well as the increasingly active Canzi River flow into eastern Lake Uberaba. The

prolonged Pantanal drought in the 1960s was not well captured in these two cores.

However, the record in Core 2 shows nearly tripled SMAR after the drought (from 0.13

to 0.31 g cm-2 yr-1). The elevated SMAR (~1.16 g cm-2 yr-1) was consistent with the

increased accumulation of organic material, forming a peaty layer approximately 34-42

years before sampling. Thus, at the site of Core 2, conditions such as water depth,

nutrient, flow, and water chemistry were suitable to host this drastically increased

vegetation growth.

Page 86: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

74

Figure 4.6: Summarized lithofacies and how the cores relate to each other spatially. The

onset of the lake is denoted by the oscillating correlation line, whereas the onset of

floodplain facies in Core 1 is marked by a dotted line and question mark. Core 2 was

considered distinct from the other cores and not included due to the presence of a peat-

rich Unit II and higher average TOC values.

Page 87: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

75

Our OSL age date is consistent with the timing of a large sandy depositional event

during the Last Glacial Maximum (23-19 ka). Ages from this period are present at the

base of paleochannel cores from the adjacent São Lourenço megafan, yielding the

following ages: 19.8 ± 3.2 ka, 20.6 ± 2.1 ka, and 21.1 ± 3.5 ka (Pupim et al., 2017). The

ancient depositional lobe of the Paraguay megafan approximately formed during the late

Pleistocene (Assine and Silva, 2009). The sand-rich facies of the ancient lobe reinforces

that the OSL age date obtained in this study likely reflects nearly the end of active lobe

deposition prior to lobe abandonment.

4.3 Distributive Fluvial Systems (DFS) and their distal environments

A number of recent studies have argued that DFS are some of the most important

depositional environments that build the continental rock record (e.g., Leier et al., 2005;

Hartley et al., 2010; Weissmann et al., 2010; Chakraborty and Ghosh, 2010; Latrubesse et

al., 2012; Assine et al., 2014; Cohen et al., 2015). Distributive fluvial systems are

increasingly recognized for their abundance on Earth’s surface, and their importance

continues to grow (Fielding et al., 2012). A DFS is characterized by increasing numbers

of small channels downstream, channel shoaling and narrowing downstream, and

decreasing flow downstream, which may result in unconfined flows spreading out across

the toe of the fan (Nichols and Fisher, 2007). A DFS is commonly characterized by three

different zones: proximal (fan apex), medial, and distal (toe). In the proximal zone, flow

is strongly channelized. Overbank deposits are rarely preserved due to intense reworking

by the migrating, often braided channels. Proximal DFS deposits in the rock record

usually consist of sandy channel deposits that are interconnected and amalgamated

(Nichols and Fisher, 2007). In the medial zone, abandoned river tracts, floodplain splays,

and avulsions are common. The medial DFS environment begins to show signs of

decreasing channel width and sediment grain size. Overbank deposits are more common

such that, in a cross section, the channel deposits can be distinguished easily from

floodplain mud and sand deposits. Several distinct depositional environments can occur

on a distal fan, where the toe of the fan encounters a transitional setting. The distal facies

host a high volume of overbank floodplain facies with few large, well-preserved channel

deposits. Many channel deposits occur as thin sheets that successively stack over time.

Page 88: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

76

These deposits may be interbedded with finer deposits of mud and silt, with potential for

paleosol development (Nichols and Fisher, 2007). Distal fluvial fans can experience

varied environments, from a channelized terminal lobe, to unvegetated, unconfined

channels that result in sheetflow deposits (Weissmann et al., 2015). Fundamentally, the

distal DFS contain valuable stratigraphic lithofacies, which can preserve evidence of

terminal lobe deposits, or changes in lacustrine basin geometry (Sáez et al., 2007). Lakes

occupy these distal zones, and vary widely in morphology and function (Gierlowski-

Kordesch et al., 2013).

Several South American examples of large DFS (fluvial megafans; Leier et al.,

2005) highlight the diversity of these landforms, particularly with respect to how they

terminate (Table 4.1) (Weissmann et al., 2010; Wilkinson, 2016). Exorheic drainage and

high precipitation/erosion (P/E) result in geomorphic forms such as wetlands, lakes, and

axial fluvial systems (Davidson et al., 2013). In one style, a dominant, channel-confined

river directly connects to a larger axial river. The Bermejo River (Argentina) runs east

from the Andes Mountains (140 m.a.s.l.), across the Chaco foreland, and terminates into

the southward-flowing, axial Paraguay River (60 m.a.s.l.). The distal DFS environment is

dominated by the tightly meandering Bermejo River, several Bermejo-parallel spring-fed

streams that also terminate into the Paraguay River, and numerous small swampy lakes

and flooded soils (McGlue et al., 2016). Similarly, the Taquari River has produced one of

the largest, most visible megafans in the Pantanal. It terminates with a dominant channel

flowing directly into the Paraguay River. In another termination style, the distal DFS

zone is marked by one or more fan splays. Adjacent to the Bermejo megafan, the

Pilcomayo River flows along the Argentina-Paraguay border, and has formed the world’s

largest DFS (Weissmann et al., 2010). The Pilcomayo River flows downstream into

increasingly smaller channels until these terminate in splays with abundant wetlands

(Cohen et al., 2015; Weissmann et al., 2015).

Still other DFS have been observed to terminate into hydrologically closed lakes

known as playas. Hartley et al. (2010) described DFS forming the margins of underfilled

playas (Iran), or into permanent lakes at the Ili River in Kazakhstan. DFS termination in a

body of water is not common; only ~10% of DFS terminate in playa lakes, and only ~8%

of DFS terminate in permanent lakes (Hartley et al., 2010).

Page 89: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

77

Table 4.1: Summary of different DFS termination styles. Marine terminations are

sometimes considered a type of DFS termination (Davidson et al., 2013). However, these

are not included in this study because marine terminations terminate in subaqueous

environments, subjected to unique processes (e.g., turbidity currents) not found in

subaerial environments.

DFS termination style Facies References

Channel-levee complex Dense, organic-rich soils McGlue et al., 2016

Fluvial splays Sandstone, siltstone,

massive conglomerate

Opluštil et al., 2005;

Weissmann et al., 2011

Playa lake Silty clays and diatom ooze

Harvey et al., 1999;

Arzani, 2012; Ribes et al.,

2015

The Paraguay megafan is unusual because much of the water flowing off of this

DFS passes through Lake Uberaba. In the lacustrine termination style presented by

Hartley et al. (2010), playas and permanent lakes form a continuous entity along most of

the distal fan fringe. In contrast, Lake Uberaba does not span the distal fringe, yet it

captures most of the axial and lateral flows of the Paraguay megafan. The presence of

Lake Uberaba creates a unique proximal-to-distal transition from distributary channels, to

wetland, to lake. In addition, a chain of swampy lakes and wetlands, linked by the

eastward flowing Corixo Grande River, form the distal toe of the Paraguay megafan. This

configuration is a previously undocumented type of megafan termination style.

The northern margins of Lake Uberaba have extensive peat-rich sediments, and

conditions there strongly point towards a high potential for organic soil formation (e.g.,

Johnston et al., 2009; Hartley et al., 2013; Davidson et al., 2013). Peats are a type of

organic soil, which by definition must be > 10 cm thick, and TOC concentrations must be

> 12% by weight if no clay is present (Osaki et al., 2016). These criteria are partly

satisfied in Core 2, where the peat-rich layer is 20 cm thick, and although clay is a minor

component (avg. 21%), TOC concentrations averaged nearly 12% (11.8%). Hence, the

potential for paleosol formation and preservation is interpreted to be high, given that

Page 90: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

78

sands (0-18 cm b.l.f.) firmly cap the peaty sediment. The presence of peat capped by

fluvial sands may be characteristic of the distal termination at the Paraguay River

megafan. Exceptionally dense macrophyte roots and leaves indicate rapid burial by

channel sands prior to more extensive decomposition.

We determined that Lake Uberaba contains a complex record of environmental

and landscape change due to its hydroclimate filter, which is strongly controlled by

fluvial processes. The indicators for environmental change are shaped by modern

processes including water chemistry, vegetation growth, hydrodynamics, and

sedimentation. The lower-resolution sedimentary records are valuable for identifying

large-scale environmental changes and minimizing background noise from the

fluctuations of the indicators. The recorded changes from the short cores and recent lake

formation suggest that the Pantanal wetlands change rapidly (over the course of years or

decades). Major changes resulting in regime shifts are strongly linked to the hydrologic

cycle.

Page 91: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

79

CHAPTER FIVE: CONCLUSIONS

Lake Uberaba is a dynamic fluvial-lacustrine system that experiences major

fluctuations in water levels and hydrologic inputs unparalleled among lakes that have

been studied in the Pantanal. Several vintages of satellite imagery attest to substantial

changes in lake surface area and fluvial connectivity over the past ~50 years. Lake

Uberaba’s eastern margin is shaped principally by the inflowing Canzi River, whereas the

Corixo Grande River controls western lake margin shape and stability. Distributary

channels from the Paraguay River megafan help define the morphology and supra-littoral

depositional environments of the northern lake shoreline. The lake’s bathymetry is

shallow (Zmax = ~3.5 m), and influenced by sediment inputs from the megafan and lateral

rivers. The lake deepens to the south near the outflowing Canal Dom Pedro II, and

isobaths indicate the presence of at least one sublacustrine channel oriented NE-SW,

originating from the eastern lake margin. The lake waters are slightly alkaline (mean pH

= 7.17), with low salinity, but are high in dissolved oxygen, TDS, and suspended

sediments, which we interpret to reflect frequent mixing of the water column by wind and

waves. Lake waters are a dilute Ca2+-Na+-HCO3- brine type.

Riverine inputs and regional base level modulate patterns of modern siliciclastic

sedimentation in Lake Uberaba. Lake floor sediments are primarily silty detritus, with

sands concentrated near the eastern edge of the lake. This distribution attests to the

influence of the active channel of the Canzi River. Inflowing rivers and their sediment

loads control the dilution and preservation of OM in modern Lake Uberaba sediments.

Areas of the lake adjacent to megafan channels are sandy, with a paucity of biogenic

components, whereas areas farther from fluvial point sources contained finer sediments,

with higher concentrations of BiSi and TOC. Most BiSi is composed of sponge spicules,

which is consistent with the modern depositional setting. Preservation of organic carbon

in the offshore environment is low, likely due to intense resuspension and mixing in the

wave-swept polymictic lake, coupled with localized reworking from bioturbating fauna.

Distribution of bioavailable elements shows that intense macrophyte growth depletes

nutrients, which is consistent with prior research on the importance of macrophytes for

shaping Lake Uberaba’s margins (Souza et al., 2011; Lo et al., 2017).

Page 92: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

80

Sediment cores collected from the deepest areas of Lake Uberaba are defined by

two lithostratigraphic units. The lower Unit 2 consisted of massive tan, clayey silts,

whereas the upper Unit 3 typically contained one or two brown, massive, sandy silt

facies, with varying amounts of organic carbon. The thicknesses of these lithofacies

varied depending on proximity to fluvial inputs and water depth, which appear to

influence sediment accumulation rates. We interpret these lithofacies to reflect two

distinct depositional environments: an older, subaerially exposed and oxidized floodplain,

and a younger hydrologically open shallow lake. The floodplain facies contained low

TOC and high BiSi concentrations, which supports periodic (rare?) inundation by flood

waters; evidence of fluvial channel incision from seismic profiling supports this

interpretation. Developing the age control necessary to pinpoint the timing of this

environmental transition was a challenge, due to resuspension and reworking of shallow

sediments in the modern lake. However, our best 14C dates point to inception of a lake

phase at ~1760 CE. At this time, the SASM briefly strengthened and brought with it

greater precipitation (e.g., Bird et al., 2011; Novello et al., 2012; Vuille et al., 2012;

Kanner et al., 2013). Our OSL date indicates that the Unit 1 basal sands of Core 1 are

thousands of years older, at 21,155 ± 1,445 years.

Global observations of modern fluvial megafans suggest that the distal zones of

these large depositional systems vary considerably. Common styles of megafan

terminations include: (a) deposition into a hydrologically closed (endorheic) lake; (b)

deposition into a hydrologically open, or series of open lakes; (c) deposition into one or

more distributary fans; and (d) connectivity between the trunk megafan stream and an

axial river. The distal fringe of the Paraguay River megafan is occupied by a chain of

paludal, hydrologically open lakes, of which Lake Uberaba is the largest, and is located

downstream of megafan distributary channels. This setting establishes a characteristic

suite of facies that could be preserved in the rock record. Several South American

megafans (e.g., Pilcomayo and Bermejo megafans of Argentina) form as sediment erodes

from the Andes Mountains, to low elevation back-bulge depocenters. However, the distal

Paraguay River megafan formed entirely in a back-bulge setting, as a consequence of

variably subsiding blocks along faults related to Andean orogenic stresses. The gradual

progradation of the fan and infilling of the lake’s accommodation is consistent with

Page 93: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

81

fluvial megafan evolution. The lake is part of a greater system of lakes, and lies at the end

of this chain of lakes. Lake Uberaba remains a unique termination style among

hydrologically open lakes, since most waters from the megafan distributary channels

converge on Lake Uberaba prior to exiting the megafan system.

Future studies could perhaps extend the record with longer cores and cores from

different locations in the basin, to provide a better spatial perspective on sedimentation.

Additional field campaigns to Lakes Orion and Piranhas, also along the fringe of the

Paraguay megafan, would be beneficial to ensuring a more accurate facies model for the

Paraguay DFS.

Page 94: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

82

APPENDICES

Appendix A: Water Column Measurements

Water samples were first titrated to pH 4, then further lowered in increments of pH 0.1.

The amount of HCl (ml) needed to reach the desired pH is indicated in the green cells.

Conversion from μeq l-1 to mg l-1 by a factor of 16.387 was referenced from the USGS

National Field Manual for the Collection of Water-Quality Data (Rounds, 2006).

LU-1 LU-13 LU-16 LU-25 LU-47 LU-50 LU-53

Latitude

Longitude

-17.455

-57.910

-17.455

-57.853

-17.51

-57.853

-17.492

-57.815

-17.546

-57.76

-17.492

-57.759

-17.442

-57.743

initial pH 7.76 7.75 7.85 7.7 7.73 7.67 7.89

pH 4.0 1.43 1.35 1.74 1.32 1.415 1.47 1.53

pH 3.9 1.485 1.415 1.825 1.355 1.47 1.515 1.6

pH 3.8 1.565 1.495 1.9 1.435 1.53 1.575 1.715

pH 3.7 1.66 1.62 2 1.535 1.605 1.665 1.83

pH 3.6 1.805 1.78 2.12 1.665 1.705 1.795 2.005

pH 3.5 1.97 2.005 2.29 1.855 1.89 1.94 2.245

pH 3.4 2.2 2.335 2.495 2.065 2.07 2.15 2.57

pH 3.3 2.545 2.77 2.765 2.37 2.295 2.425 3.025

pH 3.2 3.015 3.45 3.18 2.785 2.635 2.775 3.455

Alkalinity

(μeq l-1

HCO3-)

540.9

± 21.4

434.9

± 39.3

729.7

± 11.6

495.8

± 13.9

577.2

± 10

589.4

± 11.0

557.6

± 17.1

mg l-1

HCO3-

33.00 Error 44.52 30.25 35.22 35.96 34.02

Suspended sediments from September 2015 (TSS= total suspended solids; OSS= organic

suspended solids; ISS= inorganic suspended solids).

LU-1 LU-13 LU-16 LU-25 LU-47 LU-50 LU-53

Initial Mass (g) 0.1237 0.1211 0.1256 0.1236 0.1258 0.1252 0.1248

Mass (g) post-105°C 0.1299 0.1286 0.1321 0.1493 0.1321 0.1342 0.1253

Mass (g) post-500°C 0.1283 0.1267 0.1302 0.1465 0.1307 0.1325 0.125

Volume filtered (ml) 1,023 1,118 1,116 1,061 1,112 1,110 1,126

TSS (mg l-1) 6.061 6.708 5.824 24.222 5.665 8.108 0.444

ISS (mg l-1) 4.497 5.009 4.122 21.583 4.406 6.577 0.178

OSS (mg l-1) 1.564 1.699 1.703 2.639 1.259 1.532 0.266

Page 95: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

83

Suspended sediments from November 2015 (TSS= total suspended solids; OSS= organic

suspended solids; ISS= inorganic suspended solids, Lat. = latitude; Long. = longitude).

Lat.

Long.

Initial

Mass

(g)

Post-

105°C

mass

(g)

Post-

500°C

mass

(g)

Volume

(ml=g3)

TSS

(mg l-1)

ISS

(mg l-1)

OSS

(mg l-1)

-17.553 -57.732 0.125 0.135 0.133 500 18.2 14.8 3.4

-17.542 -57.758 0.125 0.133 0.131 500 16.2 11.6 4.6

-17.532 -57.780 0.127 0.130 0.129 500 6 4.2 1.8

-17.519 -57.804 0.124 0.133 0.132 500 16.6 14.4 2.2

-17.510 -57.825 0.127 0.136 0.135 500 19.6 16.2 3.4

-17.500 -57.846 0.126 0.130 0.129 500 9.6 6.8 2.8

-17.493 -57.865 0.127 0.137 0.135 500 19.2 15.6 3.6

-17.480 -57.890 0.126 0.135 0.133 500 18 14.6 3.4

-17.474 -57.908 0.125 0.133 0.131 500 15.2 11.8 3.4

-17.470 -57.882 0.124 0.129 0.128 500 9.6 6.2 3.4

-17.483 -57.858 0.125 0.130 0.129 500 11 8.2 2.8

-17.468 -57.852 0.128 0.131 0.130 500 5.8 3.4 2.4

-17.464 -57.828 0.124 0.128 0.127 500 8.2 5.4 2.8

-17.484 -57.828 0.127 0.131 0.130 500 9.6 6.6 3

-17.499 -57.827 0.127 0.131 0.130 500 9 6.8 2.2

-17.464 -57.802 0.124 0.126 0.125 500 3 2.2 0.8

-17.483 -57.803 0.128 0.131 0.130 500 6.4 4.2 2.2

-17.505 -57.800 0.126 0.129 0.128 500 6.8 4.4 2.4

-17.540 -57.804 0.127 0.135 0.133 500 14.8 11.2 3.6

-17.547 -57.780 0.123 0.132 0.130 500 18.2 14 4.2

-17.511 -57.774 0.124 0.129 0.128 500 10 7.2 2.8

-17.486 -57.773 0.125 0.127 0.126 500 5.8 3.8 2

-17.462 -57.779 0.123 0.126 0.125 500 6.6 4 2.6

-17.460 -57.753 0.122 0.124 0.123 500 3 1.6 1.4

-17.487 -57.750 0.124 0.128 0.127 500 7.8 6 1.8

-17.513 -57.744 0.124 0.129 0.128 500 9 6.2 2.8

-17.539 -57.769 0.122 0.127 0.126 500 11.6 9.6 2

-17.562 -57.811 0.122 0.140 0.138 500 36.6 32 4.6

-17.588 -57.788 0.125 0.132 0.131 250 26 22 4

-17.588 -57.788 0.124 0.132 0.131 250 34.8 28.8 6

-17.669 -57.759 0.123 0.128 0.127 250 22.4 18 4.4

-17.669 -57.759 0.124 0.131 0.130 250 29.2 25.2 4

Page 96: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

84

Duplicate suspended sediments from November 2015 (TSS=total suspended solids;

OSS=organic suspended solids; ISS=inorganic suspended solids; Lat. = latitude; Long. =

longitude).

Lat. Long. Initial

Mass

(g)

Post-

105°C

mass

(g)

Post-

500°C

mass

(g)

Volume

(ml=g3)

TSS

(mg l-1)

ISS

(mg l-1)

OSS

(mg l-1)

-17.553 -57.732 0.1291 0.1399 0.1378 500 21.6 17.4 4.2

-17.542 -57.758 0.125 0.1344 0.1325 500 18.8 15 3.8

-17.532 -57.780 0.1266 0.1313 0.1297 500 9.4 6.2 3.2

-17.519 -57.804 0.1232 0.1314 0.1297 500 16.4 13 3.4

-17.510 -57.825 0.1267 0.1383 0.1362 500 23.2 19 4.2

-17.500 -57.846 0.125 0.1318 0.1301 500 13.6 10.2 3.4

-17.493 -57.865 0.1251 0.131 0.1294 500 11.8 8.6 3.2

-17.480 -57.890 0.1246 0.1315 0.1299 500 13.8 10.6 3.2

-17.474 -57.908 0.1251 0.1354 0.1334 500 20.6 16.6 4

-17.470 -57.882 0.1282 0.1335 0.1319 500 10.6 7.4 3.2

-17.483 -57.858 0.1253 0.1316 0.1299 500 12.6 9.2 3.4

-17.468 -57.852 0.1264 0.131 0.1299 500 9.2 7 2.2

-17.464 -57.828 0.1253 0.1283 0.1273 500 6 4 2

-17.484 -57.828 0.123 0.1273 0.126 500 8.6 6 2.6

-17.499 -57.827 0.1268 0.1311 0.13 500 8.6 6.4 2.2

-17.464 -57.802 0.1248 0.1269 0.126 500 4.2 2.4 1.8

-17.483 -57.803 0.1245 0.1273 0.126 500 5.6 3 2.6

-17.505 -57.800 0.1251 0.1303 0.1289 500 10.4 7.6 2.8

-17.540 -57.804 0.1286 0.1371 0.1349 500 17 12.6 4.4

-17.547 -57.780 0.1228 0.1299 0.1284 500 14.2 11.2 3

-17.511 -57.774 0.1236 0.1298 0.1286 500 12.4 10 2.4

-17.486 -57.773 0.1211 0.1248 0.123 500 7.4 3.8 3.6

-17.462 -57.779 0.1222 0.1262 0.1249 500 8 5.4 2.6

-17.460 -57.753 0.1231 0.1256 0.1245 500 5 2.8 2.2

-17.487 -57.750 0.1228 0.1256 0.125 500 5.6 4.4 1.2

-17.513 -57.744 0.1233 0.1276 0.1264 500 8.6 6.2 2.4

-17.539 -57.769 0.124 0.1293 0.128 500 10.6 8 2.6

-17.562 -57.811 0.1225 0.1441 0.1409 500 43.2 36.8 6.4

-17.588 -57.788 0.1242 0.13 0.129 250 23.2 19.2 4

-17.588 -57.788 0.1231 0.13 0.1268 250 27.6 14.8 12.8

-17.669 -57.759 0.1243 0.131 0.1301 250 26.8 23.2 3.6

-17.669 -57.759 0.1231 0.1307 0.1295 250 30.4 25.6 4.8

Page 97: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

85

Appendix B: Lake Floor Sediment Analyses

Bioavailable elements from 69 surface samples. Most results of bioavailable aluminum

were excluded due to trace quantities. Cells highlighted in red (NTA) indicate sample

sizes were insufficient for analysis.

ID Latitude Longitude P Mn Fe Cu Zn K Na Ca Mg

mg l-1 mg l-1 mg l-1 mg l-1 mg l-1 mg l-1 mg l-1 mg l-1 mg l-1

1 -17.455 -57.91 6.49 399.6 1952 1.58 4.71 100 110 1046 274.3

2 -17.468 -57.921 6.96 138.2 1578 1.27 3.39 200 75 1343 413.6

3 -17.473 -57.91 7.43 180.6 1414 1.28 3.87 135 50 1469 424.2

4 -17.455 -57.891 7.66 171.2 449.1 2.13 4.58 99 43 1092 252.2

5 -17.473 -57.891 5.38 165.2 1584 1.25 3.14 89 61 975.3 211.8

6 -17.486 -57.908 5.03 58.98 1384 1.32 4.66 135 56 1277 314.1

7 -17.491 -57.893 4.5 191.3 1155 2.04 1.63 125 63 2151 665.3

8 -17.509 -57.886 5.85 33.7 1080 1.03 3.13 110 46 1084 224.7

9 -17.509 -57.872 6.79 141.9 1695 1.15 2.89 105 43 1093 323.6

10 -17.491 -57.872 10.41 146.6 1225 1.4 4.77 84 38 1168 365.1

11 -17.473 -57.872 4.5 3.33 1556 1.5 3.58 66 105 927.6 193.8

12 -17.455 -57.872 3.69 210.4 1820 1.85 3.09 95 54 857.1 265.7

13 -17.455 -57.853 3.1 168.3 1409 1.91 3.11 77 35 806.7 199.1

14 -17.473 -57.853 3.1 37.69 411.8 3.18 3.61 145 56 2182 646.3

15 -17.491 -57.853 9.77 36.94 1177 1.14 3.75 65 29 845 280

16 -17.51 -57.853 10.65 147.6 1173 1.05 4.11 95 39 1108 390

17 -17.528 -57.854 5.62 150.9 1260 1.35 5.88 130 55 1758 474.8

18 -17.528 -57.834 9.3 58.21 1309 0.9 4.17 105 54 1557 581.9

19 -17.51 -57.835 6.32 213.3 1440 0.94 5.19 82 42 1201 387

20 -17.492 -57.834 7.02 252.2 1358 1.19 4.41 62 28 994.2 298.1

21 -17.474 -57.834 6.14 149.8 1364 1.33 3.23 50 46 866.7 221.5

22 -17.455 -57.834 1.99 205.4 1226 1.75 2.58 63 30 790.7 167.5

23 -17.456 -57.815 1.29 314.9 1263 0.98 1.94 51 26 358.5 96.57

24 -17.474 -57.815 5.32 37.56 1232 0.88 2.5 29 20 511 71.25

25 -17.492 -57.815 5.91 158.2 1919 0.97 6.07 64 50 1163 344.2

26 -17.509 -57.816 6.55 210.8 1493 0.86 5.35 81 33 1227 411.8

27 -17.528 -57.816 8.19 223.8 1463 1.37 4.94 87 35 1166 385.2

28 -17.546 -57.816 9.01 214 1380 1.03 4.53 88 36 1271 463.9

29 -17.559 -57.815 9.36 168.2 1245 1.15 3.72 95 40 1332 495

30 -17.564 -57.797 7.25 169.1 1227 0.75 4.54 68 31 1081 368.7

31 -17.544 -57.797 7.37 184.6 1307 0.79 3.9 66 53 970.8 326.5

32 -17.528 -57.797 5.62 148.1 1609 0.64 5.41 55 29 947.5 309

33 -17.51 -57.797 5.79 153.9 1786 0.73 5.47 60 28 441.8 303.2

Page 98: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

86

34 -17.492 -57.796 6.03 149.1 2296 0.99 5.75 54 27 982.6 273.1

35 -17.474 -57.796 6.55 85.64 1331 0.54 1.82 21 18 313.2 71.43

36 -17.456 -57.796 2.81 202.4 1688 0.77 2.22 45 22 393.7 95.01

37 -17.444 -57.796 1.46 50.11 1270 1.63 1.81 55 67 440.2 91.49

38 -17.442 -57.779 1.76 88.81 1143 1.35 1.83 60 27 465.6 98.84

39 -17.458 -57.778 7.31 74.4 1810 2.22 2.93 46 28 755.9 127.6

40 -17.474 -57.778 5.27 68.63 1892 2.22 1.52 59 28 960.7 246.8

41 -17.492 -57.778 5.73 46.63 1488 0.9 1.39 17 19 217 41.4

42 -17.51 -57.778 5.62 51.12 1250 0.78 2.51 17 13 179.2 45.76

43 -17.528 -57.778 5.38 100.8 1897 1.86 3.59 40 83 543.3 109.3

44 -17.546 -57.778 7.25 128.9 2317 1.34 4.24 50 30 725.4 196.7

45 -17.56 -57.778 7.96 157.1 1437 1.25 3.64 64 35 963.3 316.6

46 -17.564 -57.759 8.83 157.1 1455 1.24 3.24 61 32 850.8 272.9

47 -17.546 -57.759 7.78 109.1 2457 1.28 5.3 37 33 675.1 169.9

48 -17.528 -57.759 5.62 60.09 1023 0.81 1.19 17 19 248.5 55.72

49 -17.51 -57.759 7.55 54.62 1583 0.87 1.28 16 23 252.1 43.3

50 -17.492 -57.759 6.2 43.48 1046 0.68 0.8 15 15 150.8 32.52

51 -17.474 -57.759 6.73 53.61 1542 1.31 2.22 28 23 593.5 75.99

52 -17.456 -57.759 7.49 38.5 906.7 0.8 1.09 19 26 258.6 32.79

53 -17.442 -57.759 2.28 135.9 926.7 1.7 1.26 59 31 775.9 135.8

54 -17.444 -57.74 4.1 58.24 2171 1.59 4.7 60 51 NTA NTA

55 -17.456 -57.74 6.44 191.7 2033 1.42 1.94 58 72 980.7 178

56 -17.474 -57.741 9.24 22.05 842.2 0.88 1.37 21 83 320.5 24.9

57 -17.492 -57.74 7.72 86.28 1043 1.05 0.94 45 24 516.7 98.38

58 -17.51 -57.74 9.18 108.7 2369 1.56 2.56 39 28 657.8 133.7

59 -17.528 -57.74 8.37 100.4 1920 1.32 2.09 31 32 515.4 91.19

60 -17.546 -57.74 11.88 155.6 2036 1.33 3.38 43 27 705.8 174.8

61 -17.564 -57.74 9.36 120.2 1548 1.62 3.51 54 45 959.5 277.6

62 -17.545 -57.702 9.18 78.32 1838 2.19 3.49 79 36 1258 349.8

63 -17.546 -57.721 10.12 178.5 2104 1.91 3.35 59 30 1013 231.7

64 -17.528 -57.722 8.13 135.5 1696 1.63 2.04 50 29 939.1 230.7

65 -17.51 -57.721 8.6 47.96 1403 1.6 3.87 60 37 1820 327.9

66 -17.492 -57.721 7.37 85.38 2400 2.19 6.1 50 39 1280 274.8

67 -17.474 -57.721 6.96 5.15 1547 1.64 2.87 49 26 1533 312.1

68 -17.456 -57.722 4.39 6.78 1815 1.6 4.38 76 37 1390 326.5

69 -17.445 -57.724 6.49 5.07 1623 1.78 4.61 88 140 NTA NTA

Page 99: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

87

Surface ED-XRF select major elemental abundances in percent, following Bruker’s

empirical conversion (Rowe et al., 2012). Cells in red are considered ‘zero.’

ID Latitude Longitude AlKa1 SiKa1 K Ka1 CaKa1 TiKa1 FeKa1

1 -17.455 -57.91 3.165 25.619 0.323 0.125 0.703 1.154

2 -17.468 -57.921 3.512 25.332 0.377 0.178 0.636 1.230

3 -17.473 -57.91 3.059 26.703 0.393 0.195 0.653 1.203

4 -17.455 -57.891 3.069 27.656 0.282 0.144 0.646 1.154

5 -17.473 -57.891 3.252 27.434 0.245 0.173 0.610 0.978

6 -17.486 -57.908 4.711 22.420 0.432 0.165 0.630 1.517

7 -17.491 -57.893 4.561 22.563 0.297 0.209 0.635 2.501

8 -17.509 -57.886 3.217 27.299 0.398 0.129 0.664 0.784

9 -17.509 -57.872 3.483 26.841 0.292 0.128 0.650 1.092

10 -17.491 -57.872 2.910 27.906 0.301 0.168 0.628 1.291

11 -17.473 -57.872 2.891 28.646 0.204 0.115 0.601 0.708

12 -17.455 -57.872 3.038 28.650 0.255 0.104 0.641 0.620

13 -17.455 -57.853 2.795 28.587 0.280 0.117 0.626 0.687

14 -17.473 -57.853 6.105 22.154 0.248 0.139 0.777 1.090

15 -17.491 -57.853 2.557 29.966 0.346 0.115 0.712 0.811

16 -17.51 -57.853 2.466 29.361 0.391 0.138 0.694 1.348

17 -17.528 -57.854 4.456 23.204 0.485 0.235 0.655 1.552

18 -17.528 -57.834 2.989 27.089 0.391 0.184 0.767 1.867

19 -17.51 -57.835 2.553 29.085 0.298 0.170 0.808 1.482

20 -17.492 -57.834 2.557 28.481 0.288 0.128 0.672 1.193

21 -17.474 -57.834 2.739 29.000 0.307 0.114 0.577 0.781

22 -17.455 -57.834 2.187 31.233 0.265 0.074 0.550 0.275

23 -17.456 -57.815 2.735 30.053 0.293 0.048 0.551 0.198

24 -17.474 -57.815 2.570 29.896 0.263 0.111 0.496 0.405

25 -17.492 -57.815 2.462 28.830 0.298 0.119 0.650 1.394

26 -17.509 -57.816 2.602 27.824 0.327 0.150 0.821 1.602

27 -17.528 -57.816 2.264 29.542 0.296 0.154 0.946 1.408

28 -17.546 -57.816 2.547 28.174 0.338 0.170 0.838 1.677

29 -17.559 -57.815 2.829 27.830 0.468 0.162 0.725 1.556

30 -17.564 -57.797 2.283 28.548 0.387 0.139 0.697 1.596

31 -17.544 -57.797 2.233 29.357 0.346 0.130 0.791 1.444

32 -17.528 -57.797 2.178 29.916 0.308 0.119 0.767 1.526

33 -17.51 -57.797 2.384 28.130 0.336 0.136 0.690 1.731

34 -17.492 -57.796 3.073 27.371 0.337 0.140 0.627 1.333

35 -17.474 -57.796 2.204 30.254 0.160 0.102 0.478 0.283

36 -17.456 -57.796 2.335 29.608 0.261 0.057 0.479 0.334

37 -17.444 -57.796 4.046 26.793 0.252 0.045 0.486 0.828

38 -17.442 -57.779 3.627 28.391 0.253 0.052 0.587 0.485

Page 100: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

88

39 -17.458 -57.778 3.030 27.975 0.293 0.099 0.523 0.775

40 -17.474 -57.778 3.564 26.488 0.239 0.099 0.527 1.013

41 -17.492 -57.778 1.814 30.373 0.137 0.035 0.365 0.421

42 -17.51 -57.778 2.154 30.378 0.115 0.034 0.370 0.501

43 -17.528 -57.778 2.708 27.263 0.173 0.093 0.436 1.290

44 -17.546 -57.778 2.587 27.452 0.271 0.117 0.563 1.469

45 -17.56 -57.778 2.425 29.531 0.407 0.130 0.786 1.320

46 -17.564 -57.759 2.334 29.485 0.403 0.120 0.763 1.127

47 -17.546 -57.759 2.813 27.516 0.309 0.113 0.591 1.263

48 -17.528 -57.759 2.037 30.475 -0.036 0.044 0.252 0.479

49 -17.51 -57.759 2.054 30.461 0.130 0.048 0.307 0.385

50 -17.492 -57.759 1.530 31.452 0.039 0.021 0.301 0.008

51 -17.474 -57.759 1.871 30.962 0.138 0.055 0.377 0.264

52 -17.456 -57.759 1.375 32.244 0.203 0.039 0.477 -0.103

53 -17.442 -57.759 3.604 27.849 0.150 0.066 0.558 0.352

54 -17.444 -57.74 3.314 25.294 0.253 0.112 0.762 1.040

55 -17.456 -57.74 2.804 27.802 0.252 0.159 0.554 0.891

56 -17.474 -57.741 1.533 31.454 0.358 0.085 0.397 0.184

57 -17.492 -57.74 2.904 28.095 0.104 0.118 0.338 0.645

58 -17.51 -57.74 2.867 28.191 0.219 0.117 0.438 0.955

59 -17.528 -57.74 2.488 29.231 0.191 0.074 0.441 0.627

60 -17.546 -57.74 2.979 28.389 0.351 0.116 0.530 0.903

61 -17.564 -57.74 2.598 28.546 0.253 0.123 0.581 1.345

62 -17.545 -57.702 3.579 24.501 0.354 0.204 0.755 1.736

63 -17.546 -57.721 3.025 27.501 0.357 0.155 0.681 1.151

64 -17.528 -57.722 3.126 27.933 0.215 0.115 0.575 0.983

65 -17.51 -57.721 3.774 20.573 0.276 0.446 0.698 1.646

66 -17.492 -57.721 2.977 25.880 0.294 0.233 0.483 1.493

67 -17.474 -57.721 2.682 27.482 0.503 0.179 0.367 0.791

68 -17.456 -57.722 2.928 26.604 0.354 0.210 0.525 0.993

69 -17.445 -57.724 3.135 26.094 0.284 0.301 0.555 1.258

Page 101: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

89

Surface BiSi data. ABS = absorbance at 812 nm

ID Batch

#

Weight

(mg)

ABS mM

SiO2

Dilution

Factor

BiSi(‰) BiSi(%)

1 1 54.91 0.143 0.9 0.036 40 3.97

2 1 54.56 0.149 0.9 0.038 42 4.17

3 1 54.82 0.144 0.9 0.037 40 4.01

4 1 55.54 0.140 0.9 0.036 38 3.85

5 1 55.37 0.116 0.7 0.029 32 3.20

6 1 56.80 0.151 1.0 0.038 41 4.06

7 1 55.15 0.116 0.7 0.029 32 3.21

8 1 54.38 0.127 0.8 0.032 36 3.56

9 1 54.58 0.080 0.5 0.020 22 2.24

10 1 54.07 0.091 0.6 0.023 26 2.57

11 1 55.48 0.081 0.5 0.021 22 2.23

12 1 55.46 0.109 0.7 0.028 30 3.00

13 1 54.65 0.099 0.6 0.025 28 2.76

14 1 54.65 0.102 0.6 0.026 28 2.85

15 1 54.17 0.077 0.5 0.020 22 2.17

16 1 54.48 0.097 0.6 0.025 27 2.72

17 1 55.03 0.179 1.1 0.045 50 4.96

18 1 54.48 0.111 0.7 0.028 31 3.11

19 1 53.52 0.106 0.7 0.027 30 3.02

20 1 54.90 0.101 0.6 0.026 28 2.81

21 1 55.21 0.075 0.5 0.019 21 2.07

22 1 54.77 0.086 0.5 0.022 24 2.40

23 1 55.52 0.057 0.4 0.014 16 1.57

24 1 55.32 0.043 0.3 0.011 12 1.19

25 1 54.75 0.070 0.4 0.018 20 1.95

26 1 53.18 0.087 0.6 0.022 25 2.50

27 1 54.59 0.118 0.7 0.030 33 3.30

28 1 55.02 0.111 0.7 0.028 31 3.08

29 1 54.24 0.101 0.6 0.026 28 2.84

30 1 55.50 0.085 0.5 0.022 23 2.34

31 1 55.84 0.100 0.6 0.025 27 2.73

32 1 55.68 0.067 0.4 0.017 18 1.84

33 1 54.57 0.079 0.5 0.020 22 2.21

34 1 55.75 0.082 0.5 0.021 22 2.24

35 1 55.79 0.019 0.1 0.005 5 0.52

36 2 55.05 0.055 0.3 0.014 15 1.52

37 2 55.17 0.062 0.4 0.016 17 1.72

38 2 55.58 0.067 0.4 0.017 18 1.84

39 2 55.24 0.081 0.5 0.021 22 2.24

40 2 55.24 0.061 0.4 0.015 17 1.69

41 2 55.74 0.028 0.2 0.007 8 0.77

Page 102: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

90

42 2 55.29 0.034 0.2 0.009 9 0.94

43 2 55.31 0.055 0.3 0.014 15 1.52

44 2 54.72 0.070 0.4 0.018 20 1.95

45 2 55.57 0.105 0.7 0.027 29 2.88

46 2 54.83 0.085 0.5 0.022 24 2.37

47 2 54.78 0.062 0.4 0.016 17 1.73

48 2 56.88 0.034 0.2 0.009 9 0.91

49 2 55.01 0.031 0.2 0.008 9 0.86

50 2 54.89 0.027 0.2 0.007 8 0.75

51 2 54.36 0.038 0.2 0.010 11 1.07

52 2 55.83 0.031 0.2 0.008 8 0.85

53 2 55.21 0.059 0.4 0.015 16 1.63

54 2 54.91 0.138 0.9 0.035 38 3.84

55 2 55.28 0.120 0.8 0.030 33 3.31

56 2 54.06 0.052 0.3 0.013 15 1.47

57 2 55.84 0.076 0.5 0.019 21 2.08

58 2 54.76 0.066 0.4 0.017 18 1.84

59 2 54.83 0.047 0.3 0.012 13 1.31

60 2 56.20 0.069 0.4 0.018 19 1.87

61 2 55.11 0.081 0.5 0.021 22 2.24

62 2 54.55 0.127 0.8 0.032 36 3.55

63 2 54.70 0.102 0.6 0.026 28 2.85

64 2 54.79 0.070 0.4 0.018 19 1.95

65 2 54.88 0.252 1.6 0.064 70 7.01

66 2 55.44 0.147 0.9 0.037 40 4.05

67 2 54.66 0.099 0.6 0.025 28 2.76

68 2 53.40 0.179 1.1 0.045 51 5.12

69 2 55.23 0.216 1.4 0.055 60 5.97

53

duplicate

2

55.23 0.067 0.4 0.017 19 2

65

duplicate

2

55.18 0.264 1.7 0.067 73 7

49

duplicate

2

55.36 0.038 0.2 0.010 10 1

45

duplicate

2

55.88 0.099 0.6 0.025 27 3

37

duplicate

2

53.42 0.068 0.4 0.017 19 2

5

duplicate

2

55.79 0.114 0.7 0.029 31 3

19

duplicate

2

55.26 0.110 0.7 0.028 30 3

26

duplicate

2

55.54 0.119 0.8 0.030 33 3

Page 103: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

91

33

duplicate

2

54.24 0.081 0.5 0.021 23 2

Standard

A#2

1

45.53 0.601 3.8 0.153 201 20

Standard

B#2

2

44.08 0.570 3.6 0.145 197 20

Page 104: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

92

Table listing all surface total carbon (TC), total inorganic carbon (TIC), total organic

carbon (TOC), and all associated measurements. All TIC values that are ‘zero’ were

samples that were aborted on the coulometer after 15 minutes of continuous analysis.

ID Mass TC

(g)

Mass

TIC (g)

TC

(ppm)

TIC

(ppm)

TOC

(ppm)

TOC

(%)

TIC

(%)

SARM

41

0.2846 62139

502-030 0.2177 49388

502-630 0.1147 5149.3

1 0.1615 0.1066 25765 0 25765 2.58 0.000

2 0.1493 0.1065 31633 0 31633 3.16 0.000

3 0.1511 0.1046 26672 0 26672 2.67 0.000

4 0.1519 0.1078 38101 0 38101 3.81 0.000

5 0.1547 0.1136 17725 77.75 17647.25 1.76 0.008

SARM

41

0.2541

61700

502-030 0.2525 47780

502-630 0.1256 3550.9

6 0.153 0.1053 40192 54.71 40137.29 4.01 0.005

7 0.154 0.1116 9829.7 100.07 9729.63 0.97 0.010

8 0.1561 0.1204 32843 0 32843 3.28 0.000

9 0.1558 0.1099 10292 80.47 10211.53 1.02 0.008

10 0.1563 0.1136 19012 0 19012 1.90 0.000

11 0.15 0.1102 18907 20.36 18886.64 1.89 0.002

12 0.1524 0.1152 18167 34.17 18132.83 1.81 0.003

13 0.1526 0.1089 18248 35.2 18212.8 1.82 0.004

14 0.1585 0.1042 3756.7 55.6 3701.1 0.37 0.006

15 0.1503 0.1103 14997 0 14997 1.50 0.000

16 0.1553 0.1067 12541 0 12541 1.25 0.000

17 0.1515 0.1024 53984 48.2 53935.8 5.39 0.005

18 0.1523 0.1001 15352 26.1 15325.9 1.53 0.003

19 0.1538 0.1087 17336 31.2 17304.8 1.73 0.003

20 0.155 15829 0 15829 1.58 0.000

prime 0.1515 666560

prime 0.1651 684740

prime 0.1982 674820

SARM

41

0.2839

61913

502-030 0.2303 48335

502-630 0.1484 4804.3

21 0.152 0.1123 14527 0 14527 1.45 0.000

22 0.1513 0.1145 15291 0 15291 1.53 0.000

23 0.1681 0.1168 5133 0 5133 0.51 0.000

24 0.1461 0.1003 5587.9 0 5587.9 0.56 0.000

Page 105: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

93

25 0.1466 0.1026 13780 0 13780 1.38 0.000

26 0.1529 0.1079 16798 0 16798 1.68 0.000

27 0.1572 0.1204 18307 0 18307 1.83 0.000

28 0.1489 0.1069 17321 0 17321 1.73 0.000

29 0.1596 0.1093 13942 0 13942 1.39 0.000

30 0.1566 0.1094 12790 0 12790 1.28 0.000

31 0.1501 0.1163 11731 0 11731 1.17 0.000

32 0.1437 0.1046 7594.7 0 7594.7 0.76 0.000

33 0.1573 0.1178 9957.9 0 9957.9 1.00 0.000

34 0.1434 0.1171 11984 0 11984 1.20 0.000

35 0.166 0.1129 2890.1 0 2890.1 0.29 0.000

36 0.1564 0.1034 5935.8 28.33 5907.47 0.59 0.003

37 0.1606 0.1146 7525 0 7525 0.75 0.000

38 0.1609 0.1171 6369.9 22.86 6347.04 0.63 0.002

39 0.1458 0.1055 11354 56.26 11297.74 1.13 0.006

40 0.159 0.1026 8298.4 41.89 8256.51 0.83 0.004

SARM

41

0.2622

62007

503-030 0.2379 47956

502-630 0.1315 4434.7

prime 0.1718 712220

prime 0.1473 715360

prime 0.1923 707440

SARM

41

0.2529

65142

502-030 0.2726 50611

502-630 0.139 5284.9

41 0.1628 0.1107 2881.5 86.32 2795.18 0.28 0.009

42 0.1626 0.1044 1908.2 109.92 1798.28 0.18 0.011

43 0.1762 0.1069 5495.8 0 5495.8 0.55 0.000

44 0.1778 0.1133 8423.5 64.94 8358.56 0.84 0.006

45 0.1534 0.1175 11597 78.16 11518.84 1.15 0.008

46 0.188 0.1047 11344 0 11344 1.13 0.000

47 0.1395 0.1226 6721.4 74.87 6646.53 0.66 0.007

48 0.1609 0.1115 1811.3 42.99 1768.31 0.18 0.004

49 0.1688 0.1038 2297.9 0 2297.9 0.23 0.000

50 0.1602 0.1092 1569.4 31.14 1538.26 0.15 0.003

51 0.1494 0.1215 5275.9 26.78 5249.12 0.52 0.003

52 0.1746 0.1114 2309.2 0 2309.2 0.23 0.000

53 0.1737 0.1197 4567.7 0 4567.7 0.46 0.000

54 0.155 0.1129 8767.9 36.53 8731.37 0.87 0.004

55 0.1569 0.1073 17705 56.9 17648.1 1.76 0.006

56 0.1563 0.1157 5683.7 62.86 5620.84 0.56 0.006

57 0.1501 0.1161 4662.8 26.13 4636.67 0.46 0.003

58 0.1586 0.1182 12096 0 12096 1.21 0.000

Page 106: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

94

59 0.1513 0.1129 7374.7 14.67 7360.03 0.74 0.001

60 0.1405 0.113 9207.9 0 9207.9 0.92 0.000

61 0.188 0.1037 13850 0 13850 1.39 0.000

62 0.1546 0.101 32272 39.94 32232.06 3.22 0.004

63 0.1388 0.1108 24404 0 24404 2.44 0.000

64 0.1461 0.1141 15641 73.21 15567.79 1.56 0.007

65 0.1592 0.1057 120460 0 120460 12.05 0.000

66 0.1461 0.1096 48277 43.4 48233.6 4.82 0.004

67 0.178 0.1186 26925 0 26925 2.69 0.000

68 0.1791 0.1145 50749 17.82 50731.18 5.07 0.002

69 0.1682 0.1088 58985 0 58985 5.90 0.000

SARM

41

0.2284 64732

502-030 0.2184 51072

502-630 0.1489 5116.3

Page 107: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

95

All surface grain size measurements are listed in the table below. The percentages of

sand, silt, and clay were adjusted if the sum was 99.99% (0.01% added to smallest value)

or 100.01% (0.01% subtracted from largest value).

ID % Sand % Silt % Clay Skewness D50 (µm)

1 15.45 69.14 15.41 7.834 22.285

2 23.11 59.99 16.9 7.931 26.075

3 16.53 67.99 15.48 6.706 23.489

4 38.82 49.28 11.9 1.515 41.485

5 42.84 46.47 10.69 6.955 53.798

6 24.3 53.29 22.41 1.28 22.125

7 20.37 56.09 23.54 4.019 17.093

8 29 57.98 13.02 0.777 41.632

9 44.32 45.56 10.12 0.545 56.936

10 23.52 66.14 10.34 4.025 36.489

11 36.71 53.95 9.34 1.311 45.726

12 25.92 65.28 8.8 2.35 39.216

13 19.18 73.04 7.78 2.851 34.755

14 18.65 54.21 27.14 2.158 13.557

15 20.29 71.78 7.93 1.798 32.316

16 16.32 73.59 10.09 7.29 28.631

17 14.74 61.53 23.73 6.413 14.168

18 18.1 71.17 10.73 4.776 26

19 6.86 80.59 12.55 1.664 20.408

20 13.35 76.57 10.08 4.197 27.331

21 18.94 71.66 9.4 2.989 33.867

22 18.77 73.82 7.41 1.596 36.488

23 31.45 64.24 4.31 6.01 48.262

24 40.56 53.7 5.74 7.263 53.544

25 33.78 59.03 7.19 3.914 45.179

26 9.65 77.13 13.22 2.549 20.039

27 4.43 81.42 14.15 2.209 16.212

28 7.81 79.77 12.42 2.825 18.447

29 10.42 76.78 12.8 2.559 22.925

30 19.37 71.78 8.85 3.321 28.885

31 5.7 81.61 12.69 2.811 16.853

32 25.87 64.52 9.61 8.101 30.335

33 21.77 68.47 9.76 9.099 31.003

34 22.09 65.61 12.3 8.475 29.672

35 57.57 38.55 3.88 0.531 69.977

36 52.25 45.08 2.67 0.772 64.92

37 25.77 37.1 37.13 1.342 31.376

38 38.98 55.26 5.76 3.711 52.385

39 46.96 47.39 5.65 0.661 60.12

Page 108: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

96

40 49.5 41.35 9.15 3.372 62.251

41 90.24 8.44 1.32 0.421 100.501

42 94.27 4.64 1.09 5.286 148.565

43 74.49 21.06 4.45 2.268 117.176

44 52.91 40.36 6.73 0.882 68.486

45 10.5 78.49 11.01 1.286 23.418

46 22.22 70.15 7.63 1.352 33.569

47 55.14 38.5 6.36 0.743 70.975

48 90.26 7.09 2.65 0.322 128.385

49 89.88 9.07 1.05 5.669 113.719

50 95.77 3.19 1.04 0.558 116.417

51 69.28 29.02 1.7 6.214 76.779

52 67.21 31.63 1.16 0.754 78.642

53 52.42 41.62 5.96 1.225 65.676

54 14.82 74.87 10.31 5.779 27.123

55 27.39 65.43 7.18 6.223 40.022

56 63.48 35.05 1.47 1.267 88.296

57 84.61 13.2 2.19 0.179 95.67

58 53.7 40.95 5.35 9.274 67.108

59 61.86 33.57 4.57 0.186 72.371

60 63.72 29.81 6.47 0.463 82.596

61 29.08 63.08 7.84 0.606 43.848

62 9.08 72.62 18.3 8.716 14.551

63 17.43 70.62 11.95 1.269 27.447

64 27.83 63.11 9.06 0.773 41.842

65 7.79 67.66 24.55 8.791 10.229

66 13.4 74.51 12.09 7.262 28.819

67 47.82 46.78 5.4 0.704 60.925

68 37 53.41 9.59 0.966 47.139

69 32.49 58.6 8.91 1.005 44.593

Page 109: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

97

Appendix C: Sample Statistical Analysis of Surface Sediment

This map of Lake Uberaba shows the results of the t-test for significance between

bioavailable Fe and Ca, organized by grain size contour intervals. Grain size largely

accounts for most variations among the variables, according to principal components

analysis (PCA).

Page 110: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

98

Appendix D: Sediment Core Data

Plots of RGB (depth vs. percent of RGB) with data recorded in intervals of 0.01 cm.

Camera calibration aperture was 11.0. Camera image aperture was 8.0 (core 2) or 16.0

(other cores).

0

10

20

30

40

50

60

70

80

90

100

0 50 100

Core 3

R G B

0 50 100

Core 4

R G B

0 50 100

Core 1

R G B

0 50

Core 2

Page 111: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

99

GeoTek Multi-Sensor Core Logger (MSCL) physical properties plots for Core 3

measured at 0.5 cm intervals.

-100

-90

-80

-70

-60

-50

-40

-30

-20

-10

0

0 2000

P-wave velocity

0 2 4

Gamma

Density

0 2000 4000

Impedance

-0.5 0 0.5 1

Fractional

Porosity

Page 112: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

100

Core 3 desiccated crack is outlined propagating from top to bottom in a light gray color.

The grain size of the material filling the crack is no different from its surroundings

(clayey silt). Scale has been corrected to account for floral foam at the top of the core (5.5

cm thick).

Page 113: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

101

Core 3 ED-XRF select major elemental abundances in percent, following Bruker’s

empirical conversion (Rowe et al., 2012).

ID AlKa1 SiKa1 K Ka1 CaKa1 TiKa1 FeKa1

3 2.732 29.832 0.252 0.072 0.452 1.169

5 3.039 28.692 0.336 0.092 0.526 1.386

7 3.249 31.397 0.305 0.104 0.546 1.438

9 2.983 28.600 0.283 0.093 0.466 1.543

11 3.301 27.645 0.320 0.144 0.559 1.639

13 4.245 27.344 0.476 0.102 0.438 0.992

15 4.189 27.777 0.521 0.088 0.454 0.748

17 3.640 29.195 0.412 0.071 0.384 0.574

19 3.581 28.818 0.384 0.064 0.369 0.434

21 3.167 29.643 0.355 0.056 0.410 0.412

23 4.822 27.424 0.448 0.078 0.416 0.716

25 5.809 24.345 0.571 0.109 0.497 1.129

29 6.092 23.817 0.636 0.116 0.536 1.414

33 5.953 24.926 0.663 0.110 0.496 1.171

37 6.374 24.965 0.751 0.120 0.506 1.433

41 6.060 24.268 0.687 0.108 0.470 1.597

45 6.315 23.210 0.694 0.121 0.495 2.077

49 6.009 24.658 0.737 0.114 0.527 1.851

53 6.178 23.486 0.766 0.125 0.576 2.243

57 6.374 22.540 0.846 0.148 0.672 3.436

61 6.413 23.468 0.823 0.131 0.597 2.471

65 6.122 24.034 0.809 0.114 0.581 2.056

69 6.135 24.022 0.781 0.110 0.591 2.141

73 5.952 23.763 0.880 0.122 0.630 2.404

77 5.874 23.878 0.781 0.115 0.578 2.474

81 6.307 24.230 0.821 0.110 0.585 2.071

85 5.995 25.390 0.786 0.101 0.592 1.474

89 5.805 24.257 0.735 0.107 0.567 2.356

93 5.813 23.373 0.712 0.110 0.587 2.752

97 5.311 24.910 0.674 0.104 0.535 2.384

Page 114: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

102

Core 3 BiSi values are displayed in the table below:

Depth

(cm)

Batch

#

Weight

(mg)

Absorbance at

812 nm

mM

SiO2

Dilution

Factor

BiSi

(‰)

BiSi

(%)

3 3 80.27 0.064 0.4 0.016 12 1.2

5 5 80.55 0.061 0.4 0.015 12 1.2

7 5 79.39 0.052 0.3 0.013 10 1.0

9 5 79.95 0.058 0.4 0.015 11 1.1

11 3 80.7 0.073 0.5 0.019 14 1.4

13 3 79.93 0.077 0.5 0.020 15 1.5

15 3 79.05 0.128 0.8 0.033 25 2.5

17 4 80.58 0.076 0.5 0.019 14 1.4

19 4 79.85 0.071 0.5 0.018 14 1.4

21 4 79.07 0.068 0.4 0.017 13 1.3

23 4 80.97 0.071 0.5 0.018 13 1.3

25 3 80.47 0.094 0.6 0.024 18 1.8

29 3 80.7 0.089 0.6 0.023 17 1.7

33 3 80.87 0.096 0.6 0.024 18 1.8

37 3 79.65 0.084 0.5 0.021 16 1.6

41 5 79.72 0.101 0.6 0.026 19 1.9

45 3 80.53 0.093 0.6 0.024 18 1.8

49 3 79.14 0.097 0.6 0.025 19 1.9

53 4 80.11 0.097 0.6 0.025 18 1.8

57 3 79.32 0.097 0.6 0.025 19 1.9

61 5 80.34 0.106 0.7 0.027 20 2.0

65 4 79.58 0.121 0.8 0.031 23 2.3

69 5 79.82 0.100 0.6 0.025 19 1.9

73 5 80.08 0.098 0.6 0.025 19 1.9

77 3 80.46 0.084 0.5 0.021 16 1.6

81 4 79.88 0.097 0.6 0.025 19 1.9

85 5 80.8 0.094 0.6 0.024 18 1.8

89 3 79.71 0.101 0.6 0.026 19 1.9

93 3 79.18 0.118 0.7 0.030 23 2.3

97 3 79.9 0.103 0.7 0.026 20 2.0

Page 115: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

103

Core 3 carbon analyses. Negative total carbon (TC) and negative total organic carbon

(TOC) values were treated as ‘zero.’

Depth

(cm)

Mass TC

(g)

Mass

TIC (g)

TC

(ppm)

TIC

(ppm)

TOC

(ppm)

TOC

(%)

TIC

(%)

3 0.190 0.104 4568.9 16.43 4552.47 0.46 0.002

5 0.167 0.106 5709.5 29 5680.50 0.57 0.003

7 0.273 0.117 6175.5 5.27 6170.23 0.62 0.001

9 0.194 0.119 5597.4 21.99 5575.41 0.56 0.002

11 0.237 0.118 5887.9 14.49 5873.41 0.59 0.001

13 0.209 0.112 3555.6 12.61 3542.99 0.35 0.001

15 0.243 0.118 3344 2.73 3341.27 0.33 0.000

17 0.177 0.106 2742 16.44 2725.56 0.27 0.002

19 0.222 0.113 2687.6 8.65 2678.95 0.27 0.001

21 0.183 0.116 3045 14.78 3030.22 0.30 0.001

23 0.195 0.120 2426.2 19.64 2406.56 0.24 0.002

25 0.312 0.114 1595.1 27.49 1567.61 0.16 0.003

27 0.358 0.104 1442.6 15.59 1427.01 0.14 0.002

29 0.197 0.101 1247.4 13.98 1233.42 0.12 0.001

33 0.181 0.111 482.33 30.5 451.83 0.05 0.003

37 0.221 0.115 396.15 12.66 383.49 0.04 0.001

41 0.206 0.114 154.78 0.11 154.67 0.02 0.000

45 0.185 0.110 3.1904 6.21 -3.02 0.00 0.001

49 0.241 0.114 139.29 17.65 121.64 0.01 0.002

53 0.193 0.101 -268.13 18.9 -287.03 0.00 0.002

57 0.167 0.109 320.21 23.52 296.69 0.03 0.002

61 0.229 0.116 -34.031 20.34 -54.37 0.00 0.002

65 0.194 0.108 64.709 12.94 51.77 0.01 0.001

69 0.184 0.115 299.22 24.76 274.46 0.03 0.002

73 0.248 0.108 68.646 3.37 65.28 0.01 0.000

77 0.184 0.114 126.32 24.35 101.97 0.01 0.002

81 0.188 0.110 -144.04 32.17 -176.21 0.00 0.003

85 0.234 0.103 -85.819 9.88 -95.70 0.00 0.001

89 0.197 0.107 -25.476 19.36 -44.84 0.00 0.002

93 0.166 0.106 -250.57 30.59 -281.16 0.00 0.003

97 0.211 0.103 85.005 0 85.01 0.01 0.001

Page 116: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

104

Core 3 grain size measurements are listed below. The percentages of sand, silt, and clay

were adjusted if the sum was 99.99% (0.01% added to smallest value) or 100.01%

(0.01% subtracted from largest value).

Depth (cm) % Sand % Silt % Clay Skewness D50 (µm)

3 72.44 23.35 4.21 0.566 97.353

5 75.84 20.62 3.54 0.583 103.039

7 78.6 18.27 3.13 0.551 106.255

9 74.76 21.03 4.21 0.666 108.107

11 75.49 21.08 3.43 5.597 106.579

13 50.28 41.3 8.42 8.086 63.367

15 41.94 50.69 7.37 0.756 55.349

17 42.46 51.19 6.35 8.733 55.798

19 40.47 53.27 6.26 0.743 53.987

21 51.81 43.72 4.47 0.856 64.692

23 44.29 46.96 8.75 0.912 56.027

25 39.34 49.82 10.84 0.879 49.541

29 35.94 50.81 13.25 0.872 45.392

33 41.93 47.2 10.87 0.798 52.933

37 37.06 49.37 13.57 0.869 46.315

41 39.35 46.59 14.06 0.912 47.019

45 46.68 41.22 12.1 8.573 58.282

49 43.33 43.95 12.72 6.033 53.098

53 20.99 55.02 23.99 8.698 17.398

57 14.5 57.95 27.55 10.003 10.601

61 10 55.37 34.63 7.373 6.936

65 5.19 55.32 39.49 3.307 5.67

69 16.19 57.95 25.86 7.474 10.964

73 18.35 55.9 25.75 1.706 14.019

77 11.54 54.53 33.93 7.844 7.38

81 15.73 59.16 25.11 9.43 12.171

85 19.28 57.29 23.43 1.467 15.629

89 16.1 59.46 24.44 7.358 14.363

93 26.27 54.19 19.54 1.05 28.347

97 20.91 58.48 20.61 1.363 17.329

Page 117: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

105

Core 3 radiochemical data (Bq kg-1) measured at SERRL lab, University of Kentucky.

Interval (cm) 137Cs 226Ra 228Ra 210Pbtotal 210Pbxs

0-2 0 ± 0 18.46 ± 1.18 24.83 ± 1.71 15.59 ± 0.99 10.59 ± 0.49

2-3 0 ± 0 19.63 ± 1.43 21.21 ± 1.49 17.96 ± 1.09 12.96 ± 0.59

3-4 0 ± 0 30.09 ± 1.97 23.52 ± 2.18 14.35 ± 0.71 9.35 ± 0.21

4-5 0 ± 0 13.31 ± 0.81 20.71 ± 1.07 14.71 ± 0.93 9.71 ± 0.43

5-6 0 ± 0 21.55 ± 1.50 25.83 ± 2.14 13.23 ± 0.82 8.23 ± 0.32

6-7 0 ± 0 20.03 ± 1.36 28.57 ± 1.88 12.24 ± 0.71 7.24 ± 0.21

7-8 0 ± 0 20.16 ± 1.20 27.63 ± 1.73 10.47 ± 0.66 5.47 ± 0.16

8-9 0 ± 0 21.29 ± 1.38 30.31 ± 2.01

9-10 0 ± 0 18.93 ± 1.16 20.30 ± 1.67 9.88 ± 0.59 4.88 ± 0.09

10-11 0 ± 0 16.25 ± 1.07 23.72 ± 1.54 6.53 ± 0.42 1.53 ± 0.08

11-12

6.68 ± 0.41 1.68 ± 0.09

12-13

5.55 ± 0.43

13-14

4.77 ± 0.41

14-15

15-16

16-17

17-18

18-19

19-20 0 ± 0 29.78 ± 2.01 33.32 ± 2.72

20-21

21-22

22-23

23-24

24-25

25-26

26-27

27-28

28-29

29-30

Page 118: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

106

GeoTek Multi-Sensor Core Logger (MSCL) physical properties plots for Core 4

measured at 0.5 cm intervals.

-90

-80

-70

-60

-50

-40

-30

-20

-10

0

0 1000

P-wave

Velocity

0 2 4

Gamma

Density

0 3000

Impedance

-0.5 0 0.5 1

Fractional

Porosity

Page 119: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

107

Core 4 ED-XRF select major elemental abundances in percent, following Bruker’s

empirical conversion (Rowe et al., 2012).

ID AlKa1 SiKa1 K Ka1 CaKa1 TiKa1 FeKa1

2 2.590 27.076 0.360 0.186 0.769 1.678

4 3.059 31.851 0.392 0.210 0.852 1.673

6 3.120 33.854 0.424 0.173 0.837 1.510

8 3.218 33.091 0.427 0.172 0.865 1.591

10 2.765 30.517 0.374 0.161 0.841 1.661

12 2.937 30.469 0.368 0.147 0.874 1.421

14 2.768 30.038 0.372 0.137 0.804 1.625

16 2.743 28.522 0.360 0.117 0.790 1.635

18 2.899 28.903 0.342 0.118 0.730 1.705

20 3.085 28.043 0.342 0.135 0.675 2.137

22 3.384 28.546 0.383 0.134 0.659 2.512

24 3.143 27.689 0.367 0.109 0.580 2.417

26 3.275 29.569 0.386 0.105 0.592 2.451

28 3.146 29.781 0.364 0.109 0.527 2.318

30 3.124 30.288 0.374 0.105 0.556 2.558

32 3.144 27.808 0.360 0.095 0.595 2.486

34 3.190 27.520 0.409 0.108 0.552 2.548

36 5.808 26.683 0.559 0.150 0.668 2.659

40 6.118 24.145 0.596 0.146 0.753 2.613

44 5.925 24.790 0.615 0.115 0.691 2.537

48 5.657 24.834 0.591 0.106 0.652 2.557

52 5.628 25.788 0.612 0.102 0.684 2.070

56 5.922 24.060 0.616 0.110 0.673 2.563

60 5.663 23.636 0.621 0.115 0.651 2.989

64 5.542 24.634 0.600 0.104 0.659 2.341

68 6.086 24.892 0.654 0.108 0.702 2.385

72 6.055 24.306 0.647 0.104 0.684 2.313

76 5.984 23.985 0.608 0.101 0.673 2.205

80 6.310 23.721 0.579 0.113 0.615 2.554

84 5.050 21.320 0.510 0.092 0.559 3.645

Page 120: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

108

Core 4 BiSi values are listed below:

Depth

(cm)

Batch

#

Weight

(mg)

Absorbance at

812 nm

mM

SiO2

Dilution

Factor

BiSi

(‰)

BiSi

(%)

2 5 80.02 0.126 0.8 0.032 24 2.4

4 4 79.4 0.113 0.7 0.029 22 2.2

6 4 80.54 0.113 0.7 0.029 21 2.1

8 5 80.66 0.101 0.6 0.026 19 1.9

10 5 79.36 0.120 0.8 0.030 23 2.3

12 4 79.39 0.129 0.8 0.033 25 2.5

14 5 80.39 0.112 0.7 0.028 21 2.1

16 4 80.78 0.105 0.7 0.027 20 2.0

18 3 80.5 0.096 0.6 0.024 18 1.8

20 3 79.74 0.079 0.5 0.020 15 1.5

22 4 80.05 0.09 0.6 0.023 17 1.7

24 3 80.52 0.077 0.5 0.020 15 1.5

26 3 79.96 0.09 0.6 0.023 17 1.7

28 5 80.12 0.084 0.5 0.021 16 1.6

30 3 79.52 0.073 0.5 0.019 14 1.4

32 4 79.41 0.075 0.5 0.019 14 1.4

34 5 80.53 0.058 0.4 0.015 11 1.1

36 4 79.37 0.109 0.7 0.028 21 2.1

40 5 80.54 0.160 1.0 0.041 30 3.0

44 5 79.42 0.201 1.3 0.051 39 3.9

48 4 80.26 0.195 1.2 0.050 37 3.7

52 3 80.12 0.209 1.3 0.053 40 4.0

56 3 80 0.215 1.4 0.055 41 4.1

60 5 80.15 0.199 1.3 0.051 38 3.8

64 5 79.49 0.216 1.4 0.055 41 4.1

68 3 80.5 0.179 1.1 0.045 34 3.4

72 4 80.29 0.203 1.3 0.052 39 3.9

76 5 80.46 0.191 1.2 0.049 36 3.6

80 5 79.46 0.153 1.0 0.039 29 2.9

84 4 79.6 0.151 1.0 0.038 29 2.9

Page 121: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

109

Core 4 carbon analyses. Negative total carbon (TC) and negative total organic carbon

(TOC) values were treated as ‘zero.’ Any TIC values that are ‘zero’ were aborted after 15

minutes of coulometer analysis.

Depth

(cm)

Mass TC

(g)

Mass

TIC (g)

TC

(ppm)

TIC

(ppm)

TOC

(ppm)

TOC

(%)

TIC

(%)

1 0.215 0.105 11729 0 11729 1.17 0.000

2 0.177 0.115 11518 0 11518 1.15 0.000

4 0.208 0.107 12197 0 12197 1.22 0.000

5 0.166 0.117 11806 49.14 11756.86 1.18 0.005

6 0.180 0.110 11912 0 11912 1.19 0.000

8 0.228 0.117 12602 0 12602 1.26 0.000

10 0.177 0.109 12385 34.36 12350.64 1.24 0.003

12 0.177 0.116 12599 38.19 12560.81 1.26 0.004

14 0.212 0.113 11954 0 11954 1.20 0.000

16 0.203 0.113 11414 0 11414 1.14 0.000

18 0.192 0.115 11932 0 11932 1.19 0.000

20 0.239 0.118 12661 35.11 12625.89 1.26 0.004

22 0.178 0.108 9871.7 0 9871.7 0.99 0.000

24 0.160 0.117 9454.9 0 9454.9 0.95 0.000

26 0.180 0.109 9445.8 0 9445.8 0.94 0.000

28 0.207 0.115 7611.2 14.1 7597.1 0.76 0.001

30 0.170 0.105 6826.4 30.3 6796.1 0.68 0.003

32 0.174 0.113 6629.7 9.67 6620.03 0.66 0.001

34 0.207 0.112 7155.3 22.92 7132.38 0.71 0.002

36 0.258 0.116 4094.3 0 4094.3 0.41 0.000

40 0.215 0.103 3630.6 28.17 3602.43 0.36 0.003

44 0.257 0.109 1405.4 23.83 1381.57 0.14 0.002

48 0.207 0.110 1231.7 5.21 1226.49 0.12 0.001

52 0.185 0.114 1121.1 20.83 1100.27 0.11 0.002

56 0.269 0.119 1420 13.66 1406.34 0.14 0.001

60 0.261 0.103 1217.7 6.35 1211.35 0.12 0.001

64 0.195 0.116 1092.5 22.09 1070.41 0.11 0.002

68 0.280 0.109 1189.3 5.87 1183.43 0.12 0.001

72 0.195 0.116 996.62 2.17 994.45 0.10 0.000

76 0.205 0.104 1034.9 9.19 1025.71 0.10 0.001

80 0.228 0.117 1041.5 10.21 1031.29 0.10 0.001

84 0.216 0.113 969.25 8.76 960.49 0.10 0.001

Page 122: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

110

Core 4 grain size measurements are listed below. The percentages of sand, silt, and clay

were adjusted if the sum was 99.99% (0.01% added to smallest value) or 100.01%

(0.01% subtracted from largest value).

Depth (cm) % Sand % Silt % Clay Skewness D50 (µm)

2 22.72 67.41 9.87 1.58 25.35

4 17.13 72.17 10.7 7.355 21.072

6 37.99 55.2 6.81 1.099 43.528

8 9.75 77.98 12.27 2.147 18.109

10 25.49 65.26 9.25 1.633 25.077

12 18.27 70.84 10.89 1.907 21.325

14 18.23 71.4 10.37 1.991 22.141

16 23.36 66.01 10.63 1.743 23.612

18 24.99 64.65 10.36 10.164 25.33

20 19.51 68.8 11.69 1.976 22.871

22 23.5 65.77 10.73 1.81 22.53

24 37.5 53.91 8.59 1.295 38.212

26 34.62 55.87 9.51 8.743 34.333

28 47.13 45.53 7.34 0.955 57.023

30 53.48 40.66 5.86 0.891 69.713

32 57.6 36.86 5.54 6.641 78.395

34 49.78 41.71 8.51 3.076 62.457

36 18.69 56.69 24.62 9.902 10.941

40 7.55 61.17 31.28 8.35 7.685

44 4.97 68.39 26.64 10.209 8.323

48 6.87 67.95 25.18 12.205 10.023

52 4.22 68.95 26.83 3.328 8.164

56 8.51 64.63 26.86 12.154 9.784

60 6.27 67.34 26.39 7.967 8.421

64 6.07 65.6 28.33 10.178 8.074

68 5.05 62.46 32.49 3.005 7.278

72 6.33 67.35 26.32 3.637 9.06

76 6.16 61.52 32.32 12.443 7.764

80 16.45 59.45 24.1 2.087 10.785

84 17.21 58.77 24.02 1.988 11.128

Page 123: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

111

Core 4 radiochemical data (Bq kg-1) measured at SERRL lab, University of Kentucky.

Interval (cm) 137Cs 226Ra 228Ra 210Pbtotal 210Pbxs

0-1 0 ± 0 41.42 ± 2.55 36.56 ± 2.19

1-2 0 ± 0 48.42 ± 2.79 44.58 ± 3.64 7.24 ± 0.46 4.46 ± 0.22

2-3 0 ± 0 39.48 ± 2.52 32.42 ± 2.71 7.65 ± 0.52 4.87 ± 0.28

3-4 0 ± 0 40.77 ± 2.73 41.68 ± 2.77 6.82 ± 0.42 4.04 ± 0.18

4-5 0 ± 0 37.17 ± 2.25 38.25 ± 3.38 6.38 ± 0.45 3.60 ± 0.21

5-6 0 ± 0 41.62 ± 3.04 41.70 ± 3.14 6.08 ± 0.47 3.30 ± 0.23

6-7 0 ± 0 46.18 ± 3.05 39.64 ± 2.78 6.83 ± 0.46 4.05 ± 0.22

7-8 0 ± 0 44.39 ± 2.78 38.25 ± 2.92 6.11 ± 0.41 3.33 ± 0.17

8-9 0 ± 0 35.44 ± 2.21 42.71 ± 3.54 6.53 ± 0.47 3.75 ± 0.23

9-10 0 ± 0 56.70 ± 3.57 42.26 ± 3.13 6.56 ± 0.44 3.78 ± 0.20

10-11

5.50 ± 0.58 2.72 ± 0.33

11-12

6.42 ± 0.41 3.64 ± 0.17

12-13

5.10 ± 0.35 2.32 ± 0.11

13-14

4.77 ± 0.35 1.99 ± 0.11

14-15 0 ± 0 53.42 ± 3.03 55.61 ± 4.20 4.23 ± 0.42 1.45 ± 0.18

15-16

16-17

17-18

4.16 ± 0.28 1.38 ± 0.04

18-19

19-20 2.77 ± 0.16 38.77 ± 2.24 34.60 ± 2.94

20-21

3.05 ± 0.34 0.27 ± 0.10

21-22

3.15 ± 0.34 0.37 ± 0.10

22-23

2.89 ± 0.21 0.11 ± 0.03

23-24

2.99 ± 0.33 0.21 ± 0.09

24-25

2.47 ± 0.18

25-26

2.14 ± 0.19

26-27

2.51 ± 0.22

27-28

1.96 ± 0.20

28-29

2.22 ± 0.16

29-30

1.88 ± 0.26

Page 124: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

112

GeoTek Multi-Sensor Core Logger (MSCL) physical properties plots for Core 1

measured at 0.5 cm intervals.

-80

-70

-60

-50

-40

-30

-20

-10

0

0 2 4

Gamma Density

0 1000

P-wave velocity

0 2000 4000

Impedance

-1 0 1 2

Fractional

Porosity

Page 125: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

113

Core 1 ED-XRF select major elemental abundances in percent, following Bruker’s

empirical conversion (Rowe et al., 2012).

Depth AlKa1 SiKa1 K Ka1 CaKa1 TiKa1 FeKa1

1 4.000 27.077 0.352 0.136 0.686 1.260

3 5.978 24.073 0.507 0.154 0.725 2.027

5 6.981 22.130 0.598 0.168 0.747 2.482

9 7.172 22.479 0.626 0.159 0.735 2.338

13 7.485 21.989 0.641 0.162 0.760 2.735

17 6.855 21.937 0.629 0.145 0.724 2.883

21 7.224 21.612 0.623 0.142 0.708 2.871

25 7.092 21.022 0.636 0.148 0.686 3.418

29 7.075 21.160 0.625 0.155 0.684 3.741

33 6.445 21.507 0.575 0.133 0.624 4.401

37 6.718 21.008 0.661 0.127 0.676 3.735

41 6.632 21.577 0.711 0.114 0.635 3.301

45 6.689 22.476 0.738 0.125 0.613 2.154

49 6.622 20.986 0.733 0.136 0.571 3.521

53 6.236 21.612 0.691 0.148 0.541 3.892

57 5.865 20.768 0.646 0.151 0.543 4.875

61 6.085 21.682 0.622 0.174 0.492 3.910

65 6.680 22.702 0.632 0.154 0.444 1.935

69 5.210 26.544 0.449 0.086 0.347 0.780

Page 126: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

114

Core 1 BiSi presented below:

Depth

(cm)

Batch

#

Weight

(mg)

Absorbance at

812 nm

mM

SiO2

Dilution

Factor

BiSi

(‰)

BiSi

(%)

1 5 79.83 0.116 0.7 0.029 22 2.2

3 5 81.34 0.106 0.7 0.027 20 2.0

5 5 79.76 0.128 0.8 0.033 24 2.4

9 5 80.6 0.125 0.8 0.032 24 2.4

13 4 80.03 0.08 0.5 0.020 15 1.5

17 3 80.84 0.131 0.8 0.033 25 2.5

21 4 80.75 0.138 0.9 0.035 26 2.6

25 4 80.84 0.143 0.9 0.036 27 2.7

29 3 79.71 0.154 1.0 0.039 29 2.9

33 3 79.06 0.141 0.9 0.036 27 2.7

37 4 80.39 0.131 0.8 0.033 25 2.5

41 5 79.8 0.148 0.9 0.038 28 2.8

45 4 80.46 0.128 0.8 0.033 24 2.4

49 4 79.05 0.139 0.9 0.035 27 2.7

53 5 79.47 0.146 0.9 0.037 28 2.8

57 5 79.19 0.131 0.8 0.033 25 2.5

61 3 79.67 0.126 0.8 0.032 24 2.4

65 3 80.08 0.109 0.7 0.028 21 2.1

69 5 80.57 0.099 0.6 0.025 19 1.9

Page 127: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

115

Core 1 carbon analyses. Negative total carbon (TC) and negative total organic carbon

(TOC) values were treated as ‘zero.’ Any TIC values that are ‘zero’ were aborted after 15

minutes of coulometer analysis.

Depth

(cm)

Mass TC

(g)

Mass

TIC (g)

TC

(ppm)

TIC

(ppm)

TOC

(ppm)

TOC

(%)

TIC

(%)

1 0.175 0.106 14329 26.79 14302.21 1.43 0.003

3 0.217 0.114 7181.4 35.24 7146.16 0.71 0.004

5 0.269 0.111 3329.8 6.26 3323.54 0.33 0.001

9 0.230 0.105 2823.8 0 2823.8 0.28 0.000

13 0.196 0.113 2301.8 11.34 2290.46 0.23 0.001

17 0.241 0.117 2145.6 12.44 2133.16 0.21 0.001

21 0.201 0.105 2037.2 15.19 2022.01 0.20 0.002

25 0.182 0.139 1910.9 21.84 1889.06 0.19 0.002

29 0.320 0.110 1872.1 0 1872.1 0.19 0.000

33 0.181 0.119 1398.1 0 1398.1 0.14 0.000

37 0.189 0.120 1233.5 23.76 1209.74 0.12 0.002

41 0.221 0.100 808.49 24.06 784.43 0.08 0.002

45 0.213 0.108 1154.1 0 1154.1 0.12 0.000

49 0.230 0.117 1020.6 24.43 996.17 0.10 0.002

53 0.200 0.109 657.42 0 657.42 0.07 0.000

57 0.191 0.118 878 30.72 847.28 0.08 0.003

61 0.179 0.111 2201.2 23.3 2177.9 0.22 0.002

65 0.177 0.104 221.96 27.19 194.77 0.02 0.003

69 0.193 0.126 88.361 13.23 75.131 0.01 0.001

Page 128: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

116

Core 1 grain size measurements listed below. The percentages of sand, silt, and clay were

adjusted if the sum was 99.99% (0.01% added to smallest value) or 100.01% (0.01%

subtracted from largest value).

Depth (cm) % Sand % Silt % Clay Skewness D50 (µm)

1 18.51 70.58 10.91 2.049 29.575

3 12.27 64.94 22.79 13.065 13.065

5 8.05 63.42 28.53 11.875 9.526

9 5.68 63.67 30.65 2.994 8.015

13 8.58 57.11 34.31 7.317 7.85

17 9.18 58.52 32.3 8.71 8.754

21 6.46 57.32 36.22 9.517 7.019

25 10.22 54.04 35.74 7.183 7.183

29 10.63 56.11 33.26 4.893 7.763

33 9.67 58.42 31.91 8.551 7.812

37 21.36 52.61 26.03 14.506 14.506

41 7.66 62.9 29.44 2.154 9.199

45 16.54 61.39 22.07 9.215 15.969

49 14.42 60.36 25.22 11.893 12.256

53 24.41 56.1 19.49 8.056 22.769

57 13.34 62.28 24.38 1.984 13.501

61 20.2 57.97 21.83 9.2 15.679

65 35 51.04 13.96 6.506 37.461

69 62.57 29.63 7.8 0.625 86.032

Page 129: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

117

Core 1 radiochemical data (Bq kg-1) measured in the SERRL lab, University of

Kentucky.

Interval (cm) 137Cs 226Ra 228Ra 210Pbtotal 210Pbxs

0-1 0 ± 0 42.15 ± 2.65 67.08 ± 3.86 39.81 ± 2.30 19.74 ± 1.08

1-2 0 ± 0 51.04 ± 3.34 45.68 ± 2.70 33.75 ± 2.04 13.69 ± 0.82

2-3 0 ± 0 39.70 ± 2.83 59.53 ± 3.33 25.52 ± 1.44 5.45 ± 0.22

3-4 0 ± 0 69.27 ± 4.22 58.59 ± 3.43 28.64 ± 1.74 8.58 ± 0.52

4-5 0 ± 0 57.21 ± 3.40 56.43 ± 3.86 31.91 ± 1.96 11.85 ± 0.73

5-6 0 ± 0 56.72 ± 3.55 70.98 ± 4.17 33.34 ± 2.02 13.27 ± 0.80

6-7 0 ± 0 58.70 ± 3.97 57.19 ± 4.43 29.79 ± 1.96 9.73 ± 0.73

7-8 0 ± 0 44.00 ± 3.30 68.37 ± 4.82 31.41 ± 2.09 11.35 ± 0.87

8-9 0 ± 0 71.68 ± 4.55 69.33 ± 5.14 35.59 ± 2.33 15.53 ± 1.11

9-10 0 ± 0 63.44 ± 3.89 43.16 ± 3.51 22.87 ± 1.36 2.80 ± 0.14

10-11

33.38 ± 2.20 13.31 ± 0.97

11-12

30.35 ± 1.85 10.29 ± 0.62

12-13

34.08 ± 2.28 14.02 ± 1.05

13-14

30.35 ± 1.85 10.29 ± 0.62

14-15 0 ± 0 72.38 ± 4.74 64.97 ± 4.81 30.03 ± 1.98 9.97 ± 0.75

15-16

32.59 ± 2.12 12.53 ± 0.90

16-17

29.95 ± 1.92 9.89 ± 0.69

17-18

28.41 ± 1.53 8.34 ± 0.31

18-19

25.82 ± 1.69 5.76 ± 0.46

19-20 0 ± 0 76.32 ± 4.96 64.75 ± 5.18 32.44 ± 2.12 12.38 ± 0.89

20-21

30.16 ± 1.76 10.10 ± 0.53

21-22

28.14 ± 1.82 8.08 ± 0.59

22-23

25.32 ± 1.39 5.26 ± 0.17

23-24

26.43 ± 1.76 6.37 ± 0.53

24-25 0 ± 0 58.48 ± 4.10 48.88 ± 3.97 28.64 ± 1.71 8.57 ± 0.48

25-26

28.14 ± 1.79 8.07 ± 0.56

26-27

28.37 ± 1.83 8.30 ± 0.60

27-28

30.73 ± 1.98 10.66 ± 0.76

28-29

29.49 ± 1.90 9.43 ± 0.67

29-30 0 ± 0 53.29 ± 3.11 78.80 ± 4.34 16.62 ± 1.01

30-31

22.51 ± 1.33 2.45 ± 0.10

31-32

19.38 ± 1.14

32-33

22.42 ± 1.38 2.36 ± 0.16

33-34

19.83 ± 1.19

34-35

20.00 ± 1.25

35-36

20.21 ± 1.13

36-37

21.02 ± 1.39

37-38

38-39

19.02 ± 1.13

39-40

Page 130: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

118

GeoTek Multi-Sensor Core Logger (MSCL) physical properties plots for Core 2

measured at 0.5 cm intervals.

-60

-50

-40

-30

-20

-10

0

0 2000

P-wave

Velocity

0 2

Gamma

Density

0 3000

Impedance

0 2

Fractional

Porosity

0 1 2 3 4

Magnetic

Susceptibility

Page 131: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

119

Core 2 ED-XRF select major elemental abundances in percent, following Bruker’s

empirical conversion (Rowe et al., 2012).

ID AlKa1 SiKa1 K Ka1 CaKa1 TiKa1 FeKa1

1 3.024 28.468 0.259 0.096 0.555 1.012

3 5.978 24.073 0.507 0.154 0.725 2.027

5 2.822 30.510 0.342 0.088 0.490 0.251

7 3.055 29.359 0.268 0.118 0.526 0.294

9 2.920 29.771 0.269 0.088 0.489 0.187

11 2.941 29.454 0.211 0.106 0.526 0.211

13 3.326 28.989 0.287 0.129 0.502 0.322

15 3.512 27.360 0.226 0.205 0.531 0.609

17 4.145 23.340 0.242 0.366 0.579 1.012

19 4.413 21.298 0.232 0.494 0.591 1.619

21 4.594 22.212 0.245 0.461 0.620 1.511

23 4.571 22.311 0.246 0.464 0.652 1.603

25 5.182 21.555 0.282 0.416 0.647 1.740

27 5.300 21.386 0.294 0.422 0.670 1.685

29 4.776 21.739 0.261 0.384 0.626 1.589

31 4.907 21.697 0.259 0.380 0.641 1.652

33 4.579 21.418 0.247 0.394 0.636 1.666

35 4.745 21.772 0.259 0.388 0.659 1.692

37 5.653 21.925 0.229 0.339 0.656 1.681

39 6.027 22.609 0.207 0.308 0.696 1.412

41 6.042 23.560 0.193 0.264 0.645 1.454

43 6.657 22.641 0.229 0.275 0.702 1.455

45 6.873 21.349 0.243 0.277 0.704 1.507

47 7.716 21.539 0.257 0.270 0.706 1.512

49 7.726 21.400 0.247 0.271 0.717 1.522

51 7.597 20.993 0.254 0.255 0.692 1.541

53 7.756 22.221 0.208 0.229 0.775 1.144

Page 132: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

120

Core 2 BiSi values presented below:

Depth

(cm)

Batch

#

Weight

(mg)

Absorbance at

812 nm

mM

SiO2

Dilution

Factor

BiSi

(‰)

BiSi

(%)

1 4 79.73 0.088 0.6 0.022 17 1.7

3 3 80.45 0.089 0.6 0.023 17 1.7

5 3 79.85 0.086 0.5 0.022 16 1.6

7 4 80.78 0.111 0.7 0.028 21 2.1

9 5 79.91 0.097 0.6 0.025 19 1.9

11 5 80.7 0.101 0.6 0.026 19 1.9

13 4 80.22 0.138 0.9 0.035 26 2.6

15 5 79.72 0.142 0.9 0.036 27 2.7

17 4 80.19 0.23 1.5 0.058 44 4.4

19 3 80.42 0.256 1.6 0.065 49 4.9

21 4 80.38 0.262 1.7 0.067 50 5.0

23 5 80.18 0.262 1.7 0.067 50 5.0

25 3 80.34 0.248 1.6 0.063 47 4.7

27 5 80.73 0.235 1.5 0.060 44 4.4

29 4 80.63 0.252 1.6 0.064 48 4.8

31 3 79.84 0.253 1.6 0.064 48 4.8

33 4 79.86 0.25 1.6 0.064 48 4.8

35 4 80.81 0.251 1.6 0.064 47 4.7

37 4 80.68 0.265 1.7 0.067 50 5.0

39 3 80.25 0.303 1.9 0.077 58 5.8

41 4 80.92 0.331 2.1 0.084 62 6.2

43 5 80.5 0.231 1.5 0.059 44 4.4

45 5 80.74 0.099 0.6 0.025 19 1.9

47 5 80.05 0.192 1.2 0.049 37 3.7

49 3 80.06 0.239 1.5 0.061 46 4.6

51 4 80.47 0.216 1.4 0.055 41 4.1

53 5 79.76 0.163 1.0 0.041 31 3.1

Page 133: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

121

Core 2 carbon analyses. Negative total carbon (TC) and negative total organic carbon

(TOC) values were treated as ‘zero.’ Any TIC values that are ‘zero’ were aborted after 15

minutes of coulometer analysis.

Depth

(cm)

Mass TC

(g)

Mass

TIC (g)

TC

(ppm)

TIC

(ppm)

TOC

(ppm)

TOC

(%)

TIC

(%)

1 0.166 0.103 18484 0 18484 1.85 0.000

3 0.297 0.124 17205 0 17205 1.72 0.000

5 0.212 0.139 17469 40.2 17428.8 1.74 0.004

7 0.273 0.116 27664 0 27664 2.77 0.000

9 0.151 0.106 19915 0 19915 1.99 0.000

11 0.288 0.124 22222 0 22222 2.22 0.000

13 0.255 0.139 23766 0 23766 2.38 0.000

15 0.190 0.128 51368 0 51368 5.14 0.000

17 0.235 0.115 90505 0 90505 9.05 0.000

19 0.156 0.110 147040 26.82 147013.2 14.70 0.003

21 0.214 0.120 125490 0 125490 12.55 0.000

23 0.225 0.103 134630 12.65 134617.4 13.46 0.001

25 0.207 0.121 107350 0 107350 10.74 0.000

27 0.162 0.118 109620 0 109620 10.96 0.000

29 0.187 0.114 114540 41.78 114498.2 11.45 0.004

31 0.208 0.111 109020 20.43 108999.6 10.90 0.002

33 0.212 0.112 117050 0 117050 11.71 0.000

35 0.150 0.122 124250 0 124250 12.43 0.000

37 0.225 0.104 94432 0 94432 9.44 0.000

39 0.208 0.107 75632 0 75632 7.56 0.000

41 0.153 0.107 63765 0 63765 6.38 0.000

43 0.192 0.100 58051 0 58051 5.81 0.000

45 0.192 0.113 60649 55.61 60593.39 6.06 0.006

47 0.203 0.111 36567 0 36567 3.66 0.000

49 0.251 0.107 38201 0 38201 3.82 0.000

51 0.187 0.119 31967 43.67 31923.33 3.19 0.004

53 0.192 0.116 17892 0 17892 1.79 0.000

Page 134: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

122

Core 2 grain size measurements listed below. The percentages of sand, silt, and clay were

adjusted if the sum was 99.99% (0.01% added to smallest value) or 100.01% (0.01%

subtracted from largest value).

Depth (cm) % Sand % Silt % Clay Skewness D50 (µm)

1 30.32 64.84 4.84 0.556 46.798

3 33.6 61.43 4.97 0.454 49.97

5 37.64 58.79 3.57 0.448 53.6

7 38.13 57.69 4.18 0.43 53.865

9 42.26 54.62 3.12 0.43 57.281

11 39.5 56.32 4.18 0.464 54.933

13 34.71 59.97 5.32 0.548 50.615

15 23.51 67.29 9.2 5.552 38.199

17 23.28 65.89 10.83 8.907 35.083

19 12.88 68.58 18.54 5.451 17.914

21 14.89 67.82 17.29 7.825 22.389

23 8.87 71.07 20.06 9.731 15.023

25 11.09 67.64 21.27 7.157 15.571

27 8.11 68.62 23.27 8.995 15.054

29 9.33 68.9 21.77 9.653 13.813

31 9.97 67.76 22.27 9.073 16.233

33 9.79 69.01 21.2 8.766 17.132

35 9 66.02 24.98 9.256 10.872

37 17.8 60.89 21.31 6.101 17.707

39 5.72 65.99 28.29 7.565 9.049

41 5.82 69.73 24.45 4.439 9.691

43 7.55 67.05 25.4 7.871 10.011

45 9.56 59.77 30.67 4.217 8.485

47 9.18 59.75 31.07 8.495 8.904

49 5.89 63.5 30.61 2.921 9.137

51 9.84 65.14 25.02 8.575 12.019

53 8.15 68.48 23.37 10.113 12.688

Page 135: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

123

Core 2 radiochemical data (Bq kg-1) measured in the SERRL lab, University of

Kentucky.

Interval (cm) 137Cs 226Ra 228Ra 210Pbtotal 210Pbxs

0-1 0 ± 0 36.31 ± 2.18 29.36 ± 2.03 34.28 ± 2.23 19.60 ± 1.27

1-2 0 ± 0 31.27 ± 1.98 25.17 ± 1.51 32.14 ± 2.00 17.47 ± 1.04

2-3 0 ± 0 29.09 ± 2.09 18.37 ± 1.46 29.33 ± 1.91 14.66 ± 0.95

3-4 0 ± 0 21.83 ± 1.54 29.59 ± 1.89 24.77 ± 1.62 10.10 ± 0.66

4-5 0 ± 0 34.80 ± 2.54 19.60 ± 1.22 24.44 ± 1.55 9.77 ± 0.59

5-6 0 ± 0 32.75 ± 2.16 25.76 ± 2.19 22.23 ± 1.47 7.55 ± 0.51

6-7 0 ± 0 24.78 ± 1.96 19.65 ± 1.13 19.79 ± 1.25 5.12 ± 0.29

7-8 0 ± 0 27.94 ± 1.79 27.46 ± 1.85 15.71 ± 0.97 1.04 ± 0.01

8-9 0 ± 0 39.85 ± 2.52 32.16 ± 2.13 16.01 ± 1.01 1.34 ± 0.05

9-10 0 ± 0 29.49 ± 1.86 21.81 ± 1.35 16.37 ± 0.95 1.70 ± 0.01

10-11

16.81 ± 1.02 2.14 ± 0.06

11-12

16.85 ± 1.07 2.18 ± 0.11

12-13

16.65 ± 1.04 1.98 ± 0.08

13-14

15.89 ± 0.98 1.22 ± 0.02

14-15 0 ± 0 28.29 ± 1.83 32.03 ± 1.96 19.20 ± 1.05 4.53 ± 0.09

15-16

23.44 ± 1.45 8.77 ± 0.49

16-17

25.99 ± 1.61 11.31 ± 0.65

17-18

26.79 ± 1.62 12.12 ± 0.66

18-19

19.44 ± 1.27 4.77 ± 0.31

19-20

16.33 ± 1.06 1.66 ± 0.10

20-21

14.82 ± 0.95

21-22

15.13 ± 0.99

22-23

16.93 ± 1.00 2.26 ± 0.04

23-24

15.89 ± 1.03 1.22 ± 0.07

24-25

16.17 ± 1.03 1.50 ± 0.07

25-26

11.85 ± 0.76

26-27

12.13 ± 0.81

27-28

15.40 ± 0.97

28-29

16.43 ± 1.04 1.76 ± 0.08

29-30

17.56 ± 1.22 2.89 ± 0.26

30-31

40.69 ± 2.82

31-32

50.54 ± 3.29

32-33

48.24 ± 3.39

33-34

52.47 ± 3.82

34-35

52.68 ± 3.15

35-36

54.58 ± 3.54

36-37

50.26 ± 2.82

37-38

43.26 ± 2.84

38-39

51.90 ± 3.30

39-40

62.09 ± 2.85

40-41 55.57 ± 3.17

Page 136: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

124

41-42 49.99 ± 3.25

42-43 46.20 ± 3.06

43-44 47.99 ± 2.51

44-45 52.27 ± 2.71

45-46 50.93 ± 2.69

46-47 43.39 ± 2.16

47-48 51.86 ± 3.49

48-49 52.88 ± 3.03

49-50 46.73 ± 3.32

Page 137: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

125

Biogenic silica duplicate (Dup) and standards measurements are listed:

Sample ID Batch

#

Depth

(cm)

Weight

(mg)

Absorbance

at 812 nm

mM

SiO2

Dilution

Factor

BiSi

(‰)

BiSi

(%)

4A-60 Dup 5 60 79.33 0.214 1.4 0.054 41 4.1

1A - 65 Dup 5 65 80 0.079 0.5 0.020 15 1.5

2A - 23 Dup 4 23 79.92 0.272 1.7 0.069 52 5.2

2A - 37 Dup 5 37 80.26 0.292 1.9 0.074 56 5.6

2A - 39 Dup 3 39 79.54 0.294 1.9 0.075 56 5.6

3A - 14 Dup 4 14 79.21 0.069 0.4 0.018 13 1.3

3A - 22 Dup 5 22 80.15 0.079 0.5 0.020 15 1.5

3A - 50 Dup 5 50 79.56 0.099 0.6 0.025 19 1.9

4A - 28 Dup 4 28 80.32 0.068 0.4 0.017 13 1.3

4A - 4 Dup 3 4 79.63 0.124 0.8 0.031 24 2.4

4A - 76 Dup 3 76 80.72 0.194 1.2 0.049 37 3.7

4A -68 Dup 4 68 80.66 0.223 1.4 0.057 42 4.2

4A- 84 Dup 3 84 80.31 0.137 0.9 0.035 26 2.6

Standard-a 3

46.24 0.620 3.9 0.157 205 20.5

Standard-b 5

45.78 0.600 3.8 0.152 200 20.0

Standard-c 5

46.79 0.561 3.6 0.142 183 18.3

Page 138: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

126

REFERENCES

Aalto R, Nittrouer CA. 2012. 210Pb geochronology of flood events in large tropical river

systems. Phil. Trans. R. Soc. A 370: 2040-2074. DOI: 10.1098/rsta.2011.0607

Abbott MB, Binford MW, Brenner M, Kelts KR. 1997. A 350014C yr High-Resolution

Record of Water-Level Changes in Lake Titicaca, Bolivia/Peru. Quaternary

Research 47: 169-180. DOI: 10.1006/qres.1997.1881

Aitken MJ. 1998. Introduction to optical dating: the dating of Quaternary sediments by

the use of photon-stimulated luminescence. Oxford University Press: London

Alaamer H. 2015. Characteristics of sub-bottom profile acquired in Shatt al-Arab River,

Basrah-Iraq. International Journal of Marine Science 5: 1-6. DOI:

10.5376/ijms.2015.05.0008

Al-Ghadban A, Jacob P, Abdali F. 1994. Total organic carbon in the sediments of the

Arabian Gulf and need for biological productivity investigations. Marine

Pollution Bulletin 28: 356-362. DOI: 10.1016/0025-326X(94)90272-0

Almeida TIR, Calijuri MC, Falco PB, Casali SP, Kupriyanova E, Paranhos Filho AC,

Sigolo JB, Bertolo RA. 2011. Biogeochemical processes and the diversity of

Nhecolândia lakes, Brazil. Anais da Academia Brasileira de Ciências 83: 391-

407. DOI: 10.1590/S0001-37652011000200004

Andersen TJ. 2017. Some Practical Considerations Regarding the Application of 210Pb

and 137Cs Dating to Estuarine Sediments. In Applications of Paleoenvironmental

Techniques in Estuarine Studies, Weckström K, Saunders KM, Gell PA, Skilbeck

CG (eds). Springer Netherlands: Dordrecht; 121-140. DOI: 10.1007/978-94-024-

0990-1_6

Appleby P, Flower R, Mackay A, Rose N. 1998. Paleolimnological assessment of recent

environmental change in Lake Baikal: sediment chronology. Journal of

Paleolimnology 20: 119-133. DOI: 10.1023/A:1008099631458

Appleby P, Oldfield F. 1978. The calculation of lead-210 dates assuming a constant rate

of supply of unsupported 210Pb to the sediment. CATENA 5: 1-8. DOI:

10.1016/S0341-8162(78)80002-2

Arzani N. 2012. Catchment lithology as a major control on alluvial megafan

development, Kohrud Mountain range, central Iran. Earth Surface Processes and

Page 139: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

127

Landforms 37: 726-740. DOI: 10.1002/esp.3194

Ashworth PJ, Lewin J. 2012. How do big rivers come to be different? Earth-Science

Reviews 114: 84-107. DOI: 10.1016/j.earscirev.2012.05.003

Assine ML. 2003. Sedimentação na bacia do Pantanal mato-grossense, centro-oeste do

Brasil.

Assine ML. 2005. River avulsions on the Taquari megafan, Pantanal wetland, Brazil.

Geomorphology 70: 357-371. DOI: 10.1016/j.geomorph.2005.02.013

Assine ML, Corradini FA, do Nascimento Pupim F, McGlue MM. 2014. Channel

arrangements and depositional styles in the São Lourenço fluvial megafan,

Brazilian Pantanal wetland. Sedimentary Geology 301: 172-184. DOI:

10.1016/j.sedgeo.2013.11.007

Assine ML, Macedo HA, Stevaux JC, Bergier I, Padovani CR, Silva A. 2015b. Avulsive

rivers in the hydrology of the pantanal wetland. In Dynamics of the Pantanal

Wetland in South America. Springer; 83-110. DOI: 10.1007/698_2015_351

Assine ML, Merino ER, Pupim FN, Macedo HA, Santos MGM. 2015a. The Quaternary

alluvial systems tract of the Pantanal Basin, Brazil. Brazilian Journal of Geology

45: 475-489. DOI: 10.1590/2317-4889201520150014

Assine ML, Silva A. 2009. Contrasting fluvial styles of the Paraguay River in the

northwestern border of the Pantanal wetland, Brazil. Geomorphology 113: 189-

199. DOI: 10.1016/j.geomorph.2009.03.012

Assine ML, Soares PC. 2004. Quaternary of the Pantanal, west-central Brazil. Quaternary

International 114: 23-34. DOI: 10.1016/S1040-6182(03)00039-9

Barbier EB. 1994. Valuing environmental functions: tropical wetlands. Land economics:

155-173. DOI: 10.2307/3146319

Barbiero L, Rezende Filho A, Furquim S, Furian S, Sakamoto A, Valles V, Graham R,

Fort M, Ferreira RD, Neto JQ. 2008. Soil morphological control on saline and

freshwater lake hydrogeochemistry in the Pantanal of Nhecolândia, Brazil.

Geoderma 148: 91-106. DOI: 10.1016/j.geoderma.2008.09.010

Baskaran M, Nix J, Kuyper C, Karunakara N. 2014. Problems with the dating of

sediment core using excess 210Pb in a freshwater system impacted by large scale

watershed changes. Journal of environmental radioactivity 138: 355-363. DOI:

Page 140: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

128

10.1016/j.jenvrad.2014.07.006

Bastviken D, Santoro AL, Marotta H, Pinho LQ, Calheiros DF, Crill P, Enrich-Prast A.

2010. Methane emissions from Pantanal, South America, during the low water

season: toward more comprehensive sampling. Environmental science &

technology 44: 5450-5455. DOI: 10.1021/es1005048

Bergier I, Resende Ed. 2010. Dinâmica de cheias no Pantanal do rio Paraguai de 1900 a

2009. Geopantanal. INPE/Embrapa, Cáceres, MT, Brazil: 35-43. DOI:

10.13140/2.1.2656.9289

Bergier I, Salis SM, Miranda CH, Ortega E, Luengo CA. 2012. Biofuel production from

water hyacinth in the Pantanal wetland. Ecohydrology & Hydrobiology 12: 77-84.

DOI: 10.2478/v10104-011-0041-4

Bernal JP, Cruz FW, Stríkis NM, Wang X, Deininger M, Catunda MCA, Ortega-Obregón

C, Cheng H, Edwards RL, Auler AS. 2016. High-resolution Holocene South

American monsoon history recorded by a speleothem from Botuverá Cave, Brazil.

Earth and Planetary Science Letters 450: 186-196. DOI:

10.1016/j.epsl.2016.06.008

Bertassoli Jr DJ, Sawakuchi AO, Sawakuchi HO, Pupim FN, Hartmann GA, McGlue

MM, Chiessi CM, Zabel M, Schefuß E, Pereira TS. 2017. The fate of carbon in

sediments of the Xingu and Tapajós clearwater rivers, eastern Amazon. Frontiers

in Marine Science 4: 44. DOI: 10.3389/fmars.2017.00044

Bezerra MAO, Mozeto AA. 2008. Organic carbon deposition in the Paraguay River

floodplain during the mid-Holocene. Oecologia Australis 12: 155-171

Binford MW. 1990. Calculation and uncertainty analysis of 210 Pb dates for PIRLA

project lake sediment cores. Journal of Paleolimnology 3: 253-267. DOI:

10.1007/BF00219461

Bird BW, Abbott MB, Rodbell DT, Vuille M. 2011. Holocene tropical South American

hydroclimate revealed from a decadally resolved lake sediment δ 18 O record.

Earth and Planetary Science Letters 310: 192-202. DOI:

10.1016/j.epsl.2011.08.040

Boers N, Bookhagen B, Marwan N, Kurths J. 2016. Spatiotemporal characteristics and

synchronization of extreme rainfall in South America with focus on the Andes

Page 141: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

129

Mountain range. Climate Dynamics 46: 601-617. DOI: 10.1007/s00382-015-

2601-6

Bohacs KM, Carroll AR, Neal JE. 2003. Lessons from large lake systems-thresholds,

nonlinearity, and strange attractors. Special Papers-Geological Society of

America: 75-90. DOI: 10.1130/0-8137-2370-1.75

Bohacs KM, Carroll AR, Neal JE, Mankiewicz PJ. 2000. Lake-basin type, source

potential, and hydrocarbon character: an integrated sequence-stratigraphic-

geochemical framework. Lake basins through space and time: AAPG Studies in

Geology 46: 3-34

Boix-Fayos C, Nadeu E, Quiñonero J, Martínez-Mena M, Almagro M, De Vente J. 2015.

Sediment flow paths and associated organic carbon dynamics across a

Mediterranean catchment. Hydrology and Earth System Sciences 19: 1209-1223.

DOI: 10.5194/hess-19-1209-2015

Bousquet P, Ringeval B, Pison I, Dlugokencky E, Brunke E-G, Carouge C, Chevallier F,

Fortems-Cheiney A, Frankenberg C, Hauglustaine D. 2011. Source attribution of

the changes in atmospheric methane for 2006–2008. Atmospheric Chemistry and

Physics 11: 3689-3700. DOI: 10.5194/acp-11-3689-2011

Bridge JS. 1993. Description and interpretation of fluvial deposits: a critical perspective.

Sedimentology 40: 801-810. DOI: 10.1111/j.1365-3091.1993.tb01361.x

Byrne GF, Crapper PF, Mayo KK. 1980. Monitoring land-cover change by principal

component analysis of multitemporal landsat data. Remote Sensing of

Environment 10: 175-184. DOI: http://dx.doi.org/10.1016/0034-4257(80)90021-8

Cao M, Gregson K, Marshall S. 1998. Global methane emission from wetlands and its

sensitivity to climate change. Atmospheric Environment 32: 3293-3299. DOI:

10.1016/S1352-2310(98)00105-8

Carrivick JL, Tweed FS. 2013. Proglacial lakes: character, behaviour and geological

importance. Quaternary Science Reviews 78: 34-52. DOI:

10.1016/j.quascirev.2013.07.028

Carroll AR, Bohacs KM. 1999. Stratigraphic classification of ancient lakes: Balancing

tectonic and climatic controls. Geology 27: 99-102. DOI: 10.1130/0091-

7613(1999)027<0099:SCOALB>2.3.CO;2

Page 142: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

130

Carvalho LM, Jones C, Liebmann B. 2004. The South Atlantic convergence zone:

Intensity, form, persistence, and relationships with intraseasonal to interannual

activity and extreme rainfall. Journal of Climate 17: 88-108. DOI: 10.1175/1520-

0442(2004)017<0088:TSACZI>2.0.CO;2

Chakraborty T, Ghosh P. 2010. The geomorphology and sedimentology of the Tista

megafan, Darjeeling Himalaya: implications for megafan building processes.

Geomorphology 115: 252-266. DOI: 10.1016/j.geomorph.2009.06.035

Chase C, Sussman A, Coblentz D. 2009. Curved Andes: Geoid, forebulge, and flexure.

Lithosphere 1: 358-363. DOI: 10.1130/L67.1

Chen J, Wan G, Zhang DD, Zhang F, Huang R. 2004. Environmental records of

lacustrine sediments in different time scales: Sediment grain size as an example.

Science in China Series D: Earth Sciences 47: 954-960. DOI: 10.1360/03yd0160

Citterio A, Piégay H. 2009. Overbank sedimentation rates in former channel lakes:

characterization and control factors. Sedimentology 56: 461-482. DOI:

10.1111/j.1365-3091.2008.00979.x

Cohen A, McGlue MM, Ellis GS, Zani H, Swarzenski PW, Assine ML, Silva A. 2015.

Lake formation, characteristics, and evolution in retroarc deposystems: A

synthesis of the modern Andean orogen and its associated basins. Geological

Society of America Memoirs 212. DOI: 10.1130/2015.1212(16)

Cohen AS. 2003. Paleolimnology: the history and evolution of lake systems. Oxford

University Press: New York

Cole MM. 1960. Cerrado, Caatinga and Pantanal: The Distribution and Origin of the

Savanna Vegetation of Brazil. The Geographical Journal 126: 168-179. DOI:

10.2307/1793957

Costa MLR, Henry R. 2010. Phosphorus, nitrogen, and carbon contents of macrophytes

in lakes lateral to a tropical river (Paranapanema River, São Paulo, Brazil). Acta

Limnologica Brasiliensia 22: 122-132. DOI: 10.4322/actalb.02202002

Davidson SK, Hartley AJ, Weissmann GS, Nichols GJ, Scuderi LA. 2013. Geomorphic

elements on modern distributive fluvial systems. Geomorphology 180–181: 82-

95. DOI: 10.1016/j.geomorph.2012.09.008

Dean WE, Gorham E. 1998. Magnitude and significance of carbon burial in lakes,

Page 143: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

131

reservoirs, and peatlands. Geology 26: 535-538. DOI: 10.1130/0091-

7613(1998)026<0535:MASOCB>2.3.CO;2

Dias FL, Assumpção M, Facincani EM, França GS, Assine ML, Paranhos Filho AC,

Gamarra RM. 2016. The 2009 earthquake, magnitude mb 4.8, in the Pantanal

Wetlands, west-central Brazil. Anais da Academia Brasileira de Ciências 88:

1253-1264. DOI: 10.1590/201620140507

Dubiel RF, Skipp G, Hasiotis ST. 1992. Continental depositional environments and

tropical paleosols in the Upper Triassic Chinle Formation, Eagle Basin, western

Colorado. Rocky Mountain Section (SEPM).

Eakins JD, Morrison RT. 1978. A new procedure for the determination of lead-210 in

lake and marine sediments. The International Journal of Applied Radiation and

Isotopes 29: 531-536. DOI: 10.1016/0020-708X(78)90161-8

Enters D, Kirilova E, Lotter A, Lücke A, Parplies J, Jahns S, Kuhn G, Zolitschka B.

2010. Climate change and human impact at Sacrower See (NE Germany) during

the past 13,000 years: a geochemical record. Journal of Paleolimnology 43: 719-

737. DOI: 10.1007%2Fs10933-009-9362-3

Fielding CR, Ashworth PJ, Best JL, Prokocki EW, Smith GHS. 2012. Tributary,

distributary and other fluvial patterns: What really represents the norm in the

continental rock record? Sedimentary Geology 261–262: 15-32. DOI:

10.1016/j.sedgeo.2012.03.004. DOI: 10.1038/nature06825

Fischer H, Behrens M, Bock M, Richter U, Schmitt J, Loulergue L, Chappellaz J, Spahni

R, Blunier T, Leuenberger M. 2008. Changing boreal methane sources and

constant biomass burning during the last termination. Nature 452: 864. DOI:

10.1038/nature06825

Formolo MJ, Lyons TW. 2007. Accumulation and preservation of reworked marine

pyrite beneath an oxygen-rich Devonian atmosphere: Constraints from sulfur

isotopes and framboid textures. Journal of Sedimentary Research 77: 623-633.

DOI: 10.2110/jsr.2007.062

Fornace KL, Hughen KA, Shanahan TM, Fritz SC, Baker PA, Sylva SP. 2014. A 60,000-

year record of hydrologic variability in the Central Andes from the hydrogen

isotopic composition of leaf waxes in Lake Titicaca sediments. Earth and

Page 144: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

132

Planetary Science Letters 408: 263-271. DOI: 10.1016/j.epsl.2014.10.024

Fornace KL, Whitney BS, Galy V, Hughen KA, Mayle FE. 2016. Late Quaternary

environmental change in the interior South American tropics: new insight from

leaf wax stable isotopes. Earth and Planetary Science Letters 438: 75-85. DOI:

10.1016/j.epsl.2016.01.007

Franco M, Pinheiro R. 1982. Geomorfologia, levantamento de recursos naturais: 161-

224, Vol. 27 & Folhas SE-20, 21; Corumbá. RADAMBRASIL, Ministério das

Minas e Energia: Rio de Janeiro.

Fritz SC, Metcalfe SE, Dean W. 2001. Holocene climate patterns in the Americas

inferred from paleolimnological records. Interhemispheric Climate Linkage.

Academic Press, San Diego, CA: 241-263. DOI: 10.1016/B978-012472670-

3/50018-5

Furquim SAC, Graham RC, Barbiero L, Queiroz Neto JP, Vidal-Torrado P. 2010. Soil

mineral genesis and distribution in a saline lake landscape of the Pantanal

Wetland, Brazil. Geoderma 154: 518-528. DOI: 10.1016/j.geoderma.2009.03.014

Galbraith RF, Roberts RG, Laslett GM, Yoshida H, Olley JM. 1999. Optical dating of

single and multiple grains of quartz from Jinmium rock shelter, northern

Australia: Part I, experimental design and statistical models. Archaeometry 41:

339-364. DOI: 10.1111/j.1475-4754.1999.tb00987.x

Garcia MJ, De Oliveira PE, de Siqueira E, Fernandes RS. 2004. A Holocene vegetational

and climatic record from the Atlantic rainforest belt of coastal State of São Paulo,

SE Brazil. Review of Palaeobotany and Palynology 131: 181-199. DOI:

10.1016/j.revpalbo.2004.03.007

Garreaud RD, Vuille M, Compagnucci R, Marengo J. 2009. Present-day South American

climate. Palaeogeography, Palaeoclimatology, Palaeoecology 281: 180-195. DOI:

10.1016/j.palaeo.2007.10.032

Geyh MA, Schotterer U, Grosjean M. 1997. Temporal changes of the 14 C reservoir

effect in lakes. Radiocarbon 40: 921-931. DOI: 10.1017/S0033822200018890

Ghinassi M, D'oriano F, Benvenuti M, Fedi M, Awramik S. 2015. Lacustrine Facies in

Response To Millennial–Century-Scale Climate Changes (Lake Hayk, Northern

Ethiopia). Journal of Sedimentary Research 85: 381-398. DOI:

Page 145: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

133

10.2110/jsr.2015.28

Gierlowski-Kordesch E, Finkelstein DB, Holland JJT, Kallini KD. 2013. Carbonate lake

deposits associated with distal siliciclastic perennial-river systems. Journal of

Sedimentary Research 83: 1114-1129. DOI: 10.2110/jsr.2013.81

Gran G. 1952. Determination of the equivalence point in potentiometric titrations. Part II.

Analyst 77: 661-671. DOI: 10.1039/AN9527700661

Guérin G, Mercier N, Adamiec G. 2011. Dose-rate conversion factors: update. Ancient

TL 29: 5-8

Gumbricht T, Roman‐Cuesta RM, Verchot L, Herold M, Wittmann F, Householder E,

Herold N, Murdiyarso D. 2017. An expert system model for mapping tropical

wetlands and peatlands reveals South America as the largest contributor. Global

Change Biology: 23: 3581-3599. DOI: 10.1111/gcb.13689

Gustavson TC. 1991. Buried Vertisols in lacustrine facies of the Pliocene Fort Hancock

Formation, Hueco Bolson, West Texas and Chihuahua, Mexico. Geological

Society of America Bulletin 103: 448-460. DOI: 10.1130/0016-

7606(1991)103<0448:BVILFO>2.3.CO;2

Guy HP. 1969. Laboratory theory and methods for sediment analysis: Techniques of

Water-Resources Investigrations of the U.S. Geological Survey. 58.

Hahn A, Kliem P, Oehlerich M, Ohlendorf C, Zolitschka B. 2014. Elemental composition

of the Laguna Potrok Aike sediment sequence reveals paleoclimatic changes over

the past 51 ka in southern Patagonia, Argentina. Journal of Paleolimnology 52:

349-366. DOI: 10.1007/s10933-014-9798-y

Hamilton SK. 1999. Potential effects of a major navigation project (Paraguay-Parana

Hidrovia) on inundation in the Pantanal floodplains. Regulated Rivers: Research

& Management 15: 289-299. DOI: 10.1002/(SICI)1099-

1646(199907/08)15:4<289::AID-RRR520>3.0.CO;2-I

Hamilton SK, Lewis Jr WM. 1990. Basin morphology in relation to chemical and

ecological characteristics of lakes on the Orinoco River floodplain. Venezuela.

Arch. Hydrobt'ol 119: 393-425

Hansen MC, Loveland TR. 2012. A review of large area monitoring of land cover change

using Landsat data. Remote Sensing of Environment 122: 66-74. DOI:

Page 146: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

134

10.1016/j.rse.2011.08.024

Harris MB, Tomas W, Mourao G, Silva CJ, Guimaraes E, Sonoda F, Fachim E. 2005.

Safeguarding the Pantanal wetlands: threats and conservation initiatives.

Conservation Biology 19: 714-720. DOI: 10.1111/j.1523-1739.2005.00708.x

Hartley AJ, Weissmann GS, Bhattacharayya P, Nichols GJ, Scuderi LA, Davidson SK,

Leleu S, Chakraborty T, Ghosh P, Mather AE. 2013. Soil development on modern

distributive fluvial systems: preliminary observations with implications for

interpretation of paleosols in the rock record. New Frontiers in Paleopedology and

Terrestrial Paleoclimatology: SEPM, Special Publication 104: 149-158

Hartley AJ, Weissmann GS, Nichols GJ, Warwick GL. 2010. Large distributive fluvial

systems: characteristics, distribution, and controls on development. Journal of

Sedimentary Research 80: 167-183. DOI: 10.2110/sepmsp.104.10

Harvey AM, Wigand PE, Wells SG. 1999. Response of alluvial fan systems to the late

Pleistocene to Holocene climatic transition: contrasts between the margins of

pluvial Lakes Lahontan and Mojave, Nevada and California, USA. CATENA 36:

255-281. DOI: 10.1016/S0341-8162(99)00049-1

Hatushika RS, Silva CG, Mello CL. 2007. Sismoestratigrafia de alta resolução no lago

Juparanã, Linhares (ES - Brasil) como base para estudos sobre a sedimentação e

tectônica quartenária. Revista Brasileira de Geofísica 25: 433-442. DOI:

10.1590/S0102-261X2007000400007

Hercman H, Gąsiorowski M, Pawlak J. 2014. Testing the MOD-AGE chronologies of

lake sediment sequences dated by the 210Pb method. Quaternary Geochronology

22: 155-162. DOI: 10.1016/j.quageo.2014.01.001

Hillyer R, Valencia BG, Bush MB, Silman MR, Steinitz-Kannan M. 2009. A 24,700-yr

paleolimnological history from the Peruvian Andes. Quaternary Research 71: 71-

82. DOI: 10.1016/j.yqres.2008.06.006

Horton B, DeCelles P. 2001. Modern and ancient fluvial megafans in the foreland basin

system of the central Andes, southern Bolivia: Implications for drainage network

evolution in fold‐thrust belts. Basin Research 13: 43-63. DOI: 10.1046/j.1365-

2117.2001.00137.x

Horton BK, DeCelles PG. 1997. The modern foreland basin system adjacent to the

Page 147: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

135

Central Andes. Geology 25: 895-898. DOI: 10.1130/0091-

7613(1997)025<0895:TMFBSA>2.3

Huntley SL, Wenning RJ, Su SH, Bonnevie NL, Paustenbach DJ. 1995. Geochronology

and sedimentology of the lower Passaic River, New Jersey. Estuaries and Coasts

18: 351-361. DOI: 10.2307/1352317

Ioris AAR, Irigaray CT, Girard P. 2014. Institutional responses to climate change:

opportunities and barriers for adaptation in the Pantanal and the Upper Paraguay

River Basin. Climatic change 127: 139-151. DOI: 10.1007/s10584-014-1134-z

Ivory SJ, Blome MW, King JW, McGlue MM, Cole JE, Cohen AS. 2016. Environmental

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

million years. Proceedings of the National Academy of Sciences 113: 11895-

11900. DOI: 10.1073/pnas.1611028113

Jacob J, Disnar J-R, Boussafir M, Sifeddine A, Turcq B, Spadano Albuquerque AL.

2004. Major environmental changes recorded by lacustrine sedimentary organic

matter since the last glacial maximum near the equator (Lagoa do Caçó, NE

Brazil). Palaeogeography, Palaeoclimatology, Palaeoecology 205: 183-197. DOI:

10.1016/j.palaeo.2003.12.005

Johnson RC. 2003. Northwest to southeast cross section of Cretaceous and Lower

Tertiary Rocks across the eastern part of the Uinta Basin, Utah. Petroleum

Systems and Geologic Assessment of Oil and Gas in the Uinta-Piceance Province,

Utah and Colorado. US Geological Survey Digital Data Series DDS–69–B: 1-9

Johnson TC, McCave I. 2008. Transport mechanism and paleoclimatic significance of

terrigenous silt deposited in varved sediments of an African rift lake. Limnology

and Oceanography 53: 1622. DOI: 10.4319/lo.2008.53.4.1622

Johnston AE, Poulton PR, Coleman K. 2009. Soil organic matter: its importance in

sustainable agriculture and carbon dioxide fluxes. Advances in agronomy 101: 1-

57. DOI: 10.1016/S0065-2113(08)00801-8

Junk WJ, Cunha CN, Wantzen KM, Petermann P, Strüssmann C, Marques MI, Adis J.

2006. Biodiversity and its conservation in the Pantanal of Mato Grosso, Brazil.

Aquatic Sciences-Research Across Boundaries 68: 278-309. DOI:

10.1007/s00027-006-0851-4

Page 148: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

136

Kanner LC, Burns SJ, Cheng H, Edwards RL, Vuille M. 2013. High-resolution variability

of the South American summer monsoon over the last seven millennia: insights

from a speleothem record from the central Peruvian Andes. Quaternary Science

Reviews 75: 1-10. DOI: 10.1016/j.quascirev.2013.05.008

Karp T, Scholz CA, McGlue MM. 2012. Structure and stratigraphy of the Lake Albert

Rift, East Africa: Observations from seismic reflection and gravity data. DOI:

10.1306/13291394M952903

Kelts K. 1988. Environments of deposition of lacustrine petroleum source rocks: an

introduction. Geological Society, London, Special Publications 40: 3-26. DOI:

10.1023/A:1023270324800

Kigoshi T, Kumon F, Hayashi R, Kuriyama M, Yamada K, Takemura K. 2014. Climate

changes for the past 52 ka clarified by total organic carbon concentrations and

pollen composition in Lake Biwa, Japan. Quaternary International 333: 2-12.

DOI: 10.1016/j.quaint.2014.04.028

Kodama Y. 1992. Large-scale common features of subtropical precipitation zones (the

Baiu frontal zone, the SPCZ, and the SACZ) Part I: Characteristics of subtropical

frontal zones. Journal of the Meteorological Society of Japan. Ser. II 70: 813-836.

DOI: 10.2151/jmsj1965.70.4_813

Konert M, Vandenberghe J. 1997. Comparison of laser grain size analysis with pipette

and sieve analysis: a solution for the underestimation of the clay fraction.

Sedimentology 44: 523-535. DOI: 10.1046/j.1365-3091.1997.d01-38.x

Konijnendijk T, Weber S, Tuenter E, Weele Mv. 2011. Methane variations on orbital

timescales: a transient modeling experiment. Climate of the Past 7: 635-648. DOI:

10.5194/cp-7-635-2011

Kuerten S, Parolin M, Assine ML, McGlue MM. 2012. Sponge spicules indicate

Holocene environmental changes on the Nabileque River floodplain, southern

Pantanal, Brazil. Journal of Paleolimnology 49: 171-183. DOI: 10.1007/s10933-

012-9652-z

Lacerda Filho JV, Abreu Filho W, Valente CR, Oliveira CC, Albuquerque MC. 2004.

Geologia e recursos minerais do estado de Mato Grosso.

Larsen CP, MacDonald GM. 1993. Lake morphometry, sediment mixing and the

Page 149: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

137

selection of sites for fine resolution palaeoecological studies. Quaternary Science

Reviews 12: 781-792. DOI: 10.1016/0277-3791(93)90017-G

Latrubesse EM. 2015. Large rivers, megafans and other Quaternary avulsive fluvial

systems: A potential “who's who” in the geological record. Earth-Science

Reviews 146: 1-30. DOI: 10.1016/j.earscirev.2015.03.004

Latrubesse EM, Stevaux JC, Cremon EH, May J-H, Tatumi SH, Hurtado MA, Bezada M,

Argollo JB. 2012. Late Quaternary megafans, fans and fluvio-aeolian interactions

in the Bolivian Chaco, Tropical South America. Palaeogeography,

Palaeoclimatology, Palaeoecology 356: 75-88. DOI:

10.1016/j.palaeo.2012.04.003

LECO. 2008. SC-144DR Sulfur/Carbon Determinator Specification Sheet. St. Joseph,

Michigan; 2.

LECO. 2014. Com-Cat Accelerator Safety Data Sheet. 6.

LECO. 2015. Synthetic Carbon Safety Data Sheet. 6.

Leier AL, DeCelles PG, Pelletier JD. 2005. Mountains, monsoons, and megafans.

Geology 33: 289-292. DOI: 10.1130/G21228.1

Lewis Jr WM. 1983. A revised classification of lakes based on mixing. Canadian Journal

of Fisheries and Aquatic Sciences 40: 1779-1787. DOI: 10.1139/f83-207

Lintern A, Leahy PJ, Zawadzki A, Gadd P, Heijnis H, Jacobsen G, Connor S, Deletic A,

McCarthy DT. 2016. Sediment cores as archives of historical changes in

floodplain lake hydrology. Science of the Total Environment 544: 1008-1019.

DOI: 10.1016/j.scitotenv.2015.11.153

Liu X, Battisti DS. 2015. The influence of orbital forcing of tropical insolation on the

climate and isotopic composition of precipitation in South America. Journal of

Climate 28: 4841-4862. DOI: 10.1175/JCLI-D-14-00639.1

Liu X, Colman SM, Brown ET, Minor EC, Li H. 2013. Estimation of carbonate, total

organic carbon, and biogenic silica content by FTIR and XRF techniques in

lacustrine sediments. Journal of Paleolimnology 50: 387-398. DOI:

10.1007/s10933-013-9733-7

Livingston HD, Bowen VT. 1979. Pu and137Cs in coastal sediments. Earth and Planetary

Science Letters 43: 29-45. DOI: 10.1016/0012-821X(79)90153-5

Page 150: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

138

Lo EL, Bentley SJ, Sr., Xu K. 2014. Experimental study of cohesive sediment

consolidation and resuspension identifies approaches for coastal restoration: Lake

Lery, Louisiana. Geo-Marine Letters: 1-11. DOI: 10.1007/s00367-014-0381-3

Lo EL, Silva A, Bergier I, McGlue MM, Silva BLP, Silva APS, Pereira LE, Macedo HA,

Assine ML, Silva ERS. 2017. Spatiotemporal evolution of the margins of Lake

Uberaba, Pantanal floodplain (Brazil). Geografia 42.

López Laseras P, Navarro E, Marce Romero R, Ordóñez Salinas J, Caputo Galarce L,

Armengol J. 2006. Elemental ratios in sediments as indicators of ecological

processes in Spanish reservoirs. Limnetica 25: 499-512

Loverde-Oliveira SM, Huszar VLM, Mazzeo N, Scheffer M. 2009. Hydrology-Driven

Regime Shifts in a Shallow Tropical Lake. Ecosystems 12: 807-819. DOI:

10.1007/s10021-009-9258-0

Löwemark L, Chen H-F, Yang T-N, Kylander M, Yu E-F, Hsu Y-W, Lee T-Q, Song S-R,

Jarvis S. 2011. Normalizing XRF-scanner data: a cautionary note on the

interpretation of high-resolution records from organic-rich lakes. Journal of Asian

Earth Sciences 40: 1250-1256. DOI: 10.1016/j.jseaes.2010.06.002

Macedo HA, Stevaux JC, Assine ML, Silva A, Pupim FN, Merino ER. 2017. Calculating

bedload transport in rivers: concepts, calculus routine and application. Revista

Brasileira de Geomorfologia.

Magee JW, Bowler JM, Miller GH, Williams D. 1995. Stratigraphy, sedimentology,

chronology and palaeohydrology of Quaternary lacustrine deposits at Madigan

Gulf, Lake Eyre, South Australia. Palaeogeography, Palaeoclimatology,

Palaeoecology 113: 3-42. DOI: 10.1016/0031-0182(95)00060-Y

Marani L, Alvalá PC. 2007. Methane emissions from lakes and floodplains in Pantanal,

Brazil. Atmospheric Environment 41: 1627-1633. DOI:

10.1016/j.atmosenv.2006.10.046

McGlue MM, Guerreiro RL, Bergier I, Silva A, Pupim FN, Oberc V, Assine ML. 2017.

Holocene stratigraphic evolution of saline lakes in Nhecolândia, southern

Pantanal wetlands (Brazil). Quaternary Research: 1-19. DOI:

10.1017/qua.2017.57

McGlue MM, Scholz CA, Karp T, Ongodia B, Lezzar KE. 2006. Facies architecture of

Page 151: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

139

flexural margin lowstand delta deposits in Lake Edward, East African rift:

constraints from seismic reflection imaging. Journal of Sedimentary Research 76:

942-958. DOI: 10.2110/jsr.2006.068

McGlue MM, Silva A, Assine ML, Stevaux JC, do Nascimento Pupim F. 2015.

Paleolimnology in the pantanal: using lake sediments to track quaternary

environmental change in the world’s largest tropical wetland. In Dynamics of the

Pantanal Wetland in South America. Springer; 51-81. DOI:

10.1007/698_2015_350

McGlue MM, Silva A, Corradini FA, Zani H, Trees MA, Ellis GS, Parolin M,

Swarzenski PW, Cohen AS, Assine ML. 2011. Limnogeology in Brazil’s

“forgotten wilderness”: a synthesis from the large floodplain lakes of the

Pantanal. Journal of Paleolimnology 46: 273. DOI: 10.1007/s10933-011-9538-5

McGlue MM, Silva A, Zani H, Corradini FA, Parolin M, Abel EJ, Cohen AS, Assine

ML, Ellis GS, Trees MA. 2012. Lacustrine records of Holocene flood pulse

dynamics in the Upper Paraguay River watershed (Pantanal wetlands, Brazil).

Quaternary Research 78: 285-294. DOI: 10.1016/j.yqres.2012.05.015

McGlue MM, Smith PH, Zani H, Silva A, Carrapa B, Cohen AS, Pepper MB. 2016. An

Integrated Sedimentary Systems Analysis of the Río Bermejo (Argentina):

Megafan Character in the Overfilled Southern Chaco Foreland Basin. Journal of

Sedimentary Research 86: 1359-1377. DOI: 10.2110/jsr.2016.82

Metcalfe SE, Whitney BS, Fitzpatrick KA, Mayle FE, Loader NJ, Street‐Perrott FA,

Mann DG. 2014. Hydrology and climatology at Laguna La Gaiba, lowland

Bolivia: complex responses to climatic forcings over the last 25 000 years.

Journal of Quaternary Science 29: 289-300. DOI: 10.1002/jqs.2702

Middelboe AL, Markager S. 1997. Depth limits and minimum light requirements of

freshwater macrophytes. Freshwater biology 37: 553-568. DOI: 10.1046/j.1365-

2427.1997.00183.x

Milly PC, Dunne KA, Vecchia AV. 2005. Global pattern of trends in streamflow and

water availability in a changing climate. Nature 438: 347. DOI:

10.1038/nature04312

Min J-E, Ryu J-H, Lee S, Son S. 2012. Monitoring of suspended sediment variation using

Page 152: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

140

Landsat and MODIS in the Saemangeum coastal area of Korea. Marine Pollution

Bulletin 64: 382-390. DOI: 10.1016/j.marpolbul.2011.10.025

Mitsch WJ, Nahlik A, Wolski P, Bernal B, Zhang L, Ramberg L. 2010. Tropical

wetlands: seasonal hydrologic pulsing, carbon sequestration, and methane

emissions. Wetlands Ecology and Management 18: 573-586. DOI:

10.1007/s11273-009-9164-4

Mortlock RA, Froelich PN. 1989. A simple method for the rapid determination of

biogenic opal in pelagic marine sediments. Deep Sea Research Part A.

Oceanographic Research Papers 36: 1415-1426. DOI: 10.1016/0198-

0149(89)90092-7

Mudie PJ, Rochon A, Prins MA, Soenarjo D, Troelstra SR, Levac E, Scott DB, Roncaglia

L, Kuijpers A. 2006. Late Pleistocene-Holocene marine geology of Nares Strait

region: Palaeoceanography from foraminifera and dinoflagellate cysts,

sedimentology and stable isotopes. Polarforschung 74: 169-183. DOI:

10013/epic.29931

Mueller ER, Grams PE, Schmidt JC, Hazel JE, Alexander JS, Kaplinski M. 2014. The

influence of controlled floods on fine sediment storage in debris fan-affected

canyons of the Colorado River basin. Geomorphology 226: 65-75. DOI:

10.1016/j.geomorph.2014.07.029

Murphy JT, Lowenstein TK, Pietras JT. 2014. Preservation of primary lake signatures in

alkaline earth carbonates of the Eocene Green River Wilkins Peak-Laney Member

transition zone. Sedimentary Geology 314: 75-91. DOI:

10.1016/j.sedgeo.2014.09.005

Murray AS, Wintle AG. 2000. Luminescence dating of quartz using an improved single-

aliquot regenerative-dose protocol. Radiation measurements 32: 57-73. DOI:

10.1016/S1350-4487(99)00253-X

Murray AS, Wintle AG. 2003. The single aliquot regenerative dose protocol: potential for

improvements in reliability. Radiation measurements 37: 377-381. DOI:

10.1016/S1350-4487(03)00053-2

Muttashar W, Al Mosawi W, Al-Aesawi Q, Abas N. 2012. Detection of subsurface layers

by Sub Bottom Profiling (SBP) of cross section of Shatt Al-Arab River at Al-

Page 153: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

141

Rebat branch, Basrah, southern Iraq. Mesopotamian Journal of Marine Sciences

27: 49-58

Newcombe CP, MacDonald DD. 1991. Effects of suspended sediments on aquatic

ecosystems. North American journal of fisheries management 11: 72-82. DOI:

10.1577/1548-8675(1991)011<0072:EOSSOA>2.3.CO;2

Nichols GJ, Fisher JA. 2007. Processes, facies and architecture of fluvial distributary

system deposits. Sedimentary Geology 195: 75-90. DOI:

10.1016/j.sedgeo.2006.07.004

Noller JS. 2000. Lead‐210 Geochronology. Quaternary geochronology: Methods and

applications: 115-120. DOI: 10.1029/RF004p0115

Novello VF, Cruz FW, Karmann I, Burns SJ, Stríkis NM, Vuille M, Cheng H, Lawrence

Edwards R, Santos RV, Frigo E, Barreto EAS. 2012. Multidecadal climate

variability in Brazil's Nordeste during the last 3000 years based on speleothem

isotope records. Geophysical Research Letters 39: 1-6. DOI:

10.1029/2012GL053936

Novello VF, Cruz FW, Vuille M, Stríkis NM, Edwards RL, Cheng H, Emerick S, De

Paula MS, Li X, Barreto EdS. 2017. A high-resolution history of the South

American Monsoon from Last Glacial Maximum to the Holocene. Scientific

Reports 7: 44267. DOI: 10.1038/srep44267

Novello VF, Vuille M, Cruz FW, Stríkis NM, de Paula MS, Edwards RL, Cheng H,

Karmann I, Jaqueto PF, Trindade RIF, Hartmann GA, Moquet JS. 2016.

Centennial-scale solar forcing of the South American Monsoon System recorded

in stalagmites. Scientific Reports 6: 24762. DOI: 10.1038/srep24762

Oldfield F, Appleby P. 1984. Empirical testing of 210 Pb-dating models for lake

sediments. In Lake sediments and environmental history.

Olsen J, Ascough P, Lougheed BC, Rasmussen P. 2017. Radiocarbon dating in estuarine

environments. In Applications of paleoenvironmental techniques in estuarine

studies. Springer; 141-170. DOI: 10.1007/978-94-024-0990-1_7

Opluštil S, Martínek K, Tasáryová Z. 2005. Facies and architectural analysis of fluvial

deposits of the Nýřany Member and the Týnec Formation (Westphalian D–

Barruelian) in the Kladno-Rakovník and Pilsen basins. Bulletin of Geosciences

Page 154: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

142

80: 45-66

Osaki M, Hirose K, Segah H, Helmy F. 2016. Tropical Peat and Peatland Definition in

Indonesia. In Tropical Peatland Ecosystems. Springer; 137-147. DOI:

10.1007/978-4-431-55681-7_9

Oviatt CG, Thompson RS, Kaufman DS, Bright J, Forester RM. 1999. Reinterpretation of

the Burmester core, Bonneville basin, Utah. Quaternary Research 52: 180-184.

DOI: 10.1006/qres.1999.2058

Peng Y, Xiao J, Nakamura T, Liu B, Inouchi Y. 2005. Holocene East Asian monsoonal

precipitation pattern revealed by grain-size distribution of core sediments of

Daihai Lake in Inner Mongolia of north-central China. Earth and Planetary

Science Letters 233: 467-479. DOI: 10.1016/j.epsl.2005.02.022

Pietras JT, Carroll AR, Rhodes MK. 2003. Lake basin response to tectonic drainage

diversion: Eocene Green River Formation, Wyoming. Journal of Paleolimnology

30: 115-125. DOI: 10.1023/A:1025518015341

Polissar PJ, Abbott MB, Wolfe AP, Vuille M, Bezada M. 2013. Synchronous

interhemispheric Holocene climate trends in the tropical Andes. Proceedings of

the National Academy of Sciences 110: 14551-14556. DOI:

10.1073/pnas.1219681110

Pott A, Silva JSV. 2015. Terrestrial and Aquatic Vegetation Diversity of the Pantanal

Wetland. In Dynamics of the Pantanal Wetland in South America, Bergier I,

Assine LM (eds). Springer International Publishing: Cham; 111-131. DOI:

10.1007/698_2015_352

Prescott JR, Hutton JT. 1994. Cosmic ray contributions to dose rates for luminescence

and ESR dating: large depths and long-term time variations. Radiation

measurements 23: 497-500. DOI: 10.1016/1350-4487(94)90086-8

Pupim FN, Assine ML, Sawakuchi AO. 2017. Late Quaternary Cuiabá megafan,

Brazilian Pantanal: Channel patterns and paleoenvironmental changes. Quaternary

International. DOI: 10.1016/j.quaint.2017.01.013

Pupim FN, Bierman PR, Assine ML, Rood DH, Silva A, Merino ER. 2015. Erosion rates

and landscape evolution of the lowlands of the Upper Paraguay River basin

(Brazil) from cosmogenic 10 Be. Geomorphology 234: 151-160. DOI:

Page 155: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

143

10.1016/j.geomorph.2015.01.016

Quigley HB, Crawshaw PG. 1992. A conservation plan for the jaguar Panthera onca in

the Pantanal region of Brazil. Biological Conservation 61: 149-157. DOI:

10.1016/0006-3207(92)91111-5

Resende EK. 2000. Trophic structure of fish assemblages in the Lower Miranda River,

Pantanal, Mato Grosso do Sul State, Brazil. Revista Brasileira de Biologia 60:

389-403. DOI: 10.1590/S0034-71082000000300004

Ribes C, Kergaravat C, Bonnel C, Crumeyrolle P, Callot JP, Poisson A, Temiz H,

Ringenbach JC. 2015. Fluvial sedimentation in a salt‐controlled mini‐basin:

stratal patterns and facies assemblages, Sivas Basin, Turkey. Sedimentology 62:

1513-1545. DOI: 10.1111/sed.12195

Riggs AC. 1984. Major carbon-14 deficiency in modern snail shells from southern

Nevada springs. Science 224: 58-61. DOI: 10.1126/science.224.4644.58

Ring EJ. 1993. The preparation and certification of fourteen South African silicate rocks

for use as reference materials. Geostandards Newsletter 17: 137-158. DOI:

10.1111/j.1751-908X.1993.tb00130.x

Robbins JA. 1978. Geochemical and geophysical applications of radioactive lead. The

biogeochemistry of lead in the environment 1: 285-337

Rodbell DT, Seltzer GO, Anderson DM, Abbott MB, Enfield DB, Newman JH. 1999. An

~15,000-year record of El Niño-driven alluviation in southwestern Ecuador.

Science 283: 516-520. DOI: 10.1126/science.283.5401.516

Rodríguez JM, Johns TC, Thorpe RB, Wiltshire A. 2011. Using moisture conservation to

evaluate oceanic surface freshwater fluxes in climate models. Climate Dynamics

37: 205-219. DOI: 10.1007/s00382-010-0899-7

Rounds, SA. 2006. Alkalinity and acid neutralizing capacity (ver. 3.0): U.S. Geological

Survey Techniques of Water-Resources Investigations, book 9, chap. A6., sec.

6.6.7., July 2006, accessed July 24, 2017, from

http://pubs.water.usgs.gov/twri9A6/

Rowe H, Hughes N, Robinson K. 2012. The quantification and application of handheld

energy-dispersive x-ray fluorescence (ED-XRF) in mudrock chemostratigraphy

and geochemistry. Chemical Geology 324: 122-131. DOI:

Page 156: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

144

10.1016/j.chemgeo.2011.12.023

Rowland JC, Dietrich WE, Day G, Parker G. 2009. Formation and maintenance of single-

thread tie channels entering floodplain lakes: Observations from three diverse

river systems. Journal of Geophysical Research: Earth Surface 114: n/a-n/a. DOI:

10.1029/2008JF001073

Russell JM, Johnson TC, Kelts KR, Lærdal T, Talbot MR. 2003. An 11 000-year

lithostratigraphic and paleohydrologic record from equatorial Africa: Lake

Edward, Uganda–Congo. Palaeogeography, Palaeoclimatology, Palaeoecology

193: 25-49. DOI: 10.1016/S0031-0182(02)00709-5

Sáez A, Anadón P, Herrero MJ, Moscariello A. 2007. Variable style of transition between

Palaeogene fluvial fan and lacustrine systems, southern Pyrenean foreland, NE

Spain. Sedimentology 54: 367-390. DOI: 10.1111/j.1365-3091.2006.00840.x

Sanchez-Cabeza J, Ruiz-Fernández A. 2012. 210 Pb sediment radiochronology: an

integrated formulation and classification of dating models. Geochimica et

Cosmochimica Acta 82: 183-200. DOI: 10.1016/j.gca.2010.12.024

Santos GM, Linares R, Lisi CS, Tomazello Filho M. 2015. Annual growth rings in a

sample of Paraná pine (Araucaria angustifolia): Toward improving the 14 C

calibration curve for the Southern Hemisphere. Quaternary Geochronology 25:

96-103. DOI: 10.1016/j.quageo.2014.10.004

Sawakuchi AO, Mendes VR, Pupim FN, Mineli TD, Ribeiro LMAL, Zular A, Guedes

CCF, Giannini PCF, Nogueira L, Sallun Filho W, Assine ML. 2016. Optically

stimulated luminescence and isothermal thermoluminescence dating of high

sensitivity and well bleached quartz from Brazilian sediments: from Late

Holocene to beyond the Quaternary? Brazilian Journal of Geology 46: 209-226.

DOI: 10.1590/2317-488920160030295

Schillereff DN, Chiverrell RC, Macdonald N, Hooke JM. 2014. Flood stratigraphies in

lake sediments: A review. Earth-Science Reviews 135: 17-37. DOI:

10.1016/j.earscirev.2014.03.011

Schnurrenberger D, Russell J, Kelts K. 2003. Classification of lacustrine sediments based

on sedimentary components. Journal of Paleolimnology 29: 141-154. DOI:

10.1023/A:1023270324800

Page 157: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

145

Seager R, Naik N, Vogel L. 2012. Does global warming cause intensified interannual

hydroclimate variability? Journal of Climate 25: 3355-3372. DOI: 10.1175/JCLI-

D-11-00363.1

Seidl AF, Moraes AS. 2000. Global valuation of ecosystem services: application to the

Pantanal da Nhecolandia, Brazil. Ecological economics 33: 1-6. DOI:

10.1016/S0921-8009(99)00146-9

Shanley KW, McCabe PJ. 1994. Perspectives on the sequence stratigraphy of continental

strata. AAPG bulletin 78: 544-568. DOI: 10.1306/BDFF9258-1718-11D7-

8645000102C1865D

Shen J, Yuan H, Liu E, Wang J, Wang Y. 2011. Spatial distribution and stratigraphic

characteristics of surface sediments in Taihu Lake, China. Chinese Science

Bulletin 56: 179-187. DOI: 10.1007/s11434-010-4214-0

Shindell DT, Walter BP, Faluvegi G. 2004. Impacts of climate change on methane

emissions from wetlands. Geophysical Research Letters 31: L21202. DOI:

10.1029/2004GL021009

Shynu R, Rao VP, Sarma V, Kessarkar PM, ManiMurali R. 2015. Sources and fate of

organic matter in suspended and bottom sediments of the Mandovi and Zuari

estuaries, western India.

Silva CJ, Girard P. 2004. New challenges in the management of the Brazilian Pantanal

and catchment area. Wetlands Ecology and Management 12: 553-561

Silva FC. 2009. Manual de análises químicas de solos, plantas e fertilizantes. Embrapa

Informática Tecnológica: Brasília.

Sjögersten S, Black CR, Evers S, Hoyos‐Santillan J, Wright EL, Turner BL. 2014.

Tropical wetlands: A missing link in the global carbon cycle? Global

Biogeochemical Cycles 28: 1371-1386. DOI: 10.1002/2014GB004844

Souza RCS, Vianna EF, Pellegrin LA, Salis SM, Costa M, Bergier I. 2011. Localização

de áreas permanentes de vegetação aquática na planície de inundação do Rio

Paraguai e adjacências. In Anais XV Simpósio Brasileiro de Sensoriamento

Remoto—SBSR, Curitiba, PR, Brazil.

Stevaux JC. 2000. Climatic events during the late Pleistocene and Holocene in the upper

Parana River: Correlation with NE Argentina and South-Central Brazil.

Page 158: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

146

Quaternary International 72: 73-85. DOI: 10.1016/S1040-6182(00)00023-9

Stuiver M, Polach HA. 1977. Discussion reporting of 14 C data. Radiocarbon 19: 355-

363. DOI: 10.1017/S0033822200003672

Suárez YR, Petrere Jr M, Catella AC. 2004. Factors regulating diversity and abundance

of fish communities in Pantanal lagoons, Brazil. Fisheries Management and

Ecology 11: 45-50. DOI: 10.1111/j.1365-2400.2004.00347.x

Sylvestre F. 2002. A high-resolution diatom reconstruction between 21,000 and 17, 4000

14C yr BP from the southern Bolivian Altiplano (18–23° S). Journal of

Paleolimnology 27: 45-57. DOI: 10.1023/A:1013542907394

Syvitski, JPM, Cohen, S, Kettner, AJ, Brakenridge, GR. 2014. How important and

different are tropical rivers?—An overview. Geomorphology 227: 5-17. DOI:

10.1016/j.geomorph.2014.02.029

Thermo Scientific. 2012. Dionex ICS-1100 Ion Chromatography System Operator's

Manual. Thermo Fisher Scientific Inc.; 148

Tricart J. 1982. La eco-geografía y la ordenación del medio natural. Barcelona, ES: Edit.

Anagrama

Tye RS, Coleman JM. 1989. Depositional processes and stratigraphy of fluvially

dominated lacustrine deltas: Mississippi delta plain. Journal of Sedimentary

Research 59: 973-996. DOI: 10.1306/212F90CA-2B24-11D7-

8648000102C1865D

UIC. 2017. Carbon Dioxide Coulometer. Technical description of how the Model

CM5014 CO2 coulometer functions.

Ussami N, Shiraiwa S, Dominguez JML. 1999. Basement reactivation in a sub‐Andean

foreland flexural bulge: The Pantanal wetland, SW Brazil. Tectonics 18: 25-39.

DOI: 10.1029/1998TC900004

Van Wagoner JC. 1995. Sequence stratigraphy and marine to nonmarine facies

architecture of foreland basin strata, Book Cliffs, Utah, USA.

Vega LF, Cunha CN, Rothaupt KO, Moreira MZ, Wantzen KM. 2014. Does Flood

Pulsing Act as a Switch to Store or Release Sediment-Bound Carbon in Seasonal

Floodplain Lakes? Case Study from the Colombian Orinoco-Llanos and the

Brazilian Pantanal. Wetlands 34: 177-187. DOI: 10.1007/s13157-013-0495-9

Page 159: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

147

Viana JCC, Sifeddine A, Turcq B, Albuquerque ALS, Moreira LS, Gomes DF, Cordeiro

RC. 2014. A late Holocene paleoclimate reconstruction from Boqueirao Lake

sediments, northeastern Brazil. Palaeogeography Palaeoclimatology

Palaeoecology 415: 117-126. DOI: 10.1016/j.palaeo.2014.07.010

Vogel JS, Southon JR, Nelson DE. 1987. Catalyst and binder effects in the use of

filamentous graphite for AMS. Nuclear Instruments and Methods in Physics

Research Section B: Beam Interactions with Materials and Atoms 29: 50-56. DOI:

10.1016/0168-583X(87)90202-3

Vriens B, Lenz M, Charlet L, Berg M, Winkel LHE. 2014. Natural wetland emissions of

methylated trace elements. Nature Communications 5. DOI:

10.1038/ncomms4035

Vuille M, Bradley RS, Werner M, Healy R, Keimig F. 2003. Modeling δ18O in

precipitation over the tropical Americas: 1. Interannual variability and climatic

controls. Journal of Geophysical Research: Atmospheres 108: n/a-n/a. DOI:

10.1029/2001JD002038

Vuille M, Burns SJ, Taylor BL, Cruz FW, Bird BW, Abbott MB, Kanner LC, Cheng H,

Novello VF. 2012. A review of the South American monsoon history as recorded

in stable isotopic proxies over the past two millennia. Clim. Past 8: 1309-1321.

DOI: 10.5194/cp-8-1309-2012

Wade L. 2014. Chasing South America's monsoon. Science 346: 1042-1043. DOI:

10.1126/science.346.6213.1042

Warren LV, Quaglio F, Simões MG, Freitas BT, Assine ML, Riccomini C. 2015.

Underneath the Pantanal Wetland: A Deep-Time History of Gondwana Assembly,

Climate Change, and the Dawn of Metazoan Life. In Dynamics of the Pantanal

Wetland in South America, Bergier I, Assine ML (eds). Springer International

Publishing: Cham; 1-21. DOI: 10.1007%2F698_2014_326

Weissmann G, Hartley A, Nichols G, Scuderi L, Olson M, Buehler H, Banteah R. 2010.

Fluvial form in modern continental sedimentary basins: Distributive fluvial

systems. Geology 38: 39-42. DOI: 10.1130/G30242.1

Weissmann G, Hartley A, Nichols G, Scuderi L, Olson M, Buehler H, Massengill L.

2011. Alluvial facies distributions in continental sedimentary basins—distributive

Page 160: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

148

fluvial systems. From River to Rock Record: The Preservation of Fluvial

Sediments and their Subsequent Interpretation: SEPM, Special Publication 97:

327-355. DOI: 10.2110/sepmsp.097.327

Weissmann G, Hartley A, Scuderi L, Nichols G, Owen A, Wright S, Felicia A, Holland

F, Anaya F. 2015. Fluvial geomorphic elements in modern sedimentary basins

and their potential preservation in the rock record: a review. Geomorphology 250:

187-219. DOI: 10.1016/j.geomorph.2015.09.005

Welch K, Lyons W, Graham E, Neumann K, Thomas J, Mikesell D. 1996. Determination

of major element chemistry in terrestrial waters from Antarctica by ion

chromatography. Journal of Chromatography A 739: 257-263. DOI:

10.1016/0021-9673(96)00044-1

Weninger B, Jöris O. 2008. A 14 C age calibration curve for the last 60 ka: the

Greenland-Hulu U/Th timescale and its impact on understanding the Middle to

Upper Paleolithic transition in Western Eurasia. Journal of Human Evolution 55:

772-781. DOI: 10.1016/j.jhevol.2008.08.017

Wentworth CK. 1922. A scale of grade and class terms for clastic sediments. The Journal

of Geology 30: 377-392. DOI: 10.1086/622910

Wetzel RG. 2001. Limnology: lake and river ecosystems. Academic Press: Cambridge,

MA, US.

Whitney BS, Mayle FE, Burn MJ, Guillén R, Chavez E, Pennington RT. 2014.

Sensitivity of Bolivian seasonally-dry tropical forest to precipitation and

temperature changes over glacial–interglacial timescales. Vegetation history and

archaeobotany 23: 1-14. DOI: 10.1007/s00334-013-0395-1

Whitney BS, Mayle FE, Punyasena SW, Fitzpatrick KA, Burn MJ, Guillen R, Chavez E,

Mann D, Pennington RT, Metcalfe SE. 2011. A 45 kyr palaeoclimate record from

the lowland interior of tropical South America. Palaeogeography

Palaeoclimatology Palaeoecology 307: 177-192. DOI:

10.1016/j.palaeo.2011.05.012

Wilkinson MJ. 2016. Megafans-Some New Perspectives from a Global Study. In 2017

Geological Society of America Annual Meeting: Denver, CO. DOI:

10.1130/abs/2016AM-287880

Page 161: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

149

Wolfe BB, Karst-Riddoch TL, Vardy SR, Falcone MD, Hall RI, Edwards TWD. 2005.

Impacts of climate and river flooding on the hydro-ecology of a floodplain basin,

Peace-Athabasca Delta, Canada since A.D. 1700. Quaternary Research 64: 147-

162. DOI: 10.1016/j.yqres.2005.05.001

Wright Jr H. 1967. A square-rod piston sampler for Lake Sediments: NOTES. Journal of

Sedimentary Research 37: 975-976. DOI: 10.1306/74D71807-2B21-11D7-

8648000102C1865D

Yeager KM, Brunner CA, Kulp MA, Fischer D, Feagin RA, Schindler KJ, Prouhet J,

Bera G. 2012. Significance of active growth faulting on marsh accretion processes

in the lower Pearl River, Louisiana. Geomorphology 153–154: 127-143. DOI:

10.1016/j.geomorph.2012.02.018

Yu S-Y, Shen J, Colman SM. 2007. Modeling the radiocarbon reservoir effect in

lacustrine systems. Radiocarbon 49: 1241-1254

Zhang J, Ma X, Qiang M, Huang X, Li S, Guo X, Henderson AC, Holmes JA, Chen F.

2016. Developing inorganic carbon-based radiocarbon chronologies for Holocene

lake sediments in arid NW China. Quaternary Science Reviews 144: 66-82. DOI:

10.1016/j.quascirev.2016.05.034

Zhu Q, Peng C, Ciais P, Jiang H, Liu J, Bousquet P, Li S, Chang J, Fang X, Zhou X.

2017. Inter‐annual Variation in Methane Emissions from Tropical Wetlands

Triggered by Repeated El Niño Southern Oscillation. Global Change Biology 0:

1-11. DOI: 10.1111/gcb.13726

Zhu Y, Wu F, He Z, Giesy JP, Feng W, Mu Y, Feng C, Zhao X, Liao H, Tang Z. 2015.

Influence of natural organic matter on the bioavailability and preservation of

organic phosphorus in lake sediments. Chemical Geology 397: 51-60. DOI:

10.1016/j.chemgeo.2015.01.006

Zocatelli R, Turcq B, Boussafir M, Cordeiro RC, Disnar J-R, Costa R, Sifeddine A,

Albuquerque ALS, Bernardes M, Jacob J. 2012. Late Holocene

paleoenvironmental changes in Northeast Brazil recorded by organic matter in

lacustrine sediments of Lake Boqueirão. Palaeogeography, Palaeoclimatology,

Palaeoecology 363: 127-134. DOI: 10.1016/j.palaeo.2012.08.021

Page 162: FLUVIAL-LACUSTRINE PROCESSES SHAPING THE LANDFORMS …

150

VITA

Edward Limin Lo

Education

B.S. Geology—Environmental concentration, 12/2013

Louisiana State University Agricultural and Mechanical College

Professional Positions

Graduate Teaching Assistant, Aug-Dec 2016

University of Kentucky

Graduate Research Assistant, Jan-May 2016

University of Kentucky

Fulbright Fellow, Brazil, 2014-15

Graduate Teaching Assistant, Aug-Dec 2014

University of Kentucky

Freshmen Geology Camp TA, Summer 2014

Louisiana State University Field Camp (Colorado Springs, CO)

Science Undergraduate Laboratory Intern, Spring 2014

Lawrence Berkeley National Laboratory

Professional Publications

Lo EL, Silva A, Bergier I, McGlue MM, Silva BLP, Silva APS, Pereira LE, Macedo HA, Assine

ML, Silva ERS. 2017. Spatiotemporal evolution of the margins of Lake Uberaba, Pantanal

floodplain (Brazil). Geografia, in press.

Lo EL, Bentley Sr SJ, Xu K. 2014. Experimental study of cohesive sediment consolidation and

resuspension identifies approaches for coastal restoration: Lake Lery, Louisiana. Geo-Marine

Letters 34: 1-11. DOI: 10.1007/s00367-014-0381-3

Macedo HA, Stevaux JC, Assine ML, Silva A, Pupim FN, Merino ER, Lo EL. 2017.

Calculating bedload transport in rivers: concepts, calculus routine and application.

Revista Brasileira de Geomorfologia 18: 813-824. DOI: 10.20502/rbg.v18i4.1227

Pereira LE, Lastoria G, Almeida BS, Haupenthal M, Júnior JM, Paranhos Filho AC. 2017.

Application of aerial photographs and orbital sensor to identify and delineate water bodies.

Boletim de Ciências Geodésicas 23: in press.

Silva ERS, Silva A, Silva, BLP, Pereira LE, Lo EL, Fonseca TPL, Oliveira MR. 2017. Analysis

of morphological and hydrological changes in the Correntes River. Geografia, in press.