Hydroclimatic study of Plio-Pleistocene aquatic sites in Meade County,

Kansas

A thesis submitted to the Kent State University in partial

fulfillment of the requirements for the

Degree of Master of Science

by

Marissa J. Tomin

August 2020

Thesis written by

Marissa J. Tomin

B.S., The University of Akron, 2017

M.S., Kent State University, 2020

Approved by

______, Advisor Alison J. Smith

______, Chair, Department of Geology Daniel K. Holm

______, Dr. Mandy Munro-Stasiuk, Interim Dean, College of Arts and Sciences

TABLE OF CONTENTS………………………………………………………………………... iii

LIST OF FIGURES……………………………………………………………………………… vi

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

ACKNOWLEDGEMENTS…………………………………………………………………… .viii

CHAPTERS

I. INTRODUCTION

1.1 Statement of Problem/Hypothesis to Test…………………………………………... 1

1.2 Pliocene-Pleistocene Paleoclimate and the Southern Great Plains of North America 4

1.3 Previous Work in Meade County, KS 10

II. METHODS

2.1 Fieldwork, Stratigraphy and Chronology…………………………………………….14

2.2 Sediment Processing Methods……………………………………………………….18

2.3 Vitris Benchtop Research Freeze Dryer Operation…………………………………..20

2.4 Sample Picking and Ostracode Identification………………………………………..20

2.5 Older Collections……………………………………………………………………..21

2.6 Species Identification……………………………………………………………….. .22

2.7 Statistical Analysis…………………………………………………………………...22

2.7.1 Diversity Analysis…………………………………………………23

iii

2.7.2 Classic Modern Analog Method…………………………………………... 24

2.7.3 Cluster Analysis……………………………………………………………25

2.7.4 NANODe and Neotoma databases…………………………………………26

III. RESULTS

3.1 Stratigraphy and Chronology………………………………………………………...28

3.2 Ostracode Species, stratigraphic location, and ecological significance……………...31

3.3 Constrained Cluster Analysis………………………………………………………...36

3.4 Unconstrained Cluster Analysis……………………………………………………...36

3.5 Distribution of ostracode species assemblages in solute space ………………...... 39

3.6 Shannon’s Diversity Index………………………………………………………….. 41

3.7 Modern analog estimates of precipitation with NANODe sites…………………….. 43

3.8 Summary of Results…………………………………………………………………. 45

DISSCUSSION………………………………………………………………………………….. 48

CONCLUSIONS………………………………………………………………………………... 55

REFERENCES………………………………………………………………………………….. 56

APPENDICES

A. Ostracode Counts from the Tomin 2019 Collection………………………………. 65

B. Ostracode Counts from the Gutentag and Benson (1962) Collection……………... 69

C. Ostracode Counts from the Hibbard (195-) Collection…………………………….77

iv

D. Shannon’s Diversity Index………………………………………………………… 82

E. Cluster Analysis…………………………………………………………………… 83

F. Modern Analogue…………………………………………………………………. 96

v

LIST OF FIGURES

Figure 1. Study region and hydroclimate-sensitive proxies of Ibarra et al. (2018)………………. 3

Figure 2. The LR04 from Lisiecki, & Raymo (2005)…………………………………………….. 7

Figure 3. The extent of glaciation in North America……………………………………………... 8

Figure 4. Stratigraphic column from Gutentag and Benson (1962)……………………………...12

Figure 5. Stratigraphic columns from Layzell et al. (2017)……………………………………... 13

Figure 6. Map showing sample location sites in Meade County, Kansas………………………..15

Figure 7. Stratigraphic column depicting 2019 sample site locations…………………………... 29

Figure 8. Photomicrographs of 8 common species in this study………………………………... 32

Figure 9. Constrained Cluster Analysis…………………………………………………………. 37

Figure 10. Unconstrained Cluster Analysis……………………………………………………... 38

Figure 11: Modern distribution in solute space of species assemblages………………………... 40

Figure 12: Percent Abundance of fossil ostracode species……………………………………… 47

Figure 13: Summary of Results…………………………………………………………………. 54

vi

LIST OF TABLES

Table 1: Sample names, locations, ages and formations in Meade County, KS…………………16

Table 2: Names, stages, labels, and ages of ostracode sample slides…………………………… 30

Table 3: Ostracode species and their ecological significance…………………………………… 34

Table 4: Ostracode species in stratigraphic order……………………………………………….. 35

Table 5. Shannon’s Diversity Index……………………………………………………………...42

Table 6: Modern analog of NANODe sites……………………………………………………... 44

vii

AKNOWLEDGEMENTS

Ever since I was little, I have been drawn to geology. Specifically, I have been drawn to shiny, smooth rocks that I would find in our driveway. As I grew up, I began to have an appreciation for the natural world around me, so when the time came to pick a major in undergrad, geology felt like the natural choice. As I completed my studies, I became fascinated with the world of paleontology, not only for its cool fossils, but also for how it could be used to help with modern-day environmental problems. This fascination followed me into graduate school at Kent State University, and I could not be more grateful for the opportunity be a part of this program. I would like to thank the Kent State Geology Department for giving me this opportunity, which has allowed me to become a better geologist and a better person.

To my parents, Michael, and Barbara, for their unending love and support that allowed me to be here today. No matter how difficult things became, you never let me give up on myself, and for that I will be forever grateful.

To my advisor, Dr. Alison Smith, who showed me the wonderful world of ostracodes.

Your patience, passion, and seemingly infinite knowledge of ostracodes not only inspired me but also gave me the confidence I needed to complete this thesis. I can truly never thank you enough!

To my “field advisor” Dr. Tony Layzell, who worked tirelessly to help me collect my field samples. I would never have found my field sites without you, and you made my trip to

Kansas a memory that I will never forget. To Catherine Opalka, who helped with fieldwork and was an absolute joy to work with. Thank you so much for coming to Meade County with me! I would also like to thank the Kansas Geological Survey for providing the field support needed to complete this thesis.

viii

I would also like to thank my committee members, Dr. Jefferson, and Dr. Hacker. Thank you so much for allowing me to participate in this program, and for all the help and advice you have given me.

Finally, I would like to thank my geology colleagues Jacob Bradley and Angela Lewis.

You guys have been such good friends throughout my time at Kent State, and I know that you will go on to achieve all that you set your mind to!

ix

1. Introduction 1.1 Statement of Problem/ Hypothesis to Test

One of the most pressing issues in science today is the cause and consequences of a changing global climate. Climate change itself is a hot-button topic, not just in the scientific sphere but in the political, social, and economic spheres as well. A changing climate can have long-term effects, such as rising sea level, the rapid decline of numerous plant and species, and negative effects on agriculture, human health and society (Change, et al., 2006; Karl et al., 2009). More immediate effects of climate change include the frequency and dispersion of infectious diseases and the increased intensity of hurricanes, tornadoes, droughts, and wildfires

(Karl et al., 2009; Kolivras & Comrie, 2004; Hurteau et al., 2014). One of the more serious effects of climate change that impacts society are changes in precipitation patterns. These precipitation changes, both locally and regionally, not only increase the likelihood of disaster events such as flash floods, but can also negatively impact transportation, water systems, ecology, and society (Karl et al., 2009). Decreases in precipitation can affect food security and drinking water systems for municipalities, causing water shortages in areas where freshwater is already in short supply (Mullin, 2020).

An area in the United States that is increasingly concerned about changing precipitation patterns is the southern Great Plains. This area has a Holocene history of megadroughts and dune activation at the eastern margin of the Great Plains, north-central Kansas, USA (Hanson et al.,

2010). Climate change simulations show that these dry periods could become more frequent and

1 intense in the future (Cayan et al., 2010). These intense dry periods and higher temperatures would decrease the snowpack in the Rocky Mountains, which would then decrease the meltwaters that would enter the Colorado River Basin the following spring (Cayan et al., 2010;

Karl et al., 2009). This would mean less water available for drinking, agriculture, and the economy in the Southern Great Plains of the United States. Therefore, it is vital that planning for the future water budgets in these areas occurs at the federal, state and local level in order to manage future water resources. An important method used in planning for future water resources is the examination of paleo climates and their corresponding precipitation patterns.

Recent studies indicate that isotopic evidence exists of wetter than modern conditions during episodes of the Pliocene and Pleistocene (Winnick et al. 2013; Ibarra et al. 2018; Lukens et al. 2019). For Pliocene time, the increased precipitation is hypothesized to be sourced in a persistent El Niño-like state, referred to as “El Padre” (Shukla et al., 2009), that would have dominated the eastern Pacific until mid-Pliocene time (Wycech et al., 2020).

In the northern hemisphere, glacial advances began to intensify in the late Pliocene, with a notably large glacial event known as M2 which is recorded in marine records (Lisiecki &

Raymo, 2005). The terrestrial record of the Pleistocene glacial/interglacial cycle is often associated with repeating cycles of arid and wet periods which can be seen in the form of lithologies, sediments, paleosols, isotope signatures and fossils of sedimentary rocks (Lukens et al., 2019; Shutter & Heckel, 1985; Johnson & Willey, 2000).

As mentioned above, there are many biological and chemical proxies which geoscientists use to study paleoclimates. One of these proxies is the nonmarine ostracode. Ostracodes are microcrustaceans that live in aqueous environments throughout the world. They are found in almost all aqueous environments, from saline ocean environments to freshwater streams, lakes,

2 and groundwater, to even urban environments like cisterns and birdbaths. Ostracodes also have a long geologic history, with a fossil record beginning in the Ordovician, and act as important paleoclimatic indicators (Humphreys, 2009). They vary widely in species assemblage and abundance and can be endemic to certain areas. This means that the presence of certain species of ostracodes can indicate what type of environment was previously present in the area (Forester et al. 2017). Also, the calcium carbonate shells of can be analyzed for stable isotope composition to determine groundwater flow paths, groundwater chemistry, and precipitation sources (Humphreys, 2009); (Smith and Palmer, 2012). These characteristics make ostracodes excellent paleoclimatic indicators. By conducting modern analogue analyses of fossil ostracode assemblages, paleontologists can determine many factors of paleo aquatic environments, such as water depth, salinity, dissolved ions, and the variability of these factors (Curry et al., 2012;

Mesquita-Joanes et al., 2012).

In this study, I use fossil ostracode assemblages from locations in Meade County, Kansas to test two hypotheses that relate to the hypothesis proposed by Ibarra et al. (2018). Ibarra et al.

(2018) observed through paleolake data compilation and modeling that Pleistocene glacial stages and the Pliocene epoch warm period share similar wet climatic and hydroclimatic conditions in western North America (Figure 1). They proposed that if similar increased precipitation patterns resulted from different temperature extremes (glacial stages vs Pliocene warmth), then wetter hydroclimatic conditions would occur compared to hydroclimatic conditions during interglacial periods (including present-day).

3

Figure 1: Comparison of lake extent during the Last Glacial Maximum and Mid-Pliocene. Green star indicates Meade County, Kansas, location of this study. Modified from Ibarra et al., (2018), figure 1.

Ostracodes are excellent indicators of both paleoclimate and groundwater sources

(Humphreys, 2009; Smith and Palmer, 2012) and would be an excellent proxy to compare to the

Ibarra et al. (2018) observations and test the following related hypotheses: that the Pliocene and the Pleistocene have similar hydroclimatic conditions, and that these conditions were wetter than modern hydroclimatic conditions.

I also will be addressing the following questions:

Are there patterns in ostracode species assemblage in Meade County, Kansas?

4

a. Are there spatial and temporal patterns in the ostracode species assemblages?

b. Are there species assemblages that are unique to certain stratigraphic ages?

c. Are there patterns in diversity seen in the ostracode species assemblages?

2. What do these species assemblages tell us of the paleoclimate in Meade County, Kansas?

a. Are these species assemblages similar or distinct to one another, and are they

similar or distinct to ostracode species currently living in Meade County, Kansas?

In the summer of 2019, I collected 13 sediment samples from Meade County, Kansas, 12 paleo samples and one modern sample, that provide some of the information needed to answer these questions. These 13 samples span about 4 million years of age, from the oldest sites in Fox

Canyon (~4.4my) to mid-Pleistocene Cudahy Ash (~0.62 my) (Lukens et al., 2019). These modern sample locations were combined with two other Plio-Pleistocene ostracode collections, that of Gutentag and Benson (1962) and of Claude Hibbard (1950’s), to create a dataset of 31 sample locations containing 31 ostracode species. Thus, this dataset consists of three collections

(now housed at Kent State U.): the collection I made in 2019, and the two older collections completed in the mid-20th century, by Ed Gutentag and Richard Benson (1962) and Claude

Hibbard (1950s). Stratigraphy and age dates follow Lukens et al., (2019) and species biogeography data were obtained from the Neotoma Database (Williams et al., 2018; www.neotomadb.org).

1.2 Pliocene-Pleistocene Paleoclimate and the Southern Great Plains of North America

Thompson (1991) gives a good overview of Pliocene climates throughout the western

U.S., and his findings reflect the general trend seen through the Pliocene into the Pleistocene.

The author describes how the Early and Late Pliocene saw warm and wet conditions, with frost- free winters and possible summer droughts (Thompson, 1991). What is most interesting about

5

Thompson’s (1991) work is that he describes the Great Plains paleoclimate in detail, which includes Meade County, Kansas. The paleoclimate in the Great Plains is described as an open savanna with sub humid-subtropical climates in the Early Pliocene (Thompson, 1991). These shifts in the Pliocene paleoclimate can also be seen in the evolution of hoofed mammals, as described in Janis et al. (2002). Here the authors describe the changes in hoofed mammal species from the Early Miocene through the Pleistocene as reflective of changing atmospheric CO2

(Janis et al., 2002). Species of the Great Plains are shown to have a maximum diversity in the

Middle Miocene (13.5Ma), which steadily declines throughout the rest of the Miocene, Pliocene, and Pleistocene (Janis et al., 2002). This decline reflects the declining global paleotemperature and is a good example of how paleontologists use fossil remains and other geological proxies to describe paleoclimate conditions (Janis et al., 2002).

Pliocene aquatic paleoenvironments in the western U.S. are described by the lacustrine ostracode assemblages collected by Forester (1991). In his study, Forester (1991) analyzed ostracodes from several paleo-lake sites and found that the presence of “exotic” taxa, meaning unusually ornate and endemic forms, indicated large, geologically stable lakes, which therefore indicate a geologically stable climate, at least regionally. The ostracode assemblages indicate a modern-like climate in the early Pliocene (5-4Ma) followed by a wetter than modern climate

(3.5-2.5Ma) and a return to modern conditions in the late Pliocene (2.5Ma) (Forester, 1991). This study is particularly interesting, since many of the species found by Forester (1991) in far western U.S. sites are also found in my samples from Meade County, Kansas, such as

Fabaeformiscandona rawsoni, Limnocythere ceriotuberosa, and Fabaeformiscandona caudata.

The evidence presented in the above studies concludes that the Pliocene paleoclimate in the western United States was wetter and warmer than the modern climate, with a glaciation event in

6 the Mid-Pliocene, the M2, being the first major glaciation in the Northern Hemisphere (De

Schepper et al., 2013).

In the past 2.6 million years, glaciers have expanded and retreated many times across the northern half of the continent. There were two main glacial bodies that extended into North

America during this time; the Laurentide ice sheet in the east and the Cordilleran glaciers in the west (Anderson et al., 2013). These ice sheets repeatedly expanded and contracted, resulting in multiple “ice ages” in North America. The pre-Illinoian stage saw glacial ice covering one-third of North America, extending into the northeast region of Kansas (Lyle, 2009). This ice moved forward through Kansas in an irregular pattern and was most likely only tens of meters thick at this latitude (Dort, 2006). This movement of glacial ice caused ecological shifts in Kansas from tundra to savanna-like grassland environments (Lyle, 2009).

The best detailed records of the Plio/Pleistocene glacial/interglacial cycle are marine records, which can be “stacked” to produce a truly global signal. Lisiecki & Raymo (2005) present the LR04 stack, which contains benthic δ18O marine isotope records from 57 sites from around the world. This stack shows glacial-interglacial cycles from 5.5 Ma to the present and is shown in Figure 2. We can see on the left of Figure 2 that the cycles in the Pliocene are less severe than in the Pleistocene. Then, around 3.5Ma, we see a slight increase in temperature range, which reflects the onset of the first significant glacial event in the northern hemisphere

Pliocene, known as M2 (De Schepper et al., 2013). In the Pleistocene, however, the glacial- interglacial cycles begin to fluctuate more extremely. We can also see how in the beginning of the Pleistocene, the 41kyr obliquity cycle dominated, producing more uniform temperature swings compared with the later dominance of the 100 kyr eccentricity cycle that followed it

(Lisieki & Raymo, 2005)

7

Figure 2: LR04 Stack of 57 globally distributed benthic marine δ18O records from Lisiecki, &

Raymo (2005). Samples used in this study span the period from 4.5 Ma to the present. Summary figure representing Lisiecki and Raymo (2005) LR04 Benthic Stack, With permission under

Creative Commons CC-BY-SA-3.0, “Dragons flight”(Robert A.Rohdel, svg by Jo).

Pound et al. (2014) use lake and soil data along with PRISM3 modeling to show not only how these data improve the data-to-model fit of Pliocene climates, but to also show that the

Pliocene had wetter than modern climate conditions. The authors attribute the presence of larger and more numerous lakes present at the time, especially in Australia and Africa as an indication of a global climate that is wetter than modern (Pound et al., 2014).

8

Figure 3: The extent of glaciation in North America. Figure modified from Lyle, (2009), figure

1, (modified from Illinois State Geological Survey, 2008). The red dot indicates Meade County,

Kansas, the area of focus for this study.

The advance and retreat of glacial ice in Kansas had a pronounced effect on the area’s topography. The extent of glacial ice into Kansas can be seen in Figure 3. New rivers, such as the Big and Little Blue, formed on ice margins, and proglacial lakes formed from glacial melt

9

(Aber, 1991). Sediment deposited from the glacier filled in river valleys, and ice damming formed lakes such as Kaw Lake, which spanned at least 70 miles (Dort, 2007). The finer grained sediments left from the retreat of glacial ice formed loess deposits and sand dunes that covered the Midwest (Lyle, 2009). Erosion and deposition from the glacial ice affected the entire region of Kansas, including Meade County, located in the southwestern portion of the state. The shift of the glacial boundary in Kansas is reflected in the ecological shifts in the area during the

Quaternary. As the glacial ice advanced south to northeastern Kansas, biomes further to the north, such as coniferous forests also advanced southward (Lyle, 2009). The vegetation in

Kansas during the glacial cycles shifted from open prairies and grasses to tundra environments, with the height of glaciation seeing grasslands across most of Kansas and stands of hardwood trees along lowland areas.

Meade County lies within the Meade Basin, which was formed by dissolution of underlying evaporites (Lukens et al., 2019). The stratigraphy of Meade County, Kansas, shifts between carbonate and siliciclastic rock layers. This shifting stratigraphy indicates a changing climate in the region, with rock sequences of limestones, gravel layers, and carbonate mudstones.

The shifting stratigraphy also reflects shifts from fluvial environments to semiarid river systems and wetlands. Lukens et al. (2019) used lithofacies analysis, paleopedology and ichnology to investigate the Pliocene paleoenvironment in the Meade Basin. Their findings show that the climate conditions in the Meade Basin (and to an extent the Western Great Plains) were wetter than modern during the Pliocene.

For the Meade basin, the presence of massive caliche beds indicates a more arid environment in southwestern Kansas during the Pleistocene compared to other parts of the state

(Bayne et al., 1976) . Analysis of δ13C values from paleosols in southwestern Kansas indicate a

10 shift from warm, dry climates (0.048-0.028 Ma) in the area to cooler, wetter conditions (0.028-

0.014 Ma) and back to warmer conditions (0.013Ma-present) in the late Pleistocene (Layzell et al., 2015).

1.3 Previous Work in Meade County, Kansas

The study area for my project, Meade County, Kansas, has a rich fossil history and well- defined lithology and stratigraphy. Fossil collection in Meade County, can be traced back to the late 1800s. The first person to collect fossils in the area was Orestes St. John in 1877 (Hibbard &

Taylor, 1960). Others such as Erasmus Haworth, G.I. Adams and H.T. Smith also contributed to the geologic work done in Meade County (Rempel, 2018). Ostracode microfossils were first collected in Meade County by Smith (1941) and later by Frye and Hibbard (1941; Gutentag and

Benson, 1962). Meade County is known for its abundance of rodent fossil assemblages, as well as ostracode microfossils. Hibbard worked in the area on vertebrate and invertebrate fossils, including fossil assemblages throughout the 1950s. Further work with ostracodes was done by Edwin D. Gutentag and Richard H. Benson (1962), who took sediment samples from across Meade County and analyzed them for fossil ostracodes. The field locations in this work are based on the locations used by Gutentag and Benson (1962). Their stratigraphic model is shown in Figure 4.

Geologic work in Meade County was not limited to the fossil assemblages present there.

Stratigraphic work had also been conducted in the area by Hibbard, Frye and Smith (Frye &

Hibbard, 1941; Smith, 1940). Smith (1940) did extensive research on the southwestern portion of

Kansas and describes the stratigraphy in the area from the Tertiary to the Modern. Lithologies of these layers include limestone, dolomite, sandstone, shale, chalk and volcanic ash (Smith, 1940).

Smith’s focus in his 1940 research was a general study of the area, incorporating geology,

11 topography, climate and economic resources present in southwestern Kansas. While Smith

(1940) looked at the stratigraphy of 14 counties in southwestern Kansas, including Meade, Frye and Hibbard focus on the stratigraphy of Meade Basin itself.

Recent work in Meade County has focused on revising the stratigraphic classification of the lithology in the area. Layzell et al. (2017) updated the classification of these layers from the classic Pleistocene glaciation model proposed by Zeller (1968) to one that considers pedo- stratigraphy, fossil assemblage of the layers and the identification of the Pearlette ash as made of six distinct ash beds with significant differences in age (Layzell et al., 2017). This revised stratigraphy can be seen in Figure 5. These stratigraphic columns from the Layzell et al. (2017) study reduces the number of formations and reorganizes other formations.

Previous studies such as Smith (1940), Frye and Hibbard (1941) and Gutentag and

Benson (1962) have used stratigraphy, fossil assemblage and (in the case of Gutentag and

Benson) ostracode species assemblages to recreate the paleoenvironment that existed in Meade

County KS during the Quaternary. The most recent work that has been done in this area is the work done by Lukens et al. (2019) as described previously. It is the stratigraphic column presented in Lukens et al. (2019) that I will use to place my collected samples in stratigraphic context.

12

Figure 4: Stratigraphic column from Gutentag and Benson (1962, Figure 3). Presented with permission from the publisher.

13

Figure 5: Stratigraphic columns from Layzell et al. (2017), featuring the Zeller (1968) stratigraphy and the revisions on this stratigraphy by the Kansas Geological Survey (Layzell et al., 2017, p.3). Presented with permission from the publisher.

14

2. Methods

2.1 Fieldwork, Stratigraphy and Chronology

In the summer of 2019, I traveled to Meade County, Kansas with Dr. Alison Smith of the

Kent State Geology Department, Dr. Tony Layzell of the Kansas Geologic Survey, and his graduate research assistant Catherine Opalka. We collected 12 samples of unlithified sediment, which were labeled KS01-KS12, and 1 modern vernal playa sample labeled KSM was collected from this area (Figure 6) (Table 2). These samples were placed in labeled bags and shipped to the

Kent State University Paleolimnology lab in Kent, Ohio.

The 2019 Meade County field sites were selected from locations previously sampled by

Gutentag and Benson (1962) and Hibbard (1950s). These sample sites were all in Meade County, and their stratigraphy was correlated with the updated Lukens et al. (2019) stratigraphy. The work done by Lukens et al. (2019) and by Layzell et al. (2017) not only renames the formations in the stratigraphic column, but also gives more definite ages to the stratigraphy based on rodent fossil assemblages and the dating of volcanic ash beds in the county. This allowed me, in some cases, to assign specific ages to the 12 fossil ostracode assemblage samples collected in 2019 in

Meade County.

The slides from the Gutentag and Benson (1962) and Hibbard’s (1950’s) collection was correlated to the revised stratigraphy of Layzell et al. (2017). For the Gutentag and Benson samples, ages were correlated using slide labels and the 1962 Gutentag and Benson paper, which lists the stop number for each sample site along with its approximate age. Also included in the

15

Gutentag & Benson slide collection was a list of slide numbers, their associated stop number and their township and range location in Meade County. Hibbard’s slides followed a similar pattern, with sample number and location listed on each slide. These locations did not have corresponding township and range values, but their formation name was correlated to stratigraphic columns previously published (Hibbard, 1949; Hibbard, 1958; Smith & Friedland,

1975; Miller, 1976; Martin et al., 2002).

Figure 6: Map showing location of sampling sites in May 2019 in Meade County, Kansas.

16

Table 1: Sample names, locations, ages and formations in Meade County, KS collected in June

2019. Ages of Samples and Stage/ North American Land Mammal Ages (NALMA) from

Lukens et al. (2019), Figure 2.

Sample Location of Age of Stage/

Name Sample Sample NALMA*

T-KSM Vernal Pond Modern Modern

T- Cudahy Ash 0.64Ma (Lava Creek Ash) Illinoian /

KS01 Mid-Pleistocene (1.5-0.2Ma) Irvingtonian

T- Cudahy Ash 0.64 Ma (Lava Creek Ash) Illinoian /

KS02 Mid-Pleistocene (1.5-0.2Ma) Irvingtonian

T- Cudahy Ash 0.64Ma (Lava Creek Ash) Illinoian/

KS03 Mid-Pleistocene (1.5-0.2Ma) Irvingtonian

T- Borcher's Badlands 2.11-2.14 Ma

KS04 (Huckleberry Ridge Ash) Pre-Illinoian/

Early Pleistocene 2.6 -0.77 Ma Late Blancan

T- Borcher's Badlands 2.11-2.14 Ma

KS05 (Huckleberry Ridge Ash) Pre-Illinoian/

Early Pleistocene 2.6 -0.77 Ma Late Blancan

17

Sample Location of Age of Stage/

Name Sample Sample NALMA*

T-KS06 Rexroad 3 3.04-3.11 Ma\ Wolf Gravel

(Magnetic Polarity Time Scale) Middle

Upper Pliocene (3.2-2.7Ma) Blancan

T-KS07 Rexroad 3 3.04-3.11 Ma\ Wolf Gravel Middle

(Magnetic Polarity Time Scale) Blancan

Upper Pliocene (3.2-2.7Ma)

T-KS12 Deer Park 3.04-3.11 Ma\ Wolf Gravel Middle

Upper Pliocene (3.2-2.7Ma) Blancan

T-KS08 Keefe Canyon 4.18-3.58 Ma Middle

(Magnetic Polarity Time Scale) Blancan

Mid-Pliocene (4.5-3.2Ma)

T-KS09 Keefe Canyon 4.18-3.58 Ma Middle

(Magnetic Polarity Time Scale) Blancan

Mid-Pliocene (4.5-3.2Ma)

T-KS10 Keefe Canyon 4.18-3.58 Ma Middle

(Magnetic Polarity Time Scale) Blancan

Mid-Pliocene (4.5-3.2Ma)

T-KS11 Fox Canyon 4.48-4.62 Ma Bishop Gravel Early

Mid-Pliocene (4.5-3.2Ma) Blancan

18

2.2 Sediment Processing Methods

The sediment processing took place in the paleolimnology lab at Kent State University.

All samples processed in the lab were recorded in the lab notebook. Sample processing followed the methods described in Forester (1988) and Coleman et al. (1990), with latex gloves worn for protection. When the samples first arrived at the lab, they were unpacked and set out in numerical order, with the modern sample labeled as “KSM” and with sediment bags placed in line from KS01 to KS12 with KSM at the end. This allowed for easy organization when processing sediment samples and recording them in the logbook.

First, the weight of one empty gallon-sized plastic bag was taken and recorded, and then the weight of the sediment sample bags was recorded. These numbers were then used to calculate the weight of the sediment sample in grams. Next, 20-mesh, 230-mesh, and 100-mesh sieves were used to wash and sieve the samples. These mesh sizes correspond to 850, 150 and 63 micrometer openings. The 63 micrometer (230 mesh) sieve corresponds to the sand/silt grain size boundary. Ostracodes are typically collected at the 150 micrometer (100 mesh) sieve. The sieves were stacked on top of each other, with the/ 20-mesh being on top, the 100-mesh in the middle, and the 230-mesh on the bottom. The sieves were then placed into the laboratory sink to wash the sediment.

On two Whirl-packs® bags, the sample number, date of collection, the sample location, and the sieve number 230 or 100 was written on the bags for identification (there was typically no need to prepare a 20 mesh bag-all sediments passed through that sieve). Each Whirl-pack® was then placed into a 1000mL beaker to keep them open for sample collection. The sample sediment was then poured in increments onto the 20-mesh sieve and a small showerhead attached to the sink was used to wash the sediment through the sieves. Once the sediment was washed

19 through the 20-mesh sieve, it was removed from the stack and the sediment was washed through the 100-mesh sieve. The same process was repeated until all the sediment was washed through the 100-mesh sieve, and then it was removed from the stack and set on the counter once the water had drained. Once the sediment in the 230-mesh sieve had been washed, the sieve was lifted from the sink and tilted slightly so that the showerhead could wash the remaining sediment to the bottom edge of the sieve. Then the shower head was turned off and the remaining 230- mesh sediments were washed into the bag labeled “230m” with deionized water. The same was done with the 100-mesh sediments. This process was repeated until either the entire sediment sample had been washed or a large amount of the sediment had been washed. If only some of the sediment had been washed, the weight of the remaining sediment was taken and used to determine the weight of the washed sediment. Before washing the next sample, the sieves were placed in an ultrasonic cleaner for 5 minutes each to remove any remaining sediment from the previous sample.

After all the sediment samples had been washed, the deionized water remaining in the

Whirlpak ® was drained into 200 mL beakers. Once most of the deionized water was removed, the remaining water in the Whirlpak ® was drained through a funnel containing a fine-woven polyester filter, which was placed in the Whirlpak® bag to prevent any sediment loss. This cloth was then placed into the whirl bag, and the bags were dried, sealed, and placed in a Virtis

Benchtop Research Freeze dryer.

20

2.3 Vitris Benchtop Research Freeze Dryer Operation

To prepare the freeze dryer for the sediment samples, the freeze dryer itself had to be prepared. The vacuum pump oil level was checked and adjusted if needed. This oil allows the vacuum pump to function, so it is important that the oil levels are not low. After the oil levels are double checked, the “Refrigerate” switch was turned on and when the condenser light indicated

“OK” the chamber was cool enough to proceed. After double checking that the trays were clean and free of debris or sediment, the sediment samples were added onto the trays in the vacuum chamber. Once all the samples were in place, the polycarbonate lid was placed onto the top of the chamber, and the vacuum switch was turned on. It is important to check that the neoprene ring stays on the vacuum chamber rim, and that no gaps or slips occur between the rim, the neoprene ring and the polycarbonate lid. If there are gaps present, then the vacuum will not properly seal and there is a chance that the machine could be damaged. The freeze dryer operated for 24 hours. Samples were checked to determine if freeze-drying was complete, and if so, the samples were removed. To remove the samples from the vacuum chamber, the vacuum switch was turned off, the lid was removed, the sediment samples were taken out of the chamber, and the refrigeration switch was turned off. After the process was completed, the neoprene ring, trays, and vacuum chamber were cleaned for future use.

2.4 Sample Picking and Ostracode Identification

Once the sediment samples had been freeze-dried, ostracode valve picking and identification could begin. Using a microscope, ostracode valves were separated from the sediment grains at roughly 20x magnification. A small amount of sediment (1-5 grams) was placed in a black, gridded microfossil picking tray, which was then placed under the microscope for analysis. When a valve was located, a fine (0000) sable paintbrush was dipped in deionized

21 water and used to carefully pick the ostracode valve from the sediment. The valve was then placed on a new microfossil slide. The slide was labeled the same way that the sediment bag was labeled, with sample number, the site location, the year and the mesh size. Of the 13 samples collected in 2019, only 7 samples contained ostracode valves: TKS01, TKS02, TKS08, TKS10,

TKS11, TKS12, and TKSM.

2.5 Older Collections

Two older ostracode collections were incorporated with my sample locations to build the ostracode database for this project. These older collections are the Gutentag and Benson ostracode collection (1962) and the Dr. Hibbard ostracode collection (1950’s). These microfossil slide collections are housed in the Paleolimnology lab at Kent State University. Both collections contain samples taken from locations across the Great Plains, including Meade County, Kansas.

Of these slides, each one had its name and location logged into an Excel file, and the number and type of ostracode species per slide were counted and recorded in the same file. Any sample slides that were not collected in Meade County, still had their ostracode assemblages identified and counted but were not incorporated in any statistical analysis. There were 12 microfossil slides from the Hibbard collection and 24 slides from the Gutentag and Benson collection that were counted but not used in the statistical analysis. This is because these microfossils either had no provenance or their location in the stratigraphic column was unknown.

22

2.6 Species Identification

Once separated from the sediment, fossil ostracode species were identified using a microscope at roughly 40x magnification. Ostracode species were identified using the type collections in the Paleolimnology lab at Kent State University, as well as with monographs from

Delorme (1970), Forester (1991), Forester et al., (2017), Furtos (1933), Ferguson (1967), and

Hoff (1942) and with photographic keys by Smith & Delorme, (2009) and Smith and Horne,

(2016). Each valve was compared to each type slide, monograph and photographic key, and when identified was added to the count for that species in the Excel table. Ostracode valve fragments and juveniles were also identified to the genus level but were not attributed to an individual species due to lack of carapace characteristics and were therefore not included in the statistical analysis.

2.7 Statistical Analysis

In this study, statistical analysis was run on the Multivariate Statistical Package computer program (MVSP) created by Kovach Computing Services (2002). MVSP is a computer program that performs a variety of ordination and cluster analyses, and includes distance measures used in modern analog analysis, cluster analysis and several diversity indices analysis for ecological data

(Kovach, 2002). In this study the Shannon’s diversity analysis, modern analog analysis, and cluster analysis (constrained and unconstrained) were computed on the MVSP software. This statistical software allows me to analyze the fossil ostracode assemblages and create an estimation of the paleo hydroclimatic conditions in Meade County during the Pliocene-Modern time span.

23

A cluster analysis was used to analyze the ostracode species data through time. The cluster analysis “…assemble(s) observations into relatively homogenous groups or ‘clusters,’ the members of which are at once alike and at the same time unlike members of other groups”

(Davis, 2002, p.488). Results from this analysis are shown in a dendrogram, where groups of individuals can be clearly seen. In the cluster analysis, the raw counts of the ostracode species were converted into percentages, so that the data could be better visualized in MVSP. This made it possible to visualize differences in groups and identify outliers that may be present.

2.7.1 Diversity Analysis

Shannon’s diversity index was used to determine the diversity of the ostracode species assemblages collected in this study. The analysis was run on the MVSP Version 3.1 (Kovach,

2002) software. This diversity index measures the evenness and the number of species per sample, along with the index value. Index values range from zero (no evenness, low diversity) to one (complete evenness, high diversity). The evenness of the species assemblage refers to the number of ostracode valves per species compared to the total number of valves. A high evenness indicates that the ostracode valves are evenly distributed throughout the species present. A low evenness score indicates that one or more species has a larger number of ostracode valves than the other species present in the sample. This analysis focuses on each ostracode species assemblage per sample site, listing all sites together to show which site had the highest diversity

(see Results section).

24

2.7.2 Classic Modern Analog Method

A classic modern analog analysis using Jacquard’s coefficient and the NANODe data

(Forester et al., 2005; www.kent.edu/nanode) was used to determine which modern ostracode assemblages were most like our paleo ostracode assemblages. The Neotoma database (Williams et al., 2018; www.neotomadb.org) was used to spatially map out the modern analog sites to determine which locations and climates were most like the sites in Meade County.

The modern analogue analysis of ostracode species was conducted using four of the 30 ostracode samples included in the statistical analysis. These four samples are the TKS01,

HNWSE 33-29-16db, G719561, and G630572 (10C). These samples were selected to span in age from the Upper Pliocene to the Upper Pleistocene, with one modern sample added for comparison. In this analysis, these samples were compared with over 600 ostracode surface samples from the NANODe database. The modern analogue analysis was run on the MVSP

Version 3.1 (Kovach, 2002) software. The binary measurement used for the modern analogue analysis was the Jaccard’s coefficient. This coefficient, described below, uses the presence/absence of a species in the two datasets to determine the similarity between the datasets. Jaccard’s coefficient was used because it accounts for presence and joint-presence of ostracode species in samples and omits joint-absences of species in samples from the analysis

(Weissleader et al., 1989). Below shows the formula for Jaccard’s coefficient (Kovach, 2002).

a Jc = ij (a+b+c)

This analysis identifies which assemblage samples in the modern surface sample dataset are most like the fossil samples. The environments represented by the most similar modern surface samples serve as estimates or analogs of the environments associated with the fossil

25 samples. This allows me to estimate if the paleoenvironment in Meade County, Kansas was wetter than the modern climate in the area. The modern analogue also tells me about the change in precipitation and temperature in Meade County from the Quaternary to modern times (See

Results).

2.7.3 Cluster Analysis

In this study I used a constrained cluster analysis with the Squared Chord coefficient and the farthest neighbor clustering method on the MVSP Version 3.1 (Kovach, 2002) software. This analysis measures the distance that each sample has to one another, and the distance each cluster group is to one another (Manly, 2016). In this cluster analysis, the farthest neighbor distance was used, which merges two groups based on the similarity of the two farthest individuals in the samples being compared (Manly, 2016). The coefficient used in this cluster analysis was the

Squared Chord coefficient. The Squared Chord coefficient was chosen for this analysis because it measures the distance between the ostracode samples in a way that emphasizes the “signal” component of the data and reduces the “noise” component of the data (Overpeck et al., 1985).

This means that the Squared Chord measurement places more emphasis on the abundant/common species (the “signal”) over the rare species (the “noise”). The Squared Chord distance measure is seen below (Kovach, 2002).

n 2 SCdij= ∑ (√xik -√xjk) k=1

Two cluster analyses were conducted in this study: the constrained cluster analysis and the unconstrained cluster analysis. Both analyses group the ostracode samples together based on species assemblage similarity, but the constrained cluster analysis keeps the samples in

26 stratigraphic order, and the unconstrained cluster analysis does not. This means that the constrained cluster analysis shows us how the ostracode assemblages have changed through time, and the unconstrained cluster analysis tells us the dominant habitats and environments in our data based on ostracode species assemblage.

2.7.4 NANODe and Neotoma databases

The North American Non-Marine Ostracode Database Project (Forester et al., 2005; www.kent.edu/NANODe ) is an online database that contains information on 89 species of ostracodes from over 600 North American sites. NANODe has biogeographic information on the distribution of extant species of ostracodes in the United States, heir associated major ion water chemistry, and the species distribution in solute space (Total Dissolved Solids (TDS as mg/L) vs

Carbonate Alkalinity (HCO3 + CO3/Ca as meq). Additional data from NANODe provided by the Neotoma database (Williams et al., 2018; www.neotomadb.org) includes the latitude and longitude of each site, number of species present, elevation, mean annual temperature and precipitation, major ion water composition and precise values for Total Dissolved Solids (TDS in mg/l) and bicarbonate & carbonate to calcium ratios ( (HCO3 +CO3)/ Ca in meq/l). The data from the NANODe sites are used to determine the location of the paleo ostracode species assemblages in solute space in the form of a (HCO3 +CO3)/ Ca vs. TDS graph (See Results

Section), which follows Smith, (1993), and Curry et al., (2012).

The Neotoma Paleoecology Database is a database containing a variety of biological proxy data of Pliocene-Holocene time, and modern species distributions as well. It contains information on a wide range of diatoms, pollen, insect and other proxies from a wide variety of sources. Neotoma also has information on ostracode surface samples from NANODe and from other databases, including the Delorme autecological ostracode database-now housed in the

27

Canadian Museum of Nature. As one aspect of this project, I used the Neotoma Explorer tool to visually map out the modern distribution of extant species found as fossil in Meade County samples. By doing this, I saw which areas in the United States today have similar ostracode species assemblages, and saw what the climate is like in these areas in order to estimate what the paleoclimate was like for my sites in Meade County, Kansas.

28

3. Results

3.1 Stratigraphy and Chronology

Samples from Meade County, Kansas were collected at previously chosen locations where their age and stratigraphy were known (See Methods Section 2.1). Figure 7 shows the

2019 ostracode sample sites correlated to the Lukens et al. (2019) stratigraphic column, which shows the Miocene through the Mid-Pleistocene stratigraphy in Meade County. The Hibbard

(1950’s) and Gutentag and Benson (1962) assemblages were correlated to the stratigraphy through various sources (See Methods Section 2.1) and were added to the 2019 sample pool.

Table 2 shows all the sample sites used in the analysis, as well as their stratigraphic location and age. The youngest ostracode samples are modern grabs from aquatic environments, and the oldest ostracode samples are Mid-Pliocene samples dating from 5.3-2.6 Ma. Only one sample, a

Hibbard sample named “Meade County State Park” was from the Holocene.

29

Figure 7: Stratigraphic column depicting 2019 sample site locations. Modified from the stratigraphic column in Lukens et al., 2019 (Figure 2. pg. 418).

30

Table 2: Names, stages, labels, and ages of 31 ostracode samples in this study from Meade

County, Kansas. Notice in the sample name and label that samples beginning with “T” were collected in 2019, samples beginning with “G” are from the Gutentag and Benson collection, and samples beginning with “H” are from the Hibbard collection.

Sample Name Age Sample Label Stage / NALMA*

T-Windmill-R Modern TKSMR

G-SkimSandBuriedArtesianBoil-R " GSSBABR

G- " GBSBBMR ConcreteDitchBuriedArtesianBoil-R

G-BigSpringsBola(BM)-R " GBSBBMR G-StreamfromArtesianBoil-R " G66571R

H-MeadeCounty-R " H526585R

H-MeadeCountyStatePark-H Holocene 11.7kyr HHoloH to present H-BigSprings1-PE Late Pleistocene HUM-K2-59PE Sangamonian 125-75kyr bp/ Rancholabrean H-BigSprings2-PE " HUM-K1-60PE " H-CraqinQuarry-PE " HCQPE " G-AboveCreekBed-PE " G630571,10BPE " G-CreekBed-PE " G628572PE " G-GullynearPond-PE " G630572,10CPE " G-HibbardDownCanyon-PE " G719561PE " G-JonesPas2-PE " G920572PE " G-JonesPas3-PE " G920573PE " H-ButlerSprings1-PE Middle HUM-K4-59PE Illinoian 1.5-0.2my Pleistocene bp/Irvingtonian H-ButlerSprings3-PE " HUM-K1-57PE " H-ButlerSprings4-PE " H35-29-5babPE " H-ButlerSprings5-PE " HUM-K3-61PE " H-Butler Springs6-PE " HUM-K1-55PE " T-Cudahy1-PE " TKS01PE " T-Cudahy2-PE " TKS02PE " H-CudahyAsh " HCFAMPE " H-Rexroad1-TPI Late Pliocene HUM-K1-59TPI Late Blancan H-Rexroad1-TPI " HTypeSecTPI " H-CotrellPastureRexroad-TPI " G12-35-31TPI " G-RexroadForm-TPI " G12-35-31TPI " T-KeefeCanyon8-MPI " TKS08MPI " T-KeefeCanyon10-MPI " TKS10MPI "

31

Sample Name Age Sample Label Stage / NALMA* T-FoxCanyon11-MPI Mid Pliocene TKS11MPI Middle Blancan

3.2 Ostracode Species, their stratigraphic location and ecological significance

In this study there were a total of 31 ostracode species found in the Pliocene-Holocene deposits of Meade County, KS. Photomicrographs of common species in this collection are shown in Figure 9. The fossil ostracode species collected in this study are all extant in North

America, and their bicarbonate & carbonate/calcium ratio and total dissolved solids (TDS) value ranges are listed in Table 3 (www.neotomadb.org). Table 4 shows these species in stratigraphic order, from the Mid-Pliocene to Modern times. Almost all the species collected represent freshwater environments, such as wetlands, ponds, streams, littoral zones, and springs. Some species, like Cyprideis salebrosa, Sarscypridopsis aculeata, and Limnocythere ceriotuberosa represent oligohaline to saline environments, with Pelocypris tuberculata found in modern, seasonally filled freshwater playas in the southwestern U.S. (Smith & Horne, 2016).

Looking at Table 3, we can see the bicarbonate & carbonate/calcium ratio and TDS ranges for each ostracode species, as well as species ecology. As shown in this table, these ranges overlap each other for many ostracode species, and many of the environments of individual ostracode species also overlap. For example, species such as Paracandona euplectella and Pseudocandona stagnalis have similar bicarbonate & carbonate/calcium ratios and TDS ranges (Table 3). Species assemblages, not individual species, are the most useful in identifying environmental conditions.

32

Figure 8: On Following Page: Photomicrographs of Common Species of Ostracoda in Meade

County, Kansas.

A) Paracandona euplectella (Robertson, 1889) Left valve, female, KS-01 Cudahy Ash, collected by Tomin May 2019, transmitted light.

B) Candona fluviatilis Hoff, 1942 Left valve, female, KS08, Keefe Canyon, Collected by Tomin, May 2019, transmitted light.

C) Fabaeformiscandona caudata (Kaufmann, 1900) Left valve interior, female, Butler Springs, Hibbard’s salamander site, collected by Gutentag, transmitted light.

D) Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 1995 Left valve, female, Butler Springs, XI Ranch, Hibbard’s sloth locality, collected by Gutentag, transmitted light.

E) Eucypris meadensis. Gutentag & Benson, 1962 Right valve interior, female, Skim Sand buried Artesian boil, collected by Gutentag, transmitted light.

F) Scottia pseudobrowniana Kempf, 1971 Left valve, Cudahy Ash, KS-01, collected by Tomin, May 2019, transmitted light.

G) Cyprideis salebrosa (Van den Bold, 1963), Right valve, male, 719561, Hibbard’s site down canyon, collected by Gutentag, transmitted light.

H) Pseudocandona stagnalis (Sars, 1890), Meisch and Broodbakker, 1993, Right valve, Rexroad Formation, Cottrell Pasture, 33-29-16db, collected by Hibbard, transmitted light.

33

34

Table 3: Ostracode species collected in Meade County, Kansas, and their ecological significance. References used in this table are a)

Smith and Horne (2016); b) L.D. Delorme (1970); c) Furtos (1933); e) C.C. Hoff (1942); and f) E.R. Ferguson (1967).

Ostracode species list with updated taxonomic names Carbonate Alkalinity/ Ca (meq) TDS (mg/L) Ecological Distribution Spirocypris tuberculata (Sharpe, 1908) 1.09-6.52 243-960 springs, groundwater fed ponds a, b, d Cyprideis salebrosa (Van den Bold, 1963) 0.02-2.5 190-35000 wetlands, saline springs, saline littoral a, b, d Cypridopsis vidua (O.F. Muller, 1776) 0.09-30 10-1000 aquatic settings, shallow a, b, d Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol et al., 1987 0.4-8.5 9.5-5000 wetlands, shallow ponds, lakes a, b, d Candona crogmaniana (Turner 1894) 0.2-2.5 150-1500 springs, groundwater fed ponds a, b, d Candona fluviatilis (Hoff 1942) 0.2-2.5 150-750 streams a,b, e Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 1994 0.08-50 150-45000 wetlands, shallow ponds, lakes a, b, d Candona renoensis (Gutentag and Benson, 1962) 3.0-7.0 900-1500 littoral zonesa, b, d Eucypris meadensis (Gutentag and Benson, 1962) 0.75-9 280-750 springs a, b, d Ilyocypris bradyi (Sars, 1890) .075-.25 120-2500 springs, streams, littoral zones a, b, d Ilyocypris gibba (Ramdohr, 1808) 0.075-25 120-2500 springs, streams, littoral zones a, b, d Limnocythere staplini, (Gutentag and Benson, 1962) 0.03-6.5 450-42000 pothole lakes, littoral zones a, b, d smaragdina (Vavra 1891) 0.3-5 55-3500 springs, streams a, b, d Scottia pseudobrowniana (Kempf, 1971) 0.9-1.9 200-350 hydrated fen wetland soils, charophyte mounds, weedy ponds a, b, d Dolerocypris sinensis (Sars, 1903) Data Not Available from North America Data Not Available from North America ponds, streams a, b, d Pelocypris tuberculata (Ferguson, 1967) Delorme 1969 Data Not Available from North America Data Not Available from North America seasonally filled playas of southwestern U.S. and vernal ponds of southern California a, b, f Candona inopinata (Furtos, 1933) 0.6-8 55-950 Temporary ponds and marshes a, b, c, d Paracandona euplectella (Robertson, 1889) 0.9-1.5 15-170 hyporheos of streams, springs, seeps, shallow groundwater a, b, d Fabaeformiscandona distincta (Furtos, 1933) Martens & Savatenalinton 2010 0.6-7 130-990 vernal ponds, lakes, wetlands a, b, d Pseudocandona punctata (Furtos, 1933) Swain 1994 1.0-3.0 100-450 springs, ponds, littoral zones of lakes a, b, c, d Candona acuta (Hoff, 1942) 0.55-3.5 140-1500 shallow littoral with subaquatic vegetation a, b, d Pseudocandona stagnalis (Sars, 1890) Meisch & Broodbakker 1992 0.15-5 43-2100 wetlands, springs, slow moving streams, littoral zones of lakes a, b, d Pseudocandona albacans (Brady, 1864) 1.0-1.7 69-680 springs, vernal pools, littoral zones of lakes a, b, d Sarscypridopsis aculeata 2.00 6700 saline prarie potholes and littoral zone of saline lakes a, b, d Physocypria globula (Furtos, 1933) 0.11-30 16-1000 shallow aquatic habitats, preferable with vegetationa, b, c, d Cyclocypris ovum (Jurine, 1820) 0.75-5 9.5-750 littoral zones of lakes, ponds a, b, d Herpetocypris brevicaudata (Kauffmann, 1900) 0.12-25 99-400 seeps and springs a, b, d Bradleystrandesia reticulata (Zaddach, 1854) Wouters 1989 1.11-2.96 51.9-1849 springs, ponds, ditches, littoral zones a, b, d Darwinula stevensoni (Brady and Robertson, 1890) 0.1-5 8.0-4000.0 cold and temperate lakes, nonmarine aquatic settings a, b, d Cypria exsculpta (Fischer, 1855) Data Not Available from North America Data Not Available from North America common in most fresh waterways, shady ponds with leaf litter, streams, lakes, springs a, b, d Limnocythere ceriotuberosa (Delorme, 1967) 2.8-120 450-10200 saline wetlands, lakes, prarie potholes a, b, d

35

Table 4: Ostracode species assemblages collected from Meade County, KS, and their appearance in the stratigraphic sequence from the Mid-Pliocene to the Modern.

Species Mid-Pliocene Upper Pliocene Mid-Pleistocene Upper Pleistocene Holocene Darwinula stevensoni (Brady and Robertson, 1890) x x Cyclocypris ovum (Jurine, 1820) x x x x Pseudocandona stagnalis (Sars, 1890) Meisch & Broodbakker, 1992 x x x x Candona crogmaniana (Turner, 1894) x x x Candona inopinata (Furtos, 1933) x x x Fabaeformiscandona caudata (Tressler, 1957) Griffiths, 1994 x x x Candona fluviatilis (Hoff, 1942) x x x Limnocythere ceriotuberosa (Delorme, 1967) x x x x Bradleystrandesia reticulata (Zaddach, 1854) Wouters, 1989 x x x x x Ilyocypris bradyi (Sars, 1890) and I. gibba (Ramdohr, 1808) x x x x x

Physocypria globula (Furtos, 1933) x x Fabaeformiscandona distincta (Furtos, 1933) Martens & Savatenalinton 2010 x x x Cypridopsis vidua (O.F. Muller, 1776) x x x x Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 1994 x x x x Herpetocypris brevicaudata (Kauffmann, 1900) x x x Cypria cf. exsculpta (Fischer, 1955) x x x

Paracandona euplectella (Robertson, 1889) x Sarscypridopsis aculeata (Costa, 1847) x Scottia pseudobrowniana (Kempf, 1971) x Spirocypris tuberculata (Sharpe, 1908) x x Pseudocandona albicans (Brady, 1864) x x Pseudocandona punctata (Furtos, 1933) x x Candona acuta Hoff, 1942 x x Eucypris meadensis (Gutentag and Benson, 1962) x x x Limnocythere staplini (Gutentag and Benson, 1962) x x x

Cyprideis salebrosa (Van den Bold, 1963) x Candona renoensis (Gutentag & Benson, 1962) x

Dolerocypris sinensis (Sars, 1904) x Pelocypris tuberculata (Ferguson, 1967) Delorme, 1969 x Potamocypris smaragdina (Vavra, 1891) x Physocypria globula (Furtos, 1933) x

36

3.3 Constrained Cluster Analysis

Constrained cluster analysis created classes of sample groups based on similarities in the ostracode species assemblage using the Squared Chord coefficient and the farthest neighbor distance measure. The result of this analysis can be seen in Figure 9. The analysis produced five different clusters in stratigraphic sequence, ranging from the bottom cluster (red) indicating the

Mid-Pliocene, up through to the Modern sample sites (purple). The Mid- and Upper Pliocene sites have only one main cluster, whereas the other age groups have two or more clusters. The distance between these clusters indicate different conditions with these different assemblages.

3.4 Unconstrained Cluster Analysis

The unconstrained cluster analysis illustrates similarities in species assemblages regardless of stratigraphic order. We can see in Figure 10 that there are six distinct species clusters, which indicate six distinct environments based on species ecology. The Upper Pliocene site Keefe Canyon can be seen in the bottom cluster, with the presence of Candona fluviatilis representing stream systems. Stop 5 Gutentag and Benson can be seen in its own distinct cluster, with the presence of Cyprideis salebrosa and Fabaeformiscandona rawsoni indicating shallow saline/ wetland environments. The Cudahy sites TKS01 and TKS02 have an abundance of

Paracandona euplectella, which places it in the hyporheic groundwater cluster. Other distinct environments that can be seen are shallow wetlands, ponds, springs, and ephemeral ponds.

37

Figure 9: Cluster analysis constrained by stratigraphic age using the Squared Chord coefficient and the Farthest Neighbor distance measurement. The colored sections represent the different stratigraphic ages of the samples, with red indicating the Mid-Pliocene

(4.5-3.2Ma), yellow indicating the Upper Pliocene (3.2-2.7Ma), green indicating Mid-Pleistocene (1.5-0.2Ma), blue indicating the

Upper Pleistocene (0.075-0.125Ma) and purple indicating Recent/Holocene (0.01Ma-present).

38

Figure 10: Unconstrained cluster analysis using the Squared Chord coefficient and the Farthest Neighbor distance measurement.

Notice the six distinct species assemblages present in our samples, indicating six distinct freshwater environments. Black boxes around the microfossil slide indicate areas of interest, including Keefe Canyon (TKS08), Stop 5 Gutentag and Benson (G630572

(10CPE), G628572PE, G719561PE, G630571 (10BPE)) and Cudahy (TKS01, TKS02, HCFAMPE).

39

3.5 Distribution of ostracode species assemblages in solute space

The water chemistry of extant ostracode species from NANODe (Forester et al., 2005; www.kent.edu/NANODe) was used to create a solute space graph in which the fossil ostracode assemblages of 3 sites were overlain. These three sites ranged in age from Upper Pliocene site

(Keefe Canyon), to Mid-Pleistocene (Cudahy, Pre-Illinoian) and Upper-Pleistocene (Stop 5

Gutentag and Benson, Sangamon). Water chemistry from modern sites with species that were also found in this study were plotted on a graph with the log of their bicarbonate & carbonate/calcium ratio values on the y-axis and the log of their Total Dissolved Solid (TDS) values on the x-axis. NANODe sites with ostracode species found in the three Meade County sites were plotted in solute space to see how water chemistry changed between these sites

(www.kent.edu/nanode). This can be seen in Figure 11, with each species represented in the graph, and with the total NANODe sites in grey in the background. The species found in each of the three fossil assemblages, which are also extant species, can be used to identify the position of the fossil aquatic site in solute space. To identify each fossil site being analyzed, their species assemblage has been circled in the graph, and labelled. The Pliocene site, Keefe Canyon, shows an assemblage plotted in the dilute portion of the graph, well below the calcite branch point (the first mineral branch point in natural waters, occurring at about 350 mg/l TDS (Eugster & Jones,

1979; Smith, 1993). The rising TDS beyond the calcite branch point at about 350 mg/l leads to two distinct branches, bicarbonate enriched, and bicarbonate depleted saline water. The

Sangamon site, “stop 5” collected by Gutentag & Benson (1962) plots along the rising saline and bicarbonate depleted branch of this diagram.

40

Figure 11: Modern distribution in solute space of species assemblages identified from sites in this study, from NANODe

(www.kent.edu/nanode). Both Keefe Canyon and Cudahy represent freshwater environments, whereas Gutentag and Benson Stop 5

(Sangamon, Upper Pleistocene) assemblage is more saline, with large differences in both TDS and alkalinity values, indicating a changing aquatic environment between these sites.

41

3.6 Shannon’s Diversity Index

Results from the Shannon’s Diversity analysis can be seen in Table 5. The sample with the

highest index value was HNWSE 33-29-16db, with a value of 0.718. This sample is from the

Hibbard collection, is in the late Pliocene and contains 12 ostracode species (Table 5). The

sample with the lowest index value was 0.053 for the Gutentag and Benson site G630572 10C,

in the Pleistocene. The index values vary throughout the stratigraphy, and overall show that the

samples collected in this study contain diverse fossil ostracode assemblages.

42

Table 5: Results from the Shannon’s Diversity Index analysis, with samples in stratigraphic order.

Sample Index Evenness # Species

TKSMR 0.626 0.896 5

GSSBABR 0.449 0.746 4

GBSB(BM)R 0.543 0.643 7

GCDBABR 0.479 0.796 4

G66571R 0.521 0.669 6

H526585R 0.507 0.843 4

HHoloH 0.414 0.688 4

HUM-K2-59PE 0.294 0.421 5

HUM-K1-60PE 0.594 0.764 6

HCQPE 0.449 0.942 3

G630571, 10BPE 0.331 0.549 4

G628572PE 0.062 0.206 2

G630572, 10CPE 0.053 0.175 2

G719561PE 0.349 0.58 4

HUM-K4-59PE 0.452 0.581 6

HUM-K1-57PE 0.614 0.726 7

H35-29-5babPE 0.391 0.433 8

HUM-K3-61PE 0.591 0.502 15

HUM-K1-55PE 0.688 0.815 7

TKS01PE 0.444 0.571 6

TKS02PE 0.601 0.773 6

HCFAMPE 0.366 0.766 3

HUM-K1-59TPI 0.249 0.827 2

43

Sample Index Evenness # Species

HTypeSecTPI 0.218 0.456 3

HNWSE 33-29-16db 0.718 0.665 12

TPI

G12-35-31TPI 0.452 0.946 3

TKS08MPI 0.176 0.368 3

TKS10MPI 0.301 0.999 2

TKS11MPI 0.507 0.532 9

3.7 Modern analog estimates of precipitation with NANODe sites

A modern analog analysis of NANODe sites with the sample sites was conducted to obtain estimates of modern environmental conditions that fit the fossil ostracode assemblages.

Data from the NANODe database provided calculated mean annual temperature and calculated mean annual precipitation for each NANODe site in the analog analysis. This included a calculated mean annual temperature and precipitation from Meade County. These calculated mean annual temperatures and precipitation values are 30 year averages, drawn from the USGS data and incorporated into the NANODe database (Forester et al., 2005; www.kent.edu/nanode).

This information can be seen in Table 6. The calculated mean annual temperature and mean annual precipitation for Meade County was recorded at the NANODe modern site Big Springs, with values of 13.7°C and 502mm. The NANODe sites that had ostracode assemblages most like our fossil assemblages had estimated mean annual temperatures ranging from 4.4°C to 16°C and had estimated mean annual precipitation values ranging from 265mm to 946 mm.

44

Table 6: Modern analog of NANODe sites with sample sites from Meade County, KS, including estimated mean annual temperature and precipitation.

Sites Age Stage Best Estimated Estimated Or North Modern Mean Mean American Analog Annual Annual Land (highest Temperature Precipitation Mammal Age Jaccard (⁰C) (mm) score) Big Modern Big 13.7 502 Springs Springs, (actual MAT) (actual MAP) Ranch Meade County, Kansas Gutentag Upper Sangamonian/ Horseshoe 6.1 534 Hibbard Pleistocene Rancholabrean Lake down (0.075- pothole, Canyon 0.125Ma) South Dakota (Site 206) Gutentag Upper Sangamonian Lazy 16 265 Gully Pleistocene Rancholabrean Lagoon Near (0.075- wetland, Pond 0.125Ma) New Mexico (Site 241) Tomin Mid- Illinoian/ Mantua 9.5 947 Cudahy Pleistocene Irvingtonian Bog (fen), 1 (0.62 Ma) Ohio Cisco Lake, 4.4 808 Wisconsin (Site 302) Cottrell Upper Little Wall 11.1 946 Pasture Pliocene Blancan Lake, Iowa (3.11 Ma) (Site 579)

45

3.8 Summary of Results

In this study, there are 31 ostracode microfossil samples from Meade County incorporated in the statistical analysis, with ages spanning from the Mid-Pliocene (5.3-2.6 Ma) to the present. Of these 31 samples, 6 were collected in the summer of 2019, 11 were from the

Gutentag and Benson collection, and 14 were from the Hibbard collection. The species collected from Meade County are freshwater ostracode species, and some species are extant but are not found in Meade County, KS today. Species such as Cyprideis salebrosa, Sarscypridopsis aculeata, and Limnocythere ceriotuberosa indicate shallow saline environments such as saline wetlands, seasonally filled playas, saline prairie potholes, and saline springs and littoral zones

(Pieri et al., 2009; Smith & Delorme, 2009). Ostracode species associated with springs and groundwater discharge were also identified in this collection, such as Paracandona euplectella and Eucypris meadensis (Gutentag and Benson, 1962) (Smith & Horne, 2016).

Figure 12 shows the distribution of ostracode species collected from the Meade County sites. In total, there were 7,421 ostracode valves counted , with 25% (1873) being either fragments or juvenile species that could not be identified. The most common species in the collection is the nektonic Cypridopsis vidua, an extremely common species in North America, with 1328 valves that made up 18% of all collected ostracode valves. Other common species include Paracandona euplectella (5%), Cyclocypris ovum (8%), and Bradelystrandiesa reticulata (6%) (Figure 12). These species are found in freshwater environments such as wetlands, springs, seeps, littoral zones, ponds, and streams. Other species, such as Cyprideis salebrosa (9%) and Fabaeformiscandona rawsoni (7%) indicate dryer, more saline environments. The overarching theme found in the ecology of these ostracodes is one of shallow, freshwater aquatic environments, with variations in environmental settings. All the species

46 collected indicate a wetter environment in Meade County in the Mid-Pliocene and Mid-

Pleistocene than is currently found in the area. These freshwater assemblages change throughout the Quaternary, and these changes reflect the changing paleoenvironment in Meade County, KS during this time.

47

Figure 12: The percent abundance of ostracode species collected from the Meade County sites in this study. Ostracode shell fragments and Candona juveniles make up 25% of all collected valves, with Cypridopsis vidua being the next largest making up 18% of the total collected ostracode valves.

48

4. Discussion

The change in ostracode species assemblage through geologic time is analyzed with the constrained and unconstrained cluster analysis (See Results Section 3.2-3.3). In the constrained cluster analysis (Figure 9) we see different ostracode species assemblage groups from the Mid-

Pliocene through to the Recent/Holocene. The grouping of microfossil samples is based on the similarity of ostracode assemblages, which are grouped closer together. The Mid- and Upper

Pliocene have one distinct assemblage group each, whereas the Mid-Pleistocene has three, and the Upper Pleistocene and the Recent/Modern have two. These individual groups indicate unique ostracode assemblages in each geologic stage. For example, the group containing TKS01 and

TKS02 in the Mid-Pleistocene had large amounts of Paracandona euplectella which today can be found living in springs, seeps or hyporheos of streams. These two microfossil samples are in their own node in the dendrogram since P. euplectella is absent or rare in all other samples. The different nodes in the dendrogram indicate different ostracode species assemblages, which then indicate different aquatic environments. With this understanding, when we look at Figure 9, we can see that the aquatic habitat in Meade County, KS shifted from the Mid-Pliocene through to the Recent/Holocene, and there are environmental shifts in each geologic stage.

While the constrained cluster analysis shows us how many distinct groups of ostracode assemblages were in each stratigraphic section, the unconstrained cluster analysis tells us about the aquatic environments represented in the microfossil slides (Figure 10). There are six distinct

49 aquatic environments represented in this collection; streams (Candona fluviatilis), shallow saline, wetlands (Cyprides salebrosa, Fabaefomiscandona rawsoni), springs (Eucypris meadensis,

Ilyocypris bradyi), shallow wetlands/ponds (Cypridopsis vidua, Cyclocypris ovum,

Pseudocandona stagnalis), hyporheic or groundwater discharge (Paracandona euplectella) and ephemeral ponds (Fabaeformiscandona rawsoni, Bradleystrandesia reticulata). The unconstrained cluster analysis shows not only which environments are represented in this study, but also which environment is the most common in the microfossil samples. For example, the two clusters representing springs and shallow wetlands/ponds have the largest nodes and consist of 14 of the 31 microfossil samples in the collection. Overall, except for the Sangamon interglacial samples, the aquatic environments represented by the ostracode microfossils in this collection are significantly fresher (more dilute) in the Pliocene-Pleistocene than the modern environment of Meade County.

The distribution of ostracode species in solute space from the Cudahy, Keefe Canyon and

Stop 5 Gutentag and Benson sites show patterns of change in the paleoclimate in Meade County,

KS. The Cudahy assemblage (Mid-Pleistocene) shows an ostracode species assemblage of B. reticulata and P. stagnalis (Figure 11), which indicate a freshwater environment with a TDS value below 500 mg/l. Significantly for this study, Cudahy’s associated date of 0.62 Ma corresponds to Marine Isotope Stage 16, a glacial stage (Railsback et al., 2015). The Keefe

Canyon assemblage (Late Pliocene) also shows an ostracode species assemblage with slightly lower TDS values, and includes the species P. euplectella, found in dilute flowing springs and hyporheos. We can see in Figure 11 that these two ostracode species assemblages are relatively similar and indicate an aquatic environment that decreased in TDS ( became more dilute) from the Mid-Pliocene to the Mid-Pleistocene. Stop 5 Gutentag and Benson (Upper-Pleistocene) has a

50 different ostracode species assemblage than the previous two sites, with no P. euplectella present but instead, C. salebrosa, F. rawsoni and L. staplini present along with B. reticulata and P. stagnalis. This addition of two ostracode species and the elimination of another indicates a shift in the aquatic environment towards increased salinity and bicarbonate depletion (Figure 11).

This indicates that in the Upper-Pleistocene, the ostracode species assemblage of Stop 5

Gutentag and Benson lived in a more saline aquatic environment than the environments represented by the earlier samples. A change in the aquatic environment could be from a variety of sources, such as a change in precipitation in the area leading to a rising water table and increased fresh groundwater discharge. An increase in precipitation from the Mid-Pliocene through the Mid-Pleistocene could result in the fresher aquatic conditions seen in Figure 11 in the Keefe Canyon and Cudahy assemblages. A decrease in precipitation from the Mid-

Pleistocene to the Upper Pleistocene could also explain the more saline assemblages of Stop 5

Gutentag and Benson. Previous work by Lukens et al. (2019) has shown that paleosol and isotopic analyses of Pliocene deposits in Meade County indicate wetter than modern conditions during the Pliocene, consistent with these results.

A modern analogue analysis of assemblages of ostracode species from four sites in this study shows an overall increase in precipitation in Meade County compared to modern precipitation (Table 6). The assemblage with the highest estimated mean annual precipitation was the TKS01 assemblage (Mid-Pleistocene Cudahy site), with a modern analogue value of 947 mm from the Mantua Bog site in northeast Ohio. The assemblage with the lowest estimated mean annual precipitation was the Gutentag ‘Gully near pond’ assemblage (Late Pleistocene), with a modern analogue of 265mm from the Lazy Lagoon Wetland, New Mexico (NANODe

Site 241). Table 6 shows fossil assemblages of ostracode species are most like modern wet,

51 temperate environments in sites Ohio (Mantua Bog (Fen) site), Wisconsin (NANODe Site 302) and Iowa (NANODe Site 579). Temperature ranges, as indicated in this modern analog analysis, show that the modern analog values are all within the modern ranges for Kansas today. The average precipitation of the three modern analogue sites (Sites 302, 579 and the Mantua Bog

(Fen) site) that correspond to the paleo sites (Gutentag and Hibbard Down Canyon, Tomin

Cudahy 1, and Cottrell Pasture) is 808.75mm, which is much higher than the 502mm value for the estimated mean annual precipitation in Meade County today. This tells us that the change in ostracode species assemblage is probably not due to a change in temperature range in Meade

County, but to a change in precipitation. The drier modern analog sites (Site 206 and 241) correspond to the fossil assemblages collected in Sangamonian age samples, an interglacial stage that shows evidence in this study of less precipitation than the mid-Pleistocene and

Pliocene samples. This supports the work done by Ibarra et al (2018), which concluded that the interglacial periods of the Pliocene-Pleistocene were drier than the glacial periods. We can see that Meade County Kansas received more precipitation in the Upper Pliocene and in the Mid

Pleistocene than it received in late Pleistocene Sangamon interglacial stage and today.

The ostracodes collected for this study represent freshwater environments such as streams, ponds and wetlands that are not currently recorded in Meade County today (Forester et al., 2005; www.kent.edu/nanode ), with the exception of vernal playas. These findings support other studies, in that the Pliocene/Pleistocene climate had wetter than modern conditions both in

Meade County, Kansas (Lukens et al., 2019; Ibarra et al., 2018).

This study not only supports the hypothesis proposed by Ibarra et al. (2018) but is also consistent with recent paleoclimate work for Meade County. The wetter than modern hydroclimatic conditions evidenced by the data in this study corresponds to the data presented in

52

Lukens et al. (2019), which demonstrates wetter than modern hydroclimatic conditions in Meade

County through paleopedology, ichnology, and lithofacies analysis. The paleosols analyzed by

Lukens et al. (2019) indicate a variety of wet environments, including palustrine, seasonal wetland, and flood plain environments, which are consistent with the ecology of the ostracode species of this study. Ichno-fossils found in the area, for example Planolites, also support the hypothesis of wetter-than-modern conditions in the Pliocene-Pleistocene, since these ichno- fossils indicate wetland environments (Lukens et al., 2019). Analysis of the lithofacies in Meade

County also support wet hydroclimatic conditions during the Pliocene-Pleistocene, with fluvial environments represented by the Bishop and Wolf gravels, and limestone and mudstone couplets indicating more palustrine conditions in between the two gravel members (Lukens et al., 2019).

This lithologic data corresponds to the ecology of the ostracode species in this study, which range from wetlands to rivers and other wet environments.

Data from this study also supports regional hydroclimate studies that indicate wetter- than-modern conditions during the Pliocene-Pleistocene. For example, the study by Forester

(1991) indicates a wetter and warmer climate for the Pliocene lakes of western North America, due to the presence of “exotic” ostracode species found in large, geologically stable lakes in the

Western U.S. Work done by Thompson (1991) indicates a sub-humid, open-savanna paleoclimate for the Great Plains in the early Pliocene, which is both wetter and warmer than the current climate of the Great Plains. Changing paleoclimates are evidenced in the study by Janis et al. (2002), which shows the shifting paleotemperatures through changing distributions of hoofed mammal species. These studies collectively show how the paleoclimate of the Pliocene-

Pleistocene was warmer and wetter than the modern climate, with climate shifts seen in the late

Pliocene and Pleistocene.

53

Taken together, the cluster analysis, species assemblages and modern analogue results indicate that the aquatic environments in Meade County during the Pliocene Epoch warm period and the Mid-Pleistocene glacial were fresher (more dilute) than the aquatic environments of the late Pleistocene Sangamonian and Modern environments in the same location, which were more saline. Therefore, the results of this study support the findings by Ibarra et al. (2018) that the

Pliocene Epoch warm period and the Pleistocene glacial stages have similar hydroclimatic conditions, and that these conditions were wetter than modern climatic conditions (Figure 13).

54

Figure 13. Summary of results: Pliocene sites and Mid-Pleistocene (Cudahy, 0.62 Ma, MIS 16) show dilute, freshwater assemblages associated with higher Mean Annual Precipitation compared with Late Pleistocene Sangamon (MIS 5) and Holocene assemblages. These results are consistent with the interpretations by Lukens, 2019; Snell, 2019; and Ibarra et al., 2018.

55

5. Conclusion

Ostracode specimens were collected from 12 paleo sites and 1 modern site in this county and were combined with historical collections from Gutentag and Benson (1962) and Hibbard

(1950’s) in order to test the hypothesis proposed by Ibarra et al. (2018) that the Pliocene Epoch warm period and Pleistocene glacial stages shared similar wet climatic conditions. Cluster analysis revealed patterns of ostracode assemblage from the Pliocene through to the Modern, and the ostracode species identified represent freshwater ponds, lakes, rivers, wetlands, springs, and shallow saline ponds. Modern analogue analysis showed that the paleo ostracode assemblages represent modern climates with similar mid-continent temperatures but with increased precipitation than that which is currently seen in Meade County. Analysis of these assemblages in solute space show changing aquatic environments, from fresh to more saline waters throughout the Quaternary. These results support the hypothesis proposed by Ibarra et al (2018) that the Pliocene Epoch warm period and Pleistocene glacial stages have similar hydroclimatic conditions, and that these conditions were fresher (more dilute) than supported by modern climatic conditions, and in this study, fresher than the Sangamon interglacial conditions.

More analysis is needed to determine the extent of these climatic differences, and it is my hope that future research with ostracodes in both Meade County and the North American

Southern Great Plains will continue and expand. By increasing our understanding of past climates, we can better prepare our society for future climatic conditions.

56

6. References

Aber, J. S. (1991). The glaciation of northeastern Kansas. Boreas, 20(4), 297-314.

Anderson, D. E., Anderson, D., Goudie, A., & Parker, A. (2013). Global environments through the quaternary: exploring environmental change. Oxford University Press, USA.

Bayne, C. K., Boellstorff, J., & Miller, B. B. (1976). Guidebook: 24th annual meeting, Midwestern Friends of the Pleistocene. Lawrence, KS: Kansas Geological Survey,

University of Kansas. doi:11/09/2010

Change, A. D. C., Blair, T., & Pachauri, R. (2006). Avoiding dangerous climate change.

Cambridge University Press.

Cayan, D. R., Das, T., Pierce, D. W., Barnett, T. P., Tyree, M., & Gershunov, A. (2010).

Future dryness in the southwest US and the hydrology of the early 21st century drought.

Proceedings of the National Academy of Sciences, 107(50), 21271-21276.

Colman, S.M, Jones, G.A., Forester, R.M. and D.S. Foster, 1990. Holocene paleoclimatic evidence and sedimentation rates from a core in southwestern Lake Michigan, J.

Paleolimnology, 4, 269-284.

Curry, B. B., Delorme, L. D., Smith, A. J., Palmer, D. F., & Stiff, B. J. (2012). The biogeography and physicochemical characteristics of aquatic habitats of freshwater ostracods in

Canada and the United States. In Developments in Quaternary Sciences (Vol. 17, pp. 85-115).

Elsevier.

57

Davis, J. C. (2002). Statistics and data analysis in geology (3rd ed.). John Wiley & Sons.

De Schepper S, Groeneveld J, Naafs BDA, Van Renterghem C, Hennissen J, Head M J,

Louwye, S. and Fabian, K., (2013) Northern Hemisphere Glaciation during the Globally Warm

Early Late Pliocene. PLoS ONE 8(12): e81508.

Delorme, L.D., 1970. Freshwater ostracodes of Canada, part II, subfamily Cypridopsinae and Herpetocypridinae, and family Cyclocyprididae, Canadian Journal of Zoology, 48: 253-266.

Dort, W., Jr., 2006, Multiple pre-Illinoian tills and associated sediments and paleosols, northeastern Kansas and central Missouri, AMQUA post-meeting field trip 3; in, Guidebook of the 18th biennial meeting of the American Quaternary Association, R. Mandel, ed.: Kansas

Geological Survey, Technical Series 21, p. 3-1–3-18.

Dort, W. (2007). Maximum Southwestern Extent of" Kansan" Ice Sheet and Newly

Discovered Older Till: North-central and South-central Sections, the Geological Society of

America, April 14, 2007. University Press of Kansas.

Eugster, H. & Jones, B.F., 1979. Behavior of major solutes during closed-basin brine evolution. American Journal of Science 279: 609–631.

Forester, R.M., 1988. Nonmarine calcareous microfossil sample preparation and data acquisition procedures. US Geological Survey Technical Procedure HP-78 R1: 1-9

Forester, R.M., 1991. Pliocene-climate history of the western United States derived from lacustrine ostracodes, Quaternary Science Reviews, 10 (1-2): 133-146.

58

Forester, R.M., Smith, A.J., Palmer, D.F., Curry, B.B., 2005. NANODe: North American

Nonmarine Ostracode Database, version 1, Kent State University, Kent Ohio, www.kent.edu/NANODe .

Forester, R.M., Carter, C., Quade, J., Smith, A.J., 2017. Aquifer and surface-water ostracodes in Quaternary paleowetland deposits of southern Nevada, USA, Hydrobiologia, 786:

41. doi:10.1007/s10750-016-2966-5.

Ferguson Jr, E. (1967). New ostracods from the playa lakes of eastern New Mexico and western Texas. Transactions of the American Microscopical Society, 244-250.

Frye, J. C., & Hibbard, C. W. (1941). Pliocene and Pleistocene stratigraphy and paleontology of the Meade Basin, southwestern Kansas. Kansas Geological Survey Bulletin, 38,

398-403. http://www.kgs.ku.edu/Publications/Bulletins/38_13/index.html

Furtos, N.C., 1933. The Ostracoda of Ohio, Ohio Biological Survey Bulletin 29, Ohio

State University Bulletin, 37: 413-524.

Gutentag, E. D., & Benson, R. H. (1962). Neogene (Plio-Pleistocene) fresh-water ostracodes from the central high plains. State Geol. Survey of Kansas. http://www.kgs.ku.edu/Publications/Bulletins/157_4/02_intro.html

Hanson, P.R., Arbogast, Johnson, W.C., Joeckel, and Young, A.R., 2010, Megadroughts and late Holocene dune activation at the eastern margin of the Great Plains, north-central Kansas,

USA, Aeolian Research, 1 (2010) 101–11.

Hibbard, C. W. (1949). Pleistocene stratigraphy and paleontology of Meade County,

Kansas.

59

Hibbard, C.W. (1950’s), slide collections made in the 1950’s from Meade County,

Kansas, housed in the Kent State University Paleolimnology Lab.

Hibbard, C. W. (1958). New stratigraphic names for early Pleistocene deposits in southwestern Kansas. American Journal of Science, 256(1), 54-59.

Hibbard, C. W., & Taylor, D. W. (1960). Two late Pleistocene faunas from southwestern

Kansas. Contributions from the Museum of Paleontology, The University of Michigan, XVI (1):

1-223.

Hoff, C.C., 1942. The Ostracoda of Illinois, their biology and ; Illinois biological monographs ; v. 19, no. 1-2, 196pp,

Humphreys, W. F. "Hydrogeology and groundwater ecology: Does each inform the other?" Hydrogeology Journal 17.1 (2009): 5-21

Hurteau, M. D., Bradford, J. B., Fulé, P. Z., Taylor, A. H., & Martin, K. L. (2014).

Climate change, fire management, and ecological services in the southwestern US. Forest

Ecology and Management, 327, 280-289.

Ibarra, D. E., Oster, J. L., Winnick, M. J., Caves Rugenstein, J. K., Byrne, M. P., &

Chamberlain, C. P. (2018). Warm and cold wet states in the western United States during the

Pliocene–Pleistocene. Geology, 46(4), 355-358.

Janis, C. M., Damuth, J., & Theodor, J. M. (2002). The origins and evolution of the North

American grassland biome: the story from the hoofed mammals. Paleogeography,

Paleoclimatology, Paleoecology, 177(1-2), 183-198.

60

Johnson, W. C., & Willey, K. L. (2000). Isotopic and rock magnetic expression of environmental change at the Pleistocene–Holocene transition in the central Great Plains.

Quaternary International, 67(1), 89-106.

Karl, T. R., Melillo, J. M., Peterson, T. C., & Hassol, S. J. (Eds.). (2009). Global climate change impacts in the United States. Cambridge University Press.

Kolivras, K. N., & Comrie, N. C. (2004). Climate and infectious disease in the southwestern United States. Progress in Physical Geography, 28(3), 387-398.

Kovach Computing Services (2002). Multi-Variate Statistical Package Version 3.1

[Computer software]. Pentraeth, Wales, U.K.

Layzell, A. L., Mandel, R. D., Ludvigson, G. A., Rittenour, T. M., & Smith, J. J. (2015).

Forces driving late Pleistocene (ca. 77–12 ka) landscape evolution in the Cimarron River valley, southwestern Kansas. Quaternary Research, 84(1), 106-117.

Layzell, A. L., Sawin, R. S., Mandel, R. D., Ludvigson, G. A., Franseen, E. K., West, R.

R., & Watney, W. L. (2017). Quaternary stratigraphy and stratigraphic nomenclature revisions in

Kansas. Midcontinent Geoscience, 1-8.

Lisiecki, L. & Raymo, M., 2005. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records, Paleoceanography, 20: PA1003, doi:10.1029/2004PA001071, 2005.

Lyle, S.A., 2009. Glaciers in Kansas, Kansas Geological Survey, Public Information

Circular, 28, 6pp. http://www.kgs.ku.edu/Publications/PIC/PIC28.pdf

61

Lukens, W. E., Fox, D. L., Snell, K. E., Wiest, L. A., Layzell, A. L., Uno, K. T., &

Peláez-Campomanes, P. (2019). Pliocene Paleoenvironments in the Meade Basin, Southwest

Kansas, USA. Journal of Sedimentary Research, 89(5), 416-439.

Manly, B. F., & Alberto, J. A. N. (2016). Multivariate statistical methods: a primer.

Chapman and Hall/CRC.

Martin, R. A., Honey, J. G., Peláez-Campomanes, P., Goodwin, H. T., Baskin, J. A., &

Zakrzewski, R. J. (2002). Blancan lagomorphs and rodents of the Deer Park assemblages, Meade

County, Kansas. Journal of Paleontology, 76(6), 1072-1090.

Mesquita-Joanes, F., Smith, A. J., & Viehberg, F. A. (2012). The ecology of Ostracoda across levels of biological organization from individual to ecosystem: a review of recent developments and future potential. In Developments in Quaternary Sciences (Vol. 17, pp. 15-

35). Elsevier.

Miller, B. B. (1976). The late Cenozoic molluscan succession in the Meade County,

Kansas area. In Guidebook 24th Annual Meeting Midwestern Friends of the Pleistocene.

Guidebook Series, Vol. 1, 73-85.

Mullin, M., (2020). The effects of drinking water service fragmentation on drought- related water security, Science, 368, (6488), 274-277.

Overpeck, J. T., Webb, T. I. I. I., & Prentice, I. C. (1985). Quantitative interpretation of fossil pollen spectra: dissimilarity coefficients and the method of modern analogs. Quaternary

Research, 23(1), 87-108.

62

Pieri, V., Martens, K., Stoch, F., & Rossetti, G. (2009). Distribution and ecology of non- marine ostracods (Crustacea, Ostracoda) from Friuli Venezia Giulia (NE Italy). Journal of limnology, 68(1), 1-15.

Pound, Matthew, Tindall, Julia, Pickering, Steven, Haywood, Alan, Dowsett, Harry and

Salzmann, Ulrich (2014) Late Pliocene lakes and soils: a global data set for the analysis of climate feedbacks in a warmer world. Climate of the Past, 10 (1). pp. 167-180. ISSN 1814-9332.

Railsback, L.B., Gibbard, P.I., , Head, M.J., Voarintsoa, N.R.G., and Toucanne, S., 2015.

An optimized scheme of lettered marine isotope substages for the last 1.0 million years, and the climatostratigraphic nature of isotope stages and substages, Quaternary Science Reviews,

111:94-106.

Rempel, J. A (2018). Geological History of Meade County According to Claude Hibbard. http://www.oldmeadecounty.com/hibbard.htm

Shukla, S. P., Chandler, M. A., Jonas, J., Sohl, L. E., Mankoff, K.,and Dowsett, H., 2009.

Impact of a permanent El Niño (El Padre) and Indian Ocean Dipole in warm Pliocene climates,

Paleoceanography, 24, PA2221, doi:10.1029/2008PA001682, 2009.

Smith, A.J., 1993. Lacustrine ostracodes as hydrochemical indicators in lakes of the north-central United States, J. Paleolimnology, 8: 121-134.

Smith, A. J., & Delorme, L. D. (2009). Ostracoda in: Thorp JH, Covich AP. Ecology and classification of North American freshwater invertebrates., Academic Press, New York.

63

Smith, A.J. and Palmer, D.F., 2012. "The versatility of Quaternary ostracods as paleoclimate proxies: comparative testing of geochemical, ecological and biogeographical approaches." Developments in Quaternary Sciences. Vol. 17. Elsevier, 2012. 183-203.

Smith, A.J., Horne, D.J., Martens, K., & Schön, I., 2015. Class Ostracoda. In: Thorp, J.,

Rogers, D.C. (Eds.), Ecology and General Biology: Thorp and Covich's Freshwater

Invertebrates, Academic Press, 757–780.

Smith, A.J. and Horne, D.J., 2016. Ostracoda of the Nearctic, In: Thorp, J., Rogers, D.C.

(Eds.), Ecology and General Biology: Thorp and Covich’s Freshwater Invertebrates, Volume 2,

Chapter 16, Academic Press, pp. 477-513.

Smith, G. R., & Friedland, N. E. (Eds.). (1975). Studies on Cenozoic Paleontology and

Stratigraphy: In Honor of Claude W. Hibbard. Museum of Paleontology, University of Michigan.

Smith, H. T. U. l940, Geologic studies in southwestern Kansas: Kansas State Geol.

Survey Bull, 34, 2l2.

Thompson, R.S., 1991. Pliocene environments and climates in the western United States,

Quaternary Science Reviews, 10 (2-3): 115-132.

Weissleader, L. S., Gilinsky, N. L., Ross, R. M., & Cronin, T. M. (1989). Biogeography of marine podocopid ostracodes in Micronesia. Journal of Biogeography, 103-114.

Williams, J.W., Grimm, E.G., Blois, J., Charles, D.F., Davis, E., Goring, S.J., Graham,

R., Smith, A.J., Anderson, M., Arroyo-Cabrales, J., Ashworth, A.C., Betancourt, J.L., Bills,

B.W., Booth, R.K., Buckland, P., Curry, B., Giesecke, T., Hausmann, S., Jackson, S.T., Latorre,

64

C., Nichols, J., Purdum, T., Roth, R.E., Stryker, M., Takahara, H., 2018. The Neotoma

Paleoecology Database: A multi-proxy, international community-curated data resource.

Quaternary Research 89, 156-177.

Winnick, M. J., Welker, J. M. and Chamberlain, C. P. 2013. Stable isotopic evidence of

El Niño-like atmospheric circulation in the Pliocene western United States, Climates of the Past,

9: 903–912, www.clim-past.net/9/903/2013/ doi:10.5194/cp-9-903-2013 .

Wycech, J. B., Gill, E., Rajagopalan, B.,Marchitto, T. M., Jr, & Molnar, P. H.(2020).

Multiproxy Reduced‐Dimension Reconstruction of Pliocene Equatorial Pacific Sea Surface

Temperatures. Paleoceanography andPaleoclimatology,35, e2019PA003685.https://doi.org/10.1029/2019PA003685

Zeller, D. E., 1968, The stratigraphic succession in Kansas: Kansas Geological Survey,

Bulletin 189, 81 p., 1 plate.

65

7. Appendices A. Ostracode counts from the Tomin 2019 Collection

Cudahy Cudahy Borcher's Borcher's Rexroad Location of Sample Cuday Ash Ash Ash Badlands Badlands 3 Top of Stratigraphy of Locations Pleistocene Pleistocene Pleistocene Pleistocene Pleistocene Pliocene Ostracode Species KS01 KS02 KS03 KS04 KS05 KS06 Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 0 0 0 0 Cypridopsis vidua (O.F. Muller, 1776) 35 52 0 0 0 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol et al., 1988 0 0 0 0 0 0 Candona crogmaniana Turner 1894 0 0 0 0 0 0 Candona fluviatilis Hoff 1942 0 0 0 0 0 0 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 1995 0 0 0 0 0 0 Candona renoensis (Gutentag and Benson, 1962) 0 0 0 0 0 0 Eucypris meadensis (Gutentag and Benson, 1962) 0 0 0 0 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 1808) 0 0 0 0 0 0 Limnocythere staplini, (Gutentag and Benson, 1962) 0 0 0 0 0 0 Potamocypris smaragdina Vavra 1891 0 0 0 0 0 0 Scottia pseudobrowniana 22 16 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0

66

Cudahy Cudahy Borcher's Borcher's Rexroad Location of Sample Cuday Ash Ash Ash Badlands Badlands 3 Top of Stratigraphy of Locations Pleistocene Pleistocene Pleistocene Pleistocene Pleistocene Pliocene Ostracode Species KS01 KS02 KS03 KS04 KS05 KS06 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata (inopinata) 0 0 0 0 0 0 Paracandona euplectella (Robertson, 1889) 230 127 0 0 0 0 Candona distincta 0 0 0 0 0 0 Candona punctata 0 0 0 0 0 0 Candona acuta 0 0 0 0 0 0 Pseudocandona stagnalis 22 37 0 0 0 0 Candona albicans 0 0 0 0 0 0 Sarscypridopsis aculeata 0 0 0 0 0 0 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 9 17 0 0 0 0 Herpetocypris brevicaudata 0 0 0 0 0 0 Bradleystrandesia reticulatus (Zaddach, 1854) 0 0 0 0 0 0 Darwinula stevensoni 0 0 0 0 0 0 Physocypria pustulosa (Sharpe, 1897) 0 0 0 0 0 0 Cypria exsculpta (Fischer, 1855) 6 5 0 0 0 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 0 0 0 Candona Juveniles 18 21 0 0 0 0 Ostracode Shell Fragments 30 24 0 0 0 0

67

Keefe Keefe Keefe Fox Deer Windmill Location of Sample Rexroad 3 Canyon Canyon Canyon Canyon Park Playa Top of Mid- Mid- Mid- Mid- Top of Stratigraphy of Locations Pliocene Pliocene Pliocene Pliocene Pliocene Pliocene Modern Ostracode Species KS07 KS08 KS09 KS10 KS11 KS12 TKSMR Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 0 0 0 0 0 Cypridopsis vidua (O.F. Muller, 1776) 0 0 0 0 0 0 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol et al., 1988 0 0 0 0 1 0 0 Candona crogmaniana Turner 1894 0 0 0 0 6 0 0 Candona fluviatilis Hoff 1942 0 40 0 14 0 0 45 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 1995 0 0 0 0 0 0 10 Candona renoensis (Gutentag and Benson, 1962) 0 0 0 0 0 0 0 Eucypris meadensis (Gutentag and Benson, 1962) 0 0 0 0 0 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 1808) 0 0 0 0 2 0 0 Limnocythere staplini, (Gutentag and Benson, 1962) 0 0 0 0 0 0 0 Potamocypris smaragdina Vavra 1891 0 0 0 0 0 0 20 Scottia pseudobrowniana 0 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 15 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 12 Candona truncata (inopinata) 0 0 0 0 5 0 0

68

Cudahy Cudahy Borcher's Borcher's Rexroad Location of Sample Cuday Ash Ash Ash Badlands Badlands 3 Top of Stratigraphy of Locations Pleistocene Pleistocene Pleistocene Pleistocene Pleistocene Pliocene Ostracode Species KS01 KS02 KS03 KS04 KS05 KS06 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 0 Candona distincta 0 0 0 0 0 0 0 Candona punctata 0 0 0 0 0 0 0 Candona acuta 0 0 0 0 0 0 0 Pseudocandona stagnalis 0 4 0 13 99 0 0 Candona albicans 0 0 0 0 0 0 0 Sarscypridopsis aculeata 0 0 0 0 0 0 0 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 0 0 15 0 0 Herpetocypris brevicaudata 0 0 0 0 0 0 0 Bradleystrandesia reticulatus (Zaddach, 1854) 0 1 0 0 3 0 0 Darwinula stevensoni 0 0 0 0 6 0 0 Physocypria pustulosa (Sharpe, 1897) 0 0 0 0 0 0 0 Cypria exsculpta (Fischer, 1855) 0 0 0 0 0 0 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 0 5 0 0 Candona prolata 0 0 0 0 0 0 0 Candona Species A 0 0 0 0 0 0 0 Candona Juveniles 0 1 0 11 342 1 230 Ostracode Shell Fragments 0 17 0 1 13 2 13

69

B. Ostracode counts from the Gutentag and Benson (1962) collection

Ostracode Species 719561 719562 219571 612571 613571 617571 Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 243 8 0 0 0 0 Cypridopsis vidua (O.F. Muller, 1776) 0 0 0 20 3 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol 0 0 0 0 0 0 et al., 1988 Candona crogmaniana Turner 1894 0 0 0 0 0 0 Candona fluviatilis Hoff 1942 0 0 0 0 2 0 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 146 15 3 37 1 44 1995 Candona renoensis G&B, 1962 0 0 0 0 0 0 Eucypris meadensis G&B, 1962 8 0 0 1 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 0 0 0 0 0 1 1808) Limnocythere staplini, G&B, 1962 5 0 0 3 0 91 Potamocypris smaragdina Vavra 1891 0 0 0 0 0 0 Scottia pseudobrowniana 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata 0 0 0 0 0 0 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 Candona distincta 0 0 0 0 0 0 Candona punctata 0 0 0 0 0 0 Candona acuta 0 0 0 0 0 0 Candona stagnalis 0 0 1 0 0 0 Candona albicans 0 0 0 0 0 0 Sarscypridopsis aculeata 0 0 0 0 0 0

70

Ostracode Species 719561 719562 219571 612571 613571 617571 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 10 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 0 0 0 0 Herpetocypris brevicaudata 0 0 0 0 0 0 Bradleystrandesia reticulatus (Zaddach, 1854) 0 0 0 0 0 0 Darwinula stevensoni 0 0 0 0 0 0 Physocypria pustulosa (Sharpe, 1897) 0 0 0 7 0 0 Cypria exsculpta (Fischer, 1855) 0 0 0 0 0 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 2 0 0 Candona prolata 0 0 0 0 0 0 Candona Species A 0 0 0 0 0 0 Candona Juveniles 91 8 179 254 143 197 Ostracode Shell Fragments 0 30 34 22 15 50

Ostracode Species 627571 627572, 627573 628571 628572 630571, 10B 9C Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 0 0 60 63 Cypridopsis vidua (O.F. Muller, 1776) 0 0 0 0 0 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol 0 0 0 0 0 0 et al., 1988 Candona crogmaniana Turner 1894 0 0 0 0 0 0 Candona fluviatilis Hoff 1942 0 0 0 0 0 0 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 2 0 2 51 0 23 1995 Candona renoensis G&B, 1962 0 0 0 0 0 0 Eucypris meadensis G&B, 1962 0 0 0 0 0 1

71

Ostracode Species 627571 627572, 627573 628571 628572 630571, 10B 9C Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 0 0 0 0 0 0 1808) Limnocythere staplini, G&B, 1962 0 0 7 16 0 3 Potamocypris smaragdina Vavra 1891 0 0 0 0 0 0 Scottia pseudobrowniana 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata 0 0 0 0 0 0 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 Candona distincta 0 0 0 0 0 0 Candona punctata 0 0 0 0 0 0 Candona acuta 0 0 0 0 0 0 Candona stagnalis 0 0 0 0 0 0 Candona albicans 0 0 0 0 0 0 Sarscypridopsis aculeata 0 0 0 0 0 0 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 0 0 2 0 Herpetocypris brevicaudata 0 0 0 0 0 0 Bradleystrandesia reticulatus (Zaddach, 1854) 0 0 0 0 0 0 Darwinula stevensoni 0 0 0 0 0 0 Physocypria pustulosa (Sharpe, 1897) 0 0 0 0 0 0 Cypria exsculpta (Fischer, 1855) 0 0 0 0 0 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 1 1 0 0 0 Candona prolata 0 0 0 0 0 0 Candona Species A 0 0 0 0 0 0 Candona Juveniles 20 3 9 0 2 11

72

Ostracode Species 627571 627572, 627573 628571 628572 630571, 10B 9C Ostracode Shell Fragments 0 0 0 0 0 0

630572, 79572 814571 920571 920572 920573 Ostracode Species 10C Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 260 0 0 1 0 0 Cypridopsis vidua (O.F. Muller, 1776) 0 0 0 0 0 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol 0 0 0 0 0 0 et al., 1988 Candona crogmaniana Turner 1894 0 0 0 0 0 0 Candona fluviatilis Hoff 1942 0 0 0 0 0 0 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 7 0 0 0 56 13 1995 Candona renoensis G&B, 1962 0 0 0 0 0 0 Eucypris meadensis G&B, 1962 0 0 0 0 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 0 0 0 0 0 0 1808) Limnocythere staplini, G&B, 1962 0 0 0 0 0 0 Potamocypris smaragdina Vavra 1891 0 0 0 0 0 0 Scottia pseudobrowniana 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata 0 0 0 0 0 0 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 Candona distincta 0 0 0 0 0 0 Candona punctata 0 0 0 0 0 0 Candona acuta 0 0 0 0 0 0 Candona stagnalis 0 0 0 0 0 0

73

630572, 79572 814571 920571 920572 920573 Ostracode Species 10C Candona albicans 0 0 0 0 0 0 Sarscypridopsis aculeata 0 0 0 0 0 0 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 0 0 0 0 Herpetocypris brevicaudata 0 0 0 0 0 0 Bradleystrandesia reticulatus (Zaddach, 1854) 0 0 0 0 0 0 Darwinula stevensoni 0 0 0 0 0 0 Physocypria pustulosa (Sharpe, 1897) 0 0 0 0 0 0 Cypria exsculpta (Fischer, 1855) 0 0 0 0 0 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 0 0 0 Candona prolata 0 0 0 0 0 0 Candona Species A 0 0 0 0 0 0 Candona Juveniles 44 68 0 0 0 0 Ostracode Shell Fragments 16 15 0 0 11 0

927576 1015571 1028571 Skim Big Concrete Sand Springs Ditch, buried Bola Buried Artesian (BM) Artesian Ostracode Species Boil Boil Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 0 0 0 0 Cypridopsis vidua (O.F. Muller, 1776) 0 0 0 0 1 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol 0 0 0 0 4 0 et al., 1988 Candona crogmaniana Turner 1894 0 0 0 0 0 0

74

927576 1015571 1028571 Skim Big Concrete Sand Springs Ditch, buried Bola Buried Artesian (BM) Artesian Ostracode Species Boil Boil Candona fluviatilis Hoff 1942 0 0 0 0 0 0 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 0 0 0 0 0 0 1995 Candona renoensis G&B, 1962 0 0 0 0 0 0 Eucypris meadensis G&B, 1962 0 0 0 30 0 26 Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 0 0 0 2 180 3 1808) Limnocythere staplini, G&B, 1962 0 0 0 0 0 0 Potamocypris smaragdina Vavra 1891 0 0 0 0 0 0 Scottia pseudobrowniana 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata 0 0 0 0 0 0 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 Candona distincta 0 0 0 0 0 0 Candona punctata 0 0 0 0 0 0 Candona acuta 0 0 0 0 0 0 Candona stagnalis 0 0 0 0 0 0 Candona albicans 0 0 0 0 0 0 Sarscypridopsis aculeata 0 0 0 0 0 0 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 0 0 0 0 Herpetocypris brevicaudata 0 0 0 18 5 17 Bradleystrandesia reticulatus (Zaddach, 1854) 0 0 0 5 0 5

75

927576 1015571 1028571 Skim Big Concrete Sand Springs Ditch, buried Bola Buried Artesian (BM) Artesian Ostracode Species Boil Boil Darwinula stevensoni 0 0 0 0 47 0 Physocypria pustulosa (Sharpe, 1897) 0 0 0 0 0 0 Cypria exsculpta (Fischer, 1855) 0 0 0 0 55 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 0 26 0 Candona prolata 0 0 0 0 0 0 Candona Species A 0 0 0 0 0 0 Candona Juveniles 0 0 0 4 1 0 Ostracode Shell Fragments 0 0 0 1 10 2

Ostracode Species Artesian Boil Rexroad? SE 12-35-31 Spirocypris tuberculata (Sharpe, 1908) 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 Cypridopsis vidua (O.F. Muller, 1776) 0 0 Fabaeformiscandona caudata (Kaufmann, 1900) Danielopol 0 0 et al., 1988 Candona crogmaniana Turner 1894 0 0 Candona fluviatilis Hoff 1942 0 1 Fabaeformiscandona rawsoni (Tressler, 1957), Griffiths, 0 0 1995 Candona renoensis G&B, 1962 0 0 Eucypris meadensis G&B, 1962 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris gibba (Ramdohr, 1 1 1808) Limnocythere staplini, G&B, 1962 0 0 Potamocypris smaragdina Vavra 1891 0 0

76

Ostracode Species Artesian Boil Rexroad? SE 12-35-31 Scottia pseudobrowniana 0 0 Dolerocyprus sinensis 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 Candona truncata 0 0 Paracandona euplectella (Robertson, 1889) 0 0 Candona distincta 0 0 Candona punctata 0 0 Candona acuta 0 0 Candona stagnalis 0 0 Candona albicans 0 0 Sarscypridopsis aculeata 0 0 Candona neglecta (Sars, 1887) 0 0 Physocypria globula (Furtos, 1933) 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 Herpetocypris brevicaudata 0 2 Bradleystrandesia reticulatus (Zaddach, 1854) 0 0 Darwinula stevensoni 0 0 Physocypria pustulosa (Sharpe, 1897) 0 0 Cypria exsculpta (Fischer, 1855) 0 0 Limnocythere ceriotuberosa (Delorme, 1967) 1 0 Candona prolata 0 0 Candona Species A 0 15 Candona Juveniles 0 6 Ostracode Shell Fragments 0 25

77

C. Ostracode counts from the Hibbard (1950's) collection Ostracode Species UM-K1-61 UM-K4-59 UM-K1-57 35-29- UM-K3-61 UM-K1 (34- 5bab 29-33ccc) (K561) Spirocypris tuberculata (Sharpe, 1908) 0 0 8 0 9 0 Cyprideis salebrosa (Van den Bold, 1963) 38 19 0 0 0 3 Cypridopsis vidua (O.F. Muller, 1776) 421 3 8 470 501 154 Fabaeformiscandona caudata (Kaufmann, 0 0 0 0 1 0 1900) Danielopol et al., 1988 Candona crogmaniana Turner 1894 0 0 6 0 0 0 Candona fluviatilis Hoff 1942 0 0 0 0 3 0 Fabaeformiscandona rawsoni (Tressler, 549 55 7 8 13 216 1957), Griffiths, 1995 Candona renoensis G&B, 1962 0 0 0 0 0 0 Eucypris meadensis G&B, 1962 0 0 0 0 1 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris 2 2 2 0 1 1 gibba (Ramdohr, 1808) Limnocythere staplini, G&B, 1962 0 0 0 0 0 0 Potamocypris smaragdina (Vavra 1891) 0 0 0 0 0 0 Scottia pseudobrowniana 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata 0 0 0 3 11 0 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 Candona distincta 0 1 0 19 9 0 Candona punctata 24 0 6 5 0 0 Candona acuta 0 0 0 9 2 0 Candona stagnalis 0 0 0 0 7 0 Candona albicans 2 0 0 23 64 0

78

Ostracode Species UM-K1-61 UM-K4-59 UM-K1-57 35-29- UM-K3-61 UM-K1 (34- 5bab 29-33ccc) (K561) Sarscypridopsis aculeata 0 0 0 9 0 0 Candona neglecta (Sars, 1887) 0 0 0 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 0 102 93 0 Herpetocypris brevicaudata 0 0 0 0 2 0 Bradleystrandesia reticulatus (Zaddach, 1854) 6 5 52 2 176 0 Darwinula stevensoni 0 0 0 0 0 0 Cypria exsculpta (Fischer, 1855) 0 0 0 0 3 0 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 0 0 0 Candona prolata 0 0 0 0 0 0 Candona Species A 0 0 0 0 0 0 Candona Juveniles 84 0 14 28 133 0 Ostracode Shell Fragments 49 28 15 18 15 3

UM-K2-59 NWSE 33- UM-K1-59 Type Holocene UM-K1-60 Ostracode Species 29-16db Section Spirocypris tuberculata (Sharpe, 1908) 0 0 0 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 0 0 0 0 Cypridopsis vidua (O.F. Muller, 1776) 68 150 13 18 0 1 Fabaeformiscandona caudata (Kaufmann, 0 0 0 0 0 0 1900) Danielopol et al., 1988 Candona crogmaniana Turner 1894 0 40 0 0 0 0 Candona fluviatilis Hoff 1942 0 0 0 0 0 0 Fabaeformiscandona rawsoni (Tressler, 0 4 0 0 0 0 1957), Griffiths, 1995 Candona renoensis G&B, 1962 8 0 0 0 0 0

79

UM-K2-59 NWSE 33- UM-K1-59 Type Holocene UM-K1-60 Ostracode Species 29-16db Section Eucypris meadensis G&B, 1962 0 0 0 0 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris 0 5 0 0 3 0 gibba (Ramdohr, 1808) Limnocythere staplini, G&B, 1962 0 0 0 0 0 0 Potamocypris smaragdina (Vavra 1891) 0 0 0 0 0 0 Scottia pseudobrowniana 0 0 0 0 0 0 Dolerocyprus sinensis 0 0 0 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 0 0 0 Candona truncata 0 0 0 0 0 0 Paracandona euplectella (Robertson, 1889) 0 0 0 0 0 0 Candona distincta 0 2 0 2 0 0 Candona punctata 0 0 0 0 0 0 Candona acuta 0 0 0 0 21 0 Candona stagnalis 0 69 0 1 0 2 Candona albicans 1 0 0 0 0 1 Sarscypridopsis aculeata 0 0 0 0 0 0 Candona neglecta (Sars, 1887) 0 10 0 0 0 0 Physocypria globula (Furtos, 1933) 0 23 0 0 0 0 Cyclocypris ovum (Jurine, 1820) 4 303 37 0 0 2 Herpetocypris brevicaudata 0 2 0 0 30 0 Bradleystrandesia reticulatus (Zaddach, 1854) 2 48 0 0 98 0 Darwinula stevensoni 0 8 0 0 0 0 Cypria exsculpta (Fischer, 1855) 0 19 0 0 0 5 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 0 0 12 Candona prolata 0 0 0 1 0 0 Candona Species A 0 1 0 0 0 0 Candona Juveniles 32 221 18 5 0 7

80

UM-K2-59 NWSE 33- UM-K1-59 Type Holocene UM-K1-60 Ostracode Species 29-16db Section Ostracode Shell Fragments 19 33 0 13 0 3

Ostracode Species UM-K1-55 (34-29-32d) Craqin Quarry Cuday Fauna-Ash Mine Spirocypris tuberculata (Sharpe, 1908) 0 0 0 Cyprideis salebrosa (Van den Bold, 1963) 0 0 0 Cypridopsis vidua (O.F. Muller, 1776) 0 7 0 Fabaeformiscandona caudata (Kaufmann, 9 0 0 1900) Danielopol et al., 1988 Candona crogmaniana Turner 1894 9 0 0 Candona fluviatilis Hoff 1942 0 0 0 Fabaeformiscandona rawsoni (Tressler, 11 0 15 1957), Griffiths, 1995 Candona renoensis G&B, 1962 0 0 0 Eucypris meadensis G&B, 1962 0 0 0 Ilyocypris bradyi (Sars, 1890), Ilyocypris 0 0 0 gibba (Ramdohr, 1808) Limnocythere staplini, G&B, 1962 0 0 0 Potamocypris smaragdina (Vavra 1891) 0 0 0 Scottia pseudobrowniana 0 0 0 Dolerocyprus sinensis 0 0 0 Pseudoilyocypris tuberculatum (Pelocypris) 0 0 0 Candona truncata 0 0 0 Paracandona euplectella (Robertson, 1889) 0 0 0 Candona distincta 2 0 0 Candona punctata 0 0 0 Candona acuta 29 0 0 Candona stagnalis 2 3 0

81

Ostracode Species UM-K1-55 (34-29-32d) Craqin Quarry Cuday Fauna-Ash Mine Candona albicans 0 4 0 Sarscypridopsis aculeata 0 0 0 Candona neglecta (Sars, 1887) 0 0 0 Physocypria globula (Furtos, 1933) 0 0 0 Cyclocypris ovum (Jurine, 1820) 0 0 2 Herpetocypris brevicaudata 0 0 0 Bradleystrandesia reticulatus (Zaddach, 1854) 4 0 0 Darwinula stevensoni 0 0 0 Cypria exsculpta (Fischer, 1855) 0 0 6 Limnocythere ceriotuberosa (Delorme, 1967) 0 0 0 Candona prolata 0 0 0 Candona Species A 0 0 0 Candona Juveniles 0 0 2 Ostracode Shell Fragments 0 8 8

82

D. Shannon's diversity index

Shannon's method Log base 10 Sample Index Evenness Num.Spec. TKSMR 0.626 0.896 5 GSSBABR 0.449 0.746 4 GBSB(BM)R 0.543 0.643 7 GCDBABR 0.479 0.796 4 G66571R 0.521 0.669 6 H526585R 0.507 0.843 4 HHoloH 0.414 0.688 4 HUM-K2-59PE 0.294 0.421 5 HUM-K1-60PE 0.594 0.764 6 HCQPE 0.449 0.942 3 G630571, 10BPE 0.331 0.549 4 G628572PE 0.062 0.206 2 G630572, 10CPE 0.053 0.175 2 G719561PE 0.349 0.58 4 HUM-K4-59PE 0.452 0.581 6 HUM-K1-57PE 0.614 0.726 7 H35-29-5babPE 0.391 0.433 8 HUM-K3-61PE 0.591 0.502 15 HUM-K1-55PE 0.688 0.815 7 TKS01PE 0.444 0.571 6 TKS02PE 0.601 0.773 6 HCFAMPE 0.366 0.766 3 HUM-K1-59TPI 0.249 0.827 2 HTypeSecTPI 0.218 0.456 3 HNWSE 33-29- 0.718 0.665 12 16db TPI G12-35-31TPI 0.452 0.946 3 TKS08MPI 0.176 0.368 3 TKS10MPI 0.301 0.999 2 TKS11MPI 0.507 0.532 9

83

E. Cluster Analysis

Farthest Squared Chord Constrained neighbor clustering strategy Distance matrix TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM- HUM- K2- K1- 59PE 60PE TKSMR 0 GSSBABR 200.01 0 GBSB(BM)R 200 156.966 0 GCDBABR 199.99 0.344 149.036 0 G66571R 200.01 62.221 179.361 64.209 0 H526585R 200 114.009 82.51 109.441 120.503 0 HHoloH 200 95.402 167.747 91.611 104.155 85.178 0 HUM-K2- 200 190.649 189.921 190.27 181.477 186.762 175.07 0 59PE HUM-K1- 200.01 200.02 117.585 200 170.496 131.136 200.01 144.732 0 60PE HCQPE 200 200.01 192.126 199.99 200.01 200 200 60.282 120.909 G630571, 168.346 184.447 200 184.945 179.717 200 200 200 200.01 10BPE G628572PE 200 200.01 200 199.99 200.01 200 200 192.109 189.408 G630572, 189.866 200.01 200 199.99 200.01 200 200 200 200.01 10CPE G719561PE 162.267 179.172 200 179.846 177.044 200 200 200 200.01 G920572PE 137.39 200.01 200 199.99 200.01 200 200 200 200.01 G920573PE 137.39 200.01 200 199.99 200.01 200 200 200 200.01 HUM-K4- 149.635 179.539 174.842 177.373 171.061 170.078 156.757 158.459 192.173 59PE

84

Farthest Squared Constrained clustering neighbor Chord strategy TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM- HUM- K2- K1- 59PE 60PE HUM-K1- 159.554 161.883 178.515 159.279 134.091 145.283 107.782 143.146 190.852 57PE H35-29- 192.982 196.663 190.437 196.514 193.373 195.262 182.209 20.904 132.599 5babPE HUM-K3- 184.663 161.499 179.844 160.22 140.005 147.313 119.713 30.019 127.675 61PE HUM-K1- 174.447 185.176 191.719 184.587 170.632 179.018 111.193 192.367 189.751 55PE TKS01PE 200 200.01 185.026 199.99 194.742 179.908 200 133.186 148.41 TKS02PE 200 200.01 183.286 199.99 194.574 179.267 200 106.738 130.273 HCFAMPE 149.447 200.02 157.52 200 180.237 124.559 200.01 187.059 134.988 HUM-K1- 200 200.01 194.322 199.99 200.01 200 200 69.92 127.994 59TPI HTypeSecTPI 199.99 200 189.681 199.98 200 199.99 199.99 32.393 148.512 HNWSE 33- 195.181 174.349 158.059 172.766 149.923 142.556 149.828 76.769 106.178 29-16db TPI G12-35-31TPI 133.577 138.293 124.767 131.469 115.594 109.551 72.413 178.046 200.01 TKS08MPI 74.751 191.026 200 190.661 182.223 187.294 176.073 195.374 182.421 TKS10MPI 104.342 200.01 200 199.99 200.01 200 200 200 159.076 TKS11MPI 200 186.72 153.709 185.137 179.69 180.453 173.34 181.221 104.48 TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM- HUM- K2- K1- 59PE 60PE

85

Farthest Squared Constrained neighbor Chord clustering strategy HCQPE G630571, G628572PE G630572, G719561PE G920572PE G920573PE HUM- HUM- 10BPE 10CPE K4- K1- 59PE 57PE TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM-K2- 59PE HUM-K1- 60PE HCQPE 0 G630571, 200 0 10BPE G628572PE 200 35.393 0 G630572, 200 18.508 5.851 0 10CPE G719561PE 200 1.926 47.033 27.042 0 G920572PE 200 98.886 200 167.627 79.468 0 G920573PE 200 98.886 200 167.627 79.468 0 0 HUM-K4- 173.429 39.554 106.988 80.654 29.527 39.115 39.115 0 59PE HUM-K1- 168.974 66.552 119.879 98.719 58.829 70.823 70.823 18.691 0 57PE H35-29- 58.098 188.66 185.614 196.376 186.48 177.56 177.56 143.191 139.91 5babPE 2

86

Farthest Squared Constrained neighbor Chord clustering strategy HUM-K3- 56.764 187.042 188.364 196.075 184.451 175.751 175.751 127.452 98.526 61PE HUM-K1- 183.894 158.726 200.01 186.793 150.798 118.352 118.352 118.602 106.84 55PE 7 129.399 200 194.007 200 200 200 200 187.651 185.57 TKS01PE 5 100.675 200 190.703 200 200 200 200 182.999 180.14 TKS02PE 5 HCFAMPE 200.01 118.352 189.408 173.866 102.67 38.492 38.492 70.081 95.687 HUM-K1- 127.889 200 169.079 200 200 200 200 180.84 177.62 59TPI 4 HTypeSecTP 48.863 199.99 199.99 199.99 199.99 199.99 199.99 158.498 159.37 I 1 HNWSE 33- 103.58 192.223 175.872 197.503 190.732 184.628 184.628 153.114 128.95 29-16db TPI 6 G12-35- 200 200 200 200 200 200 200 150.377 109.26 31TPI 4 172.395 200 200 200 200 200 200 192.774 183.53 TKS08MPI 8 TKS10MPI 135.755 200 200 200 200 200 200 200 199.99 122.693 200 188.319 200 200 200 200 189.315 168.83 TKS11MPI 1 HCQPE G630571, G628572PE G630572, G719561PE G920572PE G920573PE HUM- HUM- 10BPE 10CPE K4- K1- TKSMR 59PE 57PE

87

Farthest Squared Constrained clustering neighbor Chord strategy H35-29- HUM-K3-61PE HUM-K1- TKS01PE TKS02PE HCFAMPE 5babPE 55PE TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM-K2- 59PE HUM-K1- 60PE HCQPE G630571, 10BPE G628572PE G630572, 10CPE G719561PE G920572PE G920573PE HUM-K4- 59PE HUM-K1- 57PE H35-29- 0 5babPE HUM-K3- 20.816 0 61PE

88

Farthest Squared Constrained neighbor Chord clustering strategy HUM-K1- 166.295 152.779 0 55PE TKS01PE 130.153 133.591 190.938 0 TKS02PE 101.504 106.827 186.721 5.691 0 HCFAMPE 158.264 155.373 134.074 176.279 170.413 0 HUM-K1- 43.438 67.662 200.01 137.8 109.36 149.264 59TPI HTypeSecTP 30.157 50.757 181.663 127.77 99.561 200 I HNWSE 33- 58.491 50.088 149.526 125.331 93.404 130.857 29-16db TPI G12-35- 192.136 127.86 165.196 200 200 200.01 31TPI TKS08MPI 198.351 170.432 182.294 184.461 177.238 200.01 TKS10MPI 200.01 179.268 175.853 163.837 147.027 200.01 TKS11MPI 169.791 145.656 142.418 145.648 119.446 180.84 H35-29- HUM-K3-61PE HUM-K1- TKS01PE TKS02PE HCFAMPE TKSMR 5babPE 55PE

Farthest Squared Constrained neighbor Chord clustering strategy HUM-K1- HTypeSecTPI HNWSE 33-29-16db G12-35-31TPI TKS08MPI TKS10MPI TKS11MPI 59TPI TPI TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R

89

Farthest Squared Constrained neighbor Chord clustering strategy H526585R HHoloH HUM-K2-59PE HUM-K1-60PE HCQPE G630571, 10BPE G628572PE G630572, 10CPE G719561PE G920572PE G920573PE HUM-K4-59PE HUM-K1-57PE H35-29-5babPE HUM-K3-61PE HUM-K1-55PE TKS01PE TKS02PE HCFAMPE HUM-K1-59TPI 0 HTypeSecTPI 105.577 0 HNWSE 33-29- 36.405 95.212 0 16db TPI G12-35-31TPI 200 199.99 153.625 0 TKS08MPI 200 186.98 172.941 84.647 0 TKS10MPI 200 169.712 155.559 127.993 22.843 0 TKS11MPI 144.092 163.556 78.604 167.583 145.879 84.12 0 HUM-K1- HTypeSecTPI HNWSE 33-29-16db G12-35-31TPI TKS08MPI TKS10MPI TKS11MPI TKSMR 59TPI TPI

90

Farthest neighbor Squared Chord Distance matrix TKSMR GSSBABR GBSB(BM) GCDBAB G66571 H526585 HHolo HUM- HUM- R R R R H K2- K1- 59PE 60PE TKSMR 0 GSSBABR 200.01 0 GBSB(BM)R 200 156.966 0 GCDBABR 199.99 0.344 149.036 0 G66571R 200.01 62.221 179.361 64.209 0 H526585R 200 114.009 82.51 109.441 120.503 0 HHoloH 200 95.402 167.747 91.611 104.155 85.178 0 HUM-K2-59PE 200 190.649 189.921 190.27 181.477 186.762 175.07 0 HUM-K1-60PE 200.01 200.02 117.585 200 170.496 131.136 200.01 144.732 0 HCQPE 200 200.01 192.126 199.99 200.01 200 200 60.282 120.909 G630571, 10BPE 168.346 184.447 200 184.945 179.717 200 200 200 200.01 G628572PE 200 200.01 200 199.99 200.01 200 200 192.109 189.408 G630572, 10CPE 189.866 200.01 200 199.99 200.01 200 200 200 200.01 G719561PE 162.267 179.172 200 179.846 177.044 200 200 200 200.01 G920572PE 137.39 200.01 200 199.99 200.01 200 200 200 200.01 G920573PE 137.39 200.01 200 199.99 200.01 200 200 200 200.01 HUM-K4-59PE 149.635 179.539 174.842 177.373 171.061 170.078 156.757 158.459 192.173 HUM-K1-57PE 159.554 161.883 178.515 159.279 134.091 145.283 107.782 143.146 190.852 H35-29-5babPE 192.982 196.663 190.437 196.514 193.373 195.262 182.209 20.904 132.599 HUM-K3-61PE 184.663 161.499 179.844 160.22 140.005 147.313 119.713 30.019 127.675 HUM-K1-55PE 174.447 185.176 191.719 184.587 170.632 179.018 111.193 192.367 189.751 TKS01PE 200 200.01 185.026 199.99 194.742 179.908 200 133.186 148.41 TKS02PE 200 200.01 183.286 199.99 194.574 179.267 200 106.738 130.273 HCFAMPE 149.447 200.02 157.52 200 180.237 124.559 200.01 187.059 134.988

91

Farthest neighbor Squared Chord HUM-K1-59TPI 200 200.01 194.322 199.99 200.01 200 200 69.92 127.994 HTypeSecTPI 199.99 200 189.681 199.98 200 199.99 199.99 32.393 148.512 HNWSE 33-29- 195.181 174.349 158.059 172.766 149.923 142.556 149.828 76.769 106.178 16db TPI G12-35-31TPI 133.577 138.293 124.767 131.469 115.594 109.551 72.413 178.046 200.01 TKS08MPI 74.751 191.026 200 190.661 182.223 187.294 176.073 195.374 182.421 TKS10MPI 104.342 200.01 200 199.99 200.01 200 200 200 159.076 TKS11MPI 200 186.72 153.709 185.137 179.69 180.453 173.34 181.221 104.48 TKSMR GSSBABR GBSB(BM) GCDBAB G66571 H526585 HHolo HUM- HUM- R R R R H K2- K1- 59PE 60PE

Farthest Squared neighbor Chord HCQPE G630571, G628572PE G630572, G719561PE G920572PE G920573PE HUM-K4- HUM- 10BPE 10CPE 59PE K1- 57PE TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM-K2- 59PE HUM-K1- 60PE HCQPE 0

92

Farthest Squared neighbor Chord G630571, 200 0 10BPE G628572PE 200 35.393 0 G630572, 200 18.508 5.851 0 10CPE G719561PE 200 1.926 47.033 27.042 0 G920572PE 200 98.886 200 167.627 79.468 0 G920573PE 200 98.886 200 167.627 79.468 0 0 HUM-K4- 173.429 39.554 106.988 80.654 29.527 39.115 39.115 0 59PE HUM-K1- 168.974 66.552 119.879 98.719 58.829 70.823 70.823 18.691 0 57PE H35-29- 58.098 188.66 185.614 196.376 186.48 177.56 177.56 143.191 139.912 5babPE HUM-K3- 56.764 187.042 188.364 196.075 184.451 175.751 175.751 127.452 98.526 61PE HUM-K1- 183.894 158.726 200.01 186.793 150.798 118.352 118.352 118.602 106.847 55PE TKS01PE 129.399 200 194.007 200 200 200 200 187.651 185.575 TKS02PE 100.675 200 190.703 200 200 200 200 182.999 180.145 HCFAMPE 200.01 118.352 189.408 173.866 102.67 38.492 38.492 70.081 95.687 HUM-K1- 127.889 200 169.079 200 200 200 200 180.84 177.624 59TPI HTypeSecTPI 48.863 199.99 199.99 199.99 199.99 199.99 199.99 158.498 159.371 HNWSE 33- 103.58 192.223 175.872 197.503 190.732 184.628 184.628 153.114 128.956 29-16db TPI G12-35- 200 200 200 200 200 200 200 150.377 109.264 31TPI TKS08MPI 172.395 200 200 200 200 200 200 192.774 183.538 TKS10MPI 135.755 200 200 200 200 200 200 200 199.99

93

Farthest Squared neighbor Chord TKS11MPI 122.693 200 188.319 200 200 200 200 189.315 168.831 HCQPE G630571, G628572PE G630572, G719561PE G920572PE G920573PE HUM-K4- HUM- 10BPE 10CPE 59PE K1- TKSMR 57PE

Farthest neighbor Squared Chord H35-29- HUM-K3- HUM-K1- TKS01PE TKS02PE HCFAMPE HUM-K1- HTypeSecTPI 5babPE 61PE 55PE 59TPI TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM-K2-59PE HUM-K1-60PE HCQPE G630571, 10BPE G628572PE G630572, 10CPE G719561PE G920572PE G920573PE HUM-K4-59PE HUM-K1-57PE

94

Farthest neighbor Squared Chord H35-29-5babPE 0 HUM-K3-61PE 20.816 0 HUM-K1-55PE 166.295 152.779 0 TKS01PE 130.153 133.591 190.938 0 TKS02PE 101.504 106.827 186.721 5.691 0 HCFAMPE 158.264 155.373 134.074 176.279 170.413 0 HUM-K1-59TPI 43.438 67.662 200.01 137.8 109.36 149.264 0 HTypeSecTPI 30.157 50.757 181.663 127.77 99.561 200 105.577 0 HNWSE 33-29- 58.491 50.088 149.526 125.331 93.404 130.857 36.405 95.212 16db TPI G12-35-31TPI 192.136 127.86 165.196 200 200 200.01 200 199.99 TKS08MPI 198.351 170.432 182.294 184.461 177.238 200.01 200 186.98 TKS10MPI 200.01 179.268 175.853 163.837 147.027 200.01 200 169.712 TKS11MPI 169.791 145.656 142.418 145.648 119.446 180.84 144.092 163.556 H35-29- HUM-K3- HUM-K1- TKS01PE TKS02PE HCFAMPE HUM-K1- HTypeSecTPI TKSMR 5babPE 61PE 55PE 59TPI

Farthest neighbor Squared Chord HNWSE 33-29-16db TPI G12-35-31TPI TKS08MPI TKS10MPI TKS11MPI TKSMR GSSBABR GBSB(BM)R GCDBABR G66571R H526585R HHoloH HUM-K2-59PE HUM-K1-60PE HCQPE

95

Farthest neighbor Squared Chord G630571, 10BPE G628572PE G630572, 10CPE G719561PE G920572PE G920573PE HUM-K4-59PE HUM-K1-57PE H35-29-5babPE HUM-K3-61PE HUM-K1-55PE TKS01PE TKS02PE HCFAMPE HUM-K1-59TPI HTypeSecTPI HNWSE 33-29-16db TPI 0 G12-35-31TPI 153.625 0 TKS08MPI 172.941 84.647 0 TKS10MPI 155.559 127.993 22.843 0 TKS11MPI 78.604 167.583 145.879 84.12 0 TKSMR HNWSE 33-29-16db TPI G12-35-31TPI TKS08MPI TKS10MPI TKS11MPI

96

F. Modern Analog Modern Analog Analysis using the Jaccard Coefficient (Presence/Absence) and the 600 sites in the NANODe database (www.kent.edu/NANODe also available in Neotoma (www.neotomadb.org ) for the following sites in this study:

Best Modern Analog Score for Late Pleistocene site G-HibDC

0.8 0.7

0.6

0.5

0.4

0.3 Jaccard Coefficient

0.2

0.1

0 0 10 20 30 40 50 60 70 Ostracode Sample Sites in NANODe (top 70 sites of 600)

Best scores (two): Best score = 0.68, Horseshoe Lake, South Dakota, AJS-24, TDS= 19,0000 mg/L, Mean Annual Precipitation = 534 mm, Mean Annual Temperature = 6.1 ⁰C Best score = 0.68, Lazy Lagoon New Mexico, LCO-75, TDS = 36,000 mg/L, Mean Annual Precipitation = 265 mm, Mean Annual Temperature = 16 ⁰C

97

Best Modern Analog Score for Mid-Pleistocene site TCud1 (Cudahy) 0.7

0.6

0.5

0.4

0.3

Jaccard Coefficient 0.2

0.1

0 0 10 20 30 40 50 60 70 80 90 Ostracode Samples Sites in NANODe (top 86 sites of 600)

Best Score 0.6 Mantua Bog (Fen) Wet Meadow, Ohio, TDS = 262 mg/L, Mean Annual Precipitation = 947 mm, Mean Annual Temperature = 9.5 ⁰C Second best score Cisco Lake, Wisconsin , LCO Site 302 TDS = 40 mg/L , Mean Annual Precipitation = 808 mm, Mean Annual Temperature = 4.4 ⁰C

98

Best Modern Analog Score for Late Pliocene site HCotRex 0.35

0.3

0.25

0.2

0.15

Jaccard Coefficient 0.1

0.05

0 0 10 20 30 40 50 60 70 Ostracode Sample Sites in NANODe (top 62 sites of 600)

Best Score = Little Wall Lake, Iowa, LCO site 579 TDS = 264.1 mg/L, Mean Annual Precipitation = 946 mm, Mean Annual Temperature = 11.1 ⁰C

99