<<

Spatial and Temporal Geochemical Characterization of Aeolian Material from the McMurdo Dry Valleys,

A Thesis Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in the Graduate School of The Ohio State University

By Melisa A. Diaz Graduate Program in Earth Sciences

The Ohio State University 2017

Master’s Examination Committee: Dr. W. Berry Lyons, Advisor Dr. Michael Barton Dr. Michael Wilkins

Copyright by Melisa A. Diaz 2017

Abstract

Aeolian processes play an important role in the transport of both geological and biological materials globally, on the biogeochemistry of ecosystems, and in landscape evolution. As the largest ice free area on the Antarctic continent (approximately 4800 km2), the McMurdo Dry Valleys (MDV) are potentially a major source of aeolian material for Antarctica, but information on the spatial and temporal variability of this material is needed to understand its soluble and bulk geochemistry, deposition and source, and hence influence on ecosystem dynamics. 53 samples of aeolian material from

Alatna Valley, Victoria Valley, Miers Valley, and (Taylor Glacier, East

Lake Bonney, F6 (Lake Fryxell), and Explorer’s Cove) were collected at five heights (5,

10, 20, 50, 100 cm) above the surface seasonally for 2013 through 2015. The sediment was analyzed for soluble solids, total and organic carbon, minerology, and bulk

- - chemistry. Of the soluble component, the major anions varied between Cl and HCO3 , and the major cation was Na+ for all sites. Soluble N:P ratios in the aeolian material reflect nutrient limitations seen in MDV soils, where younger, coastal soils are N-limited, while older, up valley soils are P-limited. Material from East Lake Bonney was P-limited in the winter samples, but N-limited in the full year samples, suggesting different sources of material based on season. Analysis of soluble salts in aeolian material in Taylor Valley compared to published soil literature demonstrates a primarily down valley transport of materials from Taylor Glacier towards the coast. The bulk chemistry suggests that the

ii

aeolian material is highly unweathered (CIA values less than 60 %), but scanning electron microscope images show alteration for some individual sediment grains. The mineralogy was reflective of local rocks, specifically the McMurdo Volcanics, Ferrar

Dolerite, Beacon and granite, but variations in major oxide percentages and rare earth element signatures could not be explained by mixing lines between these four rock types. This potentially suggests that there may be an additional, and possibly distant, source of aeolian material to the MDV that is not accounted for. This work provides the first fully elevated spatial and temporal analysis of the geochemistry of aeolian material from the Dry Valleys, and contributes to a better understanding of sediment provenance and how aeolian deposition may affect surface biological communities.

iii

Acknowledgments

I would first like to thank Dr. Berry Lyons for his continuous efforts in helping me develop as a researcher. I have learned a copious amounts of information from him and look forward to learning even more in coming years.

Thank you to Kathy Welch and Chris Gardner for guidance in both life and science, and Sue Welch and Julie Sheets for thoughtful discussions, particularly with mineralogy.

To Sydney Olund, thank you for being a great office mate and an even greater friend.

I will miss our conversations and weekly debates.

To Cyrus, thank you for all your love and support. I will always appreciate your company during long work days.

Lastly, this work could not be possible without the help of the McMurdo LTER research team members Byron Adams, Alia Khan, and Andy Thompson, logistical support from ASC, helicopter support from PHI, and funding support from NSF ANT

1115245.

iv

Vita

October 1992 ...... Born – Queens, NY May 2014 ...... B.S. Earth and Environmental Sciences, University of Rochester January 2016 to present ...... Graduate Teaching/Research Associate, School of Earth Sciences, The Ohio State University

Field of Study

Major Field: Earth Sciences

v

Table of Contents

Abstract ...... ii Acknowledgments...... iv Vita ...... v List of Figures ...... viii List of Tables ...... x 1. Introduction ...... 1 1.1 Wind Patterns ...... 2 1.2 Aeolian Transport and Deposition ...... 3 2. Study Objectives and Questions ...... 6 3. Study Sites ...... 7 3.1 Miers Valley...... 8 3.2 Taylor Valley ...... 9 3.3 Victoria Valley ...... 10 3.4 Alatna Valley ...... 11 4. Methods...... 13 4.1 Sample Collection ...... 13 4.2 Sample Processing and “Clean” Technique Philosophy ...... 15 4.3 Water Soluble “Leaches” ...... 15 4.4 Nutrients ...... 16 4.5 Cations and Anions ...... 17 4.6 Total Nitrogen and Total and Organic Carbon ...... 18 4.7 Mineral Geochemistry ...... 19 4.8 X-ray Diffraction (XRD) ...... 19 4.9 Scanning Electron Microscopy (SEM) ...... 20 5. Results ...... 22 5.1 Wind Distribution ...... 22 5.2 Soluble Solids ...... 22

vi

5.3 Total and Organic Carbon ...... 24 5.4 Sediment Geochemistry ...... 25 5.5 Mineralogy ...... 26 5.6 Grain Size, Composition and Encrustations from SEM ...... 27 6. Discussion ...... 31 6.1.1 Aeolian Composition- Water Soluble Component ...... 32 6.1.2 Spatial and Seasonal Variability ...... 34 6.1.3 Down Valley Transport...... 38 6.2.1 Aeolian Composition- Solid Component ...... 42 6.2.2 Geographic Provenance ...... 46 6.3.1 Analysis of Variance (ANOVA) ...... 50 7. Conclusions ...... 52 8. Future Work ...... 54 9. References ...... 56 10. Figures and Tables ...... 65 Appendix A: Detection limits and Precision and Accuracy ...... 134 Appendix B: Total Mass of Aeolian Sample ...... 139 Appendix C: Analysis of Variance (ANOVA) Results ...... 142

vii

List of Figures

Figure 1. Valleys of interest in the McMurdo Dry Valleys, Antarctica ...... 66

Figure 2. Sample locations within the McMurdo Dry Valleys ...... 67

Figure 3. Isokinetic aeolian material collector ...... 68

Figure 4. Box plot of daily average wind speeds from locations in the McMurdo Dry

Valleys ...... 69

Figure 5. Total, organic, and inorganic carbon ratios with distance from the coast ...... 70

Figure 6. Major oxide, trace and rare earth element spider diagrams by location ...... 71

Figure 7. Major oxide, trace and rare earth element spider diagrams by height...... 72

Figure 8. Mineralogy via x-ray diffraction ...... 73

Figure 9. Scanning electron microscope imaging of Alatna Valley ...... 75

Figure 10. Scanning electron microscope imaging of Victoria Valley ...... 76

Figure 11. Scanning electron microscope imaging of Explorer’s Cove ...... 77

Figure 12. Scanning electron microscope imaging of F6 (Lake Fryxell) ...... 78

Figure 13. Scanning electron microscope imaging of East Lake Bonney ...... 79

Figure 14. Scanning electron microscope imaging of Taylor Glacier ...... 80

Figure 15. Scanning electron microscope imaging of Miers Valley ...... 81

Figure 16. Ternary diagrams for major anions and cations ...... 82

2+ 2+ - Figure 17. Ca and Mg versus HCO3 ...... 83

Figure 18. Na+ versus Cl- ...... 84

viii

+ 2+ - - Figure 19. Na and Ca versus Cl and HCO3 ...... 85

2- - - - Figure 20. SO4 /Cl versus NO3 /Cl for all locations ...... 86

2- - - - Figure 21. SO4 /Cl versus NO3 /Cl for Taylor Valley ...... 87

Figure 22. Cross section of major geomorphological features in Taylor Valley ...... 88

2- 3- Figure 23. NO3 to PO4 stoichiometric ratios with distance from the coast ...... 89

Figure 24. Chemical Index of Alteration (CIA) diagrams ...... 90

Figure 25. Major oxide variation diagrams...... 91

Figure 26. Major oxide potential mixing diagram ...... 92

Figure 27. ƐNd and 87Sr/86Sr fields of potential source areas ...... 93

Figure 28. ƐNd and 87Sr/86Sr isotopic signatures for McMurdo Dry Valley and McMurdo

Sound aeolian material ...... 94

Figure 29. Rare earth element signatures ...... 95

ix

List of Tables

Table 1. Sample location coordinates, elevation, and distance from coast...... 98

Table 2. Daily average wind speeds from locations in the McMurdo Dry Valleys ...... 99

Table 3. Chemical Index of Alteration (CIA) and Loss on Ignition (LOI)...... 100

Table 4. Water soluble major anions ...... 102

Table 5. Water soluble major cations ...... 106

- 3- Table 6. Water soluble nutrients (NO3 and PO4 ) ...... 110

Table 7. Total (TC), organic (OC) and inorganic (IC) carbon ...... 113

Table 8. Major oxides in bulk component ...... 116

Table 9. Trace elements in bulk component ...... 119

Table 10. Rare earth elements in bulk component ...... 129

Table 11. Total sediment flux rates in the McMurdo Dry Valleys ...... 133

Table 12. Accuracy and precision for analytes ...... 135

Table 13. Detection limits for oxides, trace elements, and rare earth elements ...... 136

Table 14. Total mass of aeolian material collected from each location ...... 140

Table 15. One-way Analysis of Variance (ANOVA) for ions...... 143

Table 16. One-way Analysis of Variance (ANOVA) for major oxides ...... 145

Table 17. One-way Analysis of Variance (ANOVA) for trace elements and REEs ...... 147

x

1. Introduction The McMurdo Dry Valleys (MDV) are a polar desert landscape of soils, perennial ice-covered lakes, ephemeral streams, and alpine glaciers located in Southern Victoria

Land in (Fountain et al., 1999). The MDV have been investigated as a

Long-Term Ecological Research site (MCM-LTER) since 1993. Over this time period, a number of locations within the Dry Valleys, especially Taylor Valley, have been the focus of integrated ecological, climatological, hydrological, limnological, and biogeochemical research. Minor, by temperate standards, climatic changes in this unique environment can induce rapid hydrological change to drastically affect the physical environment and the capabilities of organisms to grow and develop (Fountain et al.,

1999; Fountain et al., 2014).

As the largest ice free area on the Antarctic continent (approximately 4800 km2), the MDV are potentially a major dust source for Antarctica. Other than the ephemeral streams, which transmit liquid water 4-12 weeks annually connecting glaciers to primarily closed basin lakes through fixed stream channels (Fountain et al., 1999), winds are the primary mechanism of connectivity for the various landscape units in distributing organisms and particulate matter. The direction and magnitude of winds, and the mass transport of aeolian material in MDV has been previously investigated (Lancaster 2002;

Doran et al., 2002; Nylen et al., 2004; Šabacká et al., 2012), but studies focused on the geochemistry of aeolian material have relied solely on wind-blown sediment trap

1

collections at only 30 cm above the ground (Šabacká et al., 2012; Deuerling et al., 2014), or on examining sediments deposited on glacial and ice covered lake surfaces (Lyons et al., 2003; Witherow et al., 2006; Deuerling et al., 2014). Determining the composition and solubility of Antarctic dust at multiple elevations and locations will help to predict the source of this material, as well as its influence on glacier melt chemistry and controls on nutrient fluxes to the aquatic ecosystem.

1.1 Wind Patterns:

Winds in the modern McMurdo Dry Valleys originate from different locations, including down valley, coastal and drainage winds, which flow down valley sides and alpine glaciers from higher elevations (Doran et al., 2002; Nylen et al., 2004). Wind direction is generally dependent on valley orientation, with the exception of strong katabatic events blowing from the East (Clow et al., 1988; Nylen et al., 2004). In the

MDV, katabatic winds, specifically föhn winds, originate from the polar plateau (180–

315°) where cold dense air sinks and flows downslope. The air mass adiabatically warms during transport as it increases in speed (Ishikawa et al., 1982; Nylen et al., 2004), which contributes to warmer air temperatures and high velocity events. In the winter, föhn winds (identified by speeds above 5 m s-1) are dominant and reach speeds of 40 m s-1.

Coastal winds are stronger in the summer and reach speeds of 20 m s-1 (Nylen et al.,

2004).

The climate of the MDV is influenced by the magnitude and frequency of föhn events, with spatial variations dependent on proximity to the ice sheet (Nylen et al.,

2

2004). In general, winter temperatures in the Dry Valleys are higher than average coastal temperatures around the continent (Doran et al., 2002). Föhn events can increase average winter temperatures by as much as 4.3°C, but have less influence (maximum of 0.4°C) on average summer temperatures (Nylen et al., 2004). Nylen et al. (2004) observed that the frequency of föhn winter events increases by 14% for every 10 km up valley towards the

East Antarctic Ice Sheet, while summer events increase by 3%. These föhn events help to explain the discrepancies in the expected temperatures from the elevation driven adiabatic lapse rate in Taylor Valley (Doran et al., 2002; Nylen et al., 2004) and are integral to material transport (Lancaster, 2002; Šabacká et al., 2012; Gillies et al., 2013).

1.2 Aeolian Transport and Deposition:

As noted above, strong winds are an important feature of the MDV and they can lead to the transport and deposition of sediment and other materials. In the Dry Valleys, sediment is transported via three main mechanisms: suspension, saltation, and traction

(Atkins and Dunbar, 2009; Lancaster et al., 2010). Suspension is where the particles remain in the air mass they are transported in. These particles are usually small in diameter (<70 µm) and are typically associated with silt and light sand. For saltation, slightly larger and heavier grains are lifted off the bed. Lastly, during traction, larger and heavier grains, usually gravel sized, roll or slide along the surface (Pye, 1987; Atkins and

Dunbar, 2009). Aeolian deposition near the valley floor is predominately sand sized particles, while glacier and ice-covered surfaces have silt and clay sized particles

(Lancaster, 2002; Lancaster et al., 2010). Victoria Valley and East Lake Bonney are

3

dominated by sand transport at 100 cm above the surface, which is likely due to the abundance of sandy soils at higher elevations (Lancaster, 2002).

The highest particle fluxes in the MDV were measured at the highest elevation sites, such as near the Bonney Riegel (~250 m), while lower elevation measurements showed that more mass is transported near soil surfaces (30 cm) than above the ground

(100 cm) (Lancaster, 2002; Šabacká et al., 2012). The frequency of katabatic and föhn events greatly affects the volume of material transported, where 24 more hours of strong winds can increase the aeolian sediment flux by nine-fold (Šabacká et al., 2012). Sand transport wind shear threshold velocity in the MDV has been determined to be between

0.30 and 0.35 m s-1, which corresponds to wind speeds between 6.5 and 7.2 m s-1 at a collection height of 6 m (Lancaster et al., 2010). The critical moisture threshold for transport is 4-6% (Wiggs et al., 2004). These conditions are most commonly met during the Antarctic winter, but are only satisfied in Taylor Valley less than 5% of the time

(Lancaster et al., 2010; Šabacká et al., 2012).

The flux and subsequent deposition of aeolian material varies based on location and collection height. Šabacká et al. (2012) recorded the highest sediment fluxes in

Taylor Valley at a collection height of 30 cm near the center of the valley (289 g m-2 yr-1 at ), while the lowest fluxes were in the upper parts of the valley (8.5 g m-2 yr-

1 at Lake Bonney). This pattern does not apply for more elevated collection heights.

Lancaster (2002) studied aeolian material collected 100 cm above the surface and measured the highest aeolian flux at Lake Bonney and <1 g m-2 yr-1 at Lake Hoare. The material at 30 cm is predominately sand sized and is closer to the saltation zone of nearby

4

soils, which helps to account for these discrepancies (Lancaster et al., 2010; Gillies et al.,

2013; Deuerling et al., 2014). Physical obstacles in the landscape of Taylor Valley, such as the Nussbalm and Bonney Riegels, likely impede aeolian transport and also affect flux rates (Fountain et al., 1999; Šabacká et al., 2012). These landforms increase local deposition (Lyons et al., 2003), but depositional rates of 0.66 g m-1 yr-1 (Lancaster, 2002) observed in Taylor Valley are an order of magnitude lower than those in other arid regions, such as the Tibetan Plateau (Honda et al., 2004; Xu et al., 2012; Wei et al.,

2017). Therefore, although aeolian transport and deposition within the MDV are important to landscape connectivity, compared to other arid locations, there appears to be less material to be transported.

5

2. Study Objectives and Questions:

In order to understand the source, deposition and geochemistry of MDV aeolian material, temporal and spatial profiles of aeolian dust need to be characterized. This research will provide the first analysis of aeolian materials from different elevations at specific locations in the MDV on a temporal basis. Specific aims include:

1. To describe the bulk chemistry and minerology of aeolian material

collected at various heights above the soil surface.

2. To determine the chemical composition and variation of aeolian material

with respect to collection height above the surface, location, and season.

3. Identify the ultimate source of aeolian material to MDV.

Deuerling et al. (2014) concluded that the material collected at 30 cm

above the surface was derived from local soils. Material at 100 cm above the

surface is expected to be finer grained (Lancaster, 2002), and has the potential

to travel from further sources. A key question of this work is to determine

whether MDV aeolian material can be derived from outside MDV, and even

perhaps from outside the Antarctic continent.

6

3. Study Sites:

In general, mean annual MDV valley floor temperatures vary from -15°C to -

40°C and annual precipitation is generally less than 10 cm (Doran et al., 2002; Sugden and Denton, 2004; Fountain et al., 2010). The geology of MDV is predominately composed of four main rock types. These include Cambrian/Pre-Cambrian granites and metamorphic basement, - age strata in the

Transantarctic Mountains, including a collection of , shales, conglomerates, and (Craddock, 1970; Kyle, 1990), age Ferrar Dolerite of tholeiitic affinity

(Bockheim and McLeod, 2007), and the McMurdo Volcanic Group of Cenozoic age containing basalts and quartz-free basic to felsic volcanic rocks (Kyle, 1990; Marchant and Denton, 1996).

The lower elevation valleys (Miers and Taylor) have changed dramatically through the Pleistocene into the Holocene. During glacial maxima, the lakes in these valleys were larger and the tills were modified by the deposition of lacustrine sediments

(Squyres et al., 1991; Doran et al,. 1994; Lyons et al., 2000; Toner et al., 2013). In addition, during these times, the eastern ends of the valleys were inundated by the West

Antartic Ice Sheet (WAIS), blocking the outflow of water from these lakes (Denton et al.,

1989; Hall et al., 2013). During the interglacial periods, the

(EAIS) advanced seaward, while the alpine glaciers similarly advanced into the valleys

(Hall et al., 1993; Wilch et al., 1993). Throughout the Holocene, the eastern ends of these 7

valleys were again “reconnected” to the , and the lake levels decreased dramatically, but fluctuated in size depending on the variations in austral summer temperatures (Hall et al., 2013).

The valley floors consist of tills deposited by the movement of both the EAIS and

WAIS, as well as local alpline glaciers (Denton et al., 1989; Toner et al., 2013). These deposits are heavily modified by aeolian, fluvial, and fluvio-glacial and limnological activity, and therefore have higher amounts of sand compared to higher elevations

(Prentice et al., 1998). Because of the different influences of these various processes over time, the till and lacustrine sediments that are now exposed vary within each valley in

MDV and are potentially sources of local “dust”. Detailed interpretations of geologic, climatic conditions, and glacial history within each valley can provide insight into the chemistry of potential aeolian material and its impact on surface chemistry.

3.1 Miers Valley:

Located in the southern Dry Valleys, Miers Valley trends northwest to southeast at approximately 50 m above sea level. It extends from the in the west to the coast of McMurdo Sound. Unlike other valleys in the region which are isolated from the coast by glaciers or topographic highs, Miers Valley directly interacts with McMurdo Sound via discharge from Miers Stream. Two glacial streams enter Lake

Miers and one flows out to the Sound. Miers Valley contains the only large flow-through lake in the MDV. The mean annual temperature is comparable to that of other low

8

elevation southern and central valleys at approximately -20°C, and has low precipitation at less than 10 cm per year (Doran et al., 2002).

The valley floor is predominately weathered granite with quartz and marble pebbles ubiquitous in the desert pavement (Cowan et al., 2011) and is characterized by a highly eroded and porous surface (Clayton-Greene et al., 1988). The surficial materials are characteristic of a lacustrine system populated by calcite, gypsum, and aluminosilicate minerals (Clayton-Greene et al., 1988), which are likely remnants from the presence of a former proglacial lake, Glacial Lake Trowbridge (Pewe, 1960). Many of the outcrops also have carbonate coatings on their unexposed side, further supporting the inundation of the valley (Clayton-Greene et al., 1988). The presence of basaltic material, such as kenyte, indicates glacial drifts originating from the East, where the McMurdo

Volcanic Group is most prominent (Clayton-Greene et al., 1988).

3.2 Taylor Valley:

Taylor Valley is one of the most studied valleys in the McMurdo Dry Valleys. It is centrally located with respect to other valleys (i.e. Alatna and Miers), and trends east to west. The elevation ranges from sea level near the coast to over 300 m at Taylor Glacier.

Similar to Miers Valley, Taylor Valley is directly abutted to the ocean to the East. The mean annual temperature is -20°C with total annual precipitation less than 3 cm (Doran et al., 2002; Fountain et al., 2010). During the summer when solar irradiance is highest, water flows from glaciers in the valley into three major lakes: Lake Fryxell, Lake Hoare, and Lake Bonney (Figure 2). The entire valley was inundated by ancient Glacial Lake

9

Washburn between 11,000 and 24,000 years ago (Denton et al., 1989), which contributes to lacustrine sediments observed at various locations within the valley. Recent work by

Toner et al. (2013) has provided great detail on the timing and elevations of lake level fluctuations during the Last Glacial Maximum into the early Holocene, and the geochemical ramifications of these fluctuations.

The valley floor is a diverse network of tills of differing age and compositions, which is reflective of a complex glacial history with the advance and retreat of the WAIS and EAIS, and alpine glaciers (Porter and Beget, 1981; Burkins et al., 2000), and is characterized by mostly sandy gravel and exposed bedrock (Fountain et al., 1999). The central part of the valley contains metasedimentary rocks of the Ross System

(Harrington, 1958). There are large marble outcrops, some individual slabs as thick as 60 m. Along with the marble slabs are schists and quartzites with iron straining, relics from active tectonics and volcanism (Angino et al., 1962). Near the coast are granites and granodiorites which are believed to be younger than the central valley sedimentary rocks

(Angino et al., 1962).

3.3 Victoria Valley:

Victoria valley is also a centrally located valley. It is bounded to the south by the

Olympus Range, to the west by the inland ice sheet, to the north by the Clare and St. John ranges and is separated from McMurdo Sound by the Wilson Piedmont Glacier (Calkin,

1964). Victoria Valley consists of 5 main interlocking valleys totaling 90 km in length and generally trends northwest to southeast. It is a high elevation valley, with the lowest

10

elevation at approximately 350 m. The west ends of the valleys are high bedrock thresholds, which impede the advancement of glaciers, except at times when the surface of the inland ice was much higher than present (Calkin, 1964). Victoria Valley is thought to have remained largely ice free for the last 13.5 million years (Sugden and Denton,

2004). Intense thermal inversions develop from the large expanse of ice free area, which results in a mean annual temperature of -27.4°C, colder than Miers and Taylor Valleys

(Doran et al., 2002; Nylen et al., 2004).

The valley floor contains igneous, metamorphic, and sedimentary rocks of

Cambrian or Late Cambrian to Mesozoic Age (McKelvey and Webb, 1962). Larger rocks consist of white granular marbles, paragneisses, granulites, and quartzofeldspathic schists, which were subject to strong folding (McKelvey and Webb, 1962; Allen and

Gibson, 1962). The basement complex of folded metasediments of the Koettlitz Marble is cut by younger rocks of the Granite Harbor Intrusive complex (Gunn and Warren 1962).

Very large sills and dikes of Ferrar Dolerite of Jurassic-Cretaceous age cut through

Beacon rocks and the underlying basement complex. The geology consists of complex folded marble, gneisses and coarse grained granitic intrusives, and sandstones that overlie dolerite sills.

3.4 Alatna Valley:

Alatna Valley is located in the northern most part of the MDV in the Convoy

Range near Mackay Glacier. It is a northwest to southeast trending valley and has the highest elevation (>900 m) of all other study sites. The landscape achieved its present

11

form in the warmer middle Miocene through the formation and movement of both rivers and glaciers (Denton et al., 1993; Sugden and Denton, 2004). This is evidenced by the preservation of unconsolidated deposits, presence of small Miocene volcanic cones

(Wilch et al., 1993), and abundance of fjord sediments (Prentice et al., 1993). The surface also has giant potholes and meltwater channels from the warmer and moister climate

(Calkin, 1964; Sugden and Denton, 2004). As a high elevation, northern valley, the average annual temperature likely ranges from -30 to -40°C (Sugden and Denton, 2004).

This is an estimate due to the lack of long term meteorological stations in the valley.

Basement rocks are truncated by the Kukri erosional surface with interbedded tillites. Surrounding erosional surfaces have dolerite, granite, sandstone and orthoquartzite (Calkin, 1964). The valley floor is nearly covered by till and by roches moutonnees (Pocknall et al., 1994). Unlike in Victoria Valley, glacial deposits in the west suggest that much of the central valley and areas near Mt. Gran have been uncovered only during the latter part of the last glaciation (Calkin, 1964), but eastern soils in the area have been exposed for longer (Sugden and Denton, 2004). Quartz sandstone is the dominant lithologic type, while Ferrar Dolerite is the youngest unit, which is likely

Jurassic or Cretaceous in age, and occupies the largest part of this area. Hornblende- biotite gneiss is believed to be the oldest rock (Mirsky et al., 1965).

12

4. Methods:

Aeolian material from Alatna Valley, Victoria Valley, Miers Valley, and Taylor

Valley (Figure 1) was collected seasonally (summer and winter) in 2013 and 2014, and year-round for 2015, with more years planned. 53 total samples were obtained, but only

38 had 5 or more grams of material for geochemical and mineralogical studies. Samples were collected twice: during the months of November and January. November collections are termed “winter”, from January 15th to October 31st, while January collections are termed “summer”, from November 1st to January 14th. Sampler locations and coordinates are detailed in Table 1 and Figure 2. Daily average wind speed data were downloaded from the MCM-LTER website (www.mcmlter.org) and were used to calculate average seasonal wind speed (Table 2).

4.1 Sample Collection:

Aeolian materials were collected using Big Spring Number Eight (BSNE) isokinetic wind samplers at seven collection sites throughout the McMurdo Dry Valleys

(Figure 3). BSNE samplers passively collect 95% of airborne material that enters the collection box opening, regardless of wind velocity or direction. Before going into the field, sterile Whirl-Pak bags were weighed. These collection bags were brought into the field along with sterile alcohol wipes, a knife, phillips screw driver, flat head screw driver, and duct tape. At the BSNE trap, each dust collection tray was separated into two 13

components to extract and collect the dust. The bags were labeled with the appropriate location identifier and tray identifier (e.g. “ELB Top” for the 100 cm collection height at

East Lake Bonney).

Each collection tray was secured to an upright pole using a combination of bolts, thick rubber belts, and duct tape. Upon arriving at the site, the duct tape and rubber around the collection tray was removed. The bottom compartment of the tray was carefully removed, with caution to not expose the entire collection tray and re-mobilize the sediment due to the current wind conditions. The material was gathered in one corner of the tray by lightly tapping the bottom. This ensured minimal contamination. The contents of the tray were then poured into the clean, pre-weighed, and labelled Whirl-Pak bag. After all the material was gathered from the collector, the bottom tray was wiped clean with the sterile alcohol pad to remove any residue. In many cases, there was very little material in the collection tray at the glacier locations, therefore the alcohol pad was saved to be used in potential biological analyses. The cleaned collection tray was then reattached and secured to the pole with the rubber belt and new duct tape. Though the aeolian material collection trays were initially set to standardized heights (5, 10, 20, 50,

100 cm respectively), strong winds have likely affected them, effectively altering the true collection height. Therefore, relative elevation terms (bottom lower, bottom middle, bottom upper, middle, top) are used rather than absolute heights (Figure 3). This procedure was repeated for all five collection trays at each location in January and

November from 2013 until 2015. Starting in 2015, aeolian material collections were done annually instead of seasonally.

14

Upon returning to a field lab or to the Crary Laboratory at McMurdo Station, the samples were placed directly into a -20°C freezer. Once at McMurdo Station, the samples were weighed in their bags and recorded. Samples from between November 2013 to

January 2015 were sent frozen to the University of Colorado at Boulder, then transported chilled by automobile to The Ohio State University. Full year samples collected for 2015 were shipped frozen directly to The Ohio State University. All samples were stored at -

20°C before, throughout, and after analysis.

4.2 Sample Processing and “Clean” Technique Philosophy:

Throughout sample analysis, all tubes and equipment were cleaned with several deionized (DI) water rinses. Acid was generally avoided due to potential contamination from the dissolution of strong acids (i.e. HCl and HNO3), with the exception of a ceramic mortar and pestle. The mortar and pestle was rinsed with 1% HCl and subsequently rinsed several times with DI water to achieve aseptic conditions for carbon analysis.

4.3 Water Soluble “Leaches”:

For the samples with greater than 5 grams of mass, 0.5 grams of sample were weighed and placed in a Fisherbrand™ polystyrene disposable beaker. 2.5 ml of DI water chilled to 3°C was added to the beaker to obtain a water to aeolian material ratio of 5:1, consistent with soil leaches performed by MCM-LTER researchers. Samples were agitated for 1 minute by hand, then placed in a refrigerator at 3°C for 30 minutes. This was the first of two leaches. Chilled DI water was used to represent surface water

15

temperatures in MDV. After 30 minutes, the water from the first leach was extracted from the beaker using a 10 ml syringe, filtered using a 0.45 micron Whatman™ polypropylene syringe filter, and stored chilled in a Falcon™ tube at 3°C. The first 2-3 drops of leachate were discarded to minimize potential contamination from the filtration process. The same leached sediment was then rehydrated with 2.5 ml of DI water, agitated for 1 minute, and left for 24 hours in the refrigerator. After 24 hours, this second leachate was filtered using a new 0.45 micron filter and stored in a separate Falcon tube.

These two-time steps were chosen because it was thought that the 30-minute leachate ideally simulates short pulses of water hydrating aeolian sediment in MDV to dissolve rapidly soluble solids, while the 24-hour leachate simulates long term wetting. Samples with ~5 grams of mass were leached using the entire mass of the sample minus 0.5 grams that were later used for elemental analyzer analyses. The leached 5 grams was later used for X-ray fluorescence (XRF) analyses. Filter blanks were produced by filtering chilled

DI water through a clean syringe and filter, and later analyzed along with the sample leachates. Contamination from filtering was negligible, as the measured ion and nutrient concentrations were less than the concentrations found in the DI water (Table 12).

4.4 Nutrients:

The water collected from the sediment leaches was separated into 2 aliquots. The first aliquot was for soluble reactive phosphate and nitrate plus nitrite on a Skalar San++

Automated Wet Chemistry Analyzer at The Ohio State University. 1 ml of sample was diluted with DI water to a total volume of 10 ml and transferred to a new Falcon tube.

16

Samples were taken directly from the new Falcon tubes using an SA 1050 Random

Access Auto-sampler. Preliminary work showed that the samples had low phosphorus concentrations, but had nitrogen levels above the detection limit. Since the leachate was diluted, a range of standards (10 µg L-1 to 1000 µg L-1 for N and P) were analyzed with the samples. A seven-point calibration curve, using DI water as a forced zero, was generated to determine sample concentrations using the San++ Flow Access software for

Windows. DI water and duplicates were analyzed throughout the run to monitor instrument drift and to calculate precision, which was determined to be approximately

±5% for both phosphate and nitrate (Table 12).

4.5 Cations and Anions:

The remaining leachate was used to quantify major cations (K+, Na+, Ca2+, Mg2+)

- - - - 2- 3- and anions (F , Cl , Br , NO3 , SO4 , PO4 ) at The Ohio State University. Cations were analyzed using a Dionex DX-120 ion chromatographer (IC) with an AS40 automated sampler as originally described by Welch et al. (1996; 2010). Anions were analyzed using a Dionex ICS-2100 ion chromatographer and an AS-DV automated sampler. Of the remaining original sample leachate, 0.5 ml was diluted with DI water to a total volume of

5.5 ml. This diluted leachate was analyzed with six incrementally increasing standards ranging from near the detection limit, up to 10 µg L-1 for some analytes, as well as with a

US Geological Survey (USGS) standard. These standard measurements were later used to create a calibration curve from which sample concentrations were calculated. The USGS conducts a semiannual inter-laboratory comparison study to test laboratory quality

17

assurance. The results from each study are published online, of which the 2015 results were used to determine the accuracy of our measurements compared to the USGS standard. Every five samples, either a mid-range standard or a duplicate sample was analyzed to verify reproducibility of results, and to determine any instrument drift.

- Alkalinity (HCO3 ) was calculated though charge balance difference (Lyons et al., 2012).

Precision and accuracy of all measured analytes are listed in Table 12 and range from

0.13% to 21.0% for precision and from 0.18% to 18.7% for accuracy.

4.6 Total Nitrogen and Total and Organic Carbon:

Of the remaining un-leached aeolian sediment, 0.5 grams was ground to silt size using a mortar and pestle and analyzed for total and organic carbon, as well as nitrogen on a Costech ECS 4010 elemental combustion system (COHNS), located in the

Subsurface Energy Materials Characterization & Analysis Laboratory (SEMCAL) at The

Ohio State University. For organic carbon measurements, between 10 and 30 mg of sample were placed in small silver cups for analysis. All inorganic carbon is expected to be derived from carbonate minerals, therefore the samples were acidified in the silver cups using 10 to 15 µL of 1% HCl. Acid was added to samples until all carbonate was liberated, which was determined when the samples stopped forming bubbles. Known quantities of acetanilide were run with samples to create a calibration curve, as well as silver cup blanks to establish a zero point. For the total carbon and total nitrogen measurements, between 10 and 30 mg of sample were placed in small tin cups. Similar to the organic carbon measurement, known acetanilide and tin blanks were analyzed with

18

the samples. Inorganic carbon concentrations were determined by subtracting the organic carbon fraction from the total carbon. Precision ranged from 0.83% to 2.74 % and accuracy was approximately 3% (Table 12).

4.7 Mineral Geochemistry:

Between 5 and 10 g of un-leached sediment was sent to SGS Canada Inc. in

Lakefield, Canada for analysis of major oxides and a suite of 35 minor and trace elements. Five samples were analyzed on leached material due to a lack of adequate mass available to do otherwise. The samples were first crushed to where 75% of the material could pass through a 2 mm sieve, then pulverized to where 85% passed through 75 microns. Major oxide analysis was conducted on sample beads, which were made through ore grade borate fusion and analyzed via X-ray fluorescence (XRF). Major and trace elements (suite of 35) concentrations were determined through ICP-AES/ICP-MS analysis after sodium peroxide fusion. The whole rock XRF analysis required 3.0 g of material while the suite of minor and trace elements required 0.2 g. The limits for each element are listed in Table 13.

4.8 X-ray Diffraction (XRD):

XRD analyses were conducted at the Subsurface Energy Materials

Characterization & Analysis Laboratory (SEMCAL) at The Ohio State University. For the XRD analyses, approximately 2.5 grams of ground sediment were mounted on stainless steel sample holders for 14 samples. Top and bottom lower samples were

19

preferentially chosen to highlight mineralogical differences with height and were run on a

PANalytical X’pert Pro X-ray Diffractometer equipped with a high speed X’Celerator detector. The run was set to 12 minutes per sample and samples were scanned from 4° to

70° 2θ at 45 kV, with a step size of 0.020º 2θ for 2 s/step. Peaks were identified with a minimum significance of 1.00, minimum tip width of 0.10, maximum tip width of 1.00, and peak base width of 2.00, where unmatched peaks were identified manually using

PANalytical HighScore Plus software. The PDF 4+ mineral database was used to match patterns of d-spacings in the database to experimental d-spacings to identify minerals in the samples.

4.9 Scanning Electron Microscopy (SEM):

High resolution images were obtained for each site with an FEI Quanta 250 Field

Emission Gun (FEG) scanning electron microscope (SEM) equipped with a Bruker X-

Flash X-ray spectrometer and Bruker Quantax software at the Subsurface Energy

Materials Characterization and Analysis Laboratory (SEMCAL) at The Ohio State

University to examine the rounding and sorting of aeolian sediment grains, grain encrustations, and any biological material present. Parameters were set to an accelerating voltage of 15.00 keV, a working distance of 13.6 mm, and a pixel resolution of 0.2 microns. Spot chemistry was performed using energy dispersive X-ray spectroscopy

(EDXS) analyses with a working volume of 1-3 µm to identify specific elements and predict minerology. X-rays are produced as the electron beam causes specific energy transitions of electrons in elements comprising the material being analyzed, which allows

20

the user to obtain chemical elemental information on a very small sample volume.

Elemental maps were also generated to detail salt encrustations, such as NaCl and CaSO4, on grain surfaces.

21

5. Results:

5.1 Wind Distribution:

Average daily wind speeds downloaded from the MCM-LTER website for five locations in the McMurdo Dry Valleys (Explorer’s Cove, Lake Bonney and Taylor

Glacier within Taylor Valley, Miers Valley, and Victoria Valley) were plotted from

January 15th, 2013 through 2015 (Figure 4). Long term meteorological data was not available for F6 (Lake Fryxell) and Alatna Valley due to a lack of MET stations. The average seasonal wind speeds varied by season (Table 2). Lower elevation sites, such as

Explorer’s Cove and Miers Valley experienced higher seasonal averages in the summer

(November 1st to January 14th), while higher elevation sites, such as Lake Bonney, Taylor

Glacier and Victoria Valley generally had higher wind speeds in the winter (January 25th to October 31st). Lake Bonney, Taylor Glacier and Victoria Valley also had the strongest average daily wind events, with the highest at 30 m s-1 during the winter of 2013 in

Victoria Valley (Figure 4). The frequency of outlier events, likely strong föhn winds, is greatest during the winter for all sites (Figure 4).

5.2 Soluble Solids:

Ion concentrations from the two sequential aeolian material leaches varied greatly between the analytes (Tables 5 and 6). For anions, all sites, with the exception of F6

(Lake Fryxell) and Explorer’s Cove, were below the detection limit for Br- and ranged as 22

high as 19.1 µmol g-1 for Cl-. Coastal samples contained the highest amount of Cl-, while

2- - more inland samples were higher in SO4 . F6 was high in both anions. Cl concentrations were greatest in the upper most collection height (“Top”) for the inland, higher elevation sites. A similar trend was not observed for the two coastal aeolian collectors at Explorer’s

Cove. The concentrations for the second leachate were nearly as high, if not greater than

- 2- - the first leachate for F and SO4 , but the majority of Cl was dissolved in the first leach

(Table 4). This is consistent with sequential leach studies on MDV soils in Taylor Valley

(Toner et al., 2013) where >95% of Cl- was solubilized in the first leach.

Cation concentrations did not vary as greatly as anion concentrations (Table 5).

The lowest cation concentration was K+ in Alatna Valley at 0.047 µmol g-1 and the highest was 23.6 µmol g-1 of Na+ at F6. Na+ is the dominant cation for all sites, followed by Ca2+, then Mg2+ and K+. The down valley locations of F6 and Explorer’s Cove had the highest concentrations of Na+, while inland samples had relatively more Ca2+. The upper collection height samples generally had higher concentrations of Na+ and Mg2+ than lower collection height samples, but similar trends were not observed for K+ and Ca2+.

The 2015 full year samples at Explorer’s Cove, which likely reflect summer chemistry due to the strong influence of coastal breezes, contained the highest amount of Na+, while winter samples from the inland locations contained the highest concentrations. The first leach solubilized the majority of Na+ for all samples, but the second leach for K+, Mg2+, and Ca2+ liberated nearly as much, if not more than the first leach. These findings are consistent with those of Toner et al. (2013) who found that Ca2+ concentrations nearly doubled each time between three sequential leaches.

23

Nutrient concentrations varied greatly between samples, often greater than a

- magnitude in difference (Table 6). The highest NO3 concentration was in the upper most

-1 - sample at Taylor Glacier (0.411 µmol g ) while the lowest NO3 concentration was in the

-1 3- lowest sample height at Explorer’s Cove (0.006 µmol g ). PO4 concentrations were at or below the detection limit for some of the inland and high elevation locations (<0.0001

µmol g-1), but ranged as high as 0.0105 µmol g-1 at the coast. Winter and upper most

- samples had the highest concentrations of NO3 for all sites, while mid to low-level

3- (“Bottom middle”, “Bottom upper”) samples had the highest concentrations of PO4 . The

- 3- majority of NO3 was solubilized during the first leach, but for PO4 , the longer secondary leach contributed the most solute.

5.3 Total and Organic Carbon:

Total nitrogen concentrations from the elemental analysis described in Section 4.6 were all below the instrument detection limit and peaks could not be identified. Total carbon ranged from 0.31 mg g-1 in Alatna Valley to 5.8 mg g-1 in Miers Valley (Table 7).

The organic carbon and inorganic carbon distribution was similar. The mid-level samples contained the greatest amount inorganic carbon for all sites, but the largest amounts of organic carbon were found in the top most samples for all the high elevation sites (Alatna

Valley, Victoria Valley, Taylor Glacier) as well as the 2015 full year collection at East

Lake Bonney and the mid-level collection heights for the lower elevation samples (Figure

5).

24

5.4 Sediment Geochemistry:

Percent major oxides were determined for the 35 samples sent for XRF analysis

(Table 8). The most abundant oxide was SiO2 and ranged from 80.6% in Alatna Valley to

57.7% at F6. The higher elevation sites had more SiO2 than lower elevation sites, while the lower elevation sites had more P2O5, Al2O3 and K2O. Victoria Valley had the highest concentrations of Fe2O3 and CaO at 8.71% and 8.12% respectively in the top most sample. Winter samples from East Lake Bonney and Explorer’s Cove were enriched in

SiO2 compared to full year samples. The samples diverged on TiO2 percentages based on location, where the higher elevation sites of Alatna Valley, Victoria Valley and East Lake

Bonney had lower percentages compared to the lower elevation and coast sites of F6 and

Explorer’s Cove (Figure 6 a.). Aside from other minor differences, the most unique sample location is Alatna Valley, likely due to the high Si content (Figure 6 a.).

The concentrations of 35 trace elements were determined, but all samples were below the detection limit for Be, Ag, As, Bi, Cd, In, and W (Table 9). Ta was below the detection limit only for the higher elevation sampling locations. The lower elevation samples had slightly more U than the higher elevation samples. Alatna and Victoria

Valleys most strongly deviated from the lower elevation sites in their low Ga, Rb, Sr, and

Nb concentrations (7-10 ppm, 25-32 ppm, 48-126 ppm, and 3-6 ppm respectively)

(Figure 6 b.). The top most samples at East Lake Bonney were enriched in Ba in both the winter and the full year samples, while Explorer’s Cove was only enriched in Ba in the full year samples. Explorer’s Cove and F6 were very similar in their concentrations of trace elements.

25

Rare earth elements can be separated into light (Sc, La, Ce, Pr, Nd, Pm, Sm and

Eu) and heavy (Y, Gd, Tb, Dy, Ho, Er, Tm, Yb, and Lu) fractions. The samples diverged by location in the light rare earth elements (LREE) concentrations and converged in the heavy rare earth elements (HREE) concentrations, where the higher elevation and more inland samples were depleted in LREE compared to lower elevation and coastal samples

(Figure 6 c.). For Explorer’s Cove, the lowest samples were more enriched in LREEs and some HREEs than the top most samples (Figure 7 c.). This trend did not exist for the other end of Taylor Valley at East Lake Bonney for the LREEs, and was less pronounced for the HREE fraction (Table 10; Figure 7 c.)

5.5 Minerology:

2θ peaks generated through X-ray diffraction were sharp and symmetrical, suggesting a limited presence of clays, and instead well-defined crystalline mineral phases. All samples had quartz, a suite of feldspars and pyroxenes, micas and amphiboles. There were multiple pyroxenes present, mainly calcium, magnesium, and iron pyroxenes, such as pigeonite, augite, hedenbergite, and potentially clinoenstatite, within individual samples (Figure 8). Most minerals were silicates and non-silicates were not common. There was chlorapatite at F6 and East Lake Bonney, and the iron rich serpentine group mineral cronstedtite. There were some unusual minerals, such as the Bi rich pavonite group cupromakovickyite (Cu4AgPb2Bi9S18) at F6 and Explorer’s Cove, and wavellite (Al3(PO4)2(OH, F) 3· 5H2O) at Victoria Valley, which is usually found in hydrothermal regions. Both common and some unusual minerals were also identified

26

using energy dispersive X-ray spectroscopy (EDXS) on an SEM. Many of the mineral types discussed commonly occur in basalt, andesite, kenyte, and dolerite, all rocks previously catalogued in Antarctica and within the MDV region (Calkin, 1964; Kyle,

1990). Further detail of minerology can be found in Figure 8.

5.6 Grain Size, Composition and Encrustations from Scanning Electron Microscopy:

EDXS was used to examine grain composition and encrustation chemistry. Alatna

Valley had weathered pyroxene, micas, hornblende, and altered K-feldspar (Figure 9 b.).

These minerals were incorporated into pieces of rocks ranging from <0.25 mm to 0.75 mm in size (Figure 9 a.), and were fairly well-sorted and angular. Weathering products, likely illite (Figure 9 e.), were present in grain etch-pits. Many grains were encrusted with salts, predominately gypsum (CaSO4) and glauberite (Na2Ca(SO4)2) (Figure 9 e.), and had Fe staining (Figure 9 d. and f.). On some grains, there were amorphous carbon flakes which could be soot from fossil fuel combustion, or from algal mats.

Similar to Alanta Valley, Victoria Valley samples had carbon on some mineral grains, and though the source is unknown, the lack of structure in the carbon suggests that it might be coal ash or soot (Figure 10 a. and d.). The grains were generally rounded and ranged from <0.25 mm to 1 mm with few salt encrustations of mirabilite (Na2SO4) and glauberite (Figure 10 b.). Not all grains had the same degree of alteration, but some minerals, such as pyroxene, were highly weathered (Figure 10 c.). Quartz grains still generally had intact crystal faces with heavy mineral inclusions, such as chromite (Figure

10 e.) and monazite (Figure 10 f.). Other trace metals (Pb, Sn, Bi, and W) that were not

27

included in the aeolian grains were also observed as surficial flecks, with high concentrations suggesting human produced metallics.

The grains at Explorer’s Cove ranged from <0.1mm to 1mm and were highly rounded and weathered (Figure 11 a.). There was a wide range of minerals, including K- feldspar, hornblende, apatite, biotite and quartz, and some minerals, such as pyroxene, had multiple phases present (pigeonite and diopside). The grains themselves were predominately encrusted with halite (NaCl), but some contained PbBi (Figure 11 e.).

Unlike the high elevation sites of Victoria and Alatna Valleys, Explorer’s Cove had more volcanic mineral grains, including what may be ash, and many of which had been weathered on the outside edges (Figure 11 b and f.), or had more extensive alterations to needles (Figure 11 d.). Explorer’s Cove also had the shells of microorganisms, such as the diatom Hantzschia amphioxys, included with the mineral grains (Figure 11 c.).

F6 was similar in many ways to Explorer’s Cove. The grains ranged from <0.1 mm to 0.75 mm (Figure 12 a.) and were primarily comprised of biotite, serpentine, amphibole and feldspar. Grain surfaces were often complex and contained mixtures of these minerals (Figure 12 b.). There was a high degree of mineral weathering, with some grains starting to weather into clays (Figure 12 c.). Rare earth element inclusions also show weathering (Figure 12 e.). Within the small bits of weathered minerals, there were some diatoms, such as Muelleria meridionalis, present (Figure 12 d.). Unlike Explorer’s

Cove, the amount of salt encrustation at F6 varied greatly. The predominant salts were gypsum and halite (Figure 12 f.). There were also less volcanic minerals present than at

Explorer’s Cove.

28

The grains at East Lake Bonney were well-sorted and ranged from 0.25 mm to

0.75 mm Figure 13 a.). The surfaces of the grains demonstrated great alteration, where minerals such as feldspars, were weathering to needles (Figure 13 b. and c.). These alterations were uniform across the entire grain. East Lake Bonney had large pieces of micas that were slowly weathering at the mineral edges (Figure 13 e.). Other sediment grains, which had many different types of minerals, were weathering each constituent at different rates, determined by different weathering textures (Figure 13 b. and d.) and creating deep etch-pits that were filled with salts, such as halite, mirabilite, glauberite, and gypsum. Similar to F6, some grain surfaces had exposed rare earth element inclusions (Figure 13 f.).

Taylor Glacier had the smallest grain sizes, the largest being 0.5 mm and were fairly angular (Figure 14 a.). The composition of the grains was more mafic and included plagioclase, labradorite, diopside, and epidote. Some of these grains were more complex in composition (Figure 14 d.), where the outer layers of the grains were weathered away to expose a different inner composition (Figure 14 b.). On the highly weathered surfaces

(Figure 14 c.), the predominate salt was gypsum (Figure 14 e.) and unlike other locations, halite was not identified. Some surfaces also included zircons, monazite and ilmenite

(Figure 14 f.).

Miers Valley had the highest occurrence of volcanics compared to all other locations Figure 15 c.). The grains were a mixture of round and angular, and ranged from

0.1 mm to 1 mm Figure 15 a.). The degree of weathering varied as well, as seen in Figure

15 b., where a highly weathered mica is shown next to a poorly weathered piece of

29

augite. This suggests that some materials are entrained and transported longer than other materials, and therefore further exposed to subaerial processes. On some mineral surfaces, there was evidence of goethite and ilmenite (Figure 15 e. and f.). Some of the volcanic materials were weathered to expose the different inclusions (Figure 15 d.), and had many types of material that were weathering at different rates. Overall, the chemistry of grains in Miers Valley suggest an overwhelming dolerite source.

30

6. Discussion:

Aeolian processes play an important role in the transport of both geological and biological materials globally, on the biogeochemistry of ecosystems, and in the evolution of the landscape (Vitousek et al., 1997; Jickells et al., 2005; Okin et al., 2006). Previous studies on aeolian dust transport in the MDV have shown that wind is a major distributor of organic matter and nutrients, and is a modifier of biodiversity that contributes to aquatic ecosystem vitality (Šabacká et al., 2012; Deuerling et al., 2014). Aeolian deposition of potentially soluble nutrients associated with particulate matter may be important in supporting biological communities along the lake edges, where deposition to the surface of lakes is believed to be a source of organic carbon to deeper waters (Adams et al., 1998; Fristen et al., 2000). In ice-covered environments, such as lake ice and glacier surfaces, portions of the aeolian material can be solubilized, adding nutrients and other solutes to the aquatic ecosystem (Lyons et al., 2003; Porazinska et al., 2004;

Bagshaw et al., 2007; Deuerling et al., 2014).

Aeolian deposition onto annual sea ice can also play an important role in the biogeochemical dynamics of the Ross Sea. Material that is deposited on sea ice in

McMurdo Sound is eventually added to the water, with the insoluble matter being incorporated into sediments and the soluble matter being incorporated into the water column during summer melting. Because the is a location where bioavailable iron may be the most limiting nutrient to pelagic ecosystems, the terrigenous 31

derived iron released from sea ice and icebergs can aid in phytoplankton blooms and increased primary productivity (Martin and Fitzwater, 1988; Coale et al., 1996; Sedwick et al., 2000; Raiswell et al., 2008; Boyd et al., 2010), which in turn may draw down CO2 from the atmosphere, leading to burial of carbon in the sediments. Recent work in

McMurdo Sound suggests that the input of iron deposited on sea ice from the MDV may be extremely important (Bhattachan et al., 2015; Winton et al., 2016). Material found on sea, lake and glacier ice, as well as in near surface sediment traps, has been analyzed for its major oxide concentrations, but there is still uncertainty as to the concentrations of soluble constituents that are also added to these systems (Guerzoni et al., 1992; Delmonte et al., 2010; Bottos et al., 2014; Deuerling et al., 2014).

An understanding of how surface environments are being affected by external forces, such as particle transport and deposition by winds, will aid in understanding the long-term dynamics and the connectivity between various landscape units of the MDV.

There have been no studies which have analyzed the geochemistry of aeolian material collected in elevated sediment traps above 30 cm from the surface. These data are needed to understand soluble and solid material transport and its geochemical variations over the entire MDV landscape.

6.1.1 Aeolian Composition- Water Soluble Component:

Deuerling et al. (2014) analyzed aeolian material from lake and glacier ice (11 sites), elevated sediment traps at 30 cm (3 sites), and aeolian landforms (4 sites). They found the highest concentrations of total dissolved solids (TDS) in the aeolian material

32

collected in sediment traps compared to the ice surfaces. They interpreted this to mean that this aeolian trap material had not been in previous contact with the hydrological system (i.e. liquid water), while the material from the ice surfaces had been in previous contact with water and had lost some of its original soluble component. Their sediment

+ + trap water leaches were enriched in Na and K , while aeolian material collected from landforms and ice had relatively higher Ca2+ and Mg2+ concentrations (Deuerling et al.,

2014). The analysis of aeolian material presented here similarly demonstrates high

+ + concentrations of water leachable Na and K , suggesting that these cations are associated with the most soluble salts and are the first to solubilize with wetting (Figure 16 a.).

+ + Toner et al. (2013) found that the majority of Na and K from Taylor Valley soils was solubilized in the initial wetting, regardless of the water to solid ratio used to solubilize

2+ 2+ 2- the salts. These results suggest that Ca and Mg are mostly associated with SO4 and

2- - - CO3 salts, which are less soluble than Cl and NO3 salts. The cation distribution of sediment on aeolian landforms from Deuerling et al. (2014) was similar to their sediment trap material when compared to the ice surface materials. Particles on ice surfaces have the most potential for contact with liquid water, while the low precipitation and high sublimation rates of fallen precipitation on landform surfaces probably results in only minor wetting, which may explain why the landforms and entrained material are relatively more enriched in Na+ and K+ compared to these ice surfaces.

33

6.1.2 Spatial and Seasonal Variability:

Though Na+ and K+ are the major soluble cation constituents for all samples

(Figure 16 a.), there is still spatial variability among samples dependent on location, specifically with distance from McMurdo Sound, and with elevation. The samples from

Explorer’s Cove have the highest percentages of Na+ and K+ at nearly 100%, while F6 is at about 80%, East Lake Bonney at 60%, and Taylor Glacier at 70% (Figure 16 a.). This decreasing pattern generally follows a horizontal transect from the coast inland towards the . Miers, Victoria, and Alatna Valley samples were similar in their cation distribution, with minor differences in Ca2+ and Mg2+ contributions, and resemble the East Lake Bonney inland samples in composition more than the coastal samples (Figure 16 a.). When compared to winter samples, the summer and full year samples in Taylor Valley were slightly more enriched in Ca2+ (Figure 16 b.). This may suggest that there is seasonal variability in the source of materials and the difference in geochemistry reflects changes from a predominately west to east wind direction. Fortner et al. (2005) observed similar effects on the surface chemistry of where western proglacial streams had higher Ca2+:Cl- ratios. There are not enough sampling locations (i.e. up versus down valley) to conclude whether such trends exist in the other valleys.

The distributions of anions varied much more greatly than the cations, particularly

- - in Taylor Valley. The major anion alternated between Cl and HCO3 at all locations and

2- - SO4 had the lowest percentages. Even at the coastal site of Explorer’s Cove, Cl varied from between 10% to nearly 70% of the major anions (Figure 16 a.). F6 had a similar

34

range of contributions. The more distal and higher elevation sites had similar distributions between East Lake Bonney and Taylor Glacier, as well as between Miers,

- Victoria, and Alatna Valleys. In all of these locations, HCO3 is the major anion, likely associated with calcium carbonate, as this mineral is found in all MDV soils (Foley et al.,

2006; Toner et al., 2013). Deuerling et al. (2014) and Bisson et al. (2015) saw a similar variable distribution of major anions associated with aeolian material and soils in Taylor

2- Valley, but attributed higher concentrations of SO4 , particularly in the Bonney Basin, to more abundant gypsum. In general, these studies agreed with the earlier work on MDV soils by Keys and Williams (1981). Though there was gypsum observed in grain etch-pits from SEM imaging (Figure 13), it was not the only salt present at East Lake Bonney, and

- 2- the leachate from these samples suggests that Cl is the dominate anion, not SO4 .

The wide range of anion contributions is likely due to several factors: 1) changes in wind direction and hence material source throughout the year, 2) differences in the chemistry of material at different collection heights above the surface, and 3) the

“dryness” of the local/regional provenance which is usually related to the age of the surface (Lyons et al., 2016). At Explorer’s Cove, samples have a greater percentage of

- HCO3 in the winter compared to the full year samples. On the other end of Taylor

Valley, the seasonal variability at East Lake Bonney was the opposite with winter

- samples having less HCO3 compared to full year samples (Figure 16 b.). This is likely due to stronger up valley winds near Explorer’s Cove during summer months (Nylen et al., 2004), which have a more sea aerosol chemical signature (Keys and Williams, 1981).

35

Top most samples (~100 cm above the surface) also tended to deviate from this more general locational anion clustering. F6 Top is approximately 90% Cl-, where all other lower collection heights are less than 80%. Top samples from East Lake Bonney and Victoria Valley are nearly identical in their anion distributions, and have a similar

2- percentage (~20%) of SO4 as Alatna Valley Top (Figure 16 a.). Though the variations are not large, these are likely indicative of different types, either size or source, of aeolian material in Taylor Valley depending on height above the ground. Lancaster (2002) found that the grain size of aeolian sediment transported at 100 cm above the surface varied between 10.5% to 99.1% sand. Deuerling (2010) found that all aeolian collections at 30 cm above the surface were dominated by sand-sized or even larger particles, and some even had as high as 60% gravel by weight. The difference in grain size between 30 cm and 100 cm in these studies indicates a different type of material at these collection heights, and may help to explain differences in soluble chemistry between top and bottom samples in this study. Because it was not possible to determine the grain size on these samples, it is unclear if the differences are just reflected in the sizes of the materials transported or are real differences in sources.

While samples from the coastal and low elevation sites of Explorer’s Cove and F6

- were more enriched in HCO3 , the distant and higher elevation samples were more

- alkaline earth enriched (Figure 17). The higher concentrations of HCO3 corroborate a higher degree of carbonate source material. The Ca2+ and Mg2+ concentrations measured in the water leaches are likely due to the dissolution of Mg and Ca-bearing carbonate phases (Keys and Williams, 1981). Deuerling et al. (2014) attributed increases in Mg2+

36

content to the higher occurrence of Mg-bearing salts, as Keys and Williams (1981) had previously seen similar trends in Taylor Valley soils, which they attributed to enhanced

Mg-silicate weathering with distance from the coast. However, this trend of Mg increasing inland was only observed in Taylor Valley (Table 5), and not for Victoria and

Alatna Valleys.

A comparison of Na+ and Cl- concentrations in F6 and Explorer’s Cove samples suggests that the high concentrations of soluble Na+ and Cl- are from NaCl dissolution

+ 2+ - - (Figure 17). The non-coastal sites have more Na and Ca compared to Cl and HCO3 , likely from sodium and calcium sulfate salts, such as gypsum, mirabilite and glauberite

(Figures 9, 10, 13, and 15) (Bisson et al, 2015). Nearly all samples, regardless of location, have Na+ concentrations above the NaCl dissolution line (Figure 18), suggesting a relative enrichment in Na+ relative to Cl-. This relationship was also observed for aeolian samples collected at 30 cm above the surface by Deuerling et al. (2014). Since F6 and

Explorer’s Cove plot directly on the combined halite and calcium carbonate dissolution line (Figure 19), it is possible that some of the Na+ is from the dissolution of sodium bicarbonate (Na-HCO3), which Toner et al. (2013) believes could form in Eastern Taylor

Valley soils (i.e. Fryxell and Hoare basins) due to the relatively higher precipitation rates.

In these calcareous soils, sodium bicarbonate forms through the dissolution of halite, derived initially from marine aerosols and calcium carbonate, and then the exchange of cations (equation 6.1.1) (Kelley, 1951). Once Na+ displaces Ca2+, this frees Ca2+ to the

- solution and it, along with Cl , forms CaCl2 rich brines, which have been observed

37

throughout the MDV (Wilson, 1979; Lyons and Mayewski, 1993; Levy et al., 2012;

Dickson et al., 2013).

2NaCl+ (Ca,Mg)X2 = (Ca,Mg)Cl2+ 2NaX (equation 6.1.1)

*where X= ion-exchange site on a clay mineral

6.1.3 Down Valley Transport:

The presence of Na-HCO3 in aeolian material at F6 and Explorer’s Cove suggests a prominent local source of materials, with more minor influences from summer coastal breezes due to their proximity to the coast. The streams in the Fryxell Basin have high

2+ - concentrations of Na and HCO3 , which if cyroconcentrated, the subsequent brine

would form Na-CO3/HCO3 salts (Lyons et al., 1998; 2005). Thus, transport of material containing Na-HCO3 is likely derived primarily from the Fryxell Basin (Toner et al.,

2013), perhaps either suggesting down-valley material transport to the coastal site of

2- Explorer’s Cove or production of Na-HCO3 in situ there. Ratios of leachable SO4 and

- - NO3 to Cl also support this assertion, particularly when compared to local surface soils

(Toner et al., 2013) (Figures 20 and 21). The water soluble salt distribution of soils in the

Fryxell Basin is most similar to the aeolian material collected at East Lake Bonney.

Aeolian soluble material from F6 has a chemical composition between that of valley mouth or eastern most soils and seawater, which is in agreement with observed bimodal wind transport near the coast (Lancaster, 2002; Doran et al., 2002).

Low-level up valley winds near the coast may be impeded by Coral Ridge, a 78 masl saddle between F6 and Explorer’s Cove (Figure 22), which may explain why

38

aeolian material from Explorer’s Cove is the most similar to local soils, but still widely

2- varies in SO4 concentrations. Coastal samples are likely a mixture of strong down valley wind signatures, which can overprint prominent local sources (Nylen et al., 2004;

Lancaster et al., 2010), and a sea signature that remains coastal due to the transport restriction from Coral Ridge. The input of marine air, which contains dimethyl sulfide

2- (DMS) and its oxidation products, such as SO4 , varies depending on the time of year and sea ice extent (Welch et al., 1993). The composition of soils near the valley mouth is most similar to the full year, mid-level collection height samples from Explorer’s Cove, which supports the premise that locally derived resuspension of soil, likely due to up valley wind impediment from Coral Ridge and a lack of easterly wind transport beyond this geomorphologic feature, controls aeolian deposition there. Fountain et al. (1999) have previously suggested that the Nussbaum Riegel in central Taylor Valley acts as a barrier to the westward transport of marine air. It is suggested here that Coral Ridge also may similarly influence the transport of aeolian material westward into Taylor Valley, at least at low elevations. Samples from the inland locations in Taylor Valley, particularly the winter, top most collection height samples, suggest a more down-valley transport trend, where the aeolian sediment collections do not have the same soluble chemistry as nearby soils (Figure 21).

The highest elevation sites (Alatna Valley, Victoria Valley, and Taylor Glacier) can be easily distinguished from the lower elevation locations in Taylor Valley, and

2- - follow an almost linear increasing trend in both SO4 and NO3 from lowest to highest elevation (Figure 20). This is also similar to what has been observed in water soluble salts

39

in soils from throughout the Transantarctic Mountains, where the highest elevation soils

- are enriched in NO3 , while the soils close to the ocean and at the lowest elevation are enriched in Cl- (Keys and Williams, 1981; Bockheim and McLeod, 2007; Lyons et al.,

- 2016). Isotopic analysis suggests that NO3 is deposited through both tropospheric and stratospheric sources, and these soluble nitrate salts accumulate on the surface because the high elevations do not experience any significant liquid water (Lyons et al., 2016).

Interestingly, the Dry Valleys exhibit similar phosphorus and nitrogen deficiencies related to landscape age as observed in the humid tropics (Vitousek et al.,

1997; Chadwick et al., 1999;), i.e. the “younger” landscapes in Taylor Valley, such as the eastern most Ross Drift tills, are more N-limited while the “older” landscapes, such as the western most Bonney Drift tills, are more P-limited (Barrett et al., 2007). In general, the water soluble concentrations of nutrients (N and P) in the lower elevation aeolian material are at least two orders of magnitude lower than soluble major ion concentrations.

Thus, the input of aeolian material with different water soluble N:P ratios could have important ramifications on the structure and function of both the aquatic and soil ecosystems (Barrett et al., 2007). Top most samples for F6 and Miers Valley were mostly

N-enriched, while Explorer’s Cove Top samples were the most P-enriched (Figure 23 a.).

This observation further supports a down valley transport pattern from Taylor Glacier towards F6, as it reflects the observed differences in Bonney Basin versus Fryxell Basin soils (Barrett et al., 2007). As previously discussed, any wind-blown P derived from the relatively young soils of Explorer’s Cove likely remains coastal due to potential blockage of surface winds by Coral Ridge (Figure 22). The bottom most samples from F6 and

40

Miers Valley, which are intermediate between the coastal and further inland sites, and are near young soils with high P concentrations, are relatively more P- enriched than top and middle samples, and thus are likely sourced from different directions/locations. The aeolian material at East Lake Bonney differentiated based on season by its N:P ratio

(Figure 23 b.), where winter samples were above the Redfield stoichiometric ratio of

N:P= 16:1 line, implying N-enrichment, while full year samples were below the Redfield line, implying P-enrichment. This indicates that at certain times of high on-shore winds, some finer grained material from the P-enriched soils in the Fryxell Basin is transported into the Bonney Basin, or that there is another higher elevation source of P-enriched material being transported. Little information is available on P concentrations in the higher elevation Transantarctic Mountains soils, but what do exist would negate the latter idea. Therefore, the source of these higher P:N ratio materials is an enigma at this time.

- Though down valley transported NO3 could be important for young soil ecosystems, Deuerling et al. (2014) calculated that aeolian fluxes of water soluble nutrients at 30 cm above the surface are 6 times less than the flux of nutrients to aquatic systems from MDV soils (Barrett et al., 2007). Water soluble N and P data from this study and sediment flux data from Lancaster (2002) and Šabacká et al. (2012) were used to estimate N and P fluxes at 30 and 100 cm above the surface (Table 11). The 100 cm N fluxes were about 10 times lower than values from Deuerling et al. (2014), while P fluxes were similar, suggesting that much of the N may be associated with smaller grain size material. Overall, soluble N and P aeolian fluxes are between 100 and 1000 times smaller than the fluxes of nutrients from the soils, as calculated by Barrett et al. (2007). Due to

41

their lower concentrations compared to Taylor Valley soils, the input of aeolian materials to streams and lake edges probably has little ecological consequences. However, as earlier work on glacier and lake ice surfaces (Priscu, 1995; Priscu et al., 1999; Fountain et al., 2004), as well as cryoconites (Porazinska et al., 2004; Bagshaw et al., 2007) have concluded, if aeolian material deposited onto these surfaces later comes into contact with liquid water, this material has the potential to provide N and P to the cryoconite, stream, and lake surface ecosystems.

6.2.1 Aeolian Composition- Solid Component:

Although the soluble component of aeolian material provides important information on the role this material plays in adding solutes to aquatic systems, these analyses do not serve as a primary identifier of the specific sources of aeolian material, sometimes referred to as provenance. Previous studies have successfully used Chemical

Index of Alteration (CIA) calculations (Moreno et al., 2006; Deuerling, 2010), major oxide geochemical variation (Moreno et al., 2006; Ferrat et al., 2011), isotopic measurements (Delmonte et al., 2010; Deuerling et al., 2014; Winton et al., 2014; 2016), and rare earth elemental (REE) signatures (Muhs et al., 2010; Wei et al., 2017) to determine the provenance of aeolian material and thereby infer sources and pathways of dust transport. A combination of these geochemical tools have been utilized to investigate the source of aeolian material in the MDV.

An indicator of the degree of chemical weathering sediment has experienced is the Chemical Index of Alteration (CIA), calculated via equation 6.2.1. The CIA was

42

developed to measure the extent of alteration of primary igneous minerals, such as feldspars, to clays, such as kaolinite (Nesbitt and Young, 1982), but has been used more broadly to quantify silicate weathering (Deuerling, 2010; Price and Velbel, 2003; Shao and Yang, 2012). As chemical weathering occurs, alkaline earth and alkali metals are preferentially solubilized from the minerals present relative to Al, thereby increasing the

CIA value. As weathering continues, the denominator approaches the numerator, further increasing the CIA value towards 100%.

퐴푙2푂3 CIA= * 100 (equation 6.2.1) 퐴푙2푂3 + 퐶푎푂 + 푁푎2푂 + 퐾2푂

CIA values were calculated for aeolian samples in each valley and elevational profiles were developed between the bottom (~30 cm) and top (~100 cm) samples (Figure

24). These profiles were compared to average CIA values of the local rock types that are potential sources of this aeolian material, specifically McMurdo Volcanics (Cooper et al.,

2007), Ferrar Dolerite (Antonini et al., 1999) and Beacon Sandstone (Roser and Pyne,

1981). The CIAs for the average sediments on ice covered surfaces (Deuerling, 2010) and dust from Patagonia (Gaiero et al., 2003), of which has previously been identified in East

Antarctic ice cores (Basile et al., 1997), were also included (Table 3). A major conclusion from this analysis is that aeolian material from all valleys is poorly weathered and varies little with respect to height above surface and between valleys. Values ranged from

51.2% to 53.4% in Taylor Valley and 43.6% to 54.7% in the other valleys (Table 3), with

Victoria Valley being the least weathered at 43.6%. For reference, unweathered rocks, such as basalts, have low CIAs (<50%), while highly weathered rocks, such as sandstone, have higher CIAs (>60%) (Nesbitt and Young, 1982). The McMurdo Volcanics and

43

Ferrar Dolerite had CIA values of 46.9% and 55.4% respectively. Deuerling (2010) found

CIA values ranging from 44% to 57%, which were similar to the values reported here, and strongly indicates limited chemical weathering of the total material, despite evidence in some SEM images of a number of individual particles that have undergone alteration and weathering (Figures 9-15).

Compared to the potential rock sources of MDV material listed above, the CIAs of aeolian sediment were an intermediate between the poorly weathered end-member

McMurdo Volcanics and the Beacon Sandstone, and were most similar to Ferrar Dolerite

(Figure 24). Patagonian dust had higher CIA values than MDV aeolian material, and even sediment on lake ice (Deuerling, 2010) which as discussed in Section 6.1.1, can experience chemical weathering during seasonal ice cover melting in the austral summer.

To summarize, based on the placement of elevational CIA profiles in relation to other sediments and rocks types found in Antarctica, MDV aeolian CIA values suggest minor silicate chemical weathering of material, and a mixture of local source rocks.

Oxide variation diagrams plotted against SiO2 indicate the composition and mixing of the major rock types in the MDV (Figures 24 and 25). Na2O and K2O percentages remained static or increased slightly with increasing SiO2 (Figure 25), which along with the other trends of decreasing Fe2O3, CaO, TiO2 and MgO, indicates a strong alkali nature of the aeolian sediment (Kyle et al., 1970). Deuerling (2010) found a more mafic composition in the silt sized fractions and more felsic compositions in the sand fraction. The percentages of Na2O and K2O are similar between the McMurdo Volcanics and aeolian material (Figure 25), suggesting that the more mafic components of Ferrar

44

Dolerite, have been lost or were never initially a major component of the material. All samples had higher percentages of SiO2 compared to McMurdo Volcanics and Ferrar

Dolerite, which suggests another component to the material- either SiO2 material from the Beacon Supergroup, or material from the lower elevation crystalline basement rocks.

Weathered pyroxenes, likely augite and pigeonite, as identified through XRD (Figure 8), were observed in SEM images as elongated needles (Figures 9 and 10), and corroborate some Si loss.

The geochemistry of aeolian samples was compared to the geochemistry of

McMurdo Volcanics (Cooper et al., 2007), Ferrar Dolerite (Antonini et al., 1999), granitoids, (Smillie, 1992) and sandstone (Bhatia, 1983), with respect to SiO2 and Na2O +

K2O (Figure 26). Potential mixing lines were determined between possible source materials, where the chemistry of any samples that fall on or between the lines can be explained by a mixture of these four rock types. All samples from Miers and Taylor

Valleys fall between the rock mixing lines, suggesting an oxide composition that is a mixture of all four rock types and confirms the previously presented idea that the sandstone and possibly the crystalline basement rocks are contributing to the aeolian composition. Alatna and Victoria Valleys fall just outside of the mixing lines. These data suggest either that an additional unknown source of material exists at these high elevation sites, or there are regional compositional changes in the rocks that the model has not considered.

45

6.2.2 Geographic Provenance:

Rare earth elements (REE) and radiogenic isotopes, such as strontium and neodymium isotopes, can be used to determine sediment/dust sources. Strontium and

Neodymium isotopic concentrations are generally expressed as a set of ratios, specifically

87Sr/86Sr and 143Nd/144Nd, where 143Nd/144Nd is typically expressed normalized to the chondrite value as ƐNd. These isotopes fractionate based on crust-mantle differentiation age, the cycling of crustal materials and weathering, and therefore vary based on region

(Faure, 1986). Investigations on how these isotopes vary together can give important insight to sediment provenance (DePaolo, 1988).

Aeolian material on Taylor Glacier from Deuerling et al. (2014b) had a ƐNd and

87Sr/86Sr signature that was most similar to MDV rocks, but isotopic signatures of the material from the central and coastal locations in Taylor Valley do overlap with South

American, and New Zealand source signatures, as well as the MDV signature (Delmonte et al., 2004) (Figure 27). Using the Occam’s razor approach, Deuerling et al. (2014b) concluded that even with the overlap of potential sources, the material deposited on glaciers is locally derived. Aeolian sediment deposited on the sea ice of McMurdo Sound

(Winton et al., 2014; 2016) has ƐNd and 87Sr/86Sr values indicating that this material is different than material on glacier surfaces in Taylor Valley (Deuerling et al., 2014b)

(Figure 28). Aeolian material from these two studies was compared to isotopic signatures from MDV source rocks and soils (Delmonte et al., 2010) (Figure 28). The McMurdo

Sound material of Winton et al. (2014; 2016) was similar to the soils and phonolite lava of the McMurdo Volcanics, while the glacier material of Deuerling et al. (2014b) was

46

more similar to the tills, Ferrar Dolerite, and Granite Harbor Intrusives, which are the local basement rocks in much of the MDV. Although the size fraction of aeolian dust was similar in both studies, the isotopic signatures obviously suggest different sources. The

McMurdo Sound material has a primarily positive ƐNd signature, which when compared to potential dust source areas from Delmonte et al. (2004), is similar to both South

American and Dry Valley signatures (Figure 27). Recent work on southern high-latitude dust transport (Neff and Bertler, 2015), found that aeolian transport of material from

Patagonia and New Zealand is highly efficient, where >8% of trajectories reach

Antarctica after 8 days. Trajectories from Australia are much less efficient with ~2% arriving in Antarctica after 10 days. These works on trajectories and ƐNd and 87Sr/86Sr isotopic signatures suggest that some of the fine grained material in the Dry Valleys could be sourced from Patagonia and/or New Zealand. However, aeolian samples from

MDV (Deuerling et al., 2014) and Terra Nova Bay to the north (Guerzoni et al., 1992) suggest that the primary sources are local.

While ƐNd and 87Sr/86Sr signatures have not yet been measured for the aeolian material in this study, REE elemental concentrations were determined. Select ratios of these elements can give similar insight to provenance as isotopic measurements (Bhatia,

1985). The fourteen REEs have low solubility and remain primarily in the solid phase during transport, therefore maintaining the REE distribution of the source material (Piper and Bao, 2013). REE ratios can determine the types of minerals present and can be used as a tool to compare and contrast geologic materials to determine the nature of the source material. Generally, light and heavy rare earth element ratios (L/HREE) can indicate the

47

total enrichment of light over heavier REE fractions. Similarly, La/Yb ratios normalized to the upper continental crust, (La/Yb)N, (Rudnick and Gao, 2003), can indicate the enrichment of the lightest over heaviest REEs. (La/Sm)N represents the differential enrichment of LREEs, while (Gd/Yb)N ratios indicate the degree of relative depletion of

* HREEs (Bhatia, 1985). (Eu/Eu )N, calculated via equation 6.2.2, indicates the degree of

REE fractionation since Eu* represents the interpolated Eu concentrations with no enrichment or depletion (Bhatia, 1985).

* 1/2 Eu/Eu = (EuN/(SmN * GdN) ) (equation 6.2.2)

REE ratios from the MDV aeolian sediments were compared to published values of McMurdo Volcanics (Cooper et al., 2007), Ferrar Dolerite (Antonini et al., 1999),

USGS standard granite, soil from Australia (Martin and McCulloch 1999; McLennan et al., 1983; Diatloff et al., 1996), soil and aeolian dust from Patagonia (Gaiero et al., 2004;

Gaiero et al., 2003), dust from New Zealand (Marx et al., 2005), and Taylor Valley stream sediments (Dowling et al., in review) (Figure 29). The dust from areas outside of

Antarctica was included because, as discussed previously, wind trajectories from these locations suggest that fine-grained material from these regions can potentially reach

Antarctica (Li et al., 2008; Neff and Bertler, 2015).

Using the REE data, MDV aeolian samples can be separated based on elevation and proximity to the coast. Victoria and Alatna Valley samples were relatively more enriched in the heaviest REE (Yb) and had the lowest Eu/Eu* ratio (Figure 29).

Explorer’s Cove, F6, and Miers Valley all clustered with a relatively higher concentration of LREEs and a greater Eu/Eu* signature. East Lake Bonney plotted between these two

48

end members and was most similar to dust (and therefore top soil) from Patagonia

(Figure 29). Alatna and Victoria Valleys were the most similar to each other and to a few soil samples from Australia, and with the exception of the upper most samples, Miers

Valley was the most similar to F6. Aeolian sediment in the MDV is not similar to local stream sediments from the same basin (Dowling et al., in review) (Figure 29 a. and b.).

Stream sediments from the Bonney Basin were most similar to aeolian material at F6, while stream sediments from the Fryxell Basin were most similar to aeolian material at

Explorer’s Cove. This supports the down valley aeolian sediment transport trend discussed in Section 6.1.3, but the chemistry of Bonney Basin soils only partially explains the REE chemistry variability seen at F6 (Figure 29 a.- c.). The difference in

REE ratios between MDV aeolian material and stream sediments is probably not due to alterations from chemical weathering. Although the Eu/Eu* ratios in the stream sediments are variable, potentially suggesting some weathering of plagioclase with Eu inclusions and therefore resulting in Eu loss, variations in L/HREE are small and do not suggest preferential loss of either REE fraction (Figure 29 d. and e.).

Using the REE data, potential mixing lines were estimated between the local rocks, where the REE chemistry of any samples that plot on or between the lines can be explained by a mixture of MCM Volcanics, Ferrar Dolerite, and/or granites (no data for

MDV granites were available, therefore the value is from a USGS standard). Most samples plotted within these boundaries in terms of LREE geochemistry, but ratios for

HREEs, such as (Gd/Yb)N and (La/Yb)N, no locations fell fully within or on the mixing lines (Figure 29 a.- c.). Samples from Alatna and Victoria Valleys, and top most samples

49

from Miers Valley plotted outside the mixing areas for all ratios, perhaps suggesting another unknown source of material.

In conclusion, analysis of REE signatures, bulk geochemistry, and CIA values of aeolian material in the MDV suggests that this material is derived primarily from local rock sources, but deviations from MDV rock geochemistry may also suggest input from other, yet unknown and perhaps even far traveled, sources.

6.3.1 Analysis of Variance (ANOVA):

One-way Analysis of Variance (ANOVA) calculations were performed for all samples in RStudio using Fisher’s Least Significant Difference (LSD) method, where the variables were location, year, and height of collection on the BSNE trap. One-way

ANOVA compares one parameter between two or more groups to measure sources of variation in the data and to compare the relative size of groups. It is more versatile than a t-test, as it determines whether variability in a dataset is between groups or within. All the geochemical data were analyzed against location, site, and sampling height to determine which of these groups was responsible for any variance in the geochemistry. These results are detailed in Appendix C.

2+ 2- With the exception of soluble Mg and SO4 , all sites showed significant variation in soluble anion and cation concentrations between different locations and different years (Table 15). All major oxides displayed strong significant variability

(>99% confidence) in location and most had variability based on year, except P2O5, TiO2 and CaO (Table 16). Variations in trace element chemistry could not be explained by

50

either location, year, or sampling height as confidently as with the other analytes. Pb, Sb and Sn had no statistically significant variance within or between these three variables.

Several other trace elements only had variation in chemistry based on location (Table 17).

Visually, differences in ion chemistry based on height were apparent in this study, but the

LSD method, though one of the least conservative, was not sensitive enough to identify such differences between chemistry and height for any analyte. The ANOVA indicated that there is great variability in this data set that can be attributed to variations in time and with location, but not in terms of sample collection height.

51

7. Conclusions:

1. Aeolian material collected in elevated sediment traps in Taylor Valley is generally

transported down valley from Taylor Glacier towards the coast. Though there are

summer up valley winds, they mainly affect coastal areas and do not appear to

transport enough material up valley to significantly alter the composition of

aeolian material at inland locations. There is potentially inter-valley mixing

between the coastal sites of Explorer’s Cove and Miers Valley and the high

elevation sites of Alatna and Victoria Valleys based on overlaps in major oxide

and REE geochemistry between the sites.

2. The geochemistry of soluble salts and nutrients in aeolian material generally

reflects the complicated geologic and climate history of the Ross Sea region, as

well as the surficial soil ages of the valleys where some of the soluble fraction is

derived.

3. The mineralogy of the aeolian materials generally reflects the types of rocks,

having both mafic and felsic compositions, found exposed in the MDV. CIA

values suggest that the bulk material has not been chemically weathered, but SEM

images do show alteration for some samples.

4. In general, the geochemical analysis of the aeolian material supports the notion

that the majority of material is derived from local sources, but it also suggests that

there may be an additional, possibly distant, source that cannot be accounted for.

52

5. There is spatial and temporal variability in soluble solids, mineralogy, and bulk

geochemistry between locations, with height above the ground, and seasonally,

and merits further investigation.

53

8. Future Work:

Though this work provides detail to supplement previous studies on the geochemistry of aeolian material in the McMurdo Dry Valleys, there are still many questions which can be avenues for future research:

1. How does the geochemistry of material vary on smaller temporal timescales?

Specifically, how does the geochemistry of material transported during strong

föhn events compare to material from weaker coastal events? Are there

differences in source?

2. XRD cannot distinguish different clay minerals without extensive sample

preparation, therefore clay analyses were not performed in this study. Since some

chemical weathering was observed on individual particles in SEM images, what

clays are present in the aeolian material, and can the clay mineral composition aid

in our understanding of provenance and transport of the material?

3. Does the strong down valley transport and its potential impact on nutrient

limitation trends observed in Taylor Valley exist in the other valleys, such as

Miers Valley?

4. Can a physically based depositional model be developed that determines how

surface chemistry is influenced by source rocks and winds?

5. Do isotopic signatures of aeolian material confirm or refute conclusions regarding

provenance?

54

6. An initial Principal Component Analysis (PCA) test was run on the full dataset

for Cl-, and the samples clustered based on sampling height, which suggests this

test may be sensitive enough for these elevational geochemical investigations. As

a complex dataset, what other variations can be identified through detailed PCA,

Tukey and other statistical analyses?

55

9. References:

Adams E.E., Priscu J.C., Fritsen C.H., Smith S.R., Brackman S.L. 1998. Permanent ice covers of the McMurdo Dry Valleys lakes, Antarctica: bubble formation and metamorphism. In Priscu JC (ed.), Ecosystem Dynamics in a Polar Desert: The McMurdo Dry Valleys, Antarctica. Antarctic Research Series 72. American Geophysical Union: 281–295. Angino, E.E., Turner, M.D., Zeller, E.J. 1962. Reconnaissance geology of lower Taylor Valley, , Antarctica. Geological Society of America Bulletin 73: 1553- 1562. Antonini, P., Piccirillo, E.M., Petrini, R., Civetta, L., D’Antonio, M., Orsi, G. 1999. Enriched mantle - Dupal signature in the genesis of the Jurassic Ferrar tholeiites from Prince Albert Mountains (Victoria Land, Antarctica). Contributions of Mineral Petrology 136: 1-19. Atkins, C.B. and Dunbar, G.B. 2009. Aeolian sediment flux from sea ice into Southern McMurdo Sound, Antarctica. Global and Planetary Change 69: 133-141. Bagshaw, E. A., Tranter, M., Fountain, A. G., Welch, K. A, Basagic, H, Lyons,W. B. 2007. Biogeochemical evolution of cryoconite holes on Canada Glacier, Taylor Valley, Antarctica, Journal of Geophysical Research 112: G04S35, doi:10.1029/2007JG000442. Barrett, J.E., Virginia, R.A., Lyons, W.B., McKnight, D.M., Priscu, J.C., Doran, P.T., Fountain, A.G., Wall, D.H., Moorhead, D.L. 2007. Biogeochemical stoichiometry of Antarctic Dry Valley ecosystems. Journal of Geophysical Research-Biogeosciences 112: G011010. Basile, I., Grousset F. E., Revel M., Petit J. R., Biscaye P. E., Barkov N. I. 1997. Patagonian origin of glacial dust deposited in East Antarctica (Vostok and ) during glacial stages 2, 4 and 6. Earth and Planetary Science Letters 146: 573-589. Bhatia, M. R. 1983. Plate tectonics and geochemical composition of sandstones. Journal of Geology 91(6): 611–627. Bhatia., M.R. 1985. Rare earth element geochemistry of Australian Paleozoic graywackes and mudrocks: Provenance of tectonic control. Sedimentary Geology 45: 97-113. Bhattachan, A., Wang, L., Miller, M. F., Licht, K. J., D’Odorico, P. 2015. Antarctica’s Dry Valleys: A potential source of soluble iron to the Southern Ocean?, Geophysical Research Letters 42: 1912–1918, doi:10.1002/2015GL063419.

56

Bisson, K. M. Welch, K. A., Welch, S. A. Sheets, J. M., Lyons, W. B., Levy, J. S., Fountain, A. G. 2015 Patterns and Processes of Salt Efflorescences in the McMurdo region, Antarctica. Arctic, Antarctic, and Alpine Research 47(3): 407-425. Bockheim, J.G. and McLeod, M. 2007. Soil Distribution in the McMurdo Dry Valleys, Antarctica. Geoderma 144: 43-49. Bottos, E.M., Woo, A.C., Zawar-Reza, P., Pointing, S.B., Cary, S.C. 2014. Airborne Bacterial Populations Above Desert Soils of the McMurdo Dry Valleys, Antarctica. Microbial Ecology 67: 120-128. Boyd, P.W., Mackie, D.S., Hunter, K.A. 2010. Aerosol iron deposition to the surface ocean- modes of iron supply and biological responses. Marine Chemistry 120: 128- 143. Burkins, M.B., Virginia, R.A., Chamberlain, C.P., Wall, D.H. 2000. Origin and distribution of soil organic matter in Taylor Valley, Antarctica. Ecology 81(9): 2377- 2391. Calkin, P.E. 1964. Geomorphology and Glacial Geology of the Victoria Valley System, Southern Victoria Land, Antarctica. Institute of Polar Studies Report 10, Research Foundation and Institute of Polar Studies, The Ohio State University: 102 pages. Chadwick, OA, Derry LA, Vitousek PM, Huebert BJ, Hedin LO. 1999. Changing sources of nutrients during four million years of ecosystem development. Nature 397(6719):491–497. Clayton-Greene, J. M., Hendy, C. H. & Hogg, A. G. 1988. Chronology of a Wisconsin age proglacial lake in the Miers Valley, Antarctica, New Zealand Journal of Geology and Geophysics 31(3): 353-361, DOI: 10.1080/00288306.1988.10417781. Clow, G. D., McKay, C. P., Simmons, Jr., G. M., Wharton, Jr., R. A. 1988. Climatological observations and predicted sublimation rates at Lake Hoare, Antarctica. Journal of Climatology 1: 715–728. Coale, K. H., Johnson, K. S., Fitzwater, S. E., & Gordon, R. M. 1996. A massive phytoplankton bloom induced by an ecosystem-scale iron fertilization experiment in the equatorial pacific ocean. Nature 383(6600): 495-501. doi:http://dx.doi.org/10.1038/383495a0. Cooper, A.F., Adam, L.J., Coulter, R.F., Eby, G.N., McIntosh, W.C. 2007. Geology, geochronology and geochemistry of a basanitic volcano, White Island, Ross Sea, Antarctica. Journal of Volcanology and Geothermal Research 165: 189– 216. Cowan, D. A., Pointing, S. B., Stevens, M. I., Craig Cary, S., Stomeo, F., Tuffin, I. M. 2011. Distribution and abiotic influences on hypolithic microbial communities in an Antarctic Dry Valley. Polar Biology 34(2): 307–311. Craddock, C., 1970. Geologic maps of Antarctica. Antarctic Map Folio Series 12. In American Geographical Society p. 6.

57

Delmonte B., Basile-Doelsch I., Petit J., Maggi V., Revel-Rolland M., Michard A., Jagoutz E. and Grousset F. E. 2004. Comparing the Epica and Vostok dust records during the last 220,000 years: stratigraphical correlation and provenance in glacial periods. Earth-Science Reviews 66: 63-87. Delmonte, B., Baroni, C., Andersson, P. S., Schoberg, H., Hansson, M., Aciego, S., Petit, J-R., Albani, S., Mazzola, C., Maggi, V., Frezzotti, M. 2010. Aeolian dust in the Talos Dome ice core (East Antarctica, Pacific/Ross Sea sector): Victoria Land versus remote sources over the last two climate cycles. Journal of Quaternary Science 25: 1327–1337. ISSN 0267-8179. Denton, G.H., Sugden, D.E., Marchant, D.R., Hall, B.L., Wilch, T.I. 1993. East Antarctic Ice Sheet sensitivity to Pliocene climatic change from a Dry Valleys perspective. Geografiska Annalerv 75A: 155–204. Denton GH, Bockheim JG, Wilson SC, Stuiver M. 1989. Late Wisconsin and Early Holocene glacial history, Inner Ross embayment, Antarctica. Quaternary Research 31: 151-182. DePaolo, D.J. 1988. Neodymium Isotope Geochemistry, An Introduction. Springer- Verlag, Berlin, 187 p. Deuerling, K.M. 2010. Aeolian Sediments of the McMurdo Dry Valleys, Antarctica. The Ohio State University. MS Thesis. Deuerling, K.M., Lyons, W.B., Welch, S.A., Welch, K.A. 2014. The characterization and role of aeolian deposition on water quality, McMurdo Dry Valleys, Antarctica. Aeolian Research 13: 7-17. Deuerling, K.M., Baxter, E., Inglis, J., Lyons, W.B. 2014b. Geochemistry of mineral dust in the McMurdo Dry Valleys Region, Antarctica. ProScience 1: 306-311. Diatloff, E., Asher, C.J., Smith, F.W. 1996. Concentrations of rare earth elements in some Australian soils. Australian Journal of Soil Research 34: 735–747. Dickson, J.L., Head, J.W., Levy, J.S., and Marchant, D.R. 2013. Don Juan Pond, Antarctica: Near-surface CaCl2-brine feeding Earth’s most saline lake and implications for Mars. Scientific Reports 3: 1166. Doran, P.T., McKay, C.P., Clow, G.D., Dana, G.L., Fountain, A.G., Nylen, T., Lyons, W.B. 2002. Valley floor climate observations from the McMurdo Dry Valleys, Antarctica, 1986–2000. Journal of Geophysical Research-Atmospheres 107: 4772. Doran, P.T., Wharton Jr., R.A., Lyons, W.B. 1994. Paleolimnology of the McMurdo Dry Valleys, Antarctica. Journal of Paleolimnology 10: 85- 114. Dowling, C.B., Lyons, W.B., Welch, S.A. In Review. The geochemistry of glacial deposits in Taylor Valley, Antarctica: Comparison to upper continental crust abundances. Submission to Polar Science.

58

Faure, G. 1986. "Isotope geology of neodymium and strontium in igneous rocks", in Principles of Isotope Geology, second edition, John Wiley and Sons, New York: 217- 238. Foley K. K., Lyons W. B., Barrett J. E., Virginia R. A. 2006. Pedogenic carbonate distribution within glacial till in Taylor Valley, Southern Victoria Land, Antarctica. In Paleoenvironmental Record and Applications of Calcretes and Palustrine Carbonates. Alonso- Zarza AM and Tanner LH (eds.). Geological Society of America Special Paper 416: 89-103. Fountain, A.G., Levy, J.S., Gooseff, M.N., Van Horn, D. 2014. The McMurdo Dry Valleys: A landscape on the threshold of change. Geomorphology 225: 25-35. Fountain, A.G., Lyons, W.B., Burkins, M.B., Dana, G.L., Doran, P.T., Lewis, K.J., McKnight, D.M., Moorhead, D.L., Parsons, A.N., Priscu, J.C., Wall, D.H., Wharton, R.A., Virginia, R.A. 1999. Physical controls on the Taylor Valley ecosystem, Antarctica. Bioscience 49: 961–971. Fountain, A.G., Nylen, T.H., Monaghan, A., Basagic, H.J., Bromwich, D. 2010. Snow in the McMurdo Dry Valleys, Antarctica. International Journal of Climatology 30: 633– 642. Fountain, A.G., Tranter, M., Nylen, T.H., Lewis K.J., Mueller, D.R. 2004. Evolution of cryoconite holes and their contribution to meltwater runoff from glaciers in the McMurdo Dry Valleys, Antarctica. Journal of Glaciology 50(138): 35-45. Fristen, C.H., Grue, A.M., Priscu, J.C. 2000. Distribution of organic carbon and nitrogen in surface soils in the McMurdo Dry Valleys, Antarctica. Polar Biology 23: 121-128. Gaiero, D.M., Depetris, P.J., Probst, J.-L. Bidart, S.M., Leleyter., L. 2004. The signature of river-and wind-borne materials exported from Patagonia to the southern latitudes: a view from REEs and implications for paleoclimatic interpretations. Earth Planetary Science Letters 219 (3–4): 357–376. Gaiero, D.M., Probst, J.-L., Depetris, P.J., Bidart, S.M., Leleyter., L. 2003. Iron and other transition metals in Patagonian riverborne and windborne materials: geochemical control and transport to the southern South Atlantic Ocean. Geochimica et Cosmochimica Acta 67(19): 3603–3623. Gillies, J. A., Nickling, W. G., Tilson, M. 2013. Frequency, magnitude, and characteristics of aeolian sediment transport: McMurdo Dry Valleys, Antarctica. Journal of Geophysical Research: Earth Surface 118(2): 461–479. Guerzoni, S., Lenaz, R., Quarantotto, G., Taviani, M. 1992. Geochemistry of airborne particles from the lower troposphere of Terra Nova Bay, Antarctica. Tellus 44B: 304- 310. Gunn, B.M., Warren, G. 1962. Geology of Victoria Land between the Mawson and Mulock Glaciers, Antarctica. New Zealand Geological Survey Bulletin 71: 157.

59

Hall BL, Denton GH, Lux DR, Bockheim JG. 1993. Late Tertiary Antarctic paleoclimate and ice-sheet dynamics inferred from surficial deposits in Wright Valley. Geografiska Annalar 75A: 239-268. Hall, B.L., Denton, G.H., Stone, J.O., Conway, H. 2013. History of the grounded ice sheet in the Ross Sea sector of Antarctica during the Last Glacial Maximum and the last termination. Geological Society, London, Special Publications 381: 167-181. Harrington, H.J. 1958. Nomenclature of rock units in the Ross Sea region, Antarctica. Nature 182: 290. Honda, M., Yabuki, S., Shimizu, H., 2004. Geochemical and isotopic studies of aeolian sediments in China. Sedimentology 51(2): 211-230. Ishikawa, N., Kobayashi, S., Ohtake, T., Kawaguchi, S. 1982. Some radiation properties at Mizuho Station, East Antarctica in 1980, Mem. National Institute of Polar Research Special edition 24: 19–31. Jickells, T.D., An, Z.S., Andersen, K.K., Baker, A.R., Bergametti, G., Brooks, N., Cao, J.J., Boyd, P.W., Duce, R.A., Hunter, K.A., Kawahata, H., Kubilay, N., IaRoche, J., Liss, P.S., Mahowald, N., Prospero, J.M., Rigwell, A.J., Tegen, I., Torres, R. 2005. Global Iron Connections Between Desert Dust, Ocean Biogeochemistry, and Climate. Science 308(5718): 67-71. doi: 10.1126/science.1105959. Kelley, W. P. (1951), Alkali Soils: Their Formation, Properties, and Reclamation, Reinhold, New York. Keys, J.R., Williams, K., 1981. Origin of crystalline, cold desert salts in the McMurdo region, Antarctica. Geochimica et Cosmochimica Acta 45: 2299–2309. Kyle, P.R., 1990. McMurdo Volcanic Group, Western Ross Embayment. In: Le Masurier, W.E., Thomson, J.W. (Eds.), Volcanoes of the Antarctic Plate of the Southern Oceans: American Geophysical Union, Antarctic Research 48: 1–17. Lancaster, N. 2002. Flux of eolian sediment in the McMurdo Dry Valleys, Antarctica: a preliminary assessment. Arctic and Antarctic Alpine Research 34: 318–323. Lancaster, N. Nickling, W.G., Gillies, J.A. 2010. Sand transport by wind on complex surfaces: Field studies in the McMurdo Dry Valleys, Antarctica. Journal of Geophysical Research 115: F03027. Levy, J. S., Fountain, A. G., Welch, K. A., Lyons, W. B. 2012. Hypersaline “wet patches” in Taylor Valley, Antarctica. Geophysical Research Letters 39: (L05402). doi:10.1029/2012GL050898. Li, F., Ginoux P., Ramaswamy V. 2008. Distribution, transport, and deposition of mineral dust in the Southern Ocean and Antarctica: Contribution of major sources. Journal of Geophysical Research 113: D10207. Lyons, W. B., and P. A. Mayewski. 1993. The geochemical evolution of terrestrial waters in the Antarctic: The role of rock-water interactions. Physical and Biogeochemical

60

Processes in Antarctic Lakes, edited by W. J. Green and E. I. Friedman: 135-143, American Geophysical Union, Washington, D.C., Lyons, W. B., Welch, K. A., Neumann, K., Toxey, J. K., McArthur, R., Williams, C., McKnight, D. M., Moorhead, D. L. 1998. Geochemical linkages among glaciers, streams, and lakes within the Taylor Valley, Antarctica, in Ecosystem Dynamics in a Polar Desert: The McMurdo Dry Valleys, Antarctica, edited by J. C. Priscu, pp. 77– 92, American Geophysical Union, Washington, D. C. doi:10.1029/AR072p0077. Lyons, W. B., Fountain A. G., Doran P., Priscu J. C., Neumann K., Welch K. A. 2000. Importance of landscape position and legacy: the evolution of the lakes in Taylor Valley, Antarctica. Freshwater Biology 43: 355-367. Lyons, W. B., Welch, K. A., Fountain, A. G., Dana G., Vaughn., B. H. 2003. Surface glaciochemistry of Taylor Valley, Southern Victoria Land, Antarctica and its effect on stream chemistry. Hydrologic Processes 17: 115–130. Lyons, W.B., Welch, K.A., Snyder, G., Olesik, J., Graham, E.Y., Marion, G.M., Poreda, R.J. 2005. Halogen geochemistry of the McMurdo Dry Valleys Lakes, Antarctica: clues to the origin of solutes and lake evolution. Geochimica et Cosmochimica Acta 69: 305–323. Lyons, W.B., Welch, K.A., Gardner, C.B. 2012. The geochemistry of upland ponds, Taylor Valley, Antarctica. Antarctic Science 24: 3–14. Lyons, W. B., Deuerling, K., Welch, K. A., Welch, S. A., Michalski, G., Walters, W. W., Nielsen, U., Wall, D.H., Hogg, I., Adams, B. J. 2016. The Soil Geochemistry in the Beardmore Glacier Region, Antarctica: Implications for Terrestrial Ecosystem History. Scientific Reports 6: 26189. doi: 10.1038/srep26189. Marchant, D. R., Denton, G. H. (1996). Miocene and Pliocene paleoclimate of the Dry Valleys region, Southern Victoria land: a geomorphological approach. Marine Micropaleontology 27: 253–271. Martin, C.E. and McCulloch, M.T. 1999. Nd-Sr isotopic and trace element geochemistry of river sediments and soils in a fertilized catchment, New South Wales, Australia. Geochimica et Cosmochimica Acta 63(2): 287-305. Martin, J.H. and Fitzwater., S.E. 1988. Iron deficiency limits phytoplankton growth in the north-east Pacific subarctic. Nature 331: 341-343. Marx., S.K. Kamber, B.S., McGowan, H.A. 2005. Provenance of long-travelled dust determined with ultra-trace-element composition: a pilot study with samples from New Zealand glaciers. Earth Surface Processes and Landforms 30: 699-716. doi: 10.1002/esp.1169. McKelvey, B.C., Webb, P.N. 1962. Geological investigations in Southern Victoria Land, Antarctica Part 3: Geology of Wright Valley. New Zealand Journal of Geology and Geophysics 5: 143-162.

61

McLennan, S.M., Taylor, S.R., Eriksson., K.A. 1983. Geochemistry of Archean shales from the Pilbara Supergroup, Western Australia. Geochimica et Cosmochimia 47: 121l-1222. Mirsky, A. 1969. Geology of the Ohio Geology of the - Liv Glacier area. Antarctic Map Folio Series Bushnell VC (ed.) American Geographical Society, New York. Moreno, T., Querol, X., Castillo, S., Alastuey, A., Cuevas, E., Herrmann, L., Mounkaila, M., Elvira, J., Gibbons, W. 2006. Geochemical variations in aeolian mineral particles from the Sahara–Sahel Dust Corridor. Chemosphere 65: 261-270 https://doi.org/10.1016/j.chemosphere.2006.02.052. Muhs, D. R., Budahn, J., Skipp, G., Prospero, J. M., Patterson, D., Bettis, E. A. 2010. Geochemical and mineralogical evidence for Sahara and Sahel dust additions to Quaternary soils on Lanzarote, eastern Canary Islands, . Terra Nova 22(6): 399– 410. https://doi.org/10.1111/j.1365-3121.2010.00949. Neff, P.D., Bertler, N.A.N. 2015. Trajectory modeling of modern dust transport to the Southern Ocean and Antarctica. Journal of Geophysical Research- Atmospheres (120): 9303-9322. doi:10.1002/2015JD023304. Nesbitt H. W., Young G. M. 1982. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 299: 715-717. Nylen, T.H., Fountain, A.G., Doran, P.T. 2004. Climatology of katabatic winds in the McMurdo dry valleys, southern Victoria Land, Antarctica. Journal of Geophysical Research 109: D03114, doi:10.1029/2003JD003937. Okin, G.S., Gillette, D.A., Herrick, J.E. 2006. Multi-scale controls on and consequences of aeolian processes in landscape change in arid and semi-arid environments. Journal of Arid Environments 65(2): 253-275. Péwé, T. L. 1960. Multiple glaciation in the McMurdo Sound Region, Antarctica: A progress report. Journal of Geology 68(5): 498–514. doi:10.1086/ 626684. Piper, D.Z., Bau, M., 2013. Normalized rare earth elements in water, sediments, and wine: identifying sources and environmental redox conditions. American Journal Analytical Chemistry 4: 69e83. http://dx.doi.org/10.4236/ajac.2013.410A1009. Pocknall, D.T., Chinn, T.J., Sykes, R., Skinner, D.N.B. 1994. Geology of the Convoy Range area, southern Victoria Land, Antarctica: Lower Hutt, New Zealand, Institute of Geological and Nuclear Sciences. 36p. Porazinska, D.L., Fountain, A. G., Nylen, T. H., Tranter, M. Virginia, R. A., Wall., D. H. 2004. The Biodiversity and Biogeochemistry of Cryoconite Holes from McMurdo Dry Valley Glaciers, Antarctica. Arctic, Antarctic, and Alpine Research 36(1): 84-91. Porter, S. C. Beget, J. E. 1981. Provenance and Depositional Environments of Late Cenozoic Sediments in Permafrost Cores from Lower Taylor Valley, Antarctica, in Mcginnis LD (ed.), Dry Valley Drilling Project. American Geophysical Union. doi: 10.1029/AR033p0351. 62

Prentice, M.L., Kleman, J., Stroeven, J.P. 1998. The composite glacial erosional landscape of the Northern McMurdo Dry Valleys: implications for Antarctic Tertiary glacial history. In Priscu JC (ed.), Ecosystem Dynamics in a Polar Desert: The McMurdo Dry Valleys, Antarctica. Antarctic Research Series 72. American Geophysical Union: 1-38. Price, J.R., Velbel, M.A. 2003. Chemical weathering indices applied to weathering profiles developed on heterogenous felsic metamorphic parent rock. Chemical Geology 202: 397-416. Priscu J.C. 1995. Phytoplankton nutrient deficiency in lakes of the McMurdo Dry Valleys, Antarctica. Freshwater Biology 34: 215-227. Priscu J.C., Wolf D.F., Takacs C.D., Fritsen C.H., Laybourn-Parry J., Roberts E.C., Sattler B., Lyons W.B. 1999. Carbon transformations in a perennially ice-covered Antarctic lake. BioScience 49: 997-1008. Pye, K., 1987. Aeolian Dust and Dust Deposits. Academic Press, London. 334 pp. Raiswell, R., Benning, L.G., Tranter, M., Tulaczyk, S. 2008. Bioavailable iron in the Southern Ocean: the significance of the iceberg conveyor belt. Geochemical Transactions 9: 7. Roser B. P. and Pyne A. R. 1981. Whole rock geochemistry. In Antarctic Cenozoic history from the CIROS-1 drillhole, McMurdo Sound (ed. P. J. Barrett). pp. 175- 184. Rudnick, R.L, Gao, S. 2003. Composition of the Continental Crust. Composition of the continental crust. Elsevier, Oxford. Treatise on Geochemistry 3: 1-64. Šabacká, M., Priscu, J.C., Basagic, H.J., Fountain, A.G., Wall, D.H., Virginia, R.A., Greenwood, M.C. 2012. Aeolian flux of biotic and abiotic material in Taylor Valley, Antarctica. Geomorphology 155: 102-111. Sedwick, P.N., DiTullio, G.R., MacKey, D.J. 2000. Iron and manganese in the Ross Sea, Antarctica: seasonal iron limitation in Antarctic shelf waters. Journal of Geophysical Research- Oceans 105(11): 321-336. Shao J.Q., Yang S.Y. 2012. Does chemical index of alteration (CIA) reflect silicate weathering and monsoonal climate in the Changjiang River basin? Chinese Science Bulletin 57: 1179-1187, doi: 10.1007/s11434-011-4954-5. Smillie, R. W. 1992. Suite subdivision and petrological evolution of granitoids from the Taylor Valley and region, south Victoria Land. Antarctic Science 4(1): 71–87. https://doi.org/10.1017/S0954102092000130. Squyres, S. W., Andersen, D.W., Nedell, S.S., Wharton Jr., R.A. 1991. Lake Hoare Antarctica: Sedimentation through a thick perennial ice cover, Sedimentology 38: 363-379. Sugden, D., Denton, G. 2004. Cenozoic landscape evolution of the Convoy Range to Mackay Glacier area, Transantarctic Mountains : Onshore to offshore synthesis. Geological Society of America Bulletin 7/8: 840–857.

63

Toner, J. D., Sletten, R. S., Prentice, M. L. 2013. Soluble salt accumulations in Taylor Valley, Antarctica: Implications for paleolakes and Ross Sea Ice Sheet dynamics. Journal of Geophysical Research: Earth Surface 118(1): 198–215. https://doi.org/10.1029/2012JF002467. Vitousek, P., Chadwick O., Crew T., Fownes J., Hendrick D., Herbert D. 1997. Soil and ecosystem development across the Hawaiian Islands. GSA Today 7: 1-9. Wei, T., Dong, Z., Kang, S., Qin, X., Guo, Z. 2017. Geochemical evidence for sources of surface dust deposited on the Laohugou glacier, Qilian Mountains. Applied Geochemistry 79: 1-8. Welch, K.A., Mayewski, P.A., Whitlow, S.I. 1993. Methanesulfonic acid in coastal Antarctic snow related to sea-ice extent. Geophysical Research Letters 20(6): 443- 446. Wiggs, G., Baird, A., Atherton, R., 2004. The dynamic effects of moisture on the entrain- ment and transport of sand by wind. Geomorphology 59: 13–30. Wilch, T.I., Lux, D.R., Denton, G.H., McIntosh, W.C. 1993. Minimal Pliocene- Pleistocene uplift of the Dry Valleys sector of the Transantarctic mountains: A key parameter in ice sheet reconstructions. Geology 21: 841-844. Wilson, A. T. 1979. Geochemical problems of the Antarctic dry areas. Nature 280: 205– 208. Winton, V.H.L., Dunbar, G.B., Atkins, C.B., Bertler, N.A.N., Delmonte, B., Andersson, P.S., Bowie, A., Edwards, R. 2016. The origin of lithogenic sediment in the south- western Ross Sea and implications for iron fertilization. Antarctic Science 28(4): 250–260. Winton, V.H.L., Dunbar, G.B., Bertler, N.A.N., Millet, M.A., Delmonte, B., Atkins, C.B., Chewings, J.M.& Andersson, P. 2014. The contribution of aeolian sand and dust to iron fertilization of phytoplankton blooms in southwestern Ross Sea, Antarctica. Global Biogeochemical Cycles 28. Witherow, R.A., Lyons, W.B., Bertler, N.A.N., Welch, K.A., Mayewski, P.A., Sneed, S.B., Nylen, T., Handley, M.J., Fountain, A. 2006. The aeolian flux of calcium, chloride and nitrate to the McMurdo Dry Valleys landscape: evidence from snow pit analysis. Antarctic Science 18: 497-505. Xu, J., Yu, G., Kang, S. 2012. Sr-Nd isotope evidence for modern aeolian dust sources in mountain glaciers of western China. Journal of Glaciology 58(211): 859-865.

64

Section 10. Figures and Tables:

65

158°30’ E 169°30’ E

S Ross Sea

76°54’

S

78°06’

Figure 1. Satellite image of the McMurdo Dry Valleys (MDV) with the four valleys of interest highlighted. Sample locations represent a vertical transect (from Miers Valley to Alatna Valley) and a horizontal transect through Taylor Valley. McMurdo Station is identified for reference.

66

Figure 2. Satellite images of specific locations of the autonomous Big Spring Number Eight (BSNE) aeolian collectors within each valley. Numbers correspond to locational details in Table 1.

67

Figure 3. Image of aeolian collector used in this study, Big Spring Number Eight (BSNE) isokinetic wind sampler, located at F6 (Lake Fryxell). The letters represent the different collection heights: “A”- Top, “B”- Middle, “C”- Bottom upper, “D”- Bottom middle, “E”- Bottom lower. Strong winds have likely moved the bins, but the heights are roughly 100, 50, 20, 10, and 5 cm respectively.

68

Figure 4. Average daily wind speeds for Miers, Taylor, and Victoria Valleys attained from the McMurdo LTER online database (www.mcmlter.org). Wind speed information could not be found for Alatna Valley and Lake Fryxell. The black horizontal lines are the median, the colored rectangles are the 25th percentiles, and the vertical black lines indicate the spread of the 75th percentile. The black dots are outliers in the dataset and likely represent strong föhn events.

69

a.

b.

Figure 5. Ratios of inorganic carbon to organic carbon (a) and total carbon to organic carbon (b) by mass. Locations are ordered by distance from the coast (km) for Explorer’s Cove (EC), F6. Miers (M), East Lake Bonney (ELB), Alatna Valley (A), Taylor Glacier (TG), and Victoria Valley (V).

70

a.

b.

c.

Figure 6. Major oxide (a.) trace element (b.) and rare earth elements (REE) (c.) calculated by XRF and ICP-AES/ICP-MS, and normalized to the upper continental crust (Rudnick and Gao, 2003) for highest collection heights at each location. Open shapes represent full year samples, while closed are Winter 2014, except for Miers, which was a Summer 2013/2014 sample. No bulk chemistry data was attained for Taylor Glacier.

71

a.

b.

c.

Figure 7. Major oxide (a.) trace element (b.) and rare earth elements (REE) (c.) calculated by XRF and ICP-AES/ICP-MS normalized to the upper continental crust (Rudnick and Gao, 2003) for all collection heights during Winter 2014 at ELB and Explorer’s Cove.

72

Counts

Continued Figure 8. X-ray diffraction scans for the upper most samples of MDV aeolian material. The metric on the axis is 2θ spacing, which was used to determine minerology. The vertical numbered boxes correspond to highlighted areas where the minerology differed between samples and the grey vertical bar with the bolded letters are the labels for the detailed minerology in Figure 8. All samples in all highlighted areas had quartz, albite, augite, and labradorite.

73

Figure 8 Continued. 1: Muscovite (A, B, C, G), Palygorskite (C, D, F), Biotite (D, E, F), Wavellite (G) 2: Microcline (A, C, E, F, G), Anorthite (A, C, E, F, G), Orthoclase (C, D, E, F), Sanidine (A, B, C), Cupromakovickyite (B, C), Euchlorine (A) 3: Anorthite (A, C, E, F, G), Amphibole (C, E, F, G), Biotite (D, E, F), Palygorskite (C, D, F), Chlorapatite (D, E), Cupromakovickyite (B, C), Amesite (D) 4: Microcline (A, C, E, F, G), Anorthite (A, C, E, F, G), Muscovite (A, B, C, G), Orthoclase (C, D, E, F), Sanidine (A, B, C), Ilmenite (D, G), Clinopyroxene (A, C), Enstatite (F), Pigeonite (G), Wavellite (G) 5: Microcline (A, C, E, F, G), Anorthite (A, C, E, F, G), Orthoclase (C, D, E, F), Amphibole (C, E, F, G), Palygorskite (C, D, F), Clinopyroxene (A, C), Magnesio- riebeckite (A, B), Pigeonite (G), Wavellite (G), Enstatite (F) 6: Microcline (A, C, E, F, G), Muscovite (A, B, C, G), Amphibole (C, E, F, G), Sanidine (A, B, C), Clinopyroxene (A, C), Magnesio-riebeckite (A, B), Cupromakovickyite (B, C), Hedenbergite (C), Cronstedtite (C), Dolomite (E) 7: Microcline (A, C, E, F, G), Amphibole (C, E, F, G), Muscovite (A, B, C, G), Orthoclase (C, D, E, F), Sanidine (A, B, C), Palygorskite (C, D, F), Biotite (D, E, F), Forsterite (A, C), Enstatite (F), Hedenbergite (C), Pigeonite (G), Wavellite (G), 8: Microcline (A, C, E, F, G), Amphibole (C, E, F, G), Muscovite (A, B, C, G), Magnesio-hornblende (B, C, D), Palygorskite (C, D, F), Sanidine (A, B, C), Forsterite (A, C), Clinopyroxene (A, C), Ilmenite (D, G), Hedenbergite (C), Pigeonite (G), Pyrophyllite (G) 9: Microcline (A, C, E, F, G), Amphibole (C, E, F, G), Orthoclase (C, D, E, F), Chlorite- Serpentine (A, B, E, F), Muscovite (A, B, C, G), Palygorskite (C, D, F), Biotite (D, E, F), Magnesio-hornblende (B, C, D), Sanidine (A, B, C), Clinopyroxene (A, C), Forsterite (A, C), Pigeonite (G), Pyrophyllite (G), Wavellite (G), Amesite (D), Enstatite (F), Dolomite (E), Schorlomite (E), Hedenbergite (C)

74

a. b. c.

d. e. f.

Figure 9. High resolution imaging and mapping of aeolian material of Alatna Valley Mid using scanning electron microscopy (SEM). The elements from the EDSX detector in 9 f. are Ca and S. Further image details are found in Section 5.

75

a. b. c.

d. e. f.

Figure 10. High resolution imaging of aeolian material from Victoria Valley Mid using scanning electron microscopy (SEM). Further image details are found in Section 5.

76

a. b. c.

d. e. f.

Figure 11. High resolution imaging of aeolian material from Explorer’s Cove Top using scanning electron microscopy (SEM). The organism in 11 c. is the diatom Hantzschia amphioxys. Further image details are found in Section 5.

77

a. b. c.

d. e. f.

Figure 12. High resolution imaging and mapping of aeolian material from F6 (Lake Fryxell) Top using scanning electron microscopy (SEM). The organism in 12 d. is the diatom Muelleria meridionalis. The elements from the EDSX detector in 12 e. and f. are [P, La, and Ce], and [Na, Cl, S, and Mg]. Further image details are found in Section 5.

78

a. b. c.

d. e. f.

Figure 13. High resolution imaging and mapping of aeolian material from East Lake Bonney Mid using scanning electron microscopy (SEM). The elements from the EDSX detector in 13 f. are La and Ce. Further image details are found in Section 5.

79

a. b. c.

d. e. f.

Figure 14. High resolution imaging and mapping of aeolian material from Taylor Glacier Mid using scanning electron microscopy (SEM). The elements from the EDSX detector in 14 d., e. and f. are [Si, Ca, Al, K, and P], [Ca and S], and [Ti, P, and Zr]. Further image details are found in Section 5.

80

a. b. c.

d. e. f.

Figure 15. High resolution imaging and mapping of aeolian material from Miers Valley Top using scanning electron microscopy (SEM). Further image details are found in Section 5.

81

a.

b.

Figure 16. Anion and cation variation diagrams for all samples. The top two plots (a) are - by height and site, the second two plots (b) are by basin and year. HCO3 concentration was determined by charge balance. The colors correspond with the basin colors in Figure 1, where purple is Alatna valley, blue is Victoria Valley, green is Taylor Valley and red is Miers valley.

82

Figure 17. Spatial distribution of alkaline earth enriched MDV aeolian material for all samples. The black line is a carbonate dissolution line. Error bars are the standard error of duplicate samples.

83

Figure 18. Na+ and Cl- distribution in MDV aeolian material. The black line is a halite dissolution line. Error bars are the standard error of duplicate samples and are smaller than the points.

84

Figure 19. Major anions and cations associated with common salts in MDV aeolian material. The black line is a 1:1 dissolution line. Error bars are the standard error of duplicate samples.

85

Figure 20. Anion molar ratios for MDV aeolian material compared to published average soil salt distribution in Taylor Valley (Toner et al., 2013).

86

Figure 21. Anion molar ratios of aeolian material in Taylor Valley compared to nearby soil salt distributions (Toner et al., 2013).

87

Figure 22. Cross section of major features in Taylor Valley from McMurdo Sound to Taylor Glacier, including wind impeding features such as the Bonney Riegel, Nussbaum Riegel, and Coral Ridge (Toner et al., 2003). The dust collectors in Taylor Valley are superimposed over the image for reference. The large red arrow from left to right represents the relative strength of down valley winds compared to up valley winds, represented as the smaller arrow over McMurdo Sound.

88

a.

b.

2- 3- Figure 23. NO3 to PO4 molar ratios from all MDV aeolian material. Locations are ordered by distance from the coast (km) for Explorer’s Cove (EC), F6. Miers (M), East Lake Bonney (ELB), Alatna Valley (A), Taylor Glacier (TG), and Victoria Valley (V). The solid black line represents the Redfield Ratio of 16:1 for N:P. Values above the line indicate P limitations relative to N, while values below the line indicate N limitations relative to P.

89

Figure 24. Chemical Index of Alteration (CIA) values calculated using major oxide elemental data for aeolian sediment (points), average McMurdo Dry Valley source rocks (black lines), Patagonian dust (beige line), and average nearby sediment on glaciers and lakes (red line). Aeolian CIA values are plotted against collection height and compared to potential local sources. In the fourth CIA diagram, orange is Victoria Valley, green is Alatna Valley, and blue is Miers Valley.

McMurdo Volcanics: Cooper et al. (2007); Ferrar Dolerite: Antonini et al. (1999); Beacon Sandstone: Roser and Pyne (1981); glacier and lake sediment: Deuerling (2010); Patagonian dust: Gaiero et al. (2003).

90

Figure 25. Percent major oxides determined by XRF within each valley and compared to published geochemistry of potential source rocks, and foreign dust. All values were normalized to average continental crust (Rudnick and Gao, 2003). McMurdo Volcanics: Cooper et al. (2007); Ferrar Dolerite: Antonini et al. (1999); Australian soil: Martin and McCulloch (1999), McLennan et al. (1983); Patagonian dust and soil: Gaiero et al. (2003)

91

O(%)

2

O+ Na

2 K

Passive margin

SiO2 (%)

Figure 26. Major oxide data from XRF analysis from aeolian material, and potential source rocks. Red dashed lines represent potential mixing of source material, where the chemistry of any samples that fall on between the lines can be explained by a mixture McMurdo Volcanics (Cooper et al., 2007), Ferrar Dolerite (Antonini et al., 1999), granitoids (Smillie, 1992), and sandstone (Bhatia, 1983). For the sandstone, the passive margin samples were used for the mixing lines since they are more representative of Beacon Sandstone compared to active margin samples.

92

Figure 27. ƐNd(0) and 87Sr/86Sr fields of potential source areas for aeolian dust found in ice cores (Delmonte et al., 2004) Values of sediment on glacial surfaces (Deuerling et al., 2014) and sediment on the sea ice in McMurdo Sound (Winton et al., 2014; 2016) were superimposed on the original figure from Delmonte et al. (2004).

93

Nd Ɛ

87Sr/86Sr

Figure 28. ƐNd and 87Sr/86Sr isotopic signatures for sediments collected on sea ice in McMurdo Sound (Winton et al., 2014; 2016) and dust on glacial surfaces in MDV (Deuerling et al., 2014). Parent rock isotopic signatures (Delmonte et al., 2010) are also plotted to identify potential source material.

94

a.

Continued Figure 29. Rare Earth Element (REE) ratios calculated all MDV aeolian material and compared to published signatures from potential source rocks, foreign dust, and stream sediments. All values were normalized to average continental crust from Rudnick and Gao (2003). Red dashed lines represent potential mixing of source material, where the chemistry of any samples that fall on between the lines can be explained by a mixture McMurdo Volcanics, Ferrar Dolerite, and granite. Australian soil: Martin and McCulloch (1999), McLennan et al. (1983), Diatloff et al., (1996); Patagonia dust and soil: Gaiero et al. (2004), Gaiero et al. (2003); New Zealand dust: Marx et al. (2005); McMurdo Volcanics: Cooper et al. (2007); Ferrar Dolerite: Antonini et al. (1999); granite: USGS standard; soil: Dowling et al., in review.

95

Figure 29 continued.

b.

c.

Continued

96

Figure 29 continued.

d.

e.

97

Table 1. Sample site coordinates, distance from the coast and corresponding locations on Figure 1. Location Collection site Latitude Longitude Distance from Elevation on map name coast (km) (m) 1 Alatna Valley 76.90008°S 161.11102°E 36 950 2 Victoria 77.33092°S 161.60062°E 47 450 Valley 3 Explorer’s 77.58873°S 163.41752°E 3.5 24 Cove A and B 4 F6 (Lake 77.6085°S 163.2487°E 8.5 19 Fryxell) 5 East Lake 77.69263°S 162.56233°E 27 64 Bonney 6 Taylor Glacier 77.74002°S 162.13135°E 38.5 334

7 Miers Valley 78.09805°S 163.79423°E 10.5 50

98

Table 2. Average wind speeds for Miers, Taylor, and Victoria Valleys attained from the McMurdo LTER online database (www.mcmlter.org). Long term wind speed information was not available for Alatna Valley and Lake Fryxell. Miers Valley Explorer's Lake Bonney Taylor Glacier Victoria Year (m/s) Cove (m/s) (m/s) Valley (m/s) (m/s) Winter 2013 2.44 2.18 2.98 5.19 12.1 Summer 4.08 3.69 4.74 4.58 6.45 2013/2014 Winter 2014 3.35 2.73 5.97 5.71 2.19 Summer 3.46 3.63 4.43 3.95 5.19 2014/2015 Winter 2015 2.88 2.45 3.44 5.50 2.06 Full Year 3.09 2.47 3.53 5.41 2.28 2015

99

Table 3. Chemical Index of Alteration (CIA) values and Loss on Ignition (LOI) values for aeolian samples and MDV rocks. Geochemistry determined by XRF. Site Height Year LOI (%) CIA (%) Alatna Bottom Winter 2013 0.63 50.6 lower East Lake Bottom Winter 2014 1.00 53.4 Bonney lower (ELB) East Lake Bottom Winter 2014 0.92 54.0 Bonney middle (ELB) East Lake Bottom Winter 2014 0.77 53.7 Bonney upper (ELB) East Lake Middle Winter 2014 1.05 54.1 Bonney (ELB) East Lake Top Winter 2014 0.74 54.4 Bonney (ELB) East Lake Bottom 2015 0.78 53.5 Bonney lower (ELB) East Lake Bottom 2015 1.21 53.7 Bonney middle (ELB) East Lake Bottom 2015 0.85 53.6 Bonney upper (ELB) East Lake Middle 2015 0.93 53.1 Bonney (ELB) East Lake Top 2015 0.90 53.7 Bonney (ELB) Explorer's Bottom 2015 1.80 53.3 Cove A lower Explorer's Bottom 2015 1.51 53.5 Cove A middle Explorer's Bottom 2015 2.16 53.2 Cove A upper Explorer's Middle 2015 1.85 53.2 Cove A Explorer's Top 2015 1.91 52.7 Cove A Explorer's Bottom Winter 2014 1.67 53.0 Cove B lower Continued

100

Table 3 Continued. Explorer's Bottom Winter 2014 1.76 53.3 Cove B middle Explorer's Bottom Winter 2014 1.67 53.6 Cove B upper Explorer's Middle Winter 2014 1.86 52.7 Cove B Explorer's Top Winter 2014 1.90 54.1 Cove B Explorer's Bottom 2015 1.39 54.3 Cove B lower Explorer's Bottom 2015 2.17 52.6 Cove B middle Explorer's Bottom 2015 1.67 53.2 Cove B upper Explorer's Middle 2015 1.88 53.1 Cove B Explorer's Top 2015 1.97 52.9 Cove B F6 (Lake Bottom Winter 2014 2.16 51.2 Fryxell) upper F6 (Lake Middle Winter 2014 2.26 51.9 Fryxell) F6 (Lake Top Winter 2014 1.27 52.0 Fryxell) Miers Bottom Summer 2.05 51.2 upper 2013/2014 Miers Middle Summer 1.58 53.5 2013/2014 Miers Top Summer 1.27 54.7 2013/2014 Miers Bottom Winter 2014 1.83 51.3 upper Victoria Middle Winter 2013 1.09 44.1

Victoria Top Winter 2013 1.12 43.6

MDV rocks and dust McMurdo Cooper et al. - - 46.9 Volcanics* (2007) Ferrar Antonini et - - 55.4 Dolerite* al. (1999) Beacon Roser and - - 64.6 Sandstone* Pyne (1989) Patagonian Gaiero et al. - - 57.8 dust* (2004)

101

Table 4. Concentrations of water soluble major anions in MDV aeolian material, determined by ion chromatography in µmol g-1. Dashes represent measurements below the detection limit. - -1 - -1 - -1 2- -1 Site Height Year F (µmol g ) Cl (µmol g ) Br (µmol g ) SO4 (µmol g )

Leach Leach Total Leach Leach Total Leach Leach Total Leach Leach Total 1 2 1 2 1 2 1 2 Alatna Bottom Winter 0.019 0.024 0.043 0.284 0.026 0.310 - - - 0.502 0.528 1.031 lower 2013 Alatna Bottom Winter 0.012 0.031 0.043 0.274 0.149 0.424 - - - 0.193 0.202 0.395 middle 2013 Alatna Middle Winter 0.014 0.023 0.037 0.975 0.100 1.076 - - - 0.836 0.427 1.264 2013 East Lake Bottom Winter 0.010 0.009 0.020 4.959 0.620 5.579 - - - 0.868 0.760 1.628 Bonney lower 2014 (ELB) East Lake Bottom Winter 0.005 0.010 0.015 2.473 1.364 3.838 - - - 0.330 0.756 1.086 102 Bonney middle 2014

(ELB) East Lake Bottom Winter 0.015 0.011 0.027 4.997 0.568 5.565 - - - 1.155 0.383 1.539 Bonney upper 2014 (ELB) East Lake Middle Winter 0.014 0.010 0.025 3.892 0.512 4.405 - - - 0.877 0.397 1.275 Bonney 2014 (ELB) East Lake Top Winter 0.011 0.011 0.023 2.675 0.393 3.069 - - - 0.370 0.199 0.570 Bonney 2014 (ELB) East Lake Bottom 2015 0.006 0.010 0.016 2.027 0.354 2.381 - - - 0.272 0.330 0.602 Bonney lower (ELB) East Lake Bottom 2015 0.005 0.014 0.019 1.994 0.455 2.449 - - - 0.278 0.233 0.512 Bonney middle (ELB) Continued

Table 4 Continued.

East Lake Bottom 2015 0.007 0.007 0.015 2.252 0.349 2.601 - - - 0.569 0.269 0.839 Bonney upper (ELB) East Lake Middle 2015 0.010 0.012 0.022 2.564 0.359 2.923 - - - 0.550 0.648 1.198 Bonney (ELB) East Lake Top 2015 0.008 0.015 0.023 1.956 0.332 2.288 - 0.003 0.003 0.361 0.724 1.085 Bonney (ELB) Explorer’s Middle Winter 0.002 0.004 0.007 1.232 0.219 1.451 - - - 0.300 0.319 0.620 Cove B 2013 Explorer’s Bottom 2015 0.012 0.013 0.026 5.917 0.835 6.753 0.008 0.002 0.010 0.355 0.208 0.563 Cove A lower Explorer’s Bottom 2015 0.010 0.012 0.022 1.821 0.400 2.221 0.002 - 0.002 0.218 0.305 0.523

103 Cove A middle Explorer’s Bottom 2015 0.018 0.019 0.037 4.816 0.713 5.529 0.005 - 0.005 0.333 0.226 0.560

Cove A upper Explorer’s Middle 2015 0.015 0.018 0.034 2.334 0.390 2.725 0.001 - 0.001 0.246 0.225 0.471 Cove A Explorer’s Top 2015 0.012 0.016 0.029 2.422 0.334 2.756 0.002 - 0.002 0.272 0.223 0.496 Cove A Explorer’s Bottom Winter 0.012 0.012 0.024 0.571 0.164 0.735 - - - 0.121 0.150 0.271 Cove B lower 2014 Explorer’s Bottom Winter 0.011 0.012 0.023 1.183 0.148 1.331 0.010 - 0.010 0.114 0.133 0.247 Cove B middle 2014 Explorer’s Bottom Winter 0.014 0.012 0.026 0.486 0.140 0.627 0.002 - 0.002 0.146 0.116 0.262 Cove B upper 2014 Explorer’s Middle Winter 0.014 0.013 0.028 6.218 0.598 6.816 0.007 - 0.007 0.375 0.188 0.563 Cove B 2014 Explorer’s Top Winter 0.012 0.013 0.026 7.481 0.702 8.183 0.009 - 0.009 0.483 0.176 0.659 Cove B 2014 Explorer’s Bottom 2015 0.010 0.010 0.021 2.378 0.386 2.764 - 0.001 0.001 0.250 0.141 0.392 Cove B lower Continued

Table 4 Continued. Explorer’s Bottom 2015 0.019 0.014 0.033 10.00 0.757 10.76 - - - 0.527 0.110 0.638 Cove B middle Explorer’s Bottom 2015 0.012 0.009 0.022 2.118 0.336 2.454 - - - 0.230 0.157 0.387 Cove B upper Explorer’s Middle 2015 0.016 0.016 0.033 2.474 0.372 2.847 - - - 0.246 0.105 0.352 Cove B Explorer’s Top 2015 0.020 0.019 0.039 2.902 0.401 3.304 - - - 0.267 0.093 0.360 Cove B F6 (Lake Bottom Summer 0.009 0.014 0.024 4.573 0.497 5.071 - 0.015 0.015 0.620 0.218 0.838 Fryxell) lower 2013/2014 F6 (Lake Bottom Summer 0.006 0.011 0.017 3.912 0.456 4.369 - 0.016 0.016 0.590 0.182 0.773 Fryxell) middle 2013/2014 F6 (Lake Bottom Summer 0.007 0.008 0.016 1.621 0.217 1.838 - 0.016 0.016 0.397 0.170 0.568 Fryxell) upper 2013/2014

104 F6 (Lake Middle Summer 0.006 0.011 0.017 4.853 1.502 6.356 0.019 0.015 0.035 0.435 0.198 0.634 Fryxell) 2013/2014

F6 (Lake Top Summer 0.015 0.023 0.039 17.61 1.303 18.91 0.036 0.015 0.052 1.520 0.347 1.868 Fryxell) 2013/2014 F6 (Lake Bottom Winter 0.023 0.018 0.042 13.87 1.248 15.12 0.007 0.000 0.007 1.119 0.159 1.279 Fryxell) upper 2014 F6 (Lake Middle Winter 0.032 0.026 0.058 10.46 1.086 11.55 0.005 0.001 0.006 1.256 0.163 1.420 Fryxell) 2014 F6 (Lake Top Winter 0.022 0.027 0.050 17.57 1.531 19.10 0.015 0.001 0.017 0.947 0.139 1.087 Fryxell) 2014 Miers Bottom Summer 0.003 0.007 0.010 0.515 0.132 0.647 - - - 0.121 0.195 0.317 lower 2013/2014 Miers Bottom Summer 0.006 0.006 0.013 0.251 0.115 0.367 - - - 0.149 0.178 0.328 middle 2013/2014 Miers Bottom Summer 0.006 0.004 0.011 0.213 0.110 0.323 - - - 0.124 0.232 0.356 upper 2013/2014 Miers Middle Summer 0.005 0.007 0.013 0.206 0.111 0.318 - - - 0.109 0.169 0.278 2013/2014 Continued

Table 4 Continued. Miers Top Summer 0.004 0.005 0.010 0.190 0.109 0.300 - - - 0.108 0.153 0.261 2013/2014 Miers Bottom Summer 0.003 0.003 0.007 0.334 0.046 0.380 - - - 0.159 0.143 0.302 upper 2014/2015 Miers Middle Summer 0.002 0.005 0.008 0.291 0.080 0.372 - - - 0.187 0.174 0.361 2014/2015 Taylor Bottom Winter 0.003 0.004 0.007 1.943 0.098 2.041 - - - 0.236 0.250 0.486 Glacier lower 2014 Taylor Bottom Winter 0.003 0.006 0.010 1.841 0.110 1.952 - - - 0.280 0.177 0.458 Glacier middle 2014 Taylor Bottom Winter 0.002 0.004 0.007 1.724 0.107 1.831 - - - 0.268 0.187 0.455 Glacier upper 2014 Taylor Middle Winter 0.007 0.011 0.018 3.707 0.205 3.912 - - - 0.817 0.237 1.054 Glacier 2014

105 Victoria Bottom Winter 0.005 0.008 0.013 0.774 0.149 0.924 - - - 0.559 0.282 0.842 lower 2013

Victoria Bottom Winter 0.006 0.009 0.016 1.414 0.183 1.598 - - - 0.568 0.283 0.852 middle 2013 Victoria Bottom Winter 0.003 0.002 0.005 0.149 0.012 0.161 - - - 0.151 0.150 0.301 upper 2013 Victoria Middle Winter 0.002 0.004 0.007 0.122 0.905 1.027 - - - 0.416 0.087 0.503 2013 Victoria Top Winter 0.003 0.009 0.012 1.256 0.853 2.110 - - - 0.441 0.463 0.904 2013

Table 5. Concentrations of water soluble major cations in MDV aeolian material, determined by ion chromatography in µmol g-1. Dashes represent measurements below the detection limit. Site Height Year Na+ (µmol g-1) K+ (µmol g-1) Mg2+ (µmol g-1) Ca2+ (µmol g-1)

Leach Leach Total Leach Leach Total Leach Leach Total Leach Leach Total 1 2 1 2 1 2 1 2 Alatna Bottom Winter 1.471 0.882 2.352 0.025 0.022 0.047 0.178 0.201 0.380 0.320 0.322 0.642 lower 2013 Alatna Bottom Winter 0.815 0.556 1.371 0.052 0.053 0.106 0.103 0.100 0.203 0.169 0.170 0.338 middle 2013 Alatna Middle Winter 2.528 1.117 3.645 0.036 0.022 0.058 0.333 0.133 0.465 0.905 0.322 1.227 2013 East Lake Bottom Winter 3.862 1.004 4.866 0.251 0.115 0.366 1.091 0.351 1.442 1.254 0.811 2.064 Bonney lower 2014 (ELB)

106 East Lake Bottom Winter 1.970 1.585 3.554 0.147 0.136 0.283 0.555 0.494 1.049 0.641 0.874 1.515

Bonney middle 2014 (ELB) East Lake Bottom Winter 4.225 1.522 5.747 0.249 0.113 0.361 1.127 0.259 1.385 1.428 0.729 2.158 Bonney upper 2014 (ELB) East Lake Middle Winter 3.387 1.318 4.705 0.222 0.099 0.321 0.920 0.358 1.278 1.143 0.737 1.881 Bonney 2014 (ELB) East Lake Top Winter 2.467 1.022 3.489 0.162 0.086 0.248 0.544 0.294 0.838 0.656 0.613 1.269 Bonney 2014 (ELB) East Lake Bottom 2015 1.900 0.781 2.681 0.118 0.065 0.183 0.451 0.294 0.744 0.515 0.505 1.019 Bonney lower (ELB) Continued

Table 5 Continued. East Lake Bottom 2015 1.818 1.057 2.875 0.119 0.086 0.205 0.434 0.386 0.819 0.546 0.557 1.103 Bonney middle (ELB) East Lake Bottom 2015 2.119 0.903 3.022 0.144 0.075 0.219 0.579 0.215 0.794 0.876 0.581 1.457 Bonney upper (ELB) East Lake Middle 2015 2.391 1.013 3.404 0.168 0.091 0.259 0.683 0.281 0.963 0.902 0.758 1.660 Bonney (ELB) East Lake Top 2015 1.859 1.345 3.203 0.125 0.123 0.248 0.552 0.410 0.962 0.639 1.080 1.719 Bonney (ELB) Explorer’s Middle Winter 2.281 1.883 4.164 0.103 0.095 0.198 0.125 0.115 0.240 0.269 0.249 0.518 Cove B 2013

107 Explorer’s Bottom 2015 7.788 3.105 10.89 0.244 0.125 0.368 0.305 0.150 0.454 0.297 0.195 0.492 Cove A lower

Explorer’s Bottom 2015 3.206 2.573 5.779 0.126 0.112 0.237 0.096 0.106 0.203 0.144 0.192 0.336 Cove A middle Explorer’s Bottom 2015 6.900 3.438 10.34 0.217 0.125 0.343 0.200 0.098 0.298 0.243 0.179 0.422 Cove A upper Explorer’s Middle 2015 4.037 2.947 6.984 0.141 0.116 0.257 0.082 0.079 0.161 0.189 0.173 0.362 Cove A Explorer’s Top 2015 4.286 2.928 7.214 0.149 0.115 0.265 0.075 0.068 0.143 0.185 0.151 0.336 Cove A Explorer’s Bottom Winter 1.725 2.613 4.339 0.077 0.111 0.188 0.054 0.068 0.121 0.168 0.201 0.369 Cove B lower 2014 Explorer’s Bottom Winter 1.785 2.827 4.612 0.073 0.118 0.191 0.023 0.048 0.071 0.142 0.175 0.317 Cove B middle 2014 Explorer’s Bottom Winter 2.825 2.906 5.732 0.089 0.091 0.179 0.021 0.034 0.055 0.139 0.145 0.285 Cove B upper 2014 Explorer’s Middle Winter 8.690 3.176 11.87 0.230 0.124 0.355 0.170 0.057 0.227 0.354 0.166 0.520 Cove B 2014 Continued

Table 5 Continued. Explorer’s Top Winter 9.979 2.942 12.92 0.260 0.119 0.379 0.359 0.086 0.445 0.508 0.181 0.689 Cove B 2014 Explorer’s Bottom 2015 3.850 2.111 5.961 0.167 0.123 0.290 0.148 0.107 0.255 0.188 0.206 0.394 Cove B lower Explorer’s Bottom 2015 12.62 2.913 15.53 0.377 0.128 0.505 0.575 0.111 0.686 0.394 0.181 0.575 Cove B middle Explorer’s Bottom 2015 3.627 2.254 5.881 0.138 0.103 0.241 0.093 0.069 0.162 0.196 0.188 0.384 Cove B upper Explorer’s Middle 2015 4.352 2.889 7.241 0.144 0.125 0.269 0.084 0.092 0.176 0.201 0.185 0.386 Cove B Explorer’s Top 2015 4.968 3.212 8.180 0.159 0.130 0.289 0.108 0.088 0.196 0.196 0.180 0.376 Cove B F6 (Lake Bottom Summer 6.262 2.138 8.400 0.315 0.098 0.413 0.423 0.102 0.526 0.703 0.254 0.956 Fryxell) lower 2013/2014

108 F6 (Lake Bottom Summer 5.407 1.936 7.343 0.286 0.102 0.388 0.408 0.103 0.511 0.804 0.273 1.077 Fryxell) middle 2013/2014

F6 (Lake Bottom Summer 3.156 1.560 4.716 0.170 0.093 0.262 0.146 0.073 0.219 0.358 0.225 0.583 Fryxell) upper 2013/2014 F6 (Lake Middle Summer 5.270 2.258 7.529 0.273 0.130 0.403 0.581 0.169 0.750 0.649 0.275 0.923 Fryxell) 2013/2014 F6 (Lake Top Summer 14.11 3.114 17.23 0.829 0.195 1.023 2.478 0.307 2.785 2.269 0.368 2.638 Fryxell) 2013/2014 F6 (Lake Bottom Winter 15.09 3.583 18.67 0.531 0.130 0.661 0.704 0.109 0.812 0.454 0.316 0.769 Fryxell) upper 2014 F6 (Lake Middle Winter 11.96 4.865 16.83 0.434 0.198 0.632 0.403 0.092 0.495 0.350 0.171 0.521 Fryxell) 2014 F6 (Lake Top Winter 18.56 5.032 23.59 0.663 0.208 0.870 0.748 0.135 0.882 0.423 0.232 0.655 Fryxell) 2014 Miers Bottom Summer 1.057 0.812 1.869 0.097 0.086 0.184 0.066 0.045 0.111 0.275 0.384 0.659 lower 2013/2014 Miers Bottom Summer 0.599 0.712 1.311 0.086 0.094 0.180 0.050 0.054 0.104 0.308 0.492 0.800 middle 2013/2014 Continued

Table 5 Continued. Miers Bottom Summer 0.541 0.727 1.268 0.082 0.089 0.171 0.036 0.046 0.082 0.299 0.466 0.766 upper 2013/2014 Miers Middle Summer 0.505 0.667 1.172 0.079 0.086 0.165 0.031 0.042 0.073 0.292 0.460 0.752 2013/2014 Miers Top Summer 0.446 0.594 1.040 0.073 0.081 0.154 0.024 0.035 0.059 0.252 0.436 0.687 2013/2014 Miers Bottom Summer 0.634 0.560 1.193 0.056 0.052 0.108 0.047 0.040 0.088 0.288 0.478 0.766 upper 2014/2015 Miers Middle Summer 0.739 0.723 1.462 0.065 0.063 0.129 0.065 0.052 0.118 0.305 0.458 0.763 2014/2015 Taylor Bottom Winter 1.926 0.964 2.890 0.064 0.028 0.091 0.202 0.044 0.246 0.554 0.171 0.725 Glacier lower 2014 Taylor Bottom Winter 1.950 0.957 2.907 0.060 0.030 0.090 0.195 0.045 0.240 0.573 0.447 1.020 Glacier middle 2014

109 Taylor Bottom Winter 1.793 0.900 2.693 0.052 0.028 0.080 0.159 0.042 0.201 0.533 0.242 0.774 Glacier upper 2014

Taylor Middle Winter 3.341 1.128 4.468 0.104 0.034 0.138 0.407 0.064 0.471 1.350 0.447 1.797 Glacier 2014 Victoria Bottom Winter 1.621 0.950 2.571 0.061 0.043 0.104 0.178 0.091 0.269 0.943 0.509 1.452 lower 2013 Victoria Bottom Winter 2.170 1.003 3.173 0.097 0.055 0.152 0.302 0.118 0.420 0.948 0.462 1.410 middle 2013 Victoria Bottom Winter 0.511 0.439 0.951 0.035 0.028 0.063 0.031 0.041 0.072 0.224 0.381 0.605 upper 2013 Victoria Middle Winter 1.552 0.575 2.127 0.054 0.059 0.113 0.095 0.078 0.174 0.667 0.408 1.075 2013 Victoria Top Winter 1.962 2.001 3.963 0.065 0.065 0.130 0.087 0.085 0.172 0.651 0.553 1.205 2013

- 3- Table 6. Concentrations of water soluble nutrients (NO3 and PO4 ) in MDV aeolian material determined by colorimetric analysis in µmol g-1. Dashes represent measurements below the detection limit. - -1 3- -1 Site Height Year NO3 (µmol g ) PO4 (µmol g ) N:P

Leach Leach Total Leach Leach Total Norm. 1 2 1 2 to P Alatna Bottom Winter 0.235 0.023 0.258 0.0010 0.0016 0.0026 100.2 lower 2013 Alatna Bottom Winter 0.215 0.060 0.276 - 0.0003 0.0003 805.1 middle 2013 Alatna Middle Winter 0.242 0.026 0.268 - 0.0004 0.0004 672.7 2013 East Lake Bottom Winter 0.068 0.011 0.079 - - - - Bonney lower 2014 (ELB) East Lake Bottom Winter 0.039 0.022 0.061 - - - - Bonney middle 2014 (ELB) East Lake Bottom Winter 0.063 0.006 0.068 0.0012 - 0.0012 57.13 Bonney upper 2014 (ELB) East Lake Middle Winter 0.051 0.015 0.066 0.0006 0.0010 0.0016 40.45 Bonney 2014 (ELB) East Lake Top Winter 0.031 0.012 0.043 0.0008 - 0.0008 51.12 Bonney 2014 (ELB) East Lake Bottom 2015 0.025 0.010 0.035 0.0011 0.0014 0.0025 13.89 Bonney lower (ELB) East Lake Bottom 2015 0.025 0.011 0.036 0.0015 0.0013 0.0028 12.94 Bonney middle (ELB) East Lake Bottom 2015 0.028 0.008 0.036 0.0011 0.0016 0.0027 13.48 Bonney upper (ELB) East Lake Middle 2015 0.034 0.008 0.043 0.0015 0.0012 0.0027 15.79 Bonney (ELB) East Lake Top 2015 0.023 0.007 0.030 0.0017 0.0011 0.0028 10.72 Bonney (ELB) Explorer’s Middle Winter 0.026 0.010 0.035 0.0007 0.0010 0.0017 21.28 Cove B 2013 Explorer’s Bottom 2015 0.016 0.007 0.023 0.0019 0.0031 0.0050 4.661 Cove A lower Explorer’s Bottom 2015 0.011 0.006 0.017 0.0029 0.0040 0.0069 2.496 Cove A middle Explorer’s Bottom 2015 0.024 0.007 0.031 0.0027 0.0088 0.0114 2.753 Cove A upper Continued

110

Table 6 Continued. Explorer’s Middle 2015 0.020 0.007 0.027 0.0038 0.0090 0.0128 2.104 Cove A Explorer’s Top 2015 0.021 0.008 0.029 0.0044 0.0103 0.0147 1.981 Cove A Explorer’s Bottom Winter 0.003 0.003 0.006 0.0034 0.0089 0.0123 0.494 Cove B lower 2014 Explorer’s Bottom Winter 0.004 0.003 0.007 0.0056 0.0093 0.0150 0.482 Cove B middle 2014 Explorer’s Middle Winter 0.019 0.009 0.028 0.0011 0.0094 0.0105 2.679 Cove B 2014 Explorer’s Top Winter 0.024 0.011 0.035 0.0003 0.0073 0.0076 4.608 Cove B 2014 Explorer’s Bottom 2015 0.016 0.006 0.022 0.0020 0.0043 0.0063 3.532 Cove B lower Explorer’s Bottom 2015 0.042 0.007 0.049 0.0022 0.0049 0.0071 6.871 Cove B middle Explorer’s Bottom 2015 0.015 0.006 0.021 0.0028 0.0050 0.0078 2.662 Cove B upper Explorer’s Middle 2015 0.026 0.008 0.033 0.0038 0.0089 0.0127 2.617 Cove B Explorer’s Top 2015 0.028 0.009 0.037 0.0047 0.0091 0.0138 2.679 Cove B F6 (Lake Bottom Summer 0.091 0.016 0.107 - 0.0032 0.0032 33.77 Fryxell) lower 2013/2014 F6 (Lake Bottom Summer 0.079 0.014 0.093 0.0004 0.0021 0.0025 37.40 Fryxell) middle 2013/2014 F6 (Lake Bottom Summer 0.028 0.009 0.037 0.0009 0.0028 0.0037 9.918 Fryxell) upper 2013/2014 F6 (Lake Middle Summer 0.057 0.021 0.078 - 0.0003 0.0003 250.1 Fryxell) 2013/2014 F6 (Lake Top Summer 0.215 0.026 0.241 0.0019 0.0019 124.8 Fryxell) 2013/2014 F6 (Lake Bottom Winter 0.149 0.029 0.178 0.0008 0.0060 0.0068 26.26 Fryxell) upper 2014 F6 (Lake Middle Winter 0.124 0.024 0.147 0.0010 0.0102 0.0112 13.14 Fryxell) 2014 F6 (Lake Top Winter 0.162 0.022 0.184 - 0.0042 0.0042 43.79 Fryxell) 2014 Miers Bottom Summer 0.022 0.007 0.030 - - - - lower 2013/2014 Miers Bottom Summer 0.021 0.003 0.025 0.0000 0.0005 0.0006 42.99 middle 2013/2014 Miers Bottom Summer 0.018 0.004 0.022 0.0003 0.0006 0.0009 25.49 upper 2013/2014 Miers Middle Summer 0.016 - 0.016 - 0.0007 0.0007 22.72 2013/2014 Miers Top Summer 0.015 - 0.015 - 0.0002 0.0002 70.72 2013/2014 Miers Bottom Summer 0.022 0.007 0.029 0.0006 0.0003 0.0009 33.36 upper 2014/2015 Continued

111

Table 6 Continued. Miers Middle Summer 0.023 0.010 0.033 0.0003 - 0.0003 116.0 2014/2015 Taylor Glacier Bottom Winter 0.231 0.001 0.232 - 0.0002 0.0002 993.9 lower 2014 Taylor Glacier Bottom Winter 0.199 0.018 0.217 0.0003 - 0.0003 826.8 middle 2014 Taylor Glacier Bottom Winter 0.221 0.017 0.238 0.0004 0.0018 0.0023 105.2 upper 2014 Taylor Glacier Middle Winter 0.385 0.026 0.411 0.0002 0.0008 0.0010 399.6 2014 Victoria Bottom Winter 0.253 0.029 0.282 0.0000 0.0000 0.0000 5699 lower 2013 Victoria Bottom Winter 0.309 0.031 0.340 0.0002 - 0.0002 1950 middle 2013 Victoria Bottom Winter 0.034 0.007 0.041 0.0006 0.0007 0.0014 29.87 upper 2013 Victoria Middle Winter 0.138 0.014 0.151 0.0001 0.0010 0.0010 147.0 2013 Victoria Top Winter 0.168 0.112 0.280 - - - - 2013

112

Table 7. Total (TC), organic (OC) and inorganic (IC) carbon in the MDV aeolian solids. Site Height Year TC OC IC (mg/g) (mg/g) (mg/g) Alatna Bottom Winter 2013 0.316 0.282 0.034 lower Alatna Bottom Winter 2013 0.314 0.258 0.056 middle Alatna Middle Winter 2013 0.548 0.514 0.034

East Lake Bottom Winter 2014 0.887 0.381 0.506 Bonney lower (ELB) East Lake Bottom Winter 2014 0.730 0.454 0.277 Bonney middle (ELB) East Lake Bottom Winter 2014 2.019 0.307 1.712 Bonney upper (ELB) East Lake Middle Winter 2014 1.131 0.544 0.587 Bonney (ELB) East Lake Top Winter 2014 0.666 0.411 0.256 Bonney (ELB) East Lake Bottom 2015 0.604 0.409 0.196 Bonney lower (ELB) East Lake Bottom 2015 0.683 0.439 0.244 Bonney middle (ELB) East Lake Bottom 2015 0.785 0.361 0.424 Bonney upper (ELB) East Lake Middle 2015 1.765 0.380 1.385 Bonney (ELB) East Lake Top 2015 0.828 0.486 0.342 Bonney (ELB) Explorer’s Middle Winter 2013 2.319 0.876 1.443 Cove B Explorer’s Bottom 2015 1.305 0.829 0.476 Cove A lower Explorer’s Bottom 2015 1.451 0.490 0.961 Cove A middle Explorer’s Bottom 2015 2.146 0.565 1.581 Cove A upper Continued

113

Table 7 Continued. Explorer’s Middle 2015 1.639 1.046 0.593 Cove A Explorer’s Top 2015 1.615 0.500 1.115 Cove A Explorer’s Bottom Winter 2014 1.217 0.382 0.835 Cove B lower Explorer’s Bottom Winter 2014 1.389 0.461 0.929 Cove B middle Explorer’s Bottom Winter 2014 1.658 0.691 0.967 Cove B upper Explorer’s Middle Winter 2014 1.477 0.591 0.886 Cove B Explorer’s Top Winter 2014 1.561 0.652 0.910 Cove B Explorer’s Bottom 2015 0.849 0.362 0.487 Cove B lower Explorer’s Bottom 2015 1.648 0.303 1.346 Cove B middle Explorer’s Bottom 2015 1.175 0.431 0.744 Cove B upper Explorer’s Middle 2015 1.364 0.700 0.665 Cove B Explorer’s Top 2015 1.601 0.516 1.086 Cove B F6 (Lake Bottom Summer 1.603 0.768 0.835 Fryxell) lower 2013/2014 F6 (Lake Bottom Summer 2.224 1.367 0.857 Fryxell) middle 2013/2014 F6 (Lake Bottom Summer 1.367 0.747 0.620 Fryxell) upper 2013/2014 F6 (Lake Middle Summer 1.892 0.681 1.212 Fryxell) 2013/2014 F6 (Lake Top Summer 1.331 0.456 0.875 Fryxell) 2013/2014 F6 (Lake Bottom Winter 2014 1.675 0.843 0.832 Fryxell) upper F6 (Lake Middle Winter 2014 2.233 1.028 1.205 Fryxell) F6 (Lake Top Winter 2014 1.935 0.949 0.986 Fryxell) Miers Bottom Summer 3.224 0.804 2.420 lower 2013/2014 Miers Bottom Summer 2.372 0.791 1.581 middle 2013/2014 Miers Bottom Summer 3.872 0.782 3.090 upper 2013/2014 Miers Middle Summer 5.799 1.969 3.830 2013/2014 Continued

114

Table 7 Continued. Miers Top Summer 2.458 0.693 1.765 2013/2014 Miers Bottom Summer 2.650 1.627 1.023 upper 2014/2015 Miers Middle Summer 2.907 1.186 1.721 2014/2015 Taylor Bottom Winter 2014 0.513 0.311 0.203 Glacier lower Taylor Bottom Winter 2014 0.670 0.413 0.258 Glacier middle Taylor Bottom Winter 2014 0.612 0.333 0.279 Glacier upper Taylor Middle Winter 2014 0.889 0.406 0.483 Glacier Victoria Bottom Winter 2013 1.994 0.368 1.626 lower Victoria Bottom Winter 2013 2.105 0.388 1.717 middle Victoria Bottom Winter 2013 1.486 0.251 1.235 upper Victoria Middle Winter 2013 2.330 0.294 2.036

Victoria Top Winter 2013 1.990 0.487 1.503

115

Table 8. Major oxides in the bulk solid fraction of MDV aeolian material, determined by XRF.

Site Height Year SiO2 Al2O3 Fe2O3 MgO CaO K2O Na2O TiO2 MnO P2O5 Cr2O3 V2O5 (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) Alatna Bottom Winter 80.6 5.55 5.6 2.36 3.96 0.62 0.84 0.51 0.09 0.05 0.04 0.03 lower 2013 East Lake Bottom Winter 63.7 12.4 6.31 5.68 6.23 2.16 2.43 0.52 0.1 0.08 0.05 0.03 Bonney lower 2014 (ELB) East Lake Bottom Winter 64.9 12.4 5.98 5.33 5.98 2.14 2.43 0.47 0.1 0.08 0.06 0.02 Bonney middle 2014 (ELB) East Lake Bottom Winter 64.3 12.5 6.08 5.68 6.01 2.15 2.6 0.5 0.1 0.09 0.04 0.03 Bonney upper 2014 (ELB) East Lake Middle Winter 65 12.6 5.97 5.18 6.14 2.1 2.45 0.46 0.11 0.08 0.07 0.02 116 Bonney 2014

(ELB) East Lake Top Winter 65.3 12.7 5.54 5.14 5.72 2.27 2.67 0.42 0.09 0.08 0.04 0.01 Bonney 2014 (ELB) East Lake Bottom 2015 64.5 12.3 6.14 5.95 5.99 2.1 2.59 0.47 0.1 0.08 0.04 0.03 Bonney lower (ELB) East Lake Bottom 2015 63.6 12.8 6.16 5.28 6.39 2.16 2.49 0.54 0.1 0.09 0.04 0.02 Bonney middle (ELB) East Lake Bottom 2015 63.3 12.5 6.34 5.81 6.08 2.14 2.62 0.53 0.12 0.1 0.04 0.03 Bonney upper (ELB) East Lake Middle 2015 63.5 12.2 6.59 5.92 6.33 2.04 2.4 0.53 0.11 0.09 0.04 0.03 Bonney (ELB) Continued

Table 8 Continued.

East Lake Top 2015 63.4 12.4 5.98 5.49 5.88 2.18 2.63 0.48 0.1 0.08 0.05 0.02 Bonney (ELB) Explorer's Bottom 2015 59.8 13.7 6.46 4.79 5.92 2.53 3.56 1.13 0.11 0.28 0.05 0.02 Cove A lower Explorer's Bottom 2015 60.2 13.7 6.49 5.19 6.03 2.5 3.38 1.15 0.12 0.28 0.04 0.02 Cove A middle Explorer's Bottom 2015 59.4 13.6 6.95 5.46 6.29 2.41 3.27 1.33 0.12 0.31 0.03 0.02 Cove A upper Explorer's Middle 2015 59.3 13.5 6.81 5.57 6.35 2.35 3.2 1.24 0.12 0.29 0.04 0.02 Cove A Explorer's Top 2015 58.6 13.3 7.07 5.74 6.46 2.31 3.15 1.3 0.12 0.31 0.07 0.03 Cove A Explorer's Bottom Winter 59.9 13.4 6.73 5.35 6.28 2.39 3.21 1.24 0.11 0.3 0.05 0.02

117 Cove B lower 2014 Explorer's Bottom Winter 59.7 13.5 6.69 5.4 6.24 2.37 3.2 1.25 0.11 0.29 0.03 0.02

Cove B middle 2014 Explorer's Bottom Winter 60.7 13.5 6.29 5.14 5.99 2.44 3.27 1.15 0.11 0.27 0.03 0.01 Cove B upper 2014 Explorer's Middle Winter 60.4 13.2 6.79 5.37 6.09 2.42 3.34 1.18 0.1 0.28 0.06 0.03 Cove B 2014 Explorer's Top Winter 60.2 13.9 5.93 4.5 5.62 2.6 3.58 1.09 0.1 0.28 0.04 0.01 Cove B 2014 Explorer's Bottom 2015 62.2 13.8 5.6 4.45 5.55 2.58 3.47 0.97 0.1 0.25 0.03 0.01 Cove B lower Explorer's Bottom 2015 58.8 13.1 7.01 5.63 6.24 2.34 3.23 1.3 0.12 0.3 0.03 0.03 Cove B middle Explorer's Bottom 2015 60.3 13.4 6.55 5.23 6.06 2.43 3.29 1.16 0.11 0.29 0.03 0.02 Cove B upper Explorer's Middle 2015 59.5 13.4 6.73 5.53 6.3 2.38 3.16 1.23 0.12 0.29 0.03 0.03 Cove B Explorer's Top 2015 59.1 13.4 6.95 5.51 6.36 2.36 3.22 1.26 0.11 0.29 0.06 0.02 Cove B Continued

Table 8 Continued. F6 (Lake Bottom Winter 59.3 12.4 7.48 6.45 6.83 2.06 2.94 1.12 0.12 0.24 0.05 0.03 Fryxell) upper 2014 F6 (Lake Middle Winter 57.7 12.7 7.05 5.81 6.51 2.18 3.1 1.12 0.13 0.24 0.03 0.03 Fryxell) 2014 F6 (Lake Top Winter 60.6 12.2 7.06 6.46 6.5 2.04 2.74 0.91 0.13 0.19 0.05 0.02 Fryxell) 2014 Miers Bottom Summer 59.3 13.4 6.23 5.2 7.01 2.55 3.22 1.24 0.1 0.24 0.06 <0.01 upper 2013/2014 Miers Middle Summer 62.4 14.2 5.09 4.06 6.08 2.82 3.42 0.98 0.07 0.21 0.06 0.02 2013/2014 Miers Top Summer 63 14.4 4.64 4.01 5.64 2.78 3.51 0.89 0.08 0.19 0.03 <0.01 2013/2014 Miers Bottom Winter 59 13.5 6.18 5.42 6.99 2.59 3.26 1.21 0.09 0.24 0.04 0.03 upper 2014

118 Victoria Middle Winter 67.1 7.9 8.1 5.74 7.89 0.92 1.21 0.62 0.16 0.07 0.03 0.05 2013

Victoria Top Winter 65.5 7.86 8.71 5.97 8.12 0.91 1.16 0.75 0.17 0.06 0.05 0.05 2013

Table 9. Trace elements in the bulk solid fraction of MDV aeolian material, determined by ICP-AES/ICP-MS.

Site Height Year Ba Be Cr Cu Li Mn Ni Sc Sr V Zn Ag As (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) Alatna Bottom Winter 144 <5 212 40 <10 703 35 20 48 147 140 <1 <5 lower 2013 East Lake Bottom Winter 466 <5 268 21 11 804 68 20 283 133 155 <1 <5 Bonney lower 2014 (ELB) East Lake Bottom Winter 490 <5 326 20 11 813 68 20 290 121 205 <1 <5 Bonney middle 2014 (ELB) East Lake Bottom Winter 483 <5 242 20 <10 831 72 21 313 121 80 <1 <5 Bonney upper 2014 (ELB) East Lake Middle Winter 481 <5 465 25 <10 819 70 20 301 121 135 <1 <5 119 Bonney 2014

(ELB) East Lake Top Winter 517 <5 309 18 <10 765 62 19 320 111 107 <1 <5 Bonney 2014 (ELB) East Lake Bottom 2015 465 <5 269 18 11 859 75 22 307 123 94 <1 <5 Bonney lower (ELB) East Lake Bottom 2015 479 <5 259 21 11 838 62 21 300 126 129 <1 <5 Bonney middle (ELB) East Lake Bottom 2015 504 <5 273 21 10 876 77 21 324 132 82 <1 <5 Bonney upper (ELB) East Lake Middle 2015 465 <5 320 22 11 908 70 22 285 136 90 <1 <5 Bonney (ELB) Continued

Table 9 Continued.

East Lake Top 2015 506 <5 262 18 11 810 70 20 315 123 104 <1 <5 Bonney (ELB) Explorer's Bottom 2015 643 <5 380 18 14 883 73 13 573 107 313 <1 <5 Cove A lower Explorer's Bottom 2015 624 <5 248 18 13 857 85 14 550 109 132 <1 <5 Cove A middle Explorer's Bottom 2015 582 <5 242 21 13 914 81 15 565 117 175 <1 <5 Cove A upper Explorer's Middle 2015 580 <5 247 20 13 895 90 15 561 115 98 <1 <5 Cove A Explorer's Top 2015 564 <5 302 21 13 938 91 16 555 121 107 <1 <5 Cove A Explorer's Bottom Winter 576 <5 230 19 11 874 89 14 547 113 212 <1 <5

120 Cove B lower 2014 Explorer's Bottom Winter 575 <5 254 20 12 900 88 14 568 112 285 <1 <5

Cove B middle 2014 Explorer's Bottom Winter 599 <5 231 18 12 829 84 13 553 105 239 <1 <5 Cove B upper 2014 Explorer's Middle Winter 590 <5 395 19 13 899 87 14 549 111 329 <1 <5 Cove B 2014 Explorer's Top Winter 642 <5 268 19 14 777 71 12 555 97 402 <1 <5 Cove B 2014 Explorer's Bottom 2015 641 <5 223 15 11 754 72 12 546 94 118 <1 <5 Cove B lower Explorer's Bottom 2015 547 <5 256 20 12 915 89 15 545 116 234 <1 <5 Cove B middle Explorer's Bottom 2015 577 <5 267 17 11 857 80 14 548 106 125 <1 <5 Cove B upper Explorer's Middle 2015 543 <5 230 19 12 882 86 14 547 112 87 <1 <5 Cove B Explorer's Top 2015 560 <5 392 19 11 896 86 15 545 113 102 <1 <5 Cove B Continued

Table 9 Continued. F6 (Lake Bottom Winter 480 <5 341 20 11 982 97 20 445 142 287 <1 <5 Fryxell) upper 2014 F6 (Lake Middle Winter 506 <5 269 21 13 934 85 18 476 131 387 <1 <5 Fryxell) 2014 F6 (Lake Top Winter 475 <5 330 17 11 961 100 20 424 135 314 <1 <5 Fryxell) 2014 Miers Bottom Summer 546 <5 359 22 14 766 110 12 562 102 84 <1 <5 upper 2013/2014 Miers Middle Summer 645 <5 356 18 13 620 80 9 568 82 71 <1 <5 2013/2014 Miers Top Summer 629 <5 262 17 12 559 83 8 550 75 75 <1 <5 2013/2014 Miers Bottom Winter 571 <5 284 21 15 753 110 11 566 99 90 <1 <5 upper 2014

121 Victoria Middle Winter 205 <5 244 34 <10 1137 65 34 126 224 368 <1 <5 2013

Victoria Top Winter 203 <5 297 37 <10 1190 71 34 113 253 328 <1 <5 2013

Table 9 Continued.

Site Height Year Bi Cd Co Cs Ga Ge Hf In Mo Nb Pb Rb (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) Alatna Bottom Winter <0.1 <0.2 20.7 0.7 7 1 2 <0.2 9 3 13 22.5 lower 2013 East Lake Bottom Winter 0.8 <0.2 24.6 1.2 13 1 3 <0.2 4 8 751 71.8 Bonney lower 2014 (ELB) East Lake Bottom Winter <0.1 <0.2 25.6 1.2 14 1 3 <0.2 2 7 12 76.5 Bonney middle 2014 (ELB) East Lake Bottom Winter <0.1 <0.2 27 1.1 15 1 2 <0.2 4 7 10 77.3 Bonney upper 2014 (ELB) East Lake Middle Winter <0.1 <0.2 25.4 1.2 15 1 3 <0.2 <2 8 15 75.6 122 Bonney 2014

(ELB) East Lake Top Winter <0.1 <0.2 24.7 1.1 14 1 3 <0.2 <2 7 12 79.7 Bonney 2014 (ELB) East Lake Bottom 2015 <0.1 <0.2 27.4 1.1 14 1 3 <0.2 2 8 12 76.3 Bonney lower (ELB) East Lake Bottom 2015 <0.1 <0.2 25.7 1.3 15 1 3 <0.2 <2 10 11 78.8 Bonney middle (ELB) East Lake Bottom 2015 <0.1 <0.2 27.8 1.1 15 1 3 <0.2 <2 7 12 77.5 Bonney upper (ELB) East Lake Middle 2015 <0.1 <0.2 28.4 1.2 14 2 3 <0.2 <2 9 11 73.5 Bonney (ELB) Continued

Table 9 Continued.

East Lake Top 2015 <0.1 <0.2 25.7 1.2 14 1 3 <0.2 4 8 11 79.8 Bonney (ELB) Explorer's Bottom 2015 <0.1 <0.2 22.8 1.3 18 1 5 <0.2 2 38 16 78.2 Cove A lower Explorer's Bottom 2015 <0.1 <0.2 23.7 1.1 17 1 5 <0.2 <2 37 12 74.7 Cove A middle Explorer's Bottom 2015 <0.1 <0.2 26.1 1.1 17 1 5 <0.2 2 34 12 70.2 Cove A upper Explorer's Middle 2015 <0.1 <0.2 26.1 1.1 17 1 5 <0.2 2 40 13 69.3 Cove A Explorer's Top 2015 <0.1 0.3 26.9 1.1 17 1 5 <0.2 2 41 10 66.7 Cove A Explorer's Bottom Winter <0.1 <0.2 25.6 1.1 17 1 5 <0.2 2 39 23 69.9

123 Cove B lower 2014 Explorer's Bottom Winter <0.1 <0.2 25.9 1.1 17 1 5 <0.2 2 41 23 69.9

Cove B middle 2014 Explorer's Bottom Winter <0.1 <0.2 23.7 1.1 17 1 5 <0.2 4 38 15 71 Cove B upper 2014 Explorer's Middle Winter <0.1 <0.2 25.3 1.2 17 1 5 <0.2 2 39 40 71.8 Cove B 2014 Explorer's Top Winter <0.1 <0.2 21.4 1.3 17 1 5 <0.2 <2 36 25 78.6 Cove B 2014 Explorer's Bottom 2015 <0.1 <0.2 20.9 1.2 17 1 4 <0.2 2 34 13 78.5 Cove B lower Explorer's Bottom 2015 <0.1 <0.2 27.3 1.2 17 1 5 <0.2 2 41 105 69 Cove B middle Explorer's Bottom 2015 <0.1 <0.2 24.6 1.1 17 1 5 <0.2 2 38 13 72.7 Cove B upper Explorer's Middle 2015 <0.1 <0.2 26.2 1.1 17 1 5 <0.2 4 32 12 68.2 Cove B Explorer's Top 2015 <0.1 <0.2 26.2 1.2 17 1 5 <0.2 <2 40 12 69.1 Cove B Continued

Table 9 Continued. F6 (Lake Bottom Winter <0.1 <0.2 30.1 1 15 1 5 <0.2 <2 27 12 63.1 Fryxell) upper 2014 F6 (Lake Middle Winter <0.1 <0.2 28 1.2 16 1 5 <0.2 <2 35 52 67 Fryxell) 2014 F6 (Lake Top Winter <0.1 <0.2 30.2 1 15 1 4 <0.2 <2 25 49 63.8 Fryxell) 2014 Miers Bottom Summer <0.1 <0.2 24.6 1.6 18 1 5 <0.2 2 33 14 80.1 upper 2013/2014 Miers Middle Summer <0.1 <0.2 19.4 1.3 18 1 4 <0.2 <2 25 15 85.1 2013/2014 Miers Top Summer <0.1 <0.2 18.3 1.2 18 1 4 <0.2 5 23 19 84.6 2013/2014 Miers Bottom Winter <0.1 <0.2 24.7 1.5 17 1 4 <0.2 <2 30 14 82.8 upper 2014

124 Victoria Middle Winter <0.1 <0.2 35 0.8 9 2 3 <0.2 <2 5 44 32.1 2013

Victoria Top Winter <0.1 <0.2 36.6 0.9 10 2 4 <0.2 <2 6 21 31.1 2013

Table 9 Continued.

Site Height Year Sb Sn Ta Th Tl U W Y Zr (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm)

Alatna Bottom Winter 2013 0.1 5 <0.5 2.7 <0.5 0.77 <1 13.5 72.9 lower

East Lake Bottom Winter 2014 3.9 428 <0.5 6.4 <0.5 0.88 <1 16.4 102 Bonney lower (ELB)

East Lake Bottom Winter 2014 0.2 3 <0.5 4.2 <0.5 0.87 <1 15.9 90.2 Bonney middle (ELB)

125 East Lake Bottom Winter 2014 <0.1 2 <0.5 4.9 <0.5 0.88 <1 16.4 91.8 Bonney upper (ELB)

East Lake Middle Winter 2014 <0.1 5 <0.5 6.1 <0.5 0.89 1 15.5 90.4 Bonney (ELB)

East Lake Top Winter 2014 <0.1 1 <0.5 4.6 <0.5 0.81 <1 14.7 93.9 Bonney (ELB)

East Lake Bottom 2015 <0.1 2 <0.5 5.2 <0.5 0.94 <1 16.2 90 Bonney lower (ELB)

East Lake Bottom 2015 <0.1 2 0.5 6.2 <0.5 1.03 <1 18.3 109 Bonney middle (ELB) Continued

Table 9 Continued.

East Lake Bottom 2015 0.1 3 <0.5 4.7 <0.5 0.91 <1 17.1 96.4 Bonney upper (ELB)

East Lake Middle 2015 <0.1 4 <0.5 5.5 <0.5 0.9 <1 17.5 95.9 Bonney (ELB)

East Lake Top 2015 0.1 2 <0.5 4.4 <0.5 0.89 <1 15.1 90.9 Bonney (ELB)

Explorer's Bottom 2015 0.2 5 2.4 6.5 <0.5 1.42 <1 19.7 202 Cove A lower

126 Explorer's Bottom 2015 <0.1 4 2.2 6.3 <0.5 1.31 <1 19.4 193 Cove A middle

Explorer's Bottom 2015 <0.1 4 2.5 7.8 <0.5 1.38 <1 21.4 194 Cove A upper

Explorer's Middle 2015 0.1 5 2.4 6.6 <0.5 1.4 <1 21 203 Cove A

Explorer's Top 2015 0.2 3 2.5 6.2 <0.5 1.31 <1 21.7 195 Cove A

Explorer's Bottom Winter 2014 <0.1 5 2.4 5.9 <0.5 1.36 <1 20.2 187 Cove B lower

Explorer's Bottom Winter 2014 0.1 9 2.5 6.4 <0.5 1.37 <1 20.8 198 Cove B middle Continued

Table 9 Continued.

Explorer's Bottom Winter 2014 <0.1 12 2.4 6 <0.5 1.35 <1 19.1 179 Cove B upper

Explorer's Middle Winter 2014 0.1 22 2.4 6.1 <0.5 1.31 <1 19.8 189 Cove B

Explorer's Top Winter 2014 0.2 16 2.3 5.7 <0.5 1.38 <1 18.1 193 Cove B

Explorer's Bottom 2015 0.1 2 2 6.8 <0.5 1.28 <1 17.1 173 Cove B lower

Explorer's Bottom 2015 0.1 67 2.4 6.6 <0.5 1.47 <1 20.7 204 Cove B middle

127 Explorer's Bottom 2015 0.1 3 2.3 6 <0.5 1.27 <1 19.8 186

Cove B upper

Explorer's Middle 2015 <0.1 2 2.3 6 <0.5 1.4 <1 20.5 204 Cove B

Explorer's Top 2015 0.1 3 2.4 6.9 <0.5 1.35 <1 21 202 Cove B

F6 (Lake Bottom Winter 2014 0.1 3 1.8 5.9 <0.5 1.22 <1 21.4 177 Fryxell) upper

F6 (Lake Middle Winter 2014 0.3 30 2.1 6.4 <0.5 1.25 <1 21 186 Fryxell)

F6 (Lake Top Winter 2014 0.2 27 1.7 6.2 <0.5 1.08 <1 20.2 142 Fryxell)

Continued

Table 9 Continued.

Miers Bottom Summer <0.1 3 2 7.6 <0.5 1.48 <1 20.7 181 upper 2013/2014

Miers Middle Summer <0.1 2 1.5 6.6 <0.5 1.22 <1 17.2 153 2013/2014

Miers Top Summer <0.1 3 1.4 8.1 <0.5 1.1 <1 15.6 141 2013/2014

Miers Bottom Winter 2014 0.1 3 1.9 6.6 <0.5 1.44 <1 20.8 158 upper

Victoria Middle Winter 2013 0.2 17 <0.5 3.1 <0.5 0.84 <1 17.9 83.7

Victoria Top Winter 2013 0.2 6 <0.5 3.3 <0.5 0.87 <1 18.5 123

128

Table 10. Rare earth elements (REE) in the bulk solid fraction of MDV aeolian material, determined by ICP-AES/ICP-MS. Site Height Year La Ce Pr Nd Sm Eu Gd (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) Alatna Bottom Winter 9.5 19.8 2.37 9.5 2.2 0.52 2.26 lower 2013 East Lake Bottom Winter 25.4 50.5 5.59 20.2 3.8 0.9 3.4 Bonney lower 2014 (ELB) East Lake Bottom Winter 18.6 37.1 4.17 16.3 3.5 0.92 2.98 Bonney middle 2014 (ELB) East Lake Bottom Winter 22 43.3 4.89 18.9 3.7 0.94 3.3 Bonney upper 2014 (ELB) East Lake Middle Winter 26 50.1 5.32 20.3 3.7 0.95 3.13 Bonney 2014 (ELB) East Lake Top Winter 20.5 39.6 4.44 17.2 3.4 0.97 2.91 Bonney 2014 (ELB) East Lake Bottom 2015 21.6 41.9 4.58 17.6 3.6 0.94 3.29 Bonney lower (ELB) East Lake Bottom 2015 24.8 49.9 5.53 21.6 4.4 0.97 3.79 Bonney middle (ELB) East Lake Bottom 2015 20.7 41.1 4.64 18.1 3.7 1.03 3.35 Bonney upper (ELB) East Lake Middle 2015 24.5 48.1 5.29 20.4 4 0.92 3.4 Bonney (ELB) East Lake Top 2015 19.2 36.9 4.29 16.8 3.4 0.92 3.04 Bonney (ELB) Explorer's Bottom 2015 36.6 71 8.02 31.4 5.8 1.64 4.65 Cove A lower Explorer's Bottom 2015 37.4 73.4 7.98 30.7 5.8 1.59 4.61 Cove A middle Explorer's Bottom 2015 42.9 82.5 9.06 35.6 6.8 1.75 5.38 Cove A upper Explorer's Middle 2015 39.3 77.2 8.54 33.1 6.1 1.7 4.89 Cove A Explorer's Top 2015 37.7 74.7 8.46 32.4 6.1 1.71 5.05 Cove A Explorer's Bottom Winter 36.6 71 8.01 30.4 5.9 1.72 4.75 Cove B lower 2014 Explorer's Bottom Winter 38.5 75.3 8.3 32.1 6 1.67 4.73 Cove B middle 2014 Explorer's Bottom Winter 35.7 69.6 7.82 29.9 5.6 1.7 4.57 Cove B upper 2014 Continued

129

Table 10 Continued. Explorer's Middle Winter 37.1 72.1 7.98 30.8 5.8 1.65 4.54 Cove B 2014 Explorer's Top Winter 35.8 69.4 7.81 29.5 5.6 1.65 4.39 Cove B 2014 Explorer's Bottom 2015 38.8 74.4 8.07 30.2 5.5 1.51 4.2 Cove B lower Explorer's Bottom 2015 40 77.8 8.58 33 6 1.76 4.78 Cove B middle Explorer's Bottom 2015 37.1 73.1 8.01 30.6 5.6 1.65 4.57 Cove B upper Explorer's Middle 2015 37.2 73.3 8.18 30.8 6 1.68 4.82 Cove B Explorer's Top 2015 40.8 78.7 8.71 33 6.2 1.7 4.94 Cove B F6 (Lake Bottom Winter 34.4 68.4 7.59 29.9 5.7 1.46 4.6 Fryxell) upper 2014 F6 (Lake Middle Winter 37.1 73.4 8.11 31.1 5.8 1.51 4.75 Fryxell) 2014 F6 (Lake Top Winter 32.7 64.5 7.2 27.2 5.2 1.34 4.23 Fryxell) 2014 Miers Bottom Summer 37.9 74 8.21 31 6.1 1.52 4.85 upper 2013/2014 Miers Middle Summer 33.8 64.7 7.04 26.9 4.9 1.39 4.04 2013/2014 Miers Top Summer 43.5 80 8.13 29.7 5.2 1.35 3.82 2013/2014 Miers Bottom Winter 35.4 69.3 7.68 30.4 5.7 1.47 4.7 upper 2014 Victoria Middle Winter 11.5 24.2 2.87 11.9 2.8 0.62 2.83 2013 Victoria Top Winter 11.8 24.9 3.01 12.4 2.8 0.62 3.04 2013

130

Table 10 Continued.

Site Height Year Tb Dy Ho Er Tm Yb Lu (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) Alatna Bottom Winter 0.33 2.43 0.49 1.54 0.21 1.7 0.24 lower 2013 East Lake Bottom Winter 0.45 2.95 0.6 1.69 0.26 1.8 0.26 Bonney lower 2014 (ELB) East Lake Bottom Winter 0.43 3.02 0.54 1.63 0.25 1.8 0.24 Bonney middle 2014 (ELB) East Lake Bottom Winter 0.45 3.07 0.57 1.71 0.27 1.8 0.28 Bonney upper 2014 (ELB) East Lake Middle Winter 0.42 3.04 0.52 1.69 0.25 1.7 0.24 Bonney 2014 (ELB) East Lake Top Winter 0.42 2.91 0.52 1.63 0.24 1.6 0.26 Bonney 2014 (ELB) East Lake Bottom 2015 0.47 3.22 0.57 1.81 0.27 1.7 0.25 Bonney lower (ELB) East Lake Bottom 2015 0.56 3.68 0.66 1.96 0.29 1.9 0.24 Bonney middle (ELB) East Lake Bottom 2015 0.47 3.33 0.62 1.89 0.26 1.9 0.26 Bonney upper (ELB) East Lake Middle 2015 0.5 3.4 0.59 1.87 0.28 1.9 0.29 Bonney (ELB) East Lake Top 2015 0.41 2.99 0.54 1.7 0.26 1.7 0.25 Bonney (ELB) Explorer's Bottom 2015 0.64 4.09 0.72 2.11 0.29 1.9 0.28 Cove A lower Explorer's Bottom 2015 0.65 4.03 0.68 2 0.28 2 0.29 Cove A middle Explorer's Bottom 2015 0.73 4.57 0.77 2.18 0.31 2 0.31 Cove A upper Explorer's Middle 2015 0.68 4.33 0.72 2.16 0.31 2.1 0.32 Cove A Explorer's Top 2015 0.7 4.42 0.79 2.23 0.33 2.1 0.31 Cove A Explorer's Bottom Winter 0.62 4.18 0.72 2.13 0.3 2.1 0.26 Cove B lower 2014 Explorer's Bottom Winter 0.63 4.23 0.7 2.1 0.31 2 0.3 Cove B middle 2014 Explorer's Bottom Winter 0.63 4.04 0.67 2.01 0.29 1.9 0.29 Cove B upper 2014 Continued

131

Table 10 Continued. Explorer's Middle Winter 0.65 4.13 0.71 2.02 0.29 2 0.29 Cove B 2014 Explorer's Top Winter 0.58 3.89 0.65 1.9 0.27 1.8 0.32 Cove B 2014 Explorer's Bottom 2015 0.54 3.51 0.6 1.77 0.26 1.7 0.25 Cove B lower Explorer's Bottom 2015 0.7 4.27 0.72 2.11 0.3 2 0.31 Cove B middle Explorer's Bottom 2015 0.64 4.07 0.69 2.04 0.28 1.9 0.31 Cove B upper Explorer's Middle 2015 0.7 4.07 0.72 2.08 0.31 2.1 0.29 Cove B Explorer's Top 2015 0.68 4.25 0.72 2.2 0.32 2.1 0.29 Cove B F6 (Lake Bottom Winter 0.69 4.37 0.74 2.25 0.31 2.2 0.29 Fryxell) upper 2014 F6 (Lake Middle Winter 0.65 4.26 0.75 2.16 0.32 2.1 0.32 Fryxell) 2014 F6 (Lake Top Winter 0.62 4.04 0.71 2.14 0.31 2.1 0.27 Fryxell) 2014 Miers Bottom Summer 0.66 4.27 0.72 2.1 0.31 2 0.3 upper 2013/2014 Miers Middle Summer 0.54 3.34 0.6 1.78 0.25 1.7 0.27 2013/2014 Miers Top Summer 0.51 3.19 0.55 1.54 0.22 1.5 0.24 2013/2014 Miers Bottom Winter 0.69 4.37 0.72 2.25 0.33 2.1 0.31 upper 2014 Victoria Middle Winter 0.46 3.26 0.63 2.02 0.31 2.1 0.34 2013 Victoria Top Winter 0.45 3.47 0.65 2.08 0.32 2.2 0.31 2013

132

Table 11. Total sediment flux from previous studies in the McMurdo Dry Valleys. N and P fluxes were calculated using 100 cm sediment flux data from Lancaster (2002) and ~100 cm collection height chemistry from this study*, as well as 20 cm sediment flux data from (Šabacká et al., 2012) and ~30 cm collection height chemistry data from this study**. Total Soluble N Soluble P Height of Years of Location sediment flux flux (µmol flux (µmol collection (cm) collection (g m-2 a-1) m-2 a-1) m-2 a-1) Lancaster 2002

Explorer’s Cove 27.87 100 1997-2000 0.94* 0.34*

Lake Fryxell 1.00 100 1997-2000 0.21* 0.00*

Lake Hoare 0.86 100 1997-2000 - -

Lake Bonney 110.53 100 1997-2000 4.04* 0.20*

Lake Brownworth 441.80 100 1997-2000 - -

Lake Vanda 227.50 100 1997-2000 - -

Lake Vida 39.56 100 1997-2000 - -

Commonwealth 0.26 100 1997-2000 - - Glacier Canada Glacier 0.43 100 1997-2000 - -

Howard Glacier 0.47 100 1997-2000 - -

Taylor Glacier 3.73 100 1997-2000 1.53* 0.00*

Šabacká et al. 2012

Lake Fryxell 72.3 30 1998-2008 7.74** 0.23**

Lake Hoare 289 30 1999-2008

West Lake Bonney 8.50 30 1999-2008 0.48** 0.01**

Atkins and Dunbar 2009 McMurdo Sound 7.76-24.5 0 2006-2007 - -

133

Appendix A: Detection limits and Precision and Accuracy:

134

Table 12. Accuracy and precision for analytes. Precision calculated from sample duplicates, which were used for standard error bars. Filter blanks were used to determine background concentrations of analytes. Dashes represent measurements below the detection limit. Analyte Accuracy Precision Filter blank (µmol g-1) (*mg) Fluoride 4.6% 6.9% - Chloride 0.18% 0.58% 0.015 Bromide 19% 21% - Sulfate 5.4% 0.69% - Sodium 10% 0.13% 0.000039 Potassium 2.6% 4.6% 0.0043 Magnesium 5.7% 2.2% 0.000062 Calcium 0.88% 5.0% 0.0016 Nitrate 6.8% 1.7% 0.00038 Phosphate 3.6% 4.0% 0.000027 Total carbon* 2.4% 0.83% 0.0014 Organic 2.4% 2.7% 0.0018 Carbon*

135

Table 13. Detection limits for oxides, trace elements, and rare earth elements (REE). Analyte Method Detection limit Unit

Al GE_ICM90A 0.01 %

Ba GE_ICM90A 10 ppm

Be GE_ICM90A 5 ppm

Ca GE_ICM90A 0.1 %

Cr GE_ICM90A 10 ppm

Cu GE_ICM90A 10 ppm

Fe GE_ICM90A 0.01 %

K GE_ICM90A 0.1 %

Li GE_ICM90A 10 ppm

Mg GE_ICM90A 0.01 %

Mn GE_ICM90A 10 ppm

Ni GE_ICM90A 5 ppm

P GE_ICM90A 0.01 %

Sc GE_ICM90A 5 ppm

Si GE_ICM90A 0.1 %

Sr GE_ICM90A 10 ppm

Ti GE_ICM90A 0.01 %

V GE_ICM90A 5 ppm

Zn GE_ICM90A 5 ppm

Ag GE_ICM90A 1 ppm

As GE_ICM90A 5 ppm

Bi GE_ICM90A 0.1 ppm

Cd GE_ICM90A 0.2 ppm

Ce GE_ICM90A 0.1 ppm Continued

136

Table 13 Continued. Co GE_ICM90A 0.5 ppm

Cs GE_ICM90A 0.1 ppm

Dy GE_ICM90A 0.05 ppm

Er GE_ICM90A 0.05 ppm

Eu GE_ICM90A 0.05 ppm

Ga GE_ICM90A 1 ppm

Gd GE_ICM90A 0.05 ppm

Ge GE_ICM90A 1 ppm

Hf GE_ICM90A 1 ppm

Ho GE_ICM90A 0.05 ppm

In GE_ICM90A 0.2 ppm

La GE_ICM90A 0.1 ppm

Lu GE_ICM90A 0.05 ppm

Mo GE_ICM90A 2 ppm

Nb GE_ICM90A 1 ppm

Nd GE_ICM90A 0.1 ppm

Pb GE_ICM90A 5 ppm

Pr GE_ICM90A 0.05 ppm

Rb GE_ICM90A 0.2 ppm

Sb GE_ICM90A 0.1 ppm

Sm GE_ICM90A 0.1 ppm

Sn GE_ICM90A 1 ppm

Ta GE_ICM90A 0.5 ppm

Tb GE_ICM90A 0.05 ppm

Th GE_ICM90A 0.1 ppm

Tl GE_ICM90A 0.5 ppm Continued

137

Table 13 Continued. Tm GE_ICM90A 0.05 ppm

U GE_ICM90A 0.05 ppm

W GE_ICM90A 1 ppm

Y GE_ICM90A 0.5 ppm

Yb GE_ICM90A 0.1 ppm

Zr GE_ICM90A 0.5 ppm

SiO2 GO_XRF76V 0.01 %

Al2O3 GO_XRF76V 0.01 %

Fe2O3 GO_XRF76V 0.01 %

MgO GO_XRF76V 0.01 %

CaO GO_XRF76V 0.01 %

K2O GO_XRF76V 0.01 %

Na2O GO_XRF76V 0.01 %

TiO2 GO_XRF76V 0.01 %

MnO GO_XRF76V 0.01 %

P2O5 GO_XRF76V 0.01 %

Cr2O3 GO_XRF76V 0.01 %

V2O5 GO_XRF76V 0.01 %

138

Appendix B: Total Mass of Aeolian Samples

139

Table 14. Total mass (grams) of aeolian material collected from each location. Starred values are those which were collected with snow and ice, therefore the mass includes the mass of the associated water and is not solely representative of sediment mass. Winter Summer Winter Summer 2015 Full Location Height 2013 2013/2014 2014 2014/2015 Year Alatna Bottom lower 5.3* - - - -

Alatna Bottom middle 6.6 - - - -

Alatna Middle 5.1* - - - -

East Lake Bottom lower - - 8.3 - 44.9 Bonney East Lake Bottom middle - - 6.8 - 73.0 Bonney East Lake Bottom upper - - 69.1 - 55.8 Bonney East Lake Middle - - 40.7 - 37.2 Bonney East Lake Top - - 12.4 - 19.7 Bonney Explorer's Cove Bottom lower - - - - 9.1 A Explorer's Cove Bottom middle - - - - 21.2 A Explorer's Cove Bottom upper - - - - 22.5 A Explorer's Cove Middle - 1.5 - - 28.7 A Explorer's Cove Top - - - - 22.3 A Explorer's Cove Bottom lower - - 139.1* - 19.9 B Explorer's Cove Bottom middle - - 149.1* - 10.0 B Explorer's Cove Bottom upper - - 130.2* 1.2 44.8 B Explorer's Cove Middle - - 35.4* - 27.1 B Explorer's Cove Top - - 18.6* - 16.3 B F6 (Lake Bottom lower - 3.6 - - - Fryxell) F6 (Lake Bottom middle - 5.0 - - - Fryxell) F6 (Lake Bottom upper - 5.8 16.0 - - Fryxell) F6 (Lake Middle - 6.8 15.2 - - Fryxell) F6 (Lake Top - 3.2 2.3 - - Fryxell) Continued

140

Table 14 Continued. Miers Bottom lower - 14.1 - - -

Miers Bottom middle - 14.8 - - -

Miers Bottom upper - 51.3* 6.1 - -

Miers Middle - 36.1* 2.8 - -

Miers Top - 16.3 - - -

Taylor Glacier Bottom lower - - - 0.1 -

Taylor Glacier Bottom middle - - - 0.1 -

Taylor Glacier Bottom upper - - - 11.1* -

Taylor Glacier Middle - - - 6.4* -

Victoria Bottom lower 2.7 - - - -

Victoria Bottom middle 2.5 - - - -

Victoria Bottom upper 86.6* - - - -

Victoria Middle 65.0* - - - -

Victoria Top 7.8 - - - -

141

Appendix C: Analysis of Variance (ANOVA) Results

142

Table 15. One-way Analysis of Variance (ANOVA) using Fisher’s Least Significant Difference (LSD) method for ions. Variables are site, year, and height above the surface. *** is 99.9% confidence, ** is 99%, * is 95% Analyte Variable F- value p- value Significant?

F Site 9.67 <0.001*** Y

F Year 1.99 0.111 N

F Height 0.71 0.589 N

Cl Site 6.67 <0.001*** Y

Cl Year 2.21 0.082 N

Cl Height 1.335 0.271 N

NO3 Site 22.6 <0.001*** Y

NO3 Year 7.784 <0.001*** Y

NO3 Height 0.236 0.916 N

PO4 Site 12.7 <0.001*** Y

PO4 Year 5.83 0.001*** Y

PO4 Height 0.225 0.923 N

SO4 Site 6.85 <0.001*** Y

SO4 Year 1.38 0.254 N

SO4 Height 0.572 0.684 N

Na Site 9.8 <0.001*** Y

Na Year 2.34 0.0688 N

Continued

143

Table 15 Continued.

Na Height 1.14 0.349 N

K Site 12.18 <0.001*** Y

K Year 2.90 0.0315** Y

K Height 1.42 0.242 N

Mg Site 7.67 <0.001*** Y

Mg Year 1.03 0.403 N

Mg Height 0.713 0.587 N

Ca Site 9.17 <0.001*** Y

Ca Year 0.626 0.646 N

Ca Height 0.303 0.875 N

HCO3 Site 21.6 <0.001*** Y

HCO3 Year 2.39 0.064 N

HCO3 Height 0.276 0.892 N

144

Table 16. One-way Analysis of Variance (ANOVA) using Fisher’s Least Significant Difference (LSD) method for major oxides. Variables are site, year, and height above the surface. *** is 99.9% confidence, ** is 99%, * is 95% Analyte Variable F- value p- value Significant?

SiO2 Site 74.65 <0.001*** Y

SiO2 Year 9.287 <0.001*** Y

SiO2 Height 1.128 0.362 N

Al2O3 Site 291.8 <0.001*** Y

Al2O3 Year 84.87 <0.001*** Y

Al2O3 Height 0.469 0.758 N

Fe2O3 Site 13.87 <0.001*** Y

Fe2O3 Year 5.535 0.004** Y

Fe2O3 Height 0.352 0.84 N

MgO Site 13.58 <0.001*** Y

MgO Year 2.944 0.0483** Y

MgO Height 1.05 0.398 N

CaO Site 20.95 <0.001*** Y

CaO Year 0.455 0.726 N

CaO Height 1.606 0.199 N

K2O Site 171.4 <0.001*** Y

K2O Year 75.32 <0.001*** Y

K2O Height 0.357 0.837 N

Continued

145

Table 16 Continued.

Na2O Site 143.3 <0.001*** Y

Na2O Year 27.65 <0.001*** Y

Na2O Height 0.304 0.873 N

TiO2 Site 62.53 <0.001*** Y

TiO2 Year 1.075 0.374 N

TiO2 Height 0.395 0.811 N

MnO Site 27.54 <0.001*** Y

MnO Year 7 0.001*** Y

MnO Height 0.435 0.783 N

P2O5 Site 208.1 <0.001*** Y

P2O5 Year 2.757 0.059 N

P2O5 Height 0.256 0.904 N

Cr2O3 Site 0.466 0.827 N

Cr2O3 Year 0.659 0.583 N

Cr2O3 Height 0.643 0.636 N

V2O5 Site 5.542 0.001*** Y

V2O5 Year 6.988 0.00112** Y

V2O5 Height 0.544 0.705 N

146

Table 17. One-way Analysis of Variance (ANOVA) using Fisher’s Least Significant Difference (LSD) method for trace elements and REEs. Variables are site, year, and height above the surface. *** is 99.9% confidence, ** is 99%, * is 95% Analyte Variable F- value p- value Significant?

Ba Site 83.5 <0.001*** Y

Ba Year 39.0 <0.001*** Y

Ba Height 0.249 0.908 N

Cr Site 0.666 0.678 N

Cr Year 1.215 0.321 N

Cr Height 1.021 0.412 N

Cu Site 39.9 <0.001*** Y

Cu Year 67.25 <0.001*** Y

Cu Height 0.257 0.903 N

Fe Site 13.87 <0.001*** Y

Fe Year 5.635 0.003** Y

Fe Height 0.387 0.916 N

Li Site 8.44 <0.001*** Y

Li Year 1.24 0.305 N

Li Height 0.430 0.785 N

Mn Site 23.0 <0.001*** Y

Mn Year 7.98 <0.001*** Y

Mn Height 0.326 0.858 N

Continued

147

Table 17 Continued.

Ni Site 13.5 <0.001*** Y

Ni Year 4.21 <0.001*** Y

Ni Height 1.91 0.135 N

Sc Site 136.7 <0.001*** Y

Sc Year 13.3 <0.001*** Y

Sc Height 0.207 0.587 N

Si Site 5.01 0.003** Y

Si Year 0.912 0.449 N

Si Height 0.402 0.805 N

Sr Site 866 <0.001*** Y

Sr Year 10.95 <0.001*** Y

Sr Height 0.376 0.824 N

Zn Site 6.64 <0.001*** Y

Zn Year 5.71 0.003** Y

Zn Height 0.312 0.868 N

Ce Site 82.8 <0.001*** Y

Ce Year 8.31 <0.001*** Y

Ce Height 0.296 0.878 N

Co Site 15.1 <0.001*** Y

Co Year 5.10 0.006** Y

Continued

148

Table 17 Continued.

Co Height 0.655 0.628 N

Cs Site 12.4 <0.001*** Y

Cs Year 13.4 <0.001*** Y

Cs Height 0.398 0.809 N

Dy Site 16.7 <0.001*** Y

Dy Year 1.99 0.137 N

Dy Height 1.11 0.369 N

Er Site 7.50 <0.001*** Y

Er Year 0.885 0.459 N

Er Height 0.871 0.493 N

Eu Site 231 <0.001*** Y

Eu Year 5.88 0.003** Y

Eu Height 0.351 0.841 N

Ga Site 148.8 <0.001*** Y

Ga Year 30.56 <0.001*** Y

Ga Height 0.587 0.675 N

Gd Site 39.92 <0.001*** Y

Gd Year 4.263 0.0124** Y

Gd Height 0.638 0.639 N

Hf Site 38.96 <0.001*** Y

Continued

149

Table 17 Continued.

Hf Year 1.426 0.254 N

Hf Height 0.305 0.872 N

Ho Site 11.58 <0.001*** Y

Ho Year 1.088 0.369 N

Ho Height 0.693 0.602 N

La Site 76.38 <0.001*** Y

La Year 8.941 <0.001*** Y

La Height 0.249 0.908 N

Lu Site 5.484 0.001*** Y

Lu Year 0.439 0.727 N

Lu Height 2.303 0.0814 N

Mo Site 10.47 <0.001*** Y

Mo Year 12.09 <0.001*** Y

Mo Height 0.462 0.762 N

Nb Site 132 <0.001*** Y

Nb Year 2.517 0.0764 N

Nb Height 0.152 0.96 N

Nd Site 107.4 <0.001*** Y

Nd Year 6.841 0.001*** Y

Nd Height 0.407 0.802 N

Continued

150

Table 17 Continued.

Pb Site 0.273 0.945 N

Pb Year 0.547 0.654 N

Pb Height 1.119 0.366 N

Pr Site 106.9 <0.001*** Y

Pr Year 7.533 0.001*** Y

Pr Height 0.301 0.875 N

Rb Site 102.1 <0.001*** Y

Rb Year 76.33 <0.001*** Y

Rb Height 0.409 0.801 N

Sb Site 0.615 0.715 N

Sb Year 0.736 0.493 N

Sb Height 1.048 0.413 N

Sm Site 74.71 <0.001*** Y

Sm Year 5.852 0.005** Y

Sm Height 0.515 0.725 N

Sn Site 0.266 0.948 N

Sn Year 0.586 0.629 N

Sn Height 1.056 0.395 N

Ta Site 34.61 <0.001*** Y

Ta Year 1.786 0.193 N

Continued

151

Table 17 Continued.

Ta Height 0.263 0.898 N

Tb Site 21.73 <0.001*** Y

Tb Year 2.962 0.0474* Y

Tb Height 1.045 0.401 N

Th Site 17.81 <0.001*** Y

Th Year 17.55 <0.001*** Y

Th Height 0.247 0.909 N

Tm Site 5.494 0.001*** Y

Tm Year 0.81 0.498 N

Tm Height 0.894 0.48 N

U Site 42.62 <0.001*** Y

U Year 2.967 0.0471* Y

U Height 0.476 0.753 N

Y Site 11.7 <0.001*** Y

Y Year 1.098 0.365 N

Y Height 1.108 0.371 N

Yb Site 5.213 0.001*** Y

Yb Year 1.332 0.282 N

Yb Height 0.846 0.507 N

Zr Site 78.91 <0.001*** Y

Continued

152

Table 17 Continued.

Zr Year 2.042 0.128 N

Zr Height 0.188 0.943 N

153