28.11.2012 Views

keck geology consortium proceedings of the twenty-fourth annual ...

keck geology consortium proceedings of the twenty-fourth annual ...

keck geology consortium proceedings of the twenty-fourth annual ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

KECK GEOLOGY CONSORTIUM<br />

PROCEEDINGS OF THE TWENTY-FOURTH<br />

ANNUAL KECK RESEARCH SYMPOSIUM IN<br />

GEOLOGY<br />

April 2011<br />

Union College, Schenectady, NY<br />

Dr. Robert J. Varga, Editor<br />

Director, Keck Geology Consortium<br />

Pomona College<br />

Dr. Holli Frey<br />

Symposium Convenor<br />

Union College<br />

Carol Morgan<br />

Keck Geology Consortium Administrative Assistant<br />

Diane Kadyk<br />

Symposium Proceedings Layout & Design<br />

Department <strong>of</strong> Earth & Environment<br />

Franklin & Marshall College<br />

Keck Geology Consortium<br />

Geology Department, Pomona College<br />

185 E. 6 th St., Claremont, CA 91711<br />

(909) 607-0651, <strong>keck</strong><strong>geology</strong>@pomona.edu, <strong>keck</strong><strong>geology</strong>.org<br />

ISSN# 1528-7491<br />

The Consortium Colleges The National Science Foundation ExxonMobil Corporation


KECK GEOLOGY CONSORTIUM<br />

PROCEEDINGS OF THE TWENTY-FOURTH ANNUAL KECK<br />

RESEARCH SYMPOSIUM IN GEOLOGY<br />

ISSN# 1528-7491<br />

Robert J. Varga<br />

Editor and Keck Director<br />

Pomona College<br />

April 2011<br />

Keck Geology Consortium<br />

Pomona College<br />

185 E 6 th St., Claremont, CA<br />

91711<br />

Diane Kadyk<br />

Proceedings Layout & Design<br />

Franklin & Marshall College<br />

Keck Geology Consortium Member Institutions:<br />

Amherst College, Beloit College, Carleton College, Colgate University, The College <strong>of</strong> Wooster,<br />

The Colorado College, Franklin & Marshall College, Macalester College, Mt Holyoke College,<br />

Oberlin College, Pomona College, Smith College, Trinity University, Union College,<br />

Washington & Lee University, Wesleyan University, Whitman College, Williams College<br />

2010-2011 PROJECTS<br />

FORMATION OF BASEMENT-INVOLVED FORELAND ARCHES: INTEGRATED STRUCTURAL AND<br />

SEISMOLOGICAL RESEARCH IN THE BIGHORN MOUNTAINS, WYOMING<br />

Faculty: CHRISTINE SIDDOWAY, MEGAN ANDERSON, Colorado College, ERIC ERSLEV, University <strong>of</strong><br />

Wyoming<br />

Students: MOLLY CHAMBERLIN, Texas A&M University, ELIZABETH DALLEY, Oberlin College, JOHN<br />

SPENCE HORNBUCKLE III, Washington and Lee University, BRYAN MCATEE, Lafayette College, DAVID<br />

OAKLEY, Williams College, DREW C. THAYER, Colorado College, CHAD TREXLER, Whitman College, TRIANA<br />

N. UFRET, University <strong>of</strong> Puerto Rico, BRENNAN YOUNG, Utah State University.<br />

EXPLORING THE PROTEROZOIC BIG SKY OROGENY IN SOUTHWEST MONTANA<br />

Faculty: TEKLA A. HARMS, JOHN T. CHENEY, Amherst College, JOHN BRADY, Smith College<br />

Students: JESSE DAVENPORT, College <strong>of</strong> Wooster, KRISTINA DOYLE, Amherst College, B. PARKER HAYNES,<br />

University <strong>of</strong> North Carolina - Chapel Hill, DANIELLE LERNER, Mount Holyoke College, CALEB O. LUCY,<br />

Williams College, ALIANORA WALKER, Smith College.<br />

INTERDISCIPLINARY STUDIES IN THE CRITICAL ZONE, BOULDER CREEK CATCHMENT,<br />

FRONT RANGE, COLORADO<br />

Faculty: DAVID P. DETHIER, Williams College, WILL OUIMET. University <strong>of</strong> Connecticut<br />

Students: ERIN CAMP, Amherst College, EVAN N. DETHIER, Williams College, HAYLEY CORSON-RIKERT,<br />

Wesleyan University, KEITH M. KANTACK, Williams College, ELLEN M. MALEY, Smith College, JAMES A.<br />

MCCARTHY, Williams College, COREY SHIRCLIFF, Beloit College, KATHLEEN WARRELL, Georgia Tech<br />

University, CIANNA E. WYSHNYSZKY, Amherst College.<br />

SEDIMENT DYNAMICS & ENVIRONMENTS IN THE LOWER CONNECTICUT RIVER<br />

Faculty: SUZANNE O’CONNELL, Wesleyan University<br />

Students: LYNN M. GEIGER, Wellesley College, KARA JACOBACCI, University <strong>of</strong> Massachusetts (Amherst),<br />

GABRIEL ROMERO, Pomona College.<br />

GEOMORPHIC AND PALEOENVIRONMENTAL CHANGE IN GLACIER NATIONAL PARK,<br />

MONTANA, U.S.A.<br />

Faculty: KELLY MACGREGOR, Macalester College, CATHERINE RIIHIMAKI, Drew University, AMY MYRBO,<br />

LacCore Lab, University <strong>of</strong> Minnesota, KRISTINA BRADY, LacCore Lab, University <strong>of</strong> Minnesota


Students: HANNAH BOURNE, Wesleyan University, JONATHAN GRIFFITH, Union College, JACQUELINE<br />

KUTVIRT, Macalester College, EMMA LOCATELLI, Macalester College, SARAH MATTESON, Bryn Mawr<br />

College, PERRY ODDO, Franklin and Marshall College, CLARK BRUNSON SIMCOE, Washington and Lee<br />

University.<br />

GEOLOGIC, GEOMORPHIC, AND ENVIRONMENTAL CHANGE AT THE NORTHERN<br />

TERMINATION OF THE LAKE HÖVSGÖL RIFT, MONGOLIA<br />

Faculty: KARL W. WEGMANN, North Carolina State University, TSALMAN AMGAA, Mongolian University <strong>of</strong><br />

Science and Technology, KURT L. FRANKEL, Georgia Institute <strong>of</strong> Technology, ANDREW P. deWET, Franklin &<br />

Marshall College, AMGALAN BAYASAGALN, Mongolian University <strong>of</strong> Science and Technology.<br />

Students: BRIANA BERKOWITZ, Beloit College, DAENA CHARLES, Union College, MELLISSA CROSS, Colgate<br />

University, JOHN MICHAELS, North Carolina State University, ERDENEBAYAR TSAGAANNARAN, Mongolian<br />

University <strong>of</strong> Science and Technology, BATTOGTOH DAMDINSUREN, Mongolian University <strong>of</strong> Science and<br />

Technology, DANIEL ROTHBERG, Colorado College, ESUGEI GANBOLD, ARANZAL ERDENE, Mongolian<br />

University <strong>of</strong> Science and Technology, AFSHAN SHAIKH, Georgia Institute <strong>of</strong> Technology, KRISTIN TADDEI,<br />

Franklin and Marshall College, GABRIELLE VANCE, Whitman College, ANDREW ZUZA, Cornell University.<br />

LATE PLEISTOCENE EDIFICE FAILURE AND SECTOR COLLAPSE OF VOLCÁN BARÚ, PANAMA<br />

Faculty: THOMAS GARDNER, Trinity University, KRISTIN MORELL, Penn State University<br />

Students: SHANNON BRADY, Union College. LOGAN SCHUMACHER, Pomona College, HANNAH ZELLNER,<br />

Trinity University.<br />

KECK SIERRA: MAGMA-WALLROCK INTERACTIONS IN THE SEQUOIA REGION<br />

Faculty: JADE STAR LACKEY, Pomona College, STACI L. LOEWY, California State University-Bakersfield<br />

Students: MARY BADAME, Oberlin College, MEGAN D’ERRICO, Trinity University, STANLEY HENSLEY,<br />

California State University, Bakersfield, JULIA HOLLAND, Trinity University, JESSLYN STARNES, Denison<br />

University, JULIANNE M. WALLAN, Colgate University.<br />

EOCENE TECTONIC EVOLUTION OF THE TETONS-ABSAROKA RANGES, WYOMING<br />

Faculty: JOHN CRADDOCK, Macalester College, DAVE MALONE, Illinois State University<br />

Students: JESSE GEARY, Macalester College, KATHERINE KRAVITZ, Smith College, RAY MCGAUGHEY,<br />

Carleton College.<br />

Funding Provided by:<br />

Keck Geology Consortium Member Institutions<br />

The National Science Foundation Grant NSF-REU 1005122<br />

ExxonMobil Corporation


Keck Geology Consortium: Projects 2010-2011<br />

Short Contributions— Front Range, CO<br />

INTERDISCIPLINARY STUDIES IN THE CRITICAL ZONE, BOULDER CREEK CATCHMENT,<br />

FRONT RANGE, COLORADO<br />

Project Faculty: DAVID P. DETHIER: Williams College, WILL OUIMET: University <strong>of</strong> Connecticut<br />

CORING A 12KYR SPHAGNUM PEAT BOG: A SEARCH FOR MERCURY AND ITS IMPLICATIONS<br />

ERIN CAMP, Amherst College<br />

Research Advisor: Anna Martini<br />

EXAMINING KNICKPOINTS IN THE BOULDER CREEK CATCHMENT, COLORADO<br />

EVAN N. DETHIER, Williams College<br />

Research Advisor: David P. Dethier<br />

THE DISTRIBUTION OF PHOSPHORUS IN ALPINE AND UPLAND SOILS OF THE BOULDER<br />

CREEK, COLORADO CATCHMENT<br />

HAYLEY CORSON-RIKERT, Wesleyan University<br />

Research Advisor: Timothy Ku<br />

RECONSTRUCTING THE PINEDALE GLACIATION, GREEN LAKES VALLEY, COLORADO<br />

KEITH M. KANTACK, Williams College<br />

Research Advisor: David P. Dethier<br />

CHARACTERIZATION OF TRACE METAL CONCENTRATIONS AND MINING LEGACY IN SOILS,<br />

BOULDER COUNTY, COLORADO<br />

ELLEN M. MALEY, Smith College<br />

Research Advisor: Amy L. Rhodes<br />

ASSESSING EOLIAN CONTRIBUTIONS TO SOILS IN THE BOULDER CREEK CATCHMENT,<br />

COLORADO<br />

JAMES A. MCCARTHY, Williams College<br />

Research Advisor: David P. Dethier<br />

USING POLLEN TO UNDERSTAND QUATERNARY PALEOENVIRONMENTS IN BETASSO GULCH,<br />

COLORADO<br />

COREY SHIRCLIFF, Beloit College<br />

Research Advisor: Carl Mendelson<br />

STREAM TERRACES IN THE CRITICAL ZONE – LOWER GORDON GULCH, COLORADO<br />

KATHLEEN WARRELL, Georgia Tech<br />

Research Advisor: Kurt Frankel<br />

METEORIC 10 BE IN GORDON GULCH SOILS: IMPLICATIONS FOR HILLSLOPE PROCESSES AND<br />

DEVELOPMENT<br />

CIANNA E. WYSHNYSZKY, Amherst College<br />

Research Advisor: Will Ouimet and Peter Crowley<br />

Keck Geology Consortium<br />

Pomona College<br />

185 E. 6 th St., Claremont, CA 91711<br />

Keck<strong>geology</strong>.org


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

INTERDISCIPLINARY STUDIES IN THE CRITICAL ZONE,<br />

BOULDER CREEK CATCHMENT, FRONT RANGE, COLORADO<br />

DAVID P. DETHIER, Williams College<br />

WILL OUIMET, University <strong>of</strong> Connecticut<br />

INTRODUCTION<br />

Processes in <strong>the</strong> critical zone, <strong>the</strong> life-sustaining surficial<br />

mantle <strong>of</strong> <strong>the</strong> earth, involve wea<strong>the</strong>red geologic<br />

materials, water, and <strong>the</strong> biosphere, mediated by<br />

atmospheric processes that are controlled by changing<br />

climate. Field and laboratory studies that investigate<br />

geologic, hydrologic and geochemical components <strong>of</strong><br />

<strong>the</strong> critical zone provide valuable data about processes<br />

and <strong>the</strong> physical basis for <strong>the</strong>ir integration into<br />

models <strong>of</strong> short and long-term geomorphic, hydrologic<br />

and biochemical response. The Keck Colorado<br />

Project is working in cooperation with a large<br />

interdisciplinary study <strong>of</strong> <strong>the</strong> critical zone (Boulder<br />

Creek Critical Zone Observatory: Wea<strong>the</strong>red pr<strong>of</strong>ile<br />

development in a rocky environment and its influence<br />

on watershed hydrology and biogeochemistry— Suzanne<br />

Anderson, PI, Institute for Arctic and Alpine<br />

Studies, University <strong>of</strong> Colorado). The observatory<br />

(CZO) consists <strong>of</strong> 3 small, instrumented catchments<br />

in <strong>the</strong> Boulder Creek basin, Colorado Front Range:<br />

(1) Green Lakes Valley (GLV; el. 3400 m)--a steep,<br />

glacially scoured alpine area in <strong>the</strong> City <strong>of</strong> Boulder<br />

watershed; (2) Gordon Gulch (el. 2600 m)--a forested,<br />

mid-elevation catchment that exposes isolated bedrock<br />

remnants (tors) developed on a surface <strong>of</strong> low<br />

relief; and (3) Betasso gulch (el. 1950 m)--a steep,<br />

thinly forested basin that preserves thick regolith in<br />

<strong>the</strong> upper catchment and exposes extensive bedrock<br />

outcrops at lower elevations (Fig. 1).<br />

The glaciated GLV, low relief surface, and bedrock<br />

canyons are developed in granitic or gneissic rocks<br />

and are influenced by <strong>the</strong> strong gradient in elevation,<br />

climate and vegetation from west to east. Variation<br />

in critical-zone development in <strong>the</strong>se different environments<br />

allows us to test models <strong>of</strong> wea<strong>the</strong>ring and<br />

regolith generation, elemental cycling, slope evolution<br />

and sediment transport in an accessible field<br />

setting. Land-use, vegetation and hydrologic response<br />

93<br />

in each CZO catchment also reflect changes produced<br />

by anthropogenic activities such as mining and timber<br />

harvest over <strong>the</strong> past 150 years. Keck Colorado field<br />

studies focus on using a variety <strong>of</strong> techniques to map<br />

and characterize <strong>the</strong> geologic history, near-surface<br />

geologic materials and geochemical properties for<br />

each <strong>of</strong> <strong>the</strong> study catchments.<br />

Figure 1. Perspective view looking west across <strong>the</strong> Front<br />

Range from Boulder, Colorado, showing Middle Boulder<br />

Creek and location <strong>of</strong> Betasso, Gordon Gulch and Green<br />

Lakes Valley catchments, Boulder Creek Critical Zone<br />

Observatory. White filled area shows approximate extent<br />

<strong>of</strong> latest Pleistocene glaciers (after Madole et al., 1999).<br />

SETTING<br />

The middle Boulder Creek catchment (Fig. 1) extends<br />

from <strong>the</strong> glaciated alpine zone <strong>of</strong> <strong>the</strong> Continental<br />

Divide east to <strong>the</strong> semi-arid western edge <strong>of</strong> <strong>the</strong> Great<br />

Plains. The high-relief zone <strong>of</strong> cirques and deep, Ushaped<br />

valleys in <strong>the</strong> glaciated area become shallower<br />

eastward through a zone <strong>of</strong> low relief and relatively<br />

low slopes. To <strong>the</strong> east, valleys deepen into steep,<br />

narrow bedrock canyons as <strong>the</strong>y pass knickzones, and<br />

flatten to lower channel slopes near <strong>the</strong> piedmont margin.<br />

Small glaciers and late-persisting snowfields dot<br />

<strong>the</strong> alpine zone, which exposes bedrock and relatively<br />

thin deposits related to <strong>the</strong> latest Pleistocene Pinedale<br />

glaciation and to Holocene erosion. The forested zone


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

<strong>of</strong> low relief exposes local areas <strong>of</strong> thick (characteristically<br />

3 to 8 m) regolith, saprolite and oxidized bedrock,<br />

but <strong>the</strong> wea<strong>the</strong>red mantle is thin in o<strong>the</strong>r areas.<br />

Low terraces and alluvial fans as thick as 4 m line<br />

channels locally. In <strong>the</strong> vicinity <strong>of</strong> knickzones and<br />

in downstream areas such as Betasso gulch, slopes<br />

near channels are steep and fresh bedrock is exposed,<br />

whereas areas more distant from channels retain a<br />

thicker wea<strong>the</strong>red mantle.<br />

APPROACH<br />

In our third project year, we used field mapping and<br />

sampling in all three CZO catchments, supplemented<br />

by geophysical measurements, in order to provide<br />

basic data about soils and <strong>the</strong>ir geochemistry, shallow<br />

subsurface <strong>geology</strong> and erosional history <strong>of</strong><br />

<strong>the</strong> critical zone. Students supported by <strong>the</strong> Keck<br />

Geology <strong>consortium</strong> and by NSF learned geophysical<br />

techniques and initial data reduction, processing<br />

and visualization methods in <strong>the</strong>se settings. Students<br />

chose from a variety <strong>of</strong> potential projects in <strong>the</strong> study<br />

catchments; 2010 project topical areas included:<br />

1. Trace-metal studies <strong>of</strong> soils and bog sediment,<br />

Boulder Creek catchment<br />

2. Soil chemistry (Fe and P) in CZO catchments<br />

and meteoric 10Be studies <strong>of</strong> soils, focused on<br />

Gordon Gulch<br />

3. Stratigraphy and palynology <strong>of</strong> valley-fill<br />

sediment, Gordon and Betasso catchments<br />

4. Geomorphic research: Ice erosion and geo<br />

morphic evolution <strong>of</strong> <strong>the</strong> Green Lakes Valley<br />

and <strong>the</strong> evolution <strong>of</strong> knickpoints along channels<br />

in <strong>the</strong> Boulder Creek catchment<br />

We ran resistivity lines in Gordon Gulch, resistivity<br />

and ground-penetrating radar on Niwot Ridge and<br />

ground-penetrating radar in <strong>the</strong> vicinity <strong>of</strong> <strong>the</strong> bogs<br />

near <strong>the</strong> N. Branch Boulder Creek. In Gordon Gulch,<br />

we worked on regolith studies in cooperation with<br />

investigators from <strong>the</strong> University <strong>of</strong> Colorado and <strong>the</strong><br />

US Geological Survey.<br />

STUDENT PROJECTS<br />

Six Keck students joined Williams students Evan<br />

Dethier, Keith Kantack and James McCarthy and<br />

94<br />

Erin Camp (Amherst), who were supported directly<br />

by NSF funding. David Dethier and Will Ouimet<br />

supervised students on a daily basis and field teams<br />

frequently joined investigators and graduate students<br />

from <strong>the</strong> University <strong>of</strong> Colorado. Matthias Leopold<br />

(Technical University <strong>of</strong> Munich) worked with Keck<br />

students for two weeks in <strong>the</strong> field. Keck students<br />

were among <strong>the</strong> only undergraduates to present preliminary<br />

results <strong>of</strong> <strong>the</strong>ir field research at <strong>the</strong> 3rd <strong>annual</strong><br />

Boulder Creek CZO meeting on 10 August 2010.<br />

Keck Colorado students worked in pairs on a daily<br />

basis and sometimes as geophysical support teams.<br />

Geophysical data provided important background data<br />

for extensive studies in Gordon Gulch and for coring<br />

<strong>of</strong> bogs (Fig. 2).<br />

Figure 2. Coring a bog in a late Pinedale moraine complex<br />

along <strong>the</strong> N. Fork Boulder Creek, supervised by<br />

Robert Nelson, Colby College.<br />

Short papers elsewhere in this volume report results<br />

<strong>of</strong> <strong>the</strong> field and laboratory studies in some detail.<br />

We summarize and provide brief comments on this<br />

research here.<br />

Studies <strong>of</strong> trace metals in organic sediment, soil<br />

and regolith in Gordon Gulch and adjacent areas<br />

Erin Camp (Amherst) reports “Coring a 12 kyr sphagnum<br />

bog in <strong>the</strong> N. Boulder Creek valley—a search<br />

for mercury and its implications” and Ellie Maley<br />

(Smith) worked on “Characterization <strong>of</strong> trace metal<br />

concentrations and mining legacy in soils, Boulder<br />

County, Colorado”. Both <strong>of</strong> <strong>the</strong>se studies demonstrate<br />

that Hg is enriched in recent organic-rich


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

sediment <strong>of</strong> <strong>the</strong> montane zone when compared to<br />

older bog sediment or to soil C-horizons. The bog Hg<br />

pr<strong>of</strong>ile and 14 C ages show a strong correspondence<br />

between elevated Hg levels and large, silicic volcanic<br />

eruptions in Holocene time. Peak Hg levels produced<br />

by eruptions decline over narrow intervals. Mercury<br />

concentrations in bog sediments increase to a broad<br />

peak coincident with exploitation <strong>of</strong> precious metals<br />

in <strong>the</strong> western hemisphere beginning in <strong>the</strong> 16th century<br />

and peaking in <strong>the</strong> late 19th century in Colorado.<br />

Ellie’s work shows that Hg, As and possibly Pb are<br />

enriched in organic-rich O and A-horizons compared<br />

to soil parent material; o<strong>the</strong>r minor elements such as<br />

Cu and Zn do not show enrichment. Ellie’s research<br />

suggests that metal enrichment likely represents <strong>the</strong><br />

legacy <strong>of</strong> local metal mining, milling and smelting,<br />

and <strong>the</strong> affinity <strong>of</strong> organic matter for <strong>the</strong>se relatively<br />

volatile trace metals.<br />

Studies <strong>of</strong> soil and regolith geochemistry and age<br />

in Gordon Gulch and adjacent areas<br />

Four students, including Ellie Maley, studied soils<br />

and regolith exposed in pits that were dug in Gordon<br />

Gulch and at o<strong>the</strong>r nearby exposures. The “pit<br />

crew”, aided by <strong>the</strong>ir advisors and o<strong>the</strong>r CZO investigators,<br />

collected from many <strong>of</strong> <strong>the</strong> same sites and<br />

worked on separate, but complimentary topics. Cianna<br />

Wyshnytzky (Amherst) documented meteoric<br />

10 Be accumulation in soils from Gordon Gulch and<br />

at Silver Lake—“Erosion, particle paths and deposition—meteoric<br />

10 Be in Gordon Gulch”. James<br />

McCarthy (Williams) studied <strong>the</strong> texture and Fed<br />

(dithionite-extractable iron) accumulation in soils<br />

from Gordon Gulch and adjacent subalpine and alpine<br />

areas: “Assessing eolian contributions to soils in <strong>the</strong><br />

Boulder Creek catchment”, whereas Hayley Corson-<br />

Rikert (Wesleyan) studied “Extractable P in soils <strong>of</strong><br />

<strong>the</strong> Boulder Creek catchment, Colorado”.<br />

Data from <strong>the</strong>se studies demonstrate that <strong>the</strong>re are<br />

fundamental differences between soils developed on<br />

stable sites and those that formed on regolith-covered<br />

hillslopes and that dustfall and chemical wea<strong>the</strong>ring<br />

influence <strong>the</strong> partitioning <strong>of</strong> Fed and P in soils. Both<br />

erosion and wea<strong>the</strong>ring rates are related to aspect in<br />

Gordon Gulch; regolith is thicker, more wea<strong>the</strong>red<br />

and contains more meteoric 10 Be on <strong>the</strong> north-facing<br />

95<br />

Figure 3. Soil pits in Gordon Gulch. A. James McCarthy,<br />

Cianna Wyshnytzky, Hayley Corson-Rikert and Ellie<br />

Maley in a 1.8 m-deep pit in regolith, north-facing slope.<br />

B. Soil-sampling tools and thin regolith over saprolite,<br />

south-facing slope. Cianna Wyshnytzky sampling for meteoric<br />

10Be analysis.<br />

slope. Fe and P mobility also are influenced by <strong>the</strong><br />

moisture and temperature gradient between soils exposed<br />

in Gordon Gulch and soils <strong>of</strong> similar age in <strong>the</strong><br />

alpine and subalpine portions <strong>of</strong> <strong>the</strong> Boulder Creek<br />

CZO. Extractable Fe, clay and 10 Be reach peak values<br />

in <strong>the</strong> B-horizon <strong>of</strong> <strong>the</strong> till-derived soil at Silver Lake<br />

(Fig. 4), whereas P is depleted in <strong>the</strong> soil compared to<br />

unwea<strong>the</strong>red till; soils in Gordon Gulch and Betasso<br />

do not display comparable patterns.<br />

The inventory <strong>of</strong> meteoric 10 Be in Gordon Gulch is<br />

consistent with model predictions for deposition rates<br />

(Graly et al., 2011), but <strong>the</strong> inventory at Silver Lake is<br />

too high, suggesting that at this site dustfall rates or


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 4. Plot showing relationship <strong>of</strong> Fed, clay and<br />

meteoric 10 Be to depth below <strong>the</strong> surface and soil horizons<br />

at Silver Lake, Green Lakes Valley. P (not plotted) is depleted<br />

and relatively labile in <strong>the</strong> upper soil and relatively<br />

enriched and associated with inorganic material in <strong>the</strong><br />

unoxidized till.<br />

late Pleistocene precipitation may have been substantially<br />

higher than at present. Cianna’s meteoric 10 Be<br />

research represents <strong>the</strong> first application <strong>of</strong> this technique<br />

in <strong>the</strong> Boulder Creek CZO catchments.<br />

Stratigraphy and palynology <strong>of</strong> valley-fill sediment,<br />

Gordon and Betasso catchments<br />

In Gordon Gulch and in upper Betasso Gulch, colluvium<br />

and deposits beneath terraces as much as 4<br />

m above <strong>the</strong> channel comprise local valley fills <strong>of</strong><br />

Holocene and latest Pleistocene (?) age. Kathleen<br />

Warrell (Georgia Tech) studied terrace morphology<br />

and sampled deposits exposed beneath low terraces<br />

in Gordon Gulch-- “Stream terraces in <strong>the</strong> critical<br />

zone-- lower Gordon Gulch, Colorado”. Her work<br />

96<br />

shows that at least 75,000 m 3 <strong>of</strong> sediment is stored in<br />

near <strong>the</strong> channel in lower Gordon Gulch. Sediment<br />

beneath <strong>the</strong> 1-m terrace is less than 1500 years old<br />

(Fig. 5);<br />

Figure 5. Graphic log <strong>of</strong> sediment (description from C.<br />

Shircliff) exposed beneath K. Warrell’s 1-m terrace. Basal<br />

sediment is approximately 1500 years old.<br />

higher terraces are cut into middle and early Holocene<br />

alluvial and colluvial deposits. Corey Shircliff<br />

(Beloit) studied organic material in Holocene terrace<br />

deposits in Gordon Gulch and Betasso--“Using pollen<br />

to understand Quaternary paleoenvironments in<br />

Betasso Gulch, Colorado”. She was able to separate<br />

pollen from a buried soil developed on colluvium and<br />

to identify many <strong>of</strong> <strong>the</strong> pollen grains. Corey’s work<br />

indicates that early Holocene pollen at Betasso is<br />

richer in Picea (spruce) than a modern pollen sample<br />

and records a climate that was likely wetter and perhaps<br />

slightly cooler than that at present.<br />

Ice erosion and geomorphic evolution <strong>of</strong> <strong>the</strong> Green<br />

Lakes Valley and <strong>the</strong> evolution <strong>of</strong> knickpoints<br />

along channels in <strong>the</strong> Boulder Creek catchment<br />

Keith Kantack (Williams) used field measurements


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

in GLV and interpretation <strong>of</strong> DEMs derived from<br />

Lidar flown in August 2011 to map <strong>the</strong> extent <strong>of</strong> late<br />

Pinedale ice and glacial moraines in <strong>the</strong> upper North<br />

Boulder Creek catchment (Fig. 6).<br />

Figure 6. Keith Kantack, Evan Dethier and James Mc-<br />

Carthy stand on glacially sculpted and smoo<strong>the</strong>d bedrock<br />

knob, upper Green Lakes Valley.<br />

His work “Reconstructing<strong>the</strong> Pinedale glaciation in<br />

<strong>the</strong> Green Lakes valley, Colorado” shows that latest<br />

Pleistocene ice in <strong>the</strong> GLV was thin (mainly less<br />

than 150 m), but extended out <strong>of</strong> <strong>the</strong> cirques and<br />

more than 10 km to <strong>the</strong> east to elevations as low as<br />

2650 m. Measured moraine volumes suggest that <strong>the</strong><br />

average erosion rate by North Boulder cirque glaciers<br />

was about 1mm yr -1 during <strong>the</strong> maximum late Pinedale<br />

glaciation, which lasted from about 21 to 15 ka.<br />

Those rates are similar to estimates used by Ward et<br />

al. (2009) for modeling ice flow in Front Range catchments.<br />

Evan Dethier studied <strong>the</strong> evolution <strong>of</strong> channels and<br />

slopes at different scales near knickpoints in Betasso,<br />

Gordon Gulch, and along Middle Boulder Creek--<br />

“Knickpoints—a study <strong>of</strong> channels in <strong>the</strong> Boulder<br />

Creek catchment”. His field studies, combined with<br />

RiverTools interpretation <strong>of</strong> DEMs derived from August,<br />

2010 Lidar, suggest that channels and adjacent<br />

hillslopes reflect <strong>the</strong> slow migration <strong>of</strong> knickpoints<br />

in <strong>the</strong> Boulder Creek catchment, moderated by local<br />

rock strength. Betasso gulch, <strong>the</strong> smallest <strong>of</strong> <strong>the</strong><br />

Boulder Creek CZO catchments, has a channel that is<br />

97<br />

steep and rough throughout and is flanked by steep,<br />

smooth slopes that expose bedrock and thin regolith<br />

near Boulder Creek. In <strong>the</strong> upper part <strong>of</strong> <strong>the</strong> catchment,<br />

however, <strong>the</strong> channel only locally exposes<br />

bedrock, and is cut mainly in s<strong>of</strong>t, deeply wea<strong>the</strong>red<br />

saprolite and through thick colluvium that was deposited<br />

in latest Pleistocene time. At Gordon Gulch, rock<br />

strength appears to control <strong>the</strong> location <strong>of</strong> <strong>the</strong> knickzone<br />

that separates <strong>the</strong> upper and lower basin; morphology<br />

<strong>of</strong> adjacent hillslopes reflect local channel<br />

slope. At <strong>the</strong> scale <strong>of</strong> Boulder Creek, hillslope evolution<br />

appears to “lag” knickpoint migration because<br />

local rocks are strong and hillslope erosion requires<br />

removal <strong>of</strong> large volumes <strong>of</strong> rock.<br />

Figure 7. Sculpted bedrock and boulder-rich channel<br />

in <strong>the</strong> knickzone along N. Boulder Creek above Boulder<br />

Falls.<br />

CONCLUSIONS<br />

“Piggybacking” <strong>the</strong> Keck Colorado Geology Project<br />

on <strong>the</strong> NSF-Boulder Creek Critical Zone Observatory<br />

has allowed Keck undergraduates to integrate <strong>the</strong>ir<br />

projects with <strong>the</strong> research <strong>of</strong> graduate and postdoctoral<br />

students from <strong>the</strong> University <strong>of</strong> Colorado and<br />

o<strong>the</strong>r research universities. Keck student research<br />

has benefitted from <strong>the</strong> personnel, monitoring efforts,<br />

and general level <strong>of</strong> scientific interest associated with<br />

<strong>the</strong> NSF project. The Boulder Creek CZO has gained<br />

from <strong>the</strong> focused field and laboratory research <strong>of</strong> <strong>the</strong><br />

Keck students, <strong>the</strong>ir energy, and <strong>the</strong>ir collective demonstration<br />

<strong>of</strong> what can be accomplished by <strong>the</strong> best


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

undergraduates.<br />

ACKNOWLEDGMENTS<br />

Field studies and measurements in <strong>the</strong> Boulder Creek<br />

area were performed in cooperation with <strong>the</strong> Boulder<br />

Creek CZO Project (National Science Foundation),<br />

<strong>the</strong> USDA Forest Service, and <strong>the</strong> City <strong>of</strong> Boulder<br />

Watershed and Parks and Recreation Departments.<br />

Nel Caine (University <strong>of</strong> Colorado) and Craig Skeie<br />

(City Watershed Manager) guided work in <strong>the</strong> Green<br />

Lakes basin, Pete Birkeland (University <strong>of</strong> Colorado)<br />

taught us about soils. Greg Tucker, Cam Wobus and<br />

Abby Langston (all affiliated with <strong>the</strong> University <strong>of</strong><br />

Colorado) shared <strong>the</strong>ir knowledge <strong>of</strong> <strong>the</strong> Critical Zone<br />

and how to study it. Suzanne Anderson and Bob Anderson<br />

joined us for many field “teaching moments”.<br />

We gratefully acknowledge <strong>the</strong> field and laboratory<br />

skills <strong>of</strong> Bob Nelson (Colby College), and <strong>the</strong> ongoing<br />

cooperation, digging ability and cogent advice <strong>of</strong><br />

Joerg Voelkel and Matthias Leopold (Technical University<br />

<strong>of</strong> Munich). The hospitality <strong>of</strong> <strong>the</strong> Mountain<br />

Research Station made this project possible.<br />

REFERENCES<br />

Graly, J. A., Reusser, L. J., and Bierman, P. R., 2011,<br />

Short and long-term delivery rates <strong>of</strong> meteoric<br />

10Be to terrestrial soils: Earth and Planetary<br />

Science Letters 302, p. 329-336.<br />

Madole, R.F., VanSistine, D.P., and Michael, J. A.,<br />

1999, Pleistocene glaciation in <strong>the</strong> upper Platte<br />

River drainage basin, Colorado. U.S. Geol. Surv.<br />

Geol. Invest. Series I-2644.<br />

Ward, D. J., R. S. Anderson, Z. S. Guido, and J. P.<br />

Briner (2009), Numerical modeling <strong>of</strong> cosmogenic<br />

deglaciation records, Front Range and San<br />

Juan mountains, Colorado, J. Geophys. Res.,<br />

114, F01026, doi:10.1029/2008JF001057.<br />

98


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

CORING A 12KYR OMBROTROPHIC SPHAGNUM PEAT BOG:<br />

A HISTORY OF ATMOSPHERIC MERCURY<br />

ERIN CAMP, Amherst College<br />

Research Advisor: Anna Martini<br />

INTRODUCTION<br />

Elemental mercury (Hg 0 ) is primarily transported<br />

through <strong>the</strong> atmosphere, where it has an average<br />

residence time <strong>of</strong> one year and can be deposited<br />

worldwide (Bindler, 2003). Hg 0 is deposited both<br />

naturally and anthropogenically in <strong>the</strong> environment,<br />

where it can chemically transform into a highly toxic<br />

methylated form <strong>of</strong> mercury (Vandal et al., 1993).<br />

Mercury is introduced naturally into <strong>the</strong> environment<br />

through volcanism, geo<strong>the</strong>rmal activity, and emission<br />

from <strong>the</strong> biosphere and water bodies, and anthropogenically<br />

through coal combustion, waste incineration,<br />

and metal ore processing (Bindler, 2003). Additionally,<br />

mercury retention is known to increase in<br />

colder temperatures, thus can be used as a paleotemperature<br />

proxy (Martínez-Cortizas et al., 1999).<br />

Ombrotrophic peat bogs topped by Sphagnum moss<br />

are excellent archives <strong>of</strong> elemental Hg deposition<br />

because <strong>the</strong>y receive all <strong>the</strong>ir nutrients from <strong>the</strong><br />

atmosphere and allow little vertical mixing (Madsen,<br />

1981; Lodenius et al., 1983). The Colorado Front<br />

Range has a rich history <strong>of</strong> gold and silver mining,<br />

smelting and mercury amalgamation, thus it is an<br />

ideal location for mercury studies (Nriagu, 1994).<br />

This project has measured <strong>the</strong> amount <strong>of</strong> Hg deposition<br />

in North Boulder Creek Bog, CO in order to 1)<br />

identify <strong>the</strong> natural background Hg deposition for<br />

this location, 2) correlate concentrations with natural<br />

and anthropogenic historical events, and 3) calculate<br />

<strong>the</strong> amount <strong>of</strong> anthropogenically deposited mercury<br />

in this location.<br />

PROJECT LOCATION<br />

North Boulder Creek Bog (40.007349º,-105.560421º;<br />

Fig. 1) is a 3-meter deep subalpine Sphagnum mosscoated<br />

ombrotrophic bog that began to accumulate<br />

organic sediment at 12,000 cal. 14 C years BP (yBP;<br />

99<br />

Figure 1: Google Earth image <strong>of</strong> project location,<br />

illustrating <strong>the</strong> perimeter <strong>of</strong> <strong>the</strong> bog and <strong>the</strong> surrounding<br />

glacial moraine.<br />

present refers to 1950) (Leopold and Dethier, 2007;<br />

Leopold, 2010 personal communication). The bog<br />

is located in <strong>the</strong> Front Range <strong>of</strong> Colorado, situated<br />

within a kettle hole associated with <strong>the</strong> Pinedale<br />

glaciation, and flanked by a glacial moraine on its<br />

eastern side (Richmond, 1960; Fig. 1). North Boulder<br />

Creek Bog is believed to have formed after <strong>the</strong><br />

rapid drainage <strong>of</strong> Lake Devlin about 13,000 years<br />

ago (Leopold pers. comm. 2010; Madole, 1985).<br />

METHODS<br />

The coring, performed with a modified Livingstone<br />

piston corer (Livingstone, 1955), produced a complete<br />

1.8m core that had been compacted by an average<br />

<strong>of</strong> 50%—slightly less near <strong>the</strong> top and more<br />

near <strong>the</strong> bottom—representing a 3.65m core. Sampling<br />

<strong>of</strong> <strong>the</strong> core was performed at 2.5cm intervals,<br />

representing ‘expanded’ intervals <strong>of</strong> 5cm. All depths<br />

referred to in this paper are ‘expanded’ depths. All<br />

samples were dried overnight—maximum <strong>of</strong> 12<br />

hours—at 105ºC. All seventy-four samples were


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

ground with mortar and pestle, and run individually<br />

in a direct combustion cold vapor AA instrument<br />

(Hydra-C, Leeman Labs) for total mercury concentration.<br />

The machine was calibrated using a marine<br />

sediment standard with a precision <strong>of</strong> ±6.03% and an<br />

accuracy (recovery) <strong>of</strong> 103.6%. Following <strong>the</strong>se runs,<br />

higher resolution sampling was conducted at 2cm<br />

‘expanded’ intervals near <strong>the</strong> uppermost section <strong>of</strong> <strong>the</strong><br />

core, in order to obtain more precise data during<br />

modern times. Six samples were taken near 35cm<br />

depth, and five additional samples were taken from<br />

<strong>the</strong> top 5cm <strong>of</strong> <strong>the</strong> core. Additionally, five samples<br />

were sent to <strong>the</strong> Woods Hole NOSAMS Facility for<br />

AMS radiocarbon analysis. These samples were extracted<br />

from various locations along <strong>the</strong> core at 100,<br />

160, 265, 315, and 365cm depths. Radiocarbon age is<br />

calculated from <strong>the</strong> δ13C-corrected Fraction Modern<br />

(Fm) according to <strong>the</strong> following formula:<br />

Age = -8033ln (Fm).<br />

Figure 2: Photos <strong>of</strong> top two sections <strong>of</strong> <strong>the</strong> core. Top<br />

photo represents <strong>the</strong> first meter <strong>of</strong> core, compacted to<br />

48cm. Bottom picture represents <strong>the</strong> second meter <strong>of</strong> core,<br />

compacted to 46cm.<br />

RESULTS<br />

The radiocarbon data yield a typical, slightly curved<br />

age vs. depth trend. With a linear regression analysis,<br />

<strong>the</strong> bog has a deposition rate <strong>of</strong> 0.36 mm yr -1 . A poly-<br />

100<br />

Figure 3: Age-depth graphs for C-14 data <strong>of</strong> <strong>the</strong> five bog<br />

samples, calibrated at Woods Hole HOSAMS facility.<br />

Top graph is a polynomial regression; bottom graph is<br />

a linear regression. Reporting <strong>of</strong> ages and/or activities<br />

follows <strong>the</strong> convention outlined by Stuiver and Polach<br />

(1977) and Stuiver (1980). Ages are calculated using<br />

5568 years as <strong>the</strong> half-life <strong>of</strong> radiocarbon and are reported<br />

without reservoir corrections or calibration to calendar<br />

years. Boxed inset is illustrated in Figure 5.<br />

nomial regression was used in order to fit <strong>the</strong> data at<br />

<strong>the</strong> shallowest and deepest parts <strong>of</strong> <strong>the</strong> core, where<br />

<strong>the</strong> best fit line tapers <strong>of</strong>f in a convex fashion (Fig.<br />

3). The polynomial regression suggests a calibrated<br />

age <strong>of</strong> 9105 yBP at <strong>the</strong> deepest part <strong>of</strong> <strong>the</strong> core.<br />

The data from our mercury analysis, shown in Figure<br />

4, range from 5.2 ppb to 201.2 ppb (ng/g). The highest<br />

values are recorded at shallow depths in <strong>the</strong> core,<br />

at recent times with greater anthropogenic influence,<br />

while <strong>the</strong> lowest values are recorded deep within<br />

<strong>the</strong> core, during historic times. The natural mercury<br />

background concentration for <strong>the</strong> North Boulder<br />

Creek Bog was calculated at approximately 23.2 ppb.<br />

Discernable peaks in mercury concentration occur at<br />

0.5cm (158.9 ppb), 20cm (201.2 ppb), 42cm (125.8<br />

ppb), 70cm (61.9 ppb), 80cm (49.4 ppb), 105cm<br />

(58.6 ppb), 135cm (54.6 ppb), and 230cm (72.5 ppb).


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

DISCUSSION<br />

CARBON DATING<br />

The oldest date calculated using 14 C was 8,800 BP at<br />

365cm. Radiocarbon dating <strong>of</strong> this bog from previous<br />

studies demonstrated a maximum age <strong>of</strong> approximate<br />

ly 12,000 yBP (Leopold and Dethier, 2007), which<br />

suggests that our core location may not have been at<br />

<strong>the</strong> deepest or oldest part <strong>of</strong> <strong>the</strong> bog, or <strong>the</strong> core may<br />

not have reached <strong>the</strong> bottom <strong>of</strong> <strong>the</strong> bog.<br />

Accumulation rates in peat bogs from o<strong>the</strong>r studies<br />

range anywhere between 0.015-0.920 mm yr -1<br />

(Bindler, 2003; Biester et al., 2002). The average peat<br />

accumulation rate in North Boulder Creek Bog was<br />

calculated at 0.359 mm yr -1 , indicating a slow to<br />

intermediate deposition rate. This value is likely slow<br />

due to <strong>the</strong> subalpine location <strong>of</strong> <strong>the</strong> bog, where <strong>the</strong>re<br />

are minimal organic inputs.<br />

MERCURY ANALYSIS<br />

Each sample represent approximately 28 years, yet<br />

<strong>the</strong>re is a large gap <strong>of</strong> approximately 140 years be<br />

tween each sample due to <strong>the</strong> 5cm interval, due to <strong>the</strong><br />

difficulty <strong>of</strong> high-resolution sampling. Additionally,<br />

<strong>the</strong>re is typically much more compaction at depth<br />

within bogs. Thus, <strong>the</strong>re may be events within <strong>the</strong>se<br />

gaps <strong>of</strong> time that are not recorded by our mercury<br />

analysis. At <strong>the</strong> top <strong>of</strong> <strong>the</strong> core, however, sampling<br />

was conducted at tighter intervals and <strong>the</strong>re is likely<br />

to be much less compaction.<br />

The mercury concentration data from <strong>the</strong> core show<br />

reliable signals at appropriate depths, matching<br />

closely with results from o<strong>the</strong>r mercury studies <strong>of</strong><br />

peat bogs (Martínez-Cortizas et al., 1999; Biester et<br />

al., 2002; Givelet et al., 2003; Schuster et al., 2002,<br />

Bindler, 2002). Results from this core demonstrate<br />

a set <strong>of</strong> peaks in mercury concentrations in <strong>the</strong> top<br />

75cm depth, with a major drop in mercury concentration<br />

approaching <strong>the</strong> top <strong>of</strong> <strong>the</strong> core above 20cm.<br />

These peaks are all well above 50 ppb, indicating an<br />

additional source <strong>of</strong> mercury in addition to <strong>the</strong> natural<br />

deposition.<br />

From 75-70cm depth, mercury concentration rises<br />

101<br />

Natural Background Level--23.2 ppb<br />

135cm: Aniakchak, AK eruption 3435 BP<br />

Unknown<br />

See Figure 5<br />

75cm: Silver mining in <strong>the</strong> Americas 400 BP<br />

80cm: Ceboruco eruption, 1020 BP<br />

105cm: Okmok, AK eruption 2050 BP<br />

180cm: Indonesian eruptions, 5500-5300 BP<br />

305-345cm: Nor<strong>the</strong>rn Hemisphere climatic shift<br />

}<br />

Figure 4: Mercury concentrations with depth in <strong>the</strong> complete<br />

North Boulder Creek Bog core, correlated to 14 CyBP.<br />

Red dots represent individual samples, and <strong>the</strong> dashed<br />

black line is <strong>the</strong> interpreted flux in concentration.<br />

sharply to 61.9 ppb. According to <strong>the</strong> age-depth<br />

model, 75cm corresponds to an age <strong>of</strong> 509 yBP,<br />

closely following <strong>the</strong> onset <strong>of</strong> silver mining in <strong>the</strong><br />

Americas in <strong>the</strong> 1550s when mercury was used to<br />

amalgamate <strong>the</strong> silver from its natural compounds.<br />

The sharp peak above 75cm depth likely represents<br />

<strong>the</strong> first clear distinction between natural and anthropogenic<br />

mercury deposition. Below this depth, it can<br />

be inferred that <strong>the</strong> majority <strong>of</strong> mercury deposition<br />

was due to natural causes, including dust loads,<br />

volcanic events, and climatic fluxes (Pirrone et al.,<br />

2010; Martínez-Cortizas et al., 1999). Alternatively,<br />

above 75cm it can be inferred that anthropogenicallyinduced<br />

mercury deposition is a significant contribu<br />

tor in addition to <strong>the</strong> natural mercury deposition.<br />

Thus, according to historic data, <strong>the</strong> mercury depos<br />

ited above 70cm would have originated from modern<br />

mining, industrial pollution, waste incineration,<br />

WWII, coal burning, volcanic events, and climatic<br />

shifts (Hylander and Meili, 2003; Pirrone et al.,<br />

2010).<br />

In addition to <strong>the</strong> onset <strong>of</strong> ore mining and processing,<br />

<strong>the</strong> climate was also changing rapidly at this time in<br />

<strong>the</strong> Nor<strong>the</strong>rn Hemisphere. Colder temperatures are<br />

known to sequester higher quantities <strong>of</strong> elemental<br />

–60<br />

0<br />

180<br />

1,800<br />

4,000<br />

6,000<br />

7,300<br />

8,400<br />

9,000<br />

Age (cal. years BP)


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

mercury, thus shifts to a colder climate may have<br />

caused a higher retention <strong>of</strong> mercury in <strong>the</strong> bog near<br />

70 cm (Martínez-Cortizas et al., 1999). The Little<br />

Ice Age took place from 500-250 yBP (Mann et al.,<br />

2009), and could have contributed to <strong>the</strong> higher retention<br />

<strong>of</strong> mercury.<br />

PRE-ANTHROPOGENIC<br />

Below 70cm, anomalous mercury peaks occur at<br />

80cm, 105cm, 135cm, 180cm, and 230cm in <strong>the</strong> core.<br />

From 85 to 80cm depth, mercury concentration rises<br />

to a small peak <strong>of</strong> approximately 50 ppb. At 85cm,<br />

<strong>the</strong> age-depth model estimates an age <strong>of</strong> 1039 yBP,<br />

and at 80cm an estimate <strong>of</strong> 776 BP. At 1020 ± 200<br />

BP, <strong>the</strong> Mexican volcano Ceboruco erupted, releasing<br />

about 1.1 ± 0.08 x 10 10 m 3 <strong>of</strong> volcanic material and<br />

producing an explosion rated as a 6 on <strong>the</strong> Volcanic<br />

Explosivity Index (Smithsonian Institution, Global<br />

Volcanism Program). Given a time window <strong>of</strong> 1039-<br />

776 yBP, <strong>the</strong> mercury peak at 80cm likely resulted,<br />

in part, from <strong>the</strong> 1020 ± 200 BP eruption, which took<br />

place 2,090km away from <strong>the</strong> bog.<br />

The next peak <strong>of</strong> 58.6 ppb Hg is recorded at 105cm,<br />

corresponding to an age <strong>of</strong> 2048 yBP. This peak may<br />

in part be explained by <strong>the</strong> Okmok eruption in <strong>the</strong><br />

Aleutian Islands in 100 ± 50 BC (2050 BP). This<br />

eruption was a VEI 6 and ejected 5.0 ± 1.0 x 10 10 m 3<br />

<strong>of</strong> tephra (Smithsonian Institution). The Okmok Caldera<br />

is located just 4,828km from North Boulder<br />

Creek Bog and may be a contributor to <strong>the</strong> anomalous<br />

mercury peak that occurs just two years after its alculated<br />

eruption date.<br />

The 54.6 ppb Hg peak at 135cm corresponds to an<br />

age <strong>of</strong> 3437 yBP, just following <strong>the</strong> Aniakchak eruption<br />

in Alaska, US. This eruption was rated a VEI 6<br />

and released over 5 x 10 10 m 3 <strong>of</strong> tephra (Smithsonian<br />

Institution). The extreme magnitude <strong>of</strong> <strong>the</strong> Aniakchak<br />

eruption and its close proximity to <strong>the</strong> deposition site<br />

(6,700km) make it a likely candidate for <strong>the</strong> anomalous<br />

peak at 135cm.<br />

Climate may also have an effect on <strong>the</strong> amount <strong>of</strong><br />

mercury deposited in <strong>the</strong> bog between 3438 and 2048<br />

yBP, when <strong>the</strong> two previously mentioned mercury<br />

peaks were likely deposited. The cooler time period<br />

102<br />

between 3,500-2,500 yBP in <strong>the</strong> Nor<strong>the</strong>rn Hemisphere<br />

was characterized by ice rafting in <strong>the</strong> North<br />

Atlantic, high latitude cooling, and alpine glacier<br />

retreat, and is believed to be a period <strong>of</strong> Rapid Climate<br />

Change (RCC) resulting from a decline in solar<br />

output (Mayewski et al., 2004).<br />

At 180cm depth mercury rises to 58.8 ppb, but re<br />

mains at high values between 195-175cm (5932-5056<br />

yBP). This lengthy increase in mercury concentration<br />

may be partly due to <strong>the</strong> combined effects <strong>of</strong> <strong>the</strong> two<br />

eruptions <strong>of</strong> Luzon in Indonesia, which occurred at<br />

5530 and 5500 yBP at Taal and Pinatubo, respectively.<br />

Toge<strong>the</strong>r <strong>the</strong> VEI 6 eruptions released a total<br />

<strong>of</strong> 6.3 x 10 10 m 3 <strong>of</strong> tephra into <strong>the</strong> atmosphere (Smithsonian<br />

Institution). Alternatively, <strong>the</strong> sustained rise<br />

in mercury concentration in this part <strong>of</strong> <strong>the</strong> core may<br />

also be due to a shift to colder temperatures between<br />

6000 and 5000 BP—similar to <strong>the</strong> climate shift that<br />

also took place between 3,500-2,500 yBP. The colder<br />

climates in <strong>the</strong> Nor<strong>the</strong>rn Hemisphere during this<br />

time are similarly attributed to solar variability<br />

(Mayewski et al., 2004).<br />

From 245cm to 230cm depth, ano<strong>the</strong>r peak climbs<br />

from 22.4 ppb Hg to a maximum <strong>of</strong> 72.5 ppb Hg. At<br />

245cm, our age-depth model approximates an age<br />

<strong>of</strong> 7118 yBP (5,168 BC). Sufficient volcanic or climatic<br />

events cannot be correlated with <strong>the</strong> timing <strong>of</strong><br />

this peak; <strong>the</strong>refore <strong>the</strong> mercury concentration at this<br />

depth cannot be attributed to a single point source.<br />

Higher resolution mercury analyses at this depth may<br />

yield more informative data.<br />

Finally, <strong>the</strong>re is an anomalous increase in mercury<br />

concentration between 345 and 305 cm in <strong>the</strong> core,<br />

representing a plateau just slightly above <strong>the</strong> background<br />

concentration. This plateau seems to linger<br />

for approximately 500 years, between 8962 and 8477<br />

years BP, with a slight drop at 325 cm. Nor<strong>the</strong>rn<br />

Hemispheric climate experienced a rapid shift to<br />

colder temperatures at 8,200 years BP, when <strong>the</strong><br />

North Atlantic region received a large meltwater burst<br />

from proglacial lakes, causing both deepwater circulation<br />

and Nor<strong>the</strong>rn Hemisphere temperature regulation<br />

to weaken (Born and Levermann, 2010). Due to <strong>the</strong><br />

improbability <strong>of</strong> sustained volcanic influence during<br />

<strong>the</strong> time at which this plateau appears, it is highly


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

likely that Holocene climate shifts played a large role<br />

in <strong>the</strong> retention <strong>of</strong> mercury in North Boulder Creek<br />

Bog.<br />

ANTHROPOGENIC<br />

Assuming <strong>the</strong> natural background level <strong>of</strong> mercury<br />

deposition has remained constant at this location, <strong>the</strong><br />

anthropogenic input <strong>of</strong> mercury has ranged from<br />

about 13-136 ppb since <strong>the</strong> 1550s. Above 75cm in<br />

<strong>the</strong> core, <strong>the</strong> most prominent peak resides at 20cm<br />

(Fig. 5), which has been fit to o<strong>the</strong>r mercury curves in<br />

order to obtain an accurate date <strong>of</strong> 67 yBP, when<br />

Mount Krakatau erupted in Indonesia on August<br />

26, 1883. The massive eruption was rated a VEI 6<br />

and ejected 2.0 ± 0.2 x 10 10 m 3 <strong>of</strong> tephra (Smithsonian<br />

Institution), and is <strong>the</strong>refore a reliable marker with<br />

which to pinpoint <strong>the</strong> date <strong>of</strong> that peak. There is a<br />

smaller peak just below Krakatau at 42cm depth,<br />

which is interpreted as <strong>the</strong> Tambora eruption <strong>of</strong><br />

1815 in Indonesia—a VEI 7 that erupted 1.6 x 10 11<br />

m 3 <strong>of</strong> tephra. Between <strong>the</strong>se two peaks, <strong>the</strong> mercury<br />

concentration drops significantly before<br />

3.5cm: WWII<br />

20cm: Krakatau<br />

Mt. St. Helens 1980 – -30 BP<br />

38-32cm: Gold Rush<br />

42cm: Tambora<br />

– 5 BP<br />

–67 BP<br />

– 85 BP<br />

– 100 BP<br />

– 135 BP<br />

Figure 5: Zoomed inset from Figure 4. Mercury concentrations<br />

<strong>of</strong> high-resolution samples above 45cm in <strong>the</strong> core,<br />

calibrated to yBP. Red dots represent individual samples,<br />

and <strong>the</strong> dashed black line is <strong>the</strong> interpreted flux in concentration.<br />

Age (yrs BP)<br />

103<br />

Krakatau, and remains at a small plateau between 38<br />

and 32cm (137.2-120.4 ppb). This small plateau,<br />

less concentrated in mercury than <strong>the</strong> peak <strong>of</strong><br />

Krakatau but slightly more than that <strong>of</strong> Tambora, is<br />

interpreted as <strong>the</strong> signal <strong>of</strong> mercury deposited by <strong>the</strong><br />

American Gold Rush from 1850-1865, when mercury<br />

was used as an amalgamator for gold and silver ore<br />

processing.<br />

Above <strong>the</strong> 20cm peak, ano<strong>the</strong>r significant peak<br />

<strong>of</strong> 158.9 ppb resides at 0.5cm. Due to its shallow<br />

position and extremely high mercury signal, this peak<br />

is interpreted as <strong>the</strong> mercury released and deposited<br />

during <strong>the</strong> Mount St. Helens eruption in March <strong>of</strong><br />

1980. The VEI 5 eruption, just 1510km away from<br />

<strong>the</strong> deposition site, released 7.4 x 10 7 m 3 <strong>of</strong> lava and<br />

1.2 x 10 9 m 3 <strong>of</strong> tephra into <strong>the</strong> atmosphere (Smithsonian<br />

Institution). Below <strong>the</strong> Mt. St. Helens peak,<br />

<strong>the</strong>re is a smaller increase in mercury concentration at<br />

3.5cm (126.6 ppb). This small jump is likely a mercury<br />

signal deposited during WWII, when <strong>the</strong> defense<br />

industry was utilizing mercury to manufacture explosives.<br />

Above <strong>the</strong> Mt. St. Helens peak, our data record <strong>the</strong><br />

drop in atmospheric mercury concentration during<br />

<strong>the</strong> past few decades, marked by a total concentration<br />

<strong>of</strong> 86.6 ppb at 0cm, down from a peak <strong>of</strong> 158.9 ppb.<br />

This result is congruent with modern measurements<br />

that have shown a decrease in atmospheric mercury<br />

contributions from anthropogenic sources<br />

(Hylander and Meili, 2003). Given a natural background<br />

deposition <strong>of</strong> 23.2 ppb Hg, our most recent<br />

sample indicates an input <strong>of</strong> 63.4 ppb Hg from human<br />

activity at present, which includes coal combustion,<br />

industrial processes, and waste incineration.<br />

CONCLUSION<br />

As an ombrotrophic peat bog, North Boulder Creek<br />

Bog holds a well-recorded history <strong>of</strong> mercury deposition<br />

since approximately 9,000 yBP. The core used in<br />

this project did not reach <strong>the</strong> oldest portion <strong>of</strong> <strong>the</strong><br />

bog, thus analyses on an additional core in a deeper<br />

location are recommended. A search for tephra using<br />

SEM analysis is recommended at <strong>the</strong> depths<br />

where we believe volcanic signals are located.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

ACKNOWLEDGEMENTS<br />

This project was made possible by funding from <strong>the</strong><br />

National Science Foundation. I would also like to extend<br />

<strong>the</strong> utmost appreciation to Bob Nelson and Matthias<br />

Leopold for <strong>the</strong>ir assistance in tackling <strong>the</strong> bog,<br />

and to David Dethier and Will Ouimet for <strong>the</strong>ir instruction<br />

and advising in <strong>the</strong> field.<br />

REFERENCES<br />

Bindler, R., 2003, Estimating <strong>the</strong> Natural Background<br />

Atmospheric Deposition Rate <strong>of</strong> Mercury Utilizing<br />

Ombrotrophic Bogs in Sou<strong>the</strong>rn Sweden:<br />

Environ. Sci. Technol., 37 (1), p. 40–46.<br />

Biester, H., Kilian, R., Franzen, C., Woda, C., Mangini,<br />

A., and Schöler, H.F., 2002, Elevated mercury<br />

accumulation in a peat bog <strong>of</strong> <strong>the</strong> Magellanic<br />

Moorlands, Chile (53ºS) – an anthropogenic<br />

signal from <strong>the</strong> Sou<strong>the</strong>rn Hemisphere: Earth and<br />

Planetary Science Letters no. 201, p. 609-620.<br />

Born, A., Levermann, A., 2010, The 8.2 ka event:<br />

Abrupt transition <strong>of</strong> <strong>the</strong> subpolar gyre toward a<br />

modern North Atlantic circulation: Geochemistry<br />

Geophysics Geosystems, v. 11, no. 6, p. 1-8.<br />

Givelet, N., Roos-Barraclough, F., and Shotyk, W.,<br />

2003, Predominant anthropogenic sources and<br />

rates <strong>of</strong> atmospheric mercury accumulation in<br />

sou<strong>the</strong>rn Ontario recorded by peat cores from<br />

three bogs: comparison with natural “background”<br />

values (past 8000) years: J. Environ.<br />

Monit., v. 5, p. 939-949.<br />

Hylander, L.D., and Meili, M., 2003, 500 years <strong>of</strong><br />

mercury production: global <strong>annual</strong> inventory by<br />

region until 2000 and associated emissions: The<br />

Science <strong>of</strong> <strong>the</strong> Total Environment, v. 304, p. 13-<br />

27.<br />

Leopold, M., and Dethier, D., 2007, Near surface geophysics<br />

and sediment analysis to precisely date<br />

<strong>the</strong> outbreak <strong>of</strong> glacial Lake Devlin, Front Range<br />

Colorado: Eos Trans. AGU, 88(52), Fall Meet.<br />

Suppl., Abstract H51E– 0809.<br />

104<br />

Livingstone, D.A., 1955, A lightweight piston sampler<br />

for lake deposits: Ecology, v. 36, p. 137-139.<br />

Lodenius, M., Seppänen, A., and Uusi-Rauva, A.,<br />

1983, Sorption and mobilization <strong>of</strong> mercury in<br />

peat soil: Chemosphere, v. 12, p. 1575–1581.<br />

Madole, R.F., 1985, Lake Devlin and Pinedale glacial<br />

history, Front Range, Colorado: Quaternary<br />

Research, v. 25, p. 43-54.<br />

Madsen, P.P., 1981, Peat bog records <strong>of</strong> atmospheric<br />

mercury deposition: Nature, v. 293, p. 127-130.<br />

Mann, M.E., Zhang, Z., Ru<strong>the</strong>rford, S., Bradley,<br />

R.S., Hughes, M.K., Shindell, D., Ammann, C.,<br />

Faluvegi, G., and Ni, F., 2009, Global Signatures<br />

and Dynamical Origins <strong>of</strong> <strong>the</strong> Little Ice Age and<br />

Medieval Climate Anomaly: Science, v. 326, no.<br />

5957, p. 1256-1260.<br />

Martínez-Cortizas, A., Pontevedra-Pombal, X., García-Rodeja,<br />

E., Nóvoa-Muñoz, J.C., and Shotyk,<br />

W., 1999, Mercury in a Spanish peat bog: Archive<br />

<strong>of</strong> climate change and atmospheric metal<br />

deposition: Science, v. 284, p. 939-942.<br />

Mayewski, P.A., Rohling, E.E., Stager, J.C., Karlén,<br />

W., Maasch, K.A., Meeker, L.D., Meyerson,<br />

E.A., Gasse, F., van Kreveld, S., Holmgren, K.,<br />

Lee-Thorp, J., Rosqvist, G., Rack, F., Staubwasser,<br />

M., Schneider, R.R., and Steig, E.J., 2004,<br />

Holocene climate variability: Quaternary Research,<br />

v. 62 (3), p. 243-255.<br />

Nriagu, J.O., 1994, Mercury pollution from <strong>the</strong> past<br />

mining <strong>of</strong> gold and silver in <strong>the</strong> Americas: Science<br />

<strong>of</strong> <strong>the</strong> Total Environment, v. 149, p. 167-<br />

181.<br />

Pirrone, N., Cinnirella, S., Feng, X., Finkelman, R.B.,<br />

Friedli, H.R., Leaner, J., Mason, R., Mukherjee,<br />

A.B., Stracher, G.B., Streets, D.G., and Telmer,<br />

K., 2010, Global mercury emissions to <strong>the</strong> atmosphere<br />

from anthropogenic and natural sources:<br />

Atmos. Chem. Phys., v.10, p. 5951-5964.<br />

Schuster, P.F., Krabbenh<strong>of</strong>t, D.P., Naftz, D.L., Cecil,


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

L.D., Olson, M.L., Dewild, J.F., Susong, D.D.,<br />

Green, J.R., and Abbott, M.L., 2002, Atmospheric<br />

mercury depostion during <strong>the</strong> last 270 years:<br />

a glacial ice core record <strong>of</strong> natural and anthropogenic<br />

sources: Environ. Sci. Technol., v. 36, no.<br />

11, p. 2303-2310.<br />

Smithsonian Institution, National Museum <strong>of</strong> Natural<br />

History, Global Volcanism Program, “Volcanoes<br />

<strong>of</strong> <strong>the</strong> World: Large Eruptions,” http://www.<br />

volcano.si.edu/world/largeeruptions.cfm.<br />

Stuiver, M. and Polach, H. A., 1977, Discussion:<br />

Reporting <strong>of</strong> 14C data: Radiocarbon, v. 19, p.<br />

355-363.<br />

Stuiver, M., 1980. Workshop on 14C data reporting:<br />

Radiocarbon, v. 22, p. 964-966.<br />

Vandal, G.M., Fitzgerald, W.F., Boutron, C.F., and<br />

Candelone, J., 1993, Variations in mercury deposition<br />

to Antarctica over <strong>the</strong> past 34,000 years:<br />

Nature, v. 362, p. 621-623.<br />

105


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

THE DISTRIBUTION OF PHOSPHORUS IN ALPINE AND<br />

UPLAND SOILS OF THE BOULDER CREEK, COLORADO<br />

CATCHMENT<br />

HAYLEY CORSON-RIKERT, Wesleyan University<br />

Research Advisor: Timothy Ku<br />

INTRODUCTION<br />

Biological growth within terrestrial ecosystems is<br />

generally limited by <strong>the</strong> concentration <strong>of</strong> nitrogen,<br />

phosphorus, or both (Sato et al., 2009). Investigation<br />

into <strong>the</strong> availability <strong>of</strong> both <strong>the</strong>se macronutrients<br />

in modern day alpine environments is important as<br />

N + P availability determines how such ecosystems<br />

will respond to climatic changes and anthropogenic<br />

alterations <strong>of</strong> soil chemistry (Wu et al., 2006). Recent<br />

studies have shown that enhanced rates <strong>of</strong> nitrogen<br />

deposition can force alpine systems that are typically<br />

N-limited to become P-limited, especially when P is<br />

efficiently cycled, making investigation into soil P<br />

dynamics yet more important (Sievering et al., 1996;<br />

Hedin et al., 2003; Vitousek et al., 2010).<br />

Distribution <strong>of</strong> soil P occurs through geochemical<br />

and biochemical pathways, and is controlled by <strong>the</strong><br />

demand for and supply <strong>of</strong> P in soil horizons (McGill<br />

and Cole, 1981). Crystalline or primary mineral P in<br />

deeper soils represents a long-term soil P reservoir,<br />

whereas secondary mineral forms and in particular<br />

labile forms are cycled more rapidly in upper and/or<br />

surface horizons (Walker and Syers, 1976). On short<br />

timescales, <strong>the</strong> availability <strong>of</strong> labile soil P to plants is<br />

dependent on a number <strong>of</strong> factors, including temperature,<br />

moisture, aeration, and soil microorganism activity<br />

(Tate & Salcedo, 1988). In <strong>the</strong> long term, labile P<br />

availability is dependent on <strong>the</strong> state <strong>of</strong> soil development,<br />

which in turn is determined by soil residence<br />

time and <strong>the</strong> rate <strong>of</strong> chemical and physical wea<strong>the</strong>ring<br />

(Walker and Syers, 1976; Porder et al., 2007).<br />

In this study, I examine <strong>the</strong> soil P reservoirs <strong>of</strong> four<br />

soil pr<strong>of</strong>iles across an elevation gradient in Boulder<br />

County, Colorado, in order to better understand<br />

<strong>the</strong> patterns <strong>of</strong> and controls on soil P distribution in<br />

alpine environments. The four selected pr<strong>of</strong>iles are<br />

a subset <strong>of</strong> a broader set <strong>of</strong> studied soils in <strong>the</strong> Boulder<br />

Creek NCZO, and represent a range <strong>of</strong> elevations and<br />

climatic conditions. From greatest to least elevation,<br />

<strong>the</strong> sites are GLV, in <strong>the</strong> Green Lakes Valley; SLM, at<br />

<strong>the</strong> moraine below Silver Lake; UGG, in upper Gordon<br />

Gulch; and Betasso, in <strong>the</strong> Betasso Preserve (Table 1).<br />

The soils at GLV and SLM are relatively stable, with<br />

minimal soil movement, while <strong>the</strong> UGG and Betasso<br />

pr<strong>of</strong>iles are marked by buried horizons, which represent<br />

discontinuities in <strong>the</strong> soil sequence. Mean <strong>annual</strong><br />

temperature near <strong>the</strong> highest site averages -3.7ºC, while<br />

average <strong>annual</strong> temperature at Betasso are about 10ºC<br />

(Niwot Ridge LTER; NOAA). Annual precipitation at<br />

this lower altitude is about 40 cm, while precipitation at<br />

<strong>the</strong> continental divide above GLV can amount to more<br />

than 100 cm <strong>annual</strong>ly (Table 1; Birkeland et al., 2003).<br />

METHODS<br />

In July and August 2010, soils from 31 sites were collected<br />

from newly scraped exposures or fresh soil pits.<br />

Collected samples were stored in plastic bags vacated<br />

<strong>of</strong> air in order to best preserve field moisture. Soil pH<br />

in water and soil moisture were determined by standard<br />

methods (Carter and Gregorich, 2008). Total carbon<br />

and nitrogen concentrations were determined on a<br />

Thermo Flash 1112 Elemental Analyzer. Given <strong>the</strong> lack<br />

<strong>of</strong> carbonate minerals in <strong>the</strong>se soils, total carbon (TC)<br />

is assumed to equal total organic carbon (TOC). Bulk<br />

chemistry analysis for metals was determined by ICP-<br />

OES techniques at SGS Mineral Services, after dissolving<br />

soils in a four-acid digest (HCl/HNO 3 /HF/HClO 4 ).<br />

The digestion may not have completely dissolved very<br />

recalcitrant mineral phases. On 21 samples, soil P<br />

pools were determined by a modified Hedley sequential<br />

extraction procedure (Figure 1; Hedley et al., 1982;<br />

Ruttenberg, 1992; Tiessen and Moir, 1993). Inorganic<br />

(Pi) and total phosphorus (Pt) concentrations were<br />

determined by spectrophotometry methods <strong>of</strong> Murphy<br />

112


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

and Riley (1962) using a Beckman Coulter DU5300<br />

at a wavelength <strong>of</strong> 885 nm. Organic phosphorus (Po)<br />

concentrations were determined by subtraction <strong>of</strong> Pi<br />

from Pt.<br />

Figure 1. Flowchart <strong>of</strong> sequential phosphorus extraction<br />

methodology. The procedure is a modification <strong>of</strong> Tiessen<br />

and Moir (1993), with <strong>the</strong> ashing step <strong>of</strong> Ruttenberg<br />

(1992). Pi = inorganic phosphorus, and Pt = total phosphorus.<br />

Autoclaving conditions were 121ºC, 17 psi, for 50<br />

minutes.<br />

Measured extractable P fractions were grouped to<br />

obtain operationally-defined soil pools: Exchangeable<br />

P (NaHCO3 Pi); Organic P (NaHCO 3 Po + NaOH Po<br />

+ C. HCl Po); Fe-bound P (NaOH Pi); Ca-bound P<br />

(1M HCL Pi + 1M HCl Po); Recalcitrant P (C. HCl<br />

Pi); and Highly Recalcitrant P (Ashed Pi + Ashed Po)<br />

(Tiessen and Moir, 1993). The percentage <strong>of</strong> initial<br />

total P remaining in individual horizons was calculated<br />

relative to Al as follows, where X = sample<br />

horizon and Y = parent material (or deepest available<br />

horizon): Initial total P concentration equals [Al]X<br />

x ([Pt]Y/[Al]Y). The % <strong>of</strong> initial total P remaining<br />

113<br />

equals 100 x ([Pt]X/[Pt]Y) (Vitousek et al., 2004,<br />

Supp. Mat.). The percent <strong>of</strong> initial Ca-bound Pi was<br />

calculated in <strong>the</strong> same manner.<br />

RESULTS AND DISCUSSION<br />

General soil properties<br />

A summary <strong>of</strong> results is shown in Table 1. In <strong>the</strong>se<br />

four soil pr<strong>of</strong>iles, soil pH increases with depth, with<br />

surface horizons displaying pH ranging from 4.43 to<br />

5.61 and base horizons pH ranging from 5.51 to 6.05.<br />

The greater acidity <strong>of</strong> surface horizons is typically <strong>the</strong><br />

result <strong>of</strong> organic matter decay, which lowers <strong>the</strong> pH<br />

<strong>of</strong> soil pore waters (Twidale, 1990). This assumption<br />

is supported by <strong>the</strong> consistently high TOC concentrations<br />

in surface horizons (Figure 2). As expected,<br />

TOC is correlated with organic N throughout all horizons,<br />

and C:N, C:P 0 , N:P 0 , and soil moisture values<br />

decrease with depth (Table 1; Figure 2). P organic:P<br />

inorganic decreases with depth, demonstrating <strong>the</strong><br />

transition from surface horizons rich in organic and<br />

plant available P to deeper horizons dominated by<br />

primary and secondary mineral P (Figure 2). Soil<br />

concentrations <strong>of</strong> Al increase with depth at all sites<br />

(Table 1).<br />

Table 1. Table <strong>of</strong> study site information and soil properties.<br />

Soil ages from Dethier et al., unpublished data -- *<br />

denotes exposure ages measured by OSL techniques, †<br />

denotes age based on CRN techniques.<br />

Total Soil P<br />

Total P concentration pr<strong>of</strong>iles are presented in <strong>the</strong><br />

right-hand column <strong>of</strong> Figures 3 and 4. Total soil P<br />

concentration varies from 198 to 2853 ug/g across all<br />

horizons. These values are comparable to those <strong>of</strong><br />

o<strong>the</strong>r alpine soil studies, which were generally be


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 2. Soil depth vs. TOC, C:N. C:P 0 , N:P 0 , and P<br />

organic:P inorganic. Note log scale on x-axis.<br />

tween 121 to 2540 ug/g (Table 1; e.g. Makarov et al.,<br />

1997). Total soil phosphorus concentrations show no<br />

discernable correlation with elevation. It is assumed<br />

that <strong>the</strong> variation in moisture regimes, parent material,<br />

soil residence time, and wea<strong>the</strong>ring rates is too great<br />

for a clear pattern to emerge.<br />

The Development <strong>of</strong> Soil P Reservoirs<br />

Soil P is widely held to exist in three main pools:<br />

primary inorganic P, secondary inorganic P, and organic<br />

P. Walker and Syers (1976) presented a model<br />

in which primary inorganic P (mineral P) is abundant<br />

in early stages <strong>of</strong> soil development and, as wea<strong>the</strong>ring<br />

progresses, is transformed into organic forms<br />

and sorbed to secondary minerals. Thus, <strong>the</strong> primary<br />

mineral P fraction becomes depleted as <strong>the</strong> organic P<br />

and secondary mineral P fractions are enriched. At<br />

first, a portion <strong>of</strong> this sorbed inorganic secondary<br />

mineral P is exchangeable, or plant available, but this<br />

labile fraction is later diminished with <strong>the</strong> exhaustion<br />

<strong>of</strong> <strong>the</strong> primary mineral P reservoir. In more highly<br />

wea<strong>the</strong>red pr<strong>of</strong>iles, labile P is fur<strong>the</strong>r depleted due<br />

to <strong>the</strong> progressive transformation <strong>of</strong> <strong>the</strong> secondary<br />

mineral P into occluded, or recalcitrant, forms that are<br />

biologically unavailable. Within soil pr<strong>of</strong>iles, horizon<br />

development progresses vertically, with upper horizons<br />

originating from parent material. Each horizon,<br />

<strong>the</strong>refore, represents a stage and/or type <strong>of</strong> soil development,<br />

and soil P distribution within <strong>the</strong>se horizons<br />

114<br />

should reflect <strong>the</strong>ir position on <strong>the</strong> continuum. In<br />

effect, primary mineral P is expected to decrease from<br />

deeper to surface horizons, while organic, secondary<br />

mineral, and labile P are expected to increase<br />

(Walker and Syers, 1976; Stewart and Tiessen, 1987;<br />

Crews, 1995; Porder et al., 2007). In stable soils, this<br />

increase in plant-available P and decrease in primary<br />

mineral P occurs as total P is diminished due to net P<br />

removal by wea<strong>the</strong>ring. This pattern is most visible<br />

in sites that experience high <strong>annual</strong> levels <strong>of</strong> precipitation,<br />

due to <strong>the</strong> enhancement <strong>of</strong> soil redox processes<br />

and thus <strong>the</strong> quickening <strong>of</strong> mineral P dissolution and<br />

removal (Miller et al., 2001; Hedin et al., 2003).<br />

Distribution <strong>of</strong> Soil P Pools in <strong>the</strong> Boulder Creek<br />

Catchment<br />

Figures 3 and 4 show <strong>the</strong> distribution <strong>of</strong> P fractions,<br />

total P, and calculated values <strong>of</strong> % P remaining at <strong>the</strong><br />

four study sites. The importance <strong>of</strong> various soil P<br />

transformation processes is reflected in <strong>the</strong> distribution<br />

<strong>of</strong> <strong>the</strong>se soil P pools, and is impacted by <strong>the</strong> extent<br />

to which <strong>the</strong> soil pr<strong>of</strong>ile has been disturbed during<br />

development (Beck and Elsenbeer, 1999). Buried<br />

horizons at UGG and Betasso indicate that <strong>the</strong>se soils<br />

have experienced more soil movement than <strong>the</strong> more<br />

stationary pr<strong>of</strong>iles <strong>of</strong> GLV and SLM. At GLV and<br />

SLM, Ca-bound Pi generally increases with depth and<br />

exchangeable P decreases with depth. The O and A<br />

horizons at GLV are <strong>the</strong> exception to this pattern, as<br />

<strong>the</strong>y have proportionally greater concentrations <strong>of</strong> Cabound<br />

P than <strong>the</strong> horizons immediately below. These<br />

elevated levels <strong>of</strong> primary mineral P, coupled with <strong>the</strong><br />

concomitant rise <strong>of</strong> remaining initial total P to percentages<br />

greater than 100, points to an external input<br />

<strong>of</strong> comparatively unwea<strong>the</strong>red hillslope colluvium or<br />

eolian material. This is supported by <strong>the</strong> P organic:P<br />

inorganic ratios <strong>of</strong> <strong>the</strong> O and A horizons (Figure 2),<br />

which are slightly lower than in <strong>the</strong> horizons immediately<br />

below, suggesting that <strong>the</strong>se upper horizons are<br />

less wea<strong>the</strong>red than <strong>the</strong> B horizons below (Tate and<br />

Salcedo, 1988). Soil P distribution within <strong>the</strong>se GLV<br />

B horizons instead appears to be <strong>the</strong> result <strong>of</strong> continued<br />

wea<strong>the</strong>ring, as a net loss <strong>of</strong> total P due to mineral<br />

dissolution is apparent from <strong>the</strong> Cu to Bw1 horizon.<br />

Site SLM, due to its stability, has <strong>the</strong> most standard<br />

distribution <strong>of</strong> soil P fractions, with a clear inverse


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 3. GLV and SLM Soil P reservoirs, with <strong>the</strong><br />

left-hand chart depicting <strong>the</strong> relative percentage <strong>of</strong> each<br />

P-reservoir per soil horizons, and <strong>the</strong> right-hand chart<br />

illustrating total P (bars, lower x-axis) and <strong>the</strong> % <strong>of</strong> initial<br />

total P and Ca-bound Pi remaining in each horizon, relative<br />

to Al (upper x-axis).<br />

relationship between <strong>the</strong> organic P and Ca-bound P<br />

fractions. This clean pattern indicates that soil surface<br />

horizons are developed almost entirely from <strong>the</strong><br />

parent material, though some eolian deposition may<br />

have altered surface horizon composition. Here, like<br />

in <strong>the</strong> lower GLV horizons, total soil P is highest at<br />

depth and decreases above <strong>the</strong> parent material horizon.<br />

Measurements <strong>of</strong> % initial total P remaining<br />

also decrease, suggesting that P is consistently being<br />

removed from <strong>the</strong> soil system throughout all horizons,<br />

leading to a decrease in Ca-bound P and a subsequent<br />

enrichment <strong>of</strong> <strong>the</strong> organic P fraction.<br />

The lower two studies sites, UGG and Betasso, have<br />

a more complex history than GLV and SLM, as both<br />

contain buried horizons. At UGG, though organic P<br />

content decreases with depth, as expected, and Cabound<br />

P concomitantly increases, <strong>the</strong>re is a clear<br />

difference between buried and current soil horizons.<br />

Total P decreases sharply between <strong>the</strong> lower bBt2 and<br />

<strong>the</strong> bBt above it. Similarly, Ca-bound P is at least<br />

70% <strong>of</strong> total P in <strong>the</strong> lowermost horizons, but only<br />

~7% <strong>of</strong> total P in <strong>the</strong> Bt1, Cox, and Bw horizons.<br />

This low total P content in <strong>the</strong>se middle three horizons,<br />

coupled with <strong>the</strong>ir corresponding low percentage<br />

<strong>of</strong> primary mineral P, and high percentage <strong>of</strong> organic<br />

P and recalcitrant P indicate that <strong>the</strong>se horizons<br />

115<br />

are highly wea<strong>the</strong>red, despite <strong>the</strong>ir relatively young<br />

exposure age (Table 1). This conclusion, in turn, suggests<br />

that that <strong>the</strong>se upper horizons have ei<strong>the</strong>r experienced<br />

intense and rapid wea<strong>the</strong>ring, or formed from<br />

already wea<strong>the</strong>red material that was transported to<br />

this site, burying <strong>the</strong> lowermost bBt2 and CRt horizons<br />

that had formed in situ. Soil exposure ages support<br />

this last conclusion, as <strong>the</strong> two bottom horizons<br />

were last exposed 20,000 years ago, while <strong>the</strong> upper<br />

‘moved’ horizons were exposed much more recently,<br />

roughly 2,000 years ago (Table 1). Importantly,<br />

within <strong>the</strong> two soil brackets above, {CRt-bBt2} and<br />

{bBt-A}, <strong>the</strong> total soil P and % P remaining do not<br />

diminish with decreasing depth, indicating no net P<br />

loss. This is in contrast to <strong>the</strong> net P loss at <strong>the</strong> higher<br />

GLV and SLM pr<strong>of</strong>iles. The relatively larger fraction<br />

<strong>of</strong> Ca-bound P in <strong>the</strong> surface A horizon suggests that<br />

soil P distribution in <strong>the</strong> near surface environment is<br />

skewed by an external influx <strong>of</strong> unwea<strong>the</strong>red material<br />

rich in primary mineral P (Figure 4).<br />

Figure 4. UGG and Betasso Soil P reservoirs, with <strong>the</strong><br />

left-hand chart depicting <strong>the</strong> relative percentage <strong>of</strong> each<br />

P-reservoir per soil horizons, and <strong>the</strong> right-hand chart<br />

illustrating total P (bars, lower x-axis) and <strong>the</strong> % <strong>of</strong> initial<br />

total P and Ca-bound Pi remaining in each horizon, relative<br />

to Al.<br />

The Betasso soil pr<strong>of</strong>ile shows no similar enrichment<br />

<strong>of</strong> Ca-bound P in surface horizons, but contains a thin<br />

O horizon that is heavily enriched in organic P. This<br />

horizon is primarily composed <strong>of</strong> fresh and decaying<br />

plant litter and needles. Below <strong>the</strong> O horizon, organic


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

P levels do not increase greatly, total P increases only<br />

slightly, and Ca-bound P remains a dominant portion<br />

<strong>of</strong> total soil P -- suggesting that <strong>the</strong> lower A, B, bA,<br />

and bBt horizons have experienced little wea<strong>the</strong>ring<br />

despite <strong>the</strong>ir age <strong>of</strong> 5 to 12 kyr. Given <strong>the</strong> comparatively<br />

low <strong>annual</strong> levels <strong>of</strong> rainfall at this site, ~40<br />

cm, this slow soil P development is likely due to <strong>the</strong><br />

limited percolation <strong>of</strong> moisture to deeper horizons,<br />

which would limit both chemical wea<strong>the</strong>ring and soil<br />

microbial activity.<br />

CONCLUSIONS<br />

Soil P pools at <strong>the</strong> four sites can be explained by<br />

continued wea<strong>the</strong>ring and patterns <strong>of</strong> soil movement.<br />

SLM shows <strong>the</strong> most consistent trend in soil development,<br />

with surface horizons enriched in exchangeable<br />

and organic P and deeper portions <strong>of</strong> <strong>the</strong> pr<strong>of</strong>ile<br />

enriched in Ca-bound P. GLV has a similar pr<strong>of</strong>ile,<br />

except that <strong>the</strong> upper layer likely contains relocated<br />

primary mineral P. This addition <strong>of</strong> outside material<br />

is also evident in <strong>the</strong> A and Bw horizons <strong>of</strong> <strong>the</strong> UGG<br />

pr<strong>of</strong>ile, suggesting that <strong>the</strong> relocation <strong>of</strong> primary mineral<br />

P by ei<strong>the</strong>r hillslope removal or eolian deposition<br />

may be an important factor in soil P distribution and<br />

development in surface soil environments across <strong>the</strong><br />

Front Range gradient. The Betasso site is relatively<br />

unwea<strong>the</strong>red, with little accumulation <strong>of</strong> organic P in<br />

<strong>the</strong> A horizon. This may be <strong>the</strong> result <strong>of</strong> a low degree<br />

<strong>of</strong> wea<strong>the</strong>ring experienced by soils at this altitude.<br />

Overall, wea<strong>the</strong>ring appears to be more intense at <strong>the</strong><br />

higher, wetter alpine sites <strong>of</strong> GLV and SLM, where a<br />

considerable fraction <strong>of</strong> soil P has been lost, than at<br />

UGG and Betasso, though <strong>the</strong> relocation <strong>of</strong> unwea<strong>the</strong>red<br />

and wea<strong>the</strong>red soil material, as seen at both SLM<br />

and UGG, serves to complicate this trend.<br />

REFERENCES<br />

Beck, M., and Elsenbeer, H., 1999, Biogeochemical<br />

cycles <strong>of</strong> soil phosphorus in sou<strong>the</strong>rn Alpine<br />

spodosols: Geoderma, vol. 91, p. 249-260<br />

Birkeland, P., Shroba, R., Burns, S., Price, A., and<br />

Tonkin, P., 2003, Integrating soils and geomorphology<br />

in mountains—an example from <strong>the</strong><br />

Front Range <strong>of</strong> Colorado: Geomorphology, v. 55,<br />

p. 329-344<br />

116<br />

Carter, M., and Gregorich, E., editors, 2008, Soil<br />

Sampling and Methods <strong>of</strong> Analysis: CRC Press,<br />

1224 p.<br />

Crews, T., Kitayama, K., Fownes, J., Riley, R., Herbert,<br />

D., Mueller-Dombois, R., and Vitousek, P.,<br />

1994, Changes in Soil Phosphorus Fractions and<br />

Ecosystem Dynamics Across and Long Chronosequence<br />

in Hawaii: Ecology, vol. 76, no. 5,<br />

p.1407-1424<br />

Hedin, L., Vitousek, P., and Matson, P., 2003, Nutrient<br />

Losses over Four Million Years <strong>of</strong> Tropical<br />

Forest Development: Ecology, vol. 84, no. 9, p.<br />

2231-2255<br />

Hedley, M., Stewart, J., and Chauhan, B., 1982,<br />

Changes in inorganic and organic soil phosphorus<br />

fractions induced by cultivation practices and<br />

laboratory incubations: Soil Sci. Soc. Am. J., vol.<br />

46, p. 970-976<br />

McGill, W., and Cole, C., 1981, Comparative Aspects<br />

<strong>of</strong> Cycling <strong>of</strong> Organic C, N, S and P through Soil<br />

Organic Matter: Geoderma, vol. 26, p. 267-286<br />

Makarov, M., Malysheva, T., Haumaier, L., Alt, H.,<br />

and Zech, W., 1997, The forms <strong>of</strong> phosphorus<br />

in humic and fulvic acids <strong>of</strong> a toposequence <strong>of</strong><br />

alpine soils in <strong>the</strong> nor<strong>the</strong>rn Caucasus: Geoderma,<br />

vol, 80, p. 61-73<br />

McGill, W., and Cole, C., 1981, Comparative Aspects<br />

<strong>of</strong> Cycling <strong>of</strong> Organic C, N, S and P through Soil<br />

Organic Matter: Geoderma, vol. 26, p. 267-286<br />

Miller, A., Schuur, E., and Chadwick, O., 2001,<br />

Redox control <strong>of</strong> phosphorus pools in Hawaiian<br />

montane forest soils: Geoderma, vol. 102, p.<br />

219-237<br />

Murphy, J., Riley, J., 1962, A modified single solution<br />

method for <strong>the</strong> determination <strong>of</strong> phosphate<br />

in natural waters: Anal. Chim. Acta., vol. 27, p.<br />

31-36<br />

Niwot Ridge LTER, “Site Information.” Niwot Ridge<br />

LTER. Web. 12 Dec. 2010.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

.<br />

NOAA, “Boulder Monthly Mean Temperature<br />

1897-present.” NOAA Earth System Research<br />

Laboratory. Web. 13 Dec. 2010. .<br />

Porder, S., Vitousek, P., Chadwick, O., Chamberlain,<br />

C., and Hilley, G., 2007, Uplift, Erosion, and<br />

Phosphorus Limitation in Terrestrial Ecosystems:<br />

Ecosystems, vol. 10, p. 158-170<br />

Ruttenberg, K., 1992, Development <strong>of</strong> a Sequential<br />

Extraction Method for Different Forms <strong>of</strong> Phosphorus<br />

in Marine Sediments: Limnology and<br />

Oceanography, vol. 37, no. 7, p. 1460-1482<br />

Sato, S., Neves, E., Solomon, D., Liang, B., and<br />

Lehmann, J., 2009, Biogenic calcium phosphate<br />

transformation in soils over millennial time<br />

scales: J Soils Sediments, vol. 9, p. 194-205<br />

Sievering, H., Rusch, D., and Marquez, L., 1996,<br />

Nitric acid, particulate nitrate, and ammonium in<br />

<strong>the</strong> continental free troposphere: Nitrogen deposition<br />

to an alpine tundra ecosystem: Atmospheric<br />

Environment, vol. 30, no. 14, p. 2527-2537<br />

Stewart, J., and Tiessen, H., 1987, Dynamics <strong>of</strong> soil<br />

organic phosphorus: Biogeochemistry, vol. 4, p.<br />

41-60<br />

Tate, K., and Salcedo, I., 1988, Phosphorus control<br />

<strong>of</strong> soil organic matter accumulation and cycling:<br />

Biogeochemistry, vol. 5, p. 99-107<br />

Tiessen, H., and Moir, J., 1993, Characterization <strong>of</strong><br />

available phosphorus by sequential extraction, in<br />

Carter, M. (Ed.), Soil Sampling and Method <strong>of</strong><br />

Analysis: Lewis, Chelsea, MI, p. 75-86<br />

Twidale, C., 1990, Wea<strong>the</strong>ring, soil development, and<br />

landforms: GSA Special Paper 252, p. 29-50<br />

Vitousek., P., Ladeforged, T., Kirch, P., Hartshorn, A.,<br />

Graves, M., Hotchkis, S., Tuljapurkar, S., and<br />

Chadwick, O., 2004, Supplementary Material<br />

117<br />

to Soils, Agriculture, and Society in Precontact<br />

Hawaii: Science, vol. 304, p. 1665-1669<br />

Vitousek, P., Porder, S., Houlton, B., and Chadwick,<br />

O., 2010, Terrestrial phosphorus limitation:<br />

mechanisms, implications, and nitrogen-phosphorus<br />

interactions: Ecological Applications,<br />

Vol. 20, No. 1, p. 5-15<br />

Walker, T., and Syers, J., 1976, The Fate <strong>of</strong> Phosphorus<br />

During Pedogenesis: Geoderma, vol. 15, p.<br />

1-19<br />

Wu, G., Wei, J., Deng, H., and Zhao, J, 2006, Nutrient<br />

cycling in an Alpine tundra ecosystem on<br />

Changbai Mountain, Nor<strong>the</strong>ast China: Applied<br />

Soil Ecology, vol. 32, p. 199-209


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

EXAMINING KNICKPOINTS IN THE BOULDER CREEK<br />

CATCHMENT, COLORADO<br />

EVAN N. DETHIER, Williams College<br />

Research Advisor: David P. Dethier<br />

INTRODUCTION<br />

The apparent stasis <strong>of</strong> our current landscape belies <strong>the</strong><br />

constant change it has undergone for millions <strong>of</strong> years<br />

before <strong>the</strong> present. The landscape continues to transform<br />

as erosional denudation balances slow uplift <strong>of</strong><br />

rock material. Interaction between tectonics, climate,<br />

and surface processes are linked by channels, and<br />

<strong>the</strong>ir transmission <strong>of</strong> signals through <strong>the</strong> landscape.<br />

(Whipple and Tucker, 1999; Zaprowski et al, 2005;<br />

Wobus et al, 2010).<br />

In rapidly eroding landscapes, a knickpoint—a steep<br />

reach bounded on both sides by relatively shallower<br />

reaches—is a physical indicator <strong>of</strong> transient channel<br />

response to climate and tectonic signals. In postorogenic<br />

landscapes, knickpoints may also reflect<br />

rock strength or slow, complex response to external<br />

forcing. Explanations for <strong>the</strong> existence <strong>of</strong> knickpoints<br />

are varied, but consensus holds that <strong>the</strong>se features<br />

are transitory, migrating in a front from an initial<br />

source at <strong>the</strong> head or foot <strong>of</strong> <strong>the</strong> channel (Crosby and<br />

Whipple, 2006; Wobus et al, 2010). If <strong>the</strong> knickpoint<br />

is migrating upstream, <strong>the</strong> topography below will<br />

have undergone greater adjustment than <strong>the</strong> topography<br />

upstream <strong>of</strong> <strong>the</strong> knickpoint, and vice versa if<br />

<strong>the</strong> knickpoint is migrating downstream (Crosby and<br />

Whipple, 2006; Wobus et al, 2010). Studying <strong>the</strong><br />

location and characteristics <strong>of</strong> knickpoints can help us<br />

understand <strong>the</strong> dynamics <strong>of</strong> topographic response to<br />

different forcings (Crosby and Whipple, 2006; Wobus<br />

et al, 2010).<br />

Most knickpoint research has focused on regions<br />

with high uplift rates, weak rock, and rapid landscape<br />

evolution. In contrast, <strong>the</strong> Front Range in Colorado—<br />

in <strong>the</strong> interior <strong>of</strong> <strong>the</strong> North American continent—is<br />

a region with low uplift and low precipitation, relatively<br />

strong rock, and thus comparatively low rates<br />

<strong>of</strong> incision.<br />

Studying streams in <strong>the</strong> Front Range can provide<br />

insight into <strong>the</strong> behavior <strong>of</strong> a slowly evolving environment.<br />

Pr<strong>of</strong>iles <strong>of</strong> steady-state channels and hillslopes<br />

have different concavity. Steady-state channels are<br />

concave up, and steady-state hillslopes are concave<br />

down in <strong>the</strong> upper reach and concave up near <strong>the</strong> channel<br />

(Anderson, 2008). These shapes are disturbed by <strong>the</strong><br />

introduction <strong>of</strong> a knickpoint to <strong>the</strong> system, or prevented<br />

from occurring by a permanent knickpoint. As knickpoints<br />

travel through <strong>the</strong> river system, <strong>the</strong> long pr<strong>of</strong>ile<br />

<strong>of</strong> <strong>the</strong> river becomes locally convex, with shallow<br />

reaches bounding a steep section. Adjacent hillslopes<br />

respond to <strong>the</strong> resulting changes in boundary conditions,<br />

<strong>of</strong>ten exhibiting greater concavity and roughness<br />

as a result <strong>of</strong> increased incision. Examining <strong>the</strong>se features<br />

around knickpoints allows us to characterize <strong>the</strong><br />

ways a landscape moves back towards equilibrium.<br />

The thrust <strong>of</strong> this project is to identify knickpoints,<br />

characterize <strong>the</strong> morphology <strong>of</strong> <strong>the</strong> channels that include<br />

<strong>the</strong> knickpoints, and describe <strong>the</strong> nature <strong>of</strong> <strong>the</strong><br />

adjacent hillslopes. After this identification, I hope to<br />

draw conclusions by contrasting <strong>the</strong> basin area above<br />

<strong>the</strong> knickpoint with <strong>the</strong> area below, and comparing <strong>the</strong><br />

area within <strong>the</strong> knickpoint with <strong>the</strong> area that bounds it.<br />

RESEARCH AREA<br />

I conducted my field research in <strong>the</strong> Middle Boulder<br />

Creek watershed (Fig. 1). I focused on Middle Boulder<br />

Creek and its tributaries, particularly two small channels:<br />

Gordon Gulch and Betasso Gulch.<br />

The Colorado Front Range, which formed and evolved<br />

during <strong>the</strong> Laramide orogeny from 65 to 40 Ma, extends<br />

westward from <strong>the</strong> piedmont at Boulder to <strong>the</strong><br />

Continental Divide. Initially formed by rapid uplift,<br />

since 40 Ma <strong>the</strong> Front Range has been tectonically inac-<br />

106


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 1. Map showing <strong>the</strong> Boulder Creek Catchment,<br />

outlined in black. Betasso Gulch and Gordon Gulch are<br />

shown in orange. Middle Boulder Creek is traced to its<br />

headwaters at <strong>the</strong> Continental Divide, and to its outlet in<br />

<strong>the</strong> South Platte River on <strong>the</strong> plains.<br />

tive, with isostatic response driving rock uplift in <strong>the</strong><br />

region (Kellogg et al, 2008). Melt from small glaciers<br />

and seasonal snowpack in upland areas runs down<br />

numerous tributaries <strong>of</strong> Middle Boulder Creek and<br />

eventually empties onto <strong>the</strong> Great Plains. The relatively<br />

high relief alpine and subalpine zone has been<br />

sculpted by glacial and periglacial activity. Below<br />

~2500 m, a rolling upland landscape is deeply incised<br />

by narrow canyons that extend up from <strong>the</strong> piedmont.<br />

Due to high evapotranspiration rates during <strong>the</strong> summer<br />

months, many small drainages east <strong>of</strong> <strong>the</strong> glacial<br />

limit are ephemeral. The vegetated low-relief surface<br />

has not been affected by glaciation, and away from<br />

<strong>the</strong> deep canyons, locally thick wea<strong>the</strong>red deposits<br />

overlie fresh bedrock (Birkeland et al 2003). Hillslope<br />

and channel processes govern landscape evolution in<br />

this montane zone, which includes <strong>the</strong> Gordon Gulch<br />

and Betasso Gulch catchments.<br />

Climate in <strong>the</strong> Front Range is largely dependent on<br />

elevation and proximity to <strong>the</strong> continental divide,<br />

which runs North-South roughly 30 km west <strong>of</strong> Boulder.<br />

Annual precipitation is highest near <strong>the</strong> divide<br />

and low on <strong>the</strong> plains (PRISM Climate Group). The<br />

orographic effect is strong in <strong>the</strong> winter, with significant<br />

snow accumulation near <strong>the</strong> continental divide<br />

and lighter precipitation down low. Frequent, low<br />

elevation thunderstorms during <strong>the</strong> summer months<br />

mitigate this orographic effect.<br />

METHODS<br />

I surveyed channel longitudinal pr<strong>of</strong>iles in <strong>the</strong> field,<br />

while measuring numerous river parameters, hillslope<br />

character, and rock strength, and used DEMs based on<br />

Lidar (Gordon Gulch and Betasso Gulch) and USGS<br />

maps (Boulder Creek) to characterize channels and<br />

hillslopes.<br />

At <strong>the</strong> Gordon Gulch and Betasso catchments, we<br />

surveyed <strong>the</strong> longitudinal pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> stream, using<br />

a tripod-mounted Tru-Pulse 360 Laser Rangefinder.<br />

All measurements were parallel to <strong>the</strong> stream channel.<br />

I recorded <strong>the</strong> vertical distance, horizontal distance,<br />

and azimuth for each section <strong>of</strong> stream. I also measured<br />

channel width and estimated bankfull width<br />

using a tape measure I estimated d 50 and dmax grain<br />

sizes, and boulder percentage for <strong>the</strong> reach. Prominent<br />

tributaries <strong>of</strong> <strong>the</strong> main stream were also surveyed using<br />

<strong>the</strong> same process.<br />

We also conducted a series <strong>of</strong> cross-valley surveys using<br />

<strong>the</strong> channel as a base and measuring perpendicular<br />

to <strong>the</strong> channel, using a GPS point for location. At<br />

each survey point on <strong>the</strong> hillslope, I recorded bedrock<br />

and boulder percentages and noted bedrock lithology,<br />

<strong>the</strong> presence <strong>of</strong> vegetation, and local slope characteristics.<br />

We completed five cross-valley pr<strong>of</strong>iles for<br />

each main stream, and three for each <strong>of</strong> <strong>the</strong> tributaries.<br />

As we surveyed we measured rock strength with<br />

a Schmidt hammer on outcrops in <strong>the</strong> channel and at<br />

selected outcrops on <strong>the</strong> hillslopes. A Schmidt hammer<br />

measures rock strength by applying a specific<br />

amount <strong>of</strong> force to a rock with a spring loaded piston,<br />

<strong>the</strong>n recording <strong>the</strong> force <strong>of</strong> <strong>the</strong> piston as it rebounds<br />

Figure 2. Schmidt hammer measurements being taken on<br />

Betasso Gulch hillslopes.<br />

107


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

<strong>of</strong>f <strong>the</strong> outcrop.<br />

The scale <strong>of</strong> Middle Boulder Creek and North Boulder<br />

Creek precluded <strong>the</strong> comprehensive field surveys<br />

that we carried out in Gordon Gulch and Betasso<br />

Gulch. The level <strong>of</strong> detail provided by Digital Elevation<br />

Models (DEMs) is sufficient for analysis <strong>of</strong> <strong>the</strong>se<br />

channels. To supplement <strong>the</strong> digital analysis we surveyed<br />

individual, representative reaches distributed<br />

along <strong>the</strong> Boulder Creek channels. Beginning by taking<br />

a GPS point in <strong>the</strong> middle <strong>of</strong> a reach, we surveyed<br />

slope, width, and reach length with <strong>the</strong> rangefinder.<br />

We described each reach and photographed <strong>the</strong> channel<br />

and <strong>the</strong> surrounding hillslopes. We measured river<br />

and valley width with <strong>the</strong> rangefinder, estimated d 50<br />

where cobbles could be seen through <strong>the</strong> water, and<br />

estimated dmax by identifying <strong>the</strong> largest boulder in<br />

<strong>the</strong> reach. We noted <strong>the</strong> presence <strong>of</strong> hillslope features<br />

such as rock slides and falls, tributary entrances, and<br />

cliffs. We also applied <strong>the</strong> same Schmidt hammer process<br />

described above to each section that we recorded,<br />

measuring bedrock if it was present at <strong>the</strong> water level<br />

or stationary boulders where bedrock was not exposed<br />

in <strong>the</strong> channel. In sections with cliffs adjacent<br />

to <strong>the</strong> channel, we searched for evidence <strong>of</strong> sculpting,<br />

potholes, or polish on <strong>the</strong> bedrock, and used <strong>the</strong><br />

rangefinder to measure <strong>the</strong> vertical distance above <strong>the</strong><br />

current channel.<br />

I supplemented my field data with extensive remote<br />

sensing analysis. Using 1m-resolution LIDAR data<br />

for Gordon Gulch and Betasso Gulch, and 10m resolution<br />

DEMs for Middle Boulder Creek, I calculated<br />

additional cross-valley pr<strong>of</strong>iles and basin slopes. I<br />

corroborated my surveyed long pr<strong>of</strong>iles with <strong>the</strong> LI-<br />

DAR and DEMs, and calculated power relationships<br />

with drainage area, downstream distance, and slope.<br />

To ease comparison between basins <strong>of</strong> different magnitudes,<br />

I normalized longitudinal pr<strong>of</strong>iles and mean<br />

basin slope pr<strong>of</strong>iles, setting <strong>the</strong> lowest distance and<br />

elevation values equal to zero, and <strong>the</strong> highest values<br />

equal to one.<br />

RESULTS<br />

The project focuses on relationships between knickpoints,<br />

mean basin slope, channel slope, and rock<br />

strength, so I will focus on data in those areas.<br />

Distance from <strong>the</strong> headwaters is correlated with drain-<br />

Figure 3. A comparison <strong>of</strong> four basins, showing remarkable<br />

similarity in catchment shape despite a large disparity<br />

in magnitude.<br />

Figure 4. Normalized longitudinal pr<strong>of</strong>iles for <strong>the</strong> channels<br />

<strong>of</strong> Betasso Gulch, Gordon Gulch, and Middle Boulder<br />

Creek. The pr<strong>of</strong>iles are made by setting <strong>the</strong> minimum elevation<br />

and distance equal to 1 and <strong>the</strong> maximum elevation<br />

and distance equal to 0, <strong>the</strong>n adjusting each intermediate<br />

point by dividing its value by <strong>the</strong> maximum value.<br />

108


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 5. Plots <strong>of</strong> channel slope and mean basin slope orthogonal<br />

to <strong>the</strong> channel. The x-axis represents <strong>the</strong> distance<br />

in kilometers from <strong>the</strong> headwaters. Mean basin slopes are<br />

calculated by averaging <strong>the</strong> slopes <strong>of</strong> hillslope transects.<br />

The black line on <strong>the</strong> channel slopes is a moving average.<br />

109<br />

age area in each <strong>of</strong> <strong>the</strong> catchments I worked in (Fig.<br />

3).<br />

I found knickpoints in each channel that I surveyed. A<br />

graph <strong>of</strong> normalized longitudinal pr<strong>of</strong>iles shows <strong>the</strong>se<br />

knickpoints on a standard scale (Fig. 4).<br />

The largest channel, Middle Boulder Creek, contains<br />

a prominent knickpoint located on a reach ~30.5-<br />

33 kilometers from <strong>the</strong> headwaters. This knickpoint<br />

has an average slope <strong>of</strong> 0.074, higher than <strong>the</strong> 0.042<br />

channel average. Basin slopes are steepest at <strong>the</strong><br />

knickpoint and just downstream (Fig. 5). There is a<br />

second minor knickpoint near <strong>the</strong> top <strong>of</strong> <strong>the</strong> reach we<br />

surveyed, though slopes in that reach are only locally<br />

as high as 0.04.<br />

In Gordon Gulch, two knickpoints—located ~2.5 and<br />

2.8 kilometers from <strong>the</strong> headwaters and separated by<br />

300 meters <strong>of</strong> relatively low channel slope— are distinguished<br />

by an average slope <strong>of</strong> 0.153. These slopes<br />

are considerably higher than <strong>the</strong> mean reach slope <strong>of</strong><br />

0.108, and more than twice <strong>the</strong> average slope <strong>of</strong> <strong>the</strong><br />

non-knickpoint reaches (0.067) (Fig. 5). Basin slope<br />

analysis shows that <strong>the</strong> average slope on <strong>the</strong> Gordon<br />

Gulch hillsides is higher normal to <strong>the</strong> knickpoints<br />

than o<strong>the</strong>r locations on <strong>the</strong> channel. Schmidt hammer<br />

rock strength values are highest on <strong>the</strong> hillslopes and<br />

in <strong>the</strong> channel at <strong>the</strong> knickpoints: mean values at <strong>the</strong><br />

knickpoints are 45-55, as opposed to mean values <strong>of</strong><br />

30-40 elsewhere in <strong>the</strong> basin.<br />

Betasso Gulch contains three small knickpoints (Fig.<br />

5). Each knickpoint in <strong>the</strong> channel is a short, steep<br />

outcrop covered by little or no sediment. Ranging<br />

in height from 1.5-3 meters, <strong>the</strong>se bedrock steps<br />

have generally higher Schmidt values than bedrock<br />

elsewhere in <strong>the</strong> channel. Hillslope rock strength is<br />

highest between <strong>the</strong> lowest and middle knickpoint. At<br />

<strong>the</strong> two lower knickpoints, Schmidt values range from<br />

40-50, higher than <strong>the</strong> values <strong>of</strong> 25-25 elsewhere. At<br />

<strong>the</strong> upper knickpoint, <strong>the</strong> values are between 25-35,<br />

in contrast to values <strong>of</strong> 0-15 that pervade in <strong>the</strong> upper<br />

zone <strong>of</strong> Betasso Gulch. This uppermost knickpoint is<br />

near <strong>the</strong> low margin <strong>of</strong> a saprolite and colluvial zone:<br />

<strong>the</strong>se disintegrated materials provide a thick cover for<br />

solid bedrock, which rarely crops out at <strong>the</strong> surface in<br />

<strong>the</strong> upper half <strong>of</strong> <strong>the</strong> basin. In this upper section, both


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

channel slopes and hillslopes are shallowest.<br />

DISCUSSION<br />

Though knickpoints take different forms in <strong>the</strong> three<br />

study areas, <strong>the</strong>y mark <strong>the</strong> boundary between steadystate<br />

and adjusting landscapes in all catchments.<br />

Above <strong>the</strong> knickpoint, basin slopes adjust steadily to<br />

accommodate slow downcutting, and are shallower<br />

and smoo<strong>the</strong>r than <strong>the</strong> hillslopes below. The steeper<br />

channel slopes within a knickpoint focus higher<br />

stream power and greater incision rates on that reach<br />

in <strong>the</strong> channel. In <strong>the</strong> catchments in this study, <strong>the</strong> influence<br />

<strong>of</strong> a knickpoint on a longitudinal pr<strong>of</strong>ile scales<br />

with <strong>the</strong> size <strong>of</strong> <strong>the</strong> basin in which it is located (Fig.<br />

4). Middle Boulder Creek has <strong>the</strong> most prominent<br />

knickpoint: <strong>the</strong> zone <strong>of</strong> higher slopes is more than two<br />

kilometers long, and is almost twice as steep as <strong>the</strong><br />

channel average. The knickpoint significantly disrupts<br />

<strong>the</strong> concave-up shape expected for a steady-state<br />

channel. Gordon Gulch, with intermediate drainage<br />

size, includes two closely spaced knickpoints with a<br />

slope disparity comparable to that <strong>of</strong> Middle Boulder<br />

Creek. But <strong>the</strong>se knickpoints are less dominant<br />

features: <strong>the</strong>y appear as large lumps within a generally<br />

smooth pr<strong>of</strong>ile. The knickpoints in Betasso Gulch<br />

are less remarkable: <strong>the</strong>y appear as small steps in <strong>the</strong><br />

longitudinal pr<strong>of</strong>ile but do not affect <strong>the</strong> concavity <strong>of</strong><br />

<strong>the</strong> channel. The bedrock steps that mark knickpoint<br />

location are dramatic in <strong>the</strong> field but are less convincing<br />

when plotted (Fig. 4).<br />

Rapid channel lowering in <strong>the</strong> knickpoint increases<br />

adjacent basin slope: as <strong>the</strong> knickpoint moves up <strong>the</strong><br />

channel it leaves higher basin slopes in its wake. The<br />

rate at which hillslopes adjust to <strong>the</strong> new boundary<br />

conditions depends on <strong>the</strong> mobility <strong>of</strong> constituent<br />

material and <strong>the</strong> capacity <strong>of</strong> <strong>the</strong> channel to move <strong>the</strong><br />

material downstream. Stronger rock resists wea<strong>the</strong>ring<br />

and preserves steeper hillslopes, whereas weak<br />

rock and colluvium are susceptible to wea<strong>the</strong>ring and<br />

mass movements. A stream with high competence can<br />

transport wea<strong>the</strong>red debris and allow fur<strong>the</strong>r wea<strong>the</strong>ring<br />

to occur, but for streams with low stream power<br />

hillslopes only slowly return to a steady state.<br />

The most compelling relationship <strong>of</strong> hillslope to<br />

channel slope is in Middle Boulder Creek (Fig. 5).<br />

110<br />

High rock strength has produced a lag in slope response<br />

to <strong>the</strong> migration <strong>of</strong> <strong>the</strong> knickpoint. Steep basin<br />

slopes persist below <strong>the</strong> main knickpoint, despite high<br />

stream power during snowmelt and summer thunderstorms.<br />

The hillslopes above <strong>the</strong> knickpoint are<br />

smoo<strong>the</strong>r and shallower: <strong>the</strong>y are in relative steady<br />

state, having adjusted to <strong>the</strong> passage <strong>of</strong> a previous,<br />

smaller knickpoint through <strong>the</strong> system.<br />

Hillslopes in Gordon Gulch have responded similarly<br />

to <strong>the</strong> presence <strong>of</strong> knickpoints. A trend <strong>of</strong> steepening<br />

hillslopes with distance from <strong>the</strong> headwaters is broken<br />

only by <strong>the</strong> shallow section between knickpoints<br />

(Fig. 5). These low slopes can be explained by weak<br />

rock—reflected in low Schmidt hammer measurements—that<br />

underlies <strong>the</strong> area between knickpoints.<br />

Additionally, <strong>the</strong> proximity <strong>of</strong> <strong>the</strong> lower knickpoint—<br />

with its associated higher stream power—may help<br />

to efficiently transport material and accelerate <strong>the</strong><br />

hillslope adjustment to <strong>the</strong> new baselevel. Steep<br />

hillslopes below <strong>the</strong> lower knickpoint suggest that <strong>the</strong><br />

landscape <strong>the</strong>re continues to adjust.<br />

Betasso Gulch displays <strong>the</strong> most dramatic basin slope<br />

increase with distance down <strong>the</strong> channel (Fig. 5).<br />

The shallow sloped colluvial and saprolite hillslopes<br />

above <strong>the</strong> knickpoints stand in stark contrast to <strong>the</strong><br />

steep, outcrop-dominated hillslopes that flank <strong>the</strong><br />

knickpoints and below. Several factors can account<br />

for <strong>the</strong> steep lower slopes. The relative strength <strong>of</strong> <strong>the</strong><br />

rock below <strong>the</strong> knickpoints may have prevented <strong>the</strong><br />

lower hillslopes from readjusting. Even if material<br />

was available to move, <strong>the</strong> tiny channel in Betasso<br />

Gulch has a low capacity for transport, and hillslope<br />

evolution would be slowed by that limiting factor.<br />

These characteristics <strong>of</strong> Betasso Gulch have prevented<br />

its basin from adjusting to <strong>the</strong> passage <strong>of</strong> several<br />

knickpoints.<br />

CONCLUSION<br />

Catchments <strong>of</strong> dramatically different sizes can be<br />

compared effectively: general knickpoint mechanics<br />

are similar in each basin we examined. Knickpoints<br />

mark <strong>the</strong> boundary between hillslopes in steady<br />

state and hillslopes that are struggling to adjust to<br />

new boundary conditions. The disruption <strong>of</strong> previous<br />

steady state is driven by increased incision rates<br />

associated with knickpoints. Hillslopes at and below


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

knickpoints become steeper with a rapid baselevel<br />

fall. Rock strength, stream power, and continued<br />

disruption limit <strong>the</strong> ability <strong>of</strong> a hillslope to approach<br />

a new steady state after a knickpoint passes. In each<br />

study area, we found steep, rough hillslopes downstream<br />

from knickpoints, as opposed to comparatively<br />

smoo<strong>the</strong>r and flatter hillslopes above. We found high<br />

rock strength within and below knickpoints, perhaps<br />

<strong>the</strong> result <strong>of</strong> recent exposure with enhanced denudation<br />

following <strong>the</strong> knickpoint-driven baselevel lowering.<br />

The parameters involved in channel and hillslope<br />

evolution—channel slope, rock strength, hillslope<br />

shape, sediment transport, and wea<strong>the</strong>ring—all are<br />

inherently linked in <strong>the</strong>se systems to <strong>the</strong> presence <strong>of</strong><br />

knickpoints. The catchments in my study area have<br />

not fully adjusted to <strong>the</strong> passage <strong>of</strong> knickpoints. This<br />

lack <strong>of</strong> response suggests a slowly evolving landscape<br />

dominated by strong rock, subdued wea<strong>the</strong>ring, and<br />

low stream power.<br />

ACKNOWLEDGMENTS<br />

I would like to thank my advisor Dr. David P. Dethier<br />

for his guidance, advice, and helpful questions. Thank<br />

you also to Dr. William Ouimet for his assistance<br />

in <strong>the</strong> field and as an advisor through <strong>the</strong> process. I<br />

would like to thank Keith Kantack for his instrumental<br />

work assisting me in <strong>the</strong> field. I would also like to<br />

thank <strong>the</strong> Williams College Geosciences Department,<br />

<strong>the</strong> National Science Foundation, and <strong>the</strong> KECK Geology<br />

Consortium.<br />

REFERENCES<br />

Anderson, R. S. (2008), The Little Book <strong>of</strong> Geomorphology:<br />

Exercising <strong>the</strong> Principle <strong>of</strong> Conservation.<br />

Anderson, R. S. and Anderson, S. P. (2010), Geomorphology:<br />

The Mechanics and Chemistry <strong>of</strong> Landscapes<br />

(Cambridge University Press) textbook,<br />

640 pp., published June 2010.<br />

Birkeland, P.W., Shroba, R.R., Burns, S.F., Price, A.B.<br />

and Tonkin, P.J. (2003), Integrating soils and<br />

geomorphology in mountains - an example from<br />

<strong>the</strong> Front Range <strong>of</strong> Colorado, Geomorphology<br />

55, p. 329-344.<br />

111<br />

Crosby BT, Whipple KX. 2006. Knickpoint initiation<br />

and distribution within fluvial networks: 236 waterfalls<br />

in <strong>the</strong> Waipaoa River, North Island, New<br />

Zealand. Geomorphology 82: 16–38.<br />

Kellogg, K.S., Shroba, R.R., Bryant, Bruce, and<br />

Premo, W.R. (2008), Geologic map <strong>of</strong> <strong>the</strong> Denver<br />

West 30’ x 60’ quadrangle, north-central<br />

Colorado: U.S. Geological Survey Scientific<br />

Investigations Map 3000, scale 1:100,000, 48-p.<br />

pamphlet.<br />

PRISM Climate Group, Oregon State University,<br />

http://prism.oregonstate.edu, created 2011.<br />

Whipple, K. X., and G. E. Tucker (1999), Dynamics<br />

<strong>of</strong> <strong>the</strong> stream power river incision model: Implications<br />

for height limits <strong>of</strong> mountain ranges,<br />

landscape response time scales, and research<br />

needs, J. Geophys. Res., 104, 17,661–17,674.<br />

Wobus, C. W., G. E. Tucker, and R. S. Anderson<br />

(2010), Does climate change create distinctive<br />

patterns <strong>of</strong> landscape incision?, J. Geophys. Res.,<br />

115, F04008, doi:10.1029/2009JF001562.<br />

Zachos, J., Pagani, M., Sloan, L., Thomas, E., and<br />

Billups, K. (2001), Trends, rhythms, and aberrations<br />

in global climate 65 Ma to present. Science,<br />

292, p. 686-693.<br />

Zaprowski, B. J., et al. (2005), Climatic influences on<br />

pr<strong>of</strong>ile concavity and river incision, J. Geophys.<br />

Res., 110, F03004, doi:10.1029/2004JF000138.<br />

Zhang, P., P. Molnar, and W. R. Downs (2001), Increased<br />

sedimentation rates and grain sizes 2 – 4<br />

Myr ago due to <strong>the</strong> influence <strong>of</strong> climate change<br />

on erosion rates, Nature, 410, 891–897.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

RECONSTRUCTING THE PINEDALE GLACIATION, GREEN<br />

LAKES VALLEY, COLORADO<br />

KEITH M. KANTACK, Williams College<br />

Research Advisor: David P. Dethier<br />

INTRODUCTION<br />

During <strong>the</strong> Pinedale Glaciation (~32-15 kya) alpine<br />

glaciers covered much Colorado’s Front Range<br />

(Madole 1998). While <strong>the</strong>se glaciers are small and<br />

scarce today, <strong>the</strong>ir work still dominates <strong>the</strong> landscape.<br />

As powerful erosive agents, glaciers remove massive<br />

volumes <strong>of</strong> material from <strong>the</strong>ir beds and leave<br />

equally impressive volumes <strong>of</strong> sediment in <strong>the</strong> form<br />

<strong>of</strong> moraines and outwash plains. By using field and<br />

modeling techniques based on glacial evidence, we<br />

can determine glacier size and ice flow direction, as<br />

well as how much sediment <strong>the</strong>y deposited, which<br />

ultimately helps to constrain how quickly glaciers cut<br />

<strong>the</strong>ir cirques and valleys.<br />

While <strong>the</strong> scoured bedrock surface <strong>of</strong> <strong>the</strong> deglaciated<br />

alpine zone is not typical <strong>of</strong> <strong>the</strong> critical zone, glaciers<br />

are sediment factories. Much <strong>of</strong> <strong>the</strong> sediment that<br />

ends up covering <strong>the</strong> lower reaches <strong>of</strong> alpine and<br />

subalpine basins and flanking channels downstream is<br />

generated by glacial erosion <strong>of</strong> bedrock. In this way,<br />

glaciers are not just icy blocks high in <strong>the</strong> mountains,<br />

but significant players in <strong>the</strong> all important critical<br />

zone.<br />

In this paper, I use field and LIDAR evidence to model<br />

<strong>the</strong> extent <strong>of</strong> <strong>the</strong> Pinedale glaciation in <strong>the</strong> GLV, as<br />

well as its contributions to <strong>the</strong> critical zone.<br />

SETTING<br />

The Green Lakes Valley (GLV) <strong>of</strong> <strong>the</strong> Colorado Front<br />

Range is about 13 kilometers northwest <strong>of</strong> <strong>the</strong> town <strong>of</strong><br />

Nederland in <strong>the</strong> headwater area <strong>of</strong> <strong>the</strong> North Branch<br />

<strong>of</strong> Boulder Creek (Fig. 1). The valley floor runs<br />

extends from 3300 to 3900 meters in elevation and<br />

is walled by 3900 to 4100 meter peaks that form <strong>the</strong><br />

continental divide. The valley, which is part <strong>of</strong> <strong>the</strong><br />

City <strong>of</strong> Boulder watershed, is protected and access is<br />

limited to researchers working on specific topics.<br />

Figure 1. Colorado, Boulder County, and a panoramic<br />

view <strong>of</strong> <strong>the</strong> GLV, looking west toward <strong>the</strong> continental divide<br />

over <strong>the</strong> step between Green Lake 3 (left) and 4.<br />

During <strong>the</strong> peak <strong>of</strong> <strong>the</strong> Pinedale glaciation, glaciers<br />

flowed out <strong>of</strong> small cirques in <strong>the</strong> Front Range,<br />

through valleys, and joined forces, reaching lengths<br />

<strong>of</strong> 15 kilometers and elevations as low as 2500 meters<br />

(Madole et al. 1998). The GLV bears many marks<br />

<strong>of</strong> glaciation. In favorable sites, <strong>the</strong> bedrock holds<br />

polish and striations, and moraines are preserved in<br />

downvalley locations. On <strong>the</strong> valley floor, bedrock is<br />

ubiquitously smoo<strong>the</strong>d, and bears evidence <strong>of</strong> plucking.<br />

In places, <strong>the</strong> valley walls are oversteepened,<br />

reflecting undercutting by debris sheared along by <strong>the</strong><br />

moving ice. Post-glacial talus fields cover <strong>the</strong> base <strong>of</strong><br />

<strong>the</strong>se cliffs and cover large tracts <strong>of</strong> <strong>the</strong> lower walls.<br />

The valley is U-shaped, with six lakes strung along<br />

118


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

<strong>the</strong> floor. The lakes fill a series <strong>of</strong> topographic steps,<br />

<strong>the</strong> most notable dropping 125 meters from Green<br />

Lake 4 to Green Lake 3.<br />

Just south on <strong>the</strong> continental divide is <strong>the</strong> Arapaho<br />

Valley, which I did not map. Like <strong>the</strong> GLV, <strong>the</strong> floor<br />

<strong>of</strong> <strong>the</strong> Arapaho Valley is dotted with six lakes. But<br />

<strong>the</strong> valley is roughly double <strong>the</strong> area <strong>of</strong> <strong>the</strong> GLV, and<br />

ra<strong>the</strong>r than <strong>the</strong> many-stepped pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> GLV, <strong>the</strong><br />

Arapahoe Valley pr<strong>of</strong>ile has one large (500 meter)<br />

step 1.5 kilometers from <strong>the</strong> divide.<br />

Fur<strong>the</strong>r south are <strong>the</strong> Rainbow and Horseshoe cirques.<br />

These are small features, with areas < 0.5km 2 . However,<br />

<strong>the</strong>re are well-defined moraine complexes below<br />

both cirques. (Gable and Madole, 1976)<br />

The bedrock in underlying <strong>the</strong> study area is comprised<br />

<strong>of</strong> 4 main rock types: (1) cordierite and magnetitebearing<br />

sillimanite-biotite gneiss commonly referred<br />

to as Metasediments; (2) <strong>the</strong> Boulder Creek Granodiorite;<br />

(3) Silver Plume Quartz Monzonite; and (4)<br />

Monzonite .<br />

Climate in this area is cool and continental with<br />

strong seasonal variation and relatively low precipitation.<br />

Presently, mean <strong>annual</strong> temperature is -3.5ºC<br />

and mean <strong>annual</strong> precipitation is 763 mm at <strong>the</strong> Niwot<br />

Ridge D1 monitoring station, which sits atop <strong>the</strong><br />

south-facing slope <strong>of</strong> <strong>the</strong> GLV. Annual snow accumulation<br />

reaches 20 meters at <strong>the</strong> head <strong>of</strong> <strong>the</strong> valley<br />

(Caine 1995).<br />

METHODS<br />

FIELD METHODS<br />

In order to reconstruct <strong>the</strong> extent <strong>of</strong> Pinedale ice in<br />

<strong>the</strong> GLV, I examined bedrock surfaces on <strong>the</strong> floor<br />

and walls <strong>of</strong> <strong>the</strong> valley for evidence <strong>of</strong> recent glacial<br />

smoothing: primarily polish and striations. I traversed<br />

<strong>the</strong> entire valley with a Garmin Rino120 GPS and<br />

mapped evidence <strong>of</strong> glaciation and local flow directions.<br />

Trimline zones were identified on valley walls<br />

as <strong>the</strong> point where polish ceased, and rough, unglaciated<br />

outcrop began. On <strong>the</strong> south-facing slope <strong>of</strong> <strong>the</strong><br />

valley, I also mapped polished boulders.<br />

MODELING AND LABORATORY METHODS<br />

Reconstruction <strong>of</strong> ice is an important aspect <strong>of</strong> this<br />

119<br />

study, as modern glaciers in <strong>the</strong> Front Range are mere<br />

shadows <strong>of</strong> <strong>the</strong>ir former extent. Modeling an ice<br />

surface is essentially determining ice thickness at any<br />

point on <strong>the</strong> glacier and begins with concepts <strong>of</strong> ice<br />

deformation and resulting motion. I used a modeling<br />

program developed by Hulton and Benn (2010) that<br />

incorporates cross-valley pr<strong>of</strong>ile or “shape factor”,<br />

bed pr<strong>of</strong>ile, yield stress, and field observed trimline<br />

elevations.<br />

All GIS studies were done using ArcMap 10. Using<br />

<strong>the</strong> trimline points collected in <strong>the</strong> field, I made a<br />

complete trimline map <strong>of</strong> <strong>the</strong> glacier in <strong>the</strong> GLV. The<br />

Arapaho Valley glacier used Madole’s upper Platte<br />

River glaciation modeling (citation). The greatest<br />

extent <strong>of</strong> <strong>the</strong> glacier was modeled using an equilibrium<br />

line altitude (ELA- <strong>the</strong> altitude at which ablation<br />

equals accumulation) <strong>of</strong> 3350 meters (Ward et al,<br />

2009) As I moved ELA higher, I maintained an accumulation<br />

area ratio (AAR) <strong>of</strong> 0.65, which is to say <strong>the</strong><br />

area above <strong>the</strong> ELA comprised 65% <strong>of</strong> <strong>the</strong> total ice<br />

surface. Below <strong>the</strong> ELA, I modeled <strong>the</strong> glacier with a<br />

high centerline surface topography, where <strong>the</strong> surface<br />

<strong>of</strong> <strong>the</strong> glacier bowed upwards as much as 20 meters.<br />

At <strong>the</strong> ELA, <strong>the</strong> surface was flat. Above <strong>the</strong> ELA, I<br />

modeled <strong>the</strong> glacier in <strong>the</strong> accumulation zone using<br />

low centerline topography where <strong>the</strong> surface bowed<br />

down as much as 20 meters.<br />

To map glacial deposits in <strong>the</strong> Pinedale ablation zone,<br />

I used LIDAR data, which shows moraines in high<br />

detail. Light Detection and Ranging (LIDAR) data<br />

produces a digital elevation model (DEM) with a<br />

unique elevation value for every pixel. The LIDAR<br />

layer I used was flown with snow-<strong>of</strong>f in August, 2010<br />

by <strong>the</strong> National Center for Airborne Laser Mapping,<br />

and is comprised <strong>of</strong> 1-meter pixels. To map <strong>the</strong><br />

moraines, I used a hillshade layer, which accentuates<br />

terrain relief.<br />

To reconstruct <strong>the</strong> Pinedale glacier surface, I used ArcMap’s<br />

Kriging tool, which produced a layer based on<br />

<strong>the</strong> elevation <strong>of</strong> <strong>the</strong> trimline and interpolated surface<br />

points. With <strong>the</strong> layer produced by Kriging, I could<br />

reconstruct elevation contours for <strong>the</strong> Pinedale ice, as<br />

well as calculate <strong>the</strong> volume <strong>of</strong> ice held in <strong>the</strong> valley.<br />

The first step for calculating moraine volume involved<br />

mapping <strong>the</strong> moraine belts using DEMs


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

derived from <strong>the</strong> LIDAR data. Within <strong>the</strong> belts, I<br />

created a new elevation layer (again by Kriging) with<br />

a 1 m 2 pixel based on <strong>the</strong> lowest points around <strong>the</strong><br />

moraines, including kettles within <strong>the</strong> moraine belt.<br />

The points were selected so as to produce a layer that<br />

would approximate <strong>the</strong> surface elevation prior to <strong>the</strong><br />

late-glacial maximum <strong>of</strong> <strong>the</strong> Pinedale glaciation. This<br />

layer was subtracted from <strong>the</strong> LIDAR values, which<br />

(because it is a 1 m 2 pixel) yields moraine volume.<br />

Because making an elevation layer for an entire<br />

moraine belt introduces large local error, each moraine<br />

belt was divided into 4 to 7 sections. In some<br />

places, this interpolated low layer was higher than <strong>the</strong><br />

LIDAR, which produced negative volume values for<br />

those pixels. These pixels were omitted from calculations,<br />

but <strong>the</strong>ir presence indicates that this volume<br />

is a lower limit. An upper limit was established by<br />

lowering <strong>the</strong> interpolated layer 4 meters, to ensure<br />

it includes <strong>the</strong> entire moraine belt. Compared to <strong>the</strong><br />

lower limit, this equates to adding 4 m 3 to each pixel<br />

value.<br />

Erosion rate E was calculated using <strong>the</strong> following<br />

equation.<br />

Where MV is calculated moraine volume, A is area <strong>of</strong><br />

<strong>the</strong> cirque or catchment in which <strong>the</strong> glacial sediment<br />

originated, and t is time. In this case, time was 6000<br />

years, reflecting <strong>the</strong> duration <strong>of</strong> <strong>the</strong> Pinedale glacial<br />

maximum, which lasted from ~21 to 15 kya (Ward et<br />

al, 2009).<br />

RESULTS<br />

Pinedale ice reached its maximum extent at 21 kya,<br />

when <strong>the</strong> ELA stood at 3350 m in <strong>the</strong> Green Lakes<br />

area <strong>of</strong> <strong>the</strong> Front Range (Ward, 2009). At this maximum,<br />

<strong>the</strong> glacier flowing out <strong>of</strong> <strong>the</strong> Green Lakes and<br />

Arapaho Valleys covered an area <strong>of</strong> 22.3 km 2 with a<br />

volume 1.86 km 3 . By 16 kya, <strong>the</strong> ELA had risen to<br />

3650 m (Ward, 2009). With this retreat, <strong>the</strong> Green<br />

Lakes and Arapaho glaciers separated. Because I<br />

did not map <strong>the</strong> Arapaho valley in <strong>the</strong> field, I did not<br />

model its glacier at this ELA, where a more intimate<br />

knowledge <strong>of</strong> <strong>the</strong> area is required. The GLV model,<br />

however, shows a reduction in glaciated area to 2.43<br />

km 2 and volume to 0.23 km 3 . (Table 1)<br />

120<br />

Figure 2. (A) Map showing extent and thickness <strong>of</strong> Pinedale<br />

ice in Green Lakes and Arapaho Valley. Blue is modeled<br />

with an ELA <strong>of</strong> 3350 m, or <strong>the</strong> Pinedale glacial maximum.<br />

Orange is modeled with an ELA <strong>of</strong> 3650 m. Darker<br />

colors indicate greater ice thickness. 10m contours represent<br />

surface elevation <strong>of</strong> <strong>the</strong> ELA 3350 m glacier. ELAs<br />

are shown in green. Black arrows are ice flow direction<br />

indicated by striation measurements. (B) Pr<strong>of</strong>ile <strong>of</strong> glacier<br />

at ELA 3350 m (blue from “A”). Red (Arapaho) and blue<br />

(Green Lakes) are pr<strong>of</strong>iles above <strong>the</strong> confluence.<br />

Ice Thickness (m)<br />

Date ELA (m) Ice Area (km 2 ) Average Maximum Ice Volume (km 3 )<br />

21kya 3350 22.27 83.32 239.58 1.85<br />

16kya 3650 2.43 93.90 207.25 0.23<br />

Table 1. ELA and size statistics for <strong>the</strong> glaciers occupying<br />

<strong>the</strong> Green Lakes and Arapaho Valleys. Note that for 16<br />

kya, only <strong>the</strong> glacier in GLV is considered.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Area (km 2 ) Moraine volume (10 6 m 3 ) Lowering rate, in mm yr -1<br />

Cirque Catchment Moraine Minimum a Maximum b Ca d Min Ca d Max C c Min C c Max<br />

Green Lakes<br />

and Arapahoe 11.19 22.50 5.65 67.06 89.57 0.50 0.66 1.00 1.33<br />

Rainbow Lakes 0.45 0.91 0.71 0.85 3.67 0.15 0.67 0.31 1.37<br />

Horseshoe 0.30 1.45 0.77 2.20 5.28 0.25 0.61 1.23 2.98<br />

a<br />

Volume calculated using subtraction <strong>of</strong> a modeled elevation layer from lidar.<br />

b<br />

Volume with an additional 4 meters <strong>of</strong> sediment lying beneath <strong>the</strong> entire moraine<br />

complex.<br />

c<br />

Lowering rate if morainal material came only from within <strong>the</strong> cirque<br />

d<br />

Lowering rate if morainal material includes sediment from <strong>the</strong> catchment above <strong>the</strong> moraines.<br />

Table 2. Calculated moraine volume and lowering rate<br />

Figure 3. Green Lakes and Arapahoe Valleys and Rainbow and Horseshoe Cirques with associated moraine complexes.<br />

Inset shows example <strong>of</strong> hillshade view <strong>of</strong> moraines. Darker colors in <strong>the</strong> moraines reflect thicker morainal deposits. The<br />

deepest parts <strong>of</strong> <strong>the</strong> valleys and cirques are indicated by darker shades.<br />

121


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

I calculated volumes for <strong>the</strong> moraine belts left by <strong>the</strong><br />

Pinedale glaciers that flowed from <strong>the</strong> Arapaho and<br />

GLVs, along with <strong>the</strong> Rainbow Lakes and Horseshoe<br />

Cirques.<br />

The Arapaho/Green Lakes glacier deposited between<br />

67.06x10 6 and 89.57x10 6 m 3 <strong>of</strong> debris. The Horseshoe<br />

glacier deposited between 2.20 and 5.28x10 6<br />

m3. The Rainbow Lakes moraines occupied between<br />

8.5x10 5 and 3.67x10 6 m 3 . (Table 2)<br />

Erosion or lowering rates during <strong>the</strong> late Pinedale<br />

maximum were calculated for each glacier based on<br />

moraine volumes. These values are likely minima,<br />

since <strong>the</strong> glaciers also released sediment as glacial<br />

outwash, which was mainly transported downstream.<br />

I calculated rates based on <strong>the</strong> entire catchment above<br />

each moraine belt, as well as by assuming that morainal<br />

debris originated from just <strong>the</strong> cirque that held<br />

<strong>the</strong> glacier. Results suggest that bed lowering was<br />

0.25 to 2.98 mm per year for Horseshoe, 0.5 to 1.33<br />

mm per year for Green Lakes and Arapaho, and 0.15<br />

to 1.37 mm per year for Rainbow (Table 2).<br />

DISCUSSION<br />

The modeled glacier shows that while it covered an<br />

impressive area, <strong>the</strong> GLV/Arapaho ice at maximum<br />

was only 83 meters thick on average, and 240 meters<br />

at its thickest. The glacier’s thinness likely reflects<br />

relatively dry climate. Thinness may also suggest a<br />

high ice velocity, but this is not likely given <strong>the</strong> cool<br />

climate and low precipitation. High velocity could<br />

also explain <strong>the</strong> area <strong>of</strong> ice below <strong>the</strong> ELA, which is<br />

strikingly large considering <strong>the</strong> small accumulation<br />

zone in <strong>the</strong> Arapaho and Green Lakes Valleys. Given<br />

<strong>the</strong> stepped nature <strong>of</strong> <strong>the</strong> bed (slope ranges from flat<br />

to near vertical), flow velocity was likely highly varied.<br />

The erosion rates calculated here are reasonable for<br />

a small glacier in a relatively dry climate flowing<br />

over hard rocks. In <strong>the</strong>ir glacial-valley pr<strong>of</strong>ile model,<br />

MacGregor and Anderson (2000) use an erosion rate<br />

<strong>of</strong> 1 mm per year. Ward et al. (2009) use erosion<br />

rates <strong>of</strong> 0.1, 1.0, and 10 mm per year in <strong>the</strong>ir modeling<br />

<strong>of</strong> <strong>the</strong> Middle Boulder Creek Valley glaciation. It<br />

is important to realize that <strong>the</strong> rates calculated here<br />

122<br />

provide provisional upper and lower limits based on<br />

field evidence. Lidar data provide a precise DEM<br />

and GIS-based spatial analysis (kriging in this case),<br />

which is a rational method for calculating <strong>the</strong> surface<br />

beneath <strong>the</strong> late Pinedale morainal debris. We have<br />

measured <strong>the</strong> depth <strong>of</strong> sediment and water in several<br />

kettles and believe that 4 m is a reasonable estimate<br />

for an average depth and thus a thickness to add to<br />

<strong>the</strong> morainal belt. Additional field measurements and<br />

shallow geophysics would allow us to better constrain<br />

<strong>the</strong>se values. Additionally we have not estimated <strong>the</strong><br />

volume <strong>of</strong> sediment that was carried away by glaci<strong>of</strong>luvial<br />

and fluvial processes. Finally, we have not estimated<br />

what fraction <strong>of</strong> <strong>the</strong> sediment in <strong>the</strong> moraines<br />

was eroded by glaciers from <strong>the</strong> cirques and which<br />

portion comes from slope processes in <strong>the</strong> rest <strong>of</strong> <strong>the</strong><br />

catchment. However, while <strong>the</strong>se unestimated values<br />

mainly prevent me from measuring <strong>the</strong> total glacial<br />

erosion rate, <strong>the</strong> range <strong>of</strong> general lowering rates I<br />

present for each location are well constrained.<br />

One would expect <strong>the</strong> Rainbow and Horseshoe glaciers<br />

to have nearly identical erosion rates, as <strong>the</strong>y<br />

are similar size cirques located less than 1 kilometer<br />

apart on <strong>the</strong> same slope. However, Horseshoe gives<br />

a greater rate <strong>of</strong> incision. This is likely <strong>the</strong> result <strong>of</strong><br />

bedrock differences and catchment size. While Rainbow<br />

is underlain with primarily metaseds, Horseshoe<br />

is cutting into <strong>the</strong> coarser Boulder Creek Granodiorite.<br />

Additionally, Horseshoe sits in a catchment 60%<br />

larger than Rainbow’s. This suggests that <strong>the</strong> periglacial<br />

and o<strong>the</strong>r processes in <strong>the</strong> catchment contributed<br />

an important fraction <strong>of</strong> <strong>the</strong> sediment in <strong>the</strong> moraine<br />

belt.<br />

When you consider cirque volume, <strong>the</strong>se values indicate<br />

that cutting <strong>the</strong> Green Lakes and Arapaho Valleys<br />

to <strong>the</strong>ir present depth would require 11 glacier periods<br />

<strong>of</strong> <strong>the</strong> same size as <strong>the</strong> 6000 years <strong>of</strong> <strong>the</strong> late Pinedale.<br />

Cutting <strong>the</strong> Rainbow and Horseshoe cirques would<br />

require 6.5 and 3.6 such glaciations, respectively. The<br />

calculated cirque-cutting time values suggest that <strong>the</strong><br />

late Pinedale was a significant event in <strong>the</strong> morphologic<br />

evolution <strong>of</strong> <strong>the</strong>se features. These calculations<br />

assume that <strong>the</strong> erosion rate would be <strong>the</strong> same in<br />

previous glaciations.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

CONCLUSIONS<br />

Using a combination <strong>of</strong> field evidence, numerical<br />

modeling, and GIS data I was able to reconstruct <strong>the</strong><br />

Pinedale glaciation <strong>of</strong> <strong>the</strong> GLV region <strong>of</strong> <strong>the</strong> Colorado<br />

Front Range.<br />

Modeling suggests that <strong>the</strong> glacier reached an areal<br />

extent <strong>of</strong> 22 km 2 , but that <strong>the</strong> ice was generally thin,<br />

with an average thickness <strong>of</strong> 83 meters. This could<br />

indicate high flow rate, but more likely reflects <strong>the</strong><br />

dry climate.<br />

The calculated moraine volumes suggest that <strong>the</strong><br />

Pinedale represents a significant episode in <strong>the</strong> evolution<br />

<strong>of</strong> Front Range morphology. The calculated<br />

erosion rates, which are on <strong>the</strong> order <strong>of</strong> 1 mm per<br />

year, represent a substantial movement <strong>of</strong> sediment<br />

within <strong>the</strong> critical zone and exposure <strong>of</strong> large volumes<br />

<strong>of</strong> fresh surfaces.<br />

These results represent a work in progress. Fur<strong>the</strong>r<br />

work will include more detailed modeling <strong>of</strong> <strong>the</strong><br />

retreat <strong>of</strong> Pinedale ice, studying <strong>the</strong> development <strong>of</strong><br />

steps within <strong>the</strong> GLV, and determining <strong>the</strong> origin <strong>of</strong><br />

<strong>the</strong> polished boulders that sit 50 m above trimline on<br />

<strong>the</strong> south-facing slope <strong>of</strong> <strong>the</strong> valley.<br />

ACKNOWLEDGEMENTS<br />

Thanks to my advisor David Dethier for getting me to<br />

this point and somehow keeping it fun. And thanks to<br />

Evan Dethier and James McCarthy for <strong>the</strong>ir help and<br />

companionship in <strong>the</strong> field and classroom.<br />

REFERENCES<br />

Caine, N. (1995). Snowpack Influences on Geomorphic<br />

Processes in Green Lakes Valley, Colorado<br />

Front Range. The Geographic Journal 161(1):<br />

55-68.<br />

Gable, F. J. and R. F. Madole (1976). Geologic Map<br />

<strong>of</strong> <strong>the</strong> Ward Quadrangle, Boulder County, Colorado,<br />

USGS.<br />

Hulton, N. R. J. and D. I. Benn (2010). An Excel<br />

spreadsheet program for reconstructing <strong>the</strong> sur-<br />

123<br />

face pr<strong>of</strong>ile <strong>of</strong> former mountain glaciers and ice<br />

caps. Computers and Geosciences 36: 605-610.<br />

MacGregor, K.R., Anderson, R.S., Anderson, S.P.,<br />

and Waddington, E.D. Numerical simiulation<br />

<strong>of</strong> glacial-valley longitudinal pr<strong>of</strong>ile evolution.<br />

Geology. 2000.<br />

Madole, R. F., D. VanSistine, and Michael, J.A.<br />

(1998). Glaciation in <strong>the</strong> upper Platte River<br />

drainage basin. Geologic Investigations Series.<br />

USGS.<br />

Ward, D. J., Anderson, Robert S., Guido, Zackry S.,<br />

and Briner, Jason P. (2009). Numerical Modeling<br />

<strong>of</strong> Cosmogenic Deglaciation Records, Front<br />

Range and San Juan Mountains, Colorado. Journal<br />

<strong>of</strong> Geophysical Research 114.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

CHARACTERIZATION OF TRACE METAL CONCENTRATIONS<br />

AND MINING LEGACY IN SOILS, BOULDER COUNTY,<br />

COLORADO<br />

ELLEN M. MALEY, Smith College<br />

Research Advisor: Amy L. Rhodes<br />

INTRODUCTION<br />

The Boulder Creek Critical Zone Observatory (CZO)<br />

is located in an area steeped in mineral exploitation<br />

dating back to <strong>the</strong> 1880s. The Colorado Mineral Belt<br />

is comprised <strong>of</strong> igneous intrusions and ore deposits<br />

from <strong>the</strong> Laramide orogeny. This band strikes nor<strong>the</strong>ast/southwest<br />

through <strong>the</strong> region, stretching 250<br />

miles (Tweto and Sims, 1963). Beginning in <strong>the</strong> late<br />

1800s, this region was mined for gold, tungsten, lead,<br />

silver, and zinc (Fig. 1). Along with prospecting,<br />

mining, and smelting, this region was subject to deforestation<br />

and development with <strong>the</strong> flux <strong>of</strong> settlers<br />

in <strong>the</strong> late 1800s (Veblen and Lorenz, 1991). The<br />

surge <strong>of</strong> European settlers in <strong>the</strong> mining boom made<br />

this region an industrial hub, utilizing processes now<br />

known to have negative impacts on ecosystems.<br />

The Boulder Creek CZO project focuses on <strong>the</strong> balance<br />

between natural wea<strong>the</strong>ring and erosion processes<br />

in <strong>the</strong> critical zone. The purpose <strong>of</strong> this investigation<br />

is to quantify heavy metal concentrations<br />

in regional soils to identify possible contamination<br />

associated with mineral exploitation. In investigation<br />

<strong>of</strong> soil processes in <strong>the</strong> critical zone, environmental<br />

impacts associated with <strong>the</strong> flux <strong>of</strong> human settlers,<br />

such as deforestation, road construction, and mining,<br />

should not be overlooked. In <strong>the</strong> case <strong>of</strong> mining,<br />

Diawara et al. (2006) report surface soil contamination<br />

<strong>of</strong> lead (300 ppm), arsenic (30 ppm), cadmium (5<br />

ppm), and mercury (200 ppb) around major smelting<br />

sites in sou<strong>the</strong>rn Colorado. While certain elements,<br />

such as Pb, have been detected in Boulder County<br />

soils, <strong>the</strong> heavy metal signature associated with this<br />

region has not been investigated in detail (Dethier,<br />

personal communication). This investigation characterizes<br />

<strong>the</strong> regional extent <strong>of</strong> metal contamination<br />

across a range <strong>of</strong> sites in Boulder County.<br />

This investigation employs <strong>the</strong> analysis <strong>of</strong> spatial and<br />

geochemical data to determine whe<strong>the</strong>r surface soils<br />

exhibit trace metal enrichment and, if so, to characterize<br />

<strong>the</strong> nature <strong>of</strong> contamination across southwestern<br />

Boulder County. This study investigates lead, arsenic,<br />

cadmium, chromium, barium, manganese, mercury,<br />

copper, iron, and aluminum. These 10 metals include<br />

major soil constituents (Fe, Al, Mn) and common<br />

environmental contaminants (Pb, Cd, As and Hg).<br />

Trace metal enrichment in <strong>the</strong> A horizon compared to<br />

material from depth (C horizon) is expected to be a<br />

function <strong>of</strong> two variables: natural accumulation and<br />

external accumulation. Pedogenesis releases mineralbound<br />

elements to <strong>the</strong> soil through wea<strong>the</strong>ring. Biochemical<br />

processes draw elements, such as manganese,<br />

from depth and concentrate <strong>the</strong>m in surface soils<br />

via decomposition. Natural accumulation depends on<br />

pr<strong>of</strong>ile wea<strong>the</strong>ring and organic enrichment.<br />

Alternatively, external accumulation reflects anthropogenic<br />

deposition. The main control on deposition<br />

is <strong>the</strong> distance between a location and its potential<br />

contamination source, in this case, mining/smelting<br />

sites. The likelihood <strong>of</strong> contamination surrounding<br />

historic mining and smelting sites is based on <strong>the</strong><br />

observation by Kabata-Pendias (2001) that metal<br />

halos surround locations such as <strong>the</strong>se. With no o<strong>the</strong>r<br />

factors considered, metals associated with mining<br />

should be highest in concentration in soils surrounding<br />

<strong>the</strong> source <strong>of</strong> enrichment, and will decrease along<br />

a gradient with distance from <strong>the</strong> source, mostly in<br />

<strong>the</strong> direction <strong>of</strong> transport.<br />

Controls on accumulation pertain to <strong>the</strong> affinity<br />

each soil has for an element. Kabata-Pendias (2001)<br />

suggests that organic and clay content are <strong>the</strong> main<br />

controls for metal adsorption in soils, as oxide minerals<br />

and organics have partial negative-charged surfaces,<br />

which tend to form weak bonds with positively<br />

charged metals. Soil pH affects metal solubility and<br />

124


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

is a function <strong>of</strong> mineralogy and organic content. Eh/<br />

pH plots for elements demonstrate increased mobility<br />

in acidic, reducing soils. To characterize <strong>the</strong><br />

distribution <strong>of</strong> metals in Boulder County soils in this<br />

investigation, trace metal concentrations were evaluated<br />

in relation to <strong>the</strong> soil properties <strong>of</strong> pH, % organic<br />

content, and % fine-grained material. Possible spatial<br />

relationships between trace metal concentrations and<br />

locations <strong>of</strong> mining and smelting locations were also<br />

investigated.<br />

METHODS<br />

To obtain a regional spectrum <strong>of</strong> soils, samples were<br />

collected along an east-trending elevation gradient<br />

from alpine (3450 m) to lower montane (1850 m)<br />

elevations; samples included three Boulder CZO<br />

sites: Green Lakes Basin, Gordon Gulch, and Betasso<br />

Preserve. Accessible sites along major roads between<br />

CZO sites also were sampled to characterize <strong>the</strong><br />

region in higher detail (Fig. 1). Surface soils, consisting<br />

<strong>of</strong> organic and A horizons, were sampled to ~20<br />

cm depth. Unaltered parent material (C horizon) was<br />

sampled at >1 meter depth to compare surface soils<br />

with unwea<strong>the</strong>red parent material geochemically.<br />

Thirty-three air-dried samples were sieved to


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

and Hydra-C were normalized by soil digest mass.<br />

Results for As, Al, Ba, Cd, Cr, Cu, Fe, Mn, and Pb<br />

in mg/kg soil (ppm), and Hg in µg/kg soil (ppb) are<br />

reported in Table 1. Sample concentrations were used<br />

to explore how metals fractionate, how metal concentrations<br />

vary by horizon, and how concentrations <strong>of</strong><br />

one element relate to o<strong>the</strong>rs within horizons.<br />

Correlation ellipsoid matrices were used to generalize<br />

relationships between elements for each horizon.<br />

Corresponding elemental A and C horizon concentrations<br />

were compared to evaluate whe<strong>the</strong>r processes<br />

are occurring through <strong>the</strong> entire soil pr<strong>of</strong>ile or are<br />

localized within a given horizon. Surface soil concentrations<br />

were analyzed spatially in relation to mining<br />

site locations using ArcGIS. Spatial parameters<br />

<strong>of</strong> distance to nearest mine site and number <strong>of</strong> mines<br />

within 1.5 kilometer radius for each sample site were<br />

created for statistical analysis, which is not addressed<br />

in this submission.<br />

Examination <strong>of</strong> correlations between different trace<br />

metals and between trace metals and soil properties<br />

may provide insight regarding metal origin. For<br />

example, a correlation between two metals within <strong>the</strong><br />

C horizon for all samples could suggest that <strong>the</strong> metal<br />

origin is similar and may be associated with parent<br />

material breakdown. Likewise, a strong correlation<br />

between two metals in surface soils and not at depth<br />

could suggest that <strong>the</strong> metals are derived from <strong>the</strong><br />

same source, such as atmospheric deposition.<br />

Fractionation <strong>of</strong> Elements<br />

The finest material in soils is comprised <strong>of</strong> silt- and<br />

clay-sized mineral and organic particles, which bear a<br />

partial-negative surface charge and have high relative<br />

surface area (Kabata-Pendias, 2001). The charge and<br />

area draw metal cations to clay surfaces, which results<br />

in higher metal content in fine-grained fractions. In<br />

keeping with literature predictions about <strong>the</strong> behavior<br />

<strong>of</strong> trace metals in soils (Kabata-Pendias, 2001),<br />

fraction analysis <strong>of</strong> three samples showed higher trace<br />

metal concentration in fine-grained fractions (Fig. 2).<br />

Soil chemistry was comparable between 125µm.<br />

126<br />

For example, Pb and As show a negative correlation<br />

between grain-size and concentration (Fig. 2). Concentrations<br />

are higher in finer fractions than in coarse<br />

material for all elements. Analyses suggest that aluminum<br />

is lower in concentration at Magnolia Road, a<br />

21 kyr moraine soil, compared to Gordon Gulch and<br />

Betasso sites. As this soil does not vary significantly<br />

in age with o<strong>the</strong>r sites (see Wyshnytsky, this volume),<br />

this difference is likely due to variations in host rock<br />

mineralogy between sites.<br />

Figure 2. Example plots <strong>of</strong> Al, Pb, and As concentration<br />

distribution by size fraction (in µm) from three sample<br />

locations, compared with bulk concentrations in ppm (mg/<br />

kg soil) for each site.<br />

Elemental Trends by Soil Horizon<br />

Boulder County A horizons are enriched in Mn, Pb,<br />

Ba, Hg, and As. Surface soil enrichment results from<br />

accumulation and/or atmospheric deposition. Anthropogenic<br />

addition to soils from mining activity<br />

has occurred via atmospheric deposition. However,<br />

external dust input and bioaccumulation, <strong>the</strong> tendency<br />

for plants to draw elements from depth and concentrate<br />

<strong>the</strong>m in surface soils, may also be occurring.<br />

Therefore, elements enriched at <strong>the</strong> surface may be<br />

<strong>of</strong> anthropogenic or natural origin. In contrast, C<br />

horizons are enriched in Cu and Cr. Iron and Al do<br />

not vary consistently between horizons. Elements<br />

without surface enrichment are likely derived from<br />

original host rock composition.<br />

The A to C relationship across sites was evaluated by<br />

comparing elemental concentrations between A and


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Pb (ppm)<br />

Hg (ppb)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Pb<br />

Hg<br />

As<br />

Figure 3. Mercury, As, Pb and Mn, Ba as functions <strong>of</strong> % organic content with regression line equations and associated R 2<br />

values displayed. These correlations suggest that elements that are enriched in surface soils are associated with organic<br />

content.<br />

C horizons collected at <strong>the</strong> same sample site. Linear<br />

trends between A and C horizons for Fe, Cu and Cr<br />

suggest that variations by site are associated with<br />

pedogenic breakdown <strong>of</strong> host rock. Enrichment in <strong>the</strong><br />

A horizon and <strong>the</strong> relationships between A-enriched<br />

elements and soil properties, like % organics, warranted<br />

fur<strong>the</strong>r investigation.<br />

Elemental Relationships<br />

y = 7.0426 + 2.2878x R 2 = 0.51071<br />

y = 2.1103 + 2.9885x R 2 = 0.73348<br />

y = 1.1015 + 0.14777x R 2 = 0.61033<br />

0<br />

0<br />

0 5 10 15 20 25<br />

% organic content<br />

In surface soils, enrichment <strong>of</strong> elements can be<br />

derived from natural and anthropogenic processes.<br />

Elevated concentrations <strong>of</strong> Mn and Ba may result<br />

from biological surficial concentration <strong>of</strong> elements<br />

from depth or from atmospheric dust input, while <strong>the</strong><br />

source <strong>of</strong> Pb, As and Hg is anthropogenic (Kabata-<br />

Pendias, 2001). As Mn is extracted from depth over<br />

time by plants, decomposing organic matter may concentrate<br />

Mn in surface soils (Kabata-Pendias, 2001).<br />

There is a positive correlation between Mn and Ba,<br />

which suggests a common source. Mercury, Pb, As,<br />

Mn, and Ba are enriched in surface soils, and each has<br />

a positive correlation with % organic content (Fig. 3).<br />

5<br />

4<br />

3<br />

2<br />

1<br />

As (ppm)<br />

As (ppm)<br />

Mn (ppm)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

500<br />

400<br />

300<br />

200<br />

100<br />

As<br />

Hg<br />

Mn<br />

Ba<br />

y = 41.748 + 11.948x R 2 = 0.43988<br />

y = 26.236 + 1.4091x R 2 = 0.2561<br />

0<br />

0<br />

0 5 10 15 20 25<br />

% organic content<br />

y = 1.1148 + 0.044219x R 2 = 0.55056<br />

y = 2.5292 + 0.88298x R 2 = 0.67269<br />

0<br />

0<br />

0 20 40 60 80100 Pb (ppm)<br />

Figure 4. Mercury (ppb) and As (ppm) across all horizons<br />

and sites with corresponding Pb (ppm) concentrations,<br />

showing <strong>the</strong> relationship between Hg/Pb and As/Pb across<br />

southwestern Boulder County. R2 values (Hg/Pb R2 =<br />

0.67, As/Pb R2 = 0.55) suggest correlation between variables.<br />

127<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Ba (ppm)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Hg (ppb)


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

The most striking trend in trace metals occurs between<br />

Pb, As, and Hg in surface soils (Fig. 4). Strong<br />

correlation between Pb, As, and Hg (Fig. 4) in surface<br />

soils suggests that <strong>the</strong>se metals were deposited via a<br />

common atmospheric anthropogenic source.<br />

In <strong>the</strong> C horizon, Fe and Mn, Fe and Al, and Fe and<br />

Cr are positively correlated, suggesting that <strong>the</strong>se elements<br />

are associated with host rock breakdown. No<br />

relationship between Cu and Cr exists, though both<br />

<strong>of</strong> <strong>the</strong>se elements are enriched at depth. The positive<br />

correlations between Cr and Fe and Cr and Al suggest<br />

that chromium is present in host rock, and it was<br />

incorporated into soils by pedogenesis. Iron and Mn<br />

are positively correlated at depth, but not in surface<br />

soils where Mn is enriched. This difference could be<br />

attributed to Mn-rich dust inputs or bioaccumulation<br />

in organic matter (Kabata-Pendias, 2001).<br />

Mining Signal and Implications<br />

Surface soils are enriched with respect to Mn, Ba, As,<br />

and Pb, but not all A horizon enrichment can be attributed<br />

to anthropogenic activity. Barium and Mn have<br />

a weak association with organic content, and <strong>the</strong>se<br />

elements may be enriched due to bioaccumulation<br />

(Kabata-Pendias, 2001). The correlation between As,<br />

Hg, and Pb in surface soils suggests that <strong>the</strong>se metals<br />

are affiliated with atmospheric deposition. These<br />

metals are associated with ore smelting processes that<br />

may have been used in <strong>the</strong> Boulder County region<br />

(Kabata-Pendias, 2001). These trace metal signatures<br />

are likely evidence <strong>of</strong> <strong>the</strong> mining legacy in <strong>the</strong> region.<br />

CONCLUSIONS<br />

The enrichment <strong>of</strong> surface soils in elements associated<br />

with anthropogenic activity suggests that Boulder<br />

County soils bear a signature associated with historic<br />

mining activity in <strong>the</strong> region. Surface soils in sou<strong>the</strong>rn<br />

Boulder County are enriched in Hg, Mn, Ba, As,<br />

and Pb. Lead, arsenic, and mercury enrichment likely<br />

results from anthropogenic atmospheric inputs from<br />

mining activity, whereas <strong>the</strong> origin <strong>of</strong> manganese and<br />

barium enrichment is likely from bioaccumulation<br />

or dust inputs. Measured concentrations <strong>of</strong> metals<br />

reflect mechanisms <strong>of</strong> deposition and mobilization.<br />

While o<strong>the</strong>r elements may have been associated with<br />

deposition, Pb, As, and Hg are potential contaminants<br />

with a related source. This investigation <strong>of</strong> potential<br />

contamination <strong>of</strong> toxic elements bears significance for<br />

<strong>the</strong> Boulder Creek watershed, as <strong>the</strong> heavy metals in<br />

surface soils are bioavailable (U.S. EPA, 1996) and<br />

could mobilize. Fur<strong>the</strong>r exploration will quantify <strong>the</strong><br />

mass <strong>of</strong> each element by area to characterize contamination<br />

potential <strong>of</strong> <strong>the</strong> watershed.<br />

Within a soil pr<strong>of</strong>ile, trace metals form weak bonds<br />

with fine-grained (1000 ppm<br />

sample excluded from range (n=6); B: O (n=4), A (n=14),<br />

C (n=3). C: compared with mean values in U.S. soils, as<br />

reported by Kabata-Pendias (2001).<br />

128


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

this project. I also thank Dr. Robert Newton, for his<br />

encouragement and wisdom, and Dr. Anna Martini<br />

and Erin Camp at Amherst College for <strong>the</strong>ir assistance<br />

with mercury analysis.<br />

REFERENCES<br />

Causey, J. Douglas. 1998. MAS/MILS Arc/Info Point<br />

Coverage for <strong>the</strong> Western United States (Excluding<br />

Hawaii): United States Geological Survey<br />

Open-File Report 98-512, Spokane, WA.<br />

Diawara, M., J. Litt, D. Unis, N. Alfonso, L. Martinez,<br />

J. Crock, D. Smith, and J. Carsella. 2006. “Arsenic,<br />

Cadmium, Lead, and Mercury in Surface Soils,<br />

Pueblo, Colorado: Implications for Population<br />

Health Risk.” Environmental Geochemistry and<br />

Health 28 (4): 297-315.<br />

Kabata-Pendias, A. and H. Pendias. 2001. Trace Elements<br />

in Soils and Plants. 3rd ed. Boca Raton,<br />

FL: CRC Press: 331 pp.<br />

Kimbrough, David E. and Janice R. Wakakuwa. 1989.<br />

“Acid Digestion for Sediments, Sludges, Soils,<br />

and Solid Wastes. A Proposed Alternative to EPA<br />

SW 846 Method 3050.” Environmental Science &<br />

Technology 23 (7): 898-900.<br />

Robertson, G. Philip, David C. Coleman, Caroline S.<br />

Bledsoe, and Phillip Sollins, eds. 1999. Standard<br />

Soil Methods for Long-Term Ecological Research.<br />

New York: Oxford University Press.<br />

Rowell, D. L. 1994. “Section 3.3: The Determination<br />

<strong>of</strong> Water Content and Loss on Ignition.” In Soil<br />

Science: Methods & Applications, 48. Reading,<br />

Massachusetts: Prentice Hall.<br />

Tweto, O. and P. K. Sims. 1963. “Precambrian Ancestry<br />

<strong>of</strong> <strong>the</strong> Colorado Mineral Belt.” Geological<br />

Society <strong>of</strong> America Bulletin 74 (8): 991-1014.<br />

Veblen, Thomas T. and Diane C. Lorenz. 1991. The<br />

Colorado Front Range: A Century <strong>of</strong> Change.<br />

Salt Lake City, Utah: University <strong>of</strong> Utah Press.<br />

United States Environmental Protection Agency.<br />

129<br />

1996. Method 3050B: Acid Digestion <strong>of</strong> Sediments,<br />

Sludges, and Soils–Revision 2. United<br />

States EPA SW-846: Test Methods for Evaluating<br />

Solid Waste, Physical/Chemical Methods.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

INTRODUCTION<br />

The Critical Zone<br />

The Critical Zone extends from fresh rock to <strong>the</strong> top<br />

<strong>of</strong> <strong>the</strong> vegetation canopy and is where <strong>the</strong> biosphere,<br />

atmosphere, hydrosphere, and rock materials interact<br />

(Anderson et al., 2007). The complex processes<br />

occurring in <strong>the</strong> Critical Zone release raw materials<br />

from minerals and create substrates for terrestrial life,<br />

supporting microbial, plant, and faunal activity. Rates<br />

<strong>of</strong> chemical and mechanical processes vary temporally<br />

and spatially due to gradients in climate, rock<br />

materials, and slope. Soil is <strong>the</strong> highly wea<strong>the</strong>red,<br />

top-most part <strong>of</strong> <strong>the</strong> regolith, but is distinct because<br />

<strong>of</strong> its unique layered habit. These layers (horizons),<br />

record more intense wea<strong>the</strong>ring conditions at <strong>the</strong><br />

surface, and record downward transport <strong>of</strong> chemical<br />

wea<strong>the</strong>ring products and organic additions from<br />

<strong>the</strong> biosphere (Anderson and Anderson, 2010). Field<br />

and laboratory studies <strong>of</strong> soils enrich models <strong>of</strong> rock<br />

wea<strong>the</strong>ring, chemical and mechanical transport, and<br />

regolith formation. Results help to highlight <strong>the</strong> role<br />

<strong>of</strong> soil-forming processes as indicators <strong>of</strong> <strong>the</strong> geomorphic<br />

controls on Critical Zone processes.<br />

The Critical Zone can be thought <strong>of</strong> as a bottom-up<br />

feed-through reactor in which physical and chemical<br />

wea<strong>the</strong>ring processes alter fresh rock material being<br />

supplied by uplift and erosion (Anderson et al., 2007).<br />

Simultaneously, physical erosion and chemical denudation<br />

processes transport mass out <strong>of</strong> <strong>the</strong> system.<br />

Thus, <strong>the</strong> rates <strong>of</strong> wea<strong>the</strong>ring and denudation toge<strong>the</strong>r<br />

determine <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> Critical Zone (Anderson<br />

et al., 2007). The balance <strong>of</strong> transport-limited and<br />

wea<strong>the</strong>ring-limited environments is largely determined<br />

by topographic and climatic factors.<br />

Catenas<br />

ASSESSING EOLIAN CONTRIBUTIONS TO SOILS IN THE<br />

BOULDER CREEK CATCHMENT, COLORADO<br />

JAMES A. MCCARTHY, Williams College<br />

Research Advisor: David P. Dethier<br />

The bottom-up reactor model for regolith formation<br />

130<br />

assumes that parent materials in <strong>the</strong> system derive<br />

only from <strong>the</strong> wea<strong>the</strong>ring <strong>of</strong> underlying bedrock or<br />

sediment. However, soil pr<strong>of</strong>iles on slopes are distinctly<br />

related to <strong>the</strong> soils above and below because<br />

<strong>of</strong> <strong>the</strong> influence <strong>of</strong> slope-controlled transport mechanisms.<br />

The term catena describes a sequence <strong>of</strong> soils<br />

on a slope, emphasizing that <strong>the</strong>ir variation is due to<br />

changes in both slope gradient and position (Birkeland,<br />

1999). In this model, regolith and soils are<br />

thinnest at <strong>the</strong> shoulder and backslope, and gradually<br />

thicken, reaching a maximum at <strong>the</strong> toeslope. There<br />

is also a chemical gradient along slopes, driven partly<br />

by <strong>the</strong> physical transport <strong>of</strong> <strong>the</strong> mobile regolith, and<br />

enhanced by hydrologic factors; clay minerals and<br />

dissolved cations in a soil column may be transported<br />

down slope by throughflow, accumulating at <strong>the</strong> base<br />

<strong>of</strong> <strong>the</strong> slope. Climatic conditions determine both <strong>the</strong><br />

mobility <strong>of</strong> soil materials and chemical constituents,<br />

and thus <strong>the</strong> effects <strong>of</strong> throughflow on pedogenesis<br />

vary spatially. Thicker soils and better-developed<br />

horizons may occur at <strong>the</strong> base <strong>of</strong> slopes due to <strong>the</strong><br />

influx <strong>of</strong> wea<strong>the</strong>rable materials and increased soilmoisture<br />

status via throughflow water.<br />

Downslope transport <strong>of</strong> <strong>the</strong> mobile layer may be episodic,<br />

mediated by climate, and <strong>the</strong> regolith that eventually<br />

arrives at <strong>the</strong> toeslope includes a mixture <strong>of</strong><br />

soil and saprolite. Models <strong>of</strong> sediment flux generally<br />

assume that hillslope processes are constant through<br />

time, but episodic transport suggests more stochastic<br />

conditions (Anderson and Anderson, 2010). Regolith<br />

moves downslope and episodic transport results in <strong>the</strong><br />

burial <strong>of</strong> soils at <strong>the</strong> base <strong>of</strong> <strong>the</strong> slope; current soils<br />

in <strong>the</strong>se positions form from materials derived from<br />

upslope ra<strong>the</strong>r than from bedrock. Thus, morphological<br />

differences along hillslopes may represent changes<br />

in <strong>the</strong> strength <strong>of</strong> pedogenesis and changes in parent<br />

material.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Significance <strong>of</strong> dustfall<br />

Dustfall is not included explicitly in <strong>the</strong> Critical Zone<br />

reactor model; however, <strong>the</strong> enrichment <strong>of</strong> eolian silt<br />

and clay affects <strong>the</strong> “mineralogy, chemistry, nutrient<br />

status, and moisture-holding capacity <strong>of</strong> soils (Muhs<br />

and Benedict, 2006, p. 120),” and thus “may control<br />

<strong>the</strong> rate and direction <strong>of</strong> pedogenesis (Mason and Jacobs,<br />

1998, p. 1135).” For this reason, characterizing<br />

<strong>the</strong> rate <strong>of</strong> eolian inputs is important for Critical Zone<br />

studies.<br />

Transport, deposition, and subsequent wea<strong>the</strong>ring <strong>of</strong><br />

eolian materials are controlled by climate. In areas<br />

distal from eolian source materials, fine sediments are<br />

deposited at relatively low rates; soil formation exceeds<br />

deposition, and <strong>the</strong> original parent material (e.g.<br />

crystalline bedrock) primarily influences pedogenesis<br />

(Birkeland, 1999; Muhs et al., 2008). Dust inputs<br />

are rapidly wea<strong>the</strong>red at <strong>the</strong> surface, and potentially<br />

<strong>of</strong>fset losses from <strong>the</strong> wea<strong>the</strong>ring <strong>of</strong> <strong>the</strong> original parent<br />

material (Mason and Jacobs, 1998). The volume<br />

and geochemistry <strong>of</strong> <strong>the</strong> dust determine its effect on<br />

pedogenesis.<br />

Based on field observations and laboratory analysis,<br />

this study seeks to: (1) track <strong>the</strong> concentration <strong>of</strong><br />

pedogenic iron within and between soil pr<strong>of</strong>iles to<br />

characterize wea<strong>the</strong>ring patterns in high-relief environments;<br />

and (2) assess <strong>the</strong> contribution and effects<br />

<strong>of</strong> eolian materials on montane soils.<br />

METHODS<br />

Field<br />

I selected 24 sites in <strong>the</strong> Boulder Creek catchment<br />

for field description and sampling (Fig. 1). The sites<br />

toge<strong>the</strong>r represent varying elevation, slope, aspect,<br />

parent material, moisture regime, and inferred age.<br />

Sites include upper montane Gordon Gulch (14), <strong>the</strong><br />

alpine and subalpine Green Lakes basin (5), a highway<br />

road cut near <strong>the</strong> town <strong>of</strong> Ward (2), and Betasso<br />

Gulch (2). Ten <strong>of</strong> <strong>the</strong> lower Gordon Gulch sites comprise<br />

two catenas (5 sites in each) that face north and<br />

south. Lower Gordon Gulch is oriented so that slopes<br />

have near north-south aspect directions. Sites on both<br />

transects stretch from just below <strong>the</strong> catchment crest<br />

to just above <strong>the</strong> break in slope with valley fans and<br />

131<br />

terrace deposits. At each site, I dug soil pits with my<br />

Keck colleagues until excessive depth or hard saprolite<br />

prevented fur<strong>the</strong>r excavation. I followed standard<br />

soil description procedures and used a Garmin GPS<br />

device to record <strong>the</strong> site location and elevation. I collected<br />

approximately 750g <strong>of</strong> soil from all horizons<br />

(98 total samples), with <strong>the</strong> exception <strong>of</strong> O and L<br />

horizons when present.<br />

Figure 1. Locations <strong>of</strong> soil pits that were dug, described<br />

and sampled in <strong>the</strong> Boulder Creek watershed.<br />

Laboratory<br />

In <strong>the</strong> laboratory I dry-sieved all samples using<br />

USDA standard soil sieves. I <strong>the</strong>n subdivided <strong>the</strong><br />


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

obtained from D.P. Dethier), I established background<br />

values for Fe 2 O 3 for each site. I <strong>the</strong>n calculated <strong>the</strong><br />

mass <strong>of</strong> accumulated Fe 2 O 3 at each site by subtracting<br />

<strong>the</strong> background from <strong>the</strong> total mass <strong>of</strong> Fe 2 O 3 to<br />

yield total accumulated Fe 2 O 3 (in g cm -2 ). I calculated<br />

total accumulated clay by <strong>the</strong> same method. I also<br />

ran samples from previous field studies in <strong>the</strong> Front<br />

Range for comparison, including a pr<strong>of</strong>ile formed in<br />

Bull Lake (130ka) till.<br />

RESULTS<br />

Field<br />

Soil morphology varies widely across <strong>the</strong> study area,<br />

reflecting <strong>the</strong> differences in parent material, relief,<br />

and climate within <strong>the</strong> Boulder Creek CZO (Fig. 2).<br />

Soils in <strong>the</strong> Green Lakes basin formed from glacial<br />

till and periglacial debris and, at stable sites, typically<br />

exhibit A/Ej/Bw/Cox/Cu pr<strong>of</strong>iles. Soils in <strong>the</strong> Betasso<br />

Gulch derive from ei<strong>the</strong>r wea<strong>the</strong>red colluvium<br />

or highly wea<strong>the</strong>red saprolite, and sample sites show<br />

thick (>50cm) Bt horizons and, locally, a thick buried<br />

soil.<br />

In Gordon Gulch <strong>the</strong> two catenas are broadly similar,<br />

but show marked differences near <strong>the</strong> footslope.<br />

Soils on <strong>the</strong> south-facing slope transect are thin (


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

age percent clay <strong>of</strong> <strong>the</strong> surface horizon and subsurface<br />

horizons from all 24 sites is 6.17% (1 sigma = 1.71)<br />

and 5.02% (1 sigma = 3.21), respectively. This indicates<br />

slight enrichment <strong>of</strong> <strong>the</strong> surface horizons in clay.<br />

Along <strong>the</strong> Gordon Gulch transects (9 sample sites),<br />

<strong>the</strong> average percent clay <strong>of</strong> surface and subsurface horizons<br />

is 5.07% (1 sigma = 1.08) and 3.12% (1 sigma<br />

= 1.07), respectively. Differences in silt concentration<br />

between surface and subsurface horizons are<br />

insignificant across <strong>the</strong> entire catchment. However,<br />

along <strong>the</strong> two Gordon Gulch transects, surface horizons<br />

may be slightly enriched in silt.<br />

Dithionite-extractable iron (Fe d ), as %Fe 2 O 3 , is<br />

slightly lower in surface horizons than subsurface<br />

horizons; average %Fe 2 O 3 <strong>of</strong> <strong>the</strong> surface horizon and<br />

subsurface horizons from 24 sites is 1.52% (1 sigma<br />

= 0.44) and 1.81% (1 sigma = 0.65), respectively. In<br />

Gordon Gulch, average %Fe 2 O 3 <strong>of</strong> <strong>the</strong> surface and<br />

subsurface horizons is 1.43% (1 sigma = 0.22) and<br />

1.74% (1 sigma = 0.46), respectively. Total pr<strong>of</strong>ile<br />

accumulated Fe 2 O 3 values are extremely variable<br />

across <strong>the</strong> catchment, with an average value <strong>of</strong> 1.84<br />

g cm-2 and a standard deviation <strong>of</strong> 1.51 g cm -2 . The<br />

maximum value is 5.1 g cm -2 at <strong>the</strong> site “WRC-01”;<br />

<strong>the</strong> minimum accumulated Fe 2 O 3 is 0.19 g cm-2 at <strong>the</strong><br />

site “SFT-03.”<br />

DISCUSSION<br />

Data collected in this study permit evaluation <strong>of</strong><br />

wea<strong>the</strong>ring rates and eolian sedimentation in <strong>the</strong><br />

Boulder Creek catchment. Dithionite-extractable iron<br />

(Fe d ) represents <strong>the</strong> concentration <strong>of</strong> secondary iron<br />

oxides and organically bound iron complexes in a soil<br />

(McFadden and Hendricks, 1985). Higher concentrations<br />

<strong>of</strong> Fe d in a sample thus indicate higher amounts<br />

<strong>of</strong> “free” iron that have been released by wea<strong>the</strong>ring<br />

and total pr<strong>of</strong>ile Fe d content typically increases<br />

with soil age and may correlate with total pr<strong>of</strong>ile clay<br />

content (McFadden and Hendricks, 1985). This<br />

relationship is expected in soils, as clay minerals and<br />

iron oxides are both products <strong>of</strong> wea<strong>the</strong>ring, and time<br />

positively influences <strong>the</strong> degree <strong>of</strong> wea<strong>the</strong>ring (Birkeland,<br />

1999). Therefore, variation in total pr<strong>of</strong>ile Fe d<br />

may indicate relative ages <strong>of</strong> soils across a study area.<br />

133<br />

Figure 3. Fe d accumulation rate, determined by using<br />

total pr<strong>of</strong>ile Fe d content <strong>of</strong> four soils <strong>of</strong> known age. The<br />

soil from Betasso Gulch (red bullet) formed from deep,<br />

wea<strong>the</strong>red colluvium and an uncertain portion <strong>of</strong> <strong>the</strong> Fe 2 O 3<br />

content is inherited. For this reason, <strong>the</strong> pr<strong>of</strong>ile was not<br />

used to fit <strong>the</strong> curve.<br />

Total pr<strong>of</strong>ile Fe d increases with age at stable sites in<br />

<strong>the</strong> Boulder Creek catchment (Fig. 3). The strong<br />

positive correlation <strong>of</strong> Fe d with age suggests that<br />

pedogenic clay should also increase. However,<br />

textural data show that surface horizons, relative to<br />

subsurface horizons, are commonly enriched in clay<br />

but depleted in Fe d (Table 1). The enrichment <strong>of</strong> clay<br />

in surface horizons, coupled with relatively low Fe d<br />

values, suggests that clay content indicates eolian<br />

enrichment <strong>of</strong> soils in <strong>the</strong> catchment. Clay-sized sediment<br />

suggests a distal source.<br />

Total pr<strong>of</strong>ile Fe d and clay content suggest both wea<strong>the</strong>ring<br />

and transport mechanics on slopes in Gordon<br />

Gulch (Fig. 4). On both north-facing and south-facing<br />

slopes, Fe d and clay content in soil and regolith<br />

increases downslope.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 4. Total pr<strong>of</strong>ile accumulated Fed content and total<br />

pr<strong>of</strong>ile clay content along <strong>the</strong> two Gordon Gulch transects.<br />

The north-facing slope is in black and <strong>the</strong> south-facing<br />

slope in red. The filled markers and solid lines represent<br />

Fed content, and <strong>the</strong> hollow markers and dashed lines<br />

represent clay content.<br />

The correlation <strong>of</strong> clay and Fe d content on both slopes<br />

indicates that regolith is “older” downslope. Fed<br />

and clay accumulate more rapidly on <strong>the</strong> north-facing<br />

slope than on <strong>the</strong> south-facing slope. Therefore,<br />

wea<strong>the</strong>ring rates may be higher on <strong>the</strong> north-facing<br />

slope, perhaps reflecting differences in temperature<br />

and moisture content due to aspect.<br />

CONCLUSIONS<br />

Pedogenic iron in <strong>the</strong> Boulder Creek catchment increases<br />

with age. Total pr<strong>of</strong>ile Fe d content correlates<br />

with total pr<strong>of</strong>ile clay content; this suggests that clays<br />

in sampled pr<strong>of</strong>iles are primarily pedogenic. However,<br />

clay content is slightly higher in surface horizons<br />

than in subsurface horizons and Fe d content is slightly<br />

reduced, indicating that <strong>the</strong> surface horizons are<br />

enriched in clay. High clay and low iron content in<br />

surface horizons suggests an eolian source ra<strong>the</strong>r than<br />

intensified in-situ wea<strong>the</strong>ring.<br />

Future work will focus on (1) determining <strong>the</strong> significant<br />

factors affecting eolian deposition (e.g. elevation,<br />

MAP, aspect, etc.); (2) incorporating trace-element<br />

134<br />

data in surface and parent material horizons to determine<br />

local or distant provenance <strong>of</strong> eolian clays<br />

and (3) using Fe d accumulation rate to infer ages <strong>of</strong><br />

sample sites in <strong>the</strong> catchment.<br />

ACKNOWLEDGEMENTS<br />

I’d like to thank my <strong>the</strong>sis advisor, David P. Dethier<br />

for his advice, support, and enthusiasm. Secondly,<br />

I’d like to thank my fellow Keck colleagues for <strong>the</strong>ir<br />

efforts in <strong>the</strong> field. I’d also like to thank all <strong>of</strong> those<br />

working in <strong>the</strong> Boulder Creek CZO who shared <strong>the</strong>ir<br />

knowledge, <strong>the</strong> Williams College Geosciences Department,<br />

<strong>the</strong> Sperry Fund, and <strong>the</strong> Keck Geology<br />

Consortium.<br />

REFERENCES<br />

Anderson, S.P., von Blanckenburg, F., and White,<br />

A.F., 2007, Physical and chemical controls on <strong>the</strong><br />

Critical Zone: Elements, v. 3, p. 315-319.<br />

Anderson, R.S., and Anderson, S.P., 2010, Geomorphology:<br />

<strong>the</strong> mechanics and chemistry <strong>of</strong> landscapes:<br />

Cambridge, Cambridge University Press,<br />

637 p.<br />

Birkeland, P.W., 1999, Soils and Geomorphology: New<br />

York, Oxford University Press, 430 p.<br />

Gee, G.W., and J.W. Bauder. 1986. Particle-size analysis,<br />

p. 383-411. In A. Klute (ed.) Methods <strong>of</strong> soil<br />

analysis. Part 1. 2nd ed. Agron. Monogr. 9. ASA<br />

and SSSA, Madison, WI.<br />

Mason, J.A., and Jacobs, P.M., 1998, Chemical and<br />

particle-size evidence for addition <strong>of</strong> fine dust to<br />

soils <strong>of</strong> <strong>the</strong> Midwestern United States: Geology, v.<br />

26, p. 1135-1138.<br />

McFadden, L.D., and Hendricks, D.M., 1985, Changes<br />

in <strong>the</strong> content and composition <strong>of</strong> pedogenic<br />

iron oxyhydroxides in a chronosequence <strong>of</strong> soils<br />

in Sou<strong>the</strong>rn California: Quaternary Research, v.<br />

23, p. 189-204.<br />

Muhs, D.R., and Benedict, J.B., 2006, Eolian additions<br />

to Late Quaternary alpine soils, Indian Peaks


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Wilderness Area, Colorado Front Range: Artic,<br />

Antartic, and Alpince Research, v. 38, p. 120-130.<br />

Muhs, D.R., Budahn, J.R., Johnson, D.L., Reheis, M.,<br />

Beann, J., Skipp, G., Fisher, E., and Jones, J.A.,<br />

2008, geochemical evidence for airborne dust additions<br />

to soils in Channel Islands National Park,<br />

California: Geological Society <strong>of</strong> America Bulletin,<br />

v. January/February.<br />

135


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

USING POLLEN TO UNDERSTAND QUATERNARY<br />

PALEOENVIRONMENTS IN BETASSO GULCH, COLORADO<br />

COREY SHIRCLIFF, Beloit College<br />

Research Advisor: Carl Mendelson<br />

INTRODUCTION<br />

The mountain ranges <strong>of</strong> <strong>the</strong> western United States<br />

differ in climate and precipitation, and thus plant<br />

life varies as well. During <strong>the</strong> Pleistocene, oscillating<br />

climatic trends created sequences <strong>of</strong> glacial and<br />

interglacial times (Charlesworth, 1957); <strong>the</strong> Holocene<br />

is considered a modern interglacial epoch (Traverse,<br />

2007). By using palynological techniques, it is possible<br />

to understand how glacial climates have affected<br />

<strong>the</strong> environment. By documenting <strong>the</strong> current geographical<br />

trends <strong>of</strong> plant species, and <strong>the</strong>n applying<br />

that information to understanding Quaternary pollen<br />

samples, it is possible to reconstruct local environments<br />

over time.<br />

In <strong>the</strong> Rocky Mountains, <strong>the</strong> most recent glaciation is<br />

known as <strong>the</strong> Pinedale. In <strong>the</strong> Colorado Front Range,<br />

it occurred from about 30,000 YBP (years before<br />

present) to 15,000 to 12,000 years YBP (Legg and<br />

Baker, 1980). As <strong>the</strong> deglaciation occurred at higher<br />

elevations, <strong>the</strong> flora <strong>of</strong> <strong>the</strong> entire Front Range area<br />

responded. Certain floral taxa are important in determining<br />

<strong>the</strong> post-glacial environment, as discussed in<br />

detail later. One way to understand <strong>the</strong> paleoclimate<br />

<strong>of</strong> this dynamic time is by using palynological methods<br />

to examine <strong>the</strong> pollen in sediments that are known<br />

to be younger than 12,000 years in age.<br />

Betasso Gulch, located in <strong>the</strong> north-central Front<br />

Range <strong>of</strong> Colorado (Fig. 1), is a good place to collect<br />

samples for a palynological analysis due to <strong>the</strong> development<br />

<strong>of</strong> an organic-rich A soil horizon, which is<br />

buried under a modern soil pr<strong>of</strong>ile. Both radiocarbon<br />

and optically stimulated luminescence (OSL) dates<br />

are available for <strong>the</strong> soil horizons exposed in Betasso.<br />

A regional laterally continuous buried A horizon corresponds<br />

to a time between 9,000 and 6,000 years ago<br />

(see Fig. 2). This is an interesting interval to investigate<br />

palynologically, because <strong>the</strong> response <strong>of</strong><br />

<strong>the</strong> flora at this time and elevation (~ 2,000 m) is not<br />

well known in this area. Samples were taken from<br />

this buried A horizon and from modern forest litter<br />

to better understand how <strong>the</strong> climate has changed at<br />

Betasso Gulch.<br />

Figure 1. Betasso Gulch sample locality (red triangle).<br />

Previously studied sites: Devlins Park (Legg and Baker,<br />

1980) and Redrock Lake (Maher, 1972). Modified from<br />

Marr (1961) in Legg and Baker (1980, fig. 1).<br />

Although a wealth <strong>of</strong> pollen studies have been conducted<br />

in <strong>the</strong> Rocky Mountains and surrounding<br />

mountain ranges, two particularly relevant papers<br />

focus on pollen data from <strong>the</strong> Front Range, close to<br />

my sample site. Unfortunately, both localities are<br />

significantly higher in elevation (Fig. 1), but comparison<br />

with Betasso may still yield important information.<br />

Legg and Baker (1980) studied Pinedale-age<br />

lacustrine sediments from Lake Devlin; <strong>the</strong> sediments<br />

range in age from about 22,000 to 12,000 radiocarbon<br />

years before present (RCYBP). The authors<br />

found significant pollen contributions from Artemisia<br />

(sagebrush and relatives) at 42%, Poaceae (grasses)<br />

at 13%, and Cyperaceae (sedges) at 4%; Alnus (alder)<br />

and Betula (birch) appeared in small quantities at <strong>the</strong><br />

top <strong>of</strong> <strong>the</strong> section.<br />

136


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 2. Dated stratigraphic section 50 m downstream<br />

from <strong>the</strong> Betasso Gulch locality. The buried soil A horizon<br />

is labeled IIIAb. Radiocarbon dates and image from<br />

Dethier (written commun., 2011).<br />

The top <strong>of</strong> <strong>the</strong> Lake Devlin section is dated at<br />

~12,000 RCYBP, or about 14,600 calibrated years,<br />

which is slightly older than <strong>the</strong> bottom <strong>of</strong> <strong>the</strong> stratigraphic<br />

section at Betasso Gulch. The o<strong>the</strong>r relevant<br />

study is <strong>of</strong> Redrock Lake, about nine kilometers north<br />

<strong>of</strong> Devlins Park and at a similar elevation (Maher,<br />

1972). Maher investigated <strong>the</strong> time from 10,000<br />

RCYBP to <strong>the</strong> modern. He found an increase <strong>of</strong><br />

Pinus (pine) and a decrease <strong>of</strong> Picea (spruce) upsection.<br />

Additionally, maxima for deposition <strong>of</strong> pollen<br />

grains occurred from 7,500 to 3,500 RCYBP and<br />

Picea deposition peaked around 8,500 RCYBP. He<br />

concluded that <strong>the</strong> interval from 6,700 to 7,600 YBP<br />

was cooler and/or wetter than earlier post-Pinedale<br />

times.<br />

MODERN VEGETATION AND CLIMATE<br />

It is possible to classify vegetation zones on <strong>the</strong><br />

eastern slope <strong>of</strong> <strong>the</strong> Front Range (Legg and Baker,<br />

1980; see Fig. 1). Beginning on <strong>the</strong> eastern plains<br />

is <strong>the</strong> grassland zone, which is below 1,800 meters<br />

137<br />

in elevation. Moving westward, on <strong>the</strong> lower slopes<br />

<strong>of</strong> <strong>the</strong> mountains, is <strong>the</strong> montane zone, which ranges<br />

from 1,800 to 2,800 meters; Betasso Gulch is located<br />

in this zone at an elevation <strong>of</strong> about 2000 m.<br />

Higher yet are <strong>the</strong> subalpine (2,800 to 3,350 m) and<br />

alpine (above 3,350 m) zones. The montane zone is<br />

characterized by an open forest populated by Pinus<br />

ponderosa (Ponderosa Pine), Pinus contorta (Lodgepole<br />

Pine), and Pseudotsuga menziesii (Douglas Fir),<br />

along with many shrubs, grasses, and herbs (Legg<br />

and Baker, 1980). This description is reflected in <strong>the</strong><br />

modern sample I collected (see Results).<br />

METHODS<br />

Field Methods<br />

Samples were collected at Gordon Gulch and Betasso<br />

Gulch. The Gordon Gulch samples, while rich in<br />

pollen, were not included in this paper due to time<br />

constraints in pollen identification and <strong>the</strong> fact that<br />

Gordon Gulch occurs at a higher elevation, and thus<br />

may have been more affected by recent glaciation.<br />

Samples at Betasso Gulch were collected at five-cm<br />

intervals from <strong>the</strong> base (1.2 m) to <strong>the</strong> top (1.05 m) <strong>of</strong><br />

<strong>the</strong> buried A horizon (Fig. 2, level IIIAb). The succeeding<br />

meter <strong>of</strong> sediment was not sampled due to<br />

bioturbation by roots, resulting in a mix <strong>of</strong> pollen<br />

grains from different levels. The sampled horizon<br />

was located in a channel, which had been eroded by a<br />

flood from pipes for a nearby water-treatment facility.<br />

A modern sample was taken at <strong>the</strong> surface, in an area<br />

at least two meters from any tree, and without significant<br />

undergrowth. Several handfuls <strong>of</strong> surface sediment<br />

were collected in an area <strong>of</strong> about one square<br />

meter, to help rid <strong>the</strong> sample <strong>of</strong> a bias toward grasses<br />

and shrubs growing on <strong>the</strong> surface. Therefore, five<br />

total samples were collected: four from <strong>the</strong> buried A<br />

horizon and one modern sample. All <strong>of</strong> <strong>the</strong>se samples<br />

were composed <strong>of</strong> dry sediment and organic debris.<br />

Laboratory Methods<br />

Pollen was extracted from <strong>the</strong> sediment and concentrated<br />

using a variety <strong>of</strong> chemical methods (see<br />

Traverse, 2007). After washing about 2 ml <strong>of</strong> sample<br />

in 5% KOH, <strong>the</strong> residue was filtered using a 250-mm<br />

screen to exclude rock fragments and large organic<br />

debris. The sample was <strong>the</strong>n washed in HCl to dis-


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

solve carbonates, and washed again in hydr<strong>of</strong>luoric<br />

acid to dissolve silicates. After rinsing in acetic acid,<br />

residues were acetolyzed (nine parts acetic anhydride<br />

to one part sulfuric acid) in order to dissolve <strong>the</strong> part<br />

<strong>of</strong> <strong>the</strong> exine <strong>of</strong> <strong>the</strong> pollen grain that does not consist<br />

<strong>of</strong> <strong>the</strong> highly resistant sporopollenin. After diluting<br />

thoroughly in water, residues were washed through a<br />

10-mm screen. Remaining residues were successively<br />

dewatered with 95% ethanol and tert-butyl alcohol<br />

(TBA). Storage in TBA prevented <strong>the</strong> pollen grains<br />

from deflating or degrading. Slides were made using<br />

silicon oil as a mounting medium, and cover glasses<br />

were sealed with colored nail polish.<br />

Identification Methods<br />

It is necessary to identify 300 pollen grains from each<br />

sample in order to reach <strong>the</strong> 95% confidence interval<br />

for this kind <strong>of</strong> pollen analysis (Traverse, 2007).<br />

Pollen grains were identified using a Zeiss compound<br />

microscope under 40x and 100x (with oil) objectives.<br />

Identification was confirmed by referring to Kapp<br />

(1990) and to reference slides <strong>of</strong> modern pollen grains<br />

(courtesy <strong>of</strong> Robert Nelson, Colby College). Some<br />

grains were impossible to identify due to folding and<br />

exine degradation.<br />

RESULTS<br />

Modern<br />

Figure 3 shows selected pollen grains. High amounts<br />

<strong>of</strong> Pinus grains were identifiable; <strong>the</strong>y represented at<br />

least 67% <strong>of</strong> <strong>the</strong> Betasso sample. Almost all <strong>of</strong> <strong>the</strong><br />

unidentified pollen from <strong>the</strong> modern sample (9%) is<br />

arboreal; much <strong>of</strong> it probably belongs to Pinus, so<br />

67% is a minimum. Pseudotsuga menziesii is also an<br />

important arboreal component, at 3% <strong>of</strong> <strong>the</strong> modern<br />

sample. Juniperus (juniper) and Artemisia (sagebrush)<br />

toge<strong>the</strong>r represent <strong>the</strong> majority <strong>of</strong> <strong>the</strong> shrubs<br />

found in <strong>the</strong> Front Range area, and constitute 21% <strong>of</strong><br />

<strong>the</strong> Betasso modern sample. Grasses (Poaceae) are at<br />

a fairly average level (5%).<br />

Early Holocene<br />

Pinus, Artemisia, Poaceae, and Asteraceae (composites)<br />

contributed <strong>the</strong> most pollen grains to <strong>the</strong> early<br />

Holocene samples (Fig. 4). Progressing upward at<br />

138<br />

Figure 3. Light-microscope images <strong>of</strong> Betasso spores and<br />

pollen grains. A) unidentified spore found throughout <strong>the</strong><br />

buried A horizon. B) degraded Pseudotsuga menziesii<br />

(Douglas Fir) pollen grain. C) Pinus ponderosa pollen<br />

grain. D) Ambrosia artemesiifolia (common ragweed)<br />

pollen grain. (Ambrosia is a composite, and thus a member<br />

<strong>of</strong> <strong>the</strong> Asteraceae.)<br />

Betasso, <strong>the</strong>re is a slight increase <strong>of</strong> Pinus; in <strong>the</strong><br />

modern sample, Pinus is most common by far (67%).<br />

Artemisia (sagebrush and wormwood) pollen also<br />

represented a significant percent, but in <strong>the</strong> modern<br />

sample it dwindled to 18%. It was rarely possible to<br />

identify Artemisia grains to <strong>the</strong> species level. Poaceae<br />

and Asteraceae both showed decreasing trends<br />

(Fig. 4).<br />

Pseudotsuga menziesii (Douglas Fir) represented<br />

less than 1%, and in some cases 0%, in <strong>the</strong> buried A<br />

samples, but had a 3% representation in <strong>the</strong> modern<br />

sample. Because P. menziesii rarely rises above a 5%<br />

pollen representation in modern-day forests where it<br />

is known to grow in large quantities (Whitlock, 1993),<br />

I consider this result to be important.<br />

In <strong>the</strong> samples from <strong>the</strong> buried A horizon, many<br />

grains, particularly those <strong>of</strong> arboreal pollen types,<br />

were folded in a way that made identification impos-


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

sible. They were counted, and a percent <strong>of</strong> unidentifiable<br />

grains was found. The most unidentified grains<br />

were in <strong>the</strong> middle <strong>of</strong> <strong>the</strong> buried A section, rising to<br />

almost 20% in one sample.<br />

Figure 4. Pollen percentage diagram (includes identified<br />

pollen only). Note <strong>the</strong> scale break between 0 and 1.05<br />

m. Dashed lines represent unsampled interval and are an<br />

estimate only <strong>of</strong> <strong>the</strong> trend <strong>of</strong> <strong>the</strong> pollen percents. The modern<br />

sample is plotted at 0 m. The percents <strong>of</strong> total pollen<br />

grains that could not be identified were as follows: 1.20 m<br />

(13%); 1.15 m (19%); 1.10 m (17%); 1.05 m (15%); 0 m<br />

(9%).<br />

DISCUSSION<br />

Pollen percentages are typically used to indicate<br />

changes in climate (Table 1). In modern-day forests,<br />

pine trees grow in relatively dry areas. Where precipitation<br />

increases, o<strong>the</strong>r arboreal species tend to<br />

win out in <strong>the</strong> battle for forest dominance. Pines in<br />

particular flourish in dry conditions, as seen in current-day<br />

Betasso Gulch, which hosts a pine-enriched<br />

forest; significant species include Pinus ponderosa<br />

and P. contorta. Because <strong>the</strong> percentages <strong>of</strong> P. ponderosa<br />

were lower in <strong>the</strong> buried A horizon, I infer that<br />

<strong>the</strong> climate at 9 ka was probably wetter and/or cooler<br />

than today’s climate. Today in <strong>the</strong> Front Range, P.<br />

ponderosa does not grow above <strong>the</strong> montane zone<br />

due to its inability to thrive in <strong>the</strong> cold temperatures<br />

experienced in <strong>the</strong> subalpine and alpine zones (Weber,<br />

1976). Therefore, because <strong>the</strong> P. ponderosa percent<br />

increases from <strong>the</strong> buried A section to <strong>the</strong> modern, <strong>the</strong><br />

environment may have become more favorable for<br />

Ponderosa Pines. Hence, it was becoming warmer<br />

139<br />

and drier from <strong>the</strong> buried A to present. The pollen<br />

percent <strong>of</strong> total Pinus species almost doubles in <strong>the</strong><br />

modern sample; this substantial leap suggests that <strong>the</strong><br />

climate has become considerably warmer and drier<br />

than <strong>the</strong> time represented at <strong>the</strong> top <strong>of</strong> <strong>the</strong> buried A<br />

section (1.05 m below <strong>the</strong> surface), that is, since <strong>the</strong><br />

early Holocene.<br />

Taxon Common Name Climate Implications<br />

(If pollen levels increase)<br />

Picea Spruce Moister<br />

Pseudotsuga menziesii Douglas Fir Drier<br />

Poaceae Grasses Drier<br />

Chenopodiaceae Herbs and Subshrubs Drier<br />

Pinus Pine Drier and warmer<br />

Artemisia Sagebrush Drier and warmer<br />

Table 1. Climate trends associated with increases <strong>of</strong> certain<br />

taxa in pollen samples.<br />

Whitlock (1993) concluded that a decrease in Artemisia<br />

indicated a wetter, cooler climate. She also<br />

argued that an increase in Picea suggested a wetter,<br />

cooler climate. In <strong>the</strong> surface sample, Picea decreases<br />

from 10% to 2% (indicating warming) and<br />

Artemisia levels decrease by 11% (indicating cooling).<br />

Although <strong>the</strong> decrease in Artemisia indicates a<br />

wetter and cooler environment, <strong>the</strong> decrease in Picea<br />

and significant increase in Pinus suggest a warmer<br />

environment. Moreover, although <strong>the</strong> Amaranthaceae<br />

+ Chenopodiaceae pollen group stays fairly constant<br />

in <strong>the</strong> buried A horizon, its percentage rises by 5% in<br />

<strong>the</strong> modern sample. Bright (1966) and Davis et al.<br />

(1986) inferred a drier climate when Amaranthaceae +<br />

Chenopodiaceae levels increased. Lastly, an increase<br />

in Pseudotsuga menziesii (Douglas Fir) has been<br />

shown to indicate warming (Whitlock, 1993). The<br />

increase from 0% to 3% from <strong>the</strong> buried A to modern<br />

gives fur<strong>the</strong>r evidence for warming, keeping in mind<br />

that 5% is <strong>the</strong> maximum percent P. menziesii reaches<br />

in any forest (Whitlock, 1993).<br />

I conclude that <strong>the</strong> environment at Betasso Gulch<br />

changed from a wetter, cooler climate (soil horizon<br />

A) to <strong>the</strong> current warmer and more arid conditions.<br />

Maher (1972) argued that <strong>the</strong> time from about 6,700<br />

to 7,600 YBP was cool and wet, compared to <strong>the</strong> preceding<br />

~3,000 years. Although evidence for cooling<br />

within <strong>the</strong> buried A horizon is weak, <strong>the</strong>re is a pos


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

sibility that this section fits in slightly before 7,600<br />

YBP, or between <strong>the</strong>n and 6,700 YBP, and may thus<br />

be consistent with <strong>the</strong> findings <strong>of</strong> Maher (1972). Fur<strong>the</strong>r<br />

studies <strong>of</strong> this buried soil horizon should be undertaken<br />

before such a conclusion can be confirmed.<br />

My tentative conclusions are not statistically robust.<br />

For example, <strong>the</strong> percents reported for <strong>the</strong> buried<br />

A horizon are at best estimates because 15-20% <strong>of</strong><br />

<strong>the</strong> pollen grains could not be identified due to folding<br />

and degradation. In <strong>the</strong> modern sample, most<br />

unidentified pollen grains are bisaccate, so are probably<br />

Pinus species. Most significantly, <strong>the</strong> samples<br />

from <strong>the</strong> buried A horizon might not yield reliable<br />

information regarding trends in pollen percents—this<br />

horizon, like <strong>the</strong> succeeding meter <strong>of</strong> sediment, may<br />

have been bioturbated, resulting in homogenization<br />

<strong>of</strong> <strong>the</strong> micr<strong>of</strong>lora. For a more accurate understanding<br />

<strong>of</strong> climate change in this area, more samples should<br />

be taken from <strong>the</strong> buried A horizon, and additional<br />

surface samples need to be collected to confirm <strong>the</strong><br />

pollen percents representative <strong>of</strong> <strong>the</strong> current climate.<br />

It would also be beneficial to collect older samples in<br />

<strong>the</strong> section. Finally, <strong>the</strong> discrepancy between <strong>the</strong> radiocarbon<br />

and OSL ages (Fig. 2) needs to be resolved.<br />

ACKNOWLEDGMENTS<br />

I would like to pr<strong>of</strong>usely thank Dr. Bob Nelson <strong>of</strong><br />

Colby College for his tremendous assistance in this<br />

project. I would also like to thank my advisor at Beloit<br />

College, Dr. Carl Mendelson, for his mentorship<br />

and bravery in <strong>the</strong> face <strong>of</strong> a student using HF, and<br />

my advisor Dr. Jim Rougvie, for taking over advising<br />

when Carl went on sabbatical. Thanks are also in order<br />

to Dr. Carol Mankiewicz for assistance with figure<br />

preparation. Lastly, I thank both Dr. David Dethier<br />

(Williams College) and Dr. Will Ouimet (University<br />

<strong>of</strong> Connecticut) for <strong>the</strong>ir mentorship this summer and<br />

throughout <strong>the</strong> year.<br />

REFERENCES<br />

Bright, R.C., 1966, Pollen and seed stratigraphy <strong>of</strong><br />

Swan Lake, south-eastern Idaho; its relation to<br />

regional vegetational history and to Lake Bonn-<br />

140<br />

eville history: Tebiwa, v. 9, p. 1-47.<br />

Charlesworth, J.K., 1957, The Quaternary Era: with<br />

special reference to its glaciation: London, E.<br />

Arnold, 1601 p.<br />

Davis, O. K., Sheppard, J. C., and Robertson,<br />

S., 1986, Contrasting climatic histories<br />

for <strong>the</strong> Snake River Plain, Idaho, resulting from<br />

multiple <strong>the</strong>rmal maxima: Quaternary Research,<br />

v. 26, p. 321-339.<br />

Kapp, R. O., 2000, Pollen and spores (2d ed.): College<br />

Station, Texas, American Association <strong>of</strong><br />

Stratigraphic Palynologists, 279 p.<br />

Legg, T. E., and Baker, R. G., 1980, Palynology<br />

<strong>of</strong> Pinedale sediments, Devlins Park, Boulder<br />

County, Colorado: Arctic and Alpine Research,<br />

v. 12, no. 3, p. 319-333.<br />

Maher, L.J., Jr., 1972, Absolute pollen diagram <strong>of</strong> Redrock<br />

Lake, Boulder County, Colorado: Quaternary<br />

Research, v. 2, no. 4, p. 531-553.<br />

Traverse, A., 2007, Paleopalynology (2d ed.): Boston,<br />

Unwin Hyman, 600 p.<br />

Weber, W. A., 1976, Rocky Mountain flora (5th ed.):<br />

Niwot, University Press <strong>of</strong> Colorado, 479 p.<br />

Whitlock, C., 1993, Postglacial vegetation and climate<br />

<strong>of</strong> Grand Teton and sou<strong>the</strong>rn Yellowstone<br />

National Parks: Ecological Monographs, v. 63,<br />

p. 173-198.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

STREAM TERRACES IN THE CRITICAL ZONE – LOWER<br />

GORDON GULCH, COLORADO<br />

KATHLEEN WARRELL, Georgia Tech<br />

Research Advisor: Kurt Frankel<br />

INTRODUCTION<br />

Systems <strong>of</strong> stream terraces provide insight into <strong>the</strong><br />

history <strong>of</strong> a stream and how <strong>the</strong> surrounding landscape<br />

has changed throughout geologic history.<br />

Stream terraces are an integral part <strong>of</strong> <strong>the</strong> Critical<br />

Zone (CZ), which is defined as <strong>the</strong> boundary layer<br />

that extends from <strong>the</strong> buried, unwea<strong>the</strong>red bedrock up<br />

through wea<strong>the</strong>red rock and regolith to <strong>the</strong> soil where<br />

terrestrial life thrives (Anderson et al., 2007). The<br />

CZ is <strong>the</strong> vital place on Earth’s surface where rocks,<br />

soil, atmospheric gasses, and meteoric water interact.<br />

Anderson et al. (2007) described <strong>the</strong> CZ as a “feedthrough<br />

reactor” that transforms solid bedrock into<br />

soil and sediment, which is <strong>the</strong>n transported downslope<br />

into a stream channel.<br />

The morphology <strong>of</strong> a stream and its floodplain is <strong>the</strong><br />

result <strong>of</strong> a delicate balance <strong>of</strong> driving and resisting<br />

forces. Excess erosion on surrounding hillslopes can<br />

cause aggradation and increase <strong>the</strong> stream elevation.<br />

Aggraded sediment is removed when it is entrained<br />

by <strong>the</strong> stream. Sediment entrainment and deposition<br />

by a stream are driven by <strong>the</strong> depth and slope <strong>of</strong> that<br />

stream; <strong>the</strong>y are resisted by channel configuration,<br />

sediment size and sediment concentration (Ritter et<br />

al., 2002).<br />

Fill terraces are especially important in <strong>the</strong> CZ because<br />

<strong>the</strong>y store sediment and biomass eroded from<br />

surrounding hillslopes. Fill terraces are extremely<br />

productive areas in a stream valley, as <strong>the</strong>y provide<br />

a stable, flat environment with organic-rich soil on<br />

which plants and animals thrive. However, <strong>the</strong>se<br />

terraces are only temporary features in many landscapes,<br />

as stream incision and sediment entrainment<br />

are constantly removing sediment from <strong>the</strong> terraces.<br />

This study uses terrace morphology <strong>of</strong> Lower Gordon<br />

Gulch to estimate <strong>the</strong> volume <strong>of</strong> sediment stored in<br />

<strong>the</strong>se terraces and to model <strong>the</strong> timescale to remove<br />

141<br />

all <strong>of</strong> this sediment from <strong>the</strong> Gulch.<br />

GEOGRAPHIC AND GEOLOGIC SETTING<br />

The study area for this project is <strong>the</strong> 3.76 square kilometer<br />

Gordon Gulch catchment in Boulder County,<br />

Colorado. Gordon Gulch is a tributary <strong>of</strong> North<br />

Boulder Creek; it joins North Boulder Creek about 16<br />

kilometers from its headwaters. Elevations in Gordon<br />

Gulch range from 2,400 meters to 2,700 meters.<br />

Gordon Gulch is separated informally into two sections<br />

– Lower Gordon Gulch and a large tributary that<br />

constitutes Upper Gordon Gulch. A large knickpoint<br />

lies between Lower and Upper Gordon Gulch (Fig. 1).<br />

The stream in Upper Gordon Gulch is intermittent;<br />

however <strong>the</strong> majority <strong>of</strong> <strong>the</strong> stream in Lower Gordon<br />

Gulch contains water in most years.<br />

Figure 1. Map <strong>of</strong> Gordon Gulch showing location in<br />

Boulder County and Colorado. Map <strong>of</strong> Gordon Gulch is<br />

a hillshade derived from lidar flown in August 2010 with<br />

a pixel size <strong>of</strong> 1 m 2 . The start and end <strong>of</strong> <strong>the</strong> terrace map<br />

section are noted.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Stream aggradation is sensitive to local changes in<br />

land use. During <strong>the</strong> late 1800s to early 1900s as<br />

miners began working in <strong>the</strong> area surrounding Gordon<br />

Gulch, land use changed drastically. The introduction<br />

<strong>of</strong> prospecting and small-scale mining may have<br />

generated a large amount <strong>of</strong> sediment that aggraded in<br />

Gordon Gulch. The frequency <strong>of</strong> fires also increased<br />

at this time, resulting in an increase in erosion in <strong>the</strong><br />

catchment (Goldblum and Veblen, 1992).<br />

METHODS<br />

A Laser Technology Tru-Pulse 360 laser rangefinder<br />

was used to produce a detailed base map <strong>of</strong> <strong>the</strong><br />

stream. The rangefinder has an accuracy <strong>of</strong> ±0.20 meters<br />

slope distance, ±0.25 degrees slope angle, and ±1<br />

degree azimuthal angle. Azimuthal angle was used in<br />

conjunction with horizontal distance measurements<br />

to produce x and y coordinates. The z coordinate was<br />

calculated using a base-level measurement from a<br />

GPS and cumulative vertical distance measurements.<br />

These coordinates were graphed in Matlab with equal<br />

axes to produce a base map for mapping terraces.<br />

Stream morphology and a series <strong>of</strong> flags placed along<br />

<strong>the</strong> stream were used to mark terraces on <strong>the</strong> map<br />

relative to <strong>the</strong>ir location along <strong>the</strong> stream. Terraces<br />

were differentiated based upon <strong>the</strong>ir morphology and<br />

height relative to surrounding terraces and hillslopes.<br />

The rangefinder was used to measure <strong>the</strong> height <strong>of</strong><br />

each terrace above <strong>the</strong> stream channel. Seventy-five<br />

tree core ages were collected from trees growing<br />

on <strong>the</strong> terraces to approximate <strong>the</strong> age each terrace<br />

stabilized. Two samples <strong>of</strong> buried wood were also<br />

collected from <strong>the</strong> terraces for 14 C dating.<br />

A series <strong>of</strong> eight detailed cross sections were measured<br />

along <strong>the</strong> stream using <strong>the</strong> rangefinder (Fig.<br />

2C). Valley-wide cross sections were extracted from<br />

a high resolution digital elevation model to estimate<br />

<strong>the</strong> slope <strong>of</strong> <strong>the</strong> bedrock in surrounding hillslopes<br />

(Fig. 2B). Riemann sums were used to calculate cross<br />

sectional area <strong>of</strong> sediment between <strong>the</strong> bedrock slope<br />

and terrace cross section (Fig. 2C). The area was<br />

multiplied by <strong>the</strong> distance upstream to <strong>the</strong> next cross<br />

section, and all volumes were summed to obtain <strong>the</strong><br />

total volume <strong>of</strong> sediment stored in <strong>the</strong> terraces (V s ).<br />

142<br />

Figure 2. (A) Map view <strong>of</strong> terraces at KW-ST-10 with<br />

cross section X-X’ marked. Tors (Qt) are shown. (B)<br />

Valley-wide cross section derived from lidar showing<br />

estimated slope angles <strong>of</strong> <strong>the</strong> bedrock boundary. (C) Cross<br />

section and map view <strong>of</strong> stream terrace map at location<br />

KW-ST-10 showing terraces Qt4 and Qt5 and <strong>the</strong> cross<br />

sectional area <strong>of</strong> sediment.<br />

The cross sections were also used to determine bankfull<br />

width (B) and hydraulic radius (R) <strong>of</strong> <strong>the</strong> active<br />

stream. Hydraulic radius is:<br />

where:<br />

R = HB[2H + B] -1 ,<br />

H = Q w B -1 v -1 ,<br />

where H is bankfull depth, Q w is maximum discharge<br />

<strong>of</strong> water in <strong>the</strong> stream, and v is velocity <strong>of</strong> <strong>the</strong> stream.<br />

Bankfull velocity was approximated at 0.5 m/s.<br />

Maximum discharge <strong>of</strong> water from Gordon Gulch<br />

was calculated as <strong>the</strong> 90 th percentile <strong>of</strong> daily discharge<br />

data from <strong>the</strong> stream gauge over <strong>the</strong> year 2009. At<br />

each cross section location a sample <strong>of</strong> stream sediment<br />

was collected. The 84 th percentile grain size


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

(D 84 , in meters) <strong>of</strong> each sample was determined by<br />

sieving <strong>the</strong> samples.<br />

These data were used to determine <strong>the</strong> maximum<br />

sediment flux (Q s ) <strong>of</strong> <strong>the</strong> stream, which is adapted<br />

from Mueller and Pitlick (2005) as:<br />

Q = 11.2(θ-θ ) s c 4.5θ -3 3 0.5 [(s-1)gD ] B 84<br />

where θ is Shields stress, θ c is critical Shields stress<br />

(approximated at 0.03), s is specific gravity <strong>of</strong> sediment,<br />

and g is gravity. Shields stress is defined as:<br />

θ = τ[(ρ s -ρ w )gD 84 ] -1 ,<br />

where τ is shear stress (τ = ρ w gRS), ρ s and ρ w are sediment<br />

and water densities, g is gravity, and S is decimal<br />

slope (Cronin et al., 2007). Slope was measured<br />

from <strong>the</strong> base map at each sample location. When θ<br />

is equal to 0.03, <strong>the</strong> stream is capable <strong>of</strong> transporting<br />

all <strong>the</strong> sediment in its bed. Below this value, <strong>the</strong><br />

stream cannot transport all <strong>of</strong> its sediment. When θ<br />

is above 0.07 <strong>the</strong> stream is capable <strong>of</strong> carrying more<br />

sediment than its bed contains.<br />

The sediment removal time-scale (T s ) for <strong>the</strong> valley is:<br />

143<br />

T s = VsQs -1 t,<br />

Sheet1<br />

Terrace h min h max area n units age min n cores<br />

Qt1 2.2 3.3 97 2 83 1<br />

Qt2 1.2 2.1 908 8 134 6<br />

Qt3 0.9 1.7 2751 33 158 18<br />

Qt4 0.4 1.2 3043 92 162 33<br />

Qt5 0.1 0.7 1465 164 120 7<br />

Table 1. Characterization <strong>of</strong> Gordon Gulch stream terraces,<br />

with Qt1 being <strong>the</strong> oldest and Qt5 being <strong>the</strong> current<br />

floodplain. h min and h max are minimum and maximum<br />

heights <strong>of</strong> terraces above <strong>the</strong> stream channel in meters,<br />

area is total area <strong>of</strong> all units in square meters, n units is number<br />

<strong>of</strong> units mapped for each terrace, age min is minimum<br />

age obtained from tree coring in years, n cores is number <strong>of</strong><br />

tree cores obtained for each terrace.<br />

where t is a unitless time interval, defined as:<br />

t = [total years <strong>of</strong> Q w data] / [total years <strong>of</strong> Q w<br />

exceeding 90th percentile].<br />

This calculation assumes stream flow patterns remain<br />

constant over thousand year timescales.<br />

Figure 3. Stream terrace map near location KW-ST-05 (red dot). Terraces range from Qt1 (oldest) to Qt5 (youngest). Alluvial<br />

fan units (Qfa) are visible. Tor deposits (Qt) are not visible.<br />

Page 1


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

RESULTS<br />

Sheet1<br />

Sample ID Dist D 84 S H R B τ θ Q s<br />

KW-ST-04 0.00 0.002 0.08 0.09 0.075 0.9 59 1.80 710<br />

KW-ST-06 0.19 0.020 0.08 0.10 0.081 0.8 63 0.20 360<br />

KW-ST-05 0.386 0.002 0.08 0.20 0.100 0.4 79 2.40 500<br />

KW-ST-01 0.64 0.020 0.09 0.27 0.096 0.3 85 0.26 260<br />

KW-ST-07 0.771 0.010 0.09 0.09 0.075 0.9 66 0.41 650<br />

KW-ST-08 0.993 0.020 0.09 0.17 0.098 0.5 87 0.27 450<br />

KW-ST-09 1.266 0.020 0.11 0.27 0.096 0.3 100 0.32 390<br />

KW-ST-10 1.414 0.020 0.10 0.07 0.061 1.2 60 0.18 470<br />

Median 0.29 460<br />

Table 2. Shields stress and sediment flux for Gordon<br />

Gulch. Dist. is distance upstream from <strong>the</strong> beginning <strong>of</strong><br />

<strong>the</strong> mapped section in meters, D 84 is grain size in meters,<br />

S is decimal slope, H is bank-full depth in meters, R is<br />

hydraulic radius in meters, B is bank-full width in meters,<br />

τ is shear stress in Newtons per square meter, θ is Shields<br />

stress (unitless), Q s is modeled sediment flux in cubic meters<br />

per 12 year cycle.<br />

Terraces along 1.6 km <strong>of</strong> Lower Gordon Gulch were<br />

characterized into five distinct levels, which are listed<br />

in Table 1. Terrace Qt5 is <strong>the</strong> current floodplain and<br />

was vegetated by mostly grasses and young plants.<br />

There was no discernible difference in vegetation on<br />

terraces Qt1 through Qt4. Figure 3 shows a section <strong>of</strong><br />

Lower Gordon Gulch in which all five terrace levels<br />

interact with alluvial fans (Qa).<br />

Page 1<br />

Morphology <strong>of</strong> terraces in Gordon Gulch varies along<br />

<strong>the</strong> stream. Downstream, <strong>the</strong>re are more terraces<br />

flanking <strong>the</strong> stream in complex patterns. The majority<br />

<strong>of</strong> terraces are not paired. The north bank <strong>of</strong> <strong>the</strong><br />

stream <strong>of</strong>ten contains few or no terraces. Terraces on<br />

<strong>the</strong> south bank are more extensive. In some locations<br />

(Fig. 3) it is possible to find all five terraces in<br />

one location. Upstream <strong>the</strong>re may be only one or two<br />

terraces flanking <strong>the</strong> stream (Fig. 2A). No bedrock is<br />

visible in <strong>the</strong> mapped stream channel. Tor deposits<br />

are more common upstream. The overall width <strong>of</strong><br />

upstream terraces is half that <strong>of</strong> downstream terraces.<br />

The total volume <strong>of</strong> sediment stored in <strong>the</strong> terraces <strong>of</strong><br />

lower Gordon Gulch (V s ) was calculated to be 50,000<br />

m 3 in <strong>the</strong> mapped 1.6 km <strong>of</strong> <strong>the</strong> stream (Table 2).<br />

The time interval t was calculated using discharge<br />

data from Boulder Creek over <strong>the</strong> past 24 years provided<br />

by <strong>the</strong> US Geological Survey. Of <strong>the</strong> 24 years<br />

<strong>of</strong> data, two years <strong>of</strong> maximum discharge values exceeded<br />

<strong>the</strong> 90th percentile <strong>of</strong> <strong>the</strong> Boulder Creek data.<br />

144<br />

Thus, <strong>the</strong> time interval between maximum discharge<br />

events in <strong>the</strong> catchment is 12 years.<br />

Parameters for <strong>the</strong> calculation <strong>of</strong> Q s are listed in<br />

Table 2. For <strong>the</strong> 90th percentile discharge (3,500 m 3 /<br />

day from Boulder Creek CZO stream gauge data),<br />

<strong>the</strong> median θ value was 0.29, more than sufficient to<br />

mobilize terrace sediment. The θ values were used to<br />

calculate sediment flux at each sample location, with<br />

<strong>the</strong> median sediment discharge value being 460 m 3<br />

<strong>of</strong> sediment transported from Gordon Gulch every 12<br />

years. Median values were used to avoid sensitivity<br />

to outliers.<br />

At <strong>the</strong> current rates <strong>of</strong> water and sediment discharge,<br />

this model estimates that it would take 1,300 years to<br />

evacuate <strong>the</strong> sediment currently in <strong>the</strong> basin.<br />

Two radiometric 14 C dates were obtained from buried<br />

wood in terrace sediments. The first sample was 30<br />

cm above <strong>the</strong> current stream channel and was dated<br />

1,110 ± 50 years before present. The second sample<br />

was 10 cm above <strong>the</strong> current stream channel and was<br />

dated 1,520 ± 40 years before present.<br />

DISCUSSION<br />

As streams go through periods <strong>of</strong> aggradation and<br />

degradation a complex system <strong>of</strong> terraces may form.<br />

In Gordon Gulch, five terraces have formed from this<br />

process.<br />

Terrace morphology <strong>of</strong> Gordon Gulch<br />

Variations in terrace morphology along Gordon Gulch<br />

can be attributed to valley morphology. Water downstream<br />

carries more sediment, as drainage area is directly<br />

related to distance from <strong>the</strong> headwaters. Thus,<br />

more sediment is carried into <strong>the</strong> stream by erosional<br />

processes. Increased sediment is counteracted by<br />

decreased slope <strong>of</strong> <strong>the</strong> stream. The combined effect<br />

<strong>of</strong> <strong>the</strong>se factors is that a larger amount <strong>of</strong> sediment<br />

accumulated in downstream terraces versus upstream<br />

terraces. Terrace sediments in Qt1 through Qt4 were<br />

accumulated in <strong>the</strong> past 2,000 years and are currently<br />

being incised into. Qt5 may be <strong>the</strong> result <strong>of</strong> a combination<br />

<strong>of</strong> current accumulation and incision into past<br />

accumulation.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Ages obtained from tree coring are largely varied and<br />

do not accurately reflect terrace ages. This may be<br />

<strong>the</strong> result <strong>of</strong> logging and forest fires that cleared many<br />

<strong>of</strong> <strong>the</strong> trees in <strong>the</strong> past 200 years. The oldest tree was<br />

a 162 year old Ponderosa Pine on Qt4. Thus, terraces<br />

Qt1 through Qt4 stabilized at least 162 years ago.<br />

Sediment removal from Gordon Gulch<br />

Shields stress (θ) values for maximum stream discharge<br />

in Gordon Gulch have a large variation. Maximum<br />

stream discharge is more than sufficient (θ ><br />

0.07) in all sample locations to transport all sediment<br />

in <strong>the</strong> stream. In two locations (KW-ST-04 and KW-<br />

ST-05), <strong>the</strong> θ value for maximum stream discharge is<br />

very high (θ > 1.5) due to decreased grain size. These<br />

locations also have a high Q s value. In locations<br />

with a large D 84 grain size (D 84 ≥ 20 mm), θ values<br />

were below 0.40 and Q s values were below or near<br />

<strong>the</strong> median Q s value. Increased slope also resulted in<br />

increased θ and Q s values. Grain size appears to have<br />

<strong>the</strong> largest control on θ and Q s values<br />

The sediment removal timescale for terraces in Gordon<br />

Gulch calculated by this model is 1,300 years.<br />

Evacuating <strong>the</strong> sediment in this timescale would be<br />

unlikely. The model does not take into account forces<br />

holding sediments toge<strong>the</strong>r, which include roots, buried<br />

logs and o<strong>the</strong>r biologic factors, as well as compaction<br />

forces <strong>of</strong> buried sediments. The model also does<br />

not account for sediment currently being added to <strong>the</strong><br />

stream by erosion on hillslopes and from addition <strong>of</strong><br />

sediment upstream <strong>of</strong> <strong>the</strong> mapped area. Incorporating<br />

<strong>the</strong>se factors into <strong>the</strong> model would likely increase <strong>the</strong><br />

sediment removal timescale.<br />

CONCLUSION<br />

The Gordon Gulch terrace system includes five complex<br />

terrace levels that are closely related to valley<br />

morphology. Downstream terraces are wider and<br />

more complex due to aggradation from increased<br />

sediment concentration and decreased slope. Sediment<br />

stored in terraces has been accumulating for<br />

over 2,000 years. Total volume <strong>of</strong> sediment stored<br />

in <strong>the</strong> terraces was approximated to be 50,000 cubic<br />

meters. Hydrologic models applied to calculate sedi-<br />

145<br />

ment flux estimate that it would take 1,300 years to<br />

evacuate terrace sediment from Gordon Gulch. This<br />

value underestimates <strong>the</strong> time it will take to remove<br />

sediment stored in <strong>the</strong> terraces, largely because <strong>the</strong><br />

model does not take into account biologic factors and<br />

erosional input from <strong>the</strong> headwaters and hillslopes.<br />

Future research should focus on quantifying inputs<br />

<strong>of</strong> sediment into <strong>the</strong> stream by erosion on hillslopes<br />

and upstream <strong>of</strong> <strong>the</strong> mapped area. Incorporating<br />

<strong>the</strong>se factors into <strong>the</strong> model would provide a closer<br />

approximation <strong>of</strong> <strong>the</strong> sediment removal timescale.<br />

Future research should also quantify <strong>the</strong> effects <strong>of</strong><br />

biologic factors and compaction on erosion <strong>of</strong> terrace<br />

sediments. Understanding <strong>the</strong>se factors would also<br />

provide better understanding <strong>of</strong> how <strong>the</strong> complex relationships<br />

<strong>of</strong> <strong>the</strong> CZ affect sediment flux. Volume <strong>of</strong><br />

sediment should be better estimated using geophysical<br />

methods (ground penetrating radar) to measure <strong>the</strong><br />

depth to bedrock below <strong>the</strong> terraces.<br />

ACKNOWLEDGMENTS<br />

I thank David Dethier, Will Ouimet, Kurt Frankel,<br />

Corey Shircliff, Erin Camp, Reece Lyerly, Cianna<br />

Wyshnytzky, Hayley Corson-Rikert, and Ellie Maley<br />

for <strong>the</strong>ir help.<br />

REFERENCES<br />

Anderson, S. P., F. von Blanckenburg, and A. F. White<br />

(2007), Physical and chemical controls on <strong>the</strong> Critical<br />

Zone, Elements, 3(5), 315-319.<br />

Cronin, G., J. H. McCutchan, J. Pitlick, and W. M.<br />

Lewis (2007), Use <strong>of</strong> Shields stress to reconstruct and<br />

forecast changes in river metabolism, Freshw. Biol.,<br />

52(8), 1587-1601.<br />

Goldblum, D., and T. T. Veblen (1992), Fire history <strong>of</strong><br />

a Ponderosa pine Douglas-fir forest in <strong>the</strong> Colorado<br />

Front Range, Physical Geography, 13(2), 133-148.<br />

Mueller, E. R., and J. Pitlick (2005), Morphologically<br />

based model <strong>of</strong> bed load transport capacity in a headwater<br />

stream, Journal <strong>of</strong> Geophysical Research-Earth<br />

Surface, 110(F2).


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

METEORIC 10 BE IN GORDON GULCH SOILS: IMPLICATIONS<br />

FOR HILLSLOPE PROCESSES AND DEVELOPMENT<br />

CIANNA E. WYSHNYTZKY, Amherst College<br />

Research Advisors: Will Ouimet and Peter Crowley<br />

INTRODUCTION<br />

Is some agent <strong>of</strong> tectonism necessary to explain <strong>the</strong><br />

dramatic relief <strong>of</strong> <strong>the</strong> Front Range or are isostatic<br />

responses to erosion induced by climate change sufficient?<br />

Could a post-Eocene change in climate and<br />

integration <strong>of</strong> channel systems alone allow for increased<br />

incision to dominate this landscape and create<br />

<strong>the</strong> high relief <strong>of</strong> deeply incised canyons and large<br />

stream valleys, or is some tectonism necessary to account<br />

for <strong>the</strong> observed amount <strong>of</strong> warping, tilting, and<br />

erosion? Understanding <strong>the</strong> evolution <strong>of</strong> Front Range<br />

hillslopes in relation to late Cenozoic climatic and<br />

tectonic evolution, hillslope processes, and fundamental<br />

critical zone principles could provide a more<br />

thorough understanding <strong>of</strong> <strong>the</strong> modern geomorphology<br />

<strong>of</strong> <strong>the</strong> region.<br />

This research used <strong>the</strong> accumulation <strong>of</strong> meteoric 10 Be<br />

to determine <strong>the</strong> age <strong>of</strong> soils on hillslopes in Gordon<br />

Gulch and helps constrain interpretations <strong>of</strong> regolith<br />

transport, extending current information about <strong>the</strong><br />

recent evolution <strong>of</strong> Colorado’s Front Range. This<br />

research contributes to research done by <strong>the</strong> Boulder<br />

Creek Critical Zone Observatory (BC-CZO) within<br />

<strong>the</strong> extents <strong>of</strong> <strong>the</strong>ir focus area in <strong>the</strong> Front Range and<br />

complements ongoing research in Gordon Gulch.<br />

This is <strong>the</strong> first project in <strong>the</strong> region using LiDAR<br />

analysis and meteoric 10 Be as a tracer <strong>of</strong> modern hillslope<br />

evolution.<br />

Due to its adherence to sediment within soils and its<br />

constant rate <strong>of</strong> production in <strong>the</strong> atmosphere, soil<br />

age can be constrained by using <strong>the</strong> inventory <strong>of</strong> total<br />

meteoric 10 Be <strong>of</strong> a soil pr<strong>of</strong>ile. Erosion rates and<br />

estimated soil transport rates can <strong>the</strong>n be quantified<br />

(Jungers et al., 2009; Graly et al., 2010; Willenbring<br />

and von Blackenburg, 2010). Studies in North Carolina<br />

(Jungers, et al., 2009) and Australia (Fifield, et al.,<br />

2010) have traced hillslope sediment production and<br />

146<br />

transport using meteoric 10 Be, after which <strong>the</strong> <strong>the</strong>oretical<br />

framework <strong>of</strong> this research has been modeled.<br />

Graly et al. (2010) have shown that <strong>the</strong> concentration<br />

<strong>of</strong> meteoric 10 Be in soil pr<strong>of</strong>iles typically conforms<br />

to one <strong>of</strong> three general pr<strong>of</strong>ile shapes: exponentially<br />

declining, bulge, and bulge/declining (small bulge<br />

towards <strong>the</strong> top <strong>of</strong> <strong>the</strong> pr<strong>of</strong>ile). As a soil pr<strong>of</strong>ile<br />

evolves, so does its meteoric 10 Be inventory due to<br />

soil formation and mixing processes. Given a steady<br />

state hillslope, <strong>the</strong> peak concentration <strong>of</strong> meteoric<br />

10 Be is expected in one horizon (Jungers et al., 2009).<br />

Concentration <strong>the</strong>n decreases with depth, and <strong>the</strong><br />

inventory is expected to increase downslope, creating<br />

a bulge pr<strong>of</strong>ile. Given a young and eroding hillslope<br />

pr<strong>of</strong>ile, <strong>the</strong> highest concentration <strong>of</strong> meteoric 10 Be will<br />

still be in a single layer, but erosion prevents this concentration<br />

from moving to depths beyond near-surface<br />

(Graly et al., 2010).<br />

GEOGRAPHIC AND GEOLOGIC SETTING<br />

Gordon Gulch is a focus area <strong>of</strong> <strong>the</strong> BC-CZO located<br />

below and east <strong>of</strong> <strong>the</strong> modern alpine environment and<br />

late Pleistocene glacial limit and generally above and<br />

west <strong>of</strong> <strong>the</strong> deeply incised landscape that characterizes<br />

<strong>the</strong> lower portion <strong>of</strong> Front Range rivers. The<br />

degree to which <strong>the</strong> drainage basin may be affected<br />

by upstream-migrating rejuvenation from <strong>the</strong> lower<br />

portion <strong>of</strong> <strong>the</strong> range and/or alpine environmental processes<br />

(i.e. periglacial activity) is debated (Anderson<br />

et al., 2006). Gordon Gulch is a 2.75 km 2 catchment<br />

with exposed bedrock in various places. It can be<br />

subdivided into two primary floral and spatial environments:<br />

<strong>the</strong> north- and south-facing hillslopes (Fig.<br />

1).<br />

METHODS<br />

Sample Collection and Transect Selection<br />

Hillslope transects were chosen using a combination


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Figure 1. Gordon Gulch. a) air photo (1m B&W DOQ,<br />

1999), b) shaded relief map, c) slope map. (b and c derived<br />

from ~1m resolution LiDAR data). Blue dots indicates<br />

sample pit locations.<br />

<strong>of</strong> topographic pr<strong>of</strong>ile and field observations <strong>of</strong> <strong>the</strong><br />

character <strong>of</strong> hillslopes at proposed soil pit locations.<br />

Areas containing evidence <strong>of</strong> recent fire were avoided<br />

to not incorporate fire as an agent <strong>of</strong> erosion, and<br />

gullies, bedrock outcrops, trails/roads, and miners’<br />

pits were avoided to obtain <strong>the</strong> most representative<br />

and continuous downslope path from ridge to stream.<br />

These criteria affected transect selection so that <strong>the</strong><br />

smoo<strong>the</strong>st pr<strong>of</strong>ile on each hillslope was found to act<br />

as a proxy for overall hillslopes <strong>of</strong> <strong>the</strong> area. Samples<br />

from 9 hillslope pits (5 on <strong>the</strong> north-facing hillslope,<br />

4 on <strong>the</strong> south-facing hillslope) were collected in 10<br />

147<br />

cm intervals from 1-2 cm thick slots, beginning below<br />

<strong>the</strong> O-horizon (Fig. 1). Each soil pr<strong>of</strong>ile was photographed<br />

and individual horizons were described in<br />

detail.<br />

Meteoric 10 Be<br />

Samples were dried to remove excess moisture and<br />

hand sieved through a wire-mesh 2 mm sieve to<br />

remove coarse particles, since meteoric 10 Be binds<br />

to particles with high surface area (Fifield, 2010).<br />

Sieved samples were ground into fine powders using<br />

a tungsten carbide shatterbox triplet.<br />

Beryllium was extracted from samples at <strong>the</strong> Cosmogenic<br />

Nuclide Laboratory at <strong>the</strong> University <strong>of</strong><br />

Vermont from ~0.5 g aliquots by ion exchange acid<br />

elutions in a method adapted from <strong>the</strong> flux fusion<br />

methods originally presented by Stone (1998). Atoms<br />

<strong>of</strong> 10 Be were counted using an accelerator mass<br />

spectrometer (AMS) by <strong>the</strong> GeoCAMS group at <strong>the</strong><br />

Center for Accelerator Mass Spectromery (CAMS) at<br />

<strong>the</strong> Livermore National Laboratory in California.<br />

In addition to <strong>the</strong> analysis <strong>of</strong> 10 Be on hillslopes <strong>of</strong><br />

unknown age, 10 Be analysis <strong>of</strong> samples from stable<br />

reference sites (with known ages) were run to quantify<br />

how much 10 Be has been delivered to <strong>the</strong> region by<br />

precipitation and dustfall during <strong>the</strong> past ~15,000 to<br />

25,000 years. The analysis <strong>of</strong> meteoric 10 Be in <strong>the</strong>se<br />

stable reference sites also improves <strong>the</strong> understanding<br />

<strong>of</strong> correlations between o<strong>the</strong>r methods <strong>of</strong> dating,<br />

such as optically stimulated luminescence (OSL), 14 C<br />

<strong>of</strong> charcoal layers, and in situ 10 Be, something which<br />

could expand this method <strong>of</strong> dating throughout <strong>the</strong><br />

geologic research community. The reference sites are<br />

a known Pinedale moraine (~15 ka) at Silver Lake<br />

and a soil pit in Upper Gordon Gulch recently dated<br />

using OSL (~26 ka) (Völkel et al., 2010).<br />

Digital Elevation Map (DEM) Analysis<br />

Snow-<strong>of</strong>f Light Detection and Ranging (LIDAR) data<br />

(National Center for Airborne Laser Mapping, August<br />

2010) were used to produce digital elevation maps<br />

(DEM) for use in ArcMap. Pr<strong>of</strong>iles <strong>of</strong> <strong>the</strong> north- and<br />

south-facing hillslopes <strong>of</strong> Gordon Gulch were made<br />

in ArcMap using Spatial Analyst to create pr<strong>of</strong>iles


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

(Fig. 2). For each hillslope, <strong>the</strong> transect with pits<br />

was <strong>the</strong> original pr<strong>of</strong>ile produced. Transects parallel<br />

to <strong>the</strong>se were spaced approximately 120 m apart and<br />

extended to <strong>the</strong> borders <strong>of</strong> <strong>the</strong> drainage basin. All<br />

transects began in <strong>the</strong> channel and ended at <strong>the</strong> ridge<br />

and were normal to <strong>the</strong> channel. Spatial Analyst was<br />

also used to produce digital elevation, hillshade, and<br />

slope maps using LiDAR data. These two maps and<br />

<strong>the</strong> additional air photo provide three different visual<br />

190<br />

180 representations <strong>of</strong> Gordon Gulch drainage basin as a<br />

170<br />

160 whole (Fig. 2).<br />

Distance (m)<br />

150<br />

140<br />

130<br />

120<br />

110 a. North Facing Hillslope<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

F<br />

0<br />

180<br />

160<br />

140<br />

120<br />

100<br />

40<br />

20<br />

190<br />

180<br />

170<br />

160<br />

150<br />

140<br />

130<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0<br />

0<br />

D<br />

E<br />

G<br />

Figure 2. Pr<strong>of</strong>iles <strong>of</strong> <strong>the</strong> north- and south-facing hillslopes<br />

<strong>of</strong> Gordon Gulch extracted from ~1m resolution LiDAR<br />

digital elevation model. Pr<strong>of</strong>iles are labeled from west<br />

to east (downstream) beginning with “A” or “a”. Transects<br />

in bold are <strong>the</strong> focus transects <strong>of</strong> this study. Open<br />

circles indicate locations <strong>of</strong> pits sampled for meteoric 10 Be.<br />

Shaded relief map (see Fig 1. for legend) shows approximate<br />

location <strong>of</strong> transects with black lines. Dots indicate<br />

sample pit locations. Dashed line indicates approximate<br />

boundary between <strong>the</strong> upper and lower basin.<br />

RESULTS<br />

Map Analysis<br />

NGG-05<br />

NGG-04<br />

M<br />

O<br />

N R<br />

P<br />

NGG-03<br />

100 200 300 400 500<br />

NGG-02<br />

NGG-01<br />

600 700 800 900 1000<br />

Q S<br />

L<br />

K<br />

K<br />

J<br />

I<br />

H<br />

L<br />

M<br />

A B<br />

E<br />

D<br />

G<br />

F<br />

H<br />

J<br />

I<br />

C<br />

N Distance (m)<br />

S<br />

q pr<br />

s<br />

80<br />

m<br />

l<br />

g<br />

60 f<br />

t<br />

o u<br />

v<br />

n k w j<br />

e<br />

h<br />

100<br />

SGG-03<br />

i<br />

200<br />

SGG-02<br />

d<br />

300<br />

b. South Facing Hillslope<br />

c<br />

b<br />

SGG-01<br />

400<br />

a<br />

SGG-00<br />

500<br />

The air photo clearly depicts differences in present<br />

1000100 200 300 400 500 600<br />

600<br />

A BC<br />

DEF<br />

700<br />

G<br />

e<br />

f g<br />

a b c d<br />

h<br />

i<br />

j<br />

k<br />

H<br />

I<br />

J<br />

K<br />

L M N<br />

800<br />

900<br />

l<br />

V<br />

U W<br />

T<br />

m<br />

n op q r s t u v w x<br />

O P<br />

Q<br />

R<br />

S T<br />

U V<br />

W<br />

148<br />

vegetation on <strong>the</strong> north- and south-facing hillslopes <strong>of</strong><br />

lower Gordon Gulch. Vegetation is less dense on <strong>the</strong><br />

south-facing hillslope, whereas little except vegetation<br />

is visible on <strong>the</strong> north-facing hillslope.<br />

With <strong>the</strong> addition <strong>of</strong> a hillshade layer, <strong>the</strong> DEM depicts<br />

both elevation and topographic features. Bedrock<br />

outcrops are seen extruding 1-10 m above <strong>the</strong><br />

dominant soil mantled hillslopes. These bedrock outcrops<br />

are characterized by shallow slope upslope and<br />

close to vertical slope on <strong>the</strong> downslope sides. The<br />

density <strong>of</strong> bedrock outcrops between <strong>the</strong> north- and<br />

south-facing hillslopes appears similar in upper Gordon<br />

Gulch. However, <strong>the</strong>re is a disparity <strong>of</strong> bedrock<br />

outcrop density between <strong>the</strong> north- and south-facing<br />

hillslopes <strong>of</strong> lower Gordon Gulch (Trotta, 2010).<br />

More bedrock outcrops are present on <strong>the</strong> southfacing<br />

hillslope, and <strong>the</strong>se are larger than those on <strong>the</strong><br />

north-facing hillslope. The north-facing hillslope is<br />

dissected by numerous gullies, whereas few gullies<br />

are visible on <strong>the</strong> south-facing hillslope.<br />

The slope map shows that many <strong>of</strong> <strong>the</strong> north-facing<br />

hillslopes have a parabolic shape in which <strong>the</strong> highest<br />

slopes are in <strong>the</strong> middle-range <strong>of</strong> <strong>the</strong> transect and <strong>the</strong><br />

lowest slopes at <strong>the</strong> bottom <strong>of</strong> <strong>the</strong> gulch and <strong>the</strong> drainage<br />

divide. In contrast, <strong>the</strong> slope is fairly constant<br />

on <strong>the</strong> south-facing hillslope <strong>of</strong> lower Gordon Gulch,<br />

except where <strong>the</strong> slope increases in <strong>the</strong> downslope direction<br />

<strong>of</strong> bedrock outcrops and as <strong>the</strong> slope shallows<br />

near <strong>the</strong> ridge <strong>of</strong> <strong>the</strong> drainage basin.<br />

Hillslope Pr<strong>of</strong>iles<br />

The north-facing hillslope <strong>of</strong> Gordon Gulch is shorter<br />

and shallower (Fig. 2a) than <strong>the</strong> south-facing hillslope,<br />

which is longer and steeper overall (Fig. 2b).<br />

Using data drawn from ArcMap hillslope pr<strong>of</strong>iles<br />

(Fig. 2) beginning at <strong>the</strong> first stream encountered<br />

and ending at <strong>the</strong> drainage divide, <strong>the</strong> average slope<br />

<strong>of</strong> all north-facing hillslopes is 9.6º and <strong>of</strong> all southfacing<br />

hillslopes is 12.2º. Taking <strong>the</strong> average from<br />

<strong>the</strong> sampled pit transect and <strong>the</strong> two transects parallel<br />

on each side yields slopes <strong>of</strong> 15.0º on <strong>the</strong> north-facing<br />

hillslope and 19.6º on <strong>the</strong> south-facing hillslope. The<br />

steepest local (and non-bedrock outcrop) slopes are<br />

found in <strong>the</strong> lower half <strong>of</strong> <strong>the</strong> north-facing hillslope,<br />

as seen in <strong>the</strong> slope map, but not reflected in pr<strong>of</strong>ile


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

data due to pr<strong>of</strong>ile length averages. In lower Gordon<br />

Gulch, <strong>the</strong> north-facing hillslope pr<strong>of</strong>iles are characterized<br />

by a ridge creating a break in slope followed<br />

by a second slope ending at <strong>the</strong> drainage basin boundary.<br />

The first slope (before <strong>the</strong> ridge break) is parabolic<br />

in nature. Bedrock outcrops are found almost<br />

solely at <strong>the</strong> top <strong>of</strong> <strong>the</strong> pr<strong>of</strong>ile, but above <strong>the</strong> break in<br />

slope. In contrast, <strong>the</strong> south-facing hillslope pr<strong>of</strong>iles<br />

are more linear than parabolic and have a more consistent<br />

slope from ridge to stream, except where broken<br />

by bedrock outcrops. Bedrock outcrops are more<br />

abundant on this slope than <strong>the</strong> north-facing hillslope,<br />

and are seen as <strong>the</strong> many small spikes on <strong>the</strong> pr<strong>of</strong>iles.<br />

Meteoric 10 Be<br />

Lower Gordon Gulch<br />

North-facing hillslope soils have 10 Be concentrations<br />

<strong>of</strong> 5-8 x 10 8 atoms/g in <strong>the</strong> uppermost samples and<br />

decline to < 2 x 10 8 atoms/g at depths >60 cm (Table<br />

1). Pits NGG-01 and NGG-05 show exponentially<br />

declining pr<strong>of</strong>iles with some variability and small<br />

bulges, corresponding to <strong>the</strong> highest concentration<br />

<strong>of</strong> 10 Be in <strong>the</strong> B-horizon and <strong>the</strong> Cox-horizon respectively.<br />

Pits NGG-02, NGG-03, and NGG-04 show<br />

exponentially declining pr<strong>of</strong>iles (with some variability)<br />

with peak meteoric 10 Be concentrations towards<br />

<strong>the</strong> top <strong>of</strong> <strong>the</strong> pr<strong>of</strong>iles. The lowest concentration for<br />

each pit is at depth, except for NGG-01 in which <strong>the</strong><br />

lowest concentration is in <strong>the</strong> Cox2-horizon (Fig. 3a).<br />

Pits NGG-01 and NGG-02 decline to a steady concentration<br />

at depth, whereas <strong>the</strong> o<strong>the</strong>r three pits on<br />

this hillslope show no clear indication <strong>of</strong> approaching<br />

a steady concentration.<br />

South-facing hillslope soils have 10 Be concentrations<br />

<strong>of</strong> 2.5-5 x 10 8 atoms/g near <strong>the</strong> surface and decline to<br />

< 2.7 x 10 8 atoms/g at depths >32 cm (Table 1). Pits<br />

SGG-00 and SGG-02 show approximately linear declining<br />

pr<strong>of</strong>iles. Pit SGG-02 also shows a bulge, corresponding<br />

to <strong>the</strong> highest concentration <strong>of</strong> 10 Be in <strong>the</strong><br />

SGG-02 A/Cox1-horizon border. Pit SGG-03 shows<br />

an exponentially declining pr<strong>of</strong>ile with a small bulge,<br />

corresponding to <strong>the</strong> highest concentration <strong>of</strong> 10 Be in<br />

<strong>the</strong> Cox1-horizon. Pit SGG-01 shows no clear declining<br />

trend with a bulge corresponding to <strong>the</strong> highest<br />

concentrations <strong>of</strong> 10 Be in <strong>the</strong> SGG-01A-horizon. The<br />

lowest concentration for each pit is at depth (Fig. 3b).<br />

149<br />

Pits SGG-01 and SGG-03 appear to be approaching a<br />

steady concentration at depth, whereas pits SGG-00<br />

and SGG-02 show no clear indication <strong>of</strong> approaching<br />

a steady concentration.<br />

Sample<br />

10Be Bulk Density Thickness 10Be inventory Total Pit Inventory Soil Age<br />

(atoms/g) (g/cm^3) (cm) (atoms/cm^2) (atoms/cm^2) (years)<br />

NGG-01-8 8.06E+08 1.2 5 3.69E+09<br />

NGG-01-18 8.96E+08 1.4 10 9.77E+09<br />

NGG-01-28 7.54E+08 1.4 25 1.95E+10<br />

NGG-01-58 2.27E+08 1.6 35 2.26E+09<br />

NGG-01-98 2.02E+08 1.6 40 1.06E+09<br />

NGG-01-138 1.47E+08 1.6 45 -2.73E+09<br />

NGG-01-188 2.06E+08 1.6 50 1.50E+10<br />

NGG-02-5 5.73E+08 1.6 10 8.09E+09<br />

NGG-02-25 2.74E+08 1.6 25 7.97E+09<br />

NGG-02-55 1.64E+08 1.6 35 4.88E+09<br />

NGG-02-95 1.17E+08 1.6 55 3.41E+09<br />

NGG-02-165 4.07E+07 1.6 70 -4.38E+09<br />

NGG-03-10 5.92E+08 1.5 5 3.42E+09<br />

NGG-03-20 3.66E+08 1.5 15 5.23E+09<br />

NGG-03-40 2.61E+08 1.5 20 3.85E+09<br />

NGG-03-60 1.85E+08 1.5 20 1.58E+09<br />

NGG-04-5 5.48E+08 1.4 5 2.87E+09<br />

NGG-04-15 4.29E+08 1.4 10 4.10E+09<br />

NGG-04-25 4.50E+08 1.6 10 5.09E+09<br />

NGG-04-35 2.49E+08 1.6 15 2.81E+09<br />

NGG-04-55 1.80E+08 1.6 20 1.53E+09<br />

NGG-05-0 5.56E+08 1.4 5 3.00E+09<br />

NGG-05-10 6.62E+08 1.5 10 7.95E+09<br />

NGG-05-20 4.41E+08 1.5 10 4.63E+09<br />

NGG-05-30 5.31E+08 1.6 15 9.57E+09<br />

NGG-05-50 1.98E+08 1.6 20 2.10E+09<br />

SGG-00-0 4.59E+08 1.2 5 1.88E+09<br />

SGG-00-10 3.99E+08 1.2 10 3.07E+09<br />

SGG-00-20 3.95E+08 1.6 10 4.21E+09<br />

SGG-00-30 3.12E+08 1.6 15 4.32E+09<br />

SGG-00-50 1.88E+08 1.6 20 1.80E+09<br />

SGG-01-0 2.85E+08 1.2 5 8.82E+08<br />

SGG-01-10 3.51E+08 1.2 10 2.52E+09<br />

SGG-01-20 4.28E+08 1.6 15 7.11E+09<br />

SGG-01-40 3.85E+08 1.6 20 8.10E+09<br />

SGG-01-60 1.29E+08 1.6 20 -1.04E+08<br />

SGG-02-2 3.51E+08 1.5 5 1.64E+09<br />

SGG-02-12 3.69E+08 1.5 10 3.55E+09<br />

SGG-02-22 2.79E+08 1.5 10 2.20E+09<br />

SGG-02-32 2.67E+08 1.5 10 2.02E+09<br />

SGG-03-0 2.93E+08 1.3 5 1.05E+09<br />

SGG-03-10 3.56E+08 1.5 10 3.38E+09<br />

SGG-03-20 2.96E+08 1.5 15 3.72E+09<br />

SGG-03-40 2.08E+08 1.5 25 2.88E+09<br />

Table 1. 10 Be concentration data for samples chosen for<br />

this analysis. Preliminary 10 Be inventory and soil ages<br />

were calculated using equations presented in <strong>the</strong> text.<br />

Silver Lake Moraine<br />

3.53E+10 32,300<br />

2.00E+10 18,300<br />

1.41E+10 12,800<br />

1.64E+10 15,000<br />

2.73E+10 24,900<br />

1.53E+10 13,900<br />

1.85E+10 16,900<br />

9.42E+09 8,580<br />

1.19E+10 10,900<br />

The Silver Lake moraine section shows a bulge pr<strong>of</strong>ile<br />

with <strong>the</strong> highest concentration in <strong>the</strong> Bw-horizon<br />

(Fig. 4a). In <strong>the</strong> Bw-horizon at 25 cm and 35 cm<br />

deep, 10 Be concentrations (1.00 x 109 and 9.34 x 108<br />

atoms/g) are greater than twice <strong>the</strong> next highest concentration<br />

(4.35 x 10 8 atoms/g) in <strong>the</strong> Ej-horizon at 15<br />

cm deep (Tab. 1). Concentrations decrease to as low<br />

as 2.28 x 10 7 atoms/g in <strong>the</strong> Cu-horizon and appear to<br />

be approaching a steady concentration.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Depth (cm)<br />

0.00E+00 !(!!)*!!" 1.00E+08 '(!!)*!&" 2.00E+08 #(!!)*!&" 3.00E+08 +(!!)*!&" 4.00E+08 $(!!)*!&" 5.00E+08 ,(!!)*!&" 6.00E+08 %(!!)*!&" 7.00E+08 -(!!)*!&" 8.00E+08 &(!!)*!&" 9.00E+08 .(!!)*!&" 1.00E+09 '(!!)*!."<br />

0<br />

0!"<br />

20<br />

#!"<br />

40<br />

$!"<br />

60<br />

%!"<br />

60<br />

80<br />

&!"<br />

100<br />

'!!"<br />

100<br />

120<br />

'#!"<br />

140<br />

140 '$!"<br />

160<br />

'%!"<br />

180<br />

180 '&!"<br />

200<br />

#!!"<br />

0<br />

40<br />

80<br />

Figure 3. Meteoric 10 100<br />

Be from soils on <strong>the</strong> north (a.) and<br />

south (b.) facing hillslopes <strong>of</strong> lower Gordon Gulch.<br />

120<br />

140<br />

160<br />

180<br />

Upper Gordon Gulch<br />

200<br />

0<br />

20<br />

60<br />

NGG-01<br />

NGG-02<br />

NGG-03<br />

NGG-04<br />

NGG-05<br />

0 100000000 200000000 300000000 400000000 500000000 600000000 700000000 800000000 900000000 1E+09<br />

SGG-00<br />

SGG-01<br />

SGG-02<br />

SGG-03<br />

Meteoric ¹⁰Be Concentration(atoms/g)<br />

2.00 × 10⁸ 4.00 × 10⁸ 6.00 × 10⁸ 8.00 × 10⁸<br />

2.00 × 10⁸ 4.00 × 10⁸<br />

b. South Facing Hillslope<br />

The soil originally sampled for OSL dating<br />

by Jörg Völkel (2010) has a maximum 10 Be concentration<br />

<strong>of</strong> 1.66 x 10 10 atoms/g near <strong>the</strong> surface. The<br />

concentration decreases to 3.13 x 10 9 atoms/g at depth<br />

(Tab. 1) and creates an exponentially declining pr<strong>of</strong>ile<br />

(Fig. 4b).<br />

Inventory and Soil Age Calculations<br />

Meteoric 10 Be inventories I Be (atoms/cm2) were calculated<br />

using an equation taken from Jungers et al.<br />

(2009):<br />

I Be = ∑(C Be -C inh )ρ s h<br />

a. North Facing Hillslope<br />

where C Be is depth-integrated 10 Be concentration, C inh<br />

is <strong>the</strong> inherited component <strong>of</strong> meteoric 10 Be , ρ s is<br />

depth-integrated soil bulk density, and h is soil thickness<br />

for each depth subsample.<br />

Inheritance for NGG-01 was assumed as <strong>the</strong> average<br />

concentration <strong>of</strong> <strong>the</strong> three deepest samples (1.85 x 10 8<br />

atoms/g) and for NGG-02 as <strong>the</strong> average <strong>of</strong> <strong>the</strong> two<br />

150<br />

Depth (cm)<br />

NGG-01<br />

/001!'"<br />

NGG-02<br />

/001!#"<br />

NGG-03<br />

/001!+" NGG-04<br />

/001!$" NGG-05<br />

/001!,"<br />

SGG-00<br />

SGG-01<br />

SGG-02<br />

SGG-03<br />

Depth (cm)<br />

0<br />

350<br />

0<br />

0<br />

50<br />

100<br />

150<br />

200<br />

250<br />

300<br />

0<br />

20<br />

40<br />

60<br />

80<br />

100<br />

Meteoric ¹⁰Be Concentration(atoms/g)<br />

2.00 × 10⁸ 4.00 × 10⁸ 6.00 × 10⁸ 8.00 × 10⁸ 1.00 × 10⁹<br />

0 200000000 400000000 600000000 800000000 1E+09 1.2E+09<br />

Meteoric ¹⁰Be Concentration(atoms/g)<br />

a. Silver Lake Moraine<br />

1.00 × 10⁸ 2.00 × 10⁸ 3.00 × 10⁸ 4.00 × 10⁸ 5.00 × 10⁸<br />

0 100000000 200000000 300000000 400000000 500000000 600000000<br />

b. Upper Gordon Gulch OSL<br />

Dated Pit<br />

Figure 4. Meteoric 10 Be concentration plots for <strong>the</strong> Silver<br />

120<br />

Lake moraine (a.) and upper Gordon Gulch pit dated with<br />

OSL (b.).<br />

deepest samples (7.87 x 10 7 atoms/g). The average <strong>of</strong><br />

<strong>the</strong> NGG-01 and NGG-02 inheritance values (1.32 x<br />

10 8 atoms/g) was used as <strong>the</strong> value for <strong>the</strong> remaining<br />

NGG pits and <strong>the</strong> SGG pits. The deepest concentration<br />

(9.21 x 10 7 atoms/g) was used for <strong>the</strong> OSL-dated<br />

pit. Inheritance has not yet been taken into account<br />

for <strong>the</strong> Silver Lake moraine, because it is assumed<br />

to be minimal. Bulk soil density was measured in<br />

<strong>the</strong> field for several sample locations and o<strong>the</strong>r bulk<br />

density values were assumed as standard values for<br />

soil and till density. The soil thickness for each depth<br />

subsample was calculated by establishing a mid-point<br />

between sample locations. The midpoint value was<br />

added to bottom samples to account for un-collected<br />

soil.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

Soil ages (t, in years) were calculated for each pit using<br />

an equation adapted from Graly et al. (2010):<br />

t = (-1/λ)ln(1-λI Be /q)<br />

where λ is <strong>the</strong> 10 Be disintegration constant (5.1 x 10 -7 ),<br />

I Be is <strong>the</strong> total inventory <strong>of</strong> 10 Be (atoms/cm 2 ), and q is<br />

<strong>the</strong> local <strong>annual</strong> meteoric 10 Be flux (atoms/cm 2 ). The<br />

value <strong>of</strong> q was estimated from Figure 4 <strong>of</strong> Graly et al.<br />

and according to an <strong>annual</strong> precipitation rate <strong>of</strong> ~45<br />

cm in Gordon Gulch (1.1 x 10 6 atoms/cm 2 ) and ~90<br />

cm at Silver Lake (1.8 x 10 6 atoms/cm 2 ), both at ~40º<br />

latitude, (Graly et al., 2010).<br />

DISCUSSION<br />

Meteoric 10 Be<br />

Lower Gordon Gulch<br />

Meteoric 10 Be data suggest that <strong>the</strong> two opposite facing<br />

hillslopes <strong>of</strong> lower Gordon Gulch are evolving<br />

differently (Fig. 3). The three soil pits in <strong>the</strong> middle<br />

<strong>of</strong> <strong>the</strong> transect on <strong>the</strong> north-facing hillslope all show<br />

declining pr<strong>of</strong>iles <strong>of</strong> meteoric 10 Be concentration, suggesting<br />

an eroding slope because upper soil has been<br />

stripped from <strong>the</strong> column and 10 Be has not had <strong>the</strong><br />

time to concentrate in a certain layer. Potential explanations<br />

for small bulges in <strong>the</strong> lowest (NGG-01) and<br />

highest (NGG-05) pr<strong>of</strong>iles include <strong>the</strong>ir locations and<br />

horizon pr<strong>of</strong>iles. Pit NGG-05 is located above <strong>the</strong><br />

major break in slope on <strong>the</strong> hillslope, and <strong>the</strong>refore<br />

represents a different environment than <strong>the</strong> four lower<br />

pits. The highest concentrations <strong>of</strong> meteoric 10 Be in<br />

pits NGG-01 and NGG-03 are in <strong>the</strong> B-horizons.<br />

Meteoric 10 Be concentration data obtained from <strong>the</strong><br />

south-facing slope differs from <strong>the</strong> concentrations<br />

from <strong>the</strong> north-facing slope. The only declining pr<strong>of</strong>ile<br />

<strong>of</strong> <strong>the</strong> four sampled pits is <strong>the</strong> lowest (SGG-00),<br />

whereas <strong>the</strong> o<strong>the</strong>r pr<strong>of</strong>iles all contain a small bulge.<br />

Preliminary analysis suggests that perhaps <strong>the</strong> upper<br />

portion <strong>of</strong> <strong>the</strong> soil column on <strong>the</strong> south-facing hillslope<br />

has been stripped and evacuated. Had this not<br />

occurred, <strong>the</strong> pr<strong>of</strong>ile shape on <strong>the</strong> south-facing hillslope<br />

may have resembled those on <strong>the</strong> north-facing<br />

hillslope with small bulges but overall decline pr<strong>of</strong>iles<br />

151<br />

and greater inventories. Therefore consistent meteoric<br />

10 Be concentration pr<strong>of</strong>ile shapes across <strong>the</strong> two<br />

opposite facing hillslopes <strong>of</strong> Gordon Gulch would<br />

have been seen.<br />

Preliminary 10 Be inventory calculations show differences<br />

between <strong>the</strong> two hillslopes (Table 1). The<br />

north-facing hillslope has a greater inventory than <strong>the</strong><br />

south-facing hillslope, particularly in <strong>the</strong> upper ~30<br />

cm <strong>of</strong> <strong>the</strong> pr<strong>of</strong>iles. Preliminary soil age calculations<br />

(Table 1) correlate to inventory differences and suggest<br />

that <strong>the</strong> south-facing hillslope <strong>of</strong> lower Gordon<br />

Gulch is eroding faster than <strong>the</strong> north-facing hillslope.<br />

Upper Gordon Gulch<br />

The 10 Be pr<strong>of</strong>ile shows a declining pr<strong>of</strong>ile (Fig. 4b),<br />

suggesting a young, eroding landscape. The preliminary<br />

soil age calculation drawn from an inventory <strong>of</strong><br />

2.43 x 10 10 atoms/cm2 suggests an age <strong>of</strong> ~22,200<br />

years (Table 1), matching <strong>the</strong> age obtained from<br />

OSL dating <strong>of</strong> <strong>the</strong> same soil pr<strong>of</strong>ile (~26,500 years)<br />

(Völkel et al., 2010).<br />

Silver Lake<br />

The 10 Be pr<strong>of</strong>ile shows a bulge pr<strong>of</strong>ile (Fig. 4a) with<br />

<strong>the</strong> highest 10 Be concentration in <strong>the</strong> Bw-horizon, fur<strong>the</strong>r<br />

emphasizing a correlation between 10 Be and Ferich<br />

B-horizons from previous studies. Preliminary<br />

soil age calculations suggest an age <strong>of</strong> ~54 ka (Table<br />

1), older than <strong>the</strong> known age <strong>of</strong> this moraine at ~15ka.<br />

Sources <strong>of</strong> Error<br />

Potential sources <strong>of</strong> error in all inventory and soil age<br />

calculations include lack <strong>of</strong> soil bulk density measurements,<br />

estimated <strong>annual</strong> 10 Be flux rates, and estimated<br />

10 Be inheritance. Inheritance also comes into question<br />

when contemplating what this may mean for hillslope<br />

evolution. Additionally, some pits may not have been<br />

dug deep enough to sample <strong>the</strong> entire meteoric 10 Be<br />

inventory, due to regolith and potential bedrock composition.


24th Annual Keck Symposium: 2011 Union College, Schenectady, NY<br />

CONCLUSION<br />

More detailed analysis <strong>of</strong> processes, soil ages, meteoric<br />

10 Be flux, and 10 Be inventories will lead to fur<strong>the</strong>r<br />

analysis and discussion <strong>of</strong> <strong>the</strong> differences between <strong>the</strong><br />

north-and south-facing hillslopes <strong>of</strong> Gordon Gulch<br />

and more precise dating <strong>of</strong> Gordon Gulch hillslopes<br />

and <strong>the</strong> Silver Lake moraine. Preliminary age calculations<br />

show that Gordon Gulch regolith is latest<br />

Pleistocene or Holocene in age or younger, not evolving<br />

throughout <strong>the</strong> Cenozoic, and that <strong>the</strong> soil flux<br />

here is rapid.<br />

ACKNOWLEDGEMENTS<br />

Many thanks to <strong>the</strong> KECK <strong>consortium</strong> and Amherst<br />

College for funding and support. Thank you Will<br />

Ouimet and David Dethier for continuing excitement<br />

and advice. Thank you to Peter Crowley for learning<br />

a completely new subset <strong>of</strong> <strong>geology</strong> to become an<br />

effective replacement advisor. Thank you to all my<br />

fellow CO-KECK students, particularly <strong>the</strong> awesome<br />

soils group. Lastly, thank you to <strong>the</strong> many people<br />

from CU-Boulder, UMass, UVM, and <strong>the</strong> GeoCAMS<br />

group at <strong>the</strong> Livermore National Laboratory who<br />

helped in many different steps <strong>of</strong> field and laboratory<br />

work.<br />

REFERENCES<br />

Anderson, R. S., Riihimaki, C. A., Safran, E. B.,<br />

MacGregor, K. R., 2006, Facing reality: Late<br />

Cenozoic evolution <strong>of</strong> smooth peaks, glacially<br />

ornamented valleys and deep river gorges <strong>of</strong><br />

Colorado’s Front Range: in Willett, S.D., Hovius,<br />

N., Brandon, M.T., and Fisher, D.M., eds.,<br />

Tectonics, climate and landscape evolution: Geological<br />

Society <strong>of</strong> America Special Paper 398, p.<br />

397-418.<br />

Fifield L. K., Wasson R. J., Pillans B., Stone J. O. H.,<br />

2010, The longevity <strong>of</strong> hillslope soil in SE and<br />

NW Australia, CATENA, Volume 81, Issue 1, p.<br />

32-42.<br />

Graly, J. A., Bierman P. R., Reusser, L. J., Pavich,<br />

M. J., 2010, Meteoric 10Be in soil pr<strong>of</strong>iles - A<br />

152<br />

global meta-analysis, Geochimica et Cosmochimica<br />

Acta, v. 74, Issue 23, p. 6814-6829.<br />

Jungers, M. C., Bierman, P. R., Matmon, A., Nichols,<br />

K., Larsen, J., and Finkel, R., 2009, Tracing<br />

hillslope sediment production and transport with<br />

in situ and meteoric 10Be, J. Geophys. Res., v.<br />

114.<br />

Stone, J., 1998, A rapid fusion method for separation<br />

<strong>of</strong> beryllium-10 from soils and silicates: Geochimica<br />

et Cosmochimica Acta, v. 62, p. 555–561.<br />

Trotta, J. R., 2010, The Distribution <strong>of</strong> Tors in Gordon<br />

Gulch, Front Range, Colorado. Thesis submitted<br />

in partial fulfillment <strong>of</strong> <strong>the</strong> requirements for<br />

<strong>the</strong> Degree <strong>of</strong> Bachelor <strong>of</strong> Arts with Honors in<br />

Geosciences, Williams College.<br />

Völkel, J., Huber, J., Leopold, M. and Dethier, D.,<br />

2010, Young Quaternary slope sediments and<br />

paleosoils in <strong>the</strong> Colorado Front Range – process<br />

and age, Geological Society <strong>of</strong> America Abstracts<br />

with Programs, 42, no. 5, p. 469.<br />

Willenbring, J. K. and von Blanckenburg, F., 2010,<br />

Meteoric cosmogenic Beryllium-10 adsorbed to<br />

river sediment and soil: Applications for Earthsurface<br />

dynamics, Earth Science Reviews v. 98,<br />

p. 105-122.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!