13.02.2013 Views

Download the entire proceedings as an Adobe PDF - Eastern Snow ...

Download the entire proceedings as an Adobe PDF - Eastern Snow ...

Download the entire proceedings as an Adobe PDF - Eastern Snow ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Proceedings of <strong>the</strong><br />

63rd ANNUAL<br />

EASTERN SNOW CONFERENCE<br />

7–9 June 2006<br />

Newark, Delaware, USA


ISBN 0-920081-28-2<br />

ISSN: 0424-1932<br />

Proceedings of <strong>the</strong> E<strong>as</strong>tern <strong>Snow</strong> Conference<br />

Printed <strong>an</strong>d Bound in <strong>the</strong> United States of America<br />

ii


FOREWORD<br />

T<br />

his <strong>proceedings</strong> volume contains papers presented at <strong>the</strong> 63rd E<strong>as</strong>tern <strong>Snow</strong> Conference (ESC) held<br />

7–9 June 2006 at <strong>the</strong> University of Delaware in Newark, Delaware, USA. Sessions covering snow–<br />

climate interactions, snow processes, ground ice <strong>an</strong>d periglacial processes, <strong>an</strong>d remote sensing of snow<br />

were convened.<br />

The ESC is a joint United States <strong>an</strong>d C<strong>an</strong>adi<strong>an</strong> forum for discussing recent work on operational, applied, <strong>an</strong>d<br />

scientific issues related to snow <strong>an</strong>d ice. It also retains <strong>an</strong> incre<strong>as</strong>ing interest <strong>as</strong> a symposium where novel<br />

approaches to cryospheric science of international signific<strong>an</strong>ce are presented. The ESC h<strong>as</strong> published <strong>an</strong> <strong>an</strong>nual<br />

series of <strong>proceedings</strong> since 1952. Typical topics include studies of snow <strong>an</strong>d ice <strong>as</strong> materials, snow removal,<br />

meteorological forec<strong>as</strong>ting, river ice control, snow hydrology, snow chemistry, glaciology, remote sensing of snow<br />

<strong>an</strong>d ice, <strong>an</strong>d snow ecology. Membership in <strong>the</strong> ESC is open to all interested individuals <strong>an</strong>d corporations. Additional<br />

copies of <strong>the</strong> current <strong>proceedings</strong> <strong>an</strong>d all back issues c<strong>an</strong> be obtained from <strong>the</strong> Secretary. More information about <strong>the</strong><br />

E<strong>as</strong>tern <strong>Snow</strong> Conference may be found at http://www.e<strong>as</strong>ternsnow.org/.<br />

We continue with <strong>an</strong> optional review process <strong>an</strong>d authors may submit extended abstracts or full papers to <strong>the</strong><br />

<strong>proceedings</strong>. Particip<strong>an</strong>ts in <strong>the</strong> Western <strong>Snow</strong> Conference (WSC) (http://www.westernsnowconference.org/) have<br />

been invited to join <strong>the</strong> ESC particip<strong>an</strong>ts in <strong>the</strong> option of submitting papers relev<strong>an</strong>t to winter hydrology to <strong>the</strong><br />

international journal Hydrological Processes. These papers have gone through a formal journal review, revision, <strong>an</strong>d<br />

referee process, <strong>an</strong>d m<strong>an</strong>y appear both in <strong>the</strong>se <strong>proceedings</strong> <strong>an</strong>d in <strong>the</strong> December 2006 issue of <strong>the</strong> journal. John<br />

Pomeroy (University of S<strong>as</strong>katchew<strong>an</strong>), Kelly Elder (USDA Forest Service, Fort Collins, Colorado), <strong>an</strong>d Andrew<br />

Klein (Tex<strong>as</strong> A&M University) edited this special issue of Hydrological Processes dedicated to ESC/WSC papers.<br />

We th<strong>an</strong>k <strong>the</strong> editors at ERDC–CRREL. Over <strong>the</strong> years this group h<strong>as</strong> contributed considerable time to this<br />

publication <strong>an</strong>d h<strong>as</strong> enh<strong>an</strong>ced <strong>the</strong> quality of <strong>the</strong>se <strong>proceedings</strong>.<br />

The 2006 meeting of <strong>the</strong> E<strong>as</strong>tern <strong>Snow</strong> Conference <strong>an</strong>d <strong>the</strong>se <strong>proceedings</strong> were made possible by sponsorship<br />

<strong>an</strong>d corporate memberships by <strong>the</strong> following:<br />

Campbell Scientific (C<strong>an</strong>ada) Corporation ERDC–CRREL<br />

Edmonton, Alberta, C<strong>an</strong>ada H<strong>an</strong>over, New Hampshire, USA<br />

http://www.campbellsci.com/offices/csc.html http://www.crrel.usace.army.mil/<br />

University of Delaware GEONOR<br />

Newark, Delaware, USA Milford, Pennsylv<strong>an</strong>ia, USA<br />

http://www.udel.edu http://www.geonor.com<br />

Cryospheric Specialty Group of <strong>the</strong> Association of Americ<strong>an</strong> Geographers (AAG)<br />

W<strong>as</strong>hington, DC, USA<br />

http://www.geo.hunter.cuny.edu/AAG_CrSG<br />

We look forward to seeing m<strong>an</strong>y of you at <strong>the</strong> 64th E<strong>as</strong>tern <strong>Snow</strong> Conference meeting to be held jointly with<br />

<strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Geophysical Union, C<strong>an</strong>adi<strong>an</strong> Meteorological Society, <strong>an</strong>d <strong>the</strong> Americ<strong>an</strong> Meteorological Society,<br />

28 May–1 June 2007 at Memorial University, St. John’s Newfoundl<strong>an</strong>d, C<strong>an</strong>ada. The 2007 meeting <strong>the</strong>me is “Air,<br />

Oce<strong>an</strong>, Earth, <strong>an</strong>d Ice on <strong>the</strong> Rock.” It should be exciting, especially with <strong>the</strong> addition of <strong>the</strong> o<strong>the</strong>r groups.<br />

Robert Hellström <strong>an</strong>d Sus<strong>an</strong> Fr<strong>an</strong>kenstein<br />

ESC Proceedings Co-Editors<br />

Bridgewater State College <strong>an</strong>d ERDC–CRREL<br />

iii


This page is intentionally bl<strong>an</strong>k.<br />

iv


CONTENTS<br />

Foreword.......................................................................................................................................................................iii<br />

Statement of Purpose...................................................................................................................................................vii<br />

Executives for <strong>the</strong> 63rd E<strong>as</strong>tern <strong>Snow</strong> Conference.......................................................................................................ix<br />

President’s Page............................................................................................................................................................xi<br />

Weisnet Medal for Best Student Paper<br />

The Influence of <strong>Snow</strong>–Soil Moisture Flux on <strong>Snow</strong>pack Metamorphism in Late Winter <strong>an</strong>d Early Spring<br />

Y.C. Chung <strong>an</strong>d A.W. Engl<strong>an</strong>d ...............................................................................................................................3<br />

Campbell Scientific Award for Best C<strong>an</strong>adi<strong>an</strong> Student Paper<br />

Potential of a Water Bal<strong>an</strong>ce Model with High Temporal Resolution for <strong>the</strong> Distributed Modelling of Ice- <strong>an</strong>d<br />

<strong>Snow</strong>melt Processes at High Elevated Sites<br />

G. Koboltschnig, H. Holzm<strong>an</strong>n, W. Schoener, <strong>an</strong>d M. Zappa ..............................................................................19<br />

<strong>Snow</strong> <strong>an</strong>d Climate<br />

20th Century North Americ<strong>an</strong> <strong>Snow</strong> Extent Trends: Climate Ch<strong>an</strong>ge or Natural Climate Variability?<br />

A. Frei, G. Gong, D.A. Robinson, G. Choi, D. Ghatak, <strong>an</strong>d Y. Ge.......................................................................39<br />

<strong>Snow</strong> <strong>an</strong>d Climate Posters<br />

Observed Differences Between <strong>Snow</strong> Extent <strong>an</strong>d <strong>Snow</strong> Variability at Continental Scales<br />

Y. Ge <strong>an</strong>d G. Gong ...............................................................................................................................................45<br />

Climate Variability, <strong>Snow</strong>melt Distribution, <strong>an</strong>d Effects on Streamflow in a C<strong>as</strong>cades Watershed<br />

A. Jefferson, A. Nolin, S. Lewis, M. Payne, G. Gr<strong>an</strong>t, <strong>an</strong>d C. Tague ...................................................................51<br />

Synoptic Patterns Associated with <strong>the</strong> Record <strong>Snow</strong>fall of 1960 in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s<br />

L.B. Perry <strong>an</strong>d C.E. Konrad II .............................................................................................................................55<br />

Influence of <strong>Snow</strong>fall Anomalies on Summer Precipitation in <strong>the</strong> Nor<strong>the</strong>rn Great Plains of North America<br />

S.M. Quiring <strong>an</strong>d D.B. Kluver..............................................................................................................................65<br />

<strong>Snow</strong> Remote Sensing<br />

Estimating Sublimation of Intercepted <strong>an</strong>d Sub-C<strong>an</strong>opy <strong>Snow</strong> Using Eddy Covari<strong>an</strong>ce Systems<br />

N. Molotch, P. Bl<strong>an</strong>ken, M. Williams, A. Turnipseed, R. Monson, <strong>an</strong>d S. Margulis ............................................75<br />

The Retrievals of <strong>Snow</strong> Cover Extent <strong>an</strong>d <strong>Snow</strong> Water Equivalent from a Blended P<strong>as</strong>sive Microwave—Interactive<br />

Multi-Sensor <strong>Snow</strong> Product<br />

C. Kongoli, C.A. De<strong>an</strong>, S.R. Helfrich, <strong>an</strong>d R.R. Ferraro......................................................................................89<br />

Time Series Analysis <strong>an</strong>d Algorithm Development for Estimating SWE in Great Lakes Area Using Microwave Data<br />

A.E. Azar, H. Ghedira, P. Rom<strong>an</strong>ov, S. Mah<strong>an</strong>i, <strong>an</strong>d R. Kh<strong>an</strong>bilvardi ..............................................................105<br />

On <strong>the</strong> Evaluation of <strong>Snow</strong> Water Equivalent Estimates Over <strong>the</strong> Terrestrial Arctic Drainage B<strong>as</strong>in<br />

M.A. Rawlins, M. Fahnestock, S. Frolking, <strong>an</strong>d C.J. Vörösmarty......................................................................121<br />

v


<strong>Snow</strong> Remote Sensing Posters<br />

AMSR-E Algorithm for <strong>Snow</strong>melt Onset Detection in Subarctic Heterogeneous Terrain<br />

J.D. Apgar, J.M. Ramage, R.A. McKenney, <strong>an</strong>d P. Maltais...............................................................................137<br />

Combination of Active <strong>an</strong>d P<strong>as</strong>sive Microwave to Estimate <strong>Snow</strong>pack Properties in Great Lakes Area<br />

A.E. Azar, T. Lakh<strong>an</strong>kar, N. Shahroudi, <strong>an</strong>d R. Kh<strong>an</strong>bilvardi ...........................................................................153<br />

Enh<strong>an</strong>cements <strong>an</strong>d Forthcoming Developments to <strong>the</strong> Interactive Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping System (IMS)<br />

S.R. Helfrich, D. McNamara, B.H. Ramsay, T. Baldwin, <strong>an</strong>d T. K<strong>as</strong>heta..........................................................165<br />

Retreat of Tropical Glaciers in Colombia <strong>an</strong>d Venezuela from 1984 to 2004 <strong>as</strong> Me<strong>as</strong>ured from ASTER <strong>an</strong>d L<strong>an</strong>dsat<br />

Images<br />

J.N. Morris, A.J. Poole, <strong>an</strong>d A.G. Klein .............................................................................................................181<br />

<strong>Snow</strong>pack Processes<br />

<strong>Snow</strong> Cover Patterns <strong>an</strong>d Evolution at B<strong>as</strong>in Scale: GEOtop Model Simulations <strong>an</strong>d Remote Sensing Observations<br />

S. Endrizzi, G. Bertoldi, M. Neteler, <strong>an</strong>d R. Rigon.............................................................................................195<br />

Microstructural Characterization of Firn<br />

I. Baker, R. Obbard, D. Iliescu, <strong>an</strong>d D. Meese...................................................................................................211<br />

Computational Time Steps of Winter Water Bal<strong>an</strong>ce for <strong>Snow</strong> Losses at United States Meteorological Stations<br />

S.R. F<strong>as</strong>snacht ....................................................................................................................................................219<br />

Shaped Solution Domains for <strong>Snow</strong> Properties<br />

R.A. Melloh, S.A. Shoop, <strong>an</strong>d B.A. Coutermarsh................................................................................................231<br />

Glacial <strong>an</strong>d Periglacial Processes<br />

Qu<strong>an</strong>tifying <strong>the</strong> Effect of Anisotropic Properties in <strong>Snow</strong> for Modelling Meltwater Retention<br />

C.E. Bøggild .......................................................................................................................................................247<br />

The Equilibrium Flow <strong>an</strong>d M<strong>as</strong>s Bal<strong>an</strong>ce of <strong>the</strong> Taku Glacier, Al<strong>as</strong>ka, 1950–2005<br />

M.S. Pelto, G.W. Adema, M.J. Beedle, S.R. McGee, M.M. Miller, K.F. Sprenke, <strong>an</strong>d M. L<strong>an</strong>g.........................251<br />

An Embedded Sensor Network for Me<strong>as</strong>uring Hydrometeorological Variability Within <strong>an</strong> Alpine Valley<br />

R.Å. Hellström <strong>an</strong>d B.G. Mark ...........................................................................................................................263<br />

<strong>Snow</strong> <strong>an</strong>d Periglacial Processes Poster<br />

A Treatise on <strong>the</strong> Preponder<strong>an</strong>ce of Designs Over Historic <strong>an</strong>d Me<strong>as</strong>ured <strong>Snow</strong>falls, or No Two <strong>Snow</strong>flakes Are<br />

Alike: Considerations About <strong>the</strong> Formation of <strong>Snow</strong>flakes <strong>an</strong>d <strong>the</strong> Possible Numbers <strong>an</strong>d Shapes of <strong>Snow</strong>flakes<br />

M. Pilipski <strong>an</strong>d J.D. Pilipski...............................................................................................................................283<br />

62nd ESC Papers<br />

Simulations of North Americ<strong>an</strong> <strong>Snow</strong> Cover by AGCMs <strong>an</strong>d AOGCMs<br />

A. Frei, G. Gong, <strong>an</strong>d R. Brown.........................................................................................................................293<br />

The Impact of Patchy <strong>Snow</strong> Cover on <strong>Snow</strong> Water Equivalent Estimates Derived from P<strong>as</strong>sive Microwave<br />

Brightness Temperatures Over a Prairie Environment<br />

K.R. Turchenek, J.M. Piwowari, <strong>an</strong>d C. Derksen...............................................................................................297<br />

� � �<br />

Sno-Foo Award .........................................................................................................................................................311<br />

vi


STATEMENT OF PURPOSE<br />

The E<strong>as</strong>tern <strong>Snow</strong> Conference (ESC) is a joint C<strong>an</strong>adi<strong>an</strong>/U.S. org<strong>an</strong>ization founded in <strong>the</strong> 1940s, originally<br />

with members primarily from e<strong>as</strong>tern North America. Our current members are scientists, snow surveyors,<br />

engineers, technici<strong>an</strong>s, professors, students, <strong>an</strong>d operational <strong>an</strong>d mainten<strong>an</strong>ce professionals from North America, <strong>the</strong><br />

United Kingdom, Jap<strong>an</strong>, <strong>an</strong>d Germ<strong>an</strong>y. There is a western counterpart to <strong>the</strong> ESC, <strong>the</strong> Western <strong>Snow</strong> Conference<br />

(WSC), which also is a joint C<strong>an</strong>adi<strong>an</strong>/U.S. org<strong>an</strong>ization. Every fifth year <strong>the</strong> ESC <strong>an</strong>d <strong>the</strong> WSC hold joint<br />

meetings.<br />

The E<strong>as</strong>tern <strong>Snow</strong> Conference is a forum that brings <strong>the</strong> research <strong>an</strong>d operations communities toge<strong>the</strong>r to<br />

discuss recent work on scientific, applied, <strong>an</strong>d operational issues related to snow <strong>an</strong>d ice. The location of <strong>the</strong><br />

conference alternates yearly between <strong>the</strong> United States <strong>an</strong>d C<strong>an</strong>ada, <strong>an</strong>d attendees present <strong>the</strong>ir work by ei<strong>the</strong>r<br />

giving a talk or presenting a poster. Most resulting papers are reviewed, edited, <strong>an</strong>d published in our yearly<br />

Proceedings of <strong>the</strong> E<strong>as</strong>tern <strong>Snow</strong> Conference. In recent years, <strong>the</strong> ESC meetings have included sessions on snow<br />

physics, winter survival of <strong>an</strong>imals, snow <strong>an</strong>d ice loads on structures, river ice, remote sensing of snow <strong>an</strong>d ice, <strong>an</strong>d<br />

glacier processes. Volumes of <strong>the</strong> Proceedings c<strong>an</strong> be found in libraries throughout North America <strong>an</strong>d Europe, <strong>an</strong>d<br />

<strong>the</strong> papers are also available through <strong>the</strong> National Technical Information Service (NTIS) in <strong>the</strong> United States <strong>an</strong>d<br />

CISTI in C<strong>an</strong>ada.<br />

� � � � � � � � �<br />

Le Colloque sur la neige-région est (ESC) est une org<strong>an</strong>isation americain-c<strong>an</strong>adienne fondée d<strong>an</strong>s les <strong>an</strong>nées<br />

’40 et dont les membres provenaient a l’origine surtout de l’est de l’Amérique-du-Nord. Actuellement, les membres,<br />

qu’ils soient chercheurs, techniciens en enneigement, ingenieurs, techniciens, professeurs, étudi<strong>an</strong>ts, et spécialistes<br />

des services d’éxploitation et d’entretien, viennent non seulement d’Amérique-du-Nord, mais aussi du Royaume<br />

Uni, du Japon, et d’Allemagne. Le Colloque sur la neige-region ouest (WSC), aussi une org<strong>an</strong>isation americainc<strong>an</strong>adienne,<br />

est l’homologue de l’ESC pour l’ouest nord-americain. Tous les cinq <strong>an</strong>s, l’ESC et la WSC org<strong>an</strong>isent<br />

des réunions en commun.<br />

Le Colloque sur la neige-region est un forum qui r<strong>as</strong>semble chercheurs et responsables des services<br />

d’exploitation pour discuter des travaux récents sur les problemes scientifiques, operationnels, ou autres dus à la<br />

neige et à la glace. Le site de cette réunion <strong>an</strong>nuelle alterne entre les États Unis et le C<strong>an</strong>ada. Les particip<strong>an</strong>ts y<br />

présentent les résultats de leurs travaux par des communications orales ou au moyen d’affiches. Ces<br />

communications, une fois revues et éditées, sont publiées d<strong>an</strong>s les Annales de l’ESC. D<strong>an</strong>s les <strong>an</strong>nées récentes, les<br />

réunions de l’ESC ont inclus des sessions sur la physique de la neige, la survie hivernale de la faune, les forces<br />

exercées par la neige et la glace sur les structures et les batiments, la glace de rivière, la télédétection de la neige et<br />

de la glace, et les processus glaciaires. Les Annales de l’ESC sont accessibles d<strong>an</strong>s la plupart des bibliothèques<br />

scientifiques d’Amerique-du-Nord et d’Europe. Des copies d’articles peuvent être obtenues du National Technical<br />

Information Service (NTIS) aux États Unis et son équivalent au C<strong>an</strong>ada, le CISTI.<br />

vii


This page is intentionally bl<strong>an</strong>k.<br />

viii


EXECUTIVES FOR THE 63rd EASTERN SNOW CONFERENCE<br />

PRESIDENT PAST PRESIDENT<br />

Claude Duguay Sus<strong>an</strong> Taylor<br />

University of Waterloo ERDC-CRREL<br />

Waterloo, Ontario, C<strong>an</strong>ada H<strong>an</strong>over, New Hampshire, USA<br />

VICE-PRESIDENT AND PROGRAM CHAIR<br />

Andrew Klein<br />

Tex<strong>as</strong> A&M University<br />

College Station, Tex<strong>as</strong>, USA<br />

SECRETARY–TREASURER (C<strong>an</strong>ada) SECRETARY–TREASURER (United States)<br />

Miles Ecclestone Derrill Cowing<br />

Trent University U.S. Geological Service (Retired)<br />

Peterborough, Ontario, C<strong>an</strong>ada Monmouth, Maine, USA<br />

ESC PROCEEDINGS EDITORS<br />

Robert Hellström, Bridgewater State College, Bridgewater, M<strong>as</strong>sachusetts, USA<br />

Sus<strong>an</strong> Fr<strong>an</strong>kenstein, ERDC-CRREL, H<strong>an</strong>over, New Hampshire, USA<br />

HYDROLOGICAL PROCESSES SPECIAL ISSUE EDITORS<br />

John Pomeroy<br />

University of S<strong>as</strong>katchew<strong>an</strong><br />

S<strong>as</strong>katoon, S<strong>as</strong>katchew<strong>an</strong>, C<strong>an</strong>ada<br />

Andrew Klein<br />

Tex<strong>as</strong> A&M University<br />

College Station, Tex<strong>as</strong>, USA<br />

STEERING COMMITTEE<br />

Chris Derksen (Chair)<br />

Meteorological Service of C<strong>an</strong>ada<br />

Downsville, Ontario, C<strong>an</strong>ada<br />

J<strong>an</strong>et Hardy<br />

ERDC-CRREL<br />

H<strong>an</strong>over, New Hampshire, USA<br />

Steven F<strong>as</strong>snacht<br />

Colorado State University<br />

Fort Collins, Colorado, USA<br />

Mauri Pelto<br />

Nichols College<br />

Dudley, M<strong>as</strong>sachusetts, USA<br />

RESEARCH COMMITTEE<br />

Steven F<strong>as</strong>snacht (Chair)<br />

Colorado State University<br />

Fort Collins, Colorado, USA<br />

Steven Déry<br />

Princeton University<br />

Princeton, New Jersey, USA<br />

ix


RESEARCH COMMITTEE (continued from page ix)<br />

All<strong>an</strong> Frei<br />

Hunter College of <strong>the</strong> CUNY<br />

New York, New York, USA<br />

Richard Kelly<br />

Maryl<strong>an</strong>d, USA<br />

Nat<strong>as</strong>ha Neum<strong>an</strong>n<br />

Meteorological Service of C<strong>an</strong>ada<br />

S<strong>as</strong>katoon, S<strong>as</strong>katchew<strong>an</strong>, C<strong>an</strong>ada<br />

WEB MASTER<br />

Andrew Klein<br />

Tex<strong>as</strong> A&M University<br />

College Station, Tex<strong>as</strong>, USA<br />

LOCAL ARRANGEMENTS<br />

Delphis Levia (Chair)<br />

University of Delaware<br />

Newark, Delaware, USA<br />

Andrew Klein<br />

Tex<strong>as</strong> A&M University<br />

College Station, Tex<strong>as</strong>, USA<br />

Miles Ecclestone<br />

Trent University<br />

Peterborough, Ontario, C<strong>an</strong>ada<br />

LIFE MEMBERS<br />

Art Eschner<br />

Austin Hog<strong>an</strong><br />

H. Gerry Jones<br />

John Metcalfe<br />

Hilda Snelling<br />

Don Wiesnet<br />

x


THE PRESIDENT’S PAGE<br />

The 63rd <strong>an</strong>nual meeting of <strong>the</strong> E<strong>as</strong>tern <strong>Snow</strong> Conference (ESC) w<strong>as</strong> held at <strong>the</strong> University of Delaware in<br />

Newark, Delaware (7–9 June 2006). There were sessions on snow <strong>an</strong>d climate, remote sensing of snow processes,<br />

<strong>an</strong>d glacial <strong>an</strong>d periglacial processes. The quality of <strong>the</strong> presentations w<strong>as</strong> exceptional <strong>an</strong>d I th<strong>an</strong>k all who shared<br />

<strong>the</strong>ir data at <strong>the</strong> meeting.<br />

The success of <strong>the</strong> 2006 meeting rests with <strong>the</strong> Executive. I th<strong>an</strong>k Delphis Levia, who, along with Andrew<br />

Klein <strong>an</strong>d Miles Ecclestone, made all <strong>the</strong> local arr<strong>an</strong>gements; our tre<strong>as</strong>urers, Derrill Cowing <strong>an</strong>d Miles Ecclestone,<br />

for keeping us in <strong>the</strong> black; <strong>an</strong>d Claude Duguay (VP <strong>an</strong>d Program Chair) for org<strong>an</strong>izing <strong>an</strong>d running <strong>the</strong> 63rd<br />

meeting. Andrew Klein, our Webm<strong>as</strong>ter, h<strong>as</strong> helped <strong>the</strong> ESC join <strong>the</strong> digital age. I also th<strong>an</strong>k our Corporate<br />

members (listed in <strong>the</strong> Foreword) for <strong>the</strong>ir continued support of <strong>the</strong> ESC.<br />

The winner of this year’s Weisnet Medal for <strong>the</strong> best student paper w<strong>as</strong> Y.C. Chung for his work on <strong>the</strong><br />

influence of <strong>the</strong> snow–soil moisture flux on snowpack metamorphism. The Campbell Scientific Medal, which is<br />

awarded to <strong>the</strong> best student paper in which instrumentation plays a key role, w<strong>as</strong> awarded to Gernot Koboltschnig<br />

for his investigation into <strong>the</strong> potential of using a water bal<strong>an</strong>ce model with high temporal resolution for distributed<br />

modeling of melting at high elevation.<br />

The ESC Proceedings have been published for 63 years. I find this a remarkable feat for a small org<strong>an</strong>ization<br />

run by a h<strong>an</strong>dful of researchers. Editing <strong>the</strong> <strong>proceedings</strong> is one of <strong>the</strong> most time-consuming jobs <strong>an</strong>d I th<strong>an</strong>k our<br />

Editors, Dr. Robert Hellström <strong>an</strong>d Sus<strong>an</strong> Fr<strong>an</strong>kenstein, for <strong>the</strong>ir hard work. This is also <strong>the</strong> eleventh <strong>an</strong>niversary of<br />

ESC’s partnership with Hydrological Processes, in which a number of <strong>the</strong> papers presented at <strong>the</strong> conference are<br />

published. I th<strong>an</strong>k Hydrological Processes for providing ESC particip<strong>an</strong>ts <strong>the</strong> opportunity to publish in <strong>the</strong> journal,<br />

<strong>an</strong>d Dr. John Pomeroy <strong>an</strong>d Dr. Andrew Klein for h<strong>an</strong>dling <strong>the</strong> Hydrological Processes submissions.<br />

I invite you to attend <strong>the</strong> 64th meeting of <strong>the</strong> E<strong>as</strong>tern <strong>Snow</strong> Conference, to be held in conjunction with <strong>the</strong><br />

C<strong>an</strong>adi<strong>an</strong> Meteorological <strong>an</strong>d Oce<strong>an</strong>ographic Society, <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Geophysical Society, <strong>an</strong>d <strong>the</strong> Americ<strong>an</strong><br />

Meteorological Society at <strong>the</strong> end of May 2007 in St. John’s, Newfoundl<strong>an</strong>d. Ple<strong>as</strong>e visit our Web site<br />

(www.e<strong>as</strong>ternsnow.org) for up-to-date details of this meeting.<br />

xi<br />

Best Regards,<br />

Claude Duguay<br />

63rd President<br />

E<strong>as</strong>tern <strong>Snow</strong> Conference


This page is intentionally bl<strong>an</strong>k.<br />

xii


Weisnet Medal<br />

for Best Student Paper<br />

1


This page is intentionally bl<strong>an</strong>k.<br />

2


3<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

The Influence of <strong>Snow</strong>–Soil Moisture Flux<br />

on <strong>Snow</strong>pack Metamorphism in Late Winter <strong>an</strong>d Early Spring<br />

ABSTRACT<br />

Y. C. CHUNG 1 AND A. W. ENGLAND 2<br />

The import<strong>an</strong>ce of l<strong>an</strong>d–atmosphere snow cover feedback in <strong>the</strong> climate system, water storage<br />

<strong>an</strong>d rele<strong>as</strong>e from snowpacks, <strong>an</strong>d <strong>the</strong> large l<strong>an</strong>d are<strong>as</strong> covered by snow in <strong>the</strong> nor<strong>the</strong>rn hemisphere,<br />

are re<strong>as</strong>ons to accurately model late winter <strong>an</strong>d early spring snow metamorphic processes.<br />

<strong>Snow</strong>pack models like SNTHERM predict snow behavior very well during <strong>the</strong> cold periods but do<br />

not adequately capture liquid <strong>an</strong>d vapor tr<strong>an</strong>sport between soil <strong>an</strong>d snow conditions during <strong>the</strong><br />

snowmelt period. To get a more realistic description of <strong>the</strong> dynamics between snowpack <strong>an</strong>d soil,<br />

we developed a snow–soil–vegetation–atmosphere tr<strong>an</strong>sfer (ssvat) model, which combines soil<br />

processes from our l<strong>an</strong>d surface processes (lsp) with <strong>the</strong> snowpack processes of SNTHERM.<br />

We validated <strong>the</strong> model with data from <strong>the</strong> NASA cold l<strong>an</strong>d processes field experiment (CLPX)<br />

field experiment conducted in late winter <strong>an</strong>d early spring near Fr<strong>as</strong>er, Colorado, in 2003. We<br />

show <strong>an</strong> example where soil/snow fluxes incre<strong>as</strong>e <strong>the</strong> energy stored within <strong>the</strong> snowpack by <strong>as</strong><br />

much <strong>as</strong> 17%. Thermal conduction w<strong>as</strong> <strong>the</strong> predomin<strong>an</strong>t soil/snow energy flux. O<strong>the</strong>r energy<br />

fluxes, those <strong>as</strong>sociated with air convection <strong>an</strong>d vapor diffusion, were 10 –7 smaller but do<br />

contribute to <strong>the</strong> formation of depth hoar <strong>an</strong>d to grain size growth. Grain sizes in <strong>the</strong> upper<br />

snowpack are incre<strong>as</strong>ed by soil/snow fluxes. Because radiobrightness scatter darkening at<br />

frequencies above 19 GHz incre<strong>as</strong>es with grain size, static snow water equivalent (swe) algorithms<br />

that use scatter darkening to estimate swe become less reliable when <strong>the</strong> soil/snow vapor fluxes<br />

are large <strong>as</strong> <strong>the</strong>y would be when <strong>the</strong> soil is wet <strong>an</strong>d unfrozen. Comparisons of SNTHERM <strong>an</strong>d<br />

SSVAT with <strong>the</strong> CLPX data show that SSVAT provided better estimates of soil temperature by<br />

1.4 K in <strong>the</strong> late winter <strong>an</strong>d of snow temperature by 3.2 K in early spring. SSVAT also provides<br />

better moisture profiles in both snow <strong>an</strong>d soil.<br />

The combination of SSVAT <strong>an</strong>d a new radiobrightness module will enable monitoring of water<br />

stored in snow over frozen or unfrozen soil during periods when <strong>the</strong> snowpack is experiencing<br />

thawing <strong>an</strong>d refreezing. It will also contribute to <strong>the</strong> design of <strong>the</strong> NASA cold l<strong>an</strong>ds processes<br />

pathfinder (CLPP) field experiment pl<strong>an</strong>ned for <strong>the</strong> arctic <strong>an</strong>d help prepare <strong>the</strong> cold l<strong>an</strong>ds<br />

community to interpret <strong>the</strong> 1.4 GHz brightness observations from SMOS.<br />

Keywords: snow processes; soil processes; vapor diffusion; water flow; free convection; depth<br />

hoar<br />

INTRODUCTION<br />

<strong>Snow</strong>melt runoff in spring is a major source of hydroelectric <strong>an</strong>d irrigation water at middle <strong>an</strong>d<br />

high latitudes. <strong>Snow</strong> <strong>an</strong>d ice also signific<strong>an</strong>tly incre<strong>as</strong>e <strong>the</strong> Earth's albedo, promoting cooling.<br />

1<br />

Geosciences <strong>an</strong>d Remote Sensing Graduate Program, University of Michig<strong>an</strong>, Ann Arbor,<br />

Michig<strong>an</strong> 48109-2143, USA.<br />

2<br />

College of Engineering, University of Michig<strong>an</strong>, Ann Arbor, Michig<strong>an</strong> 48109-2102, USA.


Even small ch<strong>an</strong>ges in climate influence <strong>the</strong>se snow processes. The removal of <strong>the</strong> snow cover<br />

also initiates <strong>the</strong> melting of river <strong>an</strong>d lake ice, <strong>the</strong> thawing of <strong>the</strong> active layer, <strong>an</strong>d marks <strong>the</strong><br />

beginning of <strong>the</strong> evaporation se<strong>as</strong>on. Meltwater incre<strong>as</strong>es soil moisture <strong>an</strong>d initiates or incre<strong>as</strong>es<br />

streamflow. The decaying snowpack causes a major ch<strong>an</strong>ge in <strong>the</strong> surface energy bal<strong>an</strong>ce of arctic<br />

regions.<br />

A comprehensive search of <strong>the</strong> literature by <strong>the</strong> <strong>Snow</strong> Models Intercomparison Project<br />

(SNOWMIP) in 2001 identified more th<strong>an</strong> 40 snow models motivated ei<strong>the</strong>r by snow hazard<br />

investigation or <strong>as</strong> <strong>the</strong> lower boundary of <strong>an</strong> atmospheric model (Y<strong>an</strong>g, 2004). Vapor tr<strong>an</strong>sfer<br />

processes are treated in only 10% of <strong>the</strong> snow models. Most snow models are validated with field<br />

data but 10% are validated with remote sensing data. Effects of sub-grid-scale topography on<br />

distribution of precipitation, air temperature <strong>an</strong>d snow depth are considered in 25% of <strong>the</strong> snow<br />

models (Jord<strong>an</strong>, 1991; Y<strong>an</strong>g, 2004).<br />

SNTHERM w<strong>as</strong> chosen <strong>as</strong> <strong>the</strong> b<strong>as</strong>is for our snow cover-soil model because it offers <strong>the</strong> high<br />

physical fidelity (Jord<strong>an</strong>, 1991; Jord<strong>an</strong>, 1999). SNTHERM considers retention/percolation,<br />

refreezing, vapor tr<strong>an</strong>sfer, <strong>an</strong>d heat tr<strong>an</strong>sport by rainfall or snowfall for <strong>the</strong> snowpack. <strong>Snow</strong><br />

parameters, such <strong>as</strong> density <strong>an</strong>d heat capacity/conductivity evolve with <strong>the</strong> snowpack. SNTHERM<br />

predicts <strong>the</strong> temperature profiles within <strong>the</strong> frozen soil <strong>an</strong>d <strong>the</strong> stratified snow with unlimited<br />

layers. SNTHERM h<strong>as</strong> been verified with field data for several years from sites in Grayling,<br />

Michig<strong>an</strong>, <strong>an</strong>d H<strong>an</strong>over, New Hampshire. SNTHERM’s weaknesses are that it ignores <strong>the</strong> liquid<br />

water <strong>an</strong>d vapor tr<strong>an</strong>sport in <strong>the</strong> soil <strong>an</strong>d allows water to be artificially drained at <strong>the</strong> snow/soil<br />

interface.<br />

Although soil processes have been combined in some models, <strong>the</strong>se models were simplified in<br />

o<strong>the</strong>r ways that compromise overall perform<strong>an</strong>ce. Among <strong>the</strong>se models are SHAW,<br />

SNOWPACK, <strong>an</strong>d ISBA-ES (Flerchinger <strong>an</strong>d H<strong>an</strong>son. 1989; Lehning et al., 1998; Boone <strong>an</strong>d<br />

Etchevers, 2001). ISBA-ES <strong>an</strong>d SNOWPACK only consider <strong>the</strong> soil processes on <strong>the</strong> surface or<br />

subsurface. In ISBA-ES, <strong>the</strong> subsurface soil column is divided into two subsurface soil moisture<br />

reservoirs consisting of a root-zone layer <strong>an</strong>d a subroot-zone layer (Boone <strong>an</strong>d Etchevers, 2001).<br />

The purpose of <strong>the</strong> ISBA scheme is to calculate <strong>the</strong> surface radiative heat, momentum, <strong>an</strong>d<br />

moisture exch<strong>an</strong>ges with <strong>the</strong> atmosphere <strong>an</strong>d <strong>the</strong> components of <strong>the</strong> near-surface hydrological<br />

budget. SHAW considers similar soil processes to those of our LSP model but it h<strong>as</strong> less detailed<br />

snow processes th<strong>an</strong> SNTHERM. Nei<strong>the</strong>r precipitation nor litter fall are allowed <strong>an</strong>d <strong>the</strong><br />

interaction between snow grain <strong>an</strong>d vapor diffusion are ignored. SHAW w<strong>as</strong> developed for a<br />

continuous c<strong>an</strong>opy cover.<br />

Our Prairie L<strong>an</strong>d Surface Processes (LSP) model is a one-dimensional, high physical fidelity<br />

model with coupled heat <strong>an</strong>d moisture tr<strong>an</strong>sport for prairie soils (Liou, 1996; Judge, 1999). Figure<br />

1 illustrates <strong>the</strong> processes in <strong>the</strong> Prairie LSP model, including interactions between <strong>the</strong><br />

atmosphere, <strong>the</strong> c<strong>an</strong>opy <strong>an</strong>d <strong>the</strong> soil, <strong>an</strong>d <strong>the</strong> effect of tr<strong>an</strong>spiration on moisture <strong>an</strong>d energy fluxes<br />

within <strong>the</strong> root zone. To get a more realistic description of <strong>the</strong> snowpack–soil dynamics, we<br />

developed <strong>the</strong> <strong>Snow</strong>–Soil–Vegetation–Atmosphere Tr<strong>an</strong>sfer (SSVAT) model, which combines<br />

soil processes from our L<strong>an</strong>d Surface Processes (LSP) model with <strong>the</strong> snowpack processes of<br />

SNTHERM (Chung <strong>an</strong>d Engl<strong>an</strong>d, 2005; Chung et al., 2006).<br />

MODEL MODIFICATIONS<br />

The coupled SSVAT model employs <strong>an</strong> index of layers in snow <strong>an</strong>d soil in descending order<br />

from <strong>the</strong> top down <strong>as</strong> seen in Fig 1.<br />

4


Figure 1. Layer 1 represents <strong>the</strong> soil b<strong>as</strong>e; nsoil represents <strong>the</strong> soil layer under <strong>the</strong> snow/soil interface; <strong>an</strong>d n<br />

represents <strong>the</strong> top snow layer in <strong>the</strong> model structure (modified from Liou, 1996).<br />

Water Flow<br />

SNTHERM includes only gravity driven liquid water flow but liquid water is driven by both<br />

capillarity <strong>an</strong>d gravity (Tao <strong>an</strong>d Gray, 1994). Ignoring <strong>the</strong> capillary effect limits water flow to <strong>the</strong><br />

downward direction. Particularly, <strong>the</strong> capillary effect should not be neglected under non-steady<br />

state conditions (Su et al, 1999; Waldner et al, 2004). Capillary forces c<strong>an</strong> be neglected only for a<br />

stationary unsaturated water flow downwards. Water flow processes in SNTHERM have been<br />

modified in SSVAT to incorporate <strong>the</strong> capillary effect.<br />

Heat convection due to liquid water flow is described by Eqn (1) in SNTHERM where cw is<br />

specific heat of water per unit volume, Uw is <strong>the</strong> downward flow rate of liquid water, T i is <strong>the</strong><br />

temperature of <strong>the</strong> i th layer, <strong>the</strong> superscripts (I ±1/2) refer to <strong>the</strong> interfaces at <strong>the</strong> top <strong>an</strong>d bottom of<br />

i+<br />

1<br />

layer i. The heat flux due to water flow across upper boundary into layer i is H <strong>an</strong>d <strong>the</strong> heat<br />

w<br />

i<br />

flux across lower boundary out of layer i is H : w<br />

1<br />

i+<br />

1 i+<br />

1 i+<br />

2 i+<br />

1<br />

H w = cw<br />

U w T<br />

(1)<br />

1<br />

i i i−<br />

2 i<br />

H = c U T<br />

w<br />

w<br />

w<br />

Including water flow due to capillary requires consideration of <strong>the</strong> direction of water flow. For<br />

example, Eqn (2) describes flux out of <strong>the</strong> upper boundary of layer i if <strong>the</strong> water flow is upward<br />

1<br />

i+<br />

2<br />

due to capillary forces ( U w 0).<br />

1<br />

1<br />

i+<br />

1 i i+<br />

2 i i+<br />

2<br />

H w = cwU<br />

w T , if ( U w < 0)<br />

(2)<br />

H<br />

i+<br />

1<br />

w<br />

= c<br />

i+<br />

1<br />

w<br />

U<br />

1<br />

i+<br />

2<br />

w<br />

T<br />

i+<br />

1<br />

, if<br />

1<br />

i+<br />

2 ( U w > 0)<br />

T i+1<br />

T i<br />

Figure 2. Red line shows <strong>the</strong> heat tr<strong>an</strong>sport across upper boundary of <strong>the</strong> layer i.<br />

Depth Hoar<br />

Depth hoar, which is unique to SSVAT, c<strong>an</strong> ch<strong>an</strong>ge <strong>the</strong> insulating effect of <strong>the</strong> se<strong>as</strong>onal snow<br />

cover signific<strong>an</strong>tly during ablation periods because of <strong>the</strong> lower density <strong>an</strong>d <strong>the</strong>rmal conductivity<br />

5<br />

snow layer i+1<br />

snow layer i<br />

(3)


of depth hoar. The growth of depth hoar is favored when <strong>the</strong> difference between soil <strong>an</strong>d air<br />

temperatures is greater <strong>an</strong>d when vegetation buried within <strong>the</strong> snow results in a high porosity of<br />

snow cover near <strong>the</strong> snow–soil interface (Zh<strong>an</strong>g et al., 1997). Natural convection of air in snow<br />

c<strong>an</strong> occur to form depth hoar if a critical temperature gradient is exceeded. The convective flux<br />

c<strong>an</strong> be determined from <strong>the</strong> relationship between <strong>the</strong> thickness of <strong>the</strong> snow cover <strong>an</strong>d <strong>the</strong><br />

temperature gradient (Fig 44 of Akitaya, 1974).<br />

<strong>Snow</strong>/Soil Interface<br />

M<strong>as</strong>s <strong>an</strong>d energy tr<strong>an</strong>sfer at <strong>the</strong> snow/soil interface in SSVAT are b<strong>as</strong>ed on <strong>the</strong> algorithms from<br />

LSP (Liou, 1996). The infiltration model for soil c<strong>an</strong> be estimated using a qu<strong>as</strong>i-<strong>an</strong>alytic solution<br />

to Richard’s equation for vertical infiltration in a homogeneous soil (Liou, 1996; Judge, 1999). To<br />

precisely capture m<strong>as</strong>s <strong>an</strong>d energy tr<strong>an</strong>sfer at <strong>the</strong> soil/snow interface, snow <strong>an</strong>d soil layers at <strong>the</strong><br />

soil/snow interface should be thin.<br />

Figure 3. Simulations of temperature <strong>an</strong>d liquid water content for a thatch layer(modified from Chung et al.,<br />

2006)<br />

Heat tr<strong>an</strong>sfer at <strong>the</strong> snow/soil interface c<strong>an</strong> be affected by vegetation. A thatch layer in LSP w<strong>as</strong><br />

defined <strong>as</strong> a 2cm layer of org<strong>an</strong>ic matter at <strong>the</strong> b<strong>as</strong>e of <strong>the</strong> gr<strong>as</strong>s c<strong>an</strong>opy that is subject to <strong>the</strong> heat<br />

exch<strong>an</strong>ge with <strong>the</strong> atmosphere, <strong>the</strong> c<strong>an</strong>opy <strong>an</strong>d <strong>the</strong> underlying soil (Liou, 1996). To see if this<br />

thatch layer buried within <strong>the</strong> snow is import<strong>an</strong>t for heat exch<strong>an</strong>ges in SSVAT, a simulation w<strong>as</strong><br />

run for two weeks using a forcing data from Feburary 23, 2003 in CLPX (Chung et al., 2006). Our<br />

results showed that a thatch layer did not affect moisture <strong>an</strong>d temperature profiles of ei<strong>the</strong>r snow<br />

or soil when <strong>the</strong> thatch fraction varied (Fig 3). SSVAT thus ignores thatch at <strong>the</strong> snow/soil<br />

interface.<br />

RESULTS<br />

Study Area<br />

Fig 4 shows <strong>the</strong> study area at <strong>the</strong> Local Scale Observation Site (LSOS) of <strong>the</strong> NASA Cold L<strong>an</strong>d<br />

Processes Field Experiment (CLPX). SSVAT requires temperature, moisture <strong>an</strong>d grain size<br />

profiles for initialization, which were taken from <strong>the</strong> Micrometeorological Data <strong>an</strong>d <strong>Snow</strong><br />

Me<strong>as</strong>urements (Cline et al., 2002; Engl<strong>an</strong>d, 2003). SSVAT also requires precipitation data, which<br />

were taken from Ground B<strong>as</strong>ed P<strong>as</strong>sive Microwave Radiometer Data (Graf et al, 2003). Data for<br />

model validation came from snow pit me<strong>as</strong>urements of density <strong>an</strong>d grain size profiles <strong>an</strong>d subc<strong>an</strong>opy<br />

meteorological observations for late winter <strong>an</strong>d early spring (Cline et al., 2002; Hardy et<br />

al., 2002).<br />

6


Figure 4. The general layout of <strong>the</strong> LSOS during IOP3 <strong>an</strong>d IOP4 (Engl<strong>an</strong>d, 2003).<br />

The soil type in <strong>the</strong> study area is s<strong>an</strong>dy loam, whose dry subst<strong>an</strong>ce consists of 52% quartz, 7%<br />

clay, <strong>an</strong>d residual for silt (Cline et al., 2001). Trees in <strong>the</strong> study area are of mixed species<br />

(predomin<strong>an</strong>tly lodgepole pine with some Englem<strong>an</strong>n Spruce <strong>an</strong>d Subalpine Fir) with <strong>an</strong> average<br />

tree height of 7.8 m (st<strong>an</strong>dard deviation = 4.8 m; n = 88) <strong>an</strong>d heterogeneous spacing between trees<br />

(Hardy et al., 2002).<br />

Energy Stored within <strong>the</strong> <strong>Snow</strong>pack<br />

SSVAT <strong>an</strong>d SNTHERM were forced by downwelling radi<strong>an</strong>ce <strong>an</strong>d observed meteorology for<br />

<strong>the</strong> periods of IOP3 (2/19~2/24) <strong>an</strong>d IOP4 (3/25~3/29). The results were compared with<br />

observations. Figure 5 shows <strong>the</strong> model-derived heat contents stored in <strong>the</strong> snowpack over <strong>the</strong><br />

study period. The simulations show <strong>an</strong> average incre<strong>as</strong>e for SSVAT relative to SNTHERM of 0.54<br />

W/m 2 for late winter <strong>an</strong>d 1.87 W/m 2 for early spring. That is, SSVAT stored heat contents were<br />

incre<strong>as</strong>ed by soil/snow fluxes. Exceptions occurred on some afternoons (e.g. DOY 86~87 <strong>an</strong>d<br />

DOY 51~52) when <strong>the</strong> snowpack became warmer in <strong>the</strong> late afternoon tr<strong>an</strong>sporting heat to its<br />

underlying frozen soil. In general, SSVAT predicted a warmer snowpack. Soil processes incre<strong>as</strong>ed<br />

energy stored in <strong>the</strong> snowpack by <strong>as</strong> much <strong>as</strong> 17%.<br />

Figure 5. Heat content stored in <strong>the</strong> snowpack predicted by two models <strong>an</strong>d <strong>the</strong> model differences over <strong>the</strong><br />

study period.<br />

7


Thermal Conduction<br />

Heat fluxes at <strong>the</strong> snow/soil interface include <strong>the</strong>rmal conduction, vapor diffusion, <strong>an</strong>d<br />

convection from free air <strong>an</strong>d flow water (Chung et al., 2006). Figure 6 shows <strong>the</strong> <strong>the</strong>rmal<br />

conduction fluxes, with <strong>an</strong> average difference between two models of 1.43 W/m 2 for late winter<br />

(IOP3) <strong>an</strong>d 0.06 W/m 2 for early spring (IOP4). The daily oscillation w<strong>as</strong> less dramatic in early<br />

spring th<strong>an</strong> in late winter since <strong>the</strong> temperature differences at <strong>the</strong> interface w<strong>as</strong> less signific<strong>an</strong>t in<br />

early spring (soil <strong>an</strong>d snow both thawed). This suggests that soil conduction is predomin<strong>an</strong>t in <strong>the</strong><br />

soil heat contribution to <strong>the</strong> snowpack.<br />

Figure 6. Simulated soil heat fluxes at <strong>the</strong> snow/soil interface over <strong>the</strong> study period.<br />

Diffusion <strong>an</strong>d Convection<br />

SNTHERM only considers <strong>the</strong>rmal conduction from <strong>the</strong> soil. Vapor diffusion from <strong>the</strong> soil <strong>an</strong>d<br />

natural convection of free air in snow are unique to SSVAT. On average, vapor diffusion <strong>an</strong>d<br />

natural convection were 10 –7 times smaller th<strong>an</strong> <strong>the</strong>rmal conduction at <strong>the</strong> interface (Fig 7). Vapor<br />

diffusion displayed a diurnal cycle in late winter because it w<strong>as</strong> affected by <strong>the</strong> diurnal<br />

temperature cycle. Natural convection of free air in snow w<strong>as</strong> signific<strong>an</strong>t in early spring,<br />

especially on DOY 87–88. Comparing Fig 5 with Fig 7, air convection may be responsible for <strong>the</strong><br />

incre<strong>as</strong>ing variability of <strong>the</strong> heat content in snow on DOY 87–89. This suggests <strong>the</strong> import<strong>an</strong>ce of<br />

convection in snow may exceed that of vapor diffusion or <strong>the</strong> <strong>the</strong>rmal conduction.<br />

Figure 7. Simulated heat fluxes at <strong>the</strong> snow/soil interface over <strong>the</strong> study period.<br />

Air Convection<br />

Fig 8 shows <strong>the</strong> occurrence of air convection in a snowpack in late winter (IOP3) <strong>an</strong>d early<br />

spring (IOP4). Red color <strong>an</strong>d black color represented <strong>the</strong> occurrence of <strong>the</strong> air convection where<strong>as</strong><br />

8


lue color represented no convection. Air convection w<strong>as</strong> predicted to occur more frequently in<br />

<strong>the</strong> upper snowpack, especially in early spring. This may reflect on <strong>the</strong> snow <strong>an</strong>d soil<br />

characteristics, such <strong>as</strong> temperature, moisture <strong>an</strong>d grain size.<br />

Figure 8. Air convection simulated over <strong>the</strong> study period.<br />

Temperature Profiles<br />

Figure 9 compares <strong>the</strong> temperature profiles predicted by <strong>the</strong> two models with observations. The<br />

maximum difference between SSVAT <strong>an</strong>d observation of 0.3 K w<strong>as</strong> observed in soil temperature<br />

in late winter scenario, better <strong>the</strong>n that of SNTHERM (1.7 K). The maximum difference of 2.8 K<br />

w<strong>as</strong> observed by SSVAT in snow temperature, better th<strong>an</strong> 6 K by SNTHERM in early spring<br />

scenario. As expected, <strong>the</strong> bigger improvement w<strong>as</strong> displayed in early spring once <strong>the</strong> snowpack<br />

started melting. It also reflected on <strong>the</strong> simulated flux in <strong>the</strong> figures above. The temperature in <strong>the</strong><br />

upper snowpack c<strong>an</strong> be affected by <strong>the</strong> soil processes even when <strong>the</strong> soil <strong>an</strong>d lower snowpack<br />

temperature predictions by two models are similar (Fig 7 <strong>an</strong>d 8). This suggests that SSVAT better<br />

estimates temperature during <strong>the</strong> variable climate of late winter <strong>an</strong>d early spring, <strong>an</strong>d that <strong>the</strong> soil<br />

processes c<strong>an</strong>not be ignored for predicting <strong>the</strong> evolution of <strong>the</strong> snowpack.<br />

9


Figure 9(a). Simulated temperature profiles in late winter.<br />

Figure 9(b). Simulated temperature profiles in early spring.<br />

10


Total Density of Water Profiles<br />

The total density profile includes both liquid <strong>an</strong>d solid (ice) ph<strong>as</strong>es. Regression coefficients of<br />

<strong>the</strong> density predicted by two models with observations in <strong>the</strong> soil c<strong>an</strong> be seen in Table 1. It shows<br />

that SSVAT predicted soil moisture well where<strong>as</strong> SNTHERM were poor for soil moisture profiles.<br />

SNTHERM artificially draining <strong>the</strong> water at <strong>the</strong> snow/soil interface so c<strong>an</strong>not accurately predict<br />

<strong>the</strong> soil moisture in late winter.<br />

Table 1. Regression coefficients of <strong>the</strong> model predictions <strong>an</strong>d observations.<br />

SSVAT SNTHERM<br />

1.5 cm* 0.99 0.69<br />

4.5 cm* 0.89 0.68<br />

10 cm* 0.96 **<br />

27 cm* 0.98 **<br />

45 cm* 0.91 **<br />

*me<strong>an</strong>s <strong>the</strong> depth beneath <strong>the</strong> ground (snow/soil interface).<br />

** SNTHERM failed to capture <strong>the</strong> variation of <strong>the</strong> water density.<br />

Fig 10 represents that simulated density of water profiles in late winter <strong>an</strong>d early spring. It also<br />

shows that SSVAT predicted density profiles well. The differences between two model predictions<br />

were not signific<strong>an</strong>t in early spring since <strong>the</strong> model initialization in early spring w<strong>as</strong> more realistic<br />

th<strong>an</strong> that in late winter. O<strong>the</strong>rs may consider liquid water fractions to be signific<strong>an</strong>tly different.<br />

However, we do not have observations of liquid water content to compare. This suggests that <strong>the</strong><br />

SSVAT provides a realistic representation of moisture profiles in snow <strong>an</strong>d soil even under <strong>an</strong> less<br />

realistic initialization.<br />

Figure 10(a). Simulated density of water profiles in late winter.<br />

11


Figure 10(b). Simulated density of water profiles in early spring.<br />

<strong>Snow</strong> Grain Size Profiles<br />

Comparisons show that snow grain size w<strong>as</strong> predicted larger by SSVAT in response to vapor<br />

diffusion contribution from <strong>the</strong> soil that c<strong>an</strong> grow <strong>the</strong> grain size more (Fig 11). This contribution<br />

accumulates when tr<strong>an</strong>sporting through <strong>the</strong> snowpack, making <strong>the</strong> model difference more obvious<br />

near <strong>the</strong> snow surface. These differences might signific<strong>an</strong>tly affect <strong>the</strong> radiobrightness.<br />

12


Figure 11(a). Simulated snow grain size profiles in late winter<br />

Figure 11(b). Simulated snow grain size profiles in early spring.<br />

13


CONCLUSIONS<br />

The SSVAT model w<strong>as</strong> validated with data from <strong>the</strong> CLPX field experiment conducted in late<br />

winter <strong>an</strong>d early spring near Fr<strong>as</strong>er, Colorado, in 2003. We show <strong>an</strong> example where soil/snow<br />

fluxes incre<strong>as</strong>e <strong>the</strong> energy stored within <strong>the</strong> snowpack by <strong>as</strong> much <strong>as</strong> 17%. Thermal conduction<br />

w<strong>as</strong> <strong>the</strong> predomin<strong>an</strong>t soil/snow energy flux. O<strong>the</strong>r energy fluxes, those <strong>as</strong>sociated with air<br />

convection <strong>an</strong>d vapor diffusion, were 10 –7 smaller but do contribute to <strong>the</strong> formation of depth hoar<br />

<strong>an</strong>d to grain size growth. Grain sizes in <strong>the</strong> upper snowpack are incre<strong>as</strong>ed by soil/snow fluxes.<br />

Because radiobrightness scatter darkening at frequencies above 19 GHz incre<strong>as</strong>es with grain size,<br />

static <strong>Snow</strong> Water Equivalent (SWE) algorithms that use scatter darkening to estimate SWE<br />

become less reliable when <strong>the</strong> soil/snow vapor fluxes are large <strong>as</strong> <strong>the</strong>y would be when <strong>the</strong> soil is<br />

wet <strong>an</strong>d unfrozen.<br />

The maximum difference between soil temperatures of SSVAT <strong>an</strong>d observations w<strong>as</strong> 0.3 K in<br />

late winter. The maximum difference between soil temperatures for SNTHERM in that period w<strong>as</strong><br />

1.7 K. The maximum difference between snow temperatures of SSVAT <strong>an</strong>d observations w<strong>as</strong> 2.8 K<br />

in early spring. The maximum difference between snow temperatures of SNTHERM <strong>an</strong>d<br />

observations in that period w<strong>as</strong> 6 K. That is, comparisons of SNTHERM <strong>an</strong>d SSVAT with <strong>the</strong><br />

CLPX data show that SSVAT provided better estimates of soil temperature by 1.4K in late winter<br />

<strong>an</strong>d of snow temperature by 3.2 K in early spring. SSVAT also provides better moisture profiles in<br />

both snow <strong>an</strong>d soil.<br />

The combination of SSVAT <strong>an</strong>d a new radiobrightness module will enable monitoring of water<br />

stored in snow over frozen or unfrozen soil during periods when <strong>the</strong> snowpack is experiencing<br />

thawing <strong>an</strong>d refreezing. It will also contribute to <strong>the</strong> design of <strong>the</strong> NASA Cold L<strong>an</strong>ds Processes<br />

Pathfinder (CLPP) field experiment pl<strong>an</strong>ned for <strong>the</strong> Arctic <strong>an</strong>d help prepare <strong>the</strong> cold l<strong>an</strong>ds<br />

community to interpret <strong>the</strong> 1.4 GHz brightness observations from SMOS.<br />

REFERENCES<br />

Akitaya, E, 1974. Studies on depth hoar. Contributions from <strong>the</strong> Institute of Low Temperature<br />

Science, Series A, 26:1–67.<br />

Boone A, Etchevers P, 2001. An intercomparison of three snow schemes of varying complexity<br />

coupled to <strong>the</strong> same l<strong>an</strong>d surface model: Local scale evaluation at <strong>an</strong> alpine site. J.<br />

Hydrometeor., 2: 374–394.<br />

Chung YC, Engl<strong>an</strong>d, AW, 2005. A coupled soil–snow–atmosphere tr<strong>an</strong>sfer model. Proc. of <strong>the</strong><br />

IEEE International Geoscience <strong>an</strong>d Remote Sensing Symposium, IGARSS'05, Seoul. `<br />

Chung YC, Engl<strong>an</strong>d AW, De Roo RD, Weininger E, 2006(submitted). Effects of vegetation <strong>an</strong>d of<br />

heat <strong>an</strong>d vapor fluxes from soil on snowpack evolution <strong>an</strong>d radiobrightness. Proc. of <strong>the</strong> IEEE<br />

International Geoscience <strong>an</strong>d Remote Sensing Symposium, IGARSS'06, Denver, CO.<br />

Cline D (Chair, Cold L<strong>an</strong>d Processes Working Group), Armstrong R, Davis R, Elder K, Liston G.,<br />

2001. NASA Cold L<strong>an</strong>d Processes Field Experiment Pl<strong>an</strong> Home<br />

Page.<br />

Cline D, Armstrong R, Davis R, Elder K, <strong>an</strong>d Liston G., 2002, Updated July 2004. CLPX-Ground:<br />

ISA <strong>Snow</strong> Pit Me<strong>as</strong>urements. Edited by M. Parsons <strong>an</strong>d M.J. Brodzik., Boulder, CO: National<br />

<strong>Snow</strong> <strong>an</strong>d Ice Data Center. Digital Media,<br />

Engl<strong>an</strong>d AW, 2003. CLPX-Ground: Micrometeorological Data at <strong>the</strong> Local Scale Observation<br />

Site (LSOS). Boulder, CO: National <strong>Snow</strong> <strong>an</strong>d Ice Data Center. Digital Media.<br />

Graf T, Koike T, Fujii H, Brodzik M, Armstrong R, 2003. CLPX-Ground: Ground B<strong>as</strong>ed P<strong>as</strong>sive<br />

Microwave Radiometer (GBMR-7) Data. Boulder, CO: National <strong>Snow</strong> <strong>an</strong>d Ice Data Center.<br />

Digital Media.<br />

Flerchinger GN, H<strong>an</strong>son CL, 1989. Modeling soil freezing <strong>an</strong>d thawing on a r<strong>an</strong>gel<strong>an</strong>d watershed.<br />

Tr<strong>an</strong>s. Amer. Soc. of Agric. Engr., 32(5): 1551–1554.<br />

Hardy J, Melloh R, Koenig G, Pomeroy J, Rowl<strong>an</strong>ds A, Cline D, Elder K, Davis R, 2002, updated<br />

2004. Sub-c<strong>an</strong>opy energetics at <strong>the</strong> CLPX Local Scale Observation Site (LSOS). Boulder, CO:<br />

National <strong>Snow</strong> <strong>an</strong>d Ice Data Center. Digital Media.<br />

14


Jord<strong>an</strong> R, 1991. A one-dimensional temperature model for a snow cover. Technical documentation<br />

for SNTHERM.89, Special Technical Report 91-16, US Army CRREL.<br />

Jord<strong>an</strong> R, Andre<strong>as</strong> E, 1999. Heat budget of snow-covered sea ice at North Pole 4. J. Geophys.<br />

Res., 104(C4): 7785–7806.<br />

Judge J, 1999. L<strong>an</strong>d surface process <strong>an</strong>d radiobrightness modeling of <strong>the</strong> Great Plains, Ph.D.<br />

<strong>the</strong>sis, University of Michig<strong>an</strong>.<br />

Lehning et al., 1998. A network of automatic wea<strong>the</strong>r <strong>an</strong>d snow stations <strong>an</strong>d supplementary model<br />

calculations providing SNOWPACK information for aval<strong>an</strong>che warning, ISSW 98 International<br />

<strong>Snow</strong> Science Workshop, Sunriver, Oregon.<br />

Liou, YA, 1996. L<strong>an</strong>d surface process/ radiobrightness models for nor<strong>the</strong>rn prairie, Ph.D. <strong>the</strong>sis,<br />

University of Michig<strong>an</strong>.<br />

Su, GW, Geller, JT, Pruess, K, Wen, F, 1999. Experimental studies of water seepage <strong>an</strong>d<br />

intermittent flow in unsaturated, rough-walled fractures,” Water Resources Research, 35:<br />

1019–1037.<br />

Tao, Y.-X. <strong>an</strong>d Gray, D.M.. 1994. Prediction of <strong>Snow</strong>-Melt Infiltration into Frozen Soils. Numer.<br />

Heat Tr<strong>an</strong>sfer Part A, 26: 643–645.<br />

Waldner, PA, Schneebeli, M, Zimmerm<strong>an</strong>n, US, Fluhler, H, 2004. Effect of snow structure on<br />

water flow <strong>an</strong>d solute tr<strong>an</strong>sport. Hydrological Processes, 18: 1271–1290.<br />

Y<strong>an</strong>g, ZL, 2004(in press). Description of recent snow models. Book Chapter, in <strong>Snow</strong> <strong>an</strong>d<br />

Climate, E. Martin <strong>an</strong>d R. Armstrong (editors), International Committee on <strong>Snow</strong> <strong>an</strong>d Ice.<br />

Zh<strong>an</strong>g, T, Osterkamp, TE, Stamnes K, 1997. Effects of Climate on <strong>the</strong> Active Layer <strong>an</strong>d<br />

Permafrost on <strong>the</strong> North Slope of Al<strong>as</strong>ka, U.S.A. Permafrost <strong>an</strong>d Periglacial Processes, 8: 45–<br />

67.<br />

15


This page is intentionally bl<strong>an</strong>k.<br />

16


Campbell Scientific Award<br />

for<br />

Best C<strong>an</strong>adi<strong>an</strong> Student Paper<br />

17


This page is intentionally bl<strong>an</strong>k.<br />

18


19<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Potential of a Water Bal<strong>an</strong>ce Model with High Temporal<br />

Resolution for <strong>the</strong> Distributed Modelling of Ice- <strong>an</strong>d <strong>Snow</strong>melt<br />

Processes at High Elevated Sites<br />

ABSTRACT<br />

GERNOT KOBOLTSCHNIG 1 , HUBERT HOLZMANN 1 ,<br />

WOLFGANG SCHOENER 2 , AND MASSIMILIANO ZAPPA 3<br />

The potential of <strong>the</strong> distributed hydrological water bal<strong>an</strong>ce model PREVAH at <strong>the</strong> small, highly<br />

glacierized catchment area of Goldbergkees in <strong>the</strong> Austri<strong>an</strong> Alps h<strong>as</strong> been investigated. The model<br />

is driven by meteorological data from <strong>the</strong> observatory at Hoher Sonnblick, situated at <strong>the</strong> highest<br />

point of <strong>the</strong> catchment area. A dense network of field observations <strong>as</strong> additional input <strong>an</strong>d<br />

validation data h<strong>as</strong> been applied. In <strong>the</strong> final setting PREVAH h<strong>as</strong> been run in <strong>an</strong> hourly time step<br />

b<strong>as</strong>ed on 722 hydrological response units covering <strong>the</strong> catchment area. Both snow- <strong>an</strong>d icemelt<br />

have been simulated by me<strong>an</strong>s of <strong>an</strong> adv<strong>an</strong>ced air temperature-index b<strong>as</strong>ed approach taking<br />

potential direct solar radiation into account. A multi-validation approach using discharge<br />

hydrographs, me<strong>as</strong>ured snow water equivalent data (SWE), snow cover patterns derived from<br />

satellite data, <strong>an</strong>d glacier m<strong>as</strong>s bal<strong>an</strong>ce investigations have been applied to validate <strong>the</strong> water<br />

bal<strong>an</strong>ce of <strong>the</strong> hydrological year of 2004/2005. The comparison of modelled SWE with spatially<br />

dense SWE me<strong>as</strong>urements at four different dates within <strong>the</strong> period May to July 2005 shows quite<br />

good accord<strong>an</strong>ce for both individual elevation b<strong>an</strong>ds <strong>an</strong>d <strong>the</strong> <strong>entire</strong> catchment. The period of<br />

2003/2004 h<strong>as</strong> been used for cross-validating <strong>the</strong> model for discharge-hydrograph <strong>an</strong>d ice melt.<br />

Icemelt <strong>an</strong>d maximum snow accumulation have been validated against glacier m<strong>as</strong>s bal<strong>an</strong>ce<br />

me<strong>as</strong>urements. The individual components contributing to runoff such <strong>as</strong> rainfall, snow- <strong>an</strong>d<br />

icemelt have been separated for <strong>the</strong> hydrological year 2004/05 to estimate <strong>the</strong>ir fraction to total<br />

discharge which is 3.8% for rain, 86.8% for snow, <strong>an</strong>d 9.4% for ice respectively. Finally<br />

recommendations are given for a possible improvement of hydrologic models considering snow-<br />

<strong>an</strong>d icemelt at high elevated sites.<br />

Keywords: glacier melt; snowmelt; alpine hydrology; water bal<strong>an</strong>ce of high elevated sites; SWE<br />

investigation<br />

INTRODUCTION<br />

According to <strong>the</strong> location, elevation <strong>an</strong>d topography <strong>the</strong> contribution to runoff from glacier melt,<br />

snowmelt <strong>an</strong>d rainfall at glacierized, alpine watersheds varies strongly depending on climate<br />

1<br />

University of Natural Resources <strong>an</strong>d Applied Life Sciences (BOKU), Muthg<strong>as</strong>se 18, A-1190<br />

Vienna, Austria. E-mail: gernot.koboltschnig@boku.ac.at<br />

2<br />

Central Institute of Meteorology <strong>an</strong>d Geodynamics (ZAMG), Hohewarte 38, A-1190 Vienna,<br />

Austria<br />

3<br />

Swiss Federal Institute for Forest, <strong>Snow</strong> <strong>an</strong>d L<strong>an</strong>dscape Research (WSL), Zürcherstr<strong>as</strong>se 111,<br />

CH-8903 Birmensdorf, Switzerl<strong>an</strong>d


20<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

conditions (Singh <strong>an</strong>d Bengtsson, 2005). For calculating <strong>the</strong> water bal<strong>an</strong>ce of glacierized<br />

watersheds, meteorological data have to be obtained for <strong>the</strong> <strong>entire</strong> hydrological year to fully cover<br />

<strong>the</strong> process of snow accumulation during winter period <strong>as</strong> well <strong>as</strong> <strong>the</strong> snow- <strong>an</strong>d icemelt during<br />

summer period. Icemelt depends on snow depletion of <strong>the</strong> glacier <strong>an</strong>d <strong>the</strong> area of bare ice exposed<br />

to melt. Consequently, <strong>the</strong> knowledge of <strong>the</strong> distributed snow accumulation is of high import<strong>an</strong>ce<br />

to simulate <strong>the</strong> snow line retreat at <strong>the</strong> catchment area (Blöschl et al., 1991). The numerical<br />

modelling of all of <strong>the</strong> components contributing to runoff <strong>an</strong>d <strong>the</strong>ir superposition requires tools for<br />

<strong>the</strong> <strong>as</strong>sessment of spatially interpolated meteorological variables, <strong>as</strong> well <strong>as</strong> for <strong>the</strong> treatment of<br />

distributed geographical variables (Kirnbauer et al., 1994; Gurtz et al., 2003). Due to <strong>the</strong><br />

complexity of different hydrological processes at high elevated sites a multi-validation of<br />

simulation results is needed (e.g. Günter et al., 1999). Verbunt et al. (2003) recommended using<br />

discharge hydrographs, water bal<strong>an</strong>ce elements, snow water equivalent data or soil moisture<br />

values. High elevated sites with sparse hydro-meteorological observations make model<br />

<strong>as</strong>sumptions e.g. temperature-index models using wide available <strong>an</strong>d e<strong>as</strong>y to spatially interpolate<br />

air temperature data necessary. The use of temperature-index models already h<strong>as</strong> a long history.<br />

Such a relation w<strong>as</strong> first applied for <strong>an</strong> Alpine glacier by Finsterwalder <strong>an</strong>d Schunk (1887). Hock<br />

(1999) <strong>an</strong>d Pellicciotti et al. (2005) adapted <strong>the</strong> cl<strong>as</strong>sical degree-day method for glacierized are<strong>as</strong><br />

linking air temperature to global radiation respectively potential clear-sky direct solar radiation to<br />

improve <strong>the</strong> diurnal cycles of melt.<br />

The physical b<strong>as</strong>is of temperature-b<strong>as</strong>ed melt-index methods is demonstrated by Ohmura<br />

(2002). The paper concluded that <strong>the</strong> longwave atmospheric radiation is <strong>the</strong> most domin<strong>an</strong>t heat<br />

source, <strong>an</strong>d <strong>the</strong> majority of <strong>the</strong> atmospheric radiation received at <strong>the</strong> surface comes from <strong>the</strong> nearsurface<br />

layer of <strong>the</strong> atmosphere. Hence, <strong>the</strong>re is a good, physical b<strong>as</strong>ed relation between air<br />

temperature <strong>an</strong>d melt.<br />

M<strong>an</strong>y studies have been presented for <strong>the</strong> simulation of snowmelt, icemelt or both for mesoscale<br />

b<strong>as</strong>ins with physically b<strong>as</strong>ed (Cline et al., 1998; Marks et al, 1999; Lehning et al., 2006) <strong>an</strong>d<br />

conceptual approaches (Schaefli et al., 2005; Klok et al. 2001; Verbunt et al. 2003), but only few<br />

studies tested <strong>the</strong> applicability of such models for very small b<strong>as</strong>ins (Arnold et al., 1996 & 1998).<br />

Zappa et al. (2003) introduced a study comparing different temperature-index models <strong>an</strong>d <strong>an</strong><br />

energy bal<strong>an</strong>ce model for <strong>the</strong> snowmelt modelling of <strong>an</strong> alpine catchment. They showed that <strong>the</strong><br />

improved temperature-index model (according to Hock, 1999) performed better th<strong>an</strong> <strong>the</strong> energy<br />

bal<strong>an</strong>ce approach implemented in PREVAH.<br />

Thus, <strong>the</strong> main goal of this study h<strong>as</strong> been to <strong>as</strong>sess <strong>the</strong> potential of a distributed hydrological<br />

modelling in a highly glacierized, small, <strong>an</strong>d high elevated b<strong>as</strong>in. For this we simulated <strong>the</strong> water<br />

bal<strong>an</strong>ce for <strong>the</strong> hydrological year 2004/2005 of <strong>the</strong> Goldbergkees watershed (Austria) with <strong>the</strong><br />

conceptual distributed hydrological model PREVAH (Precipitation–Runoff–Evapotr<strong>an</strong>spiration–<br />

HRU model, Gurtz et al., 1999). For validation we compared <strong>the</strong> computed results with maps of<br />

<strong>the</strong> snow water equivalent derived from detailed field campaigns (similar to Elder et al. 1998),<br />

with a remotely sensed map of <strong>the</strong> snow distribution (<strong>as</strong> e.g. Cline et al., 1998; Blöschl et al.,<br />

2002), with observed data of <strong>the</strong> glacier m<strong>as</strong>s bal<strong>an</strong>ce (<strong>as</strong> e.g. Schaefli et al., 2005) <strong>an</strong>d with runoff<br />

data gauged during <strong>the</strong> summer <strong>an</strong>d fall at <strong>the</strong> b<strong>as</strong>in outlet.<br />

STUDY REGION<br />

Glacier Goldbergkees is situated directly beneath <strong>the</strong> Hoher Sonnblick observatory (3106<br />

m a.s.l., 47°03’16” N, 12°57’25” E) in <strong>the</strong> central part of <strong>the</strong> Austri<strong>an</strong> Alps (Figure 1). Hoher<br />

Sonnblick is <strong>the</strong> oldest <strong>an</strong>d highest, perm<strong>an</strong>ently staffed observatory in <strong>the</strong> Alps above 3000 m<br />

a.s.l.. Meteorological observations at <strong>the</strong> observatory are available back to 1886, detailed m<strong>as</strong>s<br />

bal<strong>an</strong>ce me<strong>as</strong>urements at <strong>the</strong> nearby glaciers started in 1983 (Auer et al., 2002), <strong>an</strong>d detailed<br />

hydrological investigations are carried out since 2002. Hence a wide r<strong>an</strong>ge of hydrological,<br />

meteorological <strong>an</strong>d glaciological observations are available for detailed melt runoff <strong>an</strong>d water<br />

bal<strong>an</strong>ce modelling. The Goldbergkees watershed h<strong>as</strong> <strong>an</strong> area of about 2.72 km² with a about 52 %<br />

glacierized (1.43 km² computed for 2003). Elevations r<strong>an</strong>ges between 2350 <strong>an</strong>d 3106 m a.s.l.. The


21<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

catchment area is well defined by <strong>an</strong> automatic discharge gauging station at about 250 meters<br />

dist<strong>an</strong>ce downstream <strong>the</strong> glacier tongue situated at <strong>the</strong> outlet of a lake (Figure 1). The solid rock<br />

river bed <strong>an</strong>d nearly laminar flow guar<strong>an</strong>tees reliable runoff gauging. The area is structured into<br />

three major topographical sections: The upper part comprises south <strong>an</strong>d sou<strong>the</strong><strong>as</strong>t facing slopes,<br />

<strong>the</strong> middle part faces e<strong>as</strong>t <strong>an</strong>d nor<strong>the</strong><strong>as</strong>t, <strong>an</strong>d <strong>the</strong> lower part comprises <strong>the</strong> tongue of <strong>the</strong> glacier<br />

which faces north <strong>an</strong>d north e<strong>as</strong>t. All parts of <strong>the</strong> catchment are above <strong>the</strong> timber line. The<br />

domin<strong>an</strong>t l<strong>an</strong>d cover is rock (central alpine gneiss), gravel <strong>an</strong>d ice (unpublished data from<br />

Koboltschnig et al., 2006). The me<strong>an</strong> air temperature at Sonnblick observatory is about –5.7°C.<br />

The <strong>an</strong>nual precipitation at Sonnblick observatory averages about 2680 mm, with 89% <strong>as</strong> snow<br />

(climate normals 1961–1990, Auer et al., 2002).<br />

Figure 1. Catchment area of Goldbergkees. Numbers from 1 to 5 indicate hydro-meteorological stations: 1<br />

observatory at <strong>the</strong> top Hoher Sonnblick; 2 air temperature station; 3 air temperature station; 4 discharge<br />

gauge, air temperature station, <strong>an</strong>d temporary tipping bucket; 5 automatic ultra sonic snow depth<br />

me<strong>as</strong>urement<br />

METHODS<br />

Field investigations <strong>an</strong>d meteorological network<br />

Hourly data of precipitation, air temperature, moisture, wind speed, sunshine duration <strong>an</strong>d<br />

global radiation were taken from <strong>the</strong> observatory at <strong>the</strong> top of Hoher Sonnblick (Figure 1, Nr. 1).


22<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Precipitation <strong>an</strong>d air temperature data observed at <strong>the</strong> observatory are shown in Figure 8 (e <strong>an</strong>d f).<br />

Three additional air temperature stations were installed during 2005 melt se<strong>as</strong>on in <strong>an</strong>d next to <strong>the</strong><br />

catchment area (Figure 1, Nr. 2, 3 <strong>an</strong>d 4). A temporary tipping bucket, for liquid precipitation<br />

me<strong>as</strong>urements only, w<strong>as</strong> installed at <strong>the</strong> catchment outlet (Figure 1, Nr. 4). The tipping bucket w<strong>as</strong><br />

fixed toge<strong>the</strong>r with <strong>the</strong> temperature station <strong>an</strong>d <strong>the</strong> data logger of <strong>the</strong> discharge gauge. Water<br />

levels could only be recorded between July <strong>an</strong>d October at a natural cross section of a lake outlet.<br />

To prevent damage <strong>the</strong> instrumentation h<strong>as</strong> to be removed during winter period. The rating curves<br />

were calculated for every melt se<strong>as</strong>on separately due to <strong>the</strong> ch<strong>an</strong>ges of <strong>the</strong> hydraulic conditions.<br />

<strong>Snow</strong> depths have been me<strong>as</strong>ured automatically in a daily interval at about hundred meters of<br />

elevation beneath <strong>the</strong> observatory (see Figure 1, Nr. 5). Starting in May 2005 four field campaigns<br />

for mapping <strong>the</strong> SWE (snow water equivalent) of <strong>the</strong> catchment area were performed in monthly<br />

steps. Aluminium probes were used to me<strong>as</strong>ure <strong>the</strong> snow depth at about 60 to 140 points<br />

irregularly distributed over <strong>the</strong> <strong>entire</strong> catchment area. The snow density h<strong>as</strong> been me<strong>as</strong>ured at two<br />

snow pits following <strong>the</strong> instructions of K<strong>as</strong>er et al. (2003). During <strong>the</strong> earliest campaign in May<br />

eight pits have been dug with regard to <strong>the</strong> higher variability of <strong>the</strong> snow layers at that time<br />

(Figure 1, grey tri<strong>an</strong>gles). From this very detailed data set we interpolated distributed maps of<br />

SWE using a spline method to generate 10 m grids of snow depth. Additional sampling points at<br />

are<strong>as</strong> with no snow cover have been set to ensure better interpolation results at border are<strong>as</strong>. The<br />

me<strong>as</strong>ured snow density data have been interpolated using inverse dist<strong>an</strong>ce weighting technique.<br />

Finally, SWE were computed from spatialized snow depth <strong>an</strong>d snow density. <strong>Snow</strong> free are<strong>as</strong><br />

were mapped in field using GPS. The maps of snow free are<strong>as</strong> have been overlaid to generate <strong>the</strong><br />

final depletion maps. This unique set of SWE grid h<strong>as</strong> been available for visual <strong>an</strong>d qu<strong>an</strong>titative<br />

comparison with <strong>the</strong> computation of <strong>the</strong> hydrological model (Figure 2).<br />

As a st<strong>an</strong>dard program of <strong>the</strong> m<strong>as</strong>s bal<strong>an</strong>ce me<strong>as</strong>urements following <strong>the</strong> glaciological method<br />

(Hoinkes, 1970; Østrem <strong>an</strong>d Brugm<strong>an</strong>, 1991; K<strong>as</strong>er et al., 2003) <strong>the</strong> ice ablation have been<br />

me<strong>as</strong>ured using ablation stakes, which have been drilled into <strong>the</strong> bare ice of <strong>the</strong> glacier. Using 17<br />

ablation stakes distributed over <strong>the</strong> <strong>entire</strong> ablation area of glacier Goldbergkees <strong>the</strong> net ablation of<br />

<strong>the</strong> glacier h<strong>as</strong> been calculated (Hynek <strong>an</strong>d Schöner, 2004).


23<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Figure 2. Simulated vs. observed distributed SWE data for different dates in 2004 <strong>an</strong>d 2005. Black are<strong>as</strong><br />

indicate snow free are<strong>as</strong>. Mentioned values of SWE me<strong>an</strong> are distributed SWE values averaged over <strong>the</strong><br />

catchment area (mm)<br />

The hydrological model PREVAH<br />

The spatially distributed hydrological model PREVAH (Precipitation–Runoff–<br />

Evapotr<strong>an</strong>spiration–HRU model, Gurtz et al., 1999) h<strong>as</strong> been used to simulate <strong>the</strong> processes<br />

contributing to runoff. PREVAH h<strong>as</strong> already been used at glacierized sites at different spatial<br />

resolution (Badoux, 1999; Gurtz et al., 2003; Zappa et al., 2003; Zappa et al., 2000). The<br />

catchment area is subdivided into HRU’s (hydrological response units, Ross et al., 1979) b<strong>as</strong>ed on<br />

DEM <strong>an</strong>d l<strong>an</strong>d use data. For every HRU <strong>the</strong> hydrological response to <strong>the</strong> meteorological input is<br />

simulated using a typical storage c<strong>as</strong>cade approach. The runoff contributions of all units are added<br />

to provide <strong>the</strong> total discharge at <strong>the</strong> outlet of <strong>the</strong> <strong>entire</strong> catchment. Due to <strong>the</strong> small size of <strong>the</strong><br />

catchment <strong>an</strong>d steep slopes no routing between <strong>the</strong> spatial units <strong>an</strong>d <strong>the</strong> river outlet is carried out.


24<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

The runoff delay caused by <strong>the</strong> different glacial reservoirs is m<strong>an</strong>aged by <strong>the</strong> main model storages<br />

for <strong>the</strong> melt simulations. This are <strong>the</strong> snow-, ice- <strong>an</strong>d firn storage, parameterized using <strong>the</strong> storage<br />

coefficient <strong>an</strong>d <strong>the</strong> tr<strong>an</strong>slation time (Badoux, 1999). PREVAH uses different approaches for <strong>the</strong><br />

snow <strong>an</strong>d icemelt simulations. For this study <strong>the</strong> radiation b<strong>as</strong>ed temperature-index approach of<br />

Hock (1999) h<strong>as</strong> been applied:<br />

⎧1<br />

⎪ (MF<br />

M = ⎨n<br />

⎪<br />

⎩<br />

snow / ice<br />

+ a<br />

0<br />

snow / ice<br />

⋅I<br />

) ⋅T<br />

:<br />

:<br />

T > 0<br />

T ≤ 0<br />

where M is <strong>the</strong> calculated melt rate (mm h –1 ), MFsnow / ice is <strong>the</strong> melt factor for snow respectively<br />

ice (mm d –1 K –1 ), <strong>as</strong>now / ice is <strong>the</strong> radiation melt factor for snow respectively ice, I is <strong>the</strong> potential<br />

clear-sky direct solar radiation at ice or snow surface (W m –2 ), T is <strong>the</strong> air temperature (°C) <strong>an</strong>d n<br />

is <strong>the</strong> number of time-steps per day, in this c<strong>as</strong>e 24 hours. The melt factors for snow <strong>an</strong>d ice are<br />

empirical coefficients <strong>an</strong>d I is a calculated value following Hock (1999).<br />

Only air temperature is needed <strong>as</strong> input to <strong>the</strong> model, radiation is calculated <strong>as</strong> <strong>the</strong> site adjusted<br />

potential direct radiation for every HRU, considering exposition <strong>an</strong>d slope. This approach h<strong>as</strong><br />

shown <strong>the</strong> best results in <strong>the</strong> comparative study of Zappa et al. (2003), where <strong>the</strong> perform<strong>an</strong>ce of<br />

<strong>an</strong> energy bal<strong>an</strong>ce model <strong>an</strong>d three degree-day-factor models have been investigated for a mesoscale<br />

b<strong>as</strong>in.<br />

For <strong>the</strong> calculation of <strong>the</strong> snow accumulation, <strong>the</strong> simulation of <strong>the</strong> surface runoff, <strong>the</strong><br />

calculation of snow- <strong>an</strong>d icemelt, <strong>an</strong>d <strong>the</strong> calculation of <strong>the</strong> evapotr<strong>an</strong>spiration <strong>the</strong> following input<br />

to <strong>the</strong> PREVAH model is required: air temperature, precipitation, water vapour pressure, global<br />

radiation, wind speed <strong>an</strong>d sunshine duration.<br />

Model application at Goldbergkees watershed<br />

As a major source for <strong>the</strong> description of <strong>the</strong> topography to <strong>the</strong> modelling system a DEM of 10 m<br />

resolution (Auer et al., 2002), covering <strong>the</strong> <strong>entire</strong> catchment area, h<strong>as</strong> been applied. Such high<br />

resolution is required to account for <strong>the</strong> small scale variability of <strong>the</strong> investigated processes for<br />

such a small b<strong>as</strong>in. The separation into HRU’s h<strong>as</strong> been realized using two l<strong>an</strong>d use cl<strong>as</strong>ses<br />

(glacier <strong>an</strong>d rock), 50 m elevation b<strong>an</strong>ds (16 cl<strong>as</strong>ses), nine <strong>as</strong>pect cl<strong>as</strong>ses <strong>an</strong>d six slope cl<strong>as</strong>ses.<br />

Thus 722 HRU’s <strong>an</strong>d 197 meteorological units (MU) were generated. MU are <strong>the</strong> spatial units<br />

covering <strong>the</strong> watershed for which <strong>the</strong> meteorological data have been interpolated b<strong>as</strong>ed on hourly<br />

data from available stations (Figure 1, stations 1, 2, 3, <strong>an</strong>d 4). Because elevation <strong>an</strong>d exposition<br />

are <strong>the</strong> main factors governing climatological <strong>an</strong>d meteorological variability in such are<strong>as</strong>, r<strong>as</strong>ter<br />

elements in <strong>the</strong> same elevation b<strong>an</strong>d <strong>an</strong>d showing similar <strong>as</strong>pect belong to <strong>the</strong> same MU. The<br />

PREVAH model h<strong>as</strong> been run in <strong>an</strong> hourly time step. For <strong>the</strong> interpolation of <strong>the</strong> meteorological<br />

input <strong>an</strong> inverse dist<strong>an</strong>ce weighting <strong>an</strong>d altitude-dependent regression approach h<strong>as</strong> been used<br />

(Klok et al. 2001). Precipitation h<strong>as</strong> been perm<strong>an</strong>ently observed at <strong>the</strong> observatory <strong>an</strong>d temporary<br />

at <strong>the</strong> catchment outlet (Figure 1, stations 1 <strong>an</strong>d 4). For a better weighting two additional virtual<br />

stations at <strong>the</strong> sites of <strong>the</strong> stations 2 <strong>an</strong>d 3 (Figure 1) have been applied, where monthly me<strong>as</strong>ured<br />

precipitation sums have been available. The laps rate for <strong>the</strong> air temperature h<strong>as</strong> been set to<br />

0.65°K/100 m. This value h<strong>as</strong> been estimated using long term air temperature me<strong>as</strong>urements of<br />

two stations in <strong>the</strong> area. The air temperature input for <strong>the</strong> melt modelling is taken from 3 stations<br />

outside <strong>the</strong> glacier, which is proposed by L<strong>an</strong>g <strong>an</strong>d Braun (1990).<br />

The simulation of one <strong>entire</strong> hydrological year with PREVAH, producing tables of hourly<br />

runoff <strong>an</strong>d storage output, <strong>an</strong>d daily maps of SWE takes less th<strong>an</strong> 3 minutes computation time<br />

using a st<strong>an</strong>dard PC. We proceeded <strong>the</strong>refore with m<strong>an</strong>ual tuning of <strong>the</strong> parameters governing <strong>the</strong><br />

processes of snow accumulation, snowmelt, icemelt <strong>an</strong>d runoff generation (Zappa et al., 2003;<br />

Gurtz et al., 2003) for <strong>the</strong> hydrological year 2004/2005, where we have additional observations on<br />

snow cover <strong>an</strong>d SWE for multi-criteria verification. The verification for <strong>the</strong> discharge simulation<br />

(1)


25<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

h<strong>as</strong> been carried out by <strong>the</strong> <strong>an</strong>alysis of <strong>the</strong> model perform<strong>an</strong>ce for <strong>the</strong> hydrological year<br />

2003/2004.<br />

Calibration procedure<br />

First step: Due to <strong>the</strong> extreme climate conditions during 2003 melt se<strong>as</strong>on <strong>the</strong> catchment area of<br />

Goldbergkees glacier w<strong>as</strong> nearly snow free. Hence it h<strong>as</strong> been possible to calibrate <strong>the</strong> melt<br />

parameters for icemelt separated from o<strong>the</strong>r processes. Therefore <strong>the</strong> observed <strong>an</strong>d simulated<br />

discharge hydrographs have been compared;<br />

Second step: The model h<strong>as</strong> been initialized using spatially distributed SWE data at <strong>the</strong> time of<br />

<strong>the</strong> maximum snow accumulation (begin of May). Thus <strong>the</strong> degree day factor for snowmelt h<strong>as</strong><br />

been calibrated comparing <strong>the</strong> daily results of <strong>the</strong> internal snow storage with <strong>the</strong> in field observed<br />

spatially distributed SWE data at different points in time;<br />

Third step: The model h<strong>as</strong> been initialized using spatially distributed SWE data at <strong>the</strong> beginning<br />

of <strong>the</strong> accumulation period (equal to <strong>the</strong> beginning of <strong>the</strong> hydrological year in <strong>the</strong> nor<strong>the</strong>rn<br />

hemisphere at <strong>the</strong> beginning of October). For this re<strong>as</strong>on <strong>the</strong> solid precipitation h<strong>as</strong> been corrected<br />

by plus 18 % more precipitation following Sevruk (1986). The snow accumulation of <strong>the</strong><br />

PREVAH model h<strong>as</strong> been compared with <strong>the</strong> in field observed spatially distributed SWE data at<br />

<strong>the</strong> time of <strong>the</strong> maximum accumulation;<br />

Fourth step: Fine tuning of <strong>the</strong> melt parameters <strong>an</strong>d adaptation of <strong>the</strong> storage parameters<br />

comparing observed <strong>an</strong>d simulated hydrographs. The storage time for snow- <strong>an</strong>d icemelt are<br />

calibrated for <strong>the</strong> Goldbergkees catchment adducting <strong>the</strong> recession curves of <strong>the</strong> observed<br />

hydrograph, which result when summer snowfalls reduce ablation by raising albedo (Collins,<br />

1982). Three events of this kind have been observed at 2005 melt se<strong>as</strong>on (see Figure 4). For <strong>the</strong><br />

determination of <strong>the</strong> tr<strong>an</strong>slation time of snow- <strong>an</strong>d icemelt <strong>the</strong> diurnal maximums of <strong>the</strong> simulated<br />

runoff have been fitted to <strong>the</strong> observed hydrograph.<br />

Figure 4. Simulated vs. observed snow cover pattern at 29 July 2005. The observed image is generated by a<br />

cl<strong>as</strong>sification of <strong>an</strong> ASTER image. Black coloured are<strong>as</strong> indicate snow free are<strong>as</strong>, white are<strong>as</strong> are still snow<br />

covered.<br />

Verification procedure<br />

First step: <strong>the</strong> distributed SWE simulation h<strong>as</strong> been verified at 4 dates of <strong>the</strong> 2004/2005 period;<br />

Second step: <strong>the</strong> simulation of <strong>the</strong> snow patterns h<strong>as</strong> been verified with satellite data from<br />

ASTER;<br />

Third step: <strong>the</strong> skill of <strong>the</strong> runoff simulation h<strong>as</strong> been independently verified without fur<strong>the</strong>r<br />

adjustment of <strong>the</strong> calibrated parameters by comparisons with <strong>the</strong> hydrological year 2003/2004;<br />

Fourth step: observed data on glacier ablation for both 2003/2004 <strong>an</strong>d 2004/2005 have been<br />

compared with <strong>the</strong> simulated ice ablation.


26<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Main model settings<br />

For simulating <strong>the</strong> hydrological year of 2004/2005 <strong>the</strong> calibrated model h<strong>as</strong> been initialized at<br />

<strong>the</strong> beginning of October 2004 by in field observed SWE data. The main calibrated parameters of<br />

<strong>the</strong> model are shown in Table 1 in comparison to similar studies (Pellicciotti et al., 2005; Hock,<br />

1999; Zappa et al., 2003). The melt factors presented in this paper are of <strong>the</strong> same size <strong>as</strong> <strong>the</strong><br />

compared values. The degree-day-factor for ice is slightly higher, which could be explained by <strong>the</strong><br />

dark ice surface <strong>an</strong>d <strong>the</strong>refore quite low albedo of Goldbergkees glacier. The snowmelt model is<br />

parameterized by <strong>the</strong> radiation melt factor <strong>an</strong>d <strong>the</strong> maximum <strong>an</strong>d <strong>the</strong> minimum temperature<br />

dependent melt factors, which define <strong>the</strong> maximum at 21 st of June <strong>an</strong>d <strong>the</strong> minimum at 21 st of<br />

December of a sinus shaped function. The temperature dependent melt factor for ice is const<strong>an</strong>t in<br />

time.<br />

Parameter<br />

Table 1. Comparison of <strong>the</strong> model parameters calibrated at Goldbergkees watershed <strong>an</strong>d <strong>the</strong><br />

parameters of <strong>the</strong> studies of Pellicciotti et al. (2005), Hock (1999) <strong>an</strong>d Zappa et al. (2003)<br />

RESULTS AND DISCUSSION<br />

Cal.<br />

Value<br />

Zappa Hock Pellicciotti<br />

Threshold temperature for snowmelt 0 0 0 1 [°C]<br />

Max. degree day factor 3.2 – – –<br />

[mm d –1 K –<br />

1<br />

]<br />

Min. degree day factor 1 – – –<br />

[mm d –1 K –<br />

1<br />

]<br />

Degree day factor – 0.8 1.8 1.97<br />

[mm d –1 K –<br />

1<br />

]<br />

Radiation melt factor for snow 0.00015 0.00027 0.0008 0.00052<br />

[mm W –1<br />

m² K –1 h –1 ]<br />

Temperature melt factor for ice 2.15 – 1.8 1.97<br />

[mm d –1 K –<br />

1<br />

]<br />

Radiation melt factor for ice 0.0003 – 0.0006 0.00106<br />

[mm W –1<br />

m² K –1 h –1 ]<br />

Storage time for snowmelt 25 – – – [h]<br />

Storage time for icemelt 2 – – – [h]<br />

Tr<strong>an</strong>slation time for snowmelt 3 – – – [h]<br />

Tr<strong>an</strong>slation time for icemelt 2 – – – [h]<br />

Distributed snow-water-equivalent (SWE)<br />

The PREVAH model makes a daily output of <strong>the</strong> distributed SWE storage possible. Thus <strong>the</strong><br />

validation using observed SWE data is <strong>an</strong> additional alternative. Figure 2 shows <strong>the</strong> validation of<br />

<strong>the</strong> SWE for four different points in time starting with <strong>the</strong> initial SWE model settings of 6 October<br />

2004. The simulated plot for this date shows <strong>the</strong> me<strong>an</strong> value of <strong>the</strong> SWE storage at <strong>the</strong> end of <strong>the</strong><br />

first simulated day described for <strong>the</strong> internal distribution of <strong>the</strong> meteorological units. Through this<br />

semi-distributed approach some information on spatial distribution is lost. The simulated <strong>an</strong>d<br />

observed maximum snow accumulation at 6 May 2005 are in a very good agreement. At 2 June<br />

2005 <strong>the</strong> simulation shows a little smaller SWE storage th<strong>an</strong> <strong>the</strong> month before but <strong>the</strong> observed<br />

SWE seems to show <strong>an</strong> overestimation through <strong>the</strong> field me<strong>as</strong>urement. There is no possibility to<br />

explain <strong>the</strong> high accumulation by observed precipitation data. At 7 July 2005 still a slightly<br />

overestimation is seen. The l<strong>as</strong>t observation at 28 July 2005 shows a good accord<strong>an</strong>ce of <strong>the</strong><br />

simulated SWE, averaged over <strong>the</strong> catchment area. The simulated distribution of snow left in <strong>the</strong><br />

Unit


27<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

catchment strongly depends on <strong>the</strong> distribution of precipitation, where<strong>as</strong> <strong>the</strong> in field observed<br />

distribution of snow depends on precipitation gradients, snow redistribution by wind <strong>an</strong>d vertical<br />

drift of snow induced by aval<strong>an</strong>ches (Hartm<strong>an</strong>n et al., 1999; Blöschl et al., 1991).<br />

The spatial <strong>an</strong>alysis of <strong>the</strong> distributed observed <strong>an</strong>d simulated SWE maps is shown in Figure 3.<br />

The catchment area h<strong>as</strong> been divided into 8 100 m elevation b<strong>an</strong>ds starting at 2300 m a.s.l.. There<br />

is a very good agreement between observation <strong>an</strong>d simulation at 6 May 2005 at <strong>the</strong> elevation<br />

b<strong>an</strong>ds between 2600 <strong>an</strong>d 2900 m (Figure 3a). The upper b<strong>an</strong>ds show <strong>an</strong> overestimation <strong>an</strong>d <strong>the</strong><br />

lower b<strong>an</strong>ds <strong>an</strong> underestimation of <strong>the</strong> simulation. This result is due to <strong>the</strong> snow redistribution<br />

from <strong>the</strong> upper to <strong>the</strong> lower parts, induced by wind or aval<strong>an</strong>ches. The problem of too high snow<br />

accumulation me<strong>as</strong>ured at 2 June 2005 c<strong>an</strong> be seen in Figure 3b, where <strong>the</strong> simulated SWE is<br />

const<strong>an</strong>tly about 10% lower th<strong>an</strong> <strong>the</strong> observed, despite <strong>the</strong> two uppermost elevation b<strong>an</strong>ds, which<br />

are fine modelled. At Figure 3c <strong>an</strong>d Figure 3d it c<strong>an</strong> be seen, that <strong>the</strong> elevation b<strong>an</strong>d at 2700 to<br />

2800 m is in a good agreement <strong>an</strong>d again <strong>the</strong> upper ones are overestimated <strong>an</strong>d <strong>the</strong> lower ones are<br />

underestimated by <strong>the</strong> simulation. This result is again due to <strong>the</strong> problem of vertical snow<br />

redistribution.<br />

Figure 3. Simulated SWE (solid line, dots) vs. observed SWE (d<strong>as</strong>hed line, x) averaged over 100 m elevation<br />

b<strong>an</strong>ds at four different points in time.<br />

Figure 8b shows <strong>the</strong> observed daily snow depth at station 5 (Figure 1) <strong>an</strong>d Figure 8c shows <strong>the</strong><br />

simulated SWE averaged over <strong>the</strong> <strong>entire</strong> catchment area. The correlation coefficient calculated for<br />

<strong>the</strong> <strong>as</strong>cending ph<strong>as</strong>e is about r²=0.91, for <strong>the</strong> descending r²=0.99, <strong>an</strong>d for <strong>the</strong> <strong>entire</strong> period r²=0.92.<br />

We imply that <strong>the</strong> good correlation at <strong>the</strong> descending ph<strong>as</strong>e is due to a quite homogeneous snow<br />

density <strong>an</strong>d quite similar snow melt processes over <strong>the</strong> <strong>entire</strong> catchment area. The <strong>as</strong>cending ph<strong>as</strong>e<br />

of <strong>the</strong> observed snow depth me<strong>as</strong>urement (Figure 8b) shows typical settling effects of <strong>the</strong><br />

snowpack, which c<strong>an</strong> not be simulated by models without physically b<strong>as</strong>ed <strong>as</strong>sumptions for <strong>the</strong><br />

snowpack modelling (Lehning et al., 2006).<br />

<strong>Snow</strong> cover patterns<br />

An ASTER (L1B) image of good quality h<strong>as</strong> been available for 29 July 2005. The image h<strong>as</strong><br />

been cl<strong>as</strong>sified (unpublished data from Vollm<strong>an</strong>n, 2006) to generate a map of snow cover patterns.<br />

Figure 4 shows <strong>the</strong> comparison of <strong>the</strong> observed <strong>an</strong>d simulated snow cover map. The lower parts of<br />

<strong>the</strong> catchment area are simulated <strong>as</strong> nearly snow free. The observation shows a more complex<br />

structure of <strong>the</strong> snow line retreat, but <strong>the</strong> simulated snow cover pattern is in a good agreement with


28<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

<strong>the</strong> observed pattern. Inadequacies arise from processes not included in <strong>the</strong> model such <strong>as</strong><br />

redistribution of snow by wind <strong>an</strong>d aval<strong>an</strong>ches (Blöschl et al., 1991; Lehning et al., 2000;<br />

Doorschot et al., 2001). At such small scale <strong>an</strong>d with <strong>the</strong> adoption of such a high spatial resolution<br />

we c<strong>an</strong> see <strong>the</strong> role of snow redistribution <strong>an</strong>d underst<strong>an</strong>d how distributed models have to be<br />

improved to account for such processes.<br />

Runoff simulations<br />

For <strong>the</strong> 2005 melt se<strong>as</strong>on (calibration period) discharge data have been available for <strong>the</strong> period<br />

of 9 July–30 September (Figure 5).<br />

Figure 5. Simulated icemelt averaged over <strong>the</strong> catchment area, M (mm h –1 ), hourly data of air temperature, T<br />

(°C), <strong>an</strong>d precipitation P (mm h –1 ) at Hoher Sonnblick observatory, <strong>an</strong>d observed <strong>an</strong>d simulated discharge Q<br />

(m³ s –1 ) at <strong>the</strong> catchment outlet of Goldbergkees. The incre<strong>as</strong>ing lines at <strong>the</strong> bottom plot indicate <strong>the</strong><br />

cumulated discharge Q cum (10 6 m³) over <strong>the</strong> period of <strong>the</strong> discharge observations from 10 July to 30<br />

September 2005.<br />

The beginning of <strong>the</strong> observations shows rain-induced discharge peaks. Typical diurnal<br />

variations of <strong>the</strong> observed runoff, induced by icemelt are seen in <strong>the</strong> period from end of August<br />

until mid of September. The perform<strong>an</strong>ce of <strong>the</strong> simulation h<strong>as</strong> been employed according to <strong>the</strong><br />

efficiency criterion R NS² (N<strong>as</strong>h <strong>an</strong>d Sutcliffe, 1970), defined <strong>as</strong>:


R<br />

∑i<br />

∑<br />

n<br />

2<br />

= 1<br />

NS = − n<br />

i = 1<br />

29<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

2<br />

( Qobs<br />

− Qsim<br />

)<br />

1 (2)<br />

2<br />

( Q − Q )<br />

obs<br />

obs<br />

where Q is <strong>the</strong> hourly value of <strong>the</strong> catchment runoff (m³ s –1 ) <strong>an</strong>d <strong>the</strong> subscripts obs <strong>an</strong>d sim refer<br />

to <strong>the</strong> observed <strong>an</strong>d simulated runoff. The bar refers to <strong>the</strong> me<strong>an</strong> of <strong>the</strong> observed runoff <strong>an</strong>d n is<br />

<strong>the</strong> number of time-steps for which R NS² is calculated. The final model perform<strong>an</strong>ce accounts for<br />

R NS² = 0.77. The scatter plot in Figure 6 shows <strong>the</strong> comparison of observed <strong>an</strong>d simulated runoffs.<br />

It is seen, that low flow <strong>an</strong>d me<strong>an</strong> flow conditions are good represented but that <strong>the</strong>re are some<br />

higher values of discharges below <strong>the</strong> 45° line which have not been simulated. Higher discharges<br />

are cushioned through <strong>the</strong> quite high value of <strong>the</strong> snow melt storage time (25 h), since rainfall on<br />

<strong>the</strong> snow surface h<strong>as</strong> to go through this storage. The lowering of <strong>the</strong> snow melt storage time would<br />

effect into a steep recession curve (e.g. recession between 4th <strong>an</strong>d 5 th of August in Figure 2) <strong>an</strong>d<br />

on <strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d <strong>the</strong> lifting of peak discharges is quite low. Hence, <strong>the</strong> simulated runoff<br />

hydrograph shown in Figure 5 is <strong>the</strong> result of <strong>an</strong> optimization. The main nature of <strong>the</strong> observed<br />

hydrograph with respect to <strong>the</strong> diurnal <strong>an</strong>d se<strong>as</strong>onal fluctuations through <strong>the</strong> superposition of<br />

melting processes are very good represented by <strong>the</strong> simulation.<br />

Figure 6. Scatter plot of <strong>the</strong> observed vs. simulated hourly discharge for <strong>the</strong> period from 10 July to 30<br />

September 2005.<br />

Looking at <strong>the</strong> <strong>entire</strong> simulation period of 2004/2005 (Figure 8, d) it is seen that <strong>the</strong>re h<strong>as</strong> been a<br />

time sp<strong>an</strong> of about five month with no flow. The earliest runoff processes of <strong>the</strong> melt se<strong>as</strong>on<br />

started at <strong>the</strong> beginning of May. During <strong>the</strong> period from May to July total discharge is mainly<br />

formed by snowmelt (see Figure 8, d <strong>an</strong>d Figure 9). Peak discharges occur during June <strong>an</strong>d July,<br />

where all components contributing to runoff, despite icemelt, reach <strong>the</strong>ir maximum.<br />

For <strong>the</strong> verification se<strong>as</strong>on 2003/2004 <strong>the</strong> model h<strong>as</strong> been run using <strong>the</strong> same set of parameters<br />

<strong>as</strong> shown in Table 1. Solid precipitation h<strong>as</strong> been corrected by plus 18% like in <strong>the</strong> calibration<br />

period 2004/2005. The model perform<strong>an</strong>ce, employed according to <strong>the</strong> efficiency criterion<br />

following N<strong>as</strong>h <strong>an</strong>d Sutcliffe (1970), h<strong>as</strong> been calculated for RNS² = 0.60. Figure 7 presents <strong>the</strong><br />

model results in detail. The hydrologic response of <strong>the</strong> Goldbergkees catchment h<strong>as</strong> been different<br />

from <strong>the</strong> year 2004/2005 presented before, but <strong>the</strong> processes of snow- <strong>an</strong>d icemelt are well<br />

simulated. A slightly overestimation of <strong>the</strong> simulated cumulated runoff c<strong>an</strong> be seen. During


30<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

2003/2004 period only <strong>the</strong> meteorological input data of <strong>the</strong> observatory at Hoher Sonnblick<br />

(Figure 1, Nr. 1) h<strong>as</strong> been available for simulations. Hence, <strong>the</strong>re is a lack of input data for <strong>the</strong><br />

interpolation <strong>an</strong>d regionalisation of meteorological variables, which results in a lower model<br />

perform<strong>an</strong>ce. Never<strong>the</strong>less this shows that <strong>the</strong> calibrated model is able to simulate different<br />

climatic conditions independently from additional validation data <strong>an</strong>d re-calibration of model<br />

parameters.<br />

Figure 7. Simulated icemelt averaged over <strong>the</strong> catchment area, M (mm h –1 ), hourly data of air temperature, T<br />

(°C), <strong>an</strong>d precipitation P (mm h –1 ) at Hoher Sonnblick observatory, <strong>an</strong>d observed <strong>an</strong>d simulated discharge Q<br />

(m³ s –1 ) at <strong>the</strong> catchment outlet of Goldbergkees. The incre<strong>as</strong>ing lines at <strong>the</strong> bottom plot indicate <strong>the</strong><br />

cumulated discharge Q cum (10 6 m³) over <strong>the</strong> period of <strong>the</strong> discharge observations from 27 July to 30<br />

September 2004.


31<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Figure 8. Main hydro-meteorological observations vs. simulated hourly discharge <strong>an</strong>d daily output of <strong>the</strong><br />

SWE storage. a: daily observed depths of fresh snow (cm) <strong>an</strong>d cumulated depths of fresh snow (maximum at<br />

1787 cm) over <strong>the</strong> period 1 October 2004 to 30 September 2005 at <strong>the</strong> automated ultr<strong>as</strong>onic station; b: daily<br />

me<strong>as</strong>urements of <strong>the</strong> snow depth (cm) at <strong>the</strong> automated ultr<strong>as</strong>onic station; c: simulated SWE storage (mm); d:<br />

simulated hourly discharge (m³ s –1 ); e: hourly air temperature me<strong>as</strong>urements (°C) at Hoher Sonnblick<br />

observatory; f: hourly precipitation (mm) at Hoher Sonnblick observatory.


32<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Water bal<strong>an</strong>ce<br />

The water bal<strong>an</strong>ce, <strong>as</strong> a final model result, states <strong>an</strong> additional possibility for <strong>the</strong> model<br />

validation. As <strong>the</strong> model generates output for all water bal<strong>an</strong>ce components using e<strong>as</strong>y bal<strong>an</strong>ce<br />

equations <strong>the</strong> results of <strong>the</strong> hydrological year 2004/2005 (see Table 2) c<strong>an</strong> be verified.<br />

dS = (SWEend<br />

- SWEbegin<br />

) - ICEmelt<br />

dS = (309 -152)<br />

- 277 = −120<br />

mm<br />

Table 2. Components of <strong>the</strong> <strong>an</strong>nual water bal<strong>an</strong>ce of Goldbergkees catchment for <strong>the</strong> hydrological year<br />

of 2004/2005<br />

WB component Values (mm) Description<br />

SWE begin 152 SWE for model initialization at begin of October 2004<br />

P adj 2991 Adjusted precipitation over <strong>the</strong> catchment area<br />

ETA 128 Real evapotr<strong>an</strong>spiration<br />

Q tot 2956 Total runoff<br />

SNOWmelt 2566 Total snowmelt<br />

ICE melt 277 Total icemelt<br />

SWEend 309 SWE left at <strong>the</strong> end of <strong>the</strong> period at end of September 2005<br />

Equation 3 calculates <strong>the</strong> ch<strong>an</strong>ge of <strong>the</strong> main storages, where dS is <strong>the</strong> ch<strong>an</strong>ge of <strong>the</strong> storages over<br />

<strong>the</strong> <strong>entire</strong> simulation period. (definitions of symbols c<strong>an</strong> be seen at Table 2)<br />

WB = Padj<br />

- Qtot<br />

WB =<br />

- ETA - dS<br />

2991 - 2956 -128<br />

+ 120 = 27 mm<br />

Equation 4 evaluates <strong>the</strong> water bal<strong>an</strong>ce, where<strong>as</strong> WB is <strong>the</strong> <strong>an</strong>nual water bal<strong>an</strong>ce. The result of<br />

WB = 27 mm is explained by <strong>the</strong> difference of <strong>the</strong> b<strong>as</strong>e flow storage component at <strong>the</strong> simulation<br />

end minus <strong>the</strong> b<strong>as</strong>e flow storage value for <strong>the</strong> initialization at <strong>the</strong> beginning. Equation 5 shows <strong>the</strong><br />

share of <strong>the</strong> components snow melt, icemelt <strong>an</strong>d rainfall contributing to runoff.<br />

Q = SNOWmelt<br />

+ ICEmelt<br />

+ RAINdirect<br />

=> RAIN = 2956 - 277 - 2566 = 113 mm<br />

RAINdirect indicates <strong>the</strong> share of <strong>the</strong> total runoff which h<strong>as</strong> been induced by effective rainfall,<br />

directly routed to discharge. Liquid precipitation (RAINdirect) accounts for 3.8%, icemelt (ICE)<br />

accounts for 9.4%, <strong>an</strong>d snowmelt (SNOW) accounts for 86.8%. The value for liquid precipitation<br />

seems to be underestimated, because snowmelt h<strong>as</strong> a very high value. <strong>Snow</strong> melt is defined <strong>as</strong> <strong>the</strong><br />

snow which is melted by <strong>the</strong> energy affecting <strong>the</strong> snow surface, where<strong>as</strong> melted snow does not<br />

have to be routed directly to discharge.<br />

The simulated snow accumulation reached its maximum at 20 May 2005 at 1507 mm (Figure 8,<br />

c), compared to SWE field me<strong>as</strong>urements at 6 May 2005 showing a value of about 1391 mm. The<br />

monthly water bal<strong>an</strong>ce is presented in Figure 9. Icemelt starts in July <strong>an</strong>d typically h<strong>as</strong> its<br />

maximum in August. Glacier melt w<strong>as</strong> also simulated for November 2004, because of quite high<br />

temperatures at that time. Using field me<strong>as</strong>urements of <strong>the</strong> ice ablation stakes during <strong>the</strong><br />

hydrological year of 2004/2005, 510 mm of ice loss are calculated over <strong>the</strong> glacierized area<br />

(unpublished data from Schöner et al., 2006). The simulated ice loss accounts for 277 mm over <strong>the</strong><br />

(3)<br />

(4)<br />

(5)


33<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

catchment area, which would be 502 mm <strong>as</strong>signed to <strong>the</strong> glacierized area. This very good<br />

agreement between simulated <strong>an</strong>d observed icemelt is due to a reliable estimation of all modelled<br />

components. Evaporation accounted for 128 mm for <strong>the</strong> <strong>entire</strong> simulated year <strong>an</strong>d is mainly<br />

affected by low temperatures at this alpine site.<br />

Figure 9. Monthly simulated bal<strong>an</strong>ce of runoff (mm), precipitation (mm), icemelt (mm), snow melt (mm),<br />

<strong>an</strong>d real evaporation (mm) averaged over <strong>the</strong> catchment area over <strong>the</strong> period October 2004 to September<br />

2005.<br />

The observed ice loss for <strong>the</strong> verification period of 2003/2004 accounted for 197 mm averaged<br />

over <strong>the</strong> glacierized area <strong>an</strong>d <strong>the</strong> snow accumulation me<strong>as</strong>ured at <strong>the</strong> beginning of May 2004<br />

accounted for 1744 mm (unpublished data from Schöner et al., 2006). The simulation results for<br />

<strong>the</strong> verification period accounted for 275 mm ice loss averaged over <strong>the</strong> glacier <strong>an</strong>d 1690 mm of<br />

SWE calculated for <strong>the</strong> beginning of May. This me<strong>an</strong>s a slightly overestimation of <strong>the</strong> ice loss <strong>an</strong>d<br />

a slightly underestimation of <strong>the</strong> snow accumulation . Once more we c<strong>an</strong> show <strong>the</strong> quality of <strong>the</strong><br />

stabile multi-validation approach used for <strong>the</strong> 2004/2005 period.<br />

CONCLUSIONS AND OUTLOOK<br />

The PREVAH model shows <strong>the</strong> ability to model <strong>the</strong> waterbal<strong>an</strong>ce <strong>as</strong> well <strong>as</strong> <strong>the</strong> discharge<br />

hydrograph of <strong>the</strong> high elevated Goldbergkees watershed. Melt processes have been simulated<br />

accurately in different temporal scales: <strong>the</strong> simulated diurnal variations of runoff matched <strong>the</strong><br />

runoff observations during melt se<strong>as</strong>on <strong>an</strong>d <strong>the</strong> se<strong>as</strong>onal bal<strong>an</strong>ce is in a good agreement with <strong>the</strong><br />

observed m<strong>as</strong>s bal<strong>an</strong>ce data. This paper h<strong>as</strong> shown <strong>the</strong> need of various observed data like<br />

discharge hydrographs, distributed snow water equivalent <strong>an</strong>d ice ablation data to guar<strong>an</strong>tee a<br />

stabile cross-validation of simulation results (Verbunt et al. 2003). The model results concerning<br />

runoff relied on <strong>the</strong> input of precipitation <strong>an</strong>d air temperature, all <strong>the</strong> o<strong>the</strong>r input h<strong>as</strong> been<br />

necessary for <strong>the</strong> evaporation simulations. Hence, <strong>the</strong>re should be a quite small operating expense<br />

for data acquisition for melt modelling, despite all <strong>the</strong> problems <strong>an</strong>d uncertainties occurring during<br />

precipitation <strong>an</strong>d air temperature me<strong>as</strong>urements (Sevruk, 1986, 1989). The watershed of<br />

Goldbergkees h<strong>as</strong> demonstrated its great adv<strong>an</strong>tage of having a meteorological observatory on site<br />

<strong>an</strong>d long term <strong>an</strong>d detailed m<strong>as</strong>s bal<strong>an</strong>ce me<strong>as</strong>urements (Auer et al., 2002). The availability of<br />

spatial dense meteorological data of good quality at this isolated location is a unique feature <strong>an</strong>d<br />

<strong>an</strong> adv<strong>an</strong>tage of <strong>the</strong> study region. Main efforts for <strong>the</strong> preparation of simulations have been due to


34<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

<strong>the</strong> acquisitions of validating data, with respect to logistic m<strong>an</strong>ner, mountainous risks <strong>an</strong>d strictly<br />

wea<strong>the</strong>r depending pl<strong>an</strong>ning of field investigations.<br />

As PREVAH does not include <strong>the</strong> redistribution of snow by wind or aval<strong>an</strong>ches <strong>the</strong> spatial<br />

distribution of <strong>the</strong> SWE c<strong>an</strong>not be matched perfectly, but <strong>the</strong> me<strong>an</strong> values of <strong>the</strong> modelled SWE<br />

are identical to me<strong>as</strong>ured catchment me<strong>an</strong>s of <strong>the</strong> SWE. Hence, it seems that <strong>the</strong> main part of <strong>the</strong><br />

drifted snow originates from <strong>the</strong> simulated watershed <strong>an</strong>d remains inside. An additional<br />

conceptual module for <strong>the</strong> snow redistribution by wind <strong>an</strong>d aval<strong>an</strong>ches (Hartm<strong>an</strong>n et al., 1999)<br />

should be applied to <strong>the</strong> model to satisfy <strong>the</strong>se claims. It seems to be obvious, that only in a very<br />

good observed catchment area reliable simulation results of high quality c<strong>an</strong> be obtained.<br />

ACKNOWLEDGEMENTS<br />

This ongoing research w<strong>as</strong> supported by a gr<strong>an</strong>t from <strong>the</strong> Austri<strong>an</strong> Academy of Sciences under<br />

<strong>the</strong> project SNOWTRANS HOE29, part of <strong>the</strong> IHP PUB (International Hydrological Program,<br />

Prediction in Ungauged B<strong>as</strong>ins). M<strong>an</strong>y th<strong>an</strong>ks to Markus Vollm<strong>an</strong>n for <strong>the</strong> cl<strong>as</strong>sification of <strong>the</strong><br />

ASTER image. The authors are grateful to all <strong>the</strong> students, colleagues <strong>an</strong>d friends who helped to<br />

carry out <strong>the</strong> exhausting field work.<br />

REFERENCES<br />

Arnold N., Richards K., Willis I., Sharp M. 1998. Initial results from a semi-distributed, physically<br />

b<strong>as</strong>ed model of glacier hydrology. Hydrological Processes 12: 191–219.<br />

Arnold N.S., Willis I. C., Sharp M., Richards K.S., Lawson W. 1996. `A distributed surface<br />

energy bal<strong>an</strong>ce model for a small valley glacier. I. Development <strong>an</strong>d testing for <strong>the</strong> Haut<br />

Glacier d'Arolla, Valais, Switzerl<strong>an</strong>d', Journal of Glaciology, 42: 77–89.<br />

Auer I., Böhm R., Leymüller M., Schöner W., Kaiser A., Scheifinger H., L<strong>an</strong>ger M., Scheider St.,<br />

Häberli Ch. 2002. D<strong>as</strong> Klima des Sonnblicks. Klimaatl<strong>as</strong> und Klimatographie der GAW<br />

Station Sonnblick einschließlich der umgebenden Gebirgsregion. Österreichische Beiträge zu<br />

Meteorologie und Geophysik. Heft 28. Zentral<strong>an</strong>stalt für Meteorologie und Geodynamik,<br />

Wien. ISSN 1016–6254.<br />

Badoux A. 1999. Untersuchung zur flächendifferenzierten Modellierung von Abfluss und<br />

Schmelze in teilvergletscherten Einzugsgebieten. Diploma <strong>the</strong>sis at <strong>the</strong> Geographical Institute<br />

of ETH, Zürich.<br />

Blöschl G., Kirnbauer R., J<strong>an</strong>sa J., Kraus K., Kuschnig D., Gutknecht D., Reszler C. 2002. Using<br />

remote sensing methods for calibrating <strong>an</strong>d verifying a spatially distributed snow model.<br />

Österreichische W<strong>as</strong>ser- und Abfallwirtschaft (ÖWW) 54: 16 pp.<br />

Blöschl G., Kirnbauer R., Gutknecht D. 1991. Distributed <strong>Snow</strong>melt Simulations in <strong>an</strong> Alpine<br />

Catchment, 1, Model Evaluation on <strong>the</strong> B<strong>as</strong>is of <strong>Snow</strong> Cover Patterns. Water Resources<br />

Research, 27(12), 3171–3179.<br />

Collins D.N., 1982. Water storage in Alpine glaciers. Hydrological Aspects of Alpine <strong>an</strong>d High<br />

Mountain Are<strong>as</strong> (Proceedings of <strong>the</strong> Exeter Symposium, July 1982) IAHS Publ, no 138: 113–<br />

122.<br />

Doorschot J., Raderschall N., Lehning M. 2001. Me<strong>as</strong>urements <strong>an</strong>d one-dimensional model<br />

calculations of snow tr<strong>an</strong>sport over a mountain ridge. Annals of Glaciology, 32: 153–158.<br />

Elder K., Rosenthal W., Davis R.E. 1998. Estimating <strong>the</strong> spatial distribution of snow water<br />

equivalence in a mont<strong>an</strong>e watershed. Hydrological Processes 12: 1793–1808.<br />

Finsterwalder S., Schunk H. 1887. Der Suldenferner. Zeitschrift des Deutschen und<br />

Oesterreichischen Alpenvereins 18: 72–89.<br />

Gurtz J., Zappa M., J<strong>as</strong>per K., L<strong>an</strong>g H., Verbunt M., Badoux A., Vitvar T. 2003. A Comparative<br />

Study in Modelling Runoff <strong>an</strong>d its Components in Two Mountainous Catchments.<br />

Hydrological Processes 17: 297–311. DOI: 10.1002/hyp.1125


35<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Güntner A., Uhlenbrook S., Seibert J., Leibundgut C. 1999. Multi-criterial validation of<br />

TOPMODEL in a mountainous catchment. Hydrological Processes 13: 1603–1620.<br />

Hartm<strong>an</strong> M.D., Baron J.S., Lammers R.B., Cline D.W., B<strong>an</strong>d L.E., Liston G.E. Tague C. 1999.<br />

Simulations of snow distribution <strong>an</strong>d hydrology in a mountain b<strong>as</strong>in. Water Resources<br />

Research, 35(5): 1587–1603.<br />

Hock R. 1999. A distributed temperature-index ice- <strong>an</strong>d snowmelt model including potential direct<br />

solar radiation. Journal of Glaciology 45(149): 101–111.<br />

Hock R., Holmgren B. 2005. A distributed surface energy-bal<strong>an</strong>ce model for complex topography<br />

<strong>an</strong>d its application to Storglaciären, Sweden. Journal of Glaciology 51(172): 25–36.<br />

Hoinkes H. 1970. Methoden und Möglichkeiten von M<strong>as</strong>senhaushaltsstudien auf Gletschern.<br />

Ergebnisse der Meßreihe Hintereisferner (Ötztaler Alpen) 1953–1968. Zeitschrift für<br />

Gletscherkunde und Glazialgeologie, 6, 37 – 90.<br />

Hynek B., Schöner W. 2004. M<strong>as</strong>senhaushalt 2002/2003 der Gletscher in der Goldberggruppe.<br />

101–102. Jahresbericht des Sonnblick-Vereines: 1–15. Eigenverlag der Zentral<strong>an</strong>stalt für<br />

Meteorologie und Geodynamik, Wien.<br />

K<strong>as</strong>er G., Fountain A., J<strong>an</strong>sson P. 2003. A m<strong>an</strong>ual for monitoring <strong>the</strong> m<strong>as</strong>s bal<strong>an</strong>ce of mountain<br />

glaciers. IHP-VI, Technical Documents in Hydrology No. 59, UNESCO, Paris, 2003.<br />

Kirnbauer R, Blöschl G, Gutknecht D. 1994. Entering <strong>the</strong> Era of Distributed <strong>Snow</strong> Models. Nordic<br />

Hydrology 25: 1–24.<br />

Klok E.J., J<strong>as</strong>per K., Roelofsma K.P., Gurtz J., Badoux A. 2001. Distributed hydrological<br />

modelling of a heavily glaciated Alpine river b<strong>as</strong>in. Hydrological Sciences Journal—Journal<br />

Des Sciences Hydrologiques 46: 553–570.<br />

L<strong>an</strong>g H., Braun L. 1990. On <strong>the</strong> Information content of air Temperature in <strong>the</strong> Context of <strong>Snow</strong><br />

Melt Estimation. In: Hydrology of Mountainous Are<strong>as</strong>, L. Molar (Ed.). IAHS Publ. no. 190:<br />

347–354.<br />

Lehning M., Voelksch I., Gustafsson D., Nguyen T.A., Stähli M., Zappa M. 2006. Influence of<br />

snow, forest, soil <strong>an</strong>d local meteorological forcing on mountain catchment hydrology.<br />

Hydrological Processes, in press: DOI: 10.1002/hyp.6204.<br />

Lehning M., Doorschot J., Bartelt P. 2000. A snowdrift index b<strong>as</strong>ed on SNOWPACK model<br />

calculations. In Annals of Glaciology, 31: 382–386.<br />

Marks D., Domingo J., Susong D., Link T., Garen D. 1999. A spatially distributed energy bal<strong>an</strong>ce<br />

snowmelt model for application in mountain b<strong>as</strong>ins. Hydrological Processes 13:1935–1959.<br />

N<strong>as</strong>h J.E., Sutcliffe J.V. 1970. River flow forec<strong>as</strong>ting through conceptual models (1), a discussion<br />

of principles. Journal of Hydrology 10 (3): 282–290.<br />

Ohmura A. 2001. Physical B<strong>as</strong>is for <strong>the</strong> Temperature-B<strong>as</strong>ed Melt-Index Method. Journal of<br />

Applied Meteorology 40, No. 4: 753–761. DOI: 10.1175/1520-<br />

0450(2001)0402.0.CO;2<br />

Østrem G., Brugm<strong>an</strong> M. 1991. Glacier m<strong>as</strong>s-bal<strong>an</strong>ce me<strong>as</strong>urements—A m<strong>an</strong>ual for field <strong>an</strong>d<br />

office work: Environment C<strong>an</strong>ada, National Hydrology Research Institute Science Report No.<br />

4, <strong>an</strong>d Norwegi<strong>an</strong> Water Resources <strong>an</strong>d Energy Administration, 224 p.<br />

Pellicciotti F., Brock B., Str<strong>as</strong>ser U., Burl<strong>an</strong>do P., Funk M. Corripio J. 2005. An enh<strong>an</strong>ced<br />

temperature-index glacier melt model including <strong>the</strong> shortwave radiation bal<strong>an</strong>ce: development<br />

<strong>an</strong>d testing for Haut Glacier d’Arolla, Switzerl<strong>an</strong>d. Journal of Glaciology 51(175): 573–587.<br />

Ross B.B., Contractor D.N., Sh<strong>an</strong>holtz V.O. 1979. A Finite Element Model Of Overl<strong>an</strong>d And<br />

Ch<strong>an</strong>nel Flow For Assessing The Hydrologic Impact Of L<strong>an</strong>duse Ch<strong>an</strong>ge. Journal of<br />

Hydrology 41: 1–30.<br />

Schaefli B., Hingray B., Niggli M., Musy A. 2005. A conceptual glacio-hydrological model for<br />

high mountainous catchments, Hydrology <strong>an</strong>d Earth System Sciences Discussions 2: 73–117.<br />

Sevruk B. (ed.) 1986. Correction of Precipitation Me<strong>as</strong>urements. ETH/IASH/WMO Workshop on<br />

<strong>the</strong> Correction of Precipitation Me<strong>as</strong>urements. Zürich, April 1–3, 1985, Zuercher<br />

Geographische Schriften, Heft 23.


36<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Sevruk B. (ed.) 1989. Precipitation Me<strong>as</strong>urement. Proc. of <strong>the</strong> WMO/IASH/ETH Workshop on<br />

Precipitation Me<strong>as</strong>urement. St. Moritz, December 3–7, 1989, WMO Instruments <strong>an</strong>d<br />

Observing Methods Rep., No. 48, WMO/TD-No. 328, 199–204, 1989, Geneva.<br />

Singh P., Bengtsson L. 2005. Impact of warmer climate on melt <strong>an</strong>d evaporation for <strong>the</strong> rainfed,<br />

snowfed <strong>an</strong>d glacierfed b<strong>as</strong>ins in <strong>the</strong> Himalay<strong>an</strong> region. Journal of Hydrology 300: 140–154.<br />

DOI: 10.1016/j.hydrol.2004.06.005<br />

Verbunt M., Gurtz J., J<strong>as</strong>per K., L<strong>an</strong>g H., Warmerdam P., Zappa M. 2003. The hydrological role<br />

of snow <strong>an</strong>d glaciers in alpine river b<strong>as</strong>ins <strong>an</strong>d <strong>the</strong>ir distributed modelling. Journal of<br />

Hydrology 282: 36–55. DOI: 10.1016/S0022-1694(03)00251-8<br />

Zappa M., Pos F., Str<strong>as</strong>ser U., Gurtz J. 2003. Se<strong>as</strong>onal water bal<strong>an</strong>ce of <strong>an</strong> alpine catchment <strong>as</strong><br />

evaluated by different methods for spatially distributed snow melt modelling. Nordic<br />

Hydrology 34(3): 179–202<br />

Zappa M, Badoux A., Gurtz J. 2000. The application of a complex distributed hydrological model<br />

in a highly glaciated alpine river catchment, in Limnological Reports, Horvatic J. (Editor), 33 rd<br />

Conference of International Association for D<strong>an</strong>ube Research, Osijek, Croatia, 3–9 Sept. 2000,<br />

Vol. 33: 23–28.


<strong>Snow</strong> <strong>an</strong>d Climate<br />

37


This page is intentionally bl<strong>an</strong>k.<br />

38


39<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

20 th Century North Americ<strong>an</strong> <strong>Snow</strong> Extent Trends:<br />

Climate Ch<strong>an</strong>ge or Natural Climate Variability?<br />

EXTENDED ABSTRACT<br />

ALLAN FREI, 1 GAVIN GONG, 2 DAVID A. ROBINSON, 3<br />

GWANGYONG CHOI, 4 DEBJANI GHATAK, 5 AND YAN GE 6<br />

The purpose of this <strong>an</strong>alysis is to test <strong>the</strong> null hypo<strong>the</strong>sis that continental scale variations in<br />

North Americ<strong>an</strong> snow cover extent (NA SCE) c<strong>an</strong> be explained by atmospheric circulation alone,<br />

without need to invoke additional expl<strong>an</strong>atory factors such <strong>as</strong> climate ch<strong>an</strong>ge. We test <strong>the</strong> null<br />

hypo<strong>the</strong>sis by (1) presenting what is known about decadal scale variations in twentieth century<br />

continental scale NA SCE, <strong>an</strong>d (2) examining historical variations in surface climate, tropospheric<br />

<strong>an</strong>d stratospheric circulation, <strong>as</strong> well <strong>as</strong> corollary evidence from arctic sea ice variations, to<br />

determine whe<strong>the</strong>r <strong>the</strong> available evidence supports or refutes <strong>the</strong> null hypo<strong>the</strong>sis. In this<br />

presentation, preliminary results are presented focusing on snow extent during spring (i.e., March).<br />

The full report is being prepared for submission for publication elsewhere.<br />

METHODS AND DATA<br />

In order to test <strong>the</strong> null hypo<strong>the</strong>sis, we utilize data sets that extend back to, <strong>an</strong>d beyond, <strong>the</strong> mid<br />

twentieth century. Variations in snow depth, surface temperature <strong>an</strong>d precipitation rate, <strong>as</strong> well <strong>as</strong><br />

upper tropospheric <strong>an</strong>d mid stratospheric geopotential heights <strong>an</strong>d wind speeds are examined<br />

using time series, composite, <strong>an</strong>d correlation <strong>an</strong>alyses. Time series <strong>an</strong>alyses are used to identify<br />

climatic features in <strong>the</strong>se fields that covary over inter<strong>an</strong>nual <strong>an</strong>d decadal time scales. To evaluate<br />

decadal ch<strong>an</strong>ges, time series were smoo<strong>the</strong>d using 9-year running me<strong>an</strong>s <strong>as</strong> a low p<strong>as</strong>s filter in<br />

order to focus on decadal <strong>an</strong>d longer scale variations. The results <strong>an</strong>d conclusions are robust with<br />

regards to ch<strong>an</strong>ges in <strong>the</strong> smoothing window size. Observations of snow extent are taken from <strong>the</strong><br />

NOAA visible b<strong>as</strong>ed satellite product (Ramsay 1998; Robinson et al. 1999; Helfrich et al. 2006);<br />

reconstructed snow extent is from two sources (Brown 2000; Frei et al. 1999); snow depth is from<br />

a new gridded product (described in Dyer <strong>an</strong>d Mote 2006); climatological fields are from <strong>the</strong><br />

NCEP/NCAR Re<strong>an</strong>alysis project (Kalnay <strong>an</strong>d co-authors 1996); <strong>an</strong>d teleconnection indices are<br />

from <strong>the</strong> NOAA Climate Diagnostic Center, <strong>the</strong> Climate Research Unit of <strong>the</strong> University of E<strong>as</strong>t<br />

Anglia, <strong>an</strong>d <strong>the</strong> National Center for Atmospheric Research.<br />

1<br />

Department of Geography, Hunter College, Graduate Program in Earth <strong>an</strong>d Environmental<br />

Sciences, City University of New York, NY.<br />

2<br />

Department of Earth <strong>an</strong>d Environmental Engineering, Columbia University, NY.<br />

3<br />

Department of Geography, Rutgers University, NJ.<br />

4<br />

Department of Geography, Rutgers University, NJ.<br />

5<br />

Department of Geography, Hunter College, Graduate Program in Earth <strong>an</strong>d Environmental<br />

Sciences, City University of New York, NY.<br />

6<br />

Department of Earth <strong>an</strong>d Environmental Engineering, Columbia University, NY.


Time series of NA SCE <strong>an</strong>d AO<br />

During spring, <strong>the</strong> AO appears to contribute signific<strong>an</strong>tly to <strong>the</strong> vari<strong>an</strong>ce in snow extent at both<br />

inter<strong>an</strong>nual <strong>an</strong>d decadal scales (table 1). Nei<strong>the</strong>r <strong>the</strong> PNA nor <strong>the</strong> PDO are signific<strong>an</strong>tly correlated<br />

to snow extent individually, but at inter<strong>an</strong>nual time scales <strong>the</strong>y do appear to contribute in <strong>the</strong><br />

multiple correlation <strong>an</strong>alysis. The smoo<strong>the</strong>d (i.e. decadal scale) correlation between AO <strong>an</strong>d snow<br />

extent is very high (rsmooth = –0.93), explaining ~86% of <strong>the</strong> vari<strong>an</strong>ce, while nei<strong>the</strong>r of <strong>the</strong> o<strong>the</strong>r<br />

two indices contribute to <strong>the</strong> expl<strong>an</strong>atory power. Note that a correlation of this magnitude (rsmooth<br />

=–0.93) is not signific<strong>an</strong>t (p=0.05) in this <strong>an</strong>alysis because of <strong>the</strong> diminished number of effective<br />

degrees of freedom due to autocorrelation in <strong>the</strong> time series. Figure 1 shows smoo<strong>the</strong>d time series<br />

that have been normalized so that units are comparable. March snow extent (inverted), including<br />

values from satellite observations <strong>an</strong>d <strong>an</strong> historical reconstruction, are shown along with AO<br />

variations <strong>an</strong>d two historical NAO time series derived from station observations. March snow<br />

extent <strong>an</strong>d <strong>the</strong> AO appear to vary inversely, <strong>an</strong>d <strong>the</strong> reconstructed snow <strong>an</strong>d NAO time series<br />

indicate that this relationship holds <strong>as</strong> far back <strong>as</strong> <strong>the</strong> mid 1940s; prior to that time <strong>the</strong> relationship<br />

appears to weaken. It is not clear whe<strong>the</strong>r <strong>the</strong> pre-1940s deterioration of <strong>the</strong> snow-circulation<br />

relationship is real, or due to (1) <strong>the</strong> difference between <strong>the</strong> EOF-b<strong>as</strong>ed AO <strong>an</strong>d <strong>the</strong> station-b<strong>as</strong>ed<br />

NAO time series, or (2) to possible inaccuracies in one or more of <strong>the</strong> historical data sets.<br />

Table 1. Pearson correlation coefficients, <strong>an</strong>d multiple correlation coefficients, between North<br />

Americ<strong>an</strong> snow extent b<strong>as</strong>ed on satellite observations <strong>an</strong>d atmospheric circulation indices for <strong>an</strong>nual<br />

(r <strong>an</strong>nual) <strong>an</strong>d 9-year running me<strong>an</strong> (r smooth) time series. Values marked with <strong>an</strong> <strong>as</strong>terisk are statistically<br />

signific<strong>an</strong>t at p=0.05. Sample sizes for <strong>an</strong>nual <strong>an</strong>d smoo<strong>the</strong>d time series are n <strong>an</strong>nual = 39 <strong>an</strong>d n smooth = 31,<br />

respectively. However, effective sample sizes are smaller due to temporal autocorrelation.<br />

SPRING<br />

J<strong>an</strong>uary–February–March Indices<br />

versus<br />

March <strong>Snow</strong> Extent<br />

r<strong>an</strong>nual rsmooth AO –0.45* –0.93<br />

PNA –0.26 –0.25<br />

PDO 0.00 –0.25<br />

AO/PNA 0.54* 0.93<br />

AO/PDO 0.46* 0.93<br />

PNA/PDO 0.38 0.25<br />

AO/PNA/PDO 0.60* 0.93<br />

Figure 1. 9-yr running me<strong>an</strong> values of JFM AO (solid), extended NAO from station observations (from <strong>the</strong><br />

Climate Research Unit at <strong>the</strong> University of E<strong>as</strong>t Anglia) (d<strong>as</strong>hed), extended NAO from station observations<br />

(from <strong>the</strong> National Center for Atmospheric Research) (dots), March snow extent from satellite observations<br />

<strong>an</strong>d reconstructed March snow extent from (Brown 2000) (<strong>as</strong>terisks) which have been inverted for e<strong>as</strong>ier<br />

comparison. All time series have been normalized to fit on <strong>the</strong> same scale.<br />

40


Spatial distributions of ch<strong>an</strong>ges in snow extent, snow depth, temperature, precipitation, <strong>an</strong>d<br />

atmospheric circulation<br />

The differences in snow extent between periods of more extensive <strong>an</strong>d less extensive snow<br />

cover occur at <strong>the</strong> sou<strong>the</strong>rn boundary of <strong>the</strong> snow pack. This is in contr<strong>as</strong>t to <strong>the</strong> differences in<br />

snow depth, which are shallower across wide swaths of <strong>the</strong> continent. These variations in <strong>the</strong><br />

March snow pack are <strong>as</strong>sociated with ch<strong>an</strong>ges in surface climate during March <strong>an</strong>d/or preceding<br />

months. Me<strong>an</strong> temperature differences on <strong>the</strong> order of 1–2 C across wide swaths of <strong>the</strong> continent<br />

are observed, <strong>an</strong>d moderate precipitation ch<strong>an</strong>ges on <strong>the</strong> order of 0.5 mm/day primarily are found<br />

near <strong>the</strong> sou<strong>the</strong>rn boundary of <strong>the</strong> snow pack. The spatial distribution of <strong>the</strong>se surface ch<strong>an</strong>ges<br />

suggests that <strong>the</strong>y might be caused by a southward displacement of <strong>the</strong> me<strong>an</strong> polar front <strong>an</strong>d jet<br />

stream. Composite difference maps of <strong>the</strong> same atmospheric fields for positive versus negative AO<br />

years show that <strong>the</strong> upper tropospheric <strong>an</strong>d stratospheric ch<strong>an</strong>ges <strong>as</strong>sociated with <strong>the</strong> AO are<br />

similar in character, <strong>an</strong>d greater in magnitude, th<strong>an</strong> ch<strong>an</strong>ges <strong>as</strong>sociated with snow cover.<br />

Links to sea ice variability<br />

In light of <strong>the</strong> well-studied impact of <strong>the</strong> AO on Arctic sea ice extent, <strong>an</strong>d in conjunction with<br />

<strong>the</strong> results of <strong>the</strong> <strong>an</strong>alyses presented in <strong>the</strong> previous sections, one would expect to find a signal in<br />

<strong>the</strong> sea ice record that is correlated to both <strong>the</strong> winter–spring AO <strong>an</strong>d to march NA SCE. However,<br />

recent reports of decre<strong>as</strong>ing sea ice suggest a possible discrep<strong>an</strong>cy. Results of <strong>an</strong> EOF <strong>an</strong>alysis of<br />

arctic sea ice extent shows that <strong>the</strong> record of sea ice extent appears to be consistent with our null<br />

hypo<strong>the</strong>sis. Winter – spring AO variations do appear to be influencing both NA SCE <strong>an</strong>d arctic<br />

sea ice. The arctic sea ice signal is regionally dependent, exerting opposite effects in different<br />

regions, <strong>an</strong>d <strong>the</strong>refore dampened in <strong>the</strong> spatially integrated signal for <strong>the</strong> <strong>entire</strong> arctic.<br />

DISCUSSION AND CONCLUSIONS<br />

Evidence presented here suggests that <strong>the</strong>se variations are m<strong>an</strong>ifestations of decadal scale<br />

variations in <strong>the</strong> position of <strong>the</strong> polar front <strong>an</strong>d jet stream, partially influenced by some of <strong>the</strong><br />

primary modes of atmospheric variability. Statistically signific<strong>an</strong>t relationships are found between<br />

March North Americ<strong>an</strong> snow extent / snow depth / surface climate <strong>an</strong>d <strong>the</strong> winter AO since <strong>the</strong><br />

mid-20 th century. These relationships are stronger at decadal th<strong>an</strong> at inter<strong>an</strong>nual time scales, <strong>an</strong>d<br />

appear to weaken prior to mid century. Fur<strong>the</strong>r supporting evidence is provided by <strong>the</strong> domin<strong>an</strong>t<br />

EOF of Nor<strong>the</strong>rn Hemisphere winter sea ice extent, which displays decadal scale variations that<br />

are also correlated to historical variations in both <strong>the</strong> AO <strong>an</strong>d snow extent, <strong>as</strong> would be expected<br />

under <strong>the</strong> null hypo<strong>the</strong>sis. Results from climate models are consistent with <strong>the</strong> empirical results<br />

presented here. Current-generation coupled atmosphere–oce<strong>an</strong> climate model results suggest that<br />

decadal scale variability of <strong>the</strong> magnitude discussed here is likely due to internal climate<br />

variability <strong>an</strong>d not to external forcing. Thus, evidence examined in this <strong>an</strong>alysis leads to <strong>the</strong><br />

conclusion that we c<strong>an</strong> not presently reject our null hypo<strong>the</strong>sis, <strong>an</strong>d that ch<strong>an</strong>ges observed thus far<br />

in <strong>the</strong> continental scale snow extent record over North America c<strong>an</strong> not be attributed to<br />

<strong>an</strong>thropogenic forcing.<br />

REFERENCES<br />

Brown, R. D. (2000). Nor<strong>the</strong>rn hemisphere snow cover variability <strong>an</strong>d ch<strong>an</strong>ge, 1915–1997.<br />

Journal of Climate 13(13): 2339–2355.<br />

Dyer, J. L. <strong>an</strong>d T. L. Mote (2006). Spatial variability <strong>an</strong>d trends in snow depth over North<br />

America. Geophysical Research Letters: in review.<br />

Frei, A., D. A. Robinson <strong>an</strong>d M. G. Hughes (1999). North Americ<strong>an</strong> snow extent: 1900–1994.<br />

International Journal of Climatology 19: 1517–1534.<br />

Helfrich, S. R., D. McNamara, B. H. Ramsay <strong>an</strong>d T. K<strong>as</strong>heta (2006). Enh<strong>an</strong>cements <strong>an</strong>d<br />

forthcoming developments to <strong>the</strong> Interactive Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping System<br />

(IMS). 63rd E<strong>as</strong>tern <strong>Snow</strong> Conference. University of Delaware, Newark Delaware.<br />

41


Kalnay, E. <strong>an</strong>d co-authors (1996). The NCEP/NCAR 40-year re<strong>an</strong>alysis project. Bulletin of <strong>the</strong><br />

Americ<strong>an</strong> Meteorological Society 77: 437–471.<br />

Ramsay, B. H. (1998). The interactive multisensor snow <strong>an</strong>d ice mapping system. Hydrological<br />

Processes 12: 1537–1546.<br />

Robinson, D. A., J. D. Tarpley <strong>an</strong>d B. H. Ramsay (1999). Tr<strong>an</strong>sition from NOAA weekly to daily<br />

hemispheric snow charts. 10th Symposium on Global Ch<strong>an</strong>ge Studies, Dall<strong>as</strong>, TX, Americ<strong>an</strong><br />

Meteorological Society.<br />

42


<strong>Snow</strong> <strong>an</strong>d Climate Posters<br />

43


This page is intentionally bl<strong>an</strong>k.<br />

44


45<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Observed Differences Between <strong>Snow</strong> Extent <strong>an</strong>d <strong>Snow</strong> Depth<br />

Variability at Continental Scales<br />

EXTENDED ABSTRACT<br />

YAN GE 1 AND GAVIN GONG 1<br />

<strong>Snow</strong> extent <strong>an</strong>d snow depth are two related characteristics of a snowpack, but <strong>the</strong>y need not be<br />

mutually consistent. Different behaviors of snow depth <strong>an</strong>d snow extent are readily apparent at<br />

local scales; for example, during conditions of uninterrupted snow cover, snow depth at a point<br />

c<strong>an</strong> vary subst<strong>an</strong>tially due to snowfall events, metamorphosis <strong>an</strong>d ablation. However, <strong>the</strong> behavior<br />

of snow extent vs. snow depth at regional to continental scales is understudied.<br />

Regional/continental-scale gridded dat<strong>as</strong>ets of snow depth <strong>an</strong>d snow water equivalent (SWE),<br />

obtained from field observations <strong>an</strong>d satellite data, were utilized to qu<strong>an</strong>titatively evaluate <strong>the</strong><br />

relationship between snow extent <strong>an</strong>d snow depth over North America. Statistical methods were<br />

applied to <strong>as</strong>sess <strong>the</strong> mutual consistency of monthly snow depth vs. snow extent. Results from all<br />

<strong>the</strong> dat<strong>as</strong>ets indicate <strong>the</strong> signific<strong>an</strong>ce of snow depth variations, <strong>an</strong>d that snow depth <strong>an</strong>omalies <strong>an</strong>d<br />

snow extent <strong>an</strong>omalies are not necessarily consistent, especially at higher latitude. This observed<br />

lack of mutual consistency at continental scales suggests that average snowpack depth variations<br />

may be of sufficiently large magnitude, spatial scope <strong>an</strong>d temporal duration to influence regionalhemispheric<br />

climate, in a m<strong>an</strong>ner unrelated to <strong>the</strong> more extensively studied snow extent variations.<br />

Keywords: <strong>Snow</strong> depth, <strong>Snow</strong> extent, Continental-Scale, Consistency<br />

Dat<strong>as</strong>ets<br />

In order to study <strong>the</strong> temporal differences between snow depth <strong>an</strong>d snow extent at regional to<br />

continental scales, dat<strong>as</strong>ets characterized by spatial coverage across North America <strong>an</strong>d/or 20th<br />

century temporal duration have been acquired <strong>an</strong>d utilized.<br />

• Gridded snow depth dat<strong>as</strong>et sp<strong>an</strong>ning North America interpolated from 349 Stations<br />

located in <strong>the</strong> United States <strong>an</strong>d sou<strong>the</strong>rn C<strong>an</strong>ada for <strong>the</strong> 1915–1997 period. (Brown,<br />

2000)<br />

• Daily 1 o x1 o grid dat<strong>as</strong>et covering all of North America interpolated from a<br />

comprehensive set of station snow depth observations from 1956 through 2000. (Dyer<br />

<strong>an</strong>d Mote, 2006)<br />

• SWE Maps for <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Prairies derived using p<strong>as</strong>sive microwave data from <strong>the</strong><br />

NASA Sc<strong>an</strong>ning Multich<strong>an</strong>nel Microwave Radiometer (SMMR) <strong>an</strong>d Special Sensor<br />

Microwave Imager (SSM/I) instruments from 1978 to 1997. (Derksen et al., 2004)<br />

Results<br />

The North America snow extent (i.e., spatial snow covered area) w<strong>as</strong> computed by Ross Brown<br />

(Brown, 2000). The corresponding me<strong>an</strong> snow depth variable is computed <strong>as</strong> <strong>the</strong> area-weighted<br />

average snow depth over this snow covered extent. Monthly statistics are applied to <strong>as</strong>sess <strong>the</strong><br />

1 Department of Earth <strong>an</strong>d Environmental Engineering, Columbia University, 500 W. 120 St.,<br />

Room 918, MC4711, New York, NY. Corresponding author: Y<strong>an</strong> Ge, yg2124@columbia.edu.


mutual consistency of snow depth vs. snow extent over North America. Results from all three of<br />

<strong>the</strong> dat<strong>as</strong>ets described above consistently indicate that at continental/regional scales, snow depth<br />

<strong>an</strong>d snow extent <strong>an</strong>omalies are not mutually consistent.<br />

Figure 1a shows that <strong>the</strong> monthly evolution of both me<strong>an</strong> <strong>an</strong>d variability of snow depth <strong>an</strong>d<br />

snow extent for North America follows different temporal patterns. <strong>Snow</strong> extent incre<strong>as</strong>es in <strong>the</strong><br />

early snow se<strong>as</strong>on <strong>an</strong>d reaches its maximum value in J<strong>an</strong>uary, but inter<strong>an</strong>nual variability of<br />

monthly snow extent is small in J<strong>an</strong>uary. On <strong>the</strong> contrary, both me<strong>an</strong> value <strong>an</strong>d r<strong>an</strong>ge of<br />

variability of monthly snow depth incre<strong>as</strong>es steadily through February/March. Ano<strong>the</strong>r indication<br />

of different behavior between <strong>the</strong>se two snow parameters is <strong>the</strong> monthly correlation between snow<br />

depth <strong>an</strong>d snow extent, presented in Figure 1b. The inter<strong>an</strong>nual snow depth <strong>an</strong>d snow extent<br />

timeseries are often poorly correlated, especially in winter. Figure 1c shows <strong>the</strong> coefficient of<br />

variation (i.e., ration of st<strong>an</strong>dard deviation to me<strong>an</strong>) for both extent <strong>an</strong>d depth, <strong>an</strong>d indicates that<br />

<strong>the</strong> inter<strong>an</strong>nual variability of snow depth comprises a larger percentage of its climatological me<strong>an</strong><br />

state th<strong>an</strong> that for snow extent, consistently across all months. Hence <strong>the</strong> magnitude of snow depth<br />

variability c<strong>an</strong> be considered more subst<strong>an</strong>tial th<strong>an</strong> <strong>the</strong> magnitude of snow extent variability.<br />

Monthly scatterplots (Figure 1d) support <strong>the</strong> correlation results, i.e., wide scatter <strong>an</strong>d a lack of <strong>an</strong><br />

apparent snow extent vs. snow depth relationship during winter, <strong>an</strong>d a more noticeable but still<br />

modest relationship during autumn <strong>an</strong>d spring.<br />

sne (10 6 km 2 )<br />

snd(cm)<br />

corr(snd,sne)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Nov Dec J<strong>an</strong> Feb Mar Apr<br />

month<br />

Nov Dec J<strong>an</strong> Feb Mar Apr<br />

month<br />

0<br />

Nov Dec J<strong>an</strong> Feb Mar Apr<br />

month<br />

a) b)<br />

c) d)<br />

Figure 1. Nov.–Apr. timeseries of North America snow extent “sne” (Brown, 2000) <strong>an</strong>d grid domain average<br />

snow-covered depth “snd” (Dyer <strong>an</strong>d Mote, 2006) from 1956 to 1997. a) Monthly box plots of inter<strong>an</strong>nual<br />

sne <strong>an</strong>d snd; b) Monthly correlations between inter<strong>an</strong>nual sne <strong>an</strong>d snd; c) Monthly cv (std/me<strong>an</strong>) of<br />

inter<strong>an</strong>nual sne <strong>an</strong>d snd time series; d) Monthly scatterplots of inter<strong>an</strong>nual sne <strong>an</strong>d snd.<br />

CV<br />

snd (cm)<br />

snd (cm)<br />

46<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

Nov Dec J<strong>an</strong> Feb Mar Apr<br />

month<br />

15<br />

snd=0.81*sne-1.12<br />

R2 Nov.<br />

=0.2220<br />

10<br />

5<br />

10 15 20<br />

45<br />

snd=1.61*sne+4.36<br />

40 R2 Feb.<br />

=0.1011<br />

35<br />

30<br />

25<br />

10 15 20<br />

30<br />

snd=0.78*sne+6.09<br />

25 R2 Dec.<br />

=0.1052<br />

20<br />

15<br />

50<br />

snd=1.86*sne+1.76<br />

45 R<br />

40<br />

35<br />

30<br />

25<br />

20<br />

10 15 20<br />

2 =0.1530<br />

sne (10 6 km 2 10<br />

10 15 20<br />

sne (10<br />

Mar.<br />

)<br />

6 km 2 )<br />

40<br />

snd=0.80*sne+13.05<br />

35 R2 J<strong>an</strong>.<br />

=0.0269<br />

30<br />

25<br />

20<br />

10 15 20<br />

35<br />

snd=2.25*sne-9.23<br />

30 R2 Apr.<br />

=0.2508<br />

25<br />

20<br />

15<br />

snd<br />

sne<br />

10<br />

10 15 20


Figure 2 presents correlation maps between gridded snow depth (Dyer <strong>an</strong>d Mote, 2006) <strong>an</strong>d<br />

North America snow extent (Brown, 2000) from November to April. All <strong>the</strong> months show strong<br />

correlation near <strong>the</strong> snowline <strong>an</strong>d weak correlation over virtually all points north of <strong>the</strong> snowline.<br />

Correlations drop off dramatically north of <strong>the</strong> snowline, resulting in a very large high-latitude<br />

region in which snow depth <strong>an</strong>d snow extent are poorly correlated. During winter months <strong>the</strong>re are<br />

even regions of negative correlation north of <strong>the</strong> snow line, which may be due to <strong>the</strong> inter<strong>an</strong>nual<br />

variation of nor<strong>the</strong><strong>as</strong>tward mid-latitude cyclone tracks.<br />

30 ° N<br />

30 ° N<br />

30 ° N<br />

45 ° N<br />

45 ° N<br />

45 ° N<br />

60 ° N<br />

60 ° N<br />

60 ° N<br />

Nov.<br />

120 ° W 90 ° W 60 ° W<br />

Dec.<br />

120 ° W 90 ° W 60 ° W<br />

J<strong>an</strong>.<br />

120 ° W 90 ° W 60 ° W<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

Figure 2. Correlation maps between gridded snow depth (Dyer <strong>an</strong>d Mote, 2006) <strong>an</strong>d North America snow<br />

extent (Brown, 2000).<br />

47<br />

30 ° N<br />

30 ° N<br />

30 ° N<br />

45 ° N<br />

45 ° N<br />

45 ° N<br />

60 ° N<br />

60 ° N<br />

60 ° N<br />

Feb.<br />

120 ° W 90 ° W 60 ° W<br />

Mar.<br />

120 ° W 90 ° W 60 ° W<br />

Apr.<br />

120 ° W 90 ° W 60 ° W<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4


Figure 3 shows <strong>the</strong> correlation maps between gridded SWE (Derksen et al., 2004) <strong>an</strong>d North<br />

America snow extent (Brown, 2000). The satellite data confirms <strong>the</strong> observed inconsistency<br />

between snow depth <strong>an</strong>d North America snow extent from station dat<strong>as</strong>ets seen in Figure 2.<br />

Correlations decline from south to north in central North America region (38N–63N) for three<br />

consecutive months (December, J<strong>an</strong>uary <strong>an</strong>d February). Regions of negative correlation also exist<br />

north of snowline during J<strong>an</strong>uary <strong>an</strong>d February.<br />

30 ° N<br />

30 ° N<br />

30 ° N<br />

45 ° N<br />

45 ° N<br />

45 ° N<br />

60 ° N<br />

60 ° N<br />

60 ° N<br />

Dec.<br />

120 ° W 90 ° W 60 ° W<br />

J<strong>an</strong>.<br />

120 ° W 90 ° W 60 ° W<br />

Feb.<br />

120 ° W 90 ° W 60 ° W<br />

Figure 3. Correlation maps between gridded SWE (Derksen et al., 2004) <strong>an</strong>d North America snow extent<br />

(Brown, 2000).<br />

48<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4


CONCLUSIONS<br />

Three regional-continental scale snow depth dat<strong>as</strong>ets have been evaluated, derived from station<br />

observational data or satellite remote sensing data. All of <strong>the</strong> dat<strong>as</strong>ets consistently exhibit differing<br />

behavior between inter<strong>an</strong>nual timeseries <strong>an</strong>omalies of average snow-covered depth <strong>an</strong>d North<br />

America snow extent at regional/continental scales, generally consistent with what is expected at<br />

local scales. Correlation maps of snow depth vs. snow extent show a broad, continental-scale<br />

region is maintained throughout much of <strong>the</strong> snow se<strong>as</strong>on in which snow depth variations are not<br />

mutually consistent with North America snow extent variations. Research also recognized that<br />

relative magnitude of snow depth <strong>an</strong>omalies is considerably larger th<strong>an</strong> that for snow extent.<br />

The observed differing behavior of <strong>the</strong> two snowpack parameters at regional/continental scales<br />

suggest that snow depth variations may be of sufficiently large magnitude, spatial scope <strong>an</strong>d<br />

temporal duration to influence regional-hemispheric climate, in a m<strong>an</strong>ner unrelated to <strong>the</strong> more<br />

extensively studied snow extent variations. Such <strong>an</strong>alysis represents <strong>the</strong> first step towards <strong>the</strong> goal<br />

of identifying explicit snow depth – climate relationships <strong>an</strong>d integrating <strong>the</strong>m with snow extent–<br />

climate relationships.<br />

REFERENCES<br />

Brown, R.D., 2000. Nor<strong>the</strong>rn Hemisphere snow cover variability <strong>an</strong>d ch<strong>an</strong>ge, 1915–1997. Journal<br />

of Climate 13: 2339–2355.<br />

Derksen, C., Brown, R. D. <strong>an</strong>d Walker, A., 2004. Merging conventional (1915–1992) <strong>an</strong>d p<strong>as</strong>sive<br />

microwave (1978–2002) estimates of snow extent <strong>an</strong>d water equivalent over central North<br />

America. Journal of Hydrometeorology 5(5): 850–861.<br />

Dyer, L.J. <strong>an</strong>d Mote,T.L., 2006. Spatial Variability And Trends In <strong>Snow</strong> Depth Over North<br />

America. Geophysical Research Letters, in press.<br />

49


This page is intentionally bl<strong>an</strong>k.<br />

50


51<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Climate Variability, <strong>Snow</strong>melt Distribution,<br />

<strong>an</strong>d Effects on Streamflow in a C<strong>as</strong>cades Watershed<br />

ANNE JEFFERSON 1 , ANNE NOLIN 1 , SARAH LEWIS 1 , MEREDITH PAYNE 2 ,<br />

GORDON GRANT 3 , AND CHRISTINA TAGUE 4<br />

EXTENDED ABSTRACT<br />

C<strong>as</strong>cades R<strong>an</strong>ge rivers provide critical water supply for agriculture, ecosystems, <strong>an</strong>d<br />

municipalities in <strong>the</strong> Pacific Northwest, <strong>an</strong>d <strong>the</strong>y derive much of <strong>the</strong>ir water from snowmelt<br />

filtered through groundwater aquifers. Recent <strong>an</strong>alyses show that this region is particularly<br />

sensitive to current <strong>an</strong>d projected climate warming trends, specifically reduced snow accumulation<br />

<strong>an</strong>d earlier spring melt (Mote, 2003; Stewart et al., 2005). By 2050, C<strong>as</strong>cades snowpacks are<br />

projected to be less th<strong>an</strong> half of what <strong>the</strong>y are today (Leung et al., 2004), potentially leading to<br />

major water shortages. Broad regional-scale characterizations identify climatic gradients <strong>as</strong> <strong>the</strong><br />

most import<strong>an</strong>t controls on spatial variability in streamflow regimes, but <strong>the</strong> potential for o<strong>the</strong>r<br />

hydrological factors, particularly groundwater, to influence this response h<strong>as</strong> received much less<br />

attention. Our objective is to develop <strong>an</strong> underst<strong>an</strong>ding of how discharge from a groundwaterdominated<br />

watershed is controlled at <strong>the</strong> event, se<strong>as</strong>onal, <strong>an</strong>d inter<strong>an</strong>nual scales by snowpack<br />

dynamics, <strong>an</strong>tecedent conditions, <strong>an</strong>d global climate signals.<br />

The study watershed is that of <strong>the</strong> McKenzie River at Clear Lake (Figure 1), in <strong>the</strong> central<br />

Oregon C<strong>as</strong>cades, includes extensive are<strong>as</strong> of high permeability Quaternary (High C<strong>as</strong>cade)<br />

b<strong>as</strong>alts that result in a subst<strong>an</strong>tial groundwater system, <strong>as</strong> well <strong>as</strong> runoff-dominated Tertiary<br />

(Western C<strong>as</strong>cades) l<strong>an</strong>dscapes (Sherrod <strong>an</strong>d Smith, 2000; Tague <strong>an</strong>d Gr<strong>an</strong>t, 2004). This 239 km 2<br />

watershed h<strong>as</strong> long-term records of streamflow from United States Geological Survey gage<br />

#14158500. It also h<strong>as</strong> record of precipitation, snow, <strong>an</strong>d temperature from three Natural<br />

Resources Conservation Service SNOTEL sites: Hogg P<strong>as</strong>s (1451 m, 21E05S), S<strong>an</strong>tiam Junction<br />

(1143 m, 21E06S) <strong>an</strong>d Jump Off Joe (1067 m, 22E07S). Annual precipitation in <strong>the</strong> watershed<br />

r<strong>an</strong>ges from ~1.8 to 3 m, <strong>an</strong>d 70% falls between November <strong>an</strong>d March. 47% of <strong>the</strong> watershed lies<br />

between 918 <strong>an</strong>d 1200 m, in <strong>the</strong> tr<strong>an</strong>sient snow zone. From 1200 m to <strong>the</strong> peak elevation (2051<br />

m), se<strong>as</strong>onal snowpacks occur from November through June. Peak snow water equivalent (SWE)<br />

occurs around April 1 st at Hogg P<strong>as</strong>s, <strong>an</strong>d around March 1 st at S<strong>an</strong>tiam Junction <strong>an</strong>d Jump Off Joe.<br />

In order to examine relationships between hydrological variables in space <strong>an</strong>d time, we<br />

performed Pearson’s correlations, autocorrelations, <strong>an</strong>d cross-correlations using 42 parameters<br />

derived from discharge, precipitation, SWE, <strong>an</strong>d temperature from <strong>the</strong> stations listed above. We<br />

also correlated discharge <strong>an</strong>d SWE with monthly values of <strong>the</strong> Niño 3.4 index of sea surface<br />

1 Department of Geosciences, Oregon State University, 104 Wilkinson Hall, Corvallis OR<br />

97331.<br />

2 College of Oce<strong>an</strong>ic <strong>an</strong>d Atmospheric Sciences, Oregon State University, 104 COAS<br />

Administration Bldg., Corvallis, OR 97331.<br />

3 USDA Forest Service, Pacific Northwest Research Station, 3200 SW Jefferson Way,<br />

Corvallis, OR 97331.<br />

4 Bren School of Environmental Science <strong>an</strong>d M<strong>an</strong>agement, University of California, S<strong>an</strong>ta<br />

Barbara, 93106.


temperature (Trenberth <strong>an</strong>d Step<strong>an</strong>iak, 2001). A simple water bal<strong>an</strong>ce w<strong>as</strong> constructed for <strong>the</strong><br />

2001–2004 water years, using values for <strong>the</strong> average b<strong>as</strong>in elevation (1215 m) interpolated from<br />

precipitation <strong>an</strong>d SWE values at Jump Off Joe <strong>an</strong>d Hogg P<strong>as</strong>s. B<strong>as</strong>in-averaged evapotr<strong>an</strong>spiration<br />

w<strong>as</strong> calulcated in RHESSys (Tague <strong>an</strong>d B<strong>an</strong>d, 2004). Predictive models of September–November<br />

minimum discharge were developed using stepwise regression in SAS 9.1.<br />

Fluctuations in discharge are muted relative to daily variability in <strong>the</strong> recharge (rain plus<br />

snowmelt) signal (Figure 2). Summer streamflows are sustained by groundwater, not snowmelt.<br />

There is a high degree of discharge auto-correlation for ~2.5 months, <strong>an</strong>d <strong>the</strong>re is a strong crosscorrelation<br />

between <strong>the</strong> previous year’s precipitation <strong>an</strong>d <strong>the</strong> current year’s discharge at a 1 year<br />

lag. The El Niño-Sou<strong>the</strong>rn Oscillation is a re<strong>as</strong>onably good predictor of SWE <strong>an</strong>d a moderate<br />

predictor of <strong>an</strong>nual discharge. Intern<strong>an</strong>ual variability in <strong>the</strong> 26-year SNOTEL record m<strong>as</strong>ks <strong>an</strong>y<br />

long-term trends in precipitation or SWE, but <strong>the</strong> longer discharge records suggests that climate<br />

warming is altering <strong>the</strong> streamflow regime at Clear Lake. The hydrograph temporal center of m<strong>as</strong>s<br />

(Stewart et al., 2005) h<strong>as</strong> moved earlier by a statistically signific<strong>an</strong>t 15 days since 1950, which is<br />

probably a function of relatively more winter rain <strong>an</strong>d earlier snowmelt. This finding is in line<br />

with o<strong>the</strong>r watersheds throughout <strong>the</strong> mountainous west (Stewart et al., 2005). September–<br />

November minimum flows have declined since 1947, probably <strong>as</strong> a result of longer summer<br />

recession periods resulting from earlier snowmelt.<br />

Clear Lake<br />

watershed<br />

Figure 1. Location map of <strong>the</strong> study watershed, relative to <strong>the</strong> topography of Oregon <strong>an</strong>d extent of High<br />

C<strong>as</strong>cades geology.<br />

52


Discharge (mm)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

10/1/03 2/1/04 6/1/04 10/1/04 2/1/05 6/1/05<br />

Figure 2. Discharge at Clear Lake for water years 2004 <strong>an</strong>d 2005 (bottom, black line), compared to <strong>the</strong><br />

interpolated rain + snowmelt time series for <strong>the</strong> medi<strong>an</strong> b<strong>as</strong>in elevation (top, blue line).<br />

The main conclusions of this study are:<br />

1. In <strong>the</strong> study watershed, pronounced se<strong>as</strong>onal variability of water inputs is damped by<br />

extensive groundwater systems. The study area may serve <strong>as</strong> a model for o<strong>the</strong>r<br />

groundwater-dominated watersheds in <strong>the</strong> mountainous west.<br />

2. Delays between precipitation <strong>an</strong>d discharge are a function of snowpack storage <strong>an</strong>d<br />

slow rele<strong>as</strong>e of groundwater. From July to October, streamflow is sustained by<br />

groundwater.<br />

3. Groundwater-dominated watersheds are somewhat buffered from


ACKNOWLEDGEMENTS<br />

This work w<strong>as</strong> funded by <strong>the</strong> Institute for Water <strong>an</strong>d Watersheds at Oregon State University,<br />

through <strong>the</strong> USGS water resources research institute program, <strong>an</strong>d by <strong>the</strong> Eugene Water <strong>an</strong>d<br />

Electric Board. This material is b<strong>as</strong>ed on work supported under a National Science Foundation<br />

Fellowship.<br />

REFERENCES<br />

Leung LR, Qi<strong>an</strong> Y, Bi<strong>an</strong> X, W<strong>as</strong>hington W, H<strong>an</strong> J, Roads JO. 2004. Mid-century ensemble<br />

regional climate ch<strong>an</strong>ge scenarios for <strong>the</strong> western United States. Climatic Ch<strong>an</strong>ge 62: 75–113.<br />

Mote PW. 2003. Trends in snow water equivalent in <strong>the</strong> Pacific Northwest <strong>an</strong>d <strong>the</strong>ir climatic<br />

causes. Geophysical Research Letters 30: 1601, doi:10.1029/2003GL017258.<br />

Sherrod DR, Smith JG. 2000. Geologic Map of Upper Eocene to Holocene Volc<strong>an</strong>ic <strong>an</strong>d Related<br />

Rocks of <strong>the</strong> C<strong>as</strong>cade R<strong>an</strong>ge, Oregon. U.S. Geological Survey: W<strong>as</strong>hington;<br />

Stewart IT, Cay<strong>an</strong> DR, Dettinger MD. 2005. Ch<strong>an</strong>ges toward earlier streamflow timing across<br />

western North America. Journal of Climate 18: 1136–1155.<br />

Tague CL, B<strong>an</strong>d LE. 2004. RHESSys: Regional Hydro-Ecologic Simulation System—An objectoriented<br />

approach to spatially distributed modeling of carbon, water, <strong>an</strong>d nutrient cycling.<br />

Earth Interactions 8: Paper No. 19, p. 1–42.<br />

Tague C, Gr<strong>an</strong>t GE. 2004. A geological framework for interpreting <strong>the</strong> low flow regimes of<br />

C<strong>as</strong>cade streams, Willamette River B<strong>as</strong>in, Oregon. Water Resources Research 40: W04303<br />

10.1029/2003WR002629.<br />

Trenberth KE, Step<strong>an</strong>iak DP. 2001. Indices of El Niño evolution. Journal of Climate 14: 1697–<br />

1701.<br />

54


ABSTRACT<br />

55<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Synoptic Patterns Associated with <strong>the</strong> Record <strong>Snow</strong>fall<br />

of 1960 in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s<br />

L. BAKER PERRY 1 AND CHARLES E. KONRAD II 2<br />

Record snowfall totals of up to 211 cm bl<strong>an</strong>keted <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s between mid-<br />

February <strong>an</strong>d late-March 1960. <strong>Snow</strong> w<strong>as</strong> reported on average every o<strong>the</strong>r day in <strong>the</strong> higher<br />

elevations, <strong>an</strong>d me<strong>an</strong> temperatures for <strong>the</strong> period were nearly 6 C below normal. <strong>Snow</strong> piled up to<br />

great depths, with Boone, NC, reporting a maximum depth of 112 cm. <strong>Snow</strong> drifts buried roads<br />

<strong>an</strong>d made travel impossible, requiring food, fuel, <strong>an</strong>d hay to be airlifted into <strong>the</strong> region. This paper<br />

<strong>an</strong>alyzes <strong>the</strong> synoptic patterns <strong>as</strong>sociated with <strong>the</strong> record snowfall of February <strong>an</strong>d March 1960.<br />

<strong>Snow</strong>fall events are identified using a combination of first order hourly observations <strong>an</strong>d<br />

cooperative observer daily snowfall totals. The spatial patterns of snowfall are mapped using a<br />

GIS, while me<strong>an</strong> values for various synoptic fields (e.g. 850 hPa temperature, 500 hPa height) are<br />

calculated <strong>an</strong>d compared to 50-year climatological me<strong>an</strong>s. <strong>Snow</strong>fall events during this period are<br />

<strong>the</strong>n cl<strong>as</strong>sified according to <strong>the</strong> pattern of cyclogenesis <strong>an</strong>d prevailing flow direction.<br />

Keywords: Synoptic patterns, snowfall, 1960, Sou<strong>the</strong>rn Appalachi<strong>an</strong>s<br />

INTRODUCTION<br />

February <strong>an</strong>d March 1960 continue to be remembered <strong>as</strong> <strong>the</strong> snowiest period on record in <strong>the</strong><br />

Sou<strong>the</strong>rn Appalachi<strong>an</strong>s. <strong>Snow</strong> w<strong>as</strong> nearly a daily occurrence between 13 February <strong>an</strong>d 26 March<br />

at higher elevations, with Boone, NC, reporting 211 cm <strong>an</strong>d m<strong>an</strong>y o<strong>the</strong>r locations in excess of 175<br />

cm. These snowfall totals are considerably greater th<strong>an</strong> <strong>the</strong> current 30-year me<strong>an</strong> <strong>an</strong>nual snowfall<br />

of 102 cm for Boone <strong>an</strong>d approach <strong>the</strong> me<strong>an</strong> <strong>an</strong>nual values of nearly 250 cm for <strong>the</strong> 2,000 m peaks<br />

(e.g Mt. Leconte <strong>an</strong>d Mt. Mitchell) of <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s. Me<strong>an</strong> temperatures for <strong>the</strong><br />

period were nearly 6 ºC below normal (Hardie 1960). The combination of frequent snowfall <strong>an</strong>d<br />

low temperatures allowed <strong>the</strong> snow to pile up to great depths, with Boone, NC, reporting a<br />

maximum depth of 112 cm on 13–14 March (NCDC 2002). Considerable blowing <strong>an</strong>d drifting of<br />

snow compounded problems, closing roads <strong>an</strong>d requiring emergency distribution of food, fuel, <strong>an</strong>d<br />

hay (Figs. 1 & 2). Although <strong>the</strong> impacts were greatest in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s, <strong>the</strong> <strong>entire</strong><br />

E<strong>as</strong>tern U.S. w<strong>as</strong> adversely affected. March alone broke more records for cold <strong>an</strong>d snow across <strong>the</strong><br />

E<strong>as</strong>tern U.S. th<strong>an</strong> <strong>an</strong>y o<strong>the</strong>r March on record until <strong>the</strong>n (Ludlum 1960a, Ludlum 1960b). It<br />

remains <strong>the</strong> coldest March on record for m<strong>an</strong>y locations in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s (NCDC<br />

2002.<br />

1<br />

Department of Geography <strong>an</strong>d Pl<strong>an</strong>ning, Box 32066, Appalachi<strong>an</strong> State University, Boone,<br />

NC 28608 (perrylb@appstate.edu)<br />

2<br />

Department of Geography, University of North Carolina, Chapel Hill, NC 27599


Figure 1. Headlines from <strong>the</strong> Watauga Democrat (Boone, NC) 10 Mar 1960 (Watauga 1960).<br />

Figure 2. <strong>Snow</strong> depth on 17 Mar 1960 in Ashe County, NC (Photo courtesy of <strong>the</strong> NC DOT).<br />

56


This paper <strong>an</strong>alyzes <strong>the</strong> synoptic patterns <strong>as</strong>sociated with <strong>the</strong> record snowfall of February <strong>an</strong>d<br />

March 1960 across <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s (Fig. 3). Three major questions serve to guide <strong>the</strong><br />

research. 1) What were <strong>the</strong> spatial patterns of snowfall during <strong>the</strong> record-breaking period between<br />

13 Feb <strong>an</strong>d 26 Mar 1960? 2) What were <strong>the</strong> synoptic field values <strong>an</strong>d map patterns for 1000 hPa<br />

height, 500 hPa height, <strong>an</strong>d 850 hPa temperature during this period <strong>an</strong>d how <strong>an</strong>omalous were <strong>the</strong>se<br />

values? 3) What synoptic patterns were primarily responsible for <strong>the</strong> snowfall, <strong>an</strong>d how did <strong>the</strong>ir<br />

relative import<strong>an</strong>ce vary across <strong>the</strong> region?<br />

DATA AND METHODS<br />

Figure 3. Location of study area.<br />

Daily snowfall records for February <strong>an</strong>d March 1960 served <strong>as</strong> <strong>the</strong> source of snowfall data. We<br />

extracted <strong>the</strong>se data from <strong>the</strong> National Climatic Data Center’s Cooperative Summary of <strong>the</strong> Day<br />

CD-Rom (NCDC 2002) for 121 cooperative observer (COOP) stations in <strong>the</strong> Sou<strong>the</strong>rn<br />

Appalachi<strong>an</strong> Region (Fig. 4). We restricted our <strong>an</strong>alyses to 43 COOP stations that had complete<br />

coverage during <strong>the</strong> study period.<br />

We defined a snowfall event <strong>as</strong> having occurred if at le<strong>as</strong>t one COOP station in <strong>the</strong> region<br />

reported snow accumulation on a given date. To improve <strong>the</strong> temporal resolution of each event,<br />

we referenced hourly surface observation summaries from four nearby first order stations (Fig. 4).<br />

From interpretation of <strong>the</strong>se data, we were able to approximate <strong>the</strong> onset, maturation, <strong>an</strong>d ending<br />

time <strong>as</strong> well <strong>as</strong> <strong>the</strong> duration of reported snowfall across <strong>the</strong> region. Assessment of <strong>the</strong> maturation<br />

time involved determining <strong>the</strong> hour in which <strong>the</strong> spatial extent of snowfall w<strong>as</strong> greatest across <strong>the</strong><br />

network of first-order stations. An event remained active if precipitation w<strong>as</strong> reported during a sixhour<br />

period. When precipitation w<strong>as</strong> no longer reported at <strong>an</strong>y of <strong>the</strong> four first-order stations for<br />

more th<strong>an</strong> six hours, we defined <strong>the</strong> event <strong>as</strong> having ended at <strong>the</strong> hour precipitation w<strong>as</strong> l<strong>as</strong>t<br />

reported. Using this approach, we identified 17 snowfall events during <strong>the</strong> 43-day period between<br />

13 February <strong>an</strong>d 26 March 1960.<br />

57


Figure 4. Topography of <strong>the</strong> region <strong>an</strong>d wea<strong>the</strong>r stations used in this study.<br />

We used gridded (2.5 by 2.5 degree latitude/longitude mesh), twice-daily synoptic fields that<br />

were extracted from CDs containing <strong>the</strong> National Center for Environmental Prediction (NCEP)<br />

Re<strong>an</strong>alysis dat<strong>as</strong>et (Kalnay et al. 1996) to <strong>an</strong>alyze <strong>the</strong> synoptic patterns during <strong>the</strong> period. These<br />

fields were spatially interpolated to <strong>the</strong> center of each snowfall event. Using <strong>the</strong> 0000 <strong>an</strong>d 1200<br />

UTC gridded synoptic fields, we undertook a temporal interpolation to estimate field values<br />

during <strong>the</strong> event maturation time. We employed <strong>an</strong> inverse dist<strong>an</strong>ce technique to carry out all<br />

spatial <strong>an</strong>d temporal interpolations. Synoptic Climatology Suite (Konrad <strong>an</strong>d Meaux 2003) w<strong>as</strong><br />

used for compositing of synoptic fields <strong>an</strong>d calculation of <strong>an</strong>omalies.<br />

Each of <strong>the</strong> 17 snowfall events identified during <strong>the</strong> period w<strong>as</strong> cl<strong>as</strong>sified m<strong>an</strong>ually into one of<br />

five synoptic types (Table 1) according to <strong>the</strong> synoptic patterns identified from twice-daily surface<br />

<strong>an</strong>alyses (NOAA 1960) <strong>an</strong>d 850 hPa wind components (Kalnay et al. 1996). Miller Type A <strong>an</strong>d<br />

Miller Type B cyclones are responsible for most of <strong>the</strong> big snowstorms across <strong>the</strong> Sou<strong>the</strong>rn<br />

Appalachi<strong>an</strong>s, mid-Atl<strong>an</strong>tic, <strong>an</strong>d into <strong>the</strong> nor<strong>the</strong><strong>as</strong>tern U.S. (Miller 1946, Kocin <strong>an</strong>d Ucellini 1990,<br />

Keeter et al. 1995, Mote et al. 1997). In fact, all of <strong>the</strong> 20 major snowstorms in <strong>the</strong> nor<strong>the</strong><strong>as</strong>tern<br />

U.S. <strong>an</strong>alyzed by Kocin <strong>an</strong>d Ucellini (1990) were of <strong>the</strong> Miller Type A or B variety. Miller Type<br />

A cyclones are characterized by <strong>the</strong> development of a surface cyclone along a frontal boundary in<br />

<strong>the</strong> Gulf of Mexico separating cold continental air from maritime tropical Gulf or Atl<strong>an</strong>tic air. The<br />

surface low tracks nor<strong>the</strong><strong>as</strong>tward out of <strong>the</strong> Gulf of Mexico, paralleling <strong>the</strong> Atl<strong>an</strong>tic co<strong>as</strong>tline <strong>an</strong>d<br />

in some c<strong>as</strong>es intensifying fur<strong>the</strong>r. Miller Type B cyclones, however, initially track west of <strong>the</strong><br />

Appalachi<strong>an</strong>s. As <strong>the</strong> primary low dissipates in <strong>the</strong> Ohio Valley, a secondary low develops along<br />

<strong>the</strong> Atl<strong>an</strong>tic co<strong>as</strong>t.<br />

Table 1. Major synoptic types observed during February <strong>an</strong>d March 1960.<br />

Cl<strong>as</strong>s Synoptic Type<br />

1 Miller Type A<br />

2 Miller Type B<br />

3 Ohio Valley<br />

4 Northwest Flow<br />

5 O<strong>the</strong>r<br />

58


Cyclones that track west of <strong>the</strong> Appalachi<strong>an</strong>s through <strong>the</strong> Ohio Valley <strong>an</strong>d do not undergo<br />

secondary cyclogenesis along <strong>the</strong> Atl<strong>an</strong>tic co<strong>as</strong>t constitute a third synoptic type responsible for<br />

producing snowfall during <strong>the</strong> period. These Ohio Valley cyclones (Knappenberger <strong>an</strong>d Michaels<br />

1993) typically result in minor snowfall accumulations <strong>as</strong> most of <strong>the</strong> region remains in <strong>the</strong> warm<br />

sector until <strong>the</strong> cold front p<strong>as</strong>ses. Northwest Flow <strong>Snow</strong> (NWFS) events result from <strong>the</strong><br />

orographic <strong>as</strong>cent of a cold <strong>an</strong>d moist low-level northwest flow, often in <strong>the</strong> absence of synopticscale<br />

lifting (Perry <strong>an</strong>d Konrad 2005, Perry <strong>an</strong>d Konrad 2006, Perry et al. 2006). In this study,<br />

NWFS were defined on <strong>the</strong> b<strong>as</strong>is of a northwest (270 to 360 degrees) 850 hPa wind at event<br />

maturation. O<strong>the</strong>r types of events (e.g. Alberta clippers, prolonged isentropic lift) were also<br />

cl<strong>as</strong>sified during <strong>the</strong> period, but contributed minimally to snowfall totals.<br />

RESULTS AND DISCUSSION<br />

The greatest snowfall totals between 13 February <strong>an</strong>d 26 March 1960 occurred in <strong>the</strong> North<br />

Carolina High Country, where Boone recorded 211 cm (Fig. 5). The 265 cm of snowfall recorded<br />

during <strong>the</strong> 1959–60 snow se<strong>as</strong>on in Boone still st<strong>an</strong>ds <strong>as</strong> <strong>the</strong> <strong>an</strong>nual record, <strong>an</strong>d only one o<strong>the</strong>r<br />

year (1967–68 at 215 cm) received more th<strong>an</strong> <strong>the</strong> 211 cm that fell during <strong>the</strong> 43-day period in<br />

1960. All of <strong>the</strong> higher elevations in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s received considerable snowfall<br />

during <strong>the</strong> period, with m<strong>an</strong>y locations me<strong>as</strong>uring two or three times <strong>as</strong> much snow during this<br />

period th<strong>an</strong> <strong>the</strong>y average in <strong>an</strong> <strong>entire</strong> year. As previously reported, <strong>the</strong> greatest snow depth<br />

me<strong>as</strong>ured w<strong>as</strong> 112 cm, also in Boone, NC. The sustained cold, coupled with signific<strong>an</strong>t blowing<br />

<strong>an</strong>d drifting snow, greatly incre<strong>as</strong>ed <strong>the</strong> societal impacts. Residents would awake to find freshly<br />

plowed roads buried once again in newly drifted snow (Minor 1960).<br />

Figure 5. Total snowfall between 13 Feb <strong>an</strong>d 26 Mar 1960.<br />

59


Figure 6. 500 hPa height (m) <strong>an</strong>d <strong>an</strong>omaly for 13 Feb to 26 Mar 1960 (reference period 1951–2000).<br />

Figure 7. 850-hPa temperature <strong>an</strong>omaly (ºC) for 13 Feb to 26 Mar 1960 (reference period 1951–2000).<br />

60


The synoptic-scale atmospheric circulation w<strong>as</strong> highly <strong>an</strong>omalous during this period. The 500hPa<br />

height, for example, w<strong>as</strong> considerably below <strong>the</strong> 50-year me<strong>an</strong> for <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s<br />

(–2.29 z) <strong>an</strong>d throughout e<strong>as</strong>tern North America. A 500 hPa trough dominated e<strong>as</strong>tern North<br />

America throughout this period, providing <strong>an</strong> active, sou<strong>the</strong><strong>as</strong>terly displaced storm track that<br />

allowed frequent intrusions of cold continental air (Fig. 6). The 850 hPa temperature <strong>an</strong>omaly w<strong>as</strong><br />

even more extreme, registering nearly 8.0 C cooler (–3.46 z) th<strong>an</strong> <strong>the</strong> 50-year me<strong>an</strong> over nor<strong>the</strong>rn<br />

portions of <strong>the</strong> study area (Fig. 7). Again, <strong>the</strong> highly <strong>an</strong>omalous values were not restricted to <strong>the</strong><br />

Sou<strong>the</strong>rn Appalachi<strong>an</strong>s, but covered <strong>the</strong> <strong>entire</strong> e<strong>as</strong>tern U.S. As a result, March 1960 w<strong>as</strong> <strong>the</strong><br />

coldest March on record across much of <strong>the</strong> e<strong>as</strong>tern U.S. (Ludlum 1960b).<br />

Three primary synoptic types contributed <strong>the</strong> bulk of <strong>the</strong> snowfall across <strong>the</strong> Sou<strong>the</strong>rn<br />

Appalachi<strong>an</strong>s (Fig. 8). The majority of <strong>the</strong> snowfall across <strong>the</strong> region occurred in <strong>as</strong>sociation with<br />

Miller Type A <strong>an</strong>d Miller Type B cyclones (Fig. 9). The two Miller Type A cyclones occurred in<br />

February, where<strong>as</strong> <strong>the</strong> three Miller Type B storms moved by in March. Interestingly, <strong>the</strong> greatest<br />

accumulations in <strong>as</strong>sociation with <strong>the</strong> Miller Type A cyclones were in <strong>the</strong> western <strong>an</strong>d nor<strong>the</strong>rn<br />

portions of <strong>the</strong> study area, while are<strong>as</strong> in <strong>the</strong> sou<strong>the</strong><strong>as</strong>t of <strong>the</strong> study area received mostly rain<br />

(NCDC 2002). Miller Type B cyclones contributed <strong>the</strong> greatest amount to snowfall totals during<br />

<strong>the</strong> period at all but two of <strong>the</strong> regions across <strong>the</strong> study area. In <strong>the</strong> south <strong>an</strong>d e<strong>as</strong>t, <strong>the</strong>se three<br />

storms contributed nearly all of <strong>the</strong> snowfall, occurring during a two-week period in early March.<br />

In <strong>the</strong> higher elevations <strong>an</strong>d along northwestern slopes, NWFS contributed subst<strong>an</strong>tially to<br />

snowfall totals, constituting <strong>as</strong> much <strong>as</strong> 35 percent of <strong>the</strong> total. Interestingly, none of <strong>the</strong> NWFS<br />

events were tied to 500 hPa cutoff lows, which were responsible for <strong>the</strong> extremely heavy spring<br />

snow events in April 1987 <strong>an</strong>d May 1992 (NWS 1987, Fishel <strong>an</strong>d Businger 1993). The NWFS<br />

events were more numerous th<strong>an</strong> <strong>the</strong> Miller Type A <strong>an</strong>d B cyclones, but <strong>the</strong> snowfall w<strong>as</strong> limited<br />

in spatial extent. Mainly <strong>as</strong> a result of <strong>the</strong> NWFS events, <strong>the</strong> cumulative percentage of total<br />

snowfall during <strong>the</strong> period contributed by each event displayed considerable variability across <strong>the</strong><br />

Sou<strong>the</strong>rn Appalachi<strong>an</strong>s (Fig. 10). At lower elevations, particularly in <strong>the</strong> sou<strong>the</strong><strong>as</strong>t, only two or<br />

three exceptionally heavy events occurred during <strong>the</strong> period. Elsewhere, <strong>an</strong>d particularly at higher<br />

elevations <strong>an</strong>d along northwestern slopes, lighter NWFS events were interspersed with <strong>the</strong> big<br />

Miller Type A <strong>an</strong>d B cyclones, reflecting a greater diversity of synoptic types responsible for <strong>the</strong><br />

snowfall totals.<br />

Figure 8. Three primary synoptic patterns responsible for <strong>the</strong> record snowfall <strong>an</strong>d <strong>as</strong>sociated dates.<br />

61


Figure 9. Total snowfall between 13 Feb <strong>an</strong>d 26 Mar 1960 by synoptic pattern.<br />

Figure 10. Cumulative percentage of total snowfall contributed by each event<br />

(from heaviest to lightest) <strong>an</strong>d number of events needed to reach 90%.<br />

62


CONCLUSIONS<br />

The record snowfall in February <strong>an</strong>d March 1960 across <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s occurred in<br />

<strong>as</strong>sociation with highly <strong>an</strong>omalous 500 hPa circulation <strong>an</strong>d 850 hPa temperatures. The persistence<br />

of a 500 hPa trough over <strong>the</strong> e<strong>as</strong>tern U.S. w<strong>as</strong> <strong>as</strong>sociated with a sou<strong>the</strong><strong>as</strong>terly displaced storm<br />

track that kept temperatures well below normal. As a result, <strong>the</strong> 1959–60 snow se<strong>as</strong>on still holds<br />

<strong>the</strong> <strong>an</strong>nual snowfall record (265 cm) <strong>an</strong>d maximum snow depth record (112 cm) for Boone, NC,<br />

<strong>an</strong>d countless o<strong>the</strong>r locations across <strong>the</strong> region. Miller Type A cyclones, Miller Type B cyclones,<br />

<strong>an</strong>d NWFS events were responsible for nearly all of <strong>the</strong> total snowfall across <strong>the</strong> region. However,<br />

NWFS made <strong>the</strong> greatest contribution to snowfall totals across <strong>the</strong> higher elevations <strong>an</strong>d along<br />

northwestern slopes, where snow fell on average every o<strong>the</strong>r day during <strong>the</strong> period. A small<br />

number of exceptionally heavy events produced <strong>the</strong> record snowfall in <strong>the</strong> Georgia <strong>an</strong>d Carolina<br />

foothills in contr<strong>as</strong>t with <strong>the</strong> elevated are<strong>as</strong> to <strong>the</strong> north <strong>an</strong>d west where numerous, relatively light<br />

NWFS events contributed subst<strong>an</strong>tially to <strong>the</strong> total snowfall. Un<strong>an</strong>swered questions include <strong>the</strong><br />

causes of <strong>the</strong> <strong>an</strong>omalous <strong>an</strong>d persistent 500 hPa circulation pattern. In particular, <strong>the</strong> role of<br />

hemispheric teleconnections versus persistent snow cover across <strong>the</strong> region in forcing this pattern<br />

needs to be explored. The influence of favorable air trajectories extending downwind from <strong>the</strong><br />

Great Lakes in periods of NWFS also deserves fur<strong>the</strong>r scrutiny, particularly since <strong>the</strong>se have been<br />

tied to enh<strong>an</strong>ced snowfall in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s (Perry <strong>an</strong>d Konrad 2005). L<strong>as</strong>tly, given<br />

<strong>the</strong> deep snowpack <strong>an</strong>d reports of flooding during rapid melting at <strong>the</strong> end of March (Ludlum<br />

1960b), <strong>an</strong> investigation of <strong>the</strong> hydrological signific<strong>an</strong>ce (e.g. SWE estimations <strong>an</strong>d flooding<br />

reports) of February <strong>an</strong>d March 1960 is also warr<strong>an</strong>ted.<br />

ACKNOWLEDGMENTS<br />

Peter Robinson, Tom Whitmore, Walt Martin, Tom Schmidlin, Laurence Lee, <strong>an</strong>d Michael<br />

Mayfield provided valuable feedback on <strong>the</strong> design of <strong>the</strong> study. Julie Whichard from <strong>the</strong> North<br />

Carolina Department of Tr<strong>an</strong>sportation kindly provided archived photos. Addie Mae Love, Louiva<br />

Ward, <strong>an</strong>d Howell Cooke shared <strong>the</strong>ir m<strong>an</strong>y memories.<br />

REFERENCES<br />

Fishel, G.B., <strong>an</strong>d S. Businger. 1993. Heavy orographic snowfall in <strong>the</strong> sou<strong>the</strong>rn Appalachi<strong>an</strong>s: a<br />

late se<strong>as</strong>on c<strong>as</strong>e study. Postprints, Third National Heavy Precipitation Workshop: 275 – 284.<br />

Pittsburgh, PA: NWS/NOAA.<br />

Hardie, A. V. 1960. Special Wea<strong>the</strong>r Summary, in Climatological Data, North Carolina: March<br />

1960. Asheville, NC: National Climatic Data Center.<br />

Kalnay, E., M. K<strong>an</strong>amitsu, R. Kistler, W. Collins, <strong>an</strong>d 18 o<strong>the</strong>rs. 1996. The NCEP / NCAR 40year<br />

re<strong>an</strong>alysis project. Bulletin of <strong>the</strong> Americ<strong>an</strong> Meteorological Society 77: 437–471.<br />

Keeter, K.K., S. Businger, L.G. Lee, J.S. Waldstreicher. 1995. Winter wea<strong>the</strong>r forec<strong>as</strong>ting through<br />

<strong>the</strong> e<strong>as</strong>tern United States. Part III: The effects of topography <strong>an</strong>d <strong>the</strong> variability of winter<br />

wea<strong>the</strong>r in <strong>the</strong> Carolin<strong>as</strong> <strong>an</strong>d Virginia. Wea<strong>the</strong>r <strong>an</strong>d Forec<strong>as</strong>ting 10: 42–60.<br />

Knappenberger, P.C., <strong>an</strong>d P.J. Michaels. 1993. Cyclone tracks <strong>an</strong>d wintertime climate in <strong>the</strong> mid-<br />

Atl<strong>an</strong>tic region of <strong>the</strong> USA. International Journal of Climatology 13: 509–531.<br />

Kocin, P.J, <strong>an</strong>d L.W. Uccelini. 1990. <strong>Snow</strong>storms along <strong>the</strong> Nor<strong>the</strong><strong>as</strong>tern Co<strong>as</strong>t of <strong>the</strong> United<br />

States: 1955 – 1985. Meteorological Monograph No. 44, Americ<strong>an</strong> Meteorological Society.<br />

Konrad, C.E., <strong>an</strong>d D.Meaux. 2003. Synoptic Suite Software Package. Chapel Hill, NC: University<br />

of North Carolina.<br />

Ludlum, D. 1960a. The great early March snowstorm of 1960. Wea<strong>the</strong>rwise 13: 59–62.<br />

Ludlum, D. 1960b. The great temperature reversal: March to April. Wea<strong>the</strong>rwise 13: 120–128.<br />

Miller, J.E. 1946. Cyclogenesis in <strong>the</strong> Atl<strong>an</strong>tic Co<strong>as</strong>tal region of <strong>the</strong> United States. Journal of<br />

Meteorology 3: 31–44.<br />

63


Minor, J. 1960. Watauga in Dis<strong>as</strong>ter Area: Red Cross, Guard Units Act to Aid <strong>Snow</strong> Victims.<br />

Watauga Democrat, March 17, 1960, p. 1.<br />

Mote, T.L., D.W. Gamble, S. J. Underwood, M.L. Bentley. 1997. Synoptic-scale features common<br />

to heavy snowstorms in <strong>the</strong> Sou<strong>the</strong><strong>as</strong>t United States. Wea<strong>the</strong>r <strong>an</strong>d Forec<strong>as</strong>ting 12: 5–23.<br />

NWS. 1987. <strong>Snow</strong>storm in <strong>the</strong> Appalachi<strong>an</strong> Region on April 2–5, 1987. Storm Data 29: 6–14.<br />

NCDC. 2002. Cooperative Summary of <strong>the</strong> Day: E<strong>as</strong>tern U.S. Asheville, NC: National Climatic<br />

Data Center.<br />

NOAA. 1960. Daily Wea<strong>the</strong>r Maps, Weekly Series. W<strong>as</strong>hington, D.C.: Department of Commerce,<br />

Environmental Data Service.<br />

Perry, L. Baker, Charles E. Konrad. 2005. The Influence of <strong>the</strong> Great Lakes on <strong>Snow</strong>fall Patterns<br />

in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s. Proceedings of <strong>the</strong> 62 nd E<strong>as</strong>tern <strong>Snow</strong> Conference: 279–289.<br />

Perry, L. Baker, <strong>an</strong>d Charles E. Konrad. 2006. Relationships between NW flow snowfall <strong>an</strong>d<br />

topography in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s, USA. Climate Research 32: 35–47.<br />

Perry, L. Baker, Charles E. Konrad, Thom<strong>as</strong> W. Schmidlin. 2006. Antecedent upstream air<br />

trajectories <strong>as</strong>sociated with northwest flow snowfall in <strong>the</strong> Sou<strong>the</strong>rn Appalachi<strong>an</strong>s, USA.<br />

Wea<strong>the</strong>r <strong>an</strong>d Forec<strong>as</strong>ting, in press.<br />

Schmidlin, T.W. 1992. Does lake-effect snow extend to <strong>the</strong> mountains of West Virginia?<br />

Proceedings of <strong>the</strong> 49 th E<strong>as</strong>tern <strong>Snow</strong> Conference: 145–148.<br />

Watauga Democrat. 1960. Watauga Digging Out Of Mountainous <strong>Snow</strong>fall: Crews Work Round<br />

Clock to Open Roads. Watauga Democrat, March 10, 1960.<br />

64


65<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Influence of <strong>Snow</strong>fall Anomalies on Summer Precipitation<br />

in <strong>the</strong> Nor<strong>the</strong>rn Great Plains of North America<br />

ABSTRACT<br />

STEVEN M. QUIRING 1 , AND DARIA B. KLUVER 2<br />

Using observations from 1929 to 1999, we examine <strong>the</strong> relationship between winter/spring<br />

snowfall <strong>an</strong>omalies <strong>an</strong>d summer precipitation over <strong>the</strong> nor<strong>the</strong>rn Great Plains of North America.<br />

Both composite <strong>an</strong>d correlation <strong>an</strong>alysis indicate that <strong>an</strong>omalously dry (wet) summers are<br />

<strong>as</strong>sociated with negative (positive) snowfall <strong>an</strong>omalies during <strong>the</strong> preceding winter <strong>an</strong>d spring. It<br />

is posited that below (above) normal snowfall is <strong>as</strong>sociated with decre<strong>as</strong>es (incre<strong>as</strong>es) in<br />

spring/early summer soil moisture <strong>an</strong>d <strong>as</strong>sociated decre<strong>as</strong>es (incre<strong>as</strong>es) in local moisture recycling<br />

during summer. It appears that <strong>the</strong> snowfall <strong>an</strong>omalies must exceed some minimum threshold<br />

before <strong>the</strong>y have a signific<strong>an</strong>t impact on atmospheric circulation <strong>an</strong>d precipitation during <strong>the</strong><br />

following summer. There is also signific<strong>an</strong>t temporal variability in <strong>the</strong> strength of <strong>the</strong> correlations<br />

between snowfall <strong>an</strong>d summer moisture. Relationships between April–May snowfall <strong>an</strong>d summer<br />

moisture were generally quite strong between 1929 <strong>an</strong>d 1954 <strong>an</strong>d 1970 to 1987, but were<br />

relatively weak during 1955 to 1969 <strong>an</strong>d after 1987. This suggests that o<strong>the</strong>r factors may be<br />

modulating <strong>the</strong> import<strong>an</strong>ce of l<strong>an</strong>d surface processes.<br />

Keywords: snowfall, precipitation, drought, Great Plains<br />

INTRODUCTION<br />

L<strong>an</strong>d surface conditions (e.g., snow cover, soil moisture) are import<strong>an</strong>t sources of se<strong>as</strong>onal<br />

climate predictability (Koster <strong>an</strong>d Suarez, 2001; Koster et al., 2003; Koster et al., 2004).<br />

Numerous studies have demonstrated that Eur<strong>as</strong>i<strong>an</strong>/Tibet<strong>an</strong> snow cover influences Indi<strong>an</strong>/Asi<strong>an</strong><br />

monsoonal circulation <strong>an</strong>d precipitation (Kripal<strong>an</strong>i et al., 2002; Robock et al., 2003; Wu <strong>an</strong>d Qi<strong>an</strong>,<br />

2003; F<strong>as</strong>ullo, 2004; Zh<strong>an</strong>g et al., 2004). <strong>Snow</strong> cover <strong>an</strong>d snow water equivalent have also been<br />

linked to variability in <strong>the</strong> North Americ<strong>an</strong> Monsoon (Gutzler, 2000; Ellis <strong>an</strong>d Hawkins, 2001;<br />

Hawkins et al., 2002; Lo <strong>an</strong>d Clark, 2002; Matsui et al., 2003). O<strong>the</strong>r studies have identified<br />

connections between Eur<strong>as</strong>i<strong>an</strong> snow cover extent <strong>an</strong>d summer air temperature in <strong>the</strong> United<br />

Kingdom (Qi<strong>an</strong> <strong>an</strong>d Saunders, 2003), <strong>an</strong>d C<strong>an</strong>adi<strong>an</strong> river discharge (Déry et al., 2005). Although<br />

both <strong>the</strong> presence <strong>an</strong>d amount of snow c<strong>an</strong> have a signific<strong>an</strong>t impact on <strong>the</strong> climate of local <strong>an</strong>d<br />

remote regions, our underst<strong>an</strong>ding of <strong>the</strong> relationship between snow cover <strong>an</strong>d climate is still<br />

incomplete. This is especially <strong>the</strong> c<strong>as</strong>e in <strong>the</strong> nor<strong>the</strong>rn Great Plains of North America where <strong>the</strong>re<br />

are a paucity of studies examining <strong>the</strong> relationship between snowfall/snow cover <strong>an</strong>d summer<br />

precipitation.<br />

While Koster et al. (2004) found a strong coupling between soil moisture <strong>an</strong>d precipitation in<br />

<strong>the</strong> Great Plains, to date no observational studies have examined <strong>the</strong> relationship between<br />

winter/spring snowfall (which contributes to soil moisture recharge) <strong>an</strong>d summer drought<br />

1 Department of Geography, Tex<strong>as</strong> A&M University, College Station, Tex<strong>as</strong>, USA<br />

2 Department of Geography, University of Delaware, Newark, Delaware, USA


conditions. Therefore <strong>the</strong> utility of l<strong>an</strong>d surface conditions for making se<strong>as</strong>onal climate forec<strong>as</strong>ts<br />

merits fur<strong>the</strong>r attention in this region. This paper utilizes observational data (1929 to 1999) to<br />

examine <strong>the</strong> relationship between winter/spring snowfall <strong>an</strong>omalies <strong>an</strong>d summer moisture<br />

(precipitation) <strong>an</strong>omalies in <strong>the</strong> nor<strong>the</strong>rn Great Plains of North America.<br />

DATA AND METHODS<br />

The nor<strong>the</strong>rn Great Plains of North America, <strong>as</strong> defined in this study, include portions of three<br />

C<strong>an</strong>adi<strong>an</strong> provinces (M<strong>an</strong>itoba, S<strong>as</strong>katchew<strong>an</strong>, <strong>an</strong>d Alberta) <strong>an</strong>d 12 US states (Colorado, Iowa,<br />

Idaho, Minnesota, Missouri, Mont<strong>an</strong>a, Nebr<strong>as</strong>ka, North Dakota, Nevada, South Dakota, Utah, <strong>an</strong>d<br />

Wyoming). The <strong>an</strong>alysis is b<strong>as</strong>ed on monthly drought index <strong>an</strong>d snowfall data (1929–1999) that<br />

have been interpolated to a one-degree grid sp<strong>an</strong>ning 40° to 54° N; 95° to 113° W (18 grid cells in<br />

<strong>the</strong> nor<strong>the</strong><strong>as</strong>tern corner of <strong>the</strong> study region were omitted due to inhomogenities in <strong>the</strong> data).<br />

Drought (moisture) indices are commonly used to qu<strong>an</strong>tify moisture conditions within a region,<br />

to detect <strong>the</strong> onset, <strong>an</strong>d to me<strong>as</strong>ure <strong>the</strong> severity <strong>an</strong>d spatial extent of drought events (Alley, 1984).<br />

The Moisture Anomaly Index (subsequently referred to <strong>as</strong> <strong>the</strong> Z-index) w<strong>as</strong> developed by Palmer<br />

(1965) <strong>an</strong>d it is calculated using a soil moisture/water bal<strong>an</strong>ce algorithm. The Z-index represents<br />

<strong>the</strong> departure from normal (or climatically appropriate) moisture conditions in a given month,<br />

when <strong>the</strong> Z-index is positive (negative) conditions are wetter (drier) th<strong>an</strong> normal (Palmer, 1965).<br />

The Z-index w<strong>as</strong> selected to represent summer moisture <strong>an</strong>omalies in <strong>the</strong> nor<strong>the</strong>rn Great Plains<br />

since previous research h<strong>as</strong> shown that this index is well-suited for monitoring moisture (drought)<br />

conditions in this region (Quiring <strong>an</strong>d Papakyriakou, 2003). Drought data for <strong>the</strong> US were<br />

provided by <strong>the</strong> National Climatic Data Center at <strong>the</strong> climate division level (available at:<br />

http://www.ncdc.noaa.gov) <strong>an</strong>d <strong>the</strong>n interpolated to a one-degree grid. Details on how <strong>the</strong><br />

C<strong>an</strong>adi<strong>an</strong> Z-index data were generated c<strong>an</strong> be found in Quiring <strong>an</strong>d Papkyriakou (2003, 2005).<br />

The snowfall data were developed by T. Mote <strong>an</strong>d collaborators at <strong>the</strong> University of Georgia (T.<br />

Mote, 2004, personal communication). It is b<strong>as</strong>ed on daily observations from <strong>the</strong> National<br />

Wea<strong>the</strong>r Service cooperative station network <strong>an</strong>d <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Daily surface observations. The<br />

data were interpolated to a one-degree latitude by one-degree longitude grid using Spheremap, a<br />

spatial interpolation program.<br />

RESULTS<br />

Only a weak linear relationship exists between snowfall <strong>an</strong>d summer moisture when conditions<br />

are averaged over <strong>the</strong> study area (248 grid cells). Correlations are weakest between fall (SON)<br />

snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies (0.05), <strong>the</strong>y incre<strong>as</strong>e to 0.17 during winter (DJF) <strong>an</strong>d<br />

spring (MAM). During <strong>the</strong> final two months of spring (April <strong>an</strong>d May) <strong>the</strong> correlation improves to<br />

0.22. Averaging over space <strong>an</strong>d time (1929 to 1999) m<strong>as</strong>ks signific<strong>an</strong>t spatial <strong>an</strong>d temporal<br />

variability present in <strong>the</strong> snowfall-summer moisture relationship.<br />

Figure 1 shows <strong>the</strong> relationship between April–May snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies<br />

in <strong>the</strong> nor<strong>the</strong>rn Great Plains (1929–1999). It is clear that <strong>the</strong> nature of this relationship h<strong>as</strong> varied<br />

over time. Linear correlations were calculated using a sliding 15 yr window (e.g., first correlation<br />

w<strong>as</strong> calculated using 1929 to 1944, <strong>the</strong> second correlation w<strong>as</strong> calculated using 1930 to 1945, <strong>an</strong>d<br />

<strong>the</strong> l<strong>as</strong>t correlation w<strong>as</strong> calculated using 1984 to 1999) (Figure 2). Correlations between April–<br />

May snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies varied from 0.82 (1971–1985) to –0.01 (1955–<br />

1969). The relationship w<strong>as</strong> quite strong in <strong>the</strong> 1930s/1940s <strong>an</strong>d 1970s/1980s, but w<strong>as</strong> relatively<br />

weak in <strong>the</strong> 1950s/1960s, <strong>an</strong>d since <strong>the</strong> late 1980s. Correlations between winter snowfall <strong>an</strong>d<br />

summer moisture <strong>an</strong>omalies varied from 0.58 (1954–1968) to –0.54 (1942–1956). During <strong>the</strong><br />

early part of <strong>the</strong> record winter snowfall w<strong>as</strong> negatively correlated with summer moisture. A<br />

signific<strong>an</strong>t shift in <strong>the</strong> relationship occurred during <strong>the</strong> 1950s <strong>an</strong>d since <strong>the</strong> 1950s <strong>the</strong>re h<strong>as</strong><br />

generally been a positive relationship between winter snowfall <strong>an</strong>d summer moisture. The<br />

relationship between winter snowfall <strong>an</strong>d summer moisture is different th<strong>an</strong> <strong>the</strong> relationship<br />

between April–May snowfall <strong>an</strong>d summer moisture. Typically when <strong>the</strong>re are relatively strong<br />

66


correlations between April–May snowfall <strong>an</strong>d summer moisture, <strong>the</strong>re are weak (or negative)<br />

correlations between winter snowfall <strong>an</strong>d summer moisture.<br />

A more detailed examination revealed that <strong>the</strong>re is a stronger relationship between April–May<br />

snowfall <strong>an</strong>d summer moisture during years that have large snowfall <strong>an</strong>omalies (snowfall<br />

<strong>an</strong>omalies that are more th<strong>an</strong> one st<strong>an</strong>dard deviation above/below <strong>the</strong> me<strong>an</strong>). There are 11 yrs with<br />

April–May snowfall <strong>an</strong>omalies more th<strong>an</strong> one st<strong>an</strong>dard deviation below <strong>the</strong> me<strong>an</strong>. Eight of <strong>the</strong>se<br />

11 yrs were followed by drier th<strong>an</strong> normal summers in <strong>the</strong> nor<strong>the</strong>rn Great Plains <strong>an</strong>d <strong>the</strong> me<strong>an</strong> Zindex<br />

for <strong>the</strong> 11 yrs is –0.44. There were also 11 yrs with April–May snowfall <strong>an</strong>omalies more<br />

th<strong>an</strong> one st<strong>an</strong>dard deviation above <strong>the</strong> me<strong>an</strong>. Nine of <strong>the</strong>se 11 yrs were followed by wetter th<strong>an</strong><br />

normal summers <strong>an</strong>d <strong>the</strong> me<strong>an</strong> Z-index for <strong>the</strong> 11 yrs is 0.46. B<strong>as</strong>ed on all 22 yrs, <strong>the</strong> linear<br />

correlation between April–May snowfall <strong>an</strong>omalies <strong>an</strong>d summer moisture is 0.49 (statistically<br />

signific<strong>an</strong>t at <strong>the</strong> 95% confidence level).<br />

Figure 1. April–May snowfall <strong>an</strong>omalies (blue line) <strong>an</strong>d summer moisture <strong>an</strong>omalies (Z-index) (red line)<br />

averaged over <strong>the</strong> nor<strong>the</strong>rn Great Plains study region (1929–1999). The two 15 yr periods with <strong>the</strong> highest<br />

(0.82) <strong>an</strong>d lowest (–0.01) linear correlations are indicated.<br />

67


Figure 2. Linear correlations between winter (DJF) (green line) <strong>an</strong>d April–May (blue line) snowfall<br />

<strong>an</strong>omalies <strong>an</strong>d summer moisture <strong>an</strong>omalies (Z-index) calculated for all 15 yr time periods between 1929 <strong>an</strong>d<br />

1999. D<strong>as</strong>hed lines indicate <strong>the</strong> 95% signific<strong>an</strong>ce level.<br />

The relationship between April–May snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies is fur<strong>the</strong>r<br />

streng<strong>the</strong>ned if <strong>the</strong> c<strong>an</strong>didate years are restricted to those occurring during <strong>the</strong> two periods (1929–<br />

1954 <strong>an</strong>d 1970–1987) when l<strong>an</strong>d surface conditions appear to be <strong>the</strong> domin<strong>an</strong>t source of se<strong>as</strong>onal<br />

climate predictability. During <strong>the</strong>se two periods <strong>the</strong>re were 8 yrs with April–May snowfall<br />

<strong>an</strong>omalies that were more th<strong>an</strong> one st<strong>an</strong>dard deviation below <strong>the</strong> me<strong>an</strong>. Seven of <strong>the</strong>se 8 yrs were<br />

<strong>as</strong>sociated with drier th<strong>an</strong> normal moisture conditions during <strong>the</strong> summer (me<strong>an</strong> Z-index = –0.73).<br />

There were also 8 yrs with April–May snowfall <strong>an</strong>omalies that were more th<strong>an</strong> one st<strong>an</strong>dard<br />

deviation above <strong>the</strong> me<strong>an</strong>. Seven of <strong>the</strong> 8 yrs with large positive April–May snowfall <strong>an</strong>omalies<br />

were followed by wetter th<strong>an</strong> normal moisture conditions during <strong>the</strong> summer (me<strong>an</strong> Z-index =<br />

0.63). B<strong>as</strong>ed on <strong>the</strong>se 16 yrs <strong>the</strong> linear correlation between April–May snowfall <strong>an</strong>d summer<br />

moisture is 0.70 (statistically signific<strong>an</strong>t at <strong>the</strong> 95% confidence level).<br />

Signific<strong>an</strong>t spatial variability is also evident in <strong>the</strong> relationship between April–May snowfall<br />

<strong>an</strong>d summer moisture <strong>an</strong>omalies (Figure 3). Linear correlations across <strong>the</strong> study region are<br />

generally positive with <strong>the</strong> highest correlations (up to 0.63) in sou<strong>the</strong>rn M<strong>an</strong>itoba <strong>an</strong>d South<br />

Dakota. There is only one grid cell in Wyoming where <strong>the</strong>re is a statistically signific<strong>an</strong>t negative<br />

correlation (–0.23). Statistically signific<strong>an</strong>t correlations between April–May snowfall <strong>an</strong>d summer<br />

moisture <strong>an</strong>omalies are found in approximately 56% of <strong>the</strong> study area. These grid cells tend to be<br />

concentrated on <strong>the</strong> e<strong>as</strong>tern side of <strong>the</strong> study region <strong>an</strong>d are notably absent from most of<br />

Wyoming, <strong>the</strong> e<strong>as</strong>tern part of Mont<strong>an</strong>a, <strong>an</strong>d sou<strong>the</strong>rn S<strong>as</strong>katchew<strong>an</strong>.<br />

68


Figure 3. Linear correlations between April–May snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies (Z-index).<br />

Colored grid cells are those with statistically signific<strong>an</strong>t correlations (95% signific<strong>an</strong>ce level).<br />

B<strong>as</strong>ed on a composite <strong>an</strong>alysis, <strong>the</strong> five driest summers between 1929 <strong>an</strong>d 1999 (Table 1) are<br />

<strong>as</strong>sociated with a me<strong>an</strong> winter (spring) snowfall <strong>an</strong>omaly of –66.7 mm (–62.4 mm) (Figure 4).<br />

Approximately 85% of <strong>the</strong> study region received below normal snowfall during <strong>the</strong> winter <strong>an</strong>d<br />

spring se<strong>as</strong>ons prior to <strong>the</strong> five driest summers. About 28% (25%) of <strong>the</strong> study region had winter<br />

(spring) snowfall <strong>an</strong>omalies that are more th<strong>an</strong> 100 mm below average <strong>an</strong>d only 2% (1%) received<br />

winter (spring) snowfall that w<strong>as</strong> more th<strong>an</strong> 100 mm above normal. The five wettest summers<br />

between 1929 <strong>an</strong>d 1999 were <strong>as</strong>sociated with a me<strong>an</strong> winter (spring) snowfall <strong>an</strong>omaly of 6.2 mm<br />

(21.6 mm). Approximately 53% (46%) of <strong>the</strong> study region received above normal snowfall during<br />

<strong>the</strong> prior winter (spring). About 15% (12.1%) of <strong>the</strong> study region had winter (spring) snowfall that<br />

w<strong>as</strong> greater th<strong>an</strong> 100 mm above normal <strong>an</strong>d only 7% (1%) received winter (spring) snowfall that<br />

w<strong>as</strong> more th<strong>an</strong> 100 mm below normal. Results of <strong>the</strong> composite <strong>an</strong>alysis indicate that <strong>an</strong>omalously<br />

dry (wet) summers are <strong>as</strong>sociated with signific<strong>an</strong>t negative (positive) snowfall <strong>an</strong>omalies during<br />

<strong>the</strong> preceding winter <strong>an</strong>d spring, which supports <strong>the</strong> results of <strong>the</strong> correlation <strong>an</strong>alysis. However,<br />

<strong>the</strong> composite <strong>an</strong>alysis demonstrated that <strong>the</strong> winter/spring snowfall <strong>an</strong>omalies <strong>as</strong>sociated with <strong>the</strong><br />

driest summers are typically greater in magnitude <strong>an</strong>d more spatially extensive th<strong>an</strong> <strong>the</strong> snowfall<br />

<strong>an</strong>omalies <strong>as</strong>sociated with wettest summers.<br />

69


Table 1. Ten driest <strong>an</strong>d wettest summers (me<strong>an</strong> Z-index) between 1929 <strong>an</strong>d 1999 (r<strong>an</strong>ked by severity).<br />

Driest Years Wettest Years<br />

Year Z-Index Year Z-Index<br />

1961 –2.52 1993 4.00<br />

1936 –2.51 1944 1.95<br />

1988 –2.48 1951 1.63<br />

1934 –2.13 1965 1.62<br />

1931 –1.73 1995 1.54<br />

1933 –1.40 1947 1.52<br />

1929 –1.24 1999 1.36<br />

1940 –1.23 1942 1.33<br />

1959 –1.09 1975 1.30<br />

1937 –0.99 1968 1.28<br />

Figure 4. Composite snowfall <strong>an</strong>omalies (mm) in winter, <strong>an</strong>d spring <strong>as</strong>sociated with <strong>the</strong> five wettest summers<br />

(1993, 1944, 1951, 1965, 1995) <strong>an</strong>d <strong>the</strong> five driest summers (1961, 1936, 1988, 1934, 1931).<br />

DISCUSSION<br />

The observational data demonstrate that below (above) normal snowfall in winter/spring is<br />

generally <strong>as</strong>sociated with <strong>an</strong>omalously dry (wet) summers in <strong>the</strong> nor<strong>the</strong>rn Great Plains. It is<br />

hypo<strong>the</strong>sized that below normal snowfall is linked to summer drought via negative soil moisture<br />

<strong>an</strong>omalies in spring <strong>an</strong>d early summer that reduce local moisture recycling. Our findings appear to<br />

support those of Nami<strong>as</strong> (1991), who suggested that reduced soil moisture during <strong>the</strong> late<br />

70


winter/early spring could contribute to a warm, dry summer in <strong>the</strong> region by reducing <strong>the</strong> amount<br />

of local moisture recycling <strong>an</strong>d by modifying <strong>the</strong> large-scale atmospheric circulation.<br />

The strength of <strong>the</strong> relationship between winter/spring snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies<br />

h<strong>as</strong> varied signific<strong>an</strong>tly over space <strong>an</strong>d time. Linear correlations between April–May snowfall<br />

<strong>an</strong>omalies <strong>an</strong>d summer moisture conditions r<strong>an</strong>ged from approximately zero (1955–1969) to 0.82<br />

(1971–1985). O<strong>the</strong>r empirical studies have also found that <strong>the</strong> strength of <strong>the</strong> relationship between<br />

l<strong>an</strong>d surface conditions (e.g., snow cover <strong>an</strong>d soil moisture) <strong>an</strong>d precipitation h<strong>as</strong> varied<br />

signific<strong>an</strong>tly during <strong>the</strong> 20 th century (Gutzler, 2000; Hu <strong>an</strong>d Feng, 2002; Zhu et al., 2005). Hu <strong>an</strong>d<br />

Feng (2004) suggest that <strong>the</strong> relationship between l<strong>an</strong>d surface conditions <strong>an</strong>d precipitation<br />

patterns over <strong>the</strong> North Americ<strong>an</strong> Monsoon region is modulated by sea surface temperature (SST)<br />

<strong>an</strong>omalies in <strong>the</strong> Pacific Oce<strong>an</strong>. They found that when SST <strong>an</strong>omalies were strong (weak), l<strong>an</strong>d<br />

surface conditions tend to have less (more) influence. Therefore it is hypo<strong>the</strong>sized that<br />

atmospheric <strong>an</strong>d/or oce<strong>an</strong>ic forcings are modulating <strong>the</strong> relationship between snowfall <strong>an</strong>d<br />

summer moisture conditions in <strong>the</strong> nor<strong>the</strong>rn Great Plains. However, <strong>the</strong> re<strong>as</strong>ons for <strong>the</strong> differential<br />

influence of winter versus spring (April–May) snowfall (Figure 2) are unknown <strong>an</strong>d merit future<br />

study.<br />

Relationships between April–May snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies also varied<br />

spatially. The strongest relationships were found in sou<strong>the</strong>rn M<strong>an</strong>itoba <strong>an</strong>d South Dakota <strong>an</strong>d<br />

statistically signific<strong>an</strong>t correlations were present across approximately 56% of <strong>the</strong> study region.<br />

Previous research h<strong>as</strong> also demonstrated that <strong>the</strong> coupling between l<strong>an</strong>d surface conditions (e.g.,<br />

soil moisture <strong>an</strong>d snow) <strong>an</strong>d precipitation c<strong>an</strong> be highly spatially variable (Lo <strong>an</strong>d Clark, 2002;<br />

Koster et al., 2004; Dominguez et al., 2006). Our results demonstrate that even within a relatively<br />

small area <strong>the</strong>re c<strong>an</strong> be subst<strong>an</strong>tial differences in <strong>the</strong> strength of <strong>the</strong> relationship between spring<br />

snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies.<br />

The relationship between spring snowfall <strong>an</strong>d summer moisture may be non-linear since it<br />

appears that snowfall <strong>an</strong>omalies must exceed some minimum threshold before <strong>the</strong>y have a<br />

signific<strong>an</strong>t (<strong>an</strong>d consistent) influence on summer moisture conditions. The me<strong>an</strong> correlation<br />

between April–May snowfall <strong>an</strong>d summer moisture <strong>an</strong>omalies incre<strong>as</strong>ed from 0.22 (all years) to<br />

0.49 when only <strong>the</strong> years with snowfall <strong>an</strong>omalies more th<strong>an</strong> one st<strong>an</strong>dard deviation above/below<br />

<strong>the</strong> me<strong>an</strong> were considered.<br />

The lack of spatial <strong>an</strong>d temporal stability in <strong>the</strong> relationship between snowfall <strong>an</strong>d summer<br />

moisture <strong>an</strong>omalies h<strong>as</strong> signific<strong>an</strong>t implications for underst<strong>an</strong>ding <strong>an</strong>d forec<strong>as</strong>ting <strong>the</strong> occurrence<br />

of severe hydrologic events (e.g., floods <strong>an</strong>d droughts). Additional study is needed to identify <strong>the</strong><br />

factors that are responsible for modulating <strong>the</strong> strength of <strong>the</strong> snowfall-summer moisture<br />

relationship over space <strong>an</strong>d time. Although spring snowfall conditions c<strong>an</strong>, in some c<strong>as</strong>es, explain<br />

more th<strong>an</strong> half of <strong>the</strong> vari<strong>an</strong>ce in summer moisture, <strong>the</strong> lack of spatial <strong>an</strong>d temporal stability in this<br />

relationship limits its utility for producing accurate forec<strong>as</strong>ts of summer droughts in <strong>the</strong> nor<strong>the</strong>rn<br />

Great Plains.<br />

ACKNOWLEDGEMENTS<br />

A version of this paper h<strong>as</strong> been submitted to Geophysical Research Letters. The authors would<br />

like to th<strong>an</strong>k Tom Mote for providing <strong>the</strong> snowfall data <strong>an</strong>d D<strong>an</strong> Lea<strong>the</strong>rs for reviewing <strong>an</strong> earlier<br />

version of this m<strong>an</strong>uscript.<br />

REFERENCES<br />

Alley WM. 1984. The Palmer Drought Severity Index: Limitation <strong>an</strong>d <strong>as</strong>sumptions. J. Clim. Appl.<br />

Meteorol. 23: 1100–1109.<br />

Déry SJ, Sheffield J, Wood EF. 2005. Connectivity between Eur<strong>as</strong>i<strong>an</strong> snow cover extent <strong>an</strong>d<br />

C<strong>an</strong>adi<strong>an</strong> snow water equivalent <strong>an</strong>d river discharge. J. Geophys. Res. 110: D23106,<br />

doi:10.1029/2005JD006173.<br />

71


Dominguez F, Kumar P, Li<strong>an</strong>g XZ, Ting MF. 2006. Impact of atmospheric moisture storage on<br />

precipitation recycling. J. Clim. 19: 1513–1530.<br />

Ellis AW, Hawkins TW. 2001. An apparent atmospheric teleconnection between snow cover <strong>an</strong>d<br />

<strong>the</strong> North Americ<strong>an</strong> monsoon. Geophys. Res. Lett. 28: 2653–2656.<br />

F<strong>as</strong>ullo J. 2004. A stratified diagnosis of <strong>the</strong> Indi<strong>an</strong> monsoon–Eur<strong>as</strong>i<strong>an</strong> snow cover relationship. J.<br />

Clim. 17: 1110–1122.<br />

Gutzler DS. 2000. Covariability of spring snowpack <strong>an</strong>d summer rainfall across <strong>the</strong> southwest<br />

United States. J. Clim. 13: 4018–4027.<br />

Hawkins TW, Ellis AW, Skindlov JA, Reigle D. 2002. Intra-<strong>an</strong>nual <strong>an</strong>alysis of <strong>the</strong> North<br />

Americ<strong>an</strong> snow cover-monsoon teleconnection: Se<strong>as</strong>onal forec<strong>as</strong>ting utility. J. Clim. 15:<br />

1743–1753.<br />

Hu Q, Feng S. 2002. Inter<strong>an</strong>nual rainfall variations in <strong>the</strong> North Americ<strong>an</strong> summer monsoon<br />

region: 1900–98. J. Clim. 15: 1189–1202.<br />

Hu Q, Feng S. 2004. Why h<strong>as</strong> <strong>the</strong> l<strong>an</strong>d memory ch<strong>an</strong>ged? J. Clim. 17: 3236–3243.<br />

Koster RD, Dirmeyer PA, Guo Z, Bon<strong>an</strong> GB, Ch<strong>an</strong> E, Cox P, Gordon CT, K<strong>an</strong>ae S, Kowalczyk E,<br />

Lawrence D, Liu P, Lu CH, Malyshev S, McAv<strong>an</strong>ey B, Oleson K, Pitm<strong>an</strong> AJ, Sud YC, Taylor<br />

CM, Verseghy D, V<strong>as</strong>ic R, Xue Y, Yamada T. 2004. Regions of strong coupling between soil<br />

moisture <strong>an</strong>d precipitation. Science 305: 1138–1140.<br />

Koster RD, Suarez MJ. 2001. Soil moisture memory in climate models. J. Hydrometeorol. 2: 558–<br />

570.<br />

Koster RD, Suarez MJ, Higgins RW, V<strong>an</strong> den Dool HM. 2003. Observational evidence that soil<br />

moisture variations affect precipitation. Geophys. Res. Lett. 30(5): 1241,<br />

doi:10.1029/2002GL016571.<br />

Kripal<strong>an</strong>i RH, Kim B-J, Oh J-H, Moon S-E. 2002. Relationship between Soviet snow <strong>an</strong>d Kore<strong>an</strong><br />

rainfall. Int. J. Climatol. 22: 1313–1325.<br />

Lo F, Clark MP. 2002. Relationships between spring snow m<strong>as</strong>s <strong>an</strong>d summer precipitation in <strong>the</strong><br />

southwestern United States <strong>as</strong>sociated with <strong>the</strong> North Americ<strong>an</strong> monsoon system. J. Clim. 15:<br />

1378–1385.<br />

Matsui T, Lakshmi V, Small E. 2003. Links between snow cover, surface skin temperature, <strong>an</strong>d<br />

rainfall variability in <strong>the</strong> North Americ<strong>an</strong> monsoon system. J. Clim. 16: 1821–1829.<br />

Nami<strong>as</strong> J. 1991. Spring <strong>an</strong>d summer 1988 drought over <strong>the</strong> contiguous United States—Causes <strong>an</strong>d<br />

prediction. J. Clim. 4: 54–65.<br />

Palmer WC. 1965. Meteorological drought. Research Paper No. 45, U.S. Wea<strong>the</strong>r Bureau:<br />

W<strong>as</strong>hington, D.C.<br />

Qi<strong>an</strong> B, Saunders MA. 2003. Summer U.K. temperature <strong>an</strong>d its links to preceding Eur<strong>as</strong>i<strong>an</strong> snow<br />

cover, North Atl<strong>an</strong>tic SSTs, <strong>an</strong>d <strong>the</strong> NAO. J. Clim. 16: 4108–4120.<br />

Quiring SM, Papakyriakou TN. 2003. An evaluation of agricultural drought indices for <strong>the</strong><br />

C<strong>an</strong>adi<strong>an</strong> prairies. Ag. For. Meteorol. 118: 49–62.<br />

Quiring SM, Papakyriakou TN. 2005. Characterizing <strong>the</strong> spatial <strong>an</strong>d temporal variability of June–<br />

July moisture conditions in <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> prairies. Int. J. Climatol. 25: 117–138.<br />

Robock A, Mu M, Vinnikov K, Robinson D. 2003. L<strong>an</strong>d surface conditions over Eur<strong>as</strong>ia <strong>an</strong>d<br />

Indi<strong>an</strong> summer monsoon rainfall. J. Geophys. Res. 108(D4): 4131,<br />

doi:10.1029/2002JD002286.<br />

Wu T, Qi<strong>an</strong> Z. 2003. The relation between <strong>the</strong> Tibet<strong>an</strong> winter snow <strong>an</strong>d <strong>the</strong> Asi<strong>an</strong> summer<br />

monsoon <strong>an</strong>d rainfall: An observational investigation. J. Clim. 16: 2038–2051.<br />

Zh<strong>an</strong>g Y, Li T, W<strong>an</strong>g B. 2004. Decadal ch<strong>an</strong>ge of <strong>the</strong> spring snow depth over <strong>the</strong> Tibet<strong>an</strong> plateau:<br />

The <strong>as</strong>sociated circulation <strong>an</strong>d influence on <strong>the</strong> e<strong>as</strong>t Asi<strong>an</strong> summer monsoon. J. Clim. 17:<br />

2780–2793.<br />

Zhu CM, Lettenmaier DP, Cavazos T. 2005. Role of <strong>an</strong>tecedent l<strong>an</strong>d surface conditions on North<br />

Americ<strong>an</strong> monsoon rainfall variability. J. Clim. 18: 3104–3121.<br />

72


<strong>Snow</strong> Remote Sensing<br />

73


This page is intentionally bl<strong>an</strong>k.<br />

74


75<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Estimating Sublimation of Intercepted <strong>an</strong>d Sub-C<strong>an</strong>opy <strong>Snow</strong><br />

Using Eddy Covari<strong>an</strong>ce Systems<br />

NOAH MOLOTCH 1 , PETER BLANKEN 2 , MARK WILLIAMS 2,3 , ANDREW TURNIPSEED 4 ,<br />

RUSSELL MONSON 5 , AND STEVEN MARGULIS 1<br />

ABSTRACT:<br />

Direct me<strong>as</strong>urements of winter water loss due to sublimation were made in a sub-alpine forest in<br />

<strong>the</strong> Rocky Mountains of Colorado. Sub-c<strong>an</strong>opy <strong>an</strong>d over-story eddy covari<strong>an</strong>ce systems indicated<br />

subst<strong>an</strong>tial losses of winter-se<strong>as</strong>on snow accumulation in <strong>the</strong> form of snowpack (0.41 mm d –1 ) <strong>an</strong>d<br />

intercepted snow (0.71 mm d –1 ) sublimation. The partitioning between <strong>the</strong>se over <strong>an</strong>d under story<br />

components of water loss w<strong>as</strong> highly dependent on atmospheric conditions <strong>an</strong>d near-surface<br />

conditions at <strong>an</strong>d below <strong>the</strong> snow / atmosphere interface. High over-story sensible heat fluxes lead<br />

to strong temperature gradients between vegetation <strong>an</strong>d <strong>the</strong> snow-surface, driving subst<strong>an</strong>tial<br />

specific humidity gradients at <strong>the</strong> snow surface <strong>an</strong>d high sublimation rates. Intercepted snowfall<br />

resulted in rapid response of over-story latent heat fluxes, high within-c<strong>an</strong>opy sublimation rates,<br />

<strong>an</strong>d diminished sub-c<strong>an</strong>opy snowpack sublimation. These results indicate that sublimation losses<br />

from <strong>the</strong> under-story snowpack are strongly dependent on <strong>the</strong> partitioning of sensible <strong>an</strong>d latent<br />

heat fluxes in <strong>the</strong> c<strong>an</strong>opy. This compels comprehensive studies of snow sublimation in forested<br />

regions that integrate sub-c<strong>an</strong>opy <strong>an</strong>d over-story processes.<br />

Keywords: vegetation c<strong>an</strong>opy; snow interception; sublimation; Rocky Mountains; eddy covari<strong>an</strong>ce<br />

INTRODUCTION<br />

Sublimation of intercepted snow constitutes a signific<strong>an</strong>t component of <strong>the</strong> overall water<br />

bal<strong>an</strong>ce in m<strong>an</strong>y se<strong>as</strong>onally snow-covered coniferous forests [Essery, et al., 2003; Lundberg <strong>an</strong>d<br />

Halldin, 1994; Pomeroy <strong>an</strong>d Gray, 1995; Schmidt <strong>an</strong>d Troendle, 1992]; sublimation losses are<br />

capable of exceeding 30% of total winter snowfall [Montesi, et al., 2004]. For a given c<strong>an</strong>opy<br />

structure <strong>an</strong>d snowfall history <strong>the</strong> distribution of radi<strong>an</strong>t <strong>an</strong>d turbulent fluxes dictates sublimation<br />

rates <strong>an</strong>d <strong>the</strong>refore strongly influences <strong>the</strong> magnitude of spring snowmelt <strong>an</strong>d subsequent growingse<strong>as</strong>on<br />

water availability. Interactions between <strong>the</strong>se fluxes <strong>an</strong>d <strong>the</strong> sublimation of intercepted<br />

snow <strong>an</strong>d <strong>the</strong> sub-c<strong>an</strong>opy snowpack are poorly understood in forested mountainous regions [Bales,<br />

et al., 2006]. This knowledge gap <strong>an</strong>d <strong>the</strong> complexity of interactions between <strong>the</strong> snowpack <strong>an</strong>d<br />

vegetation have motivated detailed <strong>an</strong>alyses of m<strong>as</strong>s <strong>an</strong>d energy fluxes between <strong>the</strong> snowpack,<br />

vegetation, <strong>an</strong>d <strong>the</strong> atmosphere [Davis, et al., 1997; Sicart, et al., 2004].<br />

1<br />

Department of Civil <strong>an</strong>d Environmental Engineering, University of California, Los Angeles,<br />

California, 90095.<br />

2<br />

Department of Geography, University of Colorado, Boulder, Colorado, 80309.<br />

3<br />

Institute for Arctic <strong>an</strong>d Alpine Research, University of Colorado, Boulder, Colorado 80309.<br />

4<br />

National Center for Atmospheric Research, Boulder, Colorado, 80305.<br />

5<br />

Department of Ecology <strong>an</strong>d Evolutionary Biology; Cooperative Institute for Research in<br />

Environmental Sciences, University of Colorado, Boulder, Colorado 80309


Various techniques have been used to estimate sublimation rates from intercepted snow.<br />

Me<strong>as</strong>urement of <strong>the</strong> components of snow sublimation is particularly challenging in forested terrain<br />

<strong>as</strong> winter-time above-c<strong>an</strong>opy water vapor flux me<strong>as</strong>urements integrate m<strong>as</strong>s loss from intercepted<br />

snow <strong>an</strong>d from <strong>the</strong> sub-c<strong>an</strong>opy snowpack. In this regard, numerous studies have focused on<br />

estimating sublimation losses from snowpacks in unforested are<strong>as</strong>. Similarly, much work h<strong>as</strong> been<br />

devoted toward estimating sublimation losses from intercepted snow [Montesi, et al., 2004;<br />

Pomeroy <strong>an</strong>d Schmidt, 1993; Schmidt <strong>an</strong>d Troendle, 1992]. Lacking is a thorough <strong>an</strong>alysis of <strong>the</strong><br />

proportion of <strong>the</strong>se two different components of snow sublimation at <strong>an</strong> individual site.<br />

Me<strong>as</strong>urement of sublimation from intercepted snow h<strong>as</strong> primarily focused on tree-weighting<br />

techniques [Montesi, et al., 2004; Nakai, et al., 1994; Schmidt, 1991; Schmidt, et al., 1988].<br />

Several factors lead to uncertainty in this approach <strong>an</strong>d toward limiting applicability at <strong>the</strong> st<strong>an</strong>d<br />

scale. First, a somewhat subjective <strong>an</strong>alysis must be used to separate unloading from sublimation.<br />

Second, sublimation of unloaded snow is not considered <strong>an</strong>d thus sublimation losses may be<br />

underestimated [Montesi, et al., 2004]. Third, tree-instability c<strong>an</strong> cause false readings. Finally,<br />

intermittent snowfall events <strong>an</strong>d small trace events c<strong>an</strong> introduce uncertainty, effectively<br />

countering sublimation losses <strong>an</strong>d leading to underestimates in sublimation losses if not<br />

considered. In terms of scaling from individual trees to <strong>the</strong> st<strong>an</strong>d scale, challenges are encountered<br />

with regard to <strong>the</strong> lack of detailed c<strong>an</strong>opy information. This lack of detailed c<strong>an</strong>opy information<br />

also complicates <strong>the</strong> use of models for estimating sublimation losses [Pomeroy, et al., 1998;<br />

Pomeroy <strong>an</strong>d Schmidt, 1993]. All of <strong>the</strong>se limitations could be accounted for in techniques that<br />

integrate all of <strong>the</strong>se processes by me<strong>as</strong>uring above <strong>an</strong>d below c<strong>an</strong>opy water vapor flux.<br />

Adv<strong>an</strong>ces in process-level knowledge have been limited <strong>as</strong> sublimation c<strong>an</strong> occur ei<strong>the</strong>r from<br />

snow intercepted by <strong>the</strong> c<strong>an</strong>opy, <strong>an</strong>d/or from <strong>the</strong> snow that reaches <strong>the</strong> ground. Coniferous forests<br />

c<strong>an</strong> intercept large qu<strong>an</strong>tities of snow, much of which sublimates from <strong>the</strong> c<strong>an</strong>opy <strong>an</strong>d does not<br />

reach <strong>the</strong> ground. Sublimation from <strong>the</strong> below-c<strong>an</strong>opy snowpack is thought to be insignific<strong>an</strong>t due<br />

to <strong>the</strong> low exposed surface area of <strong>the</strong> snowpack <strong>an</strong>d low below-c<strong>an</strong>opy wind speeds. However,<br />

<strong>the</strong>re are potentially large longwave radiation fluxes if <strong>the</strong> c<strong>an</strong>opy above is warm <strong>an</strong>d snow-free,<br />

thus promoting sublimation <strong>an</strong>d/or melting [Woo <strong>an</strong>d Giesbrecht, 2000]. Underst<strong>an</strong>ding <strong>the</strong><br />

bal<strong>an</strong>ce between sublimation from <strong>the</strong> c<strong>an</strong>opy <strong>an</strong>d snowpack is crucial to <strong>as</strong>sist water <strong>an</strong>d forest<br />

m<strong>an</strong>agers, especially in regions where forest thinning treatments are being considered to incre<strong>as</strong>e<br />

water yield.<br />

Direct me<strong>as</strong>urements of winter water loss by sublimation of snow from a subalpine forest in <strong>the</strong><br />

Rocky Mountains of Colorado are presented here. Eddy covari<strong>an</strong>ce instruments were placed both<br />

above <strong>an</strong>d beneath <strong>the</strong> c<strong>an</strong>opy during March <strong>an</strong>d early April 2002; <strong>the</strong> time before melting begins<br />

when winter sublimation is thought to be large due to <strong>the</strong> heavy late-winter snows. The above <strong>an</strong>d<br />

below-c<strong>an</strong>opy me<strong>as</strong>urements allowed sublimation of intercepted snow to be separated from that of<br />

<strong>the</strong> snowpack, <strong>an</strong>d estimates obtained over a much larger sample area th<strong>an</strong> individual trees.<br />

Simult<strong>an</strong>eous me<strong>as</strong>urements of <strong>the</strong> physical properties of <strong>the</strong> snow pack, soil moisture, <strong>as</strong> well <strong>as</strong><br />

carbon dioxide flux me<strong>as</strong>urements ensured that sublimation <strong>an</strong>d not evaporation of melting snow<br />

or tr<strong>an</strong>spiration were being me<strong>as</strong>ured. The specific objectives of this research were to: a)<br />

determine snow sublimation rates in a sub-alpine forest; b) partition snow sublimation into above<br />

<strong>an</strong>d below c<strong>an</strong>opy components; <strong>an</strong>d c) explore relationships between atmospheric <strong>an</strong>d snowpack<br />

conditions, <strong>an</strong>d snow sublimation rates.<br />

STUDY SITE<br />

This work w<strong>as</strong> conducted at <strong>the</strong> Niwot Ridge, Colorado Ameriflux site (40º 1’ 58”N; 105º 32’<br />

47” W), located at <strong>an</strong> elevation of 3050 m approximately 8 km e<strong>as</strong>t of <strong>the</strong> Continental Divide<br />

(Figure 1). The area 1 km 2 e<strong>as</strong>t of <strong>the</strong> tower is dominated by Engelm<strong>an</strong>n spruce (7 trees ha –1 ) <strong>an</strong>d<br />

lodgepole pine (27 trees ha –1 ). Rising at a slope of about 6 – 7º, <strong>the</strong> 1 km 2 area west of <strong>the</strong> tower<br />

contains subalpine fir (16 trees ha –1 ), Engelm<strong>an</strong> Spruce (10 trees ha –1 ) <strong>an</strong>d lodgepole pine (9 trees<br />

ha –1 ). Maximum leaf area index during <strong>the</strong> growing se<strong>as</strong>on is approximately 4.2 m 2 m –2 . C<strong>an</strong>opy<br />

76


height averaged 11.4 m with <strong>an</strong> average gap fraction of 17%. The site is in a state of aggradation,<br />

recovering from logging activities in <strong>the</strong> early part of <strong>the</strong> 20 th century. The hydrology of <strong>the</strong> site is<br />

dominated by moderate snowpacks that account for approximately 80% of total <strong>an</strong>nual water input<br />

to <strong>the</strong> system [Caine, 1995]. The prevailing wind direction is from <strong>the</strong> west, particularly in <strong>the</strong><br />

winter when periods of high wind speeds <strong>an</strong>d neutral atmospheric stability conditions are frequent<br />

[Turnipseed, et al., 2002]. A detailed description of <strong>the</strong> site characteristics c<strong>an</strong> be found in<br />

Turnipseed et al. [2002].<br />

Figure 1. Composite image of Niwot Ridge, LTER site <strong>an</strong>d <strong>the</strong> CU-Ameriflux tower located at C-1.<br />

METHODS<br />

Flux me<strong>as</strong>urements<br />

Water vapor fluxes, (latent heat flux; λE) were calculated <strong>as</strong> 30-min me<strong>an</strong>s of 10-Hz<br />

me<strong>as</strong>urements over a 40 d mid-winter period (DOY 60 – 100, 2002) using <strong>the</strong> eddy covari<strong>an</strong>ce<br />

(EC) method described by Turnipseed et al. [2002]:<br />

v ' v ' w L λ E = ρ<br />

where Lv is <strong>the</strong> latent heat of sublimation, w' is <strong>the</strong> deviations of vertical wind velocity (m s –1 )<br />

from <strong>the</strong> ½-hr me<strong>an</strong>, ρv’ is <strong>the</strong> deviations of <strong>the</strong> water vapor density from <strong>the</strong> ½-hr me<strong>an</strong>. The<br />

over-story <strong>an</strong>d under-story EC systems were mounted at a height of 21.5 m <strong>an</strong>d 1.7 m aboveground,<br />

respectively, from towers separated by a dist<strong>an</strong>ce of approximately 20-m. The over- <strong>an</strong>d<br />

under-story EC systems <strong>an</strong>d o<strong>the</strong>r meteorological instruments are summarized in Table 1.<br />

Components of snow sublimation were computed <strong>as</strong>:<br />

λEc,t = λEc,s + λEc,i<br />

where λEc,t is <strong>the</strong> total sublimation from <strong>the</strong> system me<strong>as</strong>ured using <strong>the</strong> over-story EC instruments<br />

(21.5 m above ground) <strong>an</strong>d λEc,s is snowpack sublimation determined from <strong>the</strong> sub-c<strong>an</strong>opy EC<br />

instruments (1.7 m above ground). Water vapor fluxes <strong>as</strong>sociated with sublimation of intercepted<br />

snow, λEc,i were determined <strong>as</strong> <strong>the</strong> difference of me<strong>as</strong>ured over-story <strong>an</strong>d sub-c<strong>an</strong>opy fluxes.<br />

Me<strong>as</strong>urements of <strong>the</strong> above-c<strong>an</strong>opy CO2 flux were used to confirm that photosyn<strong>the</strong>sis from <strong>the</strong><br />

forest c<strong>an</strong>opy w<strong>as</strong> negligible (i.e. values were positive indicating c<strong>an</strong>opy respiration but no carbon<br />

uptake) <strong>an</strong>d <strong>the</strong>refore over-story water flux observations could be inferred to be <strong>entire</strong>ly <strong>as</strong>sociated<br />

with snow sublimation since tr<strong>an</strong>spiration w<strong>as</strong> insignific<strong>an</strong>t.<br />

77<br />

(1)<br />

(2)


Atmospheric stability w<strong>as</strong> calculated by dividing <strong>the</strong> Monin-Obukhov length, L [Monin <strong>an</strong>d<br />

Obukhov, 1954] into <strong>the</strong> me<strong>as</strong>urement height (z):<br />

u * × z × ( T ) × c p ( e,<br />

p)<br />

× T<br />

L =<br />

k × z × g × H<br />

3<br />

ρ<br />

where u* is <strong>the</strong> friction velocity (m s –1 ), ρ(T) is <strong>the</strong> air density <strong>as</strong> a function of air temperature (T)<br />

(Kelvin), cp is <strong>the</strong> specific heat of dry air (kJ kg –1 K –1 ) <strong>as</strong> a function of vapor pressure, e (kPa), <strong>an</strong>d<br />

barometric pressure, p (kPa), k is von Karm<strong>an</strong>’s const<strong>an</strong>t (0.41), g is acceleration due to gravity<br />

9.81 (m s –2 ), <strong>an</strong>d H is <strong>the</strong> sensible heat flux (W m –2 (3)<br />

). Negative z/L values correspond to unstable<br />

atmospheric conditions, positive values represent stable conditions, <strong>an</strong>d values near 0 are neutral.<br />

Table 1. Observations <strong>an</strong>d instruments on <strong>the</strong> above <strong>an</strong>d below c<strong>an</strong>opy towers at <strong>the</strong> University of<br />

Colorado, Ameriflux site.<br />

observation<br />

me<strong>as</strong>urement height,<br />

meters<br />

instrument<br />

relative humidity, % 21.5 HMP-35D, Vaisala, Inc.<br />

air temperature, ºC 21.5 | 1.7 CSAT-3, Cambell Scientific<br />

pressure, kpa 18 PT101B, Vaisala, Inc.<br />

net radiation, W m -2 26 4-component CNR-1, Kipp & Zonen<br />

H2O flux, mg m -2 s -1 21.5 | 1.7 IRGA-6260, Li-Cor<br />

CO 2 flux, mg m -2 s -1 21.5 IRGA-6260, Li-Cor<br />

wind speed, m s -1 21.5 | 1.7 propv<strong>an</strong>e-09101, RM Young Inc.<br />

wind direction, degrees 21.5 | 1.7 propv<strong>an</strong>e-09101, RM Young Inc.<br />

precipitation, mm 12 385-L, Met One<br />

soil heat flux, W m -2<br />

-0.07 - -0.1 HFT-1, REBS<br />

soil moisture, % by volume 0 - -.15 CS-615, Campbell Scientific<br />

soil temperature, ºC 0- - 0.1 STP-1, REBS<br />

Note: Above <strong>an</strong>d below c<strong>an</strong>opy eddy covari<strong>an</strong>ce systems were located 21.5 <strong>an</strong>d 1.7 m above ground,<br />

respectively.<br />

Turbulent flux estimates were evaluated by exploring total energy bal<strong>an</strong>ce closure; turbulent<br />

fluxes should be equal to <strong>the</strong> available energy. A linear regression between <strong>the</strong> summation of <strong>the</strong><br />

sensible (H) <strong>an</strong>d latent heat fluxes <strong>an</strong>d <strong>the</strong> difference between <strong>the</strong> net radiation (Rn) <strong>an</strong>d ground<br />

(G) heat flux w<strong>as</strong> developed [Bl<strong>an</strong>ken, et al., 1998; Bl<strong>an</strong>ken, et al., 1997]. The relationship<br />

between <strong>the</strong> 30-min above c<strong>an</strong>opy (λE +H) <strong>an</strong>d (Rn-G) w<strong>as</strong> y = 0.77x + 13 (R 2 = 0.89; p < 0.01)<br />

indicating adequate energy bal<strong>an</strong>ce closure.<br />

The sampling area, or flux footprint, w<strong>as</strong> calculated using <strong>the</strong> method described by Schuepp et<br />

al. [Schuepp, et al., 1990]. The upwind dist<strong>an</strong>ce that <strong>the</strong> understory flux me<strong>as</strong>urements were most<br />

sensitive to occurred at a dist<strong>an</strong>ce of 23, 27, <strong>an</strong>d 29-m during typical daytime, neutral, <strong>an</strong>d<br />

nighttime atmospheric stability conditions, respectively (Figure 2a). The cumulative flux footprint,<br />

indicative of <strong>the</strong> upwind sampling area where 80% of <strong>the</strong> flux originated from, w<strong>as</strong> 207, 243, <strong>an</strong>d<br />

263 m (daytime, neutral, <strong>an</strong>d nighttime atmospheric stability conditions, respectively) (Figure 2b).<br />

Supporting Understory Me<strong>as</strong>urements<br />

Observations of soil, snow, <strong>an</strong>d air temperature from three <strong>the</strong>rmistor strings were used to<br />

develop relationships between snowpack temperature <strong>an</strong>d rates of snowpack sublimation. In this<br />

regard, we investigated relationships between snowpack temperature gradients <strong>an</strong>d diurnal<br />

variability in snow temperature, <strong>an</strong>d rates of snowpack sublimation; snowpack temperature<br />

gradients control vapor pressure gradients in <strong>the</strong> snowpack <strong>an</strong>d <strong>the</strong>refore <strong>the</strong> movement of water<br />

vapor from deeper in <strong>the</strong> snowpack toward <strong>the</strong> snowpack / atmosphere interface [McClung <strong>an</strong>d<br />

Schaerer, 1993]. The three <strong>the</strong>rmistor strings were placed along a tr<strong>an</strong>sect through a small clearing<br />

(~ 6 m in diameter) in <strong>the</strong> forest adjacent to <strong>the</strong> understory flux tower (Figure 3). The <strong>the</strong>rmistor<br />

78


strings were buried 20 – 30 cm below <strong>the</strong> soil before snow accumulation beg<strong>an</strong> <strong>an</strong>d extended to<br />

80, 180, <strong>an</strong>d 200 cm above <strong>the</strong> ground surface; a guy wire tied to two trees at opposite ends of <strong>the</strong><br />

clearing w<strong>as</strong> used to te<strong>the</strong>r <strong>the</strong> tops of <strong>the</strong> <strong>the</strong>rmistor strings. During <strong>the</strong> study period <strong>the</strong><br />

<strong>the</strong>rmistor strings provided observations of soil, snow, <strong>an</strong>d air temperature.<br />

16<br />

12<br />

8<br />

4<br />

a<br />

day<br />

neutral<br />

night<br />

0<br />

0<br />

0 100 200 300 400 500 0 100 200 300 400 500<br />

upwind dist<strong>an</strong>ce, m<br />

Figure 2. (a) Normalized ch<strong>an</strong>ge in <strong>the</strong> understory turbulent flux (Q) with upwind dist<strong>an</strong>ce (x) during typical<br />

daytime (green), neutral (blue) <strong>an</strong>d nighttime (red) atmospheric stability conditions. (b) Cumulative ch<strong>an</strong>ge in<br />

<strong>the</strong> understory turbulent flux (Q) with upwind dist<strong>an</strong>ce (x) for typical daytime (green), neutral (blue) <strong>an</strong>d<br />

nighttime (red) atmospheric stability conditions.<br />

Eight water content reflectometers (Campbell Scientific model CS-615) were used to monitor<br />

soil moisture conditions surrounding <strong>the</strong> towers (Figure 3). These observations were used to<br />

ensure that latent heat fluxes were primarily allocated to sublimation <strong>as</strong> opposed to snowmelt <strong>an</strong>d<br />

to confirm that water from snowmelt had not entered <strong>the</strong> soil horizon which might trigger <strong>the</strong><br />

onset of tr<strong>an</strong>spiration.<br />

<strong>Snow</strong>pack properties<br />

Ground observations of snow depth <strong>an</strong>d snow density were derived from snow pits excavated<br />

weekly at two different locations (sub-c<strong>an</strong>opy <strong>an</strong>d within a small clearing adjacent to <strong>the</strong> flux<br />

towers). Within each snowpit, samples were taken at 10 cm vertical intervals over <strong>the</strong> <strong>entire</strong><br />

snowpit depth using a 1000 cc stainless steel cutter. <strong>Snow</strong> density stratigraphy <strong>an</strong>d bulk density<br />

<strong>an</strong>d snow water equivalent were calculated from weighted-average density values <strong>an</strong>d total<br />

snowpack depth.<br />

Observations of precipitation were used to determine <strong>the</strong> m<strong>as</strong>s input between <strong>the</strong> weekly<br />

snowpit observations, allowing us to approximate sublimation losses; ch<strong>an</strong>ges in snow water<br />

equivalent between <strong>the</strong> weekly snowpit observations result from input of m<strong>as</strong>s due to snowfall <strong>an</strong>d<br />

reduction in m<strong>as</strong>s due to sublimation. This provides a field b<strong>as</strong>ed technique for evaluating<br />

sublimation estimates from <strong>the</strong> sub-c<strong>an</strong>opy EC system. Precipitation observations were obtained at<br />

a height of 12 m from <strong>the</strong> above-c<strong>an</strong>opy EC tower; <strong>an</strong> Alter gauge shield w<strong>as</strong> used to improve<br />

precipitation gauge catch efficiency.<br />

79<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

b


Figure 3. Location of above <strong>an</strong>d below c<strong>an</strong>opy flux towers <strong>an</strong>d supporting ground b<strong>as</strong>ed instruments. Water<br />

content reflectometers (M), <strong>the</strong>rmistor strings, ground <strong>the</strong>rmistors (T), <strong>an</strong>d soil heat flux plates (H).<br />

Additional H, M <strong>an</strong>d T me<strong>as</strong>urements were made along e<strong>as</strong>t / west (T ew) <strong>an</strong>d north / south (T NS) tr<strong>an</strong>sects.<br />

RESULTS<br />

Soil temperature, moisture <strong>an</strong>d ground heat flux were consistent with mid-winter conditions<br />

throughout <strong>the</strong> study period (Figure 4). Temporal variability in soil temperature (coefficient of<br />

variation = .77) <strong>an</strong>d ground heat flux (coefficient of variation = 2.6) w<strong>as</strong> considerably greater th<strong>an</strong><br />

that of soil moisture (coefficient of variation = 0.11). Spring onset of snowmelt percolation<br />

occurred on DOY 100; soil moisture incre<strong>as</strong>ed by threefold over <strong>the</strong> subsequent 20-d period.<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

e<strong>as</strong>t / west dist<strong>an</strong>ce, meters<br />

<strong>Snow</strong> temp.<br />

<strong>the</strong>rmistor strings<br />

Lodgepole Pine<br />

Engelm<strong>an</strong>n Spruce<br />

Subalpine Fir<br />

sub-c<strong>an</strong>opy tower<br />

25<br />

20<br />

15<br />

10<br />

MT HMT<br />

HM<br />

5<br />

60 80 100 120 60 80 100 120<br />

day of year<br />

80<br />

1<br />

.5<br />

0<br />

-.5<br />

-1<br />

60 80 100 120<br />

Figure 4. Time series of soil temperature, moisture, <strong>an</strong>d ground heat flux from day of year 60 – 120, 2002.<br />

The diurnal energy fluxes, Rn, λE <strong>an</strong>d H above <strong>an</strong>d below <strong>the</strong> c<strong>an</strong>opy are shown in Figure 5,<br />

toge<strong>the</strong>r with precipitation. The CO2 flux above <strong>the</strong> c<strong>an</strong>opy is included to show that <strong>the</strong> forests had<br />

not yet tr<strong>an</strong>sitioned from losing to gaining carbon, <strong>an</strong>d <strong>the</strong>refore tr<strong>an</strong>spiration at this time w<strong>as</strong><br />

negligible. The majority of <strong>the</strong> above-c<strong>an</strong>opy net radiation w<strong>as</strong> partitioned <strong>as</strong> H above <strong>the</strong> c<strong>an</strong>opy,<br />

<strong>an</strong>d <strong>as</strong> λE beneath <strong>the</strong> c<strong>an</strong>opy; above-c<strong>an</strong>opy ratios of <strong>the</strong> daytime me<strong>an</strong> H/Rn <strong>an</strong>d λE/Rn were<br />

0.67 <strong>an</strong>d 0.16, respectively. Beneath <strong>the</strong> c<strong>an</strong>opy, <strong>the</strong>se ratios were 0.02 (H/Rn) <strong>an</strong>d 0.06 (λE/Rn).<br />

HMT<br />

north / south dist<strong>an</strong>ce, meters<br />

T EW<br />

above-c<strong>an</strong>opy tower<br />

T NS<br />

HM<br />

N


Although <strong>the</strong> λE/Rn fraction w<strong>as</strong> on average relatively small, large incre<strong>as</strong>es in λE with a<br />

subsequent decre<strong>as</strong>e in H occurred several times in response to snowfall events.<br />

Average sublimation rates over <strong>the</strong> study period were 0.7 <strong>an</strong>d 0.41 mm d –1 for intercepted snow<br />

<strong>an</strong>d <strong>the</strong> sub-c<strong>an</strong>opy snowpack, respectively. Both fluxes exhibited considerable variability<br />

(coefficient of variation = 0.66 for both total sublimation <strong>an</strong>d snowpack sublimation), with<br />

intercepted snow sublimation rising after snowfall events (Figure 6). The ratio between subc<strong>an</strong>opy<br />

snowpack sublimation <strong>an</strong>d total sublimation averaged 0.45 during <strong>the</strong> study period,<br />

incre<strong>as</strong>ing with time after snowfall <strong>an</strong>d approaching 1 during consecutive days without snowfall;<br />

e.g. DOY 63 – 65 <strong>an</strong>d DOY 87 – 93 (Figure 6). On average snowpack to total sublimation ratios<br />

peaked 3 days after snowfall; timing to peak varied considerably with snowfall magnitude.<br />

day of year 2002 (MST)<br />

Figure 5. Diurnal variability in net radiation (R n), sensible (H) <strong>an</strong>d latent (λE) heat fluxes, carbon flux (F c),<br />

<strong>an</strong>d precipitation (P) me<strong>as</strong>ured above <strong>the</strong> c<strong>an</strong>opy (blue lines) <strong>an</strong>d beneath <strong>the</strong> c<strong>an</strong>opy (green lines). Period<br />

shown is from DOY 60 – 100 2002. Positive values represent fluxes toward <strong>the</strong> surface.<br />

A total of 34.8 mm of snow fell during <strong>the</strong> me<strong>as</strong>urement period (Fig 7). 38.5 mm of sublimation<br />

w<strong>as</strong> me<strong>as</strong>ured above <strong>the</strong> c<strong>an</strong>opy over <strong>the</strong> same time period, <strong>an</strong>d 14.8 mm sublimated from <strong>the</strong><br />

snowpack at <strong>the</strong> forest floor. These correspond to sublimation to precipitation ratios of 1.11 (total)<br />

<strong>an</strong>d 0.43 (snowpack), with <strong>the</strong> total ratio exceeding one due to sublimation of snow that fell prior<br />

to <strong>the</strong> start of <strong>the</strong> me<strong>as</strong>urements. Subtracting <strong>the</strong> above-c<strong>an</strong>opy λE me<strong>as</strong>urements from that below<br />

<strong>the</strong> c<strong>an</strong>opy (Figure 7) reveals that 23.7 mm of intercepted snow w<strong>as</strong> sublimated from <strong>the</strong> c<strong>an</strong>opy<br />

itself. This corresponds to a sublimation to precipitation ratio of 0.68.<br />

Diurnal fluctuations in snowpack <strong>an</strong>d near surface air temperatures were notably different for<br />

time periods with high snowpack sublimation rates. For example, only 0.1 mm of water<br />

81


sublimated from <strong>the</strong> snowpack on DOY 60 where<strong>as</strong> over 0.6 mm sublimated on DOY 64. At 60<br />

cm above <strong>the</strong> ground surface snow temperatures fluctuated by less th<strong>an</strong> 5˚ during DOY 60 <strong>an</strong>d by<br />

more th<strong>an</strong> 10˚ during DOY 64 (Figure 8). Similarly, diurnal variability in snow temperature w<strong>as</strong><br />

signific<strong>an</strong>tly lower on DOY 74 relative to DOY 93; sublimation rates were 0.1 versus 0.6 mm d –1<br />

for <strong>the</strong>se two days, respectively. Temperature fluctuations in <strong>the</strong> surface layers, <strong>as</strong>sociated with<br />

cool nights <strong>an</strong>d warm dry days potentially drive signific<strong>an</strong>t water vapor movement in <strong>the</strong> surface<br />

layers of <strong>the</strong> snowpack <strong>an</strong>d enh<strong>an</strong>ce sublimation rates.<br />

precipitation<br />

total sublimation<br />

snowpack / total sublimation ratio<br />

day of year<br />

82<br />

snowpack sublimation<br />

Figure 6. Time series of daily average sublimation me<strong>as</strong>ured above <strong>the</strong> c<strong>an</strong>opy (blue line) <strong>an</strong>d beneath <strong>the</strong><br />

c<strong>an</strong>opy (red line). Precipitation <strong>an</strong>d <strong>the</strong> ratio of snowpack sublimation to total sublimation are also shown.<br />

40<br />

30<br />

20<br />

10<br />

precipitation<br />

total sublimation<br />

c<strong>an</strong>opy sublimation<br />

snowpack sublimation<br />

0<br />

60 70 80 90 100<br />

day of year<br />

Figure 7. Cumulative sublimation from <strong>the</strong> snowpack, intercepted snow (c<strong>an</strong>opy), <strong>an</strong>d total sublimation<br />

throughout <strong>the</strong> study period. Cumulative precipitation is also shown.<br />

Estimates of snow depth on snow temp profile plots were derived from coincident pit<br />

observations when available. In <strong>the</strong> c<strong>as</strong>e of DOY 60, <strong>the</strong> majority of precipitation w<strong>as</strong> recorded on<br />

DOY 59 <strong>an</strong>d early hours of DOY 60 <strong>an</strong>d <strong>the</strong>refore we <strong>as</strong>sume snow depth equivalent to that<br />

me<strong>as</strong>ured in <strong>the</strong> snowpit on DOY 64. In <strong>the</strong> c<strong>as</strong>e of DOY 74, snow depth w<strong>as</strong> difficult to estimate<br />

<strong>as</strong> <strong>the</strong>re w<strong>as</strong> a large (12.45 mm) snowfall event on DOY 73. Thus, we <strong>as</strong>sumed a snow depth of 80<br />

cm, corresponding to <strong>the</strong> observed snow depth from <strong>the</strong> snowpit on DOY 84. For DOY 93, we<br />

me<strong>an</strong>


estimated snow depth b<strong>as</strong>ed on <strong>the</strong> 2:00 temperature curve which showed a distinct inflection<br />

point at <strong>the</strong> snow–atmosphere interface.<br />

Above <strong>an</strong>d below c<strong>an</strong>opy friction velocities were considerably greater for DOY 64 <strong>an</strong>d 93<br />

relative to that on DOY 60 <strong>an</strong>d 74 (Figure 9). The combination of <strong>the</strong> relatively high air<br />

temperatures on <strong>the</strong>se days with sufficient turbulence lead to enh<strong>an</strong>ced near-surface gradients in<br />

specific humidity <strong>an</strong>d sublimation.<br />

80<br />

60<br />

40<br />

20<br />

a. DOY = 60<br />

atmosphere<br />

b. DOY = 64 c. DOY = 74 d. DOY = 93<br />

snow<br />

0<br />

soil<br />

-20<br />

-25 -20 -15 -10 -5 0 -12 -8 -4 0 -20 -15 -10 -5 0 -10 -5 0 5 10 15<br />

temperature, C<br />

Figure 8. <strong>Snow</strong>pack, soil, <strong>an</strong>d air temperature profiles from 80 cm above <strong>the</strong> ground surface to 20 cm below.<br />

Profiles are shown for 4 different days for hours 2, 8, 14, <strong>an</strong>d 20 MST. Dotted horizontal lines represent <strong>the</strong><br />

snowpack / atmosphere interface. Solid horizontal lines indicate <strong>the</strong> soil / snowpack interface.<br />

below c<strong>an</strong>opy<br />

above c<strong>an</strong>opy<br />

83<br />

hour<br />

X 20:00<br />

14:00<br />

8:00<br />

2:00<br />

day of year<br />

Figure 9. Above (blue lines) <strong>an</strong>d below c<strong>an</strong>opy (green lines) diurnal variability in friction velocity <strong>an</strong>d air<br />

temperature for <strong>the</strong> same 4 days shown in Figure 8.


Unstable atmospheric conditions resulted in considerable sublimation of intercepted snow. For<br />

example, on DOY 62 me<strong>as</strong>urement-height / Monin-Obukhov ratios dropped below –200 (Figure<br />

10) <strong>an</strong>d daily sublimation w<strong>as</strong> 2.09 mm (Figure 6). Conversely, me<strong>as</strong>urement-height / Monin-<br />

Obukhov ratios on DOY 76 were less th<strong>an</strong> 0 but greater th<strong>an</strong> –3, suggesting only slight<br />

atmospheric instability. Sublimation of intercepted snow on DOY 76 w<strong>as</strong> 1.74 mm; only 17%<br />

lower th<strong>an</strong> that of DOY 62. Precipitation magnitude is likely responsible for <strong>the</strong>se differences with<br />

12 mm of precipitation falling on DOY 73 <strong>an</strong>d a combined 6 mm of precipitation falling over <strong>the</strong><br />

course of DOY 60 <strong>an</strong>d 61 (Figure 5).<br />

0<br />

-100<br />

-200<br />

5<br />

0<br />

-300<br />

60 65 70 75 80 85 90 95 100<br />

day of year<br />

84<br />

24-hour<br />

moving average<br />

-5<br />

60 70 80 90 100<br />

Figure 10. Me<strong>as</strong>urement-height / Monin-Obukhov-length ratios calculated from 30-minute, above-c<strong>an</strong>opy<br />

observations. See equation (3) for derivation of Monin-Obukhov length. Negative <strong>an</strong>d positive values<br />

represent unstable <strong>an</strong>d stable conditions, respectively.<br />

DISCUSSION<br />

A variety of techniques have been developed to estimate sublimation from snowpacks <strong>an</strong>d<br />

intercepted snow [Montesi, et al., 2004; Pomeroy, et al., 1998]. It is especially challenging to<br />

capture <strong>the</strong> impact of vegetation on variability in turbulence <strong>an</strong>d subsequent vapor fluxes. Results<br />

of previous work performed at <strong>the</strong> individual tree scale provide useful values to evaluate results of<br />

our new technique. Comparisons, however, must be made with caution <strong>as</strong> our technique integrates<br />

fluxes over <strong>the</strong> st<strong>an</strong>d scale from two systems with different flux footprints (despite re<strong>as</strong>onably<br />

homogenous st<strong>an</strong>d characteristics); tree scale studies provide limited information at <strong>the</strong> st<strong>an</strong>d scale<br />

due to introduction of uncertainty <strong>as</strong>sociated with vegetation properties. Fur<strong>the</strong>r, qu<strong>an</strong>titative<br />

comparison with previous studies is difficult given that meteorological conditions <strong>an</strong>d site specific<br />

attributes c<strong>an</strong> have dramatic impacts on <strong>the</strong> energy bal<strong>an</strong>ce of forested environments – in<br />

particular, variability in vegetation structure [Sicart, et al., 2004]. Here we compare general<br />

observations of both snowpack <strong>an</strong>d intercepted sublimation rates. Average mid-winter snowpack<br />

sublimation rates observed here (0.41 mm d –1 ) were low relative to <strong>the</strong> highest of values found<br />

within <strong>the</strong> literature (1.2 – 1.8 mm d –1 [Pomeroy <strong>an</strong>d Essery, 1999]) <strong>an</strong>d are within 14% of values<br />

observed at <strong>the</strong> nearby Fr<strong>as</strong>er Experimental Forest (e.g. 0.36 mm d –1 ) [Schmidt, et al., 1998].<br />

F<strong>as</strong>snacht [F<strong>as</strong>snacht, 2004] estimated winter sublimation rates at 0.75 mm d –1 at <strong>an</strong> open site in<br />

Leadville, Colorado; open sites are known to exhibit subst<strong>an</strong>tially greater sublimation rates [West,<br />

1962]. The weekly snowpits excavated in a clearing adjacent to <strong>the</strong> flux towers used in this<br />

research indicated a total sublimation rate of 0.8 mm d –1 . While <strong>the</strong>se estimates have inherent<br />

uncertainties, <strong>the</strong>se on-site observations <strong>an</strong>d comparisons with previous studies indicate that<br />

sublimation rates are not being overestimated using <strong>the</strong> sub-c<strong>an</strong>opy EC system. In this regard, it is


import<strong>an</strong>t to note that <strong>the</strong> average snowpack to total sublimation ratio of 0.45 (Figure 6) represents<br />

<strong>the</strong> low-end of <strong>the</strong> contribution of sub-c<strong>an</strong>opy sublimation to overall water loss; a signific<strong>an</strong>t<br />

finding given previous <strong>as</strong>sumptions that sublimation losses in forested systems are primarily <strong>the</strong><br />

result of intercepted snow sublimation [Montesi, et al., 2004].<br />

The <strong>as</strong>sessment of sub-c<strong>an</strong>opy sublimation estimates mentioned above must be considered when<br />

evaluating <strong>the</strong> EC estimates of intercepted snow sublimation <strong>as</strong> <strong>the</strong>y are calculated from <strong>the</strong><br />

residual of total sublimation <strong>an</strong>d sub-c<strong>an</strong>opy sublimation (equation (2)). Sublimation rates of<br />

intercepted snow estimated using our EC approach (0.71 mm d –1 ) compared favorably with<br />

previous works. For example, Parviainen [Parviainen <strong>an</strong>d Pomeroy, 2000] estimated intercepted<br />

snow sublimation from a boreal forest at 0.5 mm d –1 ; at higher latitudes available energy is<br />

diminished due to higher solar zenith <strong>an</strong>gles.<br />

Montesi et al., [2004] explored <strong>the</strong> impact of elevation on sublimation rates <strong>an</strong>d found that<br />

incre<strong>as</strong>ed wind speeds, lower relative humidity <strong>an</strong>d warmer air temperatures contributed to a 23%<br />

incre<strong>as</strong>e in sublimation rates at lower elevation. On average Montesi’s results indicate<br />

considerable differences between our estimates, with sublimation losses equivalent to 20 – 30% of<br />

total snow water equivalent during <strong>the</strong> 21 storms considered. These differences may be due to <strong>an</strong><br />

underestimate in sub-c<strong>an</strong>opy sublimation from <strong>the</strong> sub-c<strong>an</strong>opy EC system used here. Differences<br />

may also be due to previously mentioned sources of underestimates in sublimation using <strong>the</strong> treeweighting<br />

method of Montesi [2004].<br />

As previously found by Niu <strong>an</strong>d Y<strong>an</strong>g [2004], <strong>the</strong> relatively high over-story sensible heat fluxes<br />

lead to strong temperature differences between vegetation <strong>an</strong>d <strong>the</strong> snow-surface, driving strong<br />

specific humidity gradients at <strong>the</strong> snow / atmosphere interface <strong>an</strong>d elevated snowpack sublimation<br />

rates (e.g. DOY 78 – 82 & 88 – 95, Figure 6). When snowfall occurred, over-story available<br />

energy w<strong>as</strong> partitioned into latent heat fluxes (e.g. Figure 5, DOY 74), leading to high withinc<strong>an</strong>opy<br />

sublimation rates but diminished diurnal variability in temperatures at <strong>the</strong> snow –<br />

atmosphere interface (DOY 74, Figure 8). These results indicate that sublimation losses from <strong>the</strong><br />

under-story snowpack is strongly dependent on <strong>the</strong> partitioning of sensible <strong>an</strong>d latent heat fluxes<br />

in <strong>the</strong> c<strong>an</strong>opy.<br />

CONCLUSIONS<br />

Sub-c<strong>an</strong>opy <strong>an</strong>d over-story eddy covari<strong>an</strong>ce systems indicated subst<strong>an</strong>tial losses of winterse<strong>as</strong>on<br />

snow accumulation in <strong>the</strong> form of snowpack (0.41 mm d –1 ) <strong>an</strong>d intercepted snow (0.71 mm<br />

d –1 ) sublimation. The partitioning between <strong>the</strong>se over <strong>an</strong>d under story components of water loss<br />

w<strong>as</strong> highly dependent on atmospheric conditions <strong>an</strong>d near-surface conditions at <strong>an</strong>d below <strong>the</strong><br />

snow – atmosphere interface. High over-story sensible heat fluxes lead to strong temperature<br />

gradients between vegetation <strong>an</strong>d <strong>the</strong> snow-surface, driving subst<strong>an</strong>tial specific humidity gradients<br />

at <strong>the</strong> snow surface <strong>an</strong>d high sublimation rates. Intercepted snowfall resulted in rapid response of<br />

over-story latent heat fluxes, high within-c<strong>an</strong>opy sublimation rates, <strong>an</strong>d diminished sub-c<strong>an</strong>opy<br />

snowpack sublimation. These results indicate that sublimation losses from <strong>the</strong> under-story<br />

snowpack is strongly dependent on <strong>the</strong> partitioning of sensible <strong>an</strong>d latent heat fluxes in <strong>the</strong><br />

c<strong>an</strong>opy. This compels comprehensive studies of snow sublimation in forested regions that<br />

integrate sub-c<strong>an</strong>opy <strong>an</strong>d over-story processes.<br />

ACKNOWLEDGEMENTS<br />

This research w<strong>as</strong> supported by NASA gr<strong>an</strong>t #NNG04GO74G <strong>an</strong>d a visiting research fellowship<br />

awarded to <strong>the</strong> primary author at <strong>the</strong> Cooperative Institute for Research in Environmental<br />

Sciences, University of Colorado at Boulder. Instrumentation infr<strong>as</strong>tructure w<strong>as</strong> developed in part<br />

through <strong>the</strong> Ameriflux network <strong>an</strong>d through support of <strong>the</strong> National Science Foundation, Niwot<br />

Ridge LTER Project. Phillip Jacobson provided technical support. Andrew Rossi is acknowledged<br />

for collecting snow pit data.<br />

85


REFERENCES<br />

Bales, R., N. Molotch, T. Painter, M. Dettinger, R. Rice, <strong>an</strong>d J. Dozier (2006), Mountain<br />

hydrology of <strong>the</strong> western United States., Water Resources Research, 42, W08432,<br />

doi:08410.01029/02005WR004387.<br />

Bl<strong>an</strong>ken, P. D., T. A. Black, H. H. Neum<strong>an</strong>n, G. Den Hartog, P. C. Y<strong>an</strong>g, Z. Nesic, R. Staebler,<br />

W. Chen, <strong>an</strong>d M. D. Novak (1998), Turbulent flux me<strong>as</strong>urements above <strong>an</strong>d below <strong>the</strong><br />

overstory of a boreal <strong>as</strong>pen forest, Boundary-Layer Meteorology, 89, 109–140.<br />

Bl<strong>an</strong>ken, P. D., T. A. Black, P. C. Y<strong>an</strong>g, H. H. Neum<strong>an</strong>n, Z. Nesic, R. Staebler, G. den Hartog, M.<br />

D. Novak, <strong>an</strong>d X. Lee (1997), Energy bal<strong>an</strong>ce <strong>an</strong>d c<strong>an</strong>opy conduct<strong>an</strong>ce of a boreal <strong>as</strong>pen forest:<br />

Partitioning overstory <strong>an</strong>d understory components, Journal Of Geophysical Research-<br />

Atmospheres, 102, 28915–28927.<br />

Caine, N. (1995), Temporal trends in <strong>the</strong> quality of streamwater in <strong>an</strong> Alpine environment: Green<br />

Lakes valley, Colorado Front R<strong>an</strong>d, USA, Geografiska Annaler Series A-Physical Geography,<br />

77A, 207–220.<br />

Davis, R. E., J. P. Hardy, W. Ni, C. E. Woodcock, J. C. Mckenzie, R. Jord<strong>an</strong>, <strong>an</strong>d X. Li (1997),<br />

Variation of snow cover ablation in <strong>the</strong> boreal forest: A sensitivity study on <strong>the</strong> effects of<br />

conifer c<strong>an</strong>opy, Journal of Geophysical Research, 102, 29389–29395.<br />

Essery, R., J. Pomeroy, J. Parviainen, <strong>an</strong>d P. Storck (2003), Sublimation of snow from coniferous<br />

forests in a climate model, Journal Of Climate, 16, 1855–1864.<br />

F<strong>as</strong>snacht, S. R. (2004), Estimating Alter-shielded gauge snowfall undercatch, snowpack<br />

sublimation, <strong>an</strong>d blowing snow tr<strong>an</strong>sport at six sites in <strong>the</strong> coterminous USA, Hydrological<br />

Processes, 18, 3481–3492.<br />

Lundberg, A., <strong>an</strong>d S. Halldin (1994), Evaporation of intercepted snow: Anlaysis of governing<br />

factors., Water Resources Research, 30, 2587–2598.<br />

McClung, D., <strong>an</strong>d P. Schaerer (1993), The Aval<strong>an</strong>che H<strong>an</strong>dbook, 271 pp., The Mountaineers,<br />

Seattle, W<strong>as</strong>hington.<br />

Monin, A. S., <strong>an</strong>d A. M. Obukhov (1954), B<strong>as</strong>ic laws of turbulent mixing in <strong>the</strong> ground layer of<br />

<strong>the</strong> atmosphere., Tr. Inst. Teor. Geofiz. Akad. Nauk SSSR, 1, 95–115.<br />

Montesi, J., K. Elder, R. A. Schmidt, <strong>an</strong>d R. E. Davis (2004), Sublimation of intercepted snow<br />

within a subalpine forest c<strong>an</strong>opy at two elevations, Journal of Hydrometeorology, 5, 763–773.<br />

Nakai, Y. T., T. Sakamoto, T. Terajima, H. Kiahara, <strong>an</strong>d T. Saito (1994), <strong>Snow</strong> interception by<br />

forest c<strong>an</strong>opies: Weighing a conifer tree, meteorological observation <strong>an</strong>d <strong>an</strong>alysis by hte<br />

Penm<strong>an</strong>-Monteith formula, IAHS Publ., 223, 227–236.<br />

Niu, G. Y., <strong>an</strong>d Z. L. Y<strong>an</strong>g (2004), Effects of vegetation c<strong>an</strong>opy processes on snow surface energy<br />

<strong>an</strong>d m<strong>as</strong>s bal<strong>an</strong>ces, Journal Of Geophysical Research-Atmospheres, 109.<br />

Parviainen, J., <strong>an</strong>d J. W. Pomeroy (2000), Multiple-scale modelling of forest snow sublimation:<br />

initial findings, Hydrological Processes, 14, 2669–2681.<br />

Pomeroy, J. W., <strong>an</strong>d R. L. H. Essery (1999), Turbulent fluxes during blowing snow: field tests of<br />

model sublimation predictions, Hydrological Processes, 13, 2963–2975.<br />

Pomeroy, J. W., <strong>an</strong>d D. M. Gray (1995), <strong>Snow</strong>cover accumulation, relocation <strong>an</strong>d m<strong>an</strong>agement,<br />

NHRI Science Report 7, National Hydrology Research Institute, Environment C<strong>an</strong>ada, 134 pp.<br />

Pomeroy, J. W., N. Parviainen, N. Hedstrom, <strong>an</strong>d D. M. Gray (1998), Coupled modeling of forest<br />

snow interception <strong>an</strong>d sublimation., Hydrological Processes, 12, 2317–2337.<br />

Pomeroy, J. W., <strong>an</strong>d R. A. Schmidt (1993), The use of fractal geometry in modeling intercepted<br />

snow accumulation <strong>an</strong>d sublimation., Proceedings of <strong>the</strong> 50th E<strong>as</strong>tern <strong>Snow</strong> Conference.,<br />

Quebec City, QC, C<strong>an</strong>ada., 50, 1–10.<br />

Schmidt, R. A. (1991), Sublimation of snow intercepted by <strong>an</strong> artificial conifer, Agricultural <strong>an</strong>d<br />

Forest Meteorology, 54, 1–27.<br />

Schmidt, R. A., R. L. Jairell, <strong>an</strong>d J. W. Pomeroy (1988), Me<strong>as</strong>uring snow interception <strong>an</strong>d loss<br />

from <strong>an</strong> artificial conifer., Proceedings of <strong>the</strong> 56th Western <strong>Snow</strong> Conference., Kalispell, MT.,<br />

166–169.<br />

Schmidt, R. A., <strong>an</strong>d C. A. Troendle (1992), Sublimation of intercepted snow <strong>as</strong> a global source of<br />

water vapor., Proceedings of <strong>the</strong> 60th Western <strong>Snow</strong> Conference., Jackson, WY., 60, 1–9.<br />

86


Schmidt, R. A., C. A. Troendle, <strong>an</strong>d J. R. Meim<strong>an</strong> (1998), Sublimation of snowpacks in subalpine<br />

conifer forests, C<strong>an</strong>adi<strong>an</strong> Journal Of Forest Research-Revue C<strong>an</strong>adienne De Recherche<br />

Forestiere, 28, 501–513.<br />

Schuepp, P. H., M. Y. Leclerc, J. I. Macpherson, <strong>an</strong>d R. L. Desjardins (1990), Footprint Prediction<br />

Of Scalar Fluxes From Analytical Solutions Of The Diffusion Equation, Boundary-Layer<br />

Meteorology, 50, 353–373.<br />

Sicart, J. E., J. W. Pomeroy, R. E. Essery, J. Hardy, T. Link, <strong>an</strong>d D. Marks (2004), A sensitivity<br />

study of daytime net radiation during snowmelt to forest c<strong>an</strong>opy <strong>an</strong>d atmospheric conditions,<br />

Journal of Hydrometeorology, 5, 774–784.<br />

Turnipseed, A. A., P. D. Bl<strong>an</strong>ken, D. E. Anderson, <strong>an</strong>d R. K. Monson (2002), Energy budget<br />

above a high-elevation subalpine forest in complex topography, Agricultural And Forest<br />

Meteorology, 110, 177–201.<br />

West, A. (1962), <strong>Snow</strong> evaporation from a forested watershed in <strong>the</strong> central Sierra Nevada,<br />

Journal of Forestry, 60, 481–484.<br />

Woo, M. K., <strong>an</strong>d M. A. Giesbrecht (2000), Simulation of snowmelt in a subarctic spruce<br />

woodl<strong>an</strong>d: Scale considerations, Nordic Hydrology, 31, 301–316.<br />

87


This page is intentionally bl<strong>an</strong>k.<br />

88


89<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

The Retrievals of <strong>Snow</strong> Cover Extent <strong>an</strong>d <strong>Snow</strong> Water Equivalent<br />

from a Blended P<strong>as</strong>sive Microwave–Interactive Multi-Sensor<br />

<strong>Snow</strong> Product<br />

ABSTRACT<br />

CEZAR KONGOLI 1 , CHARLES A. DEAN 1 ,<br />

SEAN R. HELFRICH 2 , AND RALPH R. FERRARO 3<br />

The retrieval of <strong>Snow</strong> Water Equivalent (SWE) from remote sensing satellites continues to be a<br />

very challenging problem. In this paper, we evaluate a new SWE product derived from <strong>the</strong><br />

blending of a p<strong>as</strong>sive microwave <strong>Snow</strong> Water Equivalent product b<strong>as</strong>ed on <strong>the</strong> Adv<strong>an</strong>ced<br />

Microwave Sounding Unit (AMSU) with <strong>the</strong> Interactive Multi-sensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping<br />

System (IMS). The microwave me<strong>as</strong>urements have <strong>the</strong> ability to penetrate <strong>the</strong> snow pack, <strong>an</strong>d thus<br />

<strong>the</strong> retrieval of SWE is best accomplished using <strong>the</strong> AMSU me<strong>as</strong>urements. On <strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d, <strong>the</strong><br />

IMS maps snow cover more reliably due to <strong>the</strong> use of multiple satellite <strong>an</strong>d ground observations.<br />

The evolution of global snow cover extent from <strong>the</strong> blended, <strong>the</strong> AMSU <strong>an</strong>d <strong>the</strong> IMS products w<strong>as</strong><br />

examined during <strong>the</strong> 2006 snow se<strong>as</strong>on. Despite <strong>the</strong> overall good inter-product agreement, it w<strong>as</strong><br />

shown that <strong>the</strong> retrievals of snow cover extent are improved using IMS, with implications for<br />

improved microwave retrievals of SWE. In a separate investigation, <strong>the</strong> microwave retrievals of<br />

SWE were examined globally <strong>an</strong>d in Central Europe. Qualitative evaluation of global SWE<br />

patterns showed dependence on l<strong>an</strong>d surface temperature: <strong>the</strong> lower <strong>the</strong> temperature, <strong>the</strong> higher <strong>the</strong><br />

SWE retrieved. This temperature bi<strong>as</strong> w<strong>as</strong> attributed in part to temperature effects on those snow<br />

properties that impact microwave response. Therefore, algorithm modifications are needed with<br />

more dynamical adjustments for ch<strong>an</strong>ging snow cover. Qu<strong>an</strong>titative evaluation over Slovakia for a<br />

limited period in 2006 showed re<strong>as</strong>onably good perform<strong>an</strong>ce for SWE less th<strong>an</strong> 100 mm.<br />

Sensitivity to deeper snow decre<strong>as</strong>ed signific<strong>an</strong>tly.<br />

Keywords: <strong>Snow</strong> cover, <strong>Snow</strong> Water Equivalent, Adv<strong>an</strong>ced Microwave Sounding Unit (AMSU),<br />

Interactive Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping Unit (IMS)<br />

INTRODUCTION<br />

The retrievals of <strong>Snow</strong> Water Equivalent (SWE) from satellites continue to be a very difficult<br />

<strong>an</strong>d challenging problem. While mapping of global snow cover h<strong>as</strong> been accomplished using<br />

visible or p<strong>as</strong>sive microwave me<strong>as</strong>urements, <strong>the</strong> mapping of SWE from space h<strong>as</strong> long been <strong>an</strong><br />

exclusive domain of p<strong>as</strong>sive microwave sensors. Visible me<strong>as</strong>urements are typically more<br />

sensitive to snow cover surfaces th<strong>an</strong> p<strong>as</strong>sive microwave me<strong>as</strong>urements due to <strong>the</strong> high visible<br />

1<br />

QSS Group Inc. – NOAA Science Center, 5200 Auth Rd., Camp Springs, MD, 20746,<br />

cezar.kongoli@noaa.gov<br />

2<br />

NOAA/NESDIS/OSDPD - NOAA Science Center, 5200 Auth Rd., Camp Springs, MD, 20746,<br />

se<strong>an</strong>.helfrich@noaa.gov<br />

3<br />

NOAA/NESDIS/OSDPD - NOAA Science Center, 5200 Auth Rd., Camp Springs, MD, 20746,<br />

se<strong>an</strong>.helfrich@noaa.gov


eflect<strong>an</strong>ce of snow cover <strong>as</strong> compared to o<strong>the</strong>r surface features, <strong>an</strong>d <strong>the</strong>refore, <strong>the</strong>y typically<br />

provide more accurate mapping of snow cover th<strong>an</strong> <strong>the</strong> microwave imagery. On <strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d,<br />

p<strong>as</strong>sive microwaves at specific window frequencies penetrate much deeper into <strong>the</strong> snow pack due<br />

to longer wavelengths, <strong>an</strong>d have shown to provide information on <strong>the</strong> snow cover properties<br />

including SWE. Ano<strong>the</strong>r adv<strong>an</strong>tage of p<strong>as</strong>sive microwaves is <strong>the</strong>ir ability to penetrate cloudy<br />

atmospheres at specific window frequencies <strong>an</strong>d thus to provide retrievals of l<strong>an</strong>d surface<br />

parameters in near all-wea<strong>the</strong>r conditions. Despite <strong>the</strong> near all-wea<strong>the</strong>r capability, <strong>the</strong> current<br />

microwave sensors provide me<strong>as</strong>urements <strong>an</strong>d thus retrievals at much coarser resolution (25 km)<br />

th<strong>an</strong> <strong>the</strong> visible satellite sensors (1 km). Ano<strong>the</strong>r disadv<strong>an</strong>tage of p<strong>as</strong>sive microwave imagery is its<br />

limited capability to penetrate wet snow cover.<br />

The need for continuous regional <strong>an</strong>d global snow cover mapping for climate, hydrological <strong>an</strong>d<br />

wea<strong>the</strong>r applications h<strong>as</strong> led, especially in recent years, to <strong>the</strong> development of snow cover<br />

mapping products b<strong>as</strong>ed on multi-sensor data sources. For example, <strong>the</strong> National Oce<strong>an</strong>ic <strong>an</strong>d<br />

Atmospheric Administration (NOAA), <strong>the</strong> National Environmental Satellite Data <strong>an</strong>d Information<br />

Systems (NESDIS) uses <strong>the</strong> Interactive Multi-sensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping System (IMS)<br />

operationally to provide Nor<strong>the</strong>rn Hemispheric (NH) snow <strong>an</strong>d ice mapping <strong>as</strong> input to<br />

environmental prediction models. The IMS system utilizes in <strong>an</strong> interactive f<strong>as</strong>hion, through <strong>the</strong><br />

interpretation of <strong>an</strong> <strong>an</strong>alyst, visible <strong>an</strong>d microwave satellite imagery <strong>as</strong> well <strong>as</strong> ground<br />

observations. Over <strong>the</strong> years, <strong>as</strong> more satellite sensors have become available, <strong>the</strong> IMS h<strong>as</strong><br />

incorporated more satellite data sources <strong>an</strong>d h<strong>as</strong> evolved into a more efficient snow mapping<br />

system (Ramsay, 1997, Helfrich et al., this issue). Ano<strong>the</strong>r multi-sensor snow <strong>an</strong>d ice mapping<br />

product is <strong>the</strong> NOAA’s AUTOSNOW product (Rom<strong>an</strong>ov et al., 2000). It utilizes visible, infrared<br />

<strong>an</strong>d microwave satellite data to automatically generate high-resolution <strong>an</strong>d continuous NH snow<br />

<strong>an</strong>d ice mapping.<br />

Main research objective of this study is to evaluate <strong>the</strong> accuracy of a new SWE product derived<br />

from <strong>the</strong> blending of a p<strong>as</strong>sive microwave SWE product b<strong>as</strong>ed on <strong>the</strong> Adv<strong>an</strong>ced Microwave<br />

Sounding Unit (AMSU) with <strong>the</strong> IMS snow cover extent product. The blended SWE product is<br />

needed (in addition to snow cover extent) <strong>as</strong> input to environmental prediction models. Besides<br />

being delivered in compatible format with <strong>the</strong> IMS snow cover extent, <strong>the</strong> blended product would<br />

be value-added in that SWE would be improved due to more reliable IMS snow mapping. Similar<br />

approaches have been reported in literature, e.g., <strong>the</strong> blending of a microwave SWE product b<strong>as</strong>ed<br />

on <strong>the</strong> Special Sensor Microwave Imager (SSM/I) or <strong>the</strong> Adv<strong>an</strong>ced Microwave Sc<strong>an</strong>ning<br />

Radiometer-Earth Observing System (AMSR-E) with <strong>the</strong> snow cover extent product derived from<br />

<strong>the</strong> Moderate Resolution Imaging Spectrometer (MODIS) (Armstrong et al. (2003). Compared to<br />

SSM/I <strong>an</strong>d AMSR-E, <strong>the</strong> AMSU h<strong>as</strong> some unique characteristics: Despite its coarse resolution, it<br />

h<strong>as</strong> a wider spatial coverage th<strong>an</strong> AMSR-E <strong>an</strong>d SSM/I, <strong>an</strong>d h<strong>as</strong> additional microwave ch<strong>an</strong>nels at<br />

frequencies in <strong>the</strong> window, oxygen <strong>an</strong>d water vapor absorption regions. These additional features<br />

could potentially be utilized for <strong>the</strong> development of more robust SWE retrievals.<br />

In this paper, we evaluate <strong>the</strong> blended SWE product for its accuracy in mapping snow cover <strong>an</strong>d<br />

SWE. In terms of snow cover mapping, we inter-compare <strong>the</strong> AMSU snow cover extent product<br />

with that of IMS. Inter-comparability of snow cover extent retrieved from <strong>the</strong> AMSU <strong>an</strong>d from<br />

IMS is import<strong>an</strong>t to investigate <strong>as</strong> it h<strong>as</strong> implications for <strong>the</strong> accuracy of <strong>the</strong> blended SWE<br />

product. Next, <strong>the</strong> microwave-derived SWE estimates are also evaluated globally <strong>an</strong>d regionally<br />

over Central Europe (Slovakia).<br />

90


METHODS<br />

The IMS Product<br />

The IMS w<strong>as</strong> designed to allow mapping of snow cover interactively on a daily b<strong>as</strong>is using a<br />

variety of data sources within a common geographic system. Such data sources include NOAA<br />

Polar Orbiting Environmental Satellites (POES) <strong>an</strong>d Geostationary Environmental Satellites<br />

(GOES) data, Jap<strong>an</strong>ese Geostationary Meteorological Satellites (GMS), Europe<strong>an</strong> geostationary<br />

meteorological satellites (METEOSAT) <strong>an</strong>d US Department of Defense (DOD) polar orbiters<br />

(DMSP). Since its inception in 1997, <strong>the</strong> IMS system h<strong>as</strong> evolved signific<strong>an</strong>tly in terms of input<br />

data sources, production technology <strong>an</strong>d output format. For more detailed description of evolution<br />

<strong>an</strong>d capabilities of <strong>the</strong> IMS system, <strong>the</strong> reader is referred to Helfrich et al., (this issue) <strong>an</strong>d Ramsay<br />

(1998; 2000). The IMS snow <strong>an</strong>d ice mapping is currently accomplished once daily over <strong>the</strong><br />

Nor<strong>the</strong>rn Hemisphere. The product is generated at 4 km <strong>an</strong>d 24 km resolution in Polar<br />

Stereographic (PS) Projection.<br />

The AMSU Instrument<br />

The AMSU instrument contains two modules: AMSU-A <strong>an</strong>d AMSU-B. The A module h<strong>as</strong> 15<br />

ch<strong>an</strong>nels in <strong>the</strong> 23-89 GHz frequency r<strong>an</strong>ge (1-15, Table 1). The B module h<strong>as</strong> five ch<strong>an</strong>nels in <strong>the</strong><br />

89-183 GHz frequency r<strong>an</strong>ge (16–20, Table 1). The AMSU-A h<strong>as</strong> <strong>an</strong> inst<strong>an</strong>t<strong>an</strong>eous field of view<br />

(FOV) of 48 km at nadir for all frequency ch<strong>an</strong>nels, <strong>an</strong>d sc<strong>an</strong>s ±48 0 from nadir with a total of 30<br />

me<strong>as</strong>urements across <strong>the</strong> sc<strong>an</strong>. For AMSU-B, 90 me<strong>as</strong>urements are made across <strong>the</strong> sc<strong>an</strong> with a<br />

nadir resolution of 16 km for all ch<strong>an</strong>nels. The nadir resolution degrades at larger sc<strong>an</strong> <strong>an</strong>gles. The<br />

swath width of both AMSU-A <strong>an</strong>d –B is 2343 km.<br />

The AMSU-A <strong>an</strong>d -B is flown on board <strong>the</strong> NOAA 15, NOAA 16, <strong>an</strong>d NOAA 17 POES. This<br />

three-satellite suite offers a near-real time global sampling nearly every 4 hours. The unique<br />

combination of ch<strong>an</strong>nels in <strong>the</strong> microwave window (23, 31, 89, 150 GHz), opaque water vapor<br />

(183±1, ±3, ±7 GHz) <strong>an</strong>d oxygen absorption (50-60 GHz) regions h<strong>as</strong> led to <strong>the</strong> development of a<br />

variety of surface <strong>an</strong>d atmospheric products, generated in near-real time within a system called <strong>the</strong><br />

Microwave Surface <strong>an</strong>d Precipitation Product System (MSPPS). The suite of operational MSPPS<br />

products (Ferraro et al., 2002) includes rain rate, total precipitable water, cloud liquid water, ice<br />

water path, snow cover, sea ice concentration <strong>an</strong>d l<strong>an</strong>d surface temperature <strong>an</strong>d emissivities at 23,<br />

31 <strong>an</strong>d 50 GHz. Recent additions include a snowfall detection (Ferraro et al., 2004), <strong>an</strong>d <strong>the</strong><br />

extension of <strong>the</strong> snow cover extent product to include snow water equivalent (SWE) retrievals<br />

(Kongoli <strong>an</strong>d Ferraro, 2004). In May 2005, a new POES satellite, NOAA 18, w<strong>as</strong> launched <strong>an</strong>d<br />

w<strong>as</strong> also incorporated into MSPPS. This new satellite contains a new instrument, <strong>the</strong> Microwave<br />

Humidity Sensor (MHS) which replaced <strong>the</strong> AMSU-B. Similar to AMSU-B, <strong>the</strong> MHS is a fivech<strong>an</strong>nel<br />

radiometer. However, it differs from AMSU-B in that ch<strong>an</strong>nel 2 <strong>an</strong>d ch<strong>an</strong>nel 5 (ch<strong>an</strong>nels<br />

17 <strong>an</strong>d 20 in Table 1 ) are 157 GHz <strong>an</strong>d 190 GHz. Calibration <strong>an</strong>d validation of <strong>the</strong> new MHS<br />

instrument within MSPPS h<strong>as</strong> been completed <strong>an</strong>d is described in Meng et al. (2006). As a result,<br />

similar products that are generated from <strong>the</strong> three satellites are now also generated from NOAA-<br />

18.<br />

91


Table 1. AMSU-A <strong>an</strong>d –B ch<strong>an</strong>nel characteristics. Ch<strong>an</strong>nels 1-15 are AMSU-A ch<strong>an</strong>nels <strong>an</strong>d ch<strong>an</strong>nels<br />

16-20 are AMSU-B ch<strong>an</strong>nels<br />

Ch<strong>an</strong>nel<br />

number<br />

Center<br />

frequency<br />

(GHz)<br />

Number<br />

of p<strong>as</strong>s<br />

b<strong>an</strong>ds<br />

92<br />

B<strong>an</strong>d<br />

width<br />

(MHz)<br />

Center<br />

frequency<br />

stability<br />

(MHz)<br />

FOV<br />

at<br />

Nadir<br />

(km)<br />

1 23.80 1 251 10 48<br />

2 31.40 1 161 10 48<br />

3 50.30 1 161 10 48<br />

4 52.80 1 380 5 48<br />

5 53.59±0.115 2 168 5 48<br />

6 54.40 1 380 5 48<br />

7 54.94 1 380 10 48<br />

8 55.50 1 310 0.5 48<br />

9 57.29 = fo 1 310 0.5 48<br />

10 fo±0.217 2 76 0.5 48<br />

11 fo±0.322±0.048 4 34 0.5 48<br />

12 fo±0.322±0.022 4 15 0.5 48<br />

13 fo±0.322±0.010 4 8 0.5 48<br />

14 fo±0.322±0.004 4 3 0.5 48<br />

15 89.00 1 2000 50 48<br />

16 89.00 1 5000 50 16<br />

17 150 1 4000 50 16<br />

18 183±1 1 1000 50 16<br />

19 183±3 2 2000 50 16<br />

20 183±7 2 4000 50 16<br />

The AMSU <strong>Snow</strong> Cover Extent Product<br />

The identification of snow cover over l<strong>an</strong>d is b<strong>as</strong>ed on <strong>the</strong> algorithm of Grody (1991) <strong>an</strong>d Grody<br />

<strong>an</strong>d B<strong>as</strong>ist (1996). <strong>Snow</strong> is identified in a series of steps that discriminate snow from nonscattering<br />

surfaces such <strong>as</strong> wet l<strong>an</strong>d <strong>an</strong>d vegetation (Figure 1) <strong>an</strong>d from o<strong>the</strong>r scatterers such <strong>as</strong><br />

deserts <strong>an</strong>d rain. This is accomplished using a number of scattering indices that utilize a<br />

combination of <strong>the</strong> AMSU window frequency ch<strong>an</strong>nels at 23, 31, 50 <strong>an</strong>d 89 GHz (Table 1,<br />

AMSU-A ch<strong>an</strong>nels 1, 2, 3 <strong>an</strong>d AMSU-B ch<strong>an</strong>nel 16, respectively). As shown in Figure 1, snow<br />

exhibits a unique spectral signature in <strong>the</strong> 10–100 GHz microwave frequency region: The<br />

brightness temperature, <strong>an</strong>d hence, <strong>the</strong> surface emissivity decre<strong>as</strong>es with incre<strong>as</strong>ing frequency. In<br />

contr<strong>as</strong>t, o<strong>the</strong>r surfaces such <strong>as</strong> wet l<strong>an</strong>d <strong>an</strong>d vegetation exhibit a ra<strong>the</strong>r flat or reverse response.<br />

More recently, additional filters have been incorporated that utilize a combination of AMSU<br />

ch<strong>an</strong>nels at 150 GHz (Table 1, ch<strong>an</strong>nel 17) at 53.6 GHz (Table 1, ch<strong>an</strong>nel 5) <strong>an</strong>d at 183 ± 3 GHz<br />

(Table 1, ch<strong>an</strong>nel 19) for improved snow–rain discrimination (Kongoli et al. 2005).


Figure 1. Microwave spectral characteristics of vegetation <strong>an</strong>d snow cover (Matzler, 1994).<br />

The AMSU SWE Product<br />

The retrieval of SWE is b<strong>as</strong>ed on <strong>the</strong> AMSU brightness temperature me<strong>as</strong>urements at 23<br />

(TB23), 31 (TB31) <strong>an</strong>d 89 (TB89) GHz. B<strong>as</strong>ed on <strong>the</strong>se ch<strong>an</strong>nel me<strong>as</strong>urements, two scattering<br />

indices are computed:<br />

SI31 = TB23 – TB89 (1)<br />

SI89 = TB23 – TB31 (2)<br />

The SWE is computed for <strong>the</strong> snow-covered pixels by <strong>the</strong> following empirical relationships:<br />

SWE = K1 + K2* SI89 (3)<br />

SWE = K3 + K4 * SI31 (4)<br />

where K1, K2, K3, <strong>an</strong>d K4 are empirically derived coefficients. SWE is computed only over<br />

snow-covered l<strong>an</strong>d retrieved by <strong>the</strong> AMSU snow cover extent algorithm. The SWE algorithm<br />

implicitly differentiates between two snow-cover types: finer-grained fresh snow (Eq. 3) <strong>an</strong>d<br />

coarser-grained older (Eq. 4) snow-cover via <strong>the</strong> estimation of a switch (me<strong>an</strong> grain size)<br />

parameter, which is also estimated from a linear relationship with TB23, TB31 <strong>an</strong>d TB89. The<br />

coefficient values for K1, K2, K3 <strong>an</strong>d K4 <strong>an</strong>d <strong>the</strong> switch parameter are fixed. They were derived<br />

from regional studies in <strong>the</strong> U.S Great Plains are<strong>as</strong> (Kongoli et al., 2004). It is also import<strong>an</strong>t to<br />

note that atmospheric or l<strong>an</strong>d surface vegetation effects are not accounted for in <strong>the</strong> current<br />

version. The concept behind <strong>the</strong> algorithm using <strong>the</strong> scattering indices (1) <strong>an</strong>d (2) <strong>an</strong>d Equations<br />

(3) <strong>an</strong>d (4), respectively, is illustrated in Figure 2, which depicts examples of AMSU<br />

me<strong>as</strong>urements at 23, 31, 89 <strong>an</strong>d 150 GHz window frequency ch<strong>an</strong>nels. Note that microwave<br />

sensitivity to snow cover shifts from lower (20–30 GHz) to higher (89–150 GHz) frequency<br />

regions <strong>as</strong> <strong>the</strong> snow ages. For freshly fallen snow (6 hr old), <strong>the</strong> steepest microwave gradient<br />

occurs in <strong>the</strong> 89–150 GHz region (about 25 K), due to strong scattering at 150 GHz by finergrained<br />

snow cover. Note <strong>the</strong> flat response in <strong>the</strong> 20–30 GHz region, suggesting low microwave<br />

sensitivity in this lower frequency r<strong>an</strong>ge. As <strong>the</strong> snow becomes older <strong>an</strong>d <strong>the</strong> grain size incre<strong>as</strong>es,<br />

sensitivity shifts to <strong>the</strong> 30–90 GHz region which attains <strong>the</strong> steepest gradient. Only for coarsegrained,<br />

metamorphosed snow cover, referred to in Figure 2 <strong>as</strong> “old snow” does <strong>the</strong> microwave<br />

sensitivity incre<strong>as</strong>e signific<strong>an</strong>tly at 23 <strong>an</strong>d 31 GHz. In <strong>the</strong> current algorithm, 150 GHz is not<br />

utilized for <strong>the</strong> retrievals of SWE. It is import<strong>an</strong>t to note, however, that this dual snow type<br />

discrimination is a crude approximation <strong>an</strong>d representation of <strong>the</strong> much greater snow cover type<br />

<strong>an</strong>d hence grain size <strong>an</strong>d (micro)structure variability, e.g., <strong>as</strong> illustrated in Figure 1.<br />

93


Figure 2. Example of AMSU me<strong>as</strong>urements over snow covered l<strong>an</strong>d<br />

Blending Procedure<br />

Primary satellite data are not used in <strong>the</strong> blended product. Instead, <strong>the</strong> data inputs include<br />

AMSU SWE product files <strong>an</strong>d IMS snow cover product files. The SWE data are obtained from <strong>the</strong><br />

MSPPS Level 2 swath geographical products available in <strong>the</strong> Hierarchical Data Format-Earth<br />

Observing System (HDF-EOS) at NOAA/NESDIS. These swath files are generated on <strong>an</strong> orbital<br />

b<strong>as</strong>is from NOAA-15, 16, 17 <strong>an</strong>d more recently <strong>the</strong> NOAA-18 satellites. Spatial resolution of <strong>the</strong><br />

AMSU SWE swath data is that of AMSU-B: 16 km at nadir viewing <strong>an</strong>gle. Resolution degrades <strong>as</strong><br />

viewing <strong>an</strong>gle incre<strong>as</strong>es. The IMS snow cover data are obtained daily <strong>as</strong> ASCII files in 1/16 th<br />

mesh PS Projection, or approximately 24 km in resolution. The data flow <strong>an</strong>d rules for generating<br />

merged product are <strong>as</strong> follows:<br />

- Obtain daily IMS snow cover product file,<br />

- Obtain <strong>the</strong> AMSU SWE swath product files. The NOAA-18 instrument w<strong>as</strong> selected <strong>as</strong><br />

<strong>the</strong> most recent instrument, <strong>an</strong>d <strong>the</strong> descending (night-time) w<strong>as</strong> selected to minimize<br />

SWE retrieval errors due to surface melting during day-time,<br />

- Convert <strong>the</strong> AMSU swath SWE data into gridded data compatible with <strong>the</strong> IMS snow<br />

cover product file.<br />

- Inter-compare <strong>the</strong> AMSU SWE <strong>an</strong>d IMS snow cover values on a grid cell-by-cell b<strong>as</strong>is.<br />

If SWE value is missing or zero over snow cover <strong>as</strong> identified by IMS, retrieve previous<br />

day SWE value (two-day compositing). If previous-day SWE value is positive <strong>as</strong>sign that<br />

SWE value to grid cell. O<strong>the</strong>rwise, flag <strong>as</strong> “indeterminate” <strong>the</strong> grid cell that corresponds<br />

to missing or zero AMSU SWE <strong>an</strong>d IMS snow-covered l<strong>an</strong>d. Also, label <strong>as</strong><br />

“indeterminate” inst<strong>an</strong>ces of positive AMSU SWE value that correspond to IMS snowfree<br />

l<strong>an</strong>d. At this point, no revision of SWE values over <strong>the</strong>se “indeterminate” grid cells<br />

is made. This is reserved for future development,<br />

- Generate blended SWE output file, a statistics file with grid cell confusion data, <strong>an</strong>d<br />

image input <strong>an</strong>d blended output grid files for product visualization <strong>an</strong>d monitoring<br />

Ground SWE me<strong>as</strong>urements over Slovakia<br />

The SWE retrievals were qu<strong>an</strong>titatively evaluated against SWE me<strong>as</strong>urements over Slovakia<br />

during February–March 2006. The SWE data <strong>an</strong>d maps were provided by <strong>the</strong> Slovak National<br />

Hydro-meteorological Institute which maintains a dense network of ground stations that record<br />

94


snow depth <strong>an</strong>d SWE on a weekly b<strong>as</strong>is (Figure 3). Dotted green points denote station locations,<br />

where<strong>as</strong> in red are depicted grid points spaced 0.25 by 0.25 degrees Latitude <strong>an</strong>d Longitude. The<br />

are<strong>as</strong> depicted in brown in Central <strong>an</strong>d North Slovakia denotes higher elevation terrain. To<br />

statistically evaluate SWE retrievals, station SWE data were matched up with <strong>the</strong> AMSU<br />

observations <strong>an</strong>d retrieved SWE. Station SWE data were averaged over a cell of 0.5 X 0.5 degree<br />

in latitude <strong>an</strong>d longitude centered at <strong>the</strong> centroid of AMSU FOV. This size represents <strong>the</strong> average<br />

resolution of <strong>the</strong> AMSU instrument. Along with <strong>the</strong> average SWE, average values of snow depth<br />

<strong>an</strong>d elevation, <strong>the</strong> number of stations per cell, <strong>as</strong> well <strong>as</strong> st<strong>an</strong>dard deviations of SWE, depth <strong>an</strong>d<br />

elevation were computed. Only cells with over 5 stations were included in <strong>the</strong> <strong>an</strong>alysis. Due to <strong>the</strong><br />

higher density of stations, cells with over five stations were predomin<strong>an</strong>t.<br />

RESULTS AND DISCUSSION<br />

Figure 3. Map of Slovakia <strong>an</strong>d locations of ground SWE stations<br />

Inter-comparison of blended snow cover with that retrieved from <strong>the</strong> AMSU <strong>an</strong>d IMS<br />

products<br />

The blended product system h<strong>as</strong> been automated <strong>an</strong>d running experimentally on a UNIX<br />

machine since November 2005. Examples of blended SWE product are shown in Figure 4. Figure<br />

4 displays merged SWE (left) <strong>an</strong>d inter-comparison of AMSU <strong>an</strong>d IMS snow cover extent (right)<br />

on November 7, 2005 (top p<strong>an</strong>els) <strong>an</strong>d on February 15, 2006 (bottom p<strong>an</strong>els). The intercomparison<br />

map depicts <strong>the</strong> blended SWE area in blue, <strong>the</strong> underestimated microwave SWE area<br />

in green (AMSU-SWE values are zero but IMS snow cover is positive), <strong>an</strong>d <strong>the</strong> overestimated<br />

microwave SWE area in brown (AMSU-SWE is positive but IMS snow cover is zero). The red<br />

color depicts are<strong>as</strong> when AMSU-SWE is labeled <strong>as</strong> “undetermined” due to confusion with rain,<br />

<strong>an</strong>d is displayed for diagnostic purposes. Note <strong>the</strong> green-coded area over E<strong>as</strong>tern C<strong>an</strong>ada on<br />

November 7, 2005 (top right), denoting underestimation of snow cover <strong>an</strong>d hence SWE by <strong>the</strong><br />

AMSU snow cover extent product. In <strong>the</strong> blended SWE product, <strong>the</strong>se undetermined SWE are<strong>as</strong><br />

are denoted <strong>as</strong> “white”. Examination of ground observations from several C<strong>an</strong>adi<strong>an</strong><br />

meteorological stations indicated continuous snowfall occurrences prior to November 7. The<br />

deposited new snow w<strong>as</strong> not captured by <strong>the</strong> AMSU snow cover extent product, but w<strong>as</strong> identified<br />

by <strong>the</strong> IMS snow cover product. Interestingly, <strong>the</strong>se snowfall occurrences were retrieved by <strong>the</strong><br />

AMSU snowfall product. An example of <strong>the</strong> AMSU snowfall product retrievals on November 2,<br />

2005 is given in Figure 5. The AMSU snowfall algorithm utilizes frequencies at 89 GHz <strong>an</strong>d<br />

above (AMSU-B ch<strong>an</strong>nels 16-20 in Table 1) to identify precipitation in <strong>the</strong> form of snowfall<br />

95


(Kongoli et al., 2003). Examination of <strong>the</strong> AMSU me<strong>as</strong>urements on November 7, 2005 indicated<br />

that <strong>the</strong> 89 GHz frequency ch<strong>an</strong>nel w<strong>as</strong> not sufficiently sensitive to this newly deposited snow.<br />

However, <strong>the</strong> 150 GHz window frequency ch<strong>an</strong>nel exhibited scattering response, e.g., depressed<br />

brightness temperatures relative to that of 89 GHz (see also Figure 2). Note that frequency<br />

ch<strong>an</strong>nels above 89 GHz are not available for <strong>the</strong> SSM/I or AMSR-E sensors. This unique<br />

capability of <strong>the</strong> AMSU instrument to retrieve snowfall <strong>an</strong>d to detect newly deposited snow cover<br />

could be potentially incorporated <strong>an</strong>d utilized for improved SWE retrievals for new snow. On <strong>the</strong><br />

o<strong>the</strong>r h<strong>an</strong>d, in <strong>the</strong> blended SWE product, SWE over part of Mongolia <strong>as</strong> determined by <strong>the</strong> AMSU<br />

product but labeled <strong>as</strong> “snow-free” l<strong>an</strong>d by <strong>the</strong> IMS is correctly left out. This overestimation of<br />

snow cover over Mongolia by <strong>the</strong> AMSU persisted during much of <strong>the</strong> winter se<strong>as</strong>on of 2006, <strong>an</strong>d<br />

is also shown on February 15, 2006 (bottom p<strong>an</strong>els). It is shown that overall, <strong>the</strong>re is better<br />

agreement between <strong>the</strong> microwave-derived SWE <strong>an</strong>d <strong>the</strong> IMS snow cover, e.g., over E<strong>as</strong>tern<br />

C<strong>an</strong>ada, th<strong>an</strong> in early winter. Exception is are<strong>as</strong> over Mongolia where overestimation of <strong>the</strong><br />

AMSU-SWE persists throughout <strong>the</strong> winter. Examination of MODIS snow cover maps indicated<br />

that indeed <strong>the</strong>se are<strong>as</strong> did not have snow cover, <strong>an</strong>d <strong>the</strong>refore, <strong>the</strong>y represent “false” snow are<strong>as</strong><br />

by <strong>the</strong> AMSU SWE product. Examination of <strong>the</strong> AMSU me<strong>as</strong>urements over Mongolia indicated<br />

relatively small values of <strong>the</strong> scattering index at 89 GHz (SI89), which is used in <strong>the</strong> AMSU snow<br />

cover extent product to identify snow cover <strong>an</strong>d in <strong>the</strong> AMSU-SWE product to compute fresh<br />

snow SWE (Eq. 2 <strong>an</strong>d 4). This low scattering signal at 89 GHz w<strong>as</strong> present during night-time (low<br />

temperatures) <strong>an</strong>d day-time (above freezing temperatures). The presence of scattering in a wide<br />

r<strong>an</strong>ge of atmospheric conditions over Mongolia would be indicative of ground ra<strong>the</strong>r th<strong>an</strong><br />

atmospheric effects, e.g., soil grain scattering (B<strong>as</strong>ist et al., 1996). Figure 6 is a plot of <strong>the</strong><br />

percentage of <strong>the</strong> retrieved AMSU-SWE area relative to <strong>the</strong> IMS snow coverage from December<br />

1, 2005 through April 15, 2006 (left-h<strong>an</strong>d vertical axis) <strong>an</strong>d of <strong>the</strong> overestimated AMSU-SWE<br />

area <strong>as</strong> a percentage of <strong>the</strong> IMS snow coverage (right-h<strong>an</strong>d vertical axis). It is shown that <strong>the</strong> net<br />

SWE area mapped by <strong>the</strong> AMSU is at about 75% <strong>as</strong> compared to that of IMS, which is relatively<br />

high for a microwave instrument. The overestimation of snow cover by <strong>the</strong> AMSU is about 6 %<br />

relative to IMS snow coverage.<br />

96


Figure 4. Example of blended SWE product retrievals over Nor<strong>the</strong>rn Hemisphere on November 7, 2005 <strong>an</strong>d<br />

February 15, 2006. Left-h<strong>an</strong>d images depict <strong>the</strong> blended SWE <strong>an</strong>d <strong>the</strong> right-h<strong>an</strong>d images depict blended <strong>an</strong>d<br />

unblended are<strong>as</strong>.<br />

97


Figure 5. The AMSU snowfall retrievals over North Americ<strong>an</strong> on November 2, 2005.<br />

Figure 6. Inter-comparison of global SWE <strong>an</strong>d IMS snow cover extent<br />

98


Evaluation of <strong>the</strong> global distribution of retrieved SWE<br />

Figure 7 depicts example of retrievals of AMSU SWE (left) <strong>an</strong>d L<strong>an</strong>d Surface Temperature<br />

(LST) (right) in March 2002 <strong>an</strong>d 2003. Note <strong>the</strong> predomin<strong>an</strong>ce of high SWE values (depicted in<br />

red) over North-central <strong>an</strong>d E<strong>as</strong>tern Siberia, over Nor<strong>the</strong>rn C<strong>an</strong>ada <strong>an</strong>d Al<strong>as</strong>ka. Interestingly, <strong>the</strong>se<br />

patterns are similar to <strong>the</strong> ones retrieved in 2006 (Figure 4). Although global SWE data are not<br />

available to qu<strong>an</strong>titatively evaluate retrieved SWE, some general observations c<strong>an</strong> be made. For<br />

example, <strong>the</strong> incre<strong>as</strong>e in SWE from Western to E<strong>as</strong>tern Siberia may be unrealistic given <strong>the</strong> dry,<br />

low-temperature climate predomin<strong>an</strong>t in E<strong>as</strong>tern Siberia. These trends appear to be strikingly<br />

similar in 2002, 2003 <strong>an</strong>d 2006. Despite large-scale similarity, a closer visual inspection revealed<br />

some inter-<strong>an</strong>nual differences over specific regions. For inst<strong>an</strong>ce, retrieved SWE over some<br />

Western US regions, e.g., Colorado, Wyoming <strong>an</strong>d Idaho, <strong>an</strong>d over C<strong>an</strong>ada, e.g, Alberta, were<br />

higher in 2002 th<strong>an</strong> in 2003. Ano<strong>the</strong>r observed difference w<strong>as</strong> <strong>the</strong> larger extent of high SWE over<br />

Siberia in 2003 th<strong>an</strong> in 2002. This inter-<strong>an</strong>nual variability in SWE appears to follow that of LST:<br />

The lower <strong>the</strong> LST <strong>the</strong> higher <strong>the</strong> retrieved SWE. For inst<strong>an</strong>ce, lower LST in 2002 th<strong>an</strong> in 2003<br />

over Western US w<strong>as</strong> <strong>as</strong>sociated with retrieved SWE higher in 2002 th<strong>an</strong> in 2003. LST also<br />

appears to influence to a large extent <strong>the</strong> global distribution of retrieved SWE. A possible<br />

expl<strong>an</strong>ation of this temperature bi<strong>as</strong> could be <strong>the</strong> influence of temperature on <strong>the</strong> evolution of<br />

snow cover properties that impact <strong>the</strong> microwave response. As explained earlier, <strong>the</strong> AMSU SWE<br />

algorithm coefficients are static. Foster et al., 2005 report a new dynamical approach where<br />

algorithm coefficients are adjusted b<strong>as</strong>ed on a geo-referenced snow cl<strong>as</strong>sification system (Sturm et<br />

al., 1995). The authors report improved perform<strong>an</strong>ce compared to static retrievals.<br />

Figure 7. AMSU retrieval of SWE (top) <strong>an</strong>d L<strong>an</strong>d Surface Temperature (bottom) in March 2002 <strong>an</strong>d<br />

2003.<br />

99


Evaluation of SWE retrievals over Slovakia<br />

Figure 8 plots <strong>the</strong> histogram of station averaged SWE over Slovakia (top p<strong>an</strong>el) <strong>an</strong>d <strong>the</strong> average<br />

<strong>an</strong>d st<strong>an</strong>dard deviation of SWE matched up with <strong>the</strong> AMSU me<strong>as</strong>urements <strong>as</strong> a function of<br />

average elevation (bottom p<strong>an</strong>el). As shown, SWE is most frequent around 50 mm. There is,<br />

however, a good distribution of SWE in <strong>the</strong> 50- to 150-mm r<strong>an</strong>ge. Note also that <strong>the</strong> st<strong>an</strong>dard<br />

deviation of SWE incre<strong>as</strong>es with elevation <strong>an</strong>d <strong>the</strong> SWE amount. Figure 9 displays plots of<br />

retrieved SWE versus me<strong>as</strong>ured SWE for elevations less th<strong>an</strong> 400 m (top p<strong>an</strong>el left) <strong>an</strong>d for all<br />

elevation r<strong>an</strong>ges (bottom p<strong>an</strong>el left). As shown, <strong>the</strong>re is good agreement of retrieved SWE versus<br />

me<strong>as</strong>ured SWE for elevations less th<strong>an</strong> 400 m. Elevations less th<strong>an</strong> 400 mm had <strong>an</strong> average SWE<br />

of 40 mm <strong>an</strong>d maximum SWE <strong>an</strong>d snow depth of 100 mm <strong>an</strong>d 30 cm, respectively. For higher<br />

elevations, retrieved SWE is signific<strong>an</strong>tly underestimated. The right h<strong>an</strong>d p<strong>an</strong>els in Figure 9 depict<br />

plots of <strong>the</strong> bi<strong>as</strong> (retrieved SWE–me<strong>as</strong>ured SWE) <strong>as</strong> a function of elevation (top) <strong>an</strong>d <strong>as</strong> a function<br />

of me<strong>as</strong>ured SWE (bottom). As shown, <strong>the</strong> bi<strong>as</strong> exhibits stronger dependence on SWE amounts<br />

th<strong>an</strong> on elevation. Microwave signal saturation at SWE values above 100 mm is also reported in<br />

literature (Dong et al., 2005). However, given <strong>the</strong> high st<strong>an</strong>dard deviation of me<strong>as</strong>ured SWE<br />

within <strong>the</strong> AMSU FOV, a more rigorous examination of <strong>the</strong> possible error sources would have<br />

required additional information on snow cover <strong>an</strong>d ground station distribution. Figure 10 displays<br />

AMSU SWE retrievals over Central Europe including Slovakia (left p<strong>an</strong>el) <strong>an</strong>d <strong>the</strong> map of station<br />

derived SWE over Slovakia (right p<strong>an</strong>el) on February 27, 2006. As shown, <strong>the</strong> highest AMSU<br />

SWE retrieved over Slovakia is in <strong>the</strong> 100- to 120-mm r<strong>an</strong>ge in Nor<strong>the</strong><strong>as</strong>tern Slovakia, compared<br />

to station derived SWE over 180 mm. Note, however, <strong>the</strong> highly variable station-derived SWE<br />

over Nor<strong>the</strong><strong>as</strong>tern Slovakia. Large-scale SWE patterns are retrieved re<strong>as</strong>onably well, e.g., <strong>the</strong><br />

incre<strong>as</strong>e in SWE towards <strong>the</strong> Nor<strong>the</strong><strong>as</strong>t. Small-scale patterns, however, are not well captured due<br />

to <strong>the</strong> coarse AMSU resolution.<br />

Figure 8. Average me<strong>as</strong>ured SWE distribution matched up with AMSU data<br />

100


Figure 9. Inter-comparison plots of AMSU versus me<strong>as</strong>ured SWE (left p<strong>an</strong>els) <strong>an</strong>d of <strong>the</strong> bi<strong>as</strong> of retrieved<br />

SWE <strong>as</strong> a function of elevation <strong>an</strong>d me<strong>as</strong>ured SWE (right p<strong>an</strong>els)<br />

Figure 10. AMSU retrieved SWE over central Europe (left) <strong>an</strong>d station SWE over Slovakia on February 27,<br />

2006.<br />

101


SUMMARY<br />

Main research objective of this study w<strong>as</strong> to evaluate <strong>the</strong> accuracy of a new SWE product<br />

derived from <strong>the</strong> blending of a p<strong>as</strong>sive microwave SWE product b<strong>as</strong>ed on AMSU instrument with<br />

<strong>the</strong> IMS snow cover extent product. The blended SWE product is needed (in addition to snow<br />

cover extent) <strong>as</strong> input to environmental prediction models. The paper described data <strong>an</strong>d<br />

evaluation methodology. The blending procedure between <strong>the</strong> AMSU SWE <strong>an</strong>d IMS snow cover<br />

extent products w<strong>as</strong> also described. Next, snow cover extent retrievals by <strong>the</strong> AMSU, IMS <strong>an</strong>d <strong>the</strong><br />

blended product were evaluated for <strong>the</strong> 2005–2006 snow se<strong>as</strong>on. Despite good inter-product<br />

agreement, it w<strong>as</strong> shown that <strong>the</strong> IMS had better skills th<strong>an</strong> <strong>the</strong> AMSU snow cover extent product<br />

in retrieving new snow cover in early winter <strong>an</strong>d in correctly identifying snow-free l<strong>an</strong>d, e.g., over<br />

Mongolia, that w<strong>as</strong> o<strong>the</strong>rwise reported <strong>as</strong> “snow” by <strong>the</strong> AMSU product. These problem are<strong>as</strong><br />

were also evaluated using o<strong>the</strong>r observation sources, e.g., <strong>the</strong> MODIS snow cover product <strong>an</strong>d <strong>the</strong><br />

AMSU snowfall product. In a separate investigation, <strong>the</strong> paper also described <strong>the</strong> evaluation of <strong>the</strong><br />

microwave SWE product globally <strong>an</strong>d over central Europe (Slovakia). Qualitative evaluation of<br />

<strong>the</strong> large-scale, global SWE patterns showed dependence of AMSU retrieved SWE on l<strong>an</strong>d surface<br />

temperature: <strong>the</strong> lower <strong>the</strong> l<strong>an</strong>d surface temperature, <strong>the</strong> higher <strong>the</strong> SWE retrieved. This<br />

temperature bi<strong>as</strong> w<strong>as</strong> attributed in part to temperature effects on snow properties that impact<br />

microwave response. Therefore, AMSU SWE algorithm modifications are needed, e.g.,<br />

adjustment of algorithm coefficients for ch<strong>an</strong>ging snow properties. Qu<strong>an</strong>titative evaluation over<br />

Slovakia for a limited period in 2006 showed re<strong>as</strong>onably good agreement for SWE less th<strong>an</strong> 100<br />

mm <strong>an</strong>d snow depth less th<strong>an</strong> 30 cm, <strong>as</strong>sociated with low elevation terrain (less th<strong>an</strong> 400 m).<br />

Sensitivity to snow deeper th<strong>an</strong> 100 mm in SWE decre<strong>as</strong>ed signific<strong>an</strong>tly.<br />

ACKNOWLEDGEMENT<br />

The authors wish to th<strong>an</strong>k Dr. Y<strong>an</strong> K<strong>an</strong>ak <strong>an</strong>d Joseph Pecho from <strong>the</strong> Slovak National Institute<br />

of Hydrometeorology for <strong>the</strong>ir cooperation <strong>an</strong>d for graciously providing valuable station data over<br />

Slovakia.<br />

REFERENCES<br />

Armstrong, R., M.J. Brodzik, <strong>an</strong>d M. Savoie, 2003: Multisensor approach to mapping snow cover<br />

using data from NASA’s Terra <strong>an</strong>d Aqua Spacecraft (AMSR-E <strong>an</strong>d MODIS), Eos Tr<strong>an</strong>s.<br />

AGU, 84(46).<br />

B<strong>as</strong>ist, A., D. Garrett, R. Ferraro, N. Grody <strong>an</strong>d K. Mitchell, 1996: A comparison between snow<br />

cover products derived from visible <strong>an</strong>d microwave observations, J. Appl. Meteorol., 35:163–<br />

177.<br />

Dong, J., J.P. Walker, <strong>an</strong>d P. R. Houser, 2005: Factors affecting remotely sensed snow water<br />

equivalent uncertainty, Remote Sensing of Environment 97: 68–82.<br />

Ferraro, R., C. Kongoli, <strong>an</strong>d P. Pellegrino, 2004: An Evaluation of a New AMSU-Derived Falling<br />

<strong>Snow</strong> Retrieval Algorithm (Preprint), 13 th Conference on Satellite Meteorology <strong>an</strong>d<br />

Oce<strong>an</strong>ography, Americ<strong>an</strong> Meteorological Society, 20–23 September, 2004, Norfolk, VA.<br />

Ferraro, R. R., F. Weng, N.C. Grody, I. Guch, C. De<strong>an</strong>, C. Kongoli, H. Meng, P. Pellegrino, <strong>an</strong>d<br />

L. Zhao, 2002: NOAA satellite-derived hydrological products prove <strong>the</strong>ir worth, Eos Tr<strong>an</strong>s.<br />

AGU, 83: 429–437.<br />

Foster J.L., C. Sun, J.P. Walker, R. Kelly, A. Ch<strong>an</strong>g, J. Dong, <strong>an</strong>d H. Powell, 2005: Qu<strong>an</strong>tifying<br />

<strong>the</strong> uncertainty in p<strong>as</strong>sive microwave snow water equivalent observations, Remote Sensing of<br />

Environment 94: 187–203.<br />

Grody, N.C., 1991: Cl<strong>as</strong>sification of snow cover <strong>an</strong>d precipitation using <strong>the</strong> special sensor<br />

microwave imager, J. Geophys. Res., 96(D4): 7423–7435.<br />

Grody, N.C., <strong>an</strong>d A.N. B<strong>as</strong>ist, 1996: Global identification of snow cover using SSM/I<br />

Me<strong>as</strong>urements, IEEE Tr<strong>an</strong>s. Geosci. Remote Sens. 34(1): 237–249.<br />

102


Helfrich, S.R., D. McNamara, B.H.Ramsay, T. Baldwin, <strong>an</strong>d T. K<strong>as</strong>heta (2006), Enh<strong>an</strong>cements<br />

<strong>an</strong>d forthcoming developments to <strong>the</strong> Interactive Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping System<br />

(IMS), Hydrological Processes (this issue).<br />

Kongoli, C., N.C. Grody, <strong>an</strong>d R. Ferraro, 2004: Interpretation of AMSU me<strong>as</strong>urementsfor <strong>the</strong><br />

retrievals of snow water equivalent <strong>an</strong>d snow depth, J. Geophys. Res, 109:<br />

do:10.1029/2004JD004836.<br />

Kongoli, C., <strong>an</strong>d R. Ferraro, 2004: Development <strong>an</strong>d evaluation of <strong>the</strong> AMSU-b<strong>as</strong>ed SWE product<br />

(Preprint), 13 th Conference on Satellite Meteorology <strong>an</strong>d Oce<strong>an</strong>ography, Americ<strong>an</strong><br />

Meteorological Society, 20–23 September, 2004, Norfolk, VA.<br />

Kongoli, C., R. Ferraro, P. Pellegrino, <strong>an</strong>d H. Meng, 2005 (Preprint): <strong>Snow</strong> Microwave Products<br />

from <strong>the</strong> NOAA’s Adv<strong>an</strong>ced Microwave Sounding Unit, 19 th Conference on Hydrology,<br />

Americ<strong>an</strong> Meteorological Society, 8–14 J<strong>an</strong>uary, 2005, S<strong>an</strong> Diego, CA.<br />

Kongoli, C., P. Pellegrino, R.R. Ferraro, C. Grody, <strong>an</strong>d H. Meng, 2003: A new snowfall detection<br />

algorithm over l<strong>an</strong>d using me<strong>as</strong>urements from <strong>the</strong> Adv<strong>an</strong>ced Microwave Sounding Unit<br />

(AMSU), Geophys. Res. Let. 30(14): 1756–1759.<br />

Matzler, C., 1994: P<strong>as</strong>sive microwave signatures of l<strong>an</strong>dscapes in winter, Meteorol. Atmos. Phys.<br />

54: 241–260.<br />

Meng H., L. Zhao, R. Ferraro, F. Weng, Q. Liu, 2006 (Preprint): Calibration <strong>an</strong>dvalidation of N-<br />

18 AMSU-A <strong>an</strong>d MHS, 14 th Conference on Satellite Meteorology <strong>an</strong>d Oce<strong>an</strong>ography,<br />

Americ<strong>an</strong> Meteorological Society, 28 J<strong>an</strong>uary–2 February, 2006, Atl<strong>an</strong>ta, GA.<br />

Ramsay, B.H., 1998: The interactive multisensor snow <strong>an</strong>d ice mapping system, Hydrological<br />

Processes, 12:1537–1546.<br />

Ramsay, B.H., 2000: Prospects for <strong>the</strong> Interactive Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping System<br />

(IMS), Proceedings of <strong>the</strong> 57 th E<strong>as</strong>tern <strong>Snow</strong> Conference, 18–19 May 2000, Syracuse,NY,<br />

161–170.<br />

Rom<strong>an</strong>ov, P. G. Goodm<strong>an</strong> <strong>an</strong>d I. Csiszar, 2000: Automated monitoring of snow cover over North<br />

America with multispectral satellite data, J. Appl. Meteorol., 39: 1866–1880.<br />

Sturm, M., J. Holmgren, <strong>an</strong>d G.E. Liston, 1995: A se<strong>as</strong>onal snow cover cl<strong>as</strong>sification system for<br />

local to global applications, Journal of Climate, 8(5): 1261–1283.<br />

103


This page is intentionally bl<strong>an</strong>k.<br />

104


105<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Time Series Analysis <strong>an</strong>d Algorithm Development<br />

for Estimating SWE in Great Lakes Area Using Microwave Data<br />

ABSTRACT<br />

AMIR E AZAR, 1 HOSNI GHEDIRA 1 , PETER ROMANOV 2 ,<br />

SHAYESTEH MAHANI 1 , AND REZA KHANBILVARDI 1<br />

The goal of this study is to develop <strong>an</strong> algorithm to estimate <strong>Snow</strong> Water Equivalent (SWE) in<br />

Great Lakes area b<strong>as</strong>ed on a three-year of SSM/I dat<strong>as</strong>et along with corresponding ground truth<br />

data. The study area is located between latitudes 41N <strong>an</strong>d 49N <strong>an</strong>d longitudes 87W <strong>an</strong>d 98W. The<br />

area is covered by 28*35 SSM/I EASE-Grid pixels with spatial resolution of 25km. Nineteen test<br />

sites were selected b<strong>as</strong>ed on se<strong>as</strong>onal average snow depth, l<strong>an</strong>d cover type. Each of <strong>the</strong> sites<br />

covers <strong>an</strong> area of 25km*25km with minimum of one snow reporting station inside. Two types of<br />

ground truth data were used: 1) point-b<strong>as</strong>ed snow depth observations from NCDC; 2) grid b<strong>as</strong>ed<br />

SNODAS-SWE dat<strong>as</strong>et, produced by NOHRSC. To account for l<strong>an</strong>d cover variation in a<br />

qu<strong>an</strong>titative way a NDVI w<strong>as</strong> used. To do <strong>the</strong> <strong>an</strong>alysis, three scattering signatures of GTVN<br />

(19V–37V), GTH (19H–37H), <strong>an</strong>d SSI (22V–85V) were derived. The <strong>an</strong>alysis shows that at lower<br />

latitudes of <strong>the</strong> study area <strong>the</strong>re is no correlation between GTH <strong>an</strong>d GTVN versus snow depth. On<br />

<strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d SSI shows <strong>an</strong> average correlation of 75 percent with snow depth in lower latitudes<br />

which makes it suitable for shallow snow identification. In <strong>the</strong> model development a non-linear<br />

algorithm w<strong>as</strong> defined to estimate SWE using SSM/I signatures along with <strong>the</strong> NDVI values of <strong>the</strong><br />

pixels. The results show up to 60 percent correlation between <strong>the</strong> estimated SWE <strong>an</strong>d ground truth<br />

SWE. The results showed that <strong>the</strong> new algorithm improved <strong>the</strong> SWE estimation by more th<strong>an</strong> 20<br />

percent for specific test sites.<br />

Keywords: Microwave SSM/I, NDVI, SWE.<br />

INTRODUCTION<br />

Knowing <strong>the</strong> se<strong>as</strong>onal variation of snowcover <strong>an</strong>d snowpack properties is of critical import<strong>an</strong>ce<br />

for <strong>an</strong> effective m<strong>an</strong>agement of water resources. Satellites operating in <strong>the</strong> optical wavelength<br />

have monitored snowcover throughout <strong>the</strong> Nor<strong>the</strong>rn Hemisphere for more th<strong>an</strong> thirty years. These<br />

sensors c<strong>an</strong> detect snowcover during daylight <strong>an</strong>d cloud-free conditions. In contr<strong>as</strong>t to visible<br />

b<strong>an</strong>ds, remote me<strong>as</strong>urements operation in microwave region offers <strong>the</strong> potential of monitoring <strong>the</strong><br />

snowpack water equivalent <strong>an</strong>d wetness due to penetrating capability of <strong>the</strong> radiation at <strong>the</strong>se<br />

frequencies. Hallikainen et al. (1984) introduced <strong>an</strong> algorithm for estimating SWE using p<strong>as</strong>sive<br />

microwave Sc<strong>an</strong>ning Multi-ch<strong>an</strong>nel Microwave Radiometer (SMMR) data. The process involved<br />

<strong>the</strong> subtraction of a fall image from a winter image in vertical polarization of 18 <strong>an</strong>d 37 GHz<br />

frequencies. The difference, ΔT, w<strong>as</strong> used to define linear relationships between ΔT <strong>an</strong>d SWE.<br />

Aschbacher (1989) proposed <strong>an</strong> SPT algorithm for estimating snow depth <strong>an</strong>d snow water<br />

equivalent that w<strong>as</strong> b<strong>as</strong>ed on a combination of SSM/I ch<strong>an</strong>nels. Fur<strong>the</strong>r studies revealed that since<br />

1 NOAA-CREST, City University of NY, 137 th St & Convent Ave. New York, NY.<br />

2 NOAA World Wea<strong>the</strong>r Building, 5200 Auth Rd, Camp Springs, MD.


l<strong>an</strong>d cover is not considered <strong>as</strong> part of equation in <strong>the</strong> algorithm, <strong>the</strong> model is not very accurate.<br />

Ch<strong>an</strong>g et al. (1987) related <strong>the</strong> difference between <strong>the</strong> SMMR brightness temperatures in 37 GHz<br />

<strong>an</strong>d 18 GHz ch<strong>an</strong>nels to derive snow depth – brightness temperature relationship for a uniform<br />

snow field, SD=1.59 [Tb 18H–Tb37H]. Goodison <strong>an</strong>d Walker (1995) introduced <strong>an</strong>o<strong>the</strong>r<br />

algorithm to estimate SWE using SSM/I ch<strong>an</strong>nels. They used vertical gradient (GTV) between<br />

brightness temperatures at 37 GHz <strong>an</strong>d 19 GHz <strong>an</strong>d defined a linear relationship between SWE <strong>an</strong>d<br />

GTV. This gradient value is obtained by subtracting <strong>the</strong> brightness temperature, Tb at frequencies<br />

of 37 <strong>an</strong>d 19 GHz <strong>an</strong>d dividing it by a const<strong>an</strong>t (Goodison, Walker 1995). Grody (1996)<br />

developed <strong>an</strong> image cl<strong>as</strong>sification algorithm to generate global snow map from Special Sensor<br />

Microwave Imager (SSM/I) data. The algorithm employs a decision tree technique <strong>an</strong>d uses<br />

thresholds to filter out precipitation, warm desert, cold desert <strong>an</strong>d frozen surfaces. De Seve et al.<br />

(1997) applied two previously developed models by Hallikainen <strong>an</strong>d Goodison-Walker to James<br />

Bay area in La Gr<strong>an</strong>de River watershed, Quebec, C<strong>an</strong>ada to estimate SWE using SSM/I images.<br />

The investigations revealed that both models tend to underestimate SWE especially when SWE<br />

w<strong>as</strong> more th<strong>an</strong> 200mm. A modified version of Goodison-Walker algorithm w<strong>as</strong> suggested. Foster<br />

et al. (1999) have modeled various snow crystals shapes in different sizes <strong>an</strong>d concluded that <strong>the</strong><br />

shape of <strong>the</strong> crystal h<strong>as</strong> little effect on <strong>the</strong> scattering in microwave. A physically b<strong>as</strong>ed snow<br />

emission model w<strong>as</strong> introduced by Pulliainen et al. (1999) of Helsinki University of technology<br />

(HUT snow emission model). The model <strong>as</strong>sumes that scattering of <strong>the</strong> microwave radiation<br />

inside <strong>the</strong> medium is mostly in forward direction. The scattering coefficient is weighted by <strong>an</strong><br />

empirical factor. The brightness temperature is computed by solving <strong>the</strong> radiative tr<strong>an</strong>sfer<br />

equation. A boreal forest c<strong>an</strong>opy model proposed by Kurvonen et al. (1994) w<strong>as</strong> used to account<br />

for <strong>the</strong> influence of vegetation on <strong>the</strong> brightness temperature. Atmospheric effects were neglected<br />

<strong>an</strong>d <strong>the</strong> snow grain size w<strong>as</strong> allowed to vary in <strong>the</strong> model. Derksen (2004) carried out a detailed<br />

evaluation of SWE <strong>an</strong>d SCE derived using SMMR <strong>an</strong>d SSM/I data over <strong>the</strong> south Central part of<br />

C<strong>an</strong>ada. The new technique to infer SWE from satellite data incorporated different algorithms,<br />

open environments, deciduous, coniferous, <strong>an</strong>d spars forest cover <strong>an</strong>d calculated SWE <strong>as</strong> weighted<br />

average of all four estimates. SWE = FDSWED + FC SWEC + FS SWES + FOSWEO, where (F) is <strong>the</strong><br />

fraction of each l<strong>an</strong>d cover type within a pixel, D, C, S, <strong>an</strong>d O correspondingly represent<br />

deciduous forest, coniferous forest, S sparse forest, <strong>an</strong>d O open prairie environments. P<strong>as</strong>sive<br />

microwave dat<strong>as</strong>et <strong>an</strong>d in situ SWE observation were compared <strong>an</strong>d showed that <strong>the</strong> SMMR<br />

brightness temperature adjustments are required to produce SWE that would fit SWE inferred<br />

from SSM/I. SWE <strong>an</strong>d SCE time series for December through March for a period of 88 years were<br />

<strong>an</strong>alyzed to examined <strong>the</strong> variability of SWE <strong>an</strong>d SCE (Derksen, 2004). Tedesco et al. (2004)<br />

proposed <strong>an</strong> Artificial Neural Network (ANN) technique for <strong>the</strong> retrieval of SWE from SSM/I<br />

data. They have used a multilayer perceptron (MLP) with various inputs to estimate SWE. First,<br />

brightness temperatures simulated by me<strong>an</strong>s of HUT snow estimation model were employed. The<br />

second approach made use of a subset of me<strong>as</strong>ured values. The input layer consists of four<br />

neurons, made up of 19 <strong>an</strong>d 37 GHz vertical <strong>an</strong>d horizontal brightness temperatures <strong>an</strong>d <strong>the</strong> output<br />

w<strong>as</strong> snow parameters. The results showed higher perform<strong>an</strong>ce of ANN model compare to o<strong>the</strong>r<br />

methods. In 2005 Derksen conducted a study to <strong>as</strong>sess <strong>the</strong> accuracy of <strong>an</strong> inter-<strong>an</strong>nually consistent<br />

zone of high p<strong>as</strong>sive microwave derived SWE retrievals co-located with <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> nor<strong>the</strong>rn<br />

boreal forest, using extended tr<strong>an</strong>sects of in situ snow cover me<strong>as</strong>urements (Derksen, 2005). The<br />

research conducted by Kelly et al. (2001) w<strong>as</strong> focused at <strong>the</strong> development of a global snow<br />

monitoring for The Adv<strong>an</strong>ced Microwave Sc<strong>an</strong>ning Radiometer – EOS (AMSR-E) onboard Aqua<br />

satellite. The proposed algorithm had <strong>the</strong> following form: SWE (mm) = B*(TbH18–TbH37),<br />

where TbH18 <strong>an</strong>d TbH37 are horizontal polarized brightness temperature at 18 <strong>an</strong>d 37 GHz <strong>an</strong>d B<br />

coefficient h<strong>as</strong> been calibrated <strong>as</strong> 4.8 mm K –1 for SMMR data. Latter Kelly et al. (2003) described<br />

<strong>the</strong> development <strong>an</strong>d testing of <strong>an</strong> algorithm to estimate global snow cover volume from<br />

spaceborne p<strong>as</strong>sive microwave, AMSR-E.<br />

The above algorithms used <strong>the</strong> spectral difference between microwave ch<strong>an</strong>nels from various<br />

sensors to estimate SWE or snow depth. However o<strong>the</strong>r snow or l<strong>an</strong>d parameters such <strong>as</strong> snow<br />

grain size <strong>an</strong>d l<strong>an</strong>d cover type <strong>an</strong>d conditions have effects on scattering in microwave. Although,<br />

some researchers introduced l<strong>an</strong>d cover type to <strong>the</strong>ir models but <strong>the</strong>ir algorithms were developed<br />

106


<strong>an</strong>d validated regionally so <strong>the</strong>y c<strong>an</strong> not be used for o<strong>the</strong>r study are<strong>as</strong>. In addition, <strong>the</strong>se<br />

algorithms use a multi-regression approach to account for <strong>the</strong> l<strong>an</strong>d cover type variation. Then,<br />

development of <strong>an</strong> algorithm which considers variation of l<strong>an</strong>d cover qu<strong>an</strong>titatively <strong>an</strong>d c<strong>an</strong> be<br />

used <strong>an</strong>d validated for different are<strong>as</strong> is necessitated. Normalized Difference Vegetation Index<br />

(NDVI) h<strong>as</strong> been widely used to represent <strong>the</strong> health <strong>an</strong>d greenness of <strong>the</strong> vegetation. In this study<br />

a non-linear algorithm is proposed, which estimates SWE using spectral difference between SSM/I<br />

ch<strong>an</strong>nels along with NDVI data.<br />

DATA USED<br />

SSM/I Data<br />

SSM/I p<strong>as</strong>sive microwave radiometer with seven ch<strong>an</strong>nels is operating at five frequencies (19,<br />

35, 22, 37.0, <strong>an</strong>d 85.5 GHz) <strong>an</strong>d dual-polarization (except at 22GHz which is V-polarization only).<br />

The sensor spatial resolution varies for different ch<strong>an</strong>nels frequencies. In this study Scalable Equal<br />

Area Earth Grid EASE-Grid SSM/I products distributed by National <strong>Snow</strong> <strong>an</strong>d Ice Data Center<br />

(NSIDC) were used. EASE-Grid spatial resolution is slightly more th<strong>an</strong> 25km (25.06km) for all<br />

<strong>the</strong> ch<strong>an</strong>nels (NSIDC) although <strong>the</strong> recorded resolution of <strong>the</strong> microwave spectrum with longer<br />

wavelengths is more th<strong>an</strong> 50km. The three EASE-Grid projections comprise two azimuthal equalarea<br />

projections for <strong>the</strong> Nor<strong>the</strong>rn or Sou<strong>the</strong>rn hemisphere, respectively <strong>an</strong>d a global cylindrical<br />

equal area projection. In this we study have used a Nor<strong>the</strong>rn hemisphere azimuthal equal-area.<br />

Normalized Difference Vegetation Index (NDVI)<br />

NDVI is typically used to represent <strong>the</strong> vegetation cover properties <strong>an</strong>d it defines <strong>as</strong> a difference<br />

between reflect<strong>an</strong>ce in visible <strong>an</strong>d near infrared spectral b<strong>an</strong>ds divided by <strong>the</strong>ir sum (NDVI =<br />

(NIR – VIS)/(NIR + VIS)). The NDVI data for this study were obtained from <strong>the</strong> NOAA/NASA<br />

Pathfinder Adv<strong>an</strong>ced Very High Resolution Radiometer (AVHRR) which is distributed at<br />

Goddard Space Flight Center (GSFC). The spatial resolution is 8km * 8km obtained within a 10<br />

day period that h<strong>as</strong> <strong>the</strong> fewest cloud. To facilitate <strong>the</strong> comparison <strong>an</strong>d matching of <strong>the</strong> two dat<strong>as</strong>ets<br />

(NDVI <strong>an</strong>d SSM/I) NDVI data resampled <strong>an</strong>d were brought to <strong>the</strong> same EASE-Grid projection at<br />

25km spatial resolution.<br />

Ground Truth <strong>Snow</strong> Data<br />

Point Gauge Me<strong>as</strong>urements<br />

Surface observations of snow depth data for <strong>the</strong> study were obtained from point <strong>the</strong> National<br />

Climate Data Center (NCDC). The point me<strong>as</strong>urements were averaged <strong>an</strong>d gridded to 25km<br />

spatial resolution to match EASE-Grid SSM/I spatial resolution. To incre<strong>as</strong>e <strong>the</strong> reliability <strong>an</strong>d<br />

avoid errors due to interpolation, we have used only those pixels where station data were<br />

available. If more th<strong>an</strong> one station data were available in a given SSM/I pixel, <strong>the</strong> station data<br />

were averaged.<br />

SNODAS SWE<br />

<strong>Snow</strong> products generated by <strong>the</strong> <strong>Snow</strong> Data Assimilation System (SNODAS) of NOAA<br />

National Wea<strong>the</strong>r Service's National Operational Hydrologic Remote Sensing Center (NOHRSC)<br />

are available since October 2003. SNODAS includes a procedure to <strong>as</strong>similate airborne gamma<br />

radiation <strong>an</strong>d ground-b<strong>as</strong>ed observations of snow covered area <strong>an</strong>d snow water equivalent,<br />

downscaled output from Numerical Wea<strong>the</strong>r Prediction (NWP) models combined in a physically<br />

b<strong>as</strong>ed, spatially distributed energy <strong>an</strong>d m<strong>as</strong>s bal<strong>an</strong>ce model. The output products have 1km spatial<br />

<strong>an</strong>d hourly temporal resolution. In order to match <strong>the</strong> EASE-Grid pixels <strong>the</strong> SNODAS SWE data<br />

were averaged to 25km.<br />

107


ANALYSIS OF THE DATA<br />

The study area is located in Great Lakes area between latitudes 41N <strong>an</strong>d 49N <strong>an</strong>d longitudes<br />

87W <strong>an</strong>d 98W. It covers parts of Minnesota, Wisconsin <strong>an</strong>d Michig<strong>an</strong> states. The area h<strong>as</strong> various<br />

l<strong>an</strong>d covers (Fig 1). The area is covered by 980 (28 by 35) EASE-Grid pixels. To do <strong>the</strong> time<br />

series <strong>an</strong>alysis 19 test sites were selected. Each site, 25km*25km, represents <strong>an</strong> SSM/I pixel. The<br />

sites were selected b<strong>as</strong>ed on <strong>the</strong>ir latitude <strong>an</strong>d <strong>the</strong>ir l<strong>an</strong>d cover type along with <strong>the</strong> <strong>an</strong>nual snow<br />

accumulation. In order to avoid wet snow conditions we used <strong>the</strong> data starting December 1 of each<br />

year to <strong>the</strong> February 28 of <strong>the</strong> year after. Three 90 day sets of data were derived for each winter.<br />

Table 1 shows geographical location of <strong>the</strong> selected sites <strong>an</strong>d <strong>the</strong>ir NDVI characteristics<br />

including <strong>the</strong> me<strong>an</strong> value <strong>an</strong>d vari<strong>an</strong>ce. High vari<strong>an</strong>ce of NDVI indicates subst<strong>an</strong>tial ch<strong>an</strong>ges.<br />

Table 1. Coordinates of selected pixels along with NDVI values<br />

Test Site<br />

SSM/I Pixels<br />

Latitude Longitude<br />

NDVI= (P–128) × 0.008<br />

P value P me<strong>an</strong> P vari<strong>an</strong>ce<br />

1 42.33 –93.62 123.00 123.86 1.07<br />

2 42.89 –91.97 123.00 123.20 0.45<br />

3 43.63 –91.43 124.00 124.86 2.27<br />

4 44.14 –90.57 145.00 143.11 6.19<br />

5 44.39 –89.12 128.00 132.00 4.24<br />

6 45.12 –89.11 130.00 132.22 5.14<br />

7 46.07 –88.19 162.00 155.33 13.53<br />

8 45.59 –88.21 152.00 157.67 9.06<br />

9 46.09 –88.79 160.00 161.44 9.10<br />

10 46.80 –88.46 166.00 152.56 10.63<br />

11 46.80 –88.16 150.00 157.22 7.95<br />

12 46.83 –89.69 159.00 158.89 10.53<br />

13 45.36 –91.18 132.00 132.78 7.50<br />

14 45.56 –92.68 125.00 130.89 3.66<br />

15 47.26 –92.78 149.00 143.33 5.10<br />

16 48.01 –91.88 148.00 154.56 9.82<br />

17 47.92 –94.08 135.00 137.89 6.55<br />

18 48.40 –95.99 135.00 132.33 4.36<br />

19 47.47 –97.47 125.00 124.78 0.67<br />

108<br />

Dry l<strong>an</strong>d <strong>an</strong>d<br />

Cropl<strong>an</strong>d<br />

Cropl<strong>an</strong>d/Gr<strong>as</strong>sl<strong>an</strong>d<br />

Cropl<strong>an</strong>d<br />

/Woodl<strong>an</strong>d<br />

Mixed Forest<br />

Deciduous Broadleaf<br />

Evergreen<br />

Needleleaf<br />

Urb<strong>an</strong> <strong>an</strong>d Built-up<br />

L<strong>an</strong>d<br />

Water Body<br />

Figure 1: L<strong>an</strong>d Cover Image According to USGS National Atl<strong>as</strong> of L<strong>an</strong>d Cover Characteristics


109<br />

Test Site<br />

Figure 2. Variation of SWE for winter se<strong>as</strong>on 2003–2004 <strong>an</strong>d <strong>the</strong> corresponding Box plot<br />

In this study we verify <strong>the</strong> correlation between SSM/I ch<strong>an</strong>nels <strong>an</strong>d snow depth <strong>an</strong>d SWE for<br />

different types of l<strong>an</strong>d cover located in different latitudes. Three series (2002–2004) of SSM/I<br />

ch<strong>an</strong>nels versus snow depth <strong>an</strong>d SWE for each of <strong>the</strong> selected sites were derived. The SSM/I data<br />

were obtained from <strong>the</strong> descending p<strong>as</strong>s of Defense Meteorological Satellite Program (DMSP)<br />

satellites. There are three SSM/I scattering signatures used in this <strong>an</strong>alysis. The first scattering<br />

signature named GTH (19H–37H) is <strong>the</strong> difference between 19 <strong>an</strong>d 37 GHz in horizontal<br />

polarization. The second signature, GTVN (19V–37V) shows <strong>the</strong> discrep<strong>an</strong>cy between vertically<br />

polarized 19 <strong>an</strong>d 37 GHz. Finally, SSI (22V–85V) presents <strong>the</strong> difference between 22 <strong>an</strong>d 85 GHz<br />

in vertical polarization. SSI c<strong>an</strong> be used to identify shallow snowcover. The Box plot of <strong>the</strong><br />

signatures GTH <strong>an</strong>d GTVN for winter se<strong>as</strong>on 2003–2004 is shown in figure 3. The outliers in <strong>the</strong><br />

Box plots r<strong>an</strong>ge from –20 to 20 are due to ei<strong>the</strong>r sensor or data processing errors which need to be<br />

eliminated. GTVN me<strong>an</strong> r<strong>an</strong>ges from 5 to 15 for all <strong>the</strong> pixels except for <strong>the</strong> test site 12 which is<br />

very close to <strong>the</strong> lake. GTH Box plot <strong>an</strong>d me<strong>an</strong> h<strong>as</strong> <strong>the</strong> same pattern <strong>as</strong> GTVN. The test site 12 h<strong>as</strong><br />

a me<strong>an</strong> around –5 for GTVN (19V–37V) indicating of <strong>the</strong> fact that part of SSM/I pixel is water.<br />

Since <strong>the</strong> SSM/I sensor have different spatial resolution for various ch<strong>an</strong>nel, <strong>the</strong> scattering from<br />

pixel 12 is disturbed in ch<strong>an</strong>nel 19GHz (69km resolution) by <strong>the</strong> lake however it is not disturbed<br />

in ch<strong>an</strong>nel 37GHz (37km resolution). The low scattering of <strong>the</strong> water <strong>an</strong>d high scattering of <strong>the</strong><br />

l<strong>an</strong>d make <strong>the</strong> difference of 19V–37V a negative number.<br />

GTVN (19H-37H) (k)<br />

Box Plot of GTVN for Points 1-19 Box Plot of GTH for Points 1-19<br />

Test Site<br />

GTH (19H-37H) (k)<br />

Test site<br />

Figure 3. Box-Whiskers plot of GTVN <strong>an</strong>d GTH for winter se<strong>as</strong>on 2003–2004


After elimination of <strong>the</strong> outliers <strong>an</strong>d negative signatures, a three year time series of GTVN <strong>an</strong>d<br />

snow depth for each of <strong>the</strong> test sites w<strong>as</strong> produced. Figure 4 illustrates <strong>the</strong> trend of SSM/I<br />

signature of GTVN (19V–37V) versus snow depth at site 9. The plot shows that <strong>the</strong> discrep<strong>an</strong>cy<br />

GTVN (19V–37V) incre<strong>as</strong>es with incre<strong>as</strong>ing snow during <strong>the</strong> winter se<strong>as</strong>ons. This is due to <strong>the</strong><br />

high sensitivity of ch<strong>an</strong>nel 37GHz to snow. Contradictory to <strong>the</strong> high latitudes, low latitude pixels<br />

do not show a consistent se<strong>as</strong>onal pattern for snow depth <strong>an</strong>d GTVN (Figure 5).<br />

GTVN (K)<br />

GTVN (K)<br />

20<br />

10<br />

0<br />

0 100 200 300 400 500<br />

Days<br />

600 700 800 900 1000<br />

0<br />

20<br />

10<br />

110<br />

GTVN<br />

<strong>Snow</strong> Depth<br />

Figure 4. Three year time series of GTVN (19V–37V) vs. <strong>Snow</strong> Depth for point 9<br />

GTVN<br />

<strong>Snow</strong> Depth<br />

0<br />

0 100 200 300 400 500<br />

Days<br />

600 700 800 900 1000<br />

0<br />

Figure 5. Three year time series of GTVN (19V–37V) vs. <strong>Snow</strong> Depth for point 2<br />

The above figures indicate high correlations between snow depth <strong>an</strong>d GTVN for high latitudes<br />

<strong>an</strong>d lack of correlation for sites located in low latitudes. To qu<strong>an</strong>tify <strong>the</strong> relationships, correlation<br />

coefficient of various SSM/I scattering signature versus snow depth <strong>an</strong>d SWE are presented in<br />

figure (6). SSM/I signatures at sites 11, 12, <strong>an</strong>d 13 do have correlation with snow depth since <strong>the</strong><br />

scattering is disturbed by water bodies. A graphical representation of correlation coefficients for<br />

GTVN, GTH, <strong>an</strong>d SSI for all <strong>the</strong> test sites is shown in figure 6.<br />

800<br />

600<br />

400<br />

200<br />

800<br />

600<br />

400<br />

200<br />

<strong>Snow</strong> Depth (mm)<br />

<strong>Snow</strong> Depth (mm)


Correlation Coe.<br />

Correlation Coe.<br />

Correlation Coe.<br />

GTVN vs. <strong>Snow</strong> Depth SSI vs. <strong>Snow</strong> Depth GTH VS <strong>Snow</strong> Depth<br />

0.90<br />

0.80<br />

0.70<br />

0.60<br />

0.50<br />

0.40<br />

0.30<br />

0.20<br />

0.10<br />

0.00<br />

Winter 01-02<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

0.90<br />

0.80<br />

0.70<br />

0.60<br />

0.50<br />

0.40<br />

0.30<br />

0.20<br />

0.10<br />

0.00<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

0.90<br />

0.80<br />

0.70<br />

0.60<br />

0.50<br />

0.40<br />

0.30<br />

0.20<br />

0.10<br />

0.00<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

111<br />

Test Site<br />

Test Site<br />

Test Site<br />

Winter 02-03<br />

Winter 03-04<br />

Figure 6. Correlations of snow depth vs. SSM/I signatures GTVN (19v–37v), GTH (19h-37h), <strong>an</strong>d SSI (22v-<br />

85v) for various test sites (TS) for winter se<strong>as</strong>ons 01–02, 02–03, 03–04<br />

The correlation coefficients between snow depth <strong>an</strong>d scattering signatures follow a consistent<br />

pattern for all <strong>the</strong> winter se<strong>as</strong>ons. For all <strong>the</strong> sites GTVN <strong>an</strong>d GTH show <strong>the</strong> same correlation with<br />

snow depth. In o<strong>the</strong>r words, <strong>the</strong> difference between vertically <strong>an</strong>d horizontally polarized signatures<br />

is negligible in terms of correlations with snow depth. Contrary to GTVN <strong>an</strong>d GTH, SSI h<strong>as</strong> a<br />

different pattern. It h<strong>as</strong> <strong>the</strong> domin<strong>an</strong>t correlation for test sites 1 to 5 but for sites located in high<br />

latitudes GTVN becomes <strong>the</strong> domin<strong>an</strong>t. This is because of <strong>the</strong> saturation of <strong>the</strong> 85GHZ ch<strong>an</strong>nel<br />

over a deep snow pack. SSI c<strong>an</strong> be used to identify <strong>an</strong>d to estimate SWE over shallow snow. In<br />

c<strong>as</strong>e of SWE <strong>an</strong>d SSM/I signatures, Figure (7) illustrates <strong>the</strong> correlations between SWE <strong>an</strong>d<br />

different SSM/I spectral signatures. For test sites 1 to 5 SSI h<strong>as</strong> <strong>the</strong> higher correlation but for <strong>the</strong><br />

o<strong>the</strong>r sites GTVN <strong>an</strong>d GTH show better correlations with SWE. Figure 7 also shows that <strong>the</strong><br />

correlations between SWE <strong>an</strong>d scattering signatures are higher th<strong>an</strong> those for snow depth. This<br />

indicates that SSM/I signatures c<strong>an</strong> be a better estimator of SWE th<strong>an</strong> of <strong>the</strong> snow depth.


Correlation Coe.<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

Points<br />

GTVN vs SWE SSI vs SWE GTH VS SWE<br />

112<br />

Winter 03-04<br />

Figure 7. Variation of correlations of SWE vs SSM/I scattering signatures for various points for winter 03–04<br />

The scatter plots of SWE versus <strong>the</strong> three SSM/I signatures (GTVN, GTH, SSI) have been<br />

produced for <strong>the</strong> all <strong>the</strong> test sites. Figure 8 illustrates <strong>the</strong> variation of SWE versus scattering<br />

signatures for selected test sites (2, 9, <strong>an</strong>d 18).<br />

RESULTS OF THE ANALYSIS<br />

The presented correlation coefficients in figures 6, 7 represent different winter se<strong>as</strong>ons from<br />

2001–2004. The <strong>an</strong>alysis of <strong>the</strong> results indicates <strong>the</strong> following:<br />

1) For test sites located in low latitudes, below 45N, (1, 2, 3, <strong>an</strong>d 4) only SSI exhibits some<br />

correlation with <strong>the</strong> snow depth. There is no noticeable correlation of GTH <strong>an</strong>d GTVN vs. <strong>the</strong><br />

snow depth. This is due to <strong>the</strong> saturation saturation effect in of ch<strong>an</strong>nel 85GHz which makes SSI<br />

only suitable for estimating properties of a shallow snow pack.<br />

2) Sites located in mid- latitudes, 45N–46N, (sites 6, 7, 8, 9) <strong>the</strong>re is some correlation between<br />

GTH <strong>an</strong>d GTVN vs. snow depth but SSI shows no correlation with <strong>the</strong> snow depth.<br />

3) No correlation is observed at sites that are very close to <strong>the</strong> lake (10, 11, 12, 13, <strong>an</strong>d 14). This<br />

is due to <strong>the</strong> different spatial resolution of SSM/I in various spectral b<strong>an</strong>ds. The sensors field of<br />

view incre<strong>as</strong>es from 37km at 37GHz to 69km for 19GHz. Therefore if a me<strong>as</strong>urement is made<br />

close to <strong>the</strong> lake, <strong>the</strong> effect of <strong>the</strong> open water may be different in two ch<strong>an</strong>nels.<br />

4) Test sites located in forested are<strong>as</strong> away from <strong>the</strong> lake show moderate correlations of snow<br />

with GTH <strong>an</strong>d GTVN. In addition, scatter plots show <strong>an</strong> attenuation of brightness temperature due<br />

to forested l<strong>an</strong>d cover.<br />

5) Both GTH <strong>an</strong>d GTVN show high correlations with physical characteristics of <strong>the</strong> snow pack<br />

which makes <strong>the</strong>m good potential estimators for snow depth <strong>an</strong>d SWE. The highest correlations<br />

are observed in north of <strong>the</strong> US which is due to larger amount of se<strong>as</strong>onal snow, colder wea<strong>the</strong>r<br />

<strong>an</strong>d less number of freeze/thaw events during a winter se<strong>as</strong>on.<br />

6) Table 3 presents <strong>the</strong> correlation between SSM/I spectral signatures <strong>an</strong>d SWE obtained from<br />

SNODAS. The results show higher <strong>an</strong>d more consistent correlation coefficients for SWE th<strong>an</strong> for<br />

snow depth.<br />

7) Figure (8) shows <strong>the</strong> scatter plots of SSM/I Signatures versus SWE (SNODAS) <strong>an</strong>d SSM/I<br />

Signatures versus <strong>Snow</strong> Depth (stations) for winter 2003–2004 in different test sites. Lines fitted<br />

to each graph have various slopes <strong>an</strong>d intercepts. This demonstrates that having one linear<br />

algorithm (e.g. Ch<strong>an</strong>g or Goodison-Walker) may not be enough for snow depth or SWE in a<br />

variety of environmental <strong>an</strong>d geographical conditions.


Figure 8. SSM/I scattering signatures signature (GTH [19H–37H], GTVN [19H–37V], <strong>an</strong>d SSI [22H-85H])<br />

vs. SWE (SNODAS) for winter 2003–2004<br />

113


ALGORITHM DEVELOPMENT<br />

The study area, Great Lakes area, is located in <strong>the</strong> tr<strong>an</strong>sitional zone for snow which experiences<br />

snow melting during <strong>the</strong> winter se<strong>as</strong>on. In addition to complexity of snow characteristics,<br />

variability of <strong>the</strong> l<strong>an</strong>d cover types makes it more difficult to accurately estimate SWE from p<strong>as</strong>sive<br />

microwave observations with a single linear model such <strong>as</strong> <strong>the</strong> one of Ch<strong>an</strong>g or Goodison-Walker<br />

(Ch<strong>an</strong>g et al, 1987, Goodison-Walker 1995). Figure (8) shows that <strong>the</strong> slope of <strong>the</strong> best fitted<br />

equations varies for different test sites (2, 9, <strong>an</strong>d 18). These test sites have different l<strong>an</strong>d cover<br />

type <strong>an</strong>d different NDNI values. In order to qu<strong>an</strong>titatively account for scattering attenuation<br />

originating from forested are<strong>as</strong>, NDVI values are suggested to use. Figure (9) illustrates <strong>the</strong><br />

variation of slope <strong>an</strong>d NDVI for winter 2003–2004 for all <strong>the</strong> sites. Both NDVI <strong>an</strong>d <strong>the</strong> slope have<br />

<strong>the</strong> same trend. This indicates that <strong>the</strong> incre<strong>as</strong>e of NDVI makes <strong>the</strong> slope of <strong>the</strong> relationship<br />

steeper.<br />

16<br />

14<br />

12<br />

10<br />

Slope (m)<br />

8<br />

6<br />

4<br />

2<br />

Winter 03-04, Variation of NDVI with <strong>the</strong> slope of SWE vs GTVN<br />

Slope<br />

NDVI<br />

0<br />

-0.1<br />

0 5 10 15 20<br />

Test site<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

NDVI<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

114<br />

6<br />

5<br />

4<br />

Slope<br />

3<br />

2<br />

1<br />

0<br />

03-04 Slopes<br />

SD Slopes <strong>an</strong>d NDVI<br />

0 5 10 15 20<br />

Test site<br />

0.15<br />

NDVI<br />

Figure 9. Variations of <strong>the</strong> slope of <strong>the</strong> bets fitted line to scatter plots with NDVI for <strong>the</strong> test sites for winter<br />

03–04, SWE vs. GTVN (left), SD vs. GTVN (right)<br />

The scatter plots of slope versus NDVI for SWE <strong>an</strong>d snow depth are illustrated in figure 10. The<br />

modified scatter plot for 2002–2003 is for <strong>the</strong> test sites that snow depth vs scattering signature<br />

GTVN (19v–37v) showed correlation coefficient more th<strong>an</strong> 50 percent.<br />

Slope<br />

slope<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Winter 03–04 , SWE (mm) Winter 03–04, <strong>Snow</strong> depth (cm)<br />

0 0.05 0.1 0.15 0.2 0.25 0.3<br />

NDVI<br />

slope<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

NDVI<br />

0.3<br />

0.25<br />

0.2<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

0 0.05 0.1 0.15 0.2 0.25 0.3<br />

Winter 02–03, <strong>Snow</strong> depth (cm) Winter 02–03, <strong>Snow</strong> depth (cm), Modified<br />

0 0.05 0.1 0.15 0.2 0.25 0.3<br />

NDVI<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

NDVI<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25 0.3<br />

Figure 10. Variation of <strong>the</strong> derived slope from <strong>the</strong> scatter plots vs. NDVI<br />

slope<br />

NDVI<br />

-0.1


Considering <strong>the</strong> facts mentioned above we propose a new algorithm which relates SWE <strong>an</strong>d <strong>the</strong><br />

SSM/I scattering signature GTVN (19v–37v) <strong>an</strong>d accounts for possible variation of NDVI:<br />

SWE=F* (A* NDVI* + B)*GTVN, While NDVI >= 0<br />

SWE=C*SSI+D While NDVI


<strong>Snow</strong> Depth (mm)<br />

<strong>Snow</strong> Depth (cm)<br />

<strong>Snow</strong> Depth (cm)<br />

<strong>Snow</strong> Depth (cm)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Goodison-Walker Alg. Ch<strong>an</strong>g Alg. New Alg.<br />

Winter 03–04 (SWE) Winter 03–04 (SWE) Winter 03–04 (SWE)<br />

GOODISON-WALKER Linear Algorithm<br />

R=0.7734<br />

RMSE=63mm<br />

Bi<strong>as</strong>= -47mm<br />

MAX(E)=32mm<br />

MAX(S)=173mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

<strong>Snow</strong> Depth (mm)<br />

CHANG Linear Algorithm<br />

100<br />

90<br />

R=0.80306<br />

80<br />

RMSE=58mm<br />

70<br />

Bi<strong>as</strong>= -44mm<br />

60<br />

MAX(e)=43mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(s)=173mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated SWE (mm)<br />

116<br />

<strong>Snow</strong> Depth (mm)<br />

AZAR Non-linear Algorithm<br />

100<br />

90<br />

R=0.7734<br />

80<br />

RMSE=30mm<br />

70<br />

Bi<strong>as</strong>= 3mm<br />

60<br />

MAX(E)=127mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=173mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

Winter 03–04 (SD) Winter 03–04 (SD) Winter 03–04 (SD)<br />

GOODISON-WALKER Linear Algorithm<br />

100<br />

90<br />

R=0.50634<br />

80<br />

RMSE=18mm<br />

70<br />

Bi<strong>as</strong>= -13mm<br />

60<br />

MAX(E)=24mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=50mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

<strong>Snow</strong> Depth (cm)<br />

CHANG Linear Algorithm<br />

100<br />

90<br />

R=0.4396<br />

80<br />

RMSE=13mm<br />

70<br />

Bi<strong>as</strong>= 7mm<br />

60<br />

MAX(e)=43mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(s)=36mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (cm)<br />

<strong>Snow</strong> Depth (cm)<br />

AZAR Non-linear Algorithm<br />

100<br />

90<br />

R=0.50634<br />

80<br />

RMSE=14mm<br />

70<br />

Bi<strong>as</strong>= -8mm<br />

60<br />

MAX(E)=33mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=50mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

Winter 02–03 (SD) Winter 02–03 (SD) Winter 02–03 (SD)<br />

GOODISON-WALKER Linear Algorithm<br />

100<br />

90<br />

R=0.69708<br />

80<br />

RMSE=9mm<br />

70<br />

Bi<strong>as</strong>= -4mm<br />

60<br />

MAX(E)=27mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=36mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

<strong>Snow</strong> Depth (cm)<br />

CHANG Linear Algorithm<br />

100<br />

90<br />

R=0.4396<br />

80<br />

RMSE=13mm<br />

70<br />

Bi<strong>as</strong>= 7mm<br />

60<br />

MAX(e)=43mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(s)=36mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (cm)<br />

<strong>Snow</strong> Depth (cm)<br />

AZAR Non-linear Algorithm<br />

100<br />

90<br />

R=0.69708<br />

80<br />

RMSE=8mm<br />

70<br />

Bi<strong>as</strong>= 1mm<br />

60<br />

MAX(E)=38mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=36mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

Winter 01–02 (SD) Winter 01–02 (SD) Winter 01–02 (SD)<br />

AZAR Non-linear Algorithm<br />

100<br />

90<br />

R=0.62132<br />

80<br />

RMSE=7mm<br />

70<br />

Bi<strong>as</strong>= 1mm<br />

60<br />

MAX(E)=36mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=38mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (mm)<br />

<strong>Snow</strong> Depth (cm)<br />

CHANG Linear Algorithm<br />

100<br />

90<br />

R=0.53409<br />

80<br />

RMSE=16mm<br />

70<br />

Bi<strong>as</strong>= 12mm<br />

60<br />

MAX(e)=51mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(s)=38mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Estimated (cm)<br />

<strong>Snow</strong> Depth (cm)<br />

AZAR Non-linear Algorithm<br />

100<br />

90<br />

R=0.62132<br />

80<br />

RMSE=7mm<br />

70<br />

Bi<strong>as</strong>= 1mm<br />

60<br />

MAX(E)=36mm<br />

50<br />

40<br />

30<br />

20<br />

10<br />

MAX(S)=38mm<br />

0<br />

0 10 20 30 40 50 60 70 80 90 10<br />

Estimated (mm)<br />

Figure 11. Comparison of <strong>the</strong> results for different algorithms for test site 10 (Lat = 48.6N, Lon = –88.46W,<br />

<strong>an</strong>d NDVI = 0.2)


Besides <strong>the</strong> temporal validation, <strong>the</strong> new algorithm w<strong>as</strong> spatially validated for <strong>the</strong> whole study<br />

area (Latitudes: 41N to 49N & Longitudes: –87W to –98W). There were eleven days (3 days<br />

December, 4 days J<strong>an</strong>uary, <strong>an</strong>d 4 days February) in winter 2003–2004 selected. For those days <strong>the</strong><br />

full coverage of <strong>the</strong> study area from SSM/I data w<strong>as</strong> available. The ground truth data w<strong>as</strong> obtained<br />

by averaging NOHRC SNODAS dat<strong>as</strong>et. Figure 12 shows <strong>the</strong> ground truth <strong>an</strong>d estimated SWE for<br />

J<strong>an</strong>uary 25, 2004.<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Ch<strong>an</strong>g Estimated SWE (mm) Ground Truth NOHRSC SWE (mm)<br />

0<br />

0 5 10 15 20 25 30 35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Estimated SWE CHANG Algorithm (mm),J<strong>an</strong> 25,04<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-20<br />

30<br />

25<br />

20<br />

15<br />

10<br />

117<br />

5<br />

SNODAS , J<strong>an</strong>25,04<br />

0<br />

0 5 10 15 20 25 30 35<br />

Goodison-Walker Estimated SWE (mm) New Algorithm Estimated SWE (mm)<br />

Estimated SWE AZAR Algorithm (mm),J<strong>an</strong> 25,04<br />

0<br />

0 5 10 15 20 25 30 35<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Estimated SWE Goodison Walker Algorithm (mm),J<strong>an</strong> 25,04<br />

0<br />

0 5 10 15 20 25 30 35<br />

Figure 12. Comparison of estimated SWE by various algorithms with ground truth data for J<strong>an</strong>uary 25, 2004<br />

for <strong>the</strong> study area (Lat: 41N to 49N & Lon: –87W to –98W)<br />

The NDVI image of <strong>the</strong> study area (Fig 13) shows higher values of NDVI around <strong>the</strong> lake. This<br />

is <strong>the</strong> area that both Ch<strong>an</strong>g <strong>an</strong>d Goodison-Walker algorithms highly underestimate <strong>the</strong> SWE (Fig<br />

12). In contr<strong>as</strong>t, <strong>the</strong> new non-linear algorithm c<strong>an</strong> estimate SWE in <strong>the</strong> area in <strong>the</strong> vicinity of <strong>the</strong><br />

lake with much higher accuracy (Fig 13, 14). The calculated RMSE <strong>an</strong>d correlation coefficient<br />

(R 2 ) are shown for all <strong>the</strong> three algorithms. The use of NDVI in <strong>the</strong> new algorithm results in a<br />

decre<strong>as</strong>e of <strong>the</strong> RMSE <strong>an</strong>d <strong>the</strong> incre<strong>as</strong>e of <strong>the</strong> correlation coefficient. It also incre<strong>as</strong>es <strong>the</strong> r<strong>an</strong>ge for<br />

<strong>the</strong> estimated SWE. Figure 15 demonstrates a consistent improvement in <strong>the</strong> accuracy of <strong>the</strong><br />

estimated SWE for <strong>the</strong> winter se<strong>as</strong>on of 2003–2004. For all days, application of <strong>the</strong> new<br />

developed algorithm results in <strong>the</strong> highest correlation coefficient between SSM/I <strong>an</strong>d SWE. At <strong>the</strong><br />

same time, <strong>the</strong> RMSE of SWE derived with <strong>the</strong> new algorithm is lower for all days but one. There<br />

80<br />

60<br />

40<br />

20<br />

0<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

180<br />

160<br />

140<br />

120<br />

100


is a decre<strong>as</strong>ing trend of in correlations <strong>an</strong>d incre<strong>as</strong>ing trend in SWE in February. The most<br />

probable re<strong>as</strong>on for this trend is snow melt. In February, <strong>the</strong> study area <strong>an</strong>d especially its sou<strong>the</strong>rn<br />

part experienced several melts <strong>an</strong>d refreeze of snow. Estimates of snow depth <strong>an</strong>d SWE with<br />

satellite observations in microwave become practically impossible when snow is wet.<br />

SNODAS SWE (mm)<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

NDVI Image New Algorithm Estimated SWE vs Ground Truth<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

118<br />

SNODAS SWE (mm)<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

AZAR Non-linear Algorithm<br />

R=0.69786<br />

RMSE=19mm<br />

Bi<strong>as</strong>= 0mm<br />

Ave(E)=14mm<br />

Ave(S)=15mm<br />

MAX(E)=181mm<br />

MAX(S)=177mm<br />

0<br />

0 20 40 60 80 100 120 140 160 180<br />

Estimated SWE (mm)<br />

Figure 13. NDVI image <strong>an</strong>d results of estimated SWE vs. ground truth for J<strong>an</strong>uary 25, 2004<br />

Ch<strong>an</strong>ge Alg. Goodison-Walker Alg.<br />

CHANG Linear Algorithm<br />

R=0.53679<br />

RMSE=24mm<br />

Ave(E)=22mm<br />

Ave(S)=15mm<br />

MAX(e)=72mm<br />

MAX(s)=177mm<br />

0<br />

0 20 40 60 80 100 120 140 160 180<br />

Estimated SWE (mm)<br />

SNODAS SWE (mm)<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

GOODISON-WALKER Linear Algorithm<br />

R=0.51162<br />

RMSE=23mm<br />

Ave(E)=18mm<br />

Ave(S)=15mm<br />

MAX(E)=51mm<br />

MAX(S)=177mm<br />

0<br />

0 20 40 60 80 100 120 140 160 180<br />

Estimated SWE (mm)<br />

Figure 14. Results of estimated SWE using Ch<strong>an</strong>g <strong>an</strong>d Goodison-Walker algorithm vs ground truth for<br />

J<strong>an</strong>uary 25, 2004


RMSE<br />

Non linear Azar<br />

Ch<strong>an</strong>g<br />

Goodison-Walker<br />

Non linear Azar<br />

Ch<strong>an</strong>g<br />

Goodison-Walker<br />

CONCLUSIONS<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0.80<br />

0.70<br />

0.60<br />

0.50<br />

0.40<br />

0.30<br />

0.20<br />

0.10<br />

0.00<br />

6-Dec-03<br />

6-Dec-03<br />

Days<br />

Days<br />

13-Dec-03<br />

13-Dec-03<br />

20-Dec-03<br />

20-Dec-03<br />

27-Dec-03<br />

27-Dec-03<br />

3-J<strong>an</strong>-04<br />

3-J<strong>an</strong>-04<br />

Figure 15. Correlation <strong>an</strong>d RMSE variations for selected days in winter 2003–2004<br />

A non-linear method w<strong>as</strong> developed to estimate SWE using SSM/I scattering Signatures <strong>an</strong>d<br />

NDVI. The study h<strong>as</strong> shown that current linear algorithms such <strong>as</strong> Goodison-Walker <strong>an</strong>d Ch<strong>an</strong>g<br />

algorithms are not sufficient for accurate estimations of SWE. In order to resolve this problem<br />

three winter se<strong>as</strong>ons were studied. SSM/I data with corresponding snow depth, <strong>an</strong>d snow water<br />

equivalent (SWE) were used to examine <strong>the</strong> sensors response to <strong>the</strong> ch<strong>an</strong>ges in snow pack<br />

properties. SSM/I response in GTVN (19V–37V), GTH (19H–37H), <strong>an</strong>d SSI (22V–85V) to snow<br />

depth or water equivalent ch<strong>an</strong>ges were <strong>an</strong>alyzed. The <strong>an</strong>alysis h<strong>as</strong> revealed that in low latitudes<br />

with shallow snow SSI h<strong>as</strong> <strong>the</strong> highest correlation with SWE. In higher latitudes GTVN <strong>an</strong>d GTH<br />

are better estimators of SWE however <strong>the</strong> slope of <strong>the</strong> relationship between <strong>the</strong> spectral signature<br />

<strong>an</strong>d SWE varies with location. It is found that <strong>the</strong> variation of <strong>the</strong> slope of this relationship is<br />

correlated with NDVI. This fact w<strong>as</strong> used to propose <strong>the</strong> new algorithm to estimate SWE using<br />

SSM/I data <strong>an</strong>d NDVI. Validation of <strong>the</strong> new algorithm h<strong>as</strong> shown that it allows reducing of <strong>the</strong><br />

error of SWE estimates by more th<strong>an</strong> 20 percent <strong>as</strong> compared to earlier linear algorithms. The<br />

<strong>an</strong>alysis of derived SWE distributions over <strong>the</strong> study area h<strong>as</strong> revealed a consistent improvement<br />

of retrieval accuracy of SWE with <strong>the</strong> new algorithm.<br />

10-J<strong>an</strong>-04<br />

10-J<strong>an</strong>-04<br />

119<br />

17-J<strong>an</strong>-04<br />

17-J<strong>an</strong>-04<br />

24-J<strong>an</strong>-04<br />

24-J<strong>an</strong>-04<br />

31-J<strong>an</strong>-04<br />

31-J<strong>an</strong>-04<br />

7-Feb-04<br />

7-Feb-04<br />

14-Feb-04<br />

14-Feb-04<br />

21-Feb-04<br />

21-Feb-04


ACKNOWLEDMENTS<br />

The authors express <strong>the</strong>ir gratitude to Dr C. Derksen of Meteorological Service of C<strong>an</strong>ada<br />

(MSC) <strong>an</strong>d Dr A. Frei of <strong>the</strong> City University of New York. The SNODAS dat<strong>as</strong>ets produced by<br />

NOHRSC, were obtained through NSIDC. Th<strong>an</strong>ks to Dr T. Carrol at NOHRSC <strong>an</strong>d L. Ballagh at<br />

NOAA at NSIDC, University of Colorado.<br />

REFRENCES<br />

Aschbacher, J. (1989). L<strong>an</strong>d surface studies <strong>an</strong>d atmospheric effects by satellite microwave<br />

radiometry, university of Innsbruk.<br />

Ch<strong>an</strong>g A.T.C., J. L. F., <strong>an</strong>d D.K.Hall. (1987). "Nimbus-7 SMMR derived global snow cover<br />

parameters." Annals Glaciology 9: 39–44.<br />

Derksen, R. B., A. Walker (2004). "Merging Conventional (1915–92) <strong>an</strong>d P<strong>as</strong>sive Microwave<br />

(1978–2002) Estimates of <strong>Snow</strong> Extent <strong>an</strong>d Water Equivalent over Central North America."<br />

Journal of Hydrometeorology 5: 850–861.<br />

Derksen C., A. W., B.Goodison (2005). "Evaluation of p<strong>as</strong>sive microwave snow water equivalent<br />

retrievals across <strong>the</strong> boreal forest/tundra tr<strong>an</strong>sition of western C<strong>an</strong>ada." Remote Sensing of<br />

Environment 96: 315–327.<br />

De Seve D., B. M., Fortin J. P. , <strong>an</strong>d Walker A., (1997). "Preliminary <strong>an</strong>alysis of snow microwave<br />

radiometry using <strong>the</strong> SSM/I p<strong>as</strong>sive-microwave data: The c<strong>as</strong>e of La Gr<strong>an</strong>de River watershed<br />

(Quebec)." Annals of Geology 25.<br />

Foster J. L., H. D. K., Ch<strong>an</strong>g A.T., R<strong>an</strong>go A., Wergin W., <strong>an</strong>d Erbe E., (1999). "Effect of <strong>Snow</strong><br />

Crystal Shape on <strong>the</strong> Scattering of P<strong>as</strong>sive Microwave Radiation." IEEE Tr<strong>an</strong>saction on<br />

Geosciences <strong>an</strong>d Remote sensing 37(2).<br />

Grody N., B. N. (1996). "Global Identification of <strong>Snow</strong>cover Using SSM/I Me<strong>as</strong>urements." IEEE<br />

Tr<strong>an</strong>saction on Geosciences <strong>an</strong>d Remote sensing 34(1).<br />

Hallikainen, M. T. (1984). "Retrieval of snow water equivalent from Nimbus-7 SSMR data: effect<br />

of l<strong>an</strong>d cover categories <strong>an</strong>d wea<strong>the</strong>r conditions." IEEE Oce<strong>an</strong>ic Eng 9(5): 372–376.<br />

Kelly R.E., A. T. C., L.Ts<strong>an</strong>g, J.L.Foster (2003). "A prototype AMSR-E Global <strong>Snow</strong> Area <strong>an</strong>d<br />

<strong>Snow</strong> depth Algorithm." IEEE Tr<strong>an</strong>saction on Geosciences <strong>an</strong>d Remote Sensing, 41(2): 230–<br />

242.<br />

Kelly R.E.J. , A. T. C. C., J.L. Foster <strong>an</strong>d D.K. Hall (2001). Development of a p<strong>as</strong>sive microwave<br />

global snow monitoring algorithm for <strong>the</strong> Adv<strong>an</strong>ced Microwave Sc<strong>an</strong>ning Radiometer-EOS.<br />

Kurvonen, L. (1994). Radiometer me<strong>as</strong>urements of snow in Sod<strong>an</strong>kyla, Helsinki University of<br />

Technology, Laboratory of Space Technology.<br />

Pulli<strong>an</strong>en, J. T., Gr<strong>an</strong>dell J., <strong>an</strong>d Hallikainen M (1999). "HUT snow emission model <strong>an</strong>d its<br />

applicability to snow water equivalent retrieval." IEEE Tr<strong>an</strong>saction on Geosciences <strong>an</strong>d<br />

Remote sensing 37(3): 1378–1390.<br />

Tedesco M., P. J., Takala M., Hallikainen M., <strong>an</strong>d Pampaloni P (2004). "Artificial neural networkb<strong>as</strong>ed<br />

techniques for <strong>the</strong> retrieval of SWE <strong>an</strong>d snow depth from SSM/I data." Remote sensing<br />

of Environment 90.<br />

Walker, G. B. E. a. A. E. (1995). C<strong>an</strong>adi<strong>an</strong> development <strong>an</strong>d use of snow cover information from<br />

p<strong>as</strong>sive microwave satellite data. P<strong>as</strong>sive microwave remote sensing of l<strong>an</strong>d–atmosphere<br />

interactions. B. J. Choudhury, Y. H. Kerr, E. G. Njoku <strong>an</strong>d P. Pampaloni. The Ne<strong>the</strong>rl<strong>an</strong>ds,<br />

VSP BV 245–262.<br />

120


63 nd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

On <strong>the</strong> Evaluation of <strong>Snow</strong> Water Equivalent Estimates over <strong>the</strong><br />

Terrestrial Arctic Drainage B<strong>as</strong>in<br />

ABSTRACT<br />

Michael A. Rawlins 1 , Mark Fahnestock 1,2 ,SteveFrolking 1,2 ,<br />

Charles J. Vörösmarty 1,2<br />

Comparisons between snow water equivalent (SWE) <strong>an</strong>d river discharge estimates are import<strong>an</strong>t<br />

in evaluating <strong>the</strong> SWE fields <strong>an</strong>d to our underst<strong>an</strong>ding of linkages in <strong>the</strong> freshwater cycle. In<br />

this study we compared SWE drawn from l<strong>an</strong>d surface models <strong>an</strong>d remote sensing observations<br />

with me<strong>as</strong>ured river discharge (Q) across 179 arctic river b<strong>as</strong>ins. Over <strong>the</strong> period 1988-2000,<br />

b<strong>as</strong>in-averaged SWE prior to snowmelt explains a relatively small (yet statistically signific<strong>an</strong>t)<br />

fraction of inter<strong>an</strong>nual variability in spring (April–June) Q, <strong>as</strong> <strong>as</strong>sessed using <strong>the</strong> coefficient of<br />

determination (R 2 ). Over all river b<strong>as</strong>ins, me<strong>an</strong> R 2 s vary from 0.20 to 0.28, with <strong>the</strong> best agreement<br />

noted for SWE drawn from simulations of <strong>the</strong> P<strong>an</strong>-Arctic Water Bal<strong>an</strong>ce Model (PWBM) that<br />

are forced with data from <strong>the</strong> National Center for Environmental Prediction / National Center<br />

for Atmospheric Research (NCEP-NCAR) Re<strong>an</strong>alysis. Variability <strong>an</strong>d magnitude in SWE derived<br />

from Special Sensor Microwave Imager (SSM/I) data are considerably lower th<strong>an</strong> <strong>the</strong> variability<br />

<strong>an</strong>d magnitude in SWE drawn from <strong>the</strong> l<strong>an</strong>d surface models, <strong>an</strong>d generally poor agreement is noted<br />

between SSM/I SWE <strong>an</strong>d spring Q. We find that <strong>the</strong> SWE vs. Q comparisons are no better when<br />

alternate temporal integrations—using <strong>an</strong> estimate of <strong>the</strong> timing in b<strong>as</strong>in thaw—are used to define<br />

pre-melt SWE <strong>an</strong>d spring Q. Thus, a majority of <strong>the</strong> variability in spring discharge must arise<br />

from factors o<strong>the</strong>r th<strong>an</strong> b<strong>as</strong>in snowpack water storage. This study suggests that SWE estimated<br />

from remote sensing observations or general circulation models (GCMs) c<strong>an</strong> be evaluated effectively<br />

using monthly discharge data or SWE from a hydrological model. The relatively small fraction of Q<br />

variability explained by b<strong>as</strong>in SWE warr<strong>an</strong>ts fur<strong>the</strong>r investigation using daily discharge observations<br />

to more accurately define <strong>the</strong> snowmelt contribution to river runoff.<br />

Keywords: SWE; River Discharge; Remote Sensing; SSM/I<br />

1<br />

Water Systems Analysis Group, Institute for <strong>the</strong> Study of Earth, Oce<strong>an</strong>s, <strong>an</strong>d Space, University of New<br />

Hampshire, Durham, NH 03824 (USA)<br />

2 Department of Earth Sciences, University of New Hampshire, Durham, NH 03824 (USA)<br />

121


INTRODUCTION<br />

Winter snow storage <strong>an</strong>d its subsequent melt are integral components of <strong>the</strong> climate system.<br />

Much remains unknown regarding <strong>the</strong> magnitudes <strong>an</strong>d inter<strong>an</strong>nual variations of this key feature of <strong>the</strong><br />

arctic water <strong>an</strong>d energy cycles. Across large parts of <strong>the</strong> terrestrial Arctic direct snow observations<br />

are unavailable, <strong>an</strong>d this lack of information limits our ability to monitor a region which is exhibiting<br />

signs of ch<strong>an</strong>ge (Peterson et al., 2002; Vörösmarty et al., 2001). Yet, amid declines in P<strong>an</strong>-Arctic<br />

station observations (Shiklom<strong>an</strong>ov et al., 2002), a growing number of models <strong>an</strong>d remote sensing<br />

data are being brought to bear for studying <strong>the</strong> arctic hydrological cycle. Retrospective <strong>an</strong>alysis<br />

or “re<strong>an</strong>alysis” of <strong>the</strong> atmospheric state such <strong>as</strong> <strong>the</strong> National Centers for Environmental Prediction<br />

(NCEP) <strong>an</strong>d <strong>the</strong> National Center for Atmospheric Research (NCAR) Re<strong>an</strong>alysis Project (Kalnay<br />

et al., 1996) provide benchmark, temporally-consistent data sets for water cycle studies. Remote<br />

sensing techniques offer <strong>the</strong> potential for more complete coverage at regional scales (Derksen et al.,<br />

2003; McDonald et al., 2004).<br />

High quality estimates of snow storage <strong>an</strong>d melt c<strong>an</strong> be used to validate <strong>the</strong> behavior of hydrological<br />

models <strong>an</strong>d GCMs, which have difficulty reproducing solid precipitation dynamics (Waliser et al.,<br />

2005). <strong>Snow</strong> cover <strong>an</strong>d snow water equivalent (SWE) estimates are also needed for climate ch<strong>an</strong>ge<br />

<strong>an</strong>alysis <strong>an</strong>d flood prediction studies. Approximately 8000 (daily) snow depth observations were<br />

<strong>an</strong>alyzed to create monthly snow depth <strong>an</strong>d SWE climatologies for North America (Brown et al.,<br />

2003) for use in evaluating Atmospheric Model Intercomparison Project II (AMIP II) snow cover<br />

simulations. Comparisons of continental-scale snow parameters with river discharge time series are<br />

useful to improving our underst<strong>an</strong>ding of <strong>the</strong> role of snow accumulation <strong>an</strong>d melt in runoff generation<br />

processes. Y<strong>an</strong>g et al. (2002), examining <strong>the</strong> snow-discharge relationship, noted a weak correlation<br />

(R = 0.14 to 0.27) between winter precipitation (a proxy for snow thickness) <strong>an</strong>d streamflow between<br />

May <strong>an</strong>d July across <strong>the</strong> Lena river b<strong>as</strong>in in Siberia. Across <strong>the</strong> Ob b<strong>as</strong>in, winter snow depth derived<br />

from Special Sensor Microwave Imager (SSM/I) agrees well with runoff in June (R=0.61), with lower<br />

correlations for comparisons using May or July discharge (Grippa et al., 2005). Strong links have<br />

been reported between end-of-winter SWE <strong>an</strong>d spring/early summer river discharge in <strong>the</strong> Churchill<br />

River <strong>an</strong>d Chesterfield Inlet B<strong>as</strong>ins of Nor<strong>the</strong>rn C<strong>an</strong>ada (Déry et al., 2005). Frappart et al. (2006)<br />

recently compared snow m<strong>as</strong>s derived from SSM/I data <strong>an</strong>d three l<strong>an</strong>d surface models with snow<br />

solutions derived from GRACE geoid data. GRACE (Gravity Recovery <strong>an</strong>d Climate Experiment)<br />

is a geodesy mission to qu<strong>an</strong>tify <strong>the</strong> terrestrial hydrological cycle through me<strong>as</strong>urements of Earth’s<br />

gravity field. They found that GRACE solutions correlate well with <strong>the</strong> high-latitude zones of strong<br />

accumulation of snow at <strong>the</strong> se<strong>as</strong>onal scale.<br />

To better underst<strong>an</strong>d <strong>the</strong> agreement between SWE <strong>an</strong>d observed river discharge, we examine<br />

comparisons of <strong>the</strong>ir year-to-year ch<strong>an</strong>ges across 179 river b<strong>as</strong>ins over <strong>the</strong> period 1988–2000. Gridded<br />

SWE estimates across <strong>the</strong> P<strong>an</strong>-Arctic drainage b<strong>as</strong>in are taken from both satellite microwave data<br />

<strong>an</strong>d l<strong>an</strong>d surface model estimates. The objective of our study is to evaluate several common SWE<br />

data sets using monthly discharge for watersheds across <strong>the</strong> terrestrial arctic b<strong>as</strong>in.<br />

DATA AND METHODS<br />

Spatial, gridded estimates of monthly SWE <strong>an</strong>d discharge for river b<strong>as</strong>ins across <strong>the</strong> P<strong>an</strong>-Arctic<br />

were <strong>an</strong>alyzed for <strong>the</strong> period 1988–2000. Monthly SWE is drawn from <strong>the</strong> <strong>an</strong>alysis scheme described<br />

by Brown et al. (2003) <strong>an</strong>d archived at <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Cryospheric Information Network (CCIN,<br />

122


http://www.ccin.ca); from simulations using <strong>the</strong> P<strong>an</strong>-Arctic Water Bal<strong>an</strong>ce Model (PWBM) (Rawlins<br />

et al., 2003); from snowpack water storage in <strong>the</strong> L<strong>an</strong>d Dynamics Model (LaD) (Milly <strong>an</strong>d<br />

Shmakin, 2002); <strong>an</strong>d from SSM/I brightness temperatures (Armstrong <strong>an</strong>d Brodzik, 1995; Armstrong<br />

et al., 2006). PWBM uses gridded fields of pl<strong>an</strong>t rooting depth, soil characteristics (texture,<br />

org<strong>an</strong>ic content), vegetation, <strong>an</strong>d is driven with daily time series of climate (precipitation (P )<strong>an</strong>dair<br />

temperature (T )) variables. Monthly PWBM SWE is obtained from model runs using P <strong>an</strong>d T from<br />

3 different sources (i) ERA-40 (ECMWF, 2002), (ii) NCEP-NCAR Re<strong>an</strong>alysis (NNR) (Kalnay et al.,<br />

1996), (iii) Willmott-Matsuura (WM) (Willmott <strong>an</strong>d Matsuura, 2001). We refer to <strong>the</strong>se SWE estimates<br />

<strong>as</strong> PWBM/ERA-40, PWBM/NNR, <strong>an</strong>d PWBM/WM, respectively. The NNR P data have<br />

been adjusted b<strong>as</strong>ed on a statistical downscaling approach (Serreze et al., 2003). Implemented in <strong>an</strong><br />

effort to minimize bi<strong>as</strong>es through <strong>the</strong> use of observed P data, this method involved (1) interpolation<br />

of observed monthly totals from available station records with bi<strong>as</strong> adjustments <strong>an</strong>d (2) disaggregation<br />

of <strong>the</strong> monthly totals to daily totals, making use of daily P forec<strong>as</strong>ts from <strong>the</strong> NCEP/NCAR<br />

Re<strong>an</strong>alysis. PWBM simulations were performed on <strong>the</strong> 25×25 km Equal Area Scalable Earth Grid<br />

(EASE-Grid) (Brodzik <strong>an</strong>d Knowles, 2002). The LaD h<strong>as</strong> previously been found to explain half of<br />

<strong>the</strong> inter<strong>an</strong>nual vari<strong>an</strong>ce of <strong>the</strong> runoff/precipitation ratio of 44 major river b<strong>as</strong>ins (Shmakin et al.,<br />

2002). In a study of SWE derived from GRACE <strong>an</strong>d from three LSMs, LaD estimates most closely<br />

matched those from GRACE, with a good correspondence at se<strong>as</strong>onal time scales (Frappart et al.,<br />

2006). Our <strong>an</strong>alysis also includes SWE (0.25 ◦ resolution for years 1988-1997) from <strong>the</strong> <strong>an</strong>alysis<br />

scheme described by Brown et al. (2003) <strong>an</strong>d archived at <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Cryospheric Information<br />

Network (CCIN, http://www.ccin.ca). LaD SWE at 1 ◦ resolution w<strong>as</strong> mapped to <strong>the</strong> EASE-Grid<br />

using inverse-dist<strong>an</strong>ce weighted interpolation, while CCIN SWE w<strong>as</strong> aggregated to <strong>the</strong> EASE-Grid<br />

using <strong>the</strong> average of all 0.25 ◦ grids falling within each EASE-Grid cell. The PWBM-, LaD-, <strong>an</strong>d<br />

SSM/I-derived SWE estimates are P<strong>an</strong>-Arctic in nature, ie. defined at all 39,926 EASE-Grid cells<br />

encomp<strong>as</strong>sing <strong>the</strong> P<strong>an</strong>-Arctic drainage b<strong>as</strong>in (Figure 1). CCIN SWE are available across EASE-Grid<br />

cells over North America only.<br />

P<strong>as</strong>sive microwave radi<strong>an</strong>ces from SSM/I—aboard <strong>the</strong> Defense Meteorological Satellite Program<br />

satellite series since 1987—have been used to produce maps of SWE across large regions (Armstrong<br />

<strong>an</strong>d Brodzik, 2001; Armstrong <strong>an</strong>d Brodzik, 2002; Derksen et al., 2003; Goita et al., 2003). Monthly<br />

SWE estimated from SSM/I radi<strong>an</strong>ces for <strong>the</strong> period 1988–1999 (on <strong>the</strong> EASE-Grid) were acquired<br />

from <strong>the</strong> US National <strong>Snow</strong> <strong>an</strong>d Ice Data Center (M. J. Brodzik, personal communication, March<br />

2, 2004) <strong>an</strong>d are archived under <strong>the</strong> ArcticRIMS project (http://RIMS.unh.edu). The snow depth<br />

algorithm (Armstrong <strong>an</strong>d Brodzik, 2001) is: snow depth (cm) = 1.59 * [(T19H − 6) − (T37H − 1)],<br />

where T19H is <strong>the</strong> brightness temperature at 19 GHz <strong>an</strong>d T19H is <strong>the</strong> brightness temperature at 37<br />

GHz. Water equivalent is obtained from <strong>the</strong> product of snow depth <strong>an</strong>d density.<br />

Our <strong>an</strong>alysis involves <strong>the</strong> use of what we term “pre-melt” SWE (<strong>the</strong> average of February <strong>an</strong>d<br />

March monthly SWE) <strong>an</strong>d spring total Q (<strong>the</strong> total discharge flow over <strong>the</strong> months April–June). We<br />

chose <strong>an</strong> average of two months of SWE over one month (or maximal monthly) to better represent<br />

mid-winter conditions. The SWE <strong>an</strong>d Q time series are prewhitened to remove <strong>an</strong>y trends prior<br />

to <strong>the</strong> covari<strong>an</strong>ce <strong>an</strong>alysis. The Q records are drawn from <strong>an</strong> updated version of R-ArcticNET<br />

(Lammers et al., 2001). Although SWE estimates are available for more recent years, our <strong>an</strong>alysis<br />

here ends in 2000 due to a lack of more recent river discharge data for river b<strong>as</strong>ins across Eur<strong>as</strong>ia.<br />

Alternate comparisons using SWE <strong>an</strong>d Q which vary depending thaw timing derived from SSM/I<br />

data, <strong>an</strong>d by simulated snowmelt, are described in <strong>the</strong> Results section. In this study, all SWE<br />

data sets have valid data at each grid defining <strong>the</strong> P<strong>an</strong>-Arctic drainage b<strong>as</strong>in. For each of <strong>the</strong> 179<br />

river b<strong>as</strong>ins, pre-melt SWE is <strong>the</strong>n determined <strong>as</strong> <strong>an</strong> average over all EASE-Grid cells defining <strong>the</strong><br />

123


Figure 1: P<strong>an</strong>-Arctic l<strong>an</strong>d m<strong>as</strong>s (north of 45 ◦ N, dark gray), <strong>the</strong> arctic drainage b<strong>as</strong>in (light gray),<br />

<strong>an</strong>d locations of 179 river b<strong>as</strong>ins used in <strong>the</strong> study. Dot sizes are scaled by b<strong>as</strong>in area. A total of<br />

39,926 EASE-Grid cells comprise <strong>the</strong> approximately 25 million km 2 drainage b<strong>as</strong>in. Are<strong>as</strong> for <strong>the</strong><br />

179 river b<strong>as</strong>ins r<strong>an</strong>ge from 20,000 km 2 to 486,000 km 2 .<br />

respective <strong>the</strong> b<strong>as</strong>in.<br />

Satellite-borne remote sensing at microwave wavelengths c<strong>an</strong> be used to monitor l<strong>an</strong>dscape<br />

freeze/thaw state (Ulaby et al., 1986; Way et al., 1997; Frolking et al., 1999; Kimball et al., 2001).<br />

A step edge detection scheme applied to SSM/I brightness temperatures (McDonald et al., 2004)<br />

w<strong>as</strong> used to identify <strong>the</strong> predomin<strong>an</strong>t springtime thaw tr<strong>an</strong>sition event for each EASE-Grid cell.<br />

As with SWE, we derived a b<strong>as</strong>in average date of thaw by averaging thaw event dates across <strong>the</strong><br />

b<strong>as</strong>in grid cells. <strong>Snow</strong> thaw across arctic b<strong>as</strong>ins often c<strong>an</strong> occur over a period of weeks or months.<br />

Therefore, for large watersheds, our timing estimates derived from SSM/I brightness temperatrures<br />

must be interpreted with caution. None<strong>the</strong>less <strong>the</strong>y provide a general approximation of <strong>the</strong> timing<br />

in l<strong>an</strong>dscape thaw for use in estimating pre-melt SWE <strong>an</strong>d spring Q. As <strong>an</strong> illustration, monthly<br />

river discharge, SWE, <strong>an</strong>d thaw date for <strong>the</strong> Yukon b<strong>as</strong>in are shown in Figure 2a–d).<br />

A simulated topological network (Vörösmarty et al., 2000), recently implemented at 6 minute<br />

resolution, defines river b<strong>as</strong>ins over <strong>the</strong> approximately 25 million km 2 of <strong>the</strong> P<strong>an</strong>-Arctic b<strong>as</strong>in. The<br />

degree to which SWE <strong>an</strong>d Q covary over <strong>the</strong> period 1988-2000 is evaluated using <strong>the</strong> coefficient of<br />

determination, R 2 (squared correlation). Throughout our <strong>an</strong>alysis we <strong>as</strong>sume a signific<strong>an</strong>ce level<br />

of 0.05 (5%) <strong>as</strong> <strong>the</strong> cutoff to determine whe<strong>the</strong>r a given SWE vs. Q comparison is statistically<br />

signific<strong>an</strong>t, <strong>an</strong>d not due to ch<strong>an</strong>ce. For a sample size of 13 years this correspondes to R 2 ≥ 0.22.<br />

RESULTS<br />

Inter<strong>an</strong>nual variability in b<strong>as</strong>in averaged, pre-melt SWE is compared with spring Q for 179 b<strong>as</strong>ins<br />

124


over <strong>the</strong> period 1988–2000. With <strong>the</strong> exception of SWE derived from SSM/I data, inter<strong>an</strong>nual<br />

variability in pre-melt SWE agrees well with spring Q variability across <strong>the</strong> Yukon b<strong>as</strong>in in Al<strong>as</strong>ka<br />

(Figure 2). Variability in b<strong>as</strong>in SWE from <strong>the</strong> CCIN <strong>an</strong>alysis scheme explains nearly 75% of <strong>the</strong><br />

variability in spring Q. When re<strong>an</strong>alysis data drives <strong>the</strong> PWBM (PWBM/ERA-40 or PWBM/NNR),<br />

pre-melt SWE explains well over 50% of <strong>the</strong> variability in spring Q. Across <strong>the</strong> Yukon b<strong>as</strong>in, <strong>the</strong><br />

greater SWE variability <strong>an</strong>d magnitude (among all SWE products) is noted for LaD SWE, along<br />

with a lower R 2 (Figure 2c–d). B<strong>as</strong>in averaged SWE derived from SSM/I, however, shows little<br />

inter<strong>an</strong>nual variability <strong>an</strong>d relatively low magnitude. For snow packs above 100 mm, <strong>the</strong> bi<strong>as</strong> in<br />

SWE estimated from Sc<strong>an</strong>ning Multich<strong>an</strong>nel Microwave Radiometer (SMMR) data w<strong>as</strong> shown to be<br />

linearly related to <strong>the</strong> snow pack m<strong>as</strong>s, with root-me<strong>an</strong>-square errors approaching 150 mm (Dong<br />

et al., 2005).<br />

In contr<strong>as</strong>t to <strong>the</strong> result over <strong>the</strong> Yukon b<strong>as</strong>in, strong agreements in pre-melt SWE <strong>an</strong>d spring Q<br />

variability are not noted across m<strong>an</strong>y of <strong>the</strong> study b<strong>as</strong>ins. Although R 2 s for a majority of <strong>the</strong> North<br />

Americ<strong>an</strong> b<strong>as</strong>ins are signific<strong>an</strong>t (me<strong>an</strong> values between 0.26 <strong>an</strong>d 0.36, Table 1), agreements across<br />

e<strong>as</strong>tern Eur<strong>as</strong>ia are generally low (Figure 3, Table 1). Of <strong>the</strong> comparisons involving SWE from<br />

PWBM, more th<strong>an</strong> half (130 of 231) are signific<strong>an</strong>t. Me<strong>an</strong> R 2 s from comparisons using <strong>the</strong> PWBM<br />

are comparable to those involving CCIN SWE estimates, which were developed using observed snow<br />

depth observations (Brown et al., 2003). Across all b<strong>as</strong>ins <strong>an</strong>alyzed, <strong>the</strong> highest proportion of<br />

negative correlations (very poor agreement) <strong>an</strong>d lowest overall R 2 are <strong>as</strong>sociated with SSM/I SWE.<br />

The algorithm used to produce <strong>the</strong>se estimates, like m<strong>an</strong>y of <strong>the</strong> early p<strong>as</strong>sive-microwave SWE<br />

algorithms, tends to underestimate SWE in forested regions. Models which account for <strong>the</strong> differing<br />

influences on <strong>the</strong> microwave signature have shown promise in reducing errors in forested regions<br />

(Goita et al., 2003). The best agreements involving SSM/I SWE are found across <strong>the</strong> prairies of<br />

south-central C<strong>an</strong>ada. This is expected, <strong>as</strong> <strong>the</strong> SSM/I SWE algorithm w<strong>as</strong> developed for application<br />

across <strong>the</strong> non-forested prairie provinces of C<strong>an</strong>ada.<br />

Comparisons using SWE from <strong>the</strong> PWBM simulations (PWBM/ERA-40, PWBM/NNR, <strong>an</strong>d<br />

PWBM/WM), produce similar R 2 values across each region, with me<strong>an</strong> value by region r<strong>an</strong>ging<br />

from 0.15 to 0.36. Given that water budget models like PWBM are most sensitive to time-varying<br />

climatic inputs (Rawlins et al., 2003), small differences in R 2 among <strong>the</strong>se SWE estimates suggest<br />

similar spatial <strong>an</strong>d temporal variability among <strong>the</strong> underlying precipitation data. B<strong>as</strong>in R 2 s obtained<br />

from comparisons using LaD SWE are comparable with those from <strong>the</strong> comparisons using PWBM<br />

SWE across North America, while lower correlations are noted for Eur<strong>as</strong>ia. Me<strong>an</strong> R 2 sarehigher<br />

across e<strong>as</strong>tern Eur<strong>as</strong>ia (e<strong>as</strong>t of longitude 90 ◦ E) <strong>as</strong> compared with western Eur<strong>as</strong>ia. The better<br />

agreement across Siberia is likely attributable to <strong>the</strong> higher fraction of precipitation which falls <strong>as</strong><br />

snow <strong>an</strong>d <strong>the</strong> higher discharge/precipitation ratios across <strong>the</strong> colder e<strong>as</strong>t. When <strong>the</strong> PWBM is<br />

driven with precipitation data from a new gauge-corrected archive for <strong>the</strong> former USSR (“Daily <strong>an</strong>d<br />

Sub-daily Precipitation for <strong>the</strong> Former USSR”) (National Climatic Data Center, 2005), b<strong>as</strong>in R 2 s<br />

are generally no higher (figure not shown). This suggests that precipitation-gauge undercatch is not<br />

a signific<strong>an</strong>t influence on <strong>the</strong> computed SWE vs. Q agreements.<br />

<strong>Snow</strong>melt <strong>an</strong>d subsequent rises in river Q begins in sou<strong>the</strong>rly regions of <strong>the</strong> terrestrial arctic <strong>an</strong>d<br />

progresses northward each spring. Comparisons of winter SWE storage <strong>an</strong>d Q over a fixed interval<br />

(e.g. April–June) are complicated when inputs from rainfall are signific<strong>an</strong>t, or a large fraction of <strong>the</strong><br />

snowmelt occurs outside of <strong>the</strong> April–June period. A more me<strong>an</strong>ingful comparison of SWE <strong>an</strong>d river<br />

Q wouldberestrictedtothatfractionofQ which is attributable to <strong>the</strong> melting of snow. For example,<br />

simulated spring Q from PWBM—driven by ERA-40 data—explains a much higher proportion of<br />

observed spring Q th<strong>an</strong> does <strong>the</strong> pre-melt SWE across <strong>the</strong> study b<strong>as</strong>ins (Figure 4, Table 1). The<br />

125


Figure 2: Monthly total SWE (mm) <strong>an</strong>d me<strong>an</strong> discharge (Q, mmday −1 ) across <strong>the</strong> Yukon b<strong>as</strong>in for<br />

(a) 1992, a year with relatively high SWE <strong>an</strong>d Q, <strong>an</strong>d (b) 1998, a year with low SWE <strong>an</strong>d Q totals.<br />

Vertical bars show SWE— in this c<strong>as</strong>e from PWBM driven with ERA-40 data (PWBM/ERA-40).<br />

February <strong>an</strong>d March SWE values (gray bars here) are averaged to give “pre-melt” SWE in this<br />

study. April–June SWE are depicted by white bars. Spring Q (monthly values indicated by dots at<br />

middle of month) is <strong>the</strong> integration of <strong>the</strong> monthly Q for April through June (hatched area), <strong>an</strong>d is<br />

used for comparisons with <strong>the</strong> pre-melt SWE. A “thaw date” (marked Thaw) estimated from SSM/I<br />

data are used in alternate Q integrations. (c) Scatterplot of pre-melt SWE from each data set, for<br />

years 1988-2000. The best fit line b<strong>as</strong>ed on linear le<strong>as</strong>t squares regression is shown. (d) Time series<br />

of SWE <strong>an</strong>d Q. Statistics (R 2 <strong>an</strong>d <strong>as</strong>sociated p-value) for each covari<strong>an</strong>ce comparison are shown in<br />

paren<strong>the</strong>sis.<br />

126


Figure 3: Explained vari<strong>an</strong>ce (R 2 ) for pre-melt SWE <strong>an</strong>d spring Q comparisons (1988–2000) at <strong>the</strong><br />

179 river b<strong>as</strong>ins <strong>an</strong>d for <strong>the</strong> 6 SWE products. SWE is taken from <strong>the</strong> CCIN SWE <strong>an</strong>alysis; PWBM<br />

simulations driven by ERA-40, NNR <strong>an</strong>d WM; LaD model; <strong>an</strong>d SSM/I data. The ’X’s mark b<strong>as</strong>ins<br />

with a negative correlation. Average R 2 values across all b<strong>as</strong>ins, North America (NA), western<br />

Eur<strong>as</strong>ia (WE), <strong>an</strong>d e<strong>as</strong>tern Eur<strong>as</strong>ia (EE) are shown in Table 1. The vertical line in colorbar is level<br />

(R 2 = 0.22) at which R 2 is signific<strong>an</strong>t at 5% level. P-values <strong>as</strong>sociated with each R 2 interval are<br />

shown above <strong>the</strong> colorbar. Note that p < 0.01 for all R 2 ≥ 0.40.<br />

127


Feb–Mar Min, Max, <strong>an</strong>d Me<strong>an</strong> Coefficient of Variation, R2 SWE Data SWE (mm) %neg. All North Am. W. Eur<strong>as</strong>ia E. Eur<strong>as</strong>ia<br />

CCIN N/A 15.6 N/A 0.00, 0.87, 0.35 N/A N/A<br />

PWBM/ERA-40 103 8.5 0.00, 0.91, 0.28 0.00, 0.91, 0.36 0.00, 0.45, 0.15 0.00, 0.66, 0.27<br />

PWBM/NNR 109 12.6 0.00, 0.87, 0.25 0.00, 0.87, 0.33 0.00, 0.56, 0.15 0.00, 0.71, 0.23<br />

PWBM/WM 109 11.9 0.00, 0.91, 0.26 0.00, 0.91, 0.33 0.00, 0.75, 0.22 0.00, 0.53, 0.17<br />

LaD 144 20.1 0.00, 0.83, 0.24 0.00, 0.83, 0.33 0.00, 0.49, 0.12 0.00, 0.69, 0.16<br />

SSM/I 80 72.1 0.00, 0.76, 0.20 0.00, 0.76, 0.26 0.00, 0.40, 0.10 0.00, 0.57, 0.14<br />

SimRO 103 5.3 0.00, 0.92, 0.46 0.01, 0.91, 0.44 0.00, 0.79, 0.35 0.05, 0.92, 0.57<br />

PWBM/ERA-40a 103 10.0 0.00, 0.80, 0.27 0.00, 0.80, 0.33 0.00, 0.61, 0.22 0.00, 0.64, 0.22<br />

PWBM/ERA-40b 103 18.7 0.00, 0.93, 0.34 0.00, 0.93, 0.37 0.00, 0.86, 0.38 0.00, 0.70, 0.26<br />

PWBM/ERA-40c 103 10.0 0.00, 0.80, 0.27 0.80, 0.00, 0.33 0.61, 0.00, 0.22 0.00, 0.64, 0.22<br />

PWBM/ERA-40d 103 10.0 0.00, 0.80, 0.27 0.80, 0.00, 0.33 0.61, 0.00, 0.22 0.00, 0.64, 0.22<br />

PWBM/ERA-40e 103 22.8 0.00, 0.76, 0.25 0.00, 0.76, 0.35 0.00, 0.28, 0.12 0.00, 0.58, 0.19<br />

128<br />

Table 1: Me<strong>an</strong> February-March SWE (mm), percent negative correlations, <strong>an</strong>d minimum, maximum, <strong>an</strong>d me<strong>an</strong> coefficient<br />

of determination (R2 ) from <strong>the</strong> pre-melt SWE <strong>an</strong>d spring Q comparisons. SWE is taken from data sets described in Data<br />

<strong>an</strong>d Methods <strong>an</strong>d shown in Figure 3. Percentage of negative correlations, <strong>an</strong>d me<strong>an</strong> explained vari<strong>an</strong>ce is also tabulated for<br />

simulated spring Q vs. observed spring Q (row SimRO), where simulated Q is from PWBM/ERA-40 model simulation. Me<strong>an</strong><br />

February-March SWE is averaged across <strong>the</strong> terrestrial arctic b<strong>as</strong>in, excluding Greenl<strong>an</strong>d. Individual R2 values for each study<br />

b<strong>as</strong>in (shown in Figure 3) are averaged (excluding negative correlations) over all 179 river b<strong>as</strong>ins (All), <strong>an</strong>d <strong>the</strong> b<strong>as</strong>ins across<br />

North America, western Eur<strong>as</strong>ia, <strong>an</strong>d e<strong>as</strong>tern Eur<strong>as</strong>ia, with <strong>the</strong> latter two separated by <strong>the</strong> 90◦E meridi<strong>an</strong>. Me<strong>an</strong> R2 for CCIN<br />

SWE are determined for North Americ<strong>an</strong> sector only. PWBM/ERA-40a−e represent <strong>the</strong> alternate comparisons, defined in<br />

Results section.


correspondence between simulated <strong>an</strong>d observed spring Q suggests <strong>the</strong> model—to some degree—is<br />

accounting for processes connecting <strong>the</strong> snowpack <strong>an</strong>d spring river flow, e.g. sublimation, rainfall,<br />

<strong>an</strong>d soil infiltration.<br />

To better underst<strong>an</strong>d <strong>the</strong> covari<strong>an</strong>ce between SWE <strong>an</strong>d Q, alternate comparisons were made<br />

using PWBM/ERA-40 monthly SWE <strong>an</strong>d <strong>an</strong> estimate of when thaw is <strong>as</strong>sumed to have occurred.<br />

The month of thaw (TM,withTM− 1, <strong>an</strong>d TM+ 1 indicating <strong>the</strong> month preceding <strong>an</strong>d postceding<br />

<strong>the</strong> thaw month, respectively) w<strong>as</strong> determined with a step edge detection scheme applied to SSM/I<br />

brightness temperatures (McDonald et al., 2004). Then, SWETM becomes monthly b<strong>as</strong>in SWE<br />

during TM (or TM− 1), <strong>an</strong>d QTM is discharge in month TM. These alternate comparisons (across<br />

all 179 b<strong>as</strong>ins) are defined (a) SWETM vs. spring Q, (b)SWETM-1 vs. spring Q, (c)SWETM-1 vs.<br />

QTM+1, (d)SWETM-1 vs. QTM+1,2. R 2 s are highest for alternate comparison (b), which compared<br />

SWE in <strong>the</strong> month before thaw (TM − 1) with spring (April–June) Q (Table 1). Yet, despite <strong>the</strong><br />

fact that <strong>the</strong> me<strong>an</strong> R 2 across western Eur<strong>as</strong>ia improves from 0.15 (using default PWBM/ERA-40)<br />

to 0.38 (alternate comparison b), little difference is noted with <strong>the</strong> remi<strong>an</strong>ing alternate comparisons<br />

<strong>an</strong>d o<strong>the</strong>r regions.<br />

L<strong>as</strong>tly, we scaled spring Q using a factor S, whereS = PWBM monthly snow melt–runoff ratio,<br />

with 0 < S < 1. Then, snowmelt Q each month is Qs =Q·S. Each occurrence of QS w<strong>as</strong> <strong>the</strong>n<br />

summed resulting in a total QS each spring, for each b<strong>as</strong>in. Using QS in place of <strong>the</strong> default Q (<strong>an</strong>d<br />

PWBM/ERA-40 SWE), we note a decre<strong>as</strong>e in agreement across e<strong>as</strong>tern Eur<strong>as</strong>ia, with no ch<strong>an</strong>ge<br />

across most of <strong>the</strong> domain. And although SWE from simulations with ERA-40, in general, explains<br />

more th<strong>an</strong> a third of <strong>the</strong> variation in Q, a large proportion of <strong>the</strong> inter<strong>an</strong>nual variability is not<br />

due to SWE variability. When considering <strong>the</strong>se results, it is interesting to note that Lammers et<br />

al. (2006) recently found that <strong>an</strong>nual simulated discharge across Al<strong>as</strong>ka (drawn from three separate<br />

models) w<strong>as</strong> in poor agreement with observed discharge data between 1980–2001. Better agreements<br />

across northwestern North America, e<strong>as</strong>tern Eur<strong>as</strong>ia (EE in Figure 3), <strong>an</strong>d parts of western Eur<strong>as</strong>ia<br />

(WE) in this study are attributable to relatively higher snowfall rates <strong>an</strong>d a greater inter<strong>an</strong>nual<br />

variability in spring discharge (Figure 4b). Conversely, <strong>the</strong> region of e<strong>as</strong>tern Eur<strong>as</strong>ia with numerous<br />

negative correlations is characterized by low spring discharge variability. Delays in snowmelt water<br />

reaching river systems, which c<strong>an</strong> be signific<strong>an</strong>t (Hinzm<strong>an</strong> <strong>an</strong>d K<strong>an</strong>e, 1991), are likely <strong>an</strong> additional<br />

influence on <strong>the</strong>se reported correlations. For large arctic b<strong>as</strong>ins, comparisons between snow storage<br />

<strong>an</strong>d discharge volume are complicated by <strong>the</strong> large temporal variation in b<strong>as</strong>in thaw <strong>an</strong>d <strong>the</strong> delays<br />

in snowmelt water reaching <strong>the</strong> gauge. More me<strong>an</strong>ingful comparisons between spatial SWE <strong>an</strong>d<br />

river discharge are possible through <strong>the</strong> use of hydrograph separation to partition discharge into<br />

overl<strong>an</strong>d <strong>an</strong>d b<strong>as</strong>eflow components. This, however, requires <strong>the</strong> use of daily discharge data which<br />

are more limited for <strong>the</strong> P<strong>an</strong>-Arctic region.<br />

CONCLUSIONS<br />

In our comparisons of inter<strong>an</strong>nual variations in pre-melt SWE <strong>an</strong>d spring Q, R 2 values are<br />

highest (me<strong>an</strong> of 0.25 to 0.28 over all b<strong>as</strong>ins) when PWBM is driven by ERA-40, NNR or WM<br />

climate data. Similar agreements are noted when SWE from <strong>the</strong> observed data <strong>an</strong>alysis scheme are<br />

used, which suggests that <strong>the</strong> hydrological model is capturing <strong>as</strong> much variability in <strong>the</strong> spring flow<br />

<strong>as</strong> does <strong>the</strong> observed SWE scheme. Average R 2 determined from <strong>the</strong> SSM/I SWE <strong>an</strong>d spring Q<br />

comparisons are generally low, <strong>an</strong>d a sizable majority (over 72%) of <strong>the</strong>se correlations are negative.<br />

The low variability <strong>an</strong>d magnitude is likely related to saturation of <strong>the</strong> SSM/I algorithm at high SWE<br />

129


Figure 4: (a) R 2 for PWBM simulated spring total Q vs. observed spring total Q. The vertical<br />

line in colorbar is level at which R 2 is signific<strong>an</strong>t. Minimum, maximum, <strong>an</strong>d me<strong>an</strong> R 2 sacrossall<br />

b<strong>as</strong>ins, North America (NA), western Eur<strong>as</strong>ia (WE), <strong>an</strong>d e<strong>as</strong>tern Eur<strong>as</strong>ia (EE) are shown in Table<br />

1. The ’X’s mark b<strong>as</strong>ins with a negative correlation. (b) St<strong>an</strong>dard deviation of spring (April–June)<br />

discharge for <strong>the</strong> period 1988-2000.<br />

130


values. Continued development of new regional schemes which account for microwave emmission<br />

from forests should improve large-scale SWE estimates. Poor agreements among all SWE products<br />

—particularly across parts of western Eur<strong>as</strong>ia—are noted in are<strong>as</strong> with low discharge variability.<br />

Pre-screening to eliminate b<strong>as</strong>ins with low flow or insufficient variability would likely improve <strong>the</strong><br />

SWE vs. Q agreements.<br />

Results of <strong>the</strong> covari<strong>an</strong>ce <strong>an</strong>alysis using alternate temporal integrations to derive pre-melt SWE<br />

(or Q) suggest that our choice of a fixed interval for spring, ie. April–June, is not <strong>the</strong> primary cause<br />

of <strong>the</strong> relatively low R 2 s. Fur<strong>the</strong>rmore, we conclude that much of <strong>the</strong> inter<strong>an</strong>nual variability in river<br />

discharge must be influenced by factors o<strong>the</strong>r th<strong>an</strong> b<strong>as</strong>in SWE storage variations. The unexplained<br />

variability is likely due to a combination of effects from physical processes (sublimation, infiltration)<br />

<strong>an</strong>d errors in spatial SWE. Relatively good agreement between simulated <strong>an</strong>d observed spring Q<br />

suggests that hydrological models c<strong>an</strong> be useful in underst<strong>an</strong>ding <strong>the</strong> SWE-to-Q linkages. Our<br />

results provide a benchmark of <strong>the</strong> relationship between <strong>the</strong> solid precipitation input <strong>an</strong>d spring<br />

discharge flux, <strong>an</strong>d demonstrate that hydrological models driven with re<strong>an</strong>alysis data c<strong>an</strong> provide<br />

SWE estimates sufficient for use in validation of remote-sensing <strong>an</strong>d GCM SWE fields. Additional<br />

studies using daily discharge data to better qu<strong>an</strong>tify snowmelt runoff will fur<strong>the</strong>r facilitate SWE<br />

product evaluations <strong>an</strong>d <strong>the</strong> underst<strong>an</strong>ding of linkages in arctic hydrological system.<br />

ACKNOWLEDGEMENTS<br />

The authors th<strong>an</strong>k Richard Lammers, Ernst Linder, <strong>an</strong>d Dominik Wisser (University of New<br />

Hampshire) for frequent useful discussions, <strong>an</strong>d Kyle McDonald (Jet Propulsion Laboratory) for <strong>the</strong><br />

thaw timing data. We also th<strong>an</strong>k Chris Milly (NOAA/GFDL) for providing <strong>the</strong> LaD SWE estimates,<br />

<strong>an</strong>d Ross Brown (Environnement C<strong>an</strong>ada) for <strong>the</strong> CCIN data. This study w<strong>as</strong> supported by <strong>the</strong><br />

NSF ARCSS program <strong>an</strong>d NSF gr<strong>an</strong>ts OPP-9910264, OPP-0230243, OPP-0094532, <strong>an</strong>d NASA gr<strong>an</strong>t<br />

NAG5-9617.<br />

REFERENCES<br />

Armstrong, R. L. <strong>an</strong>d Brodzik, M. J. 1995. An earth-gridded SSM/I data set for cryospheric studies<br />

<strong>an</strong>d global ch<strong>an</strong>ge monitoring. Adv<strong>an</strong>ces in Space Research, 16(10):155–163.<br />

Armstrong, R. L. <strong>an</strong>d Brodzik, M. J. 2001. Recent Nor<strong>the</strong>rn Hemisphere snow extent: a comparison<br />

of data derived from visible <strong>an</strong>d microwave sensors. Geophys. Res. Lett., 28(19):3673–3676.<br />

Armstrong, R. L. <strong>an</strong>d Brodzik, M. J. 2002. Hemispheric-scale comparison <strong>an</strong>d evaluation of p<strong>as</strong>sivemicrowave<br />

snow algorithms. Annals of Glaciology, 34(1):38–44.<br />

Armstrong, R. L., Knowles K. W., Brodzik M. J., Hardm<strong>an</strong> M. A. 2006. DMSP SSM/I Pathfinder<br />

daily EASE-Grid brightness temperatures. Technical report, National <strong>Snow</strong> <strong>an</strong>d Ice Center.<br />

CDROM, Boulder, CO, USA. http://nsidc.org/data/nsidc-0032.html.<br />

Brodzik, M. J. (2004). personal communication, 2 March, 2004.<br />

Brodzik, M. J. <strong>an</strong>d Knowles K. 2002. EASE-Grid: A versatile set of equal-area projections <strong>an</strong>d<br />

grids. in M. Goodchild (Ed.) Discrete Global Grids. S<strong>an</strong>ta Barbara, CA, USA: National Center<br />

for Geographic Information <strong>an</strong>d Analysis.<br />

Brown, R. D., Br<strong>as</strong>nett, B., Robinson, D. (2003). Gridded North Americ<strong>an</strong> monthly snow depth<br />

<strong>an</strong>d snow water equivalent for GCM evaluation. Atmosphere–Oce<strong>an</strong>, 41(1):1–14.<br />

131


Derksen, C., Walker, A., LeDrew, E., Goodison, B. (2003). Combining SMMR <strong>an</strong>d SSM/I data<br />

for time series <strong>an</strong>alysis of central North Americ<strong>an</strong> snow water equivalent. J. Hydrometeorol.,<br />

4(2):304–316.<br />

Déry, S. J., Sheffield, J., Wood, E. F. (2005). Connectivity between Eur<strong>as</strong>i<strong>an</strong> snow cover extent <strong>an</strong>d<br />

C<strong>an</strong>adi<strong>an</strong> snow water equivalent <strong>an</strong>d river discharge. J. Geophys. Res., 110(D23). D23106,<br />

doi:10.1029/2005JD006173.<br />

Dong, J., Walker, J. P., Houser, P. R. (2005). Factors affecting remotely sensed snow water equivalent<br />

uncertainty. Remote Sens. Environ., 97:68–82.<br />

Europe<strong>an</strong> Centre for Medium R<strong>an</strong>ge Wea<strong>the</strong>r Forec<strong>as</strong>ts (ECMWF) (2002). ERA-40 Project Report<br />

Series. 3. Workshop om Re-<strong>an</strong>alysis, 5-9 november 2001. Technical report, Europe<strong>an</strong> Centre for<br />

Medium R<strong>an</strong>ge Wea<strong>the</strong>r Forec<strong>as</strong>ts. 443 pp.<br />

Frappart, F., Ramillien, G., Bi<strong>an</strong>camaria, S., Mognard, N. M., Cazenave, A. (2006). Evolution of<br />

high-latitude snow m<strong>as</strong>s derived from <strong>the</strong> GRACE gravimetry mission (2002–2004). Geophys.<br />

Res. Lett., 33. L02501, doi:10.10292005GL024778.<br />

Frolking, S., McDonald, K. C., Kimball, J., Way, J. B., Zimmerm<strong>an</strong>n, R., Running, S. W. (1999).<br />

Using <strong>the</strong> space-borne NASA scatterometer (NSCAT) to determine <strong>the</strong> frozen <strong>an</strong>d thawed<br />

se<strong>as</strong>ons of a boreal l<strong>an</strong>dscape. J. Geophys. Res., 104(D22):27,895–27,907.<br />

Goita, K., Walker, A. E., Goodison, B. E. (2003). Algorithm development for <strong>the</strong> estimation of<br />

snow water equivalent in <strong>the</strong> boreal forest using p<strong>as</strong>sive microwave data. Int. J. Remote Sens.,<br />

24(5):1097–1102.<br />

Grippa, M., Mognard, N., To<strong>an</strong>a, T. L. (2005). Comparison between <strong>the</strong> inter<strong>an</strong>nual variability<br />

of snow parameters derived from SSM/I <strong>an</strong>d <strong>the</strong> Ob river discharge. Remote Sens. Environ.,<br />

98:35–44.<br />

Hinzm<strong>an</strong>, L. D. <strong>an</strong>d K<strong>an</strong>e, D. L. (1991). <strong>Snow</strong> hydrology of a headwater Arctic b<strong>as</strong>in 2. Conceptual<br />

<strong>an</strong>alysis <strong>an</strong>d computer modeling. Water Resour. Res., 27(6):1111–1121.<br />

Kalnay, E., K<strong>an</strong>amitsu, M., Kistler, R., Collins, W., Deaven, D., G<strong>an</strong>din, L., Iredell, M., Saha,<br />

S., White, G., Woolen, J., Zhu, Y., Chelliah, M., Ebisuzaki, W., Higgens, W., J<strong>an</strong>owiak, J.,<br />

Ropelewski, K. C., W<strong>an</strong>g, J., Leetma, A., Reynolds, R., Jenne, R., Joseph, D. (1996). The<br />

NCEP/NCAR 40-year re<strong>an</strong>alysis project. Bull. Am. Meteorol. Soc., 77:437–471.<br />

Kimball, J. S., McDonald, K. C., Keyser, A. R., Frolking, S., Running, S. W. (2001). Application<br />

of <strong>the</strong> NASA scatterometer (NSCAT) for determining <strong>the</strong> daily frozen <strong>an</strong>d nonfrozen l<strong>an</strong>dscape<br />

of Al<strong>as</strong>ka. Remote Sens. Environ., 75:113–126.<br />

Lammers, R. B., Rawlins, M., McGuire, D., Clein, J., Kimball, J., Wu, W. (2006). Water budget<br />

closure over <strong>the</strong> western arctic <strong>an</strong>d Yukon river b<strong>as</strong>in - a model inter-comparison. Earth<br />

Interactions. in preparation.<br />

Lammers, R. B., Shiklom<strong>an</strong>ov, A. I., Vörösmarty, C. J., Fekete, B. M., Peterson, B. J. (2001).<br />

Assessment of contemporary Arctic river runoff b<strong>as</strong>ed on observational discharge records. J.<br />

Geophys. Res., 106(D4):3321–3334.<br />

McDonald, K. C., Kimball, J. S., Njoku, E., Zimmerm<strong>an</strong>n, R., Zhao, M. (2004). Variability in<br />

springtime thaw in <strong>the</strong> terrestrial high latitudes: Monitoring a major control on biospheric<br />

<strong>as</strong>similation of atmospheric CO2 with spaceborne microwave remote sensing. Earth Interactions,<br />

7:1–23.<br />

Milly, P. C. D. <strong>an</strong>d Shmakin, A. B. (2002). Global modeling of l<strong>an</strong>d water <strong>an</strong>d energy bal<strong>an</strong>ces: I.<br />

The L<strong>an</strong>d Dynamics (LaD) model. J. Hydrometeorol., 3:283–299.<br />

132


National Climatic Data Center (2005). Daily <strong>an</strong>d Sub-daily Precipitation for <strong>the</strong> Former<br />

USSR. Technical report. Available from National Geophysical Data Center,<br />

http://www.ncdc.noaa.gov/oa/documentlibrary/surface-doc.html9813.<br />

Peterson, B. J., Holmes, R. M., McClell<strong>an</strong>d, J. W., Vörösmarty, C. J., Lammers, R. B., Shiklom<strong>an</strong>ov,<br />

A. I., Shiklom<strong>an</strong>ov, I. A., Rahmstorf, S. (2002). Incre<strong>as</strong>ing river discharge to <strong>the</strong> Arctic Oce<strong>an</strong>.<br />

Science, 298:2171–2173.<br />

Rawlins, M. A., Lammers, R. B., Frolking, S., Fekete, B. M., Vörösmarty, C. J. (2003). Simulating<br />

p<strong>an</strong>-Arctic runoff with a macro-scale terrestrial water bal<strong>an</strong>ce model. Hydrol. Processes,<br />

17:2521–2539.<br />

Serreze, M. C., Clark, M. P., Bromwich, D. H., Etringer, A. J., Zh<strong>an</strong>g, T., Lammers, R. (2003).<br />

Monitoring precipitation over <strong>the</strong> Arctic terrestrial drainage system: Data requirements, shortcomings,<br />

<strong>an</strong>d applications of atmospheric re<strong>an</strong>alysis. J. Hydrometeorol., 4(2):387–407.<br />

Shiklom<strong>an</strong>ov, A. I., Lammers, R. B., Vörösmarty, C. J. (2002). Widespread decline in hydrological<br />

monitoring threatens P<strong>an</strong>-Arctic research. Eos Tr<strong>an</strong>s. AGU, 83(2):13–17.<br />

Shmakin, A. B., Milly, P. C. D., Dunne, K. (2002). Global modeling of l<strong>an</strong>d water <strong>an</strong>d energy<br />

bal<strong>an</strong>ces. part III: Inter<strong>an</strong>nual variability. J. Hydrometeorol., 3(3):311–321.<br />

Ulaby, F. T., Moore, R. K., Fung, A. K. (1986). Microwave Remote Sensing: Active <strong>an</strong>d P<strong>as</strong>sive,<br />

Vol. III – Volume Scattering <strong>an</strong>d Emission Theory, Adv<strong>an</strong>ced Systems <strong>an</strong>d Applications. Artec<br />

House Inc., Dedham, MA.<br />

Vörösmarty, C. J., Fekete, B. M., Maybeck, M., Lammers, R. B. (2000). Geomorphometric attributes<br />

of <strong>the</strong> gloabl system of rivers at 30-min spatial resolution. J. Hydrol., 237:17–39.<br />

Vörösmarty, C. J., Hinzm<strong>an</strong>, L. D., Peterson, B. J., Bromwich, D. H., Hamilton, L. C., Morrison,<br />

J., Rom<strong>an</strong>ovsky, V. E., Sturm, M., Webb, R. S. (2001). The hydrologic cycle <strong>an</strong>d its role in<br />

arctic <strong>an</strong>d global environmental ch<strong>an</strong>ge: A rationale <strong>an</strong>d strategy for syn<strong>the</strong>sis study. Technical<br />

report, Fairb<strong>an</strong>ks, AK: Arctic Research Consortium of <strong>the</strong> U.S.<br />

Waliser, D. E., Waliser, S. E., Seo, K., Enjoku, E. (2005). Evaluation <strong>an</strong>d climate ch<strong>an</strong>ge projections<br />

of <strong>the</strong> global hydrological cycle in IPCC AR4 model simulations. In Eos Tr<strong>an</strong>s. AGU,volume86.<br />

Fall Meet. Suppl., Abstract H51I-07.<br />

Way, J. B., Zimmerm<strong>an</strong>n, R., Rignot, E., McDonald, K., Oren, R. (1997). Winter <strong>an</strong>d spring thaw<br />

<strong>as</strong> observed with imaging radar at BOREAS. J. Geophys. Res., 102(D24):29673–29684.<br />

Willmott, C. J. <strong>an</strong>d Matsuura, K. (2001). Arctic terrestrial air temperature <strong>an</strong>d precipitation:<br />

Monthly <strong>an</strong>d <strong>an</strong>nual time series (1930–2000) version 1. available online at:<br />

http://climate.geog.udel.edu/ climate/.<br />

Y<strong>an</strong>g, D., K<strong>an</strong>e, D. L., Hinzm<strong>an</strong>, L. D. (2002). Siberi<strong>an</strong> Lena River hydrologic regime <strong>an</strong>d recent<br />

ch<strong>an</strong>ge. J. Geophys. Res., 107(D23). 4694, doi:10.1029/2002JD002542.<br />

133


This page is intentionally bl<strong>an</strong>k.<br />

134


<strong>Snow</strong> Remote Sensing Posters<br />

135


This page is intentionally bl<strong>an</strong>k.<br />

136


ABSTRACT<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

AMSR-E Algorithm for <strong>Snow</strong>melt Onset Detection<br />

in Subarctic Heterogeneous Terrain<br />

J.D. APGAR 1 , J.M. RAMAGE 1 , R.A. MCKENNEY 2 , AND P. MALTAIS 3<br />

<strong>Snow</strong>melt onset in <strong>the</strong> upper Yukon River b<strong>as</strong>in, C<strong>an</strong>ada, c<strong>an</strong> be derived from brightness<br />

temperatures (Tb) obtained by <strong>the</strong> Adv<strong>an</strong>ced Microwave Sc<strong>an</strong>ning Radiometer for EOS (AMSR-<br />

E) on NASA’s Aqua satellite. This sensor, with a resolution of 14 x 8 km 2 for <strong>the</strong> 36.5 GHz<br />

frequency <strong>an</strong>d two to four observations per day, improves upon <strong>the</strong> twice-daily coverage <strong>an</strong>d 37 x<br />

28 km 2 spatial resolution of <strong>the</strong> Special Sensor Microwave Imager (SSM/I). The onset of melt<br />

within a snowpack causes <strong>an</strong> incre<strong>as</strong>e in <strong>the</strong> average daily 36.5 GHz vertically polarized Tb <strong>as</strong> well<br />

<strong>as</strong> a shift to high diurnal amplitude variations (DAV) <strong>as</strong> <strong>the</strong> snow melts during <strong>the</strong> day <strong>an</strong>d<br />

refreezes at night. The higher temporal <strong>an</strong>d spatial resolution makes AMSR-E more sensitive to<br />

sub-daily Tb oscillations, resulting in DAV that often show a greater daily r<strong>an</strong>ge compared to<br />

SSM/I. Therefore, thresholds of Tb > 246 K <strong>an</strong>d DAV > ±10 K developed for use with SSM/I<br />

have been adjusted for detecting melt onset with AMSR-E using ground-b<strong>as</strong>ed surface<br />

temperature <strong>an</strong>d snowpack wetness relationships. Using newly developed thresholds of Tb > 252<br />

K <strong>an</strong>d DAV > ±18 K, AMSR-E determined snowmelt onset correlates well with SSM/I<br />

observations in <strong>the</strong> small subarctic Wheaton River b<strong>as</strong>in through <strong>the</strong> 2004 <strong>an</strong>d 2005 winter/spring<br />

tr<strong>an</strong>sition. In addition, snowmelt onset derived from AMSR-E data gridded at a higher resolution<br />

th<strong>an</strong> <strong>the</strong> SSM/I data indicates that finer-scale differences in elevation <strong>an</strong>d l<strong>an</strong>d cover affect<br />

snowmelt onset <strong>an</strong>d are detectable with <strong>the</strong> AMSR-E sensor. B<strong>as</strong>ed on <strong>the</strong>se observations, <strong>the</strong><br />

enh<strong>an</strong>ced resolution of AMSR-E is more effective th<strong>an</strong> SSM/I at delineating spatial <strong>an</strong>d temporal<br />

snowmelt dynamics in <strong>the</strong> heterogeneous terrain of <strong>the</strong> upper Yukon River b<strong>as</strong>in.<br />

Keywords: snowmelt, p<strong>as</strong>sive microwave, SSM/I, AMSR-E, Yukon River, Wheaton River<br />

INTRODUCTION<br />

This Hydrologic <strong>an</strong>d atmospheric processes are closely tied with snow characteristics<br />

throughout <strong>the</strong> arctic <strong>an</strong>d subarctic regions of <strong>the</strong> Nor<strong>the</strong>rn Hemisphere. The extent of snow<br />

cover, snow water equivalent (SWE), <strong>an</strong>d <strong>the</strong> timing of spring snowmelt influence local hydrology<br />

<strong>as</strong> well <strong>as</strong> regional <strong>an</strong>d global water budgets, <strong>an</strong>d <strong>the</strong> high albedo of snow h<strong>as</strong> <strong>an</strong> import<strong>an</strong>t<br />

influence on <strong>the</strong> global energy budget (Vorosmarty et al., 2001). The spring snowmelt is <strong>an</strong><br />

<strong>an</strong>nual event in snow-covered, high-latitude drainage b<strong>as</strong>ins that sends signific<strong>an</strong>t qu<strong>an</strong>tities of<br />

meltwater downstream during a relatively short period of time. The timing <strong>an</strong>d magnitude of<br />

<strong>the</strong>se events are variable, but <strong>the</strong>y c<strong>an</strong> create a large impact on <strong>the</strong> b<strong>as</strong>in hydrology <strong>an</strong>d influence<br />

geomorphic ch<strong>an</strong>ge within <strong>the</strong> b<strong>as</strong>in. The use of p<strong>as</strong>sive microwave data to examine <strong>the</strong><br />

1 Earth <strong>an</strong>d Environmental Sciences, Lehigh University, Bethlehem, PA 18015, USA<br />

2 Geosciences/Environmental Studies, Pacific Lu<strong>the</strong>r<strong>an</strong> University, Tacoma, WA 98447, USA<br />

3 Water Survey of C<strong>an</strong>ada, Environment C<strong>an</strong>ada, 91782 Al<strong>as</strong>ka Highway, Whitehorse, Yukon<br />

Territory, C<strong>an</strong>ada Y1A 5B7<br />

137


characteristics <strong>as</strong>sociated with this se<strong>as</strong>onal snowmelt may provide a better underst<strong>an</strong>ding of <strong>the</strong><br />

sensitivity of snowmelt timing to climatic conditions <strong>as</strong> well <strong>as</strong> <strong>the</strong> effects of <strong>the</strong> snowmelt upon<br />

b<strong>as</strong>in hydrology.<br />

The Yukon River extends from <strong>the</strong> Juneau Icefield, British Columbia, to <strong>the</strong> Bering Sea, p<strong>as</strong>sing<br />

through <strong>the</strong> Yukon Territory <strong>an</strong>d Al<strong>as</strong>ka. It receives drainage from a b<strong>as</strong>in more th<strong>an</strong> 850,000 km 2<br />

in area. The majority of this drainage flows from <strong>the</strong> Bering Sea northward to provide 8% of <strong>the</strong><br />

total fluvial freshwater input to <strong>the</strong> Arctic Oce<strong>an</strong> (Aagaard <strong>an</strong>d Carmack, 1989). The b<strong>as</strong>in<br />

includes a variety of ecosystems <strong>as</strong> well <strong>as</strong> a r<strong>an</strong>ge of elevations <strong>an</strong>d local relief. This region h<strong>as</strong> a<br />

heterogeneous terrain marked by boreal forests, mountain r<strong>an</strong>ges, low-lying valleys, <strong>an</strong>d a variety<br />

of lakes. Much of <strong>the</strong> upper b<strong>as</strong>in is se<strong>as</strong>onally snow covered <strong>as</strong>ide from some glaciated regions<br />

near <strong>the</strong> headwaters.<br />

This paper will focus on <strong>the</strong> timing of <strong>the</strong> spring snowmelt tr<strong>an</strong>sition in <strong>the</strong> Wheaton River subb<strong>as</strong>in<br />

of <strong>the</strong> Yukon River (Fig. 1), from which <strong>the</strong> findings may <strong>the</strong>n be applied to larger subb<strong>as</strong>ins<br />

<strong>an</strong>d eventually <strong>the</strong> upper Yukon River b<strong>as</strong>in <strong>as</strong> a whole. The Wheaton River b<strong>as</strong>in is a<br />

small (875 km 2 ) upl<strong>an</strong>d b<strong>as</strong>in that h<strong>as</strong> heterogeneous topography <strong>an</strong>d terrain with only a small<br />

glacial input.<br />

Two p<strong>as</strong>sive microwave sensors, <strong>the</strong> Special Sensor Microwave Imager (SSM/I) <strong>an</strong>d Adv<strong>an</strong>ced<br />

Microwave Sc<strong>an</strong>ning Radiometer for EOS (AMSR-E), provide at le<strong>as</strong>t twice daily observations in<br />

<strong>the</strong> high latitude regions, <strong>an</strong>d <strong>the</strong> record of SSM/I observations extends back nearly two decades.<br />

Due to differences in brightness temperature emitted from dry versus wet snow, <strong>the</strong>se p<strong>as</strong>sive<br />

microwave sensors are effective at differentiating between snow that is melting <strong>an</strong>d snow that is<br />

not melting, or frozen. In addition, <strong>the</strong> ability of <strong>the</strong>se sensors to observe snow characteristics<br />

through clouds <strong>an</strong>d most o<strong>the</strong>r atmospheric conditions, <strong>as</strong> well <strong>as</strong> during <strong>the</strong> night, proves useful<br />

for examining snowmelt in high latitude are<strong>as</strong> that may have incre<strong>as</strong>ed cloud cover <strong>an</strong>d darkness<br />

during <strong>the</strong> winter <strong>an</strong>d spring months.<br />

The purpose of this study is to derive <strong>the</strong> timing of spring snowmelt in <strong>the</strong> Wheaton River b<strong>as</strong>in<br />

with recently acquired AMSR-E observations <strong>an</strong>d to test <strong>the</strong> sensitivity of this sensor to <strong>the</strong><br />

dynamic <strong>an</strong>d varied regional snowpack characteristics. A snowmelt onset algorithm developed by<br />

Ramage et al. (2006) for use with SSM/I data is modified for use with AMSR-E data to allow for<br />

a direct comparison of snowmelt onset timing from both of <strong>the</strong>se p<strong>as</strong>sive microwave sensors. In<br />

addition, higher-resolution AMSR-E data <strong>an</strong>d elevation data are used to show improvements of<br />

this sensor over <strong>the</strong> SSM/I sensor in detecting snowmelt within heterogeneous terrain. The<br />

algorithm is fine-tuned using hourly near-surface air temperature data <strong>an</strong>d ground-b<strong>as</strong>ed snow<br />

wetness data from within <strong>the</strong> Wheaton River b<strong>as</strong>in. Due to <strong>the</strong> greater spatial <strong>an</strong>d temporal<br />

resolution of <strong>the</strong> AMSR-E sensor compared to SSM/I, snowmelt dynamics c<strong>an</strong> be more accurately<br />

described with AMSR-E data in <strong>the</strong> mixed terrain of <strong>the</strong> upper Yukon River b<strong>as</strong>in.<br />

DATA<br />

P<strong>as</strong>sive microwave satellites provide a me<strong>an</strong>s by which to differentiate between snow that is<br />

melting <strong>an</strong>d snow that is frozen. Both AMSR-E <strong>an</strong>d SSM/I brightness temperature data were used<br />

for this <strong>an</strong>alysis. In addition, 90 meter digital elevation models of <strong>the</strong> Yukon Territory were<br />

utilized (Environment Yukon, 2000), <strong>as</strong> well <strong>as</strong> hourly temperature me<strong>as</strong>urements <strong>an</strong>d snow<br />

wetness me<strong>as</strong>urements from within <strong>the</strong> Wheaton River b<strong>as</strong>in.<br />

AMSR-E swath data, in <strong>the</strong> form of Level-2A Global Swath Spatially-Resampled Brightness<br />

Temperatures (Tb) V001, are supplied by <strong>the</strong> National <strong>Snow</strong> <strong>an</strong>d Ice Data Center (NSIDC)<br />

through <strong>the</strong> EOS Data Gateway with a record of observation that extends back to 2002 (Ashcroft<br />

<strong>an</strong>d Wentz, 2004). The AMSR-E sensor provides two to four observations of <strong>the</strong> study area per<br />

138


day <strong>an</strong>d h<strong>as</strong> a me<strong>an</strong> pixel resolution at 36.5 GHz of 12 x 12 km 2 (actual footprint size is 14 x 8<br />

km 2 ). Twice-daily SSM/I data extending back to 1987 come from <strong>the</strong> Defense Meteorological<br />

Satellite Program (DMSP) SSM/I <strong>an</strong>d have a nominal resolution for <strong>the</strong> 37 GHz ch<strong>an</strong>nel of 37 x<br />

28 km 2 . SSM/I images gridded to <strong>the</strong> Equal Area Scalable Earth Grid (EASE-Grid) 25 x 25 km 2<br />

resolution are provided by <strong>the</strong> NSIDC, <strong>an</strong>d this product separates <strong>the</strong> <strong>as</strong>cending <strong>an</strong>d descending<br />

p<strong>as</strong>ses. The AMSR-E data were gridded first at <strong>the</strong> EASE-Grid 25 x 25 km 2 resolution to compare<br />

directly with <strong>the</strong> SSM/I data <strong>an</strong>d establish to what degree <strong>the</strong>y are similar with a minimum of<br />

complicating factors. The AMSR-E data <strong>the</strong>n were gridded at <strong>the</strong> finer EASE-Grid 12.5 x 12.5<br />

km 2 resolution to examine improvements in snowmelt detection of AMSR-E upon <strong>the</strong> SSM/I<br />

sensor.<br />

P<strong>as</strong>sive microwave sensors have been used to examine characteristics of snow such <strong>as</strong> snow<br />

extent (e.g. Abdalati <strong>an</strong>d Steffen, 1997; Walker <strong>an</strong>d Goodison, 1993; W<strong>an</strong>g et al., 2005), snow<br />

depth (e.g. Josberger <strong>an</strong>d Mognard, 2002; Kelly et al., 2003), <strong>an</strong>d snow water equivalent (e.g.<br />

Derksen et al., 2005; Foster et al., 2005; Goita et al., 2003). The SSM/I sensor h<strong>as</strong> been used in<br />

previous studies to establish <strong>the</strong> timing of <strong>the</strong> spring melt tr<strong>an</strong>sition in <strong>the</strong> upper Yukon River<br />

b<strong>as</strong>in (Ramage et al., 2006) <strong>as</strong> well <strong>as</strong> in <strong>the</strong> Juneau Icefield (Ramage <strong>an</strong>d Isacks, 2002, 2003).<br />

Even though <strong>the</strong> SSM/I sensor provides twice-daily observations <strong>an</strong>d h<strong>as</strong> been shown to correlate<br />

well with ground-b<strong>as</strong>ed brightness temperature me<strong>as</strong>urements over fairly homogeneous terrain<br />

such <strong>as</strong> <strong>the</strong> Al<strong>as</strong>k<strong>an</strong> North Slope (Kim <strong>an</strong>d Engl<strong>an</strong>d, 2003), <strong>the</strong> pixel resolution of greater th<strong>an</strong> 25<br />

x 25 km 2 that results from <strong>the</strong> p<strong>as</strong>sive nature of <strong>the</strong> sensor is a problematic issue in monitoring<br />

dynamic ch<strong>an</strong>ges over heterogeneous terrain. The AMSR-E sensor, recently launched aboard<br />

NASA’s Aqua satellite in 2002, provides more observations of <strong>the</strong> study area per day <strong>an</strong>d, with <strong>an</strong><br />

improved pixel resolution over SSM/I, c<strong>an</strong> help to provide a more accurate examination of snow<br />

characteristics over mixed terrain.<br />

A signific<strong>an</strong>t difference in brightness temperature (Tb) between dry <strong>an</strong>d wet snow occurs at<br />

frequencies greater th<strong>an</strong> 10 GHz. The Tb of a material is related to its surface temperature (Ts) <strong>an</strong>d<br />

emissivity (E),<br />

Tb = ETs. (1)<br />

A rapid incre<strong>as</strong>e in emissivity occurs <strong>as</strong> a result of a small amount (~ 1-2%) of liquid water within<br />

<strong>the</strong> snowpack, causing <strong>the</strong> Tb to incre<strong>as</strong>e for wet snow (Ulaby et al., 1986). The Tb in <strong>the</strong> 19 <strong>an</strong>d<br />

37 GHz frequencies <strong>as</strong>sociated with <strong>the</strong> SSM/I sensor are useful in detecting melt on glaciers<br />

(Ramage <strong>an</strong>d Isacks, 2002, 2003) <strong>an</strong>d on heterogeneous terrain (Ramage et al., 2006) since <strong>the</strong> Tb<br />

tr<strong>an</strong>sition from dry to wet snow occurs <strong>as</strong> surface temperatures approach 0°C. From <strong>the</strong> AMSR-E<br />

sensor, <strong>the</strong> Tb from <strong>the</strong> vertically polarized 36.5 GHz frequency (wavelength of 0.82 cm) is<br />

comparable to <strong>the</strong> SSM/I sensor for <strong>the</strong> detection of snowmelt. Here, we focus on comparing <strong>an</strong>d<br />

contr<strong>as</strong>ting <strong>the</strong> Tb from <strong>the</strong> two sensors <strong>an</strong>d comparing <strong>the</strong> ability to detect <strong>the</strong> onset of snowmelt.<br />

During <strong>the</strong> spring snowmelt, <strong>the</strong> snowpack experiences cyclical daytime melt <strong>an</strong>d nighttime<br />

freeze, ch<strong>an</strong>ges that are m<strong>an</strong>ifested <strong>as</strong> diurnal differences in Tb. As a result of <strong>the</strong> at le<strong>as</strong>t twicedaily<br />

observations by <strong>the</strong> sensors, <strong>the</strong>se high diurnal amplitude variations of Tb (referred to <strong>as</strong><br />

DAV) c<strong>an</strong> be detected <strong>an</strong>d are useful in identification of <strong>the</strong>se dynamic tr<strong>an</strong>sition periods.<br />

The Wheaton River b<strong>as</strong>in is covered by six EASE-Grid 25 x 25 km 2 pixels (Fig. 1a <strong>an</strong>d Table<br />

1). Due to <strong>the</strong> coarse pixel size in relation to <strong>the</strong> size of <strong>the</strong> b<strong>as</strong>in, some pixels only cover a small<br />

percentage of <strong>the</strong> b<strong>as</strong>in (see Table 1). The Wheaton b<strong>as</strong>in h<strong>as</strong> a r<strong>an</strong>ge of elevations <strong>an</strong>d l<strong>an</strong>d cover<br />

that include bare mountain slopes, dense boreal forest, <strong>an</strong>d slopes of varying <strong>as</strong>pect (Brabets et al.,<br />

2000; Ramage et al., 2006). Pixels A02 <strong>an</strong>d A03 cover <strong>the</strong> upl<strong>an</strong>d headwaters of <strong>the</strong> Wheaton<br />

River <strong>an</strong>d have high me<strong>an</strong> elevations. The rest of <strong>the</strong> b<strong>as</strong>in is mostly covered by pixels B03 <strong>an</strong>d<br />

B02, representing <strong>the</strong> middle <strong>an</strong>d lower parts of <strong>the</strong> b<strong>as</strong>in respectively. Pixels C02 <strong>an</strong>d C03 cover<br />

a small fraction of <strong>the</strong> b<strong>as</strong>in but c<strong>an</strong> be used to examine <strong>the</strong> lowest elevations of <strong>the</strong> b<strong>as</strong>in. In<br />

139


elation to <strong>the</strong> AMSR-E data gridded to <strong>the</strong> EASE-Grid 12.5 x 12.5 km 2 resolution, <strong>the</strong> Wheaton<br />

River b<strong>as</strong>in is covered by parts of 14 pixels, improving <strong>the</strong> spatial scale at which differences in<br />

snowmelt characteristics c<strong>an</strong> be investigated (Fig. 1b). Adjacent pixels in this area appear to have<br />

similar Tb signatures that differ only slightly due to differences in l<strong>an</strong>d cover <strong>as</strong> well <strong>as</strong> elevation<br />

<strong>an</strong>d topography. Se<strong>as</strong>onal variations related to frost, vegetation, <strong>an</strong>d lakes are also factors in this<br />

complex l<strong>an</strong>dscape, but snowmelt h<strong>as</strong> <strong>the</strong> largest influence on Tb.<br />

Figure 1. Wheaton River b<strong>as</strong>in EASE-Grid (a) 25 x 25 km 2 pixels <strong>an</strong>d (b) 12.5 x 12.5 km 2 pixels<br />

superimposed on a 1:250,000 (90 m) digital elevation model (Environment Yukon, 2000). While <strong>the</strong> actual<br />

sensor footprints are more elliptical in shape, <strong>the</strong>se square pixels provide a good representation for <strong>the</strong> l<strong>an</strong>d<br />

area covered. The inset map shows <strong>the</strong> location of <strong>the</strong> b<strong>as</strong>in within <strong>the</strong> Yukon Territory. The b<strong>as</strong>in (black<br />

outline) is represented by parts of six 25 x 25 km 2 pixels <strong>an</strong>d fourteen 12.5 x 12.5 km 2 pixels. The Wheaton<br />

River near Carcross stream gauge is located near <strong>the</strong> river’s mouth at <strong>an</strong> elevation of 668 m (white dot in<br />

pixels B02 <strong>an</strong>d B02B).<br />

140


Table 1. Wheaton River b<strong>as</strong>in EASE-Grid 25 x 25 km 2 pixel characteristics (from Ramage et al., 2006)<br />

Near-surface air temperatures for <strong>the</strong> Wheaton River b<strong>as</strong>in were acquired in 2004 <strong>an</strong>d 2005<br />

using a HOBO automatic temperature logger located within pixel B02 at 668 m, approximately 10<br />

m from Environment C<strong>an</strong>ada’s “Wheaton River near Carcross” stream gauge (60°08′05″N,<br />

134°53′45″W). The HOBO recorded hourly air temperature <strong>an</strong>d relative humidity <strong>an</strong>d w<strong>as</strong><br />

situated about 1.5 m above <strong>the</strong> forest floor on <strong>the</strong> north facing side of a spruce tree (shaded most<br />

of <strong>the</strong> time). It w<strong>as</strong> installed under a protective platform in August 2004, <strong>an</strong>d data were<br />

downloaded periodically by Pat Maltais of <strong>the</strong> Water Survey of C<strong>an</strong>ada.<br />

A field campaign w<strong>as</strong> carried out during <strong>the</strong> spring of 2005 to collect a variety of in situ<br />

snowpack me<strong>as</strong>urements from snowpits, including snow density <strong>an</strong>d <strong>the</strong> dielectric properties of<br />

<strong>the</strong> snow at various depths, at times that coincided with AMSR-E satellite overp<strong>as</strong>ses. The<br />

dielectric properties of <strong>the</strong> snow, <strong>an</strong>d ultimately <strong>the</strong> snow liquid water content <strong>as</strong> a percentage of<br />

<strong>the</strong> volume, were derived from <strong>the</strong>se me<strong>as</strong>urements to relate to AMSR-E brightness temperature<br />

data. O<strong>the</strong>r observations of <strong>the</strong> snowpack in <strong>the</strong> upper Yukon River b<strong>as</strong>in provided support to<br />

show that <strong>the</strong> response in Tb could be attributed to snowmelt instead of liquid precipitation events.<br />

METHODS<br />

Ramage et al. (2006) developed a snowmelt onset algorithm for use with SSM/I 37V GHz data<br />

to detect <strong>the</strong> timing of snowmelt onset within <strong>the</strong> Wheaton River b<strong>as</strong>in. This algorithm<br />

determines <strong>the</strong> onset of snowmelt b<strong>as</strong>ed on <strong>the</strong> day of <strong>the</strong> year when certain thresholds for both<br />

brightness temperature (Tb) <strong>an</strong>d diurnal amplitude variation (DAV) are met (Table 2). For <strong>the</strong><br />

SSM/I data, ‘DAV Melt’ occurs when DAV is greater th<strong>an</strong> ±10 K, <strong>an</strong>d ‘Melt Onset’ occurs on <strong>the</strong><br />

day when Tb is greater th<strong>an</strong> 246 K <strong>an</strong>d DAV is greater th<strong>an</strong> ±10 K, indicating <strong>the</strong> snow is both<br />

melting periodically <strong>an</strong>d experiencing strong melt-refreeze cycles (Ramage <strong>an</strong>d Isacks, 2002,<br />

2003). In addition, ‘End DAV’ represents <strong>the</strong> end of <strong>the</strong> tr<strong>an</strong>sition when DAV decre<strong>as</strong>es below<br />

±10 K, <strong>an</strong>d <strong>the</strong> duration of <strong>the</strong> tr<strong>an</strong>sition period is calculated <strong>as</strong> <strong>the</strong> number of days between ‘Melt<br />

Onset’ <strong>an</strong>d ‘End DAV’.<br />

Table 2. <strong>Snow</strong>melt onset thresholds<br />

141


A number of <strong>as</strong>sumptions must be made in detecting snowmelt onset using this p<strong>as</strong>sive<br />

microwave data. The heterogeneous <strong>an</strong>d mountainous terrain within <strong>the</strong> upper Yukon River b<strong>as</strong>in<br />

most likely creates a complex Tb signal, but b<strong>as</strong>ed largely on field observations <strong>an</strong>d local wea<strong>the</strong>r<br />

records, it is <strong>as</strong>sumed that <strong>the</strong> majority of <strong>the</strong> l<strong>an</strong>d is covered with snow during <strong>the</strong> spring<br />

tr<strong>an</strong>sition period. In addition, since <strong>the</strong> methods focus on snowmelt events during <strong>the</strong> spring<br />

tr<strong>an</strong>sition from March through May, intermittent snowmelt <strong>an</strong>d thaw events during <strong>the</strong> winter<br />

months are largely ignored. It is also <strong>as</strong>sumed, b<strong>as</strong>ed on support from field data, that <strong>the</strong> events<br />

identified with <strong>the</strong> p<strong>as</strong>sive microwave data <strong>as</strong> melting snow are representative of snowmelt <strong>an</strong>d<br />

not liquid precipitation events.<br />

AMSR-E swath data were processed using <strong>the</strong> AMSR-E Swath-to-Grid Toolkit (AS2GT) in<br />

NSIDC’s P<strong>as</strong>sive Microwave Swath Data Tools (PMSDT). The data were <strong>the</strong>n gridded to <strong>the</strong><br />

EASE-Grid Nor<strong>the</strong>rn Hemisphere projection with a nominal resolution of 25 x 25 km 2 , <strong>the</strong> same<br />

format <strong>an</strong>d pixel resolution <strong>as</strong> <strong>the</strong> SSM/I data provided by <strong>the</strong> NSIDC. The Tb were extracted<br />

from each ch<strong>an</strong>nel using Interactive Data L<strong>an</strong>guage (IDL) code, creating a chronological time<br />

series of all Tb observations for 2004 <strong>an</strong>d 2005.<br />

Since <strong>the</strong>re are two to four AMSR-E observations per day, DAV c<strong>an</strong>not simply be calculated <strong>as</strong><br />

<strong>the</strong> difference from one observation to <strong>the</strong> next, <strong>as</strong> is done with <strong>the</strong> twice-daily SSM/I<br />

observations. Extra observations in a given day are often less th<strong>an</strong> 2.5 hours from <strong>the</strong> previous<br />

one <strong>an</strong>d have fairly similar Tb (Fig. 2). These extra values cause <strong>the</strong> DAV to artificially approach<br />

0 K <strong>as</strong> a result of similar Tb values even if <strong>the</strong> actual daily oscillation is much greater. An<br />

alternative to <strong>the</strong> raw data is to average <strong>the</strong> observations that are less th<strong>an</strong> 2.5 hours apart (Fig. 2).<br />

This creates a total of two observations for each day <strong>an</strong>d eliminates extr<strong>an</strong>eous DAV values that<br />

approach 0 K, allowing for more representative DAV values to be calculated <strong>an</strong>d used for<br />

examining snowmelt characteristics in a similar way to SSM/I DAV. O<strong>the</strong>r techniques used to<br />

calculate DAV, including using daily minimum <strong>an</strong>d maximum values, did not provide<br />

representative diurnal variations <strong>as</strong> well <strong>as</strong> <strong>the</strong> above method. The absolute value of DAV is used<br />

since DAV c<strong>an</strong> be ei<strong>the</strong>r positive or negative. For <strong>the</strong> purposes of examining snowmelt, only <strong>the</strong><br />

36.5 GHz vertically polarized observations were used.<br />

Figure 2. Calculation of DAV from AMSR-E 36.5V GHz Tb. In order to examine daily fluctuations, Tb less<br />

th<strong>an</strong> 2.5 hours apart (medium gray circles) were averaged. The running differences of <strong>the</strong>se observations<br />

with <strong>the</strong> o<strong>the</strong>r observations produces a more accurate DAV representation (solid black line) th<strong>an</strong> if all<br />

observations are used to calculate daily variations (gray line).<br />

142


The SSM/I snowmelt onset algorithm w<strong>as</strong> modified for use with AMSR-E 36.5 GHz vertically<br />

polarized Tb data. Initial examination of this data from <strong>the</strong> AMSR-E sensor for 2005 indicated<br />

that <strong>the</strong> observations were more sensitive to daily fluctuations <strong>an</strong>d that <strong>the</strong> previous Tb snowmelt<br />

threshold for SSM/I of 246 K w<strong>as</strong> too low. To develop a new threshold for <strong>the</strong> snowmelt onset<br />

algorithm, <strong>the</strong> AMSR-E 36.5V GHz Tb for <strong>the</strong> six Wheaton River b<strong>as</strong>in pixels were plotted<br />

against temporally-corresponding near-surface air temperature data from <strong>the</strong> Wheaton River for<br />

<strong>the</strong> period of August 2004 to December 2005 (Fig. 3). When <strong>the</strong> air temperature is near 0°C, <strong>the</strong><br />

Tb rapidly incre<strong>as</strong>es due to <strong>the</strong> presence of liquid water in <strong>the</strong> snowpack. The Tb at which this<br />

tr<strong>an</strong>sition from frozen to melting snow occurs is considered <strong>the</strong> threshold. A histogram of <strong>an</strong>nual<br />

AMSR-E 36.5V GHz Tb data w<strong>as</strong> also produced to identify <strong>the</strong> new Tb threshold (Fig. 4). A<br />

typical year of observations in this se<strong>as</strong>onally-snow covered region h<strong>as</strong> a bimodal distribution of<br />

Tb related to frozen, dry snow during <strong>the</strong> cold months <strong>an</strong>d wet snow or no snow during <strong>the</strong> warmer<br />

months, <strong>an</strong>d <strong>the</strong> Tb that separates <strong>the</strong>se two conditions is <strong>the</strong> Tb threshold (Fig. 4, Table 2).<br />

Figure 3. Scatterplots of Wheaton River near-surface air temperatures compared to AMSR-E 36.5 GHz<br />

vertically polarized Tb for <strong>the</strong> six 25 x 25 km 2 Wheaton River b<strong>as</strong>in pixels for 2004-2005. The tr<strong>an</strong>sition in<br />

Tb from frozen (1,2) to melting (3) snow, which appears <strong>as</strong> <strong>an</strong> inflection in <strong>the</strong> slope near a surface<br />

temperature of 0°C, helps to establish <strong>the</strong> Tb threshold of snowmelt <strong>as</strong> 252 K. The two unique distributions<br />

that exist at temperatures below 0°C in most of <strong>the</strong> pixels (1,2) may be related to snowpack grain size growth<br />

or <strong>the</strong> presence of ice on vegetation. The area of high temperatures (4) represents snow-free surfaces during<br />

<strong>the</strong> summer months.<br />

143


Figure 4. Frequency histogram of all AMSR-E 36.5V GHz Tb for <strong>the</strong> six Wheaton River b<strong>as</strong>in 25 x 25 km 2<br />

pixels during <strong>the</strong> 2004-2005 period. The Tb are grouped in bins of 2 K. The bimodal distribution suggests a<br />

threshold of 252 K is more suitable for distinguishing between frozen (left distribution) <strong>an</strong>d wet (right<br />

distribution) snow using AMSR-E in <strong>the</strong> Wheaton River b<strong>as</strong>in th<strong>an</strong> <strong>the</strong> SSM/I threshold of 246 K. The<br />

individual pixel sets, while not <strong>as</strong> clearly defined, also have minima near <strong>the</strong> threshold of 252 K.<br />

Me<strong>as</strong>urements of snowpack properties were obtained in <strong>the</strong> Wheaton River b<strong>as</strong>in during <strong>the</strong><br />

spring of 2005. More th<strong>an</strong> fifty snowpits were dug in a variety of locations, encomp<strong>as</strong>sing a r<strong>an</strong>ge<br />

of elevations, terrains, <strong>an</strong>d forest cover in <strong>an</strong> attempt to represent <strong>the</strong> heterogeneous nature of <strong>the</strong><br />

region. Characteristics of <strong>the</strong> snowpack were described <strong>an</strong>d me<strong>as</strong>ured, including <strong>the</strong> general<br />

stratigraphy <strong>as</strong> well <strong>as</strong> grain size, air <strong>an</strong>d snow temperature, <strong>an</strong>d snow structure at various depths.<br />

A comprehensive <strong>an</strong>alysis of <strong>the</strong>se parameters is beyond <strong>the</strong> scope of this paper. A snow surface<br />

dielectric capacit<strong>an</strong>ce probe <strong>an</strong>d me<strong>as</strong>urements of snow density were used to derive snow wetness<br />

<strong>as</strong> a percentage of <strong>the</strong> volume at different depths (Denoth, 1989). Surface <strong>an</strong>d near-surface<br />

wetness values were averaged from multiple me<strong>as</strong>urements at each site, <strong>an</strong>d <strong>the</strong>se wetness values<br />

were later compared with AMSR-E 36.5V GHz Tb from overp<strong>as</strong>ses obtained less th<strong>an</strong> one hour<br />

before or after <strong>the</strong> ground observations (Figs. 5a <strong>an</strong>d 5b, Table 3). In addition to supporting <strong>the</strong><br />

new Tb threshold, <strong>the</strong>se observations helped show that <strong>the</strong> observed Tb response is linked to<br />

melting snow <strong>an</strong>d is not <strong>the</strong> result of liquid precipitation or o<strong>the</strong>r events.<br />

Once <strong>the</strong> new Tb threshold w<strong>as</strong> determined, <strong>the</strong> AMSR-E 36.5V GHz data, along with melt<br />

onset <strong>an</strong>d spring tr<strong>an</strong>sition data, were compared with <strong>the</strong> melt onset results of <strong>the</strong> SSM/I 37V GHz<br />

data for all six of <strong>the</strong> Wheaton River b<strong>as</strong>in 25 x 25 km 2 pixels (Figs. 6a <strong>an</strong>d 6b, Table 4). This<br />

allowed for a new DAV threshold to be created for <strong>the</strong> AMSR-E data (Table 2), since <strong>the</strong> daily<br />

fluctuations appear to have incre<strong>as</strong>ed with <strong>the</strong> AMSR-E sensor, <strong>an</strong>d also helped to determine if <strong>the</strong><br />

new thresholds are effective at detecting snowmelt <strong>as</strong> related to previous snowmelt onset detection<br />

studies using SSM/I Tb data.<br />

After determining that similar snow melt onset results could be obtained from AMSR-E <strong>an</strong>d<br />

SSM/I Tb data gridded at <strong>the</strong> same 25 x 25 km 2 resolution, <strong>the</strong> effectiveness of <strong>the</strong> AMSR-E<br />

sensor’s improved spatial resolution w<strong>as</strong> investigated. The AMSR-E swath data were gridded to<br />

<strong>the</strong> EASE-Grid with a resolution of 12.5 x 12.5 km 2 , which is more representative of <strong>the</strong> 12 km<br />

me<strong>an</strong> spatial resolution of <strong>the</strong> 36.5V GHz ch<strong>an</strong>nel (Ashcroft <strong>an</strong>d Wentz, 2004). This gridded<br />

resolution allowed for <strong>the</strong> creation of four finer resolution pixels of information for each 25 x 25<br />

km 2 pixel (Fig. 1b). Elevation data were used to identify <strong>the</strong> fine-resolution pixels within <strong>an</strong>y<br />

144


coarse-resolution pixel with <strong>the</strong> highest <strong>an</strong>d lowest me<strong>an</strong> elevations (Table 5). The AMSR-E<br />

36.5V GHz Tb for <strong>the</strong>se pixels, <strong>as</strong> well <strong>as</strong> <strong>the</strong> snowmelt onset timing, were <strong>the</strong>n compared with <strong>the</strong><br />

coarser 25 x 25 km 2 data to examine <strong>the</strong> effects of <strong>the</strong> AMSR-E sensor’s incre<strong>as</strong>ed spatial<br />

resolution on improving <strong>the</strong> ability to detect snowmelt in heterogeneous terrain (Figs. 7a <strong>an</strong>d 7b,<br />

Table 5).<br />

Table 3. Spring 2005 Wheaton River b<strong>as</strong>in snow wetness me<strong>as</strong>urements <strong>an</strong>d AMSR-E 25 km 2 Tb<br />

RESULTS<br />

The two main techniques used to determine a new snowmelt onset Tb threshold for <strong>the</strong> AMSR-E<br />

36.5V GHz ch<strong>an</strong>nel both found <strong>the</strong> threshold between non-melting <strong>an</strong>d melting snow to be 252 K.<br />

The first method, in which <strong>the</strong> Tb for <strong>the</strong> six Wheaton b<strong>as</strong>in pixels were compared with air<br />

temperature data for 2004 <strong>an</strong>d 2005 (refer to Fig. 3), illustrates that a Tb threshold of around 252 K<br />

is <strong>an</strong> effective boundary between (1,2) non-melting <strong>an</strong>d (3) melting snow near 0°C. In <strong>the</strong>se nonlinear<br />

plots, it c<strong>an</strong> be noted that two distinct relationships occur on ei<strong>the</strong>r side of <strong>the</strong> 0°C mark.<br />

Below 0°C, <strong>the</strong> Tb are below 252 K <strong>an</strong>d incre<strong>as</strong>e very slightly with <strong>an</strong> incre<strong>as</strong>e in temperature. As<br />

<strong>the</strong> temperature approaches <strong>an</strong>d exceeds 0°C, <strong>the</strong> Tb rapidly incre<strong>as</strong>es above 252 K (Fig. 3). This<br />

rapid ch<strong>an</strong>ge in <strong>the</strong> slope of Tb when <strong>the</strong> air temperature is near 0°C at Tb = 252 K is due to <strong>the</strong><br />

high sensitivity of microwaves to liquid water within <strong>the</strong> melting snowpack. The tr<strong>an</strong>sition from<br />

frozen to melting snow is especially clear for pixel B02, which contains <strong>the</strong> site of <strong>the</strong> near-surface<br />

air temperature observations, <strong>as</strong> <strong>the</strong> ch<strong>an</strong>ge in slope between <strong>the</strong>se two conditions at 0°C occurs at<br />

<strong>the</strong> Tb threshold of 252 K (Fig. 3).<br />

It is also worth noting that two distinct distributions of Tb occur when compared to air<br />

temperatures below 0°C (Fig. 3). The upper Tb response (Fig. 3, #1) is from periods in November<br />

from both 2004 <strong>an</strong>d 2005, while <strong>the</strong> lower response (Fig. 3, #2) is from periods in December of<br />

<strong>the</strong> same years. The cause of <strong>the</strong> upper response may be related to ice accumulation on trees<br />

during <strong>the</strong> earlier winter months, a condition that w<strong>as</strong> observed in <strong>the</strong> 2006 winter se<strong>as</strong>on. During<br />

<strong>the</strong> winter, especially during unusually warm years, open lakes that have not yet frozen c<strong>an</strong> create<br />

145


<strong>an</strong> icy fog that creeps up valleys <strong>an</strong>d allows for ice to accumulate on vegetation. These distinct<br />

distributions may also be <strong>an</strong> effect of grain size growth during <strong>the</strong> early winter months,<br />

influencing <strong>the</strong> scattering properties of <strong>the</strong> snowpack. At <strong>the</strong> 36.5 GHz frequency, emissivity<br />

decre<strong>as</strong>es proportional to <strong>an</strong> incre<strong>as</strong>e in grain size (Mätzler, 1994), <strong>an</strong>d so <strong>as</strong> <strong>the</strong> grains within <strong>the</strong><br />

snowpack grow early in <strong>the</strong> winter, <strong>the</strong>ir Tb decre<strong>as</strong>es from <strong>the</strong> upper distribution to <strong>the</strong> lower<br />

distribution (Fig. 3). As <strong>the</strong>se events both occur within <strong>the</strong> same time frame, fur<strong>the</strong>r investigation<br />

is required to identify <strong>the</strong> source of <strong>the</strong>se differing relationships.<br />

The second technique, examining a histogram of AMSR-E 36.5V GHz Tb distribution from <strong>the</strong><br />

Wheaton b<strong>as</strong>in for all 2004 <strong>an</strong>d 2005 data, also produced a Tb threshold of 252 K (Fig. 4). The<br />

frequency histogram, with a bin size of 2 K, displays a bimodal distribution from which it is<br />

interpreted that brightness temperatures less th<strong>an</strong> 252 K relate to frozen, non-melting snow <strong>an</strong>d<br />

brightness temperatures greater th<strong>an</strong> 252 K relate to wet, melting snow <strong>as</strong> well <strong>as</strong> no snow during<br />

<strong>the</strong> summer months. The sets of individual pixels, representing <strong>the</strong> upper (pixels A02 <strong>an</strong>d A03),<br />

middle (pixels B02 <strong>an</strong>d B03), <strong>an</strong>d lower (pixels C02 <strong>an</strong>d C03) portions of <strong>the</strong> Wheaton River<br />

b<strong>as</strong>in, do not have <strong>as</strong> clear a bimodal distribution but do have comparable minimum values around<br />

252 K (Fig. 4). This, along with <strong>the</strong> plots of Tb against air temperature, suggests that <strong>the</strong> Tb<br />

threshold is effective in differentiating between dry <strong>an</strong>d wet snow at <strong>the</strong> drainage b<strong>as</strong>in <strong>an</strong>d subdrainage<br />

b<strong>as</strong>in scales in this heterogeneous terrain.<br />

The DAV calculation, which involves taking <strong>the</strong> difference between night <strong>an</strong>d day observations,<br />

shows <strong>the</strong> daily contr<strong>as</strong>t in brightness temperatures at this time of year. The timing of <strong>the</strong> AMSR-<br />

E overp<strong>as</strong>ses for this region are around 03:30 <strong>an</strong>d 13:30 local time (PST), which are likely near<br />

<strong>the</strong> times of minimum <strong>an</strong>d maximum daily melt, <strong>an</strong>d so <strong>the</strong> difference in <strong>the</strong> Tb from <strong>the</strong>se times<br />

provides a good indicator of melting <strong>an</strong>d refreezing snow. The DAV threshold incre<strong>as</strong>ed due to<br />

<strong>the</strong> apparent incre<strong>as</strong>ed sensitivity of <strong>the</strong> AMSR-E sensor to daily Tb fluctuations (Fig. 6a). A new<br />

DAV threshold of ±18 K, which w<strong>as</strong> determined by comparing <strong>the</strong> AMSR-E DAV with SSM/I<br />

DAV for <strong>the</strong> Wheaton River pixels in 2005 (Fig. 6b), represents times when strong melt-refreeze<br />

cycles are occurring during <strong>the</strong> spring snowmelt tr<strong>an</strong>sition.<br />

The ch<strong>an</strong>ge in Tb <strong>an</strong>d DAV thresholds from SSM/I to AMSR-E (Table 2) is most likely <strong>the</strong><br />

result of differences in <strong>the</strong> times of data acquisition for <strong>the</strong> two sensors. Daily AMSR-E<br />

overp<strong>as</strong>ses for <strong>the</strong> area occur around 03:30 <strong>an</strong>d 13:30 PST, where<strong>as</strong> SSM/I overp<strong>as</strong>ses occur<br />

around 08:30 <strong>an</strong>d 18:30 PST. The AMSR-E sensor makes observations during <strong>the</strong> early morning,<br />

when temperatures <strong>an</strong>d corresponding Tb would be near <strong>the</strong>ir daily minimum, <strong>an</strong>d again in <strong>the</strong><br />

early afternoon, when much of <strong>the</strong> mountainous terrain would be near <strong>the</strong> daily maximum air <strong>an</strong>d<br />

brightness temperatures. This creates larger DAV in relation to <strong>the</strong> DAV from SSM/I, since <strong>the</strong><br />

timing of SSM/I overp<strong>as</strong>ses is fur<strong>the</strong>r from <strong>the</strong>se daily minimum <strong>an</strong>d maximum, <strong>an</strong>d <strong>the</strong>refore<br />

requires a larger DAV threshold. In addition, <strong>the</strong> DAV threshold must incre<strong>as</strong>e so <strong>as</strong> to not<br />

indicate that melting is occurring when large DAV are observed during <strong>the</strong> summer, when much<br />

of <strong>the</strong> l<strong>an</strong>d is snow-free (Fig. 6b).<br />

Field-observed snow liquid-water content me<strong>as</strong>urements from <strong>the</strong> Wheaton River b<strong>as</strong>in during<br />

<strong>the</strong> spring of 2005 support <strong>the</strong> Tb threshold of 252 K <strong>an</strong>d also show that <strong>the</strong> <strong>as</strong>sumed response in<br />

Tb to melting snow c<strong>an</strong> be directly attributed to snowmelt instead of o<strong>the</strong>r effects, such <strong>as</strong> liquid<br />

precipitation events. Surface (0 cm) <strong>an</strong>d near surface (0.5-2 cm) wetness me<strong>as</strong>urements relate<br />

very well to AMSR-E 36.5V GHz Tb observations obtained at coincident times (Figs. 5a <strong>an</strong>d 5b,<br />

Table 3). At wetness values below 2%, <strong>the</strong> Tb are clustered between 230 K <strong>an</strong>d 245 K, but <strong>as</strong> <strong>the</strong><br />

liquid-water content of <strong>the</strong> snowpack incre<strong>as</strong>es above 2% by volume, <strong>the</strong> Tb appear to incre<strong>as</strong>e<br />

near <strong>an</strong>d above <strong>the</strong> threshold of 252 K (Fig. 5a). For <strong>the</strong> period prior to <strong>an</strong>d including <strong>the</strong><br />

beginning of <strong>the</strong> spring tr<strong>an</strong>sition, <strong>the</strong> ch<strong>an</strong>ges in <strong>the</strong> snowpack wetness through time closely<br />

follow that of <strong>the</strong> AMSR-E Tb, <strong>an</strong>d at wetness values above 2%, <strong>the</strong> Tb are at or above <strong>the</strong> 252 K<br />

threshold (Fig. 5b). This relationship verifies that <strong>the</strong> Tb response is due to <strong>the</strong> presence of liquid<br />

water within melting snow.<br />

146


Figure 5. (a) Scatterplot of snowpack wetness compared to AMSR-E 36.5V GHz Tb within <strong>the</strong> Wheaton<br />

River b<strong>as</strong>in. Below 2% wetness, Tb are clustered between 230 K <strong>an</strong>d 245 K, but <strong>an</strong> incre<strong>as</strong>e in liquid-water<br />

content beyond 2% creates <strong>an</strong> abrupt rise in Tb around <strong>the</strong> threshold of 252 K. (b) The snowpack wetness<br />

(gray) relates well to <strong>the</strong> AMSR-E Tb (black) during <strong>the</strong> period prior to <strong>the</strong> spring tr<strong>an</strong>sition.<br />

With <strong>the</strong>se new AMSR-E snowmelt onset thresholds (Tb > 252 K <strong>an</strong>d DAV > ±18 K), <strong>the</strong><br />

AMSR-E spring tr<strong>an</strong>sition in 2004 <strong>an</strong>d 2005 w<strong>as</strong> characterized for <strong>the</strong> six Wheaton River 25 x 25<br />

km 2 pixels <strong>an</strong>d compared with SSM/I data for <strong>the</strong> same pixels <strong>an</strong>d years (Figs. 6a <strong>an</strong>d 6b, Table<br />

4). The snowmelt onset algorithm developed by Ramage et al. (2006) w<strong>as</strong> used for <strong>the</strong> SSM/I<br />

data, using thresholds of Tb > 246 K <strong>an</strong>d DAV > ±10 K (Table 2). For pixel B02, <strong>the</strong> onset of<br />

snowmelt compares very well between AMSR-E <strong>an</strong>d SSM/I, <strong>as</strong> does <strong>the</strong> end of <strong>the</strong> spring<br />

tr<strong>an</strong>sition <strong>an</strong>d <strong>the</strong> total duration (Figs. 6a <strong>an</strong>d 6b). It should be noted that although <strong>the</strong> AMSR-E<br />

thresholds were met for this pixel around day 95, <strong>the</strong>y were not met for a consistent length of time,<br />

<strong>an</strong>d so <strong>the</strong> spring tr<strong>an</strong>sition is interpreted <strong>as</strong> beginning about 10 days later (Figs. 6a <strong>an</strong>d 6b). The<br />

DAV of <strong>the</strong> AMSR-E observations are higher th<strong>an</strong> <strong>the</strong> corresponding SSM/I observations<br />

throughout <strong>the</strong> spring tr<strong>an</strong>sition <strong>as</strong> well <strong>as</strong> for much of <strong>the</strong> summer. Preliminary investigations of<br />

AMSR-E derived snowmelt in <strong>the</strong> larger Pelly River b<strong>as</strong>in, approximately 250 km northwest of<br />

<strong>the</strong> Wheaton River b<strong>as</strong>in, indicate <strong>the</strong> AMSR-E Tb <strong>an</strong>d DAV thresholds are effective for detecting<br />

snowmelt on a regional scale.<br />

147


Figure 6. (a) Comparison of pixel B02 AMSR-E <strong>an</strong>d SSM/I Tb for 2005. The AMSR-E Tb (black) match<br />

closely with <strong>the</strong> SSM/I Tb (gray), although <strong>the</strong>y appear to show greater variability during <strong>the</strong> spring tr<strong>an</strong>sition<br />

<strong>an</strong>d summer months. The melt onset date <strong>an</strong>d duration for <strong>the</strong> two records are also similar. (b) Comparison<br />

of pixel B02 AMSR-E <strong>an</strong>d SSM/I DAV for 2005. The AMSR-E DAV (black) is consistently higher th<strong>an</strong> <strong>the</strong><br />

SSM/I DAV (gray) throughout <strong>the</strong> spring <strong>an</strong>d summer. The AMSR-E DAV threshold of ±18 K appears to<br />

match with <strong>the</strong> SSM/I DAV threshold of ±10 K <strong>an</strong>d <strong>the</strong> timing of SSM/I melt onset.<br />

Table 4. AMSR-E <strong>an</strong>d SSM/I melt dates (day of year) for <strong>the</strong> Wheaton River B<strong>as</strong>in<br />

148


For both 2004 <strong>an</strong>d 2005, <strong>the</strong> AMSR-E <strong>an</strong>d SSM/I melt onset dates for <strong>the</strong> six Wheaton River 25<br />

x 25 km 2 pixels match up very well, with <strong>the</strong> dates differing by less th<strong>an</strong> two days (Table 4). Pixel<br />

B03 in 2004 w<strong>as</strong> found to have <strong>the</strong> same melt onset date for both sensors, <strong>as</strong> w<strong>as</strong> pixel C02 in<br />

2005. These well correlated melt onset dates indicate that <strong>the</strong> new AMSR-E data <strong>an</strong>d thresholds<br />

are effective in detecting snowmelt in <strong>the</strong> Wheaton River b<strong>as</strong>in at <strong>the</strong> same resolution <strong>as</strong> SSM/I.<br />

In addition, this relationship may eventually allow for <strong>the</strong> observations from both sensors to be<br />

combined to create <strong>an</strong> even more robust satellite record of observation.<br />

Since <strong>the</strong> AMSR-E sensor provides enh<strong>an</strong>ced resolutions over <strong>the</strong> SSM/I sensor, especially in<br />

<strong>the</strong> 36.5 GHz ch<strong>an</strong>nel, <strong>the</strong> timing of snowmelt within <strong>the</strong> Wheaton River b<strong>as</strong>in w<strong>as</strong> also<br />

determined using Tb data gridded to <strong>the</strong> EASE-Grid 12.5 x 12.5 km 2 resolution (Fig. 1b). Pixel<br />

B02 (25 x 25 km 2 resolution) <strong>an</strong>d <strong>the</strong> four finer-resolution pixels within it were examined in <strong>an</strong><br />

attempt to te<strong>as</strong>e out <strong>the</strong> effects of <strong>the</strong> heterogeneous terrain on snowmelt onset (Table 5).<br />

Differences between <strong>the</strong> plots of AMSR-E Tb in <strong>the</strong> first half of 2005 for pixels B02C <strong>an</strong>d B02B,<br />

which represent <strong>the</strong> highest <strong>an</strong>d lowest me<strong>an</strong> elevations within pixel B02 respectively, indicate<br />

that <strong>the</strong> spatial resolution of <strong>the</strong> AMSR-E sensor improves upon <strong>the</strong> SSM/I sensor’s ability to<br />

detect snowmelt in mixed terrain, especially in are<strong>as</strong> with a r<strong>an</strong>ge of elevations (Figs. 7a <strong>an</strong>d 7b,<br />

Table 5).<br />

Table 5. EASE-Grid 12.5 x 12.5 km 2 pixel characteristics <strong>an</strong>d melt onset dates for 2005<br />

In relation to <strong>the</strong> 2005 melt onset timing <strong>an</strong>d spring tr<strong>an</strong>sition duration for pixel B02, <strong>the</strong> timing<br />

of <strong>the</strong>se events for <strong>the</strong> finer-resolution pixels is similar (Table 5). Two notable exceptions are<br />

pixel B02B, <strong>the</strong> lowest in me<strong>an</strong> elevation, <strong>an</strong>d pixel B02C, <strong>the</strong> highest in me<strong>an</strong> elevation. Melt<br />

onset occurs much earlier in B02B th<strong>an</strong> in <strong>the</strong> o<strong>the</strong>r pixels, most likely <strong>as</strong> a result of warmer<br />

temperatures within its lower elevations. For B02C, <strong>the</strong> spring tr<strong>an</strong>sition period l<strong>as</strong>ts more th<strong>an</strong> 20<br />

days longer th<strong>an</strong> <strong>the</strong> o<strong>the</strong>r pixels, <strong>an</strong>d this is due to colder temperatures in <strong>the</strong> higher elevations<br />

keeping <strong>the</strong> nighttime snowpack frozen <strong>an</strong>d DAV high well into <strong>the</strong> spring. Examining <strong>the</strong> time<br />

during which snowmelt onset h<strong>as</strong> begun for B02B (Fig. 7b), it c<strong>an</strong> be noted that <strong>the</strong> Tb for B02B<br />

are consistently above <strong>the</strong> threshold from day 94 to 100, where<strong>as</strong> <strong>the</strong> Tb for B02C never reach <strong>the</strong><br />

threshold during this period. Fur<strong>the</strong>rmore, Tb for pixel B02 indicate melting during <strong>the</strong> day from<br />

days 96 to 100, <strong>an</strong>d while this concurs with B02B, it disagrees with B02C, <strong>an</strong>d so <strong>the</strong> higher<br />

resolution AMSR-E data is more effective at differentiating melting <strong>an</strong>d non-melting are<strong>as</strong> in this<br />

heterogeneous terrain th<strong>an</strong> lower resolution AMSR-E <strong>an</strong>d SSM/I data.<br />

149


Figure 7. (a) Comparison of higher resolution AMSR-E Tb from 2005 for pixels B02B <strong>an</strong>d B02C. The lower<br />

elevation B02B (black) h<strong>as</strong> higher overall Tb th<strong>an</strong> B02C (gray) that relate to warmer temperatures. This<br />

results in <strong>an</strong> earlier melt onset for B02B in relation to B02C <strong>an</strong>d <strong>the</strong> o<strong>the</strong>r pixels (Table 5). (b) A detailed<br />

look at days 90 to 100 reveals that <strong>the</strong> Tb threshold is met consistently in <strong>the</strong> lower elevations of B02B<br />

(black) but not in <strong>the</strong> higher elevations of B02C (thick gray). Pixel B02 (thin gray) is more closely related to<br />

B02B, but <strong>the</strong> Tb threshold is not met <strong>as</strong> early <strong>as</strong> in B02B.<br />

CONCLUSIONS<br />

AMSR-E p<strong>as</strong>sive microwave data c<strong>an</strong> be used to investigate snowmelt at <strong>the</strong> b<strong>as</strong>in <strong>an</strong>d regional<br />

scales. In addition, <strong>the</strong> finer resolution of AMSR-E over that of SSM/I allows for <strong>the</strong> snowpack<br />

characteristics of are<strong>as</strong> with heterogeneous terrain to be more accurately examined. For <strong>the</strong><br />

Wheaton River dat<strong>as</strong>et, AMSR-E is more sensitive to daily brightness temperature fluctuations<br />

th<strong>an</strong> SSM/I at <strong>the</strong> same spatial scale. Relating <strong>the</strong> two dat<strong>as</strong>ets at <strong>the</strong> same scale provides <strong>the</strong><br />

foundation for a longer-term study of snowmelt dynamics combining <strong>the</strong> historical record of both<br />

sensors. Plots of AMSR-E Tb <strong>an</strong>d near-surface air temperatures, <strong>as</strong> well <strong>as</strong> a bimodal distribution<br />

of Tb <strong>an</strong>d comparisons with snow liquid-water content, suggest a Tb threshold of 252 K c<strong>an</strong> be<br />

used to make <strong>the</strong> distinction between melting <strong>an</strong>d non-melting snow in combination with a DAV<br />

150


threshold of ±18 K that reflects <strong>the</strong> spring tr<strong>an</strong>sition when large melt-refreeze cycles are occurring.<br />

The larger oscillations in daily Tb are most likely due to <strong>the</strong> timing of data acquisition being closer<br />

to <strong>the</strong> actual maximum <strong>an</strong>d minimum daily Tb. The snowmelt onset algorithm will also be<br />

applicable in most arctic <strong>an</strong>d subarctic l<strong>an</strong>dscapes.<br />

<strong>Snow</strong>melt onset dates from AMSR-E, <strong>as</strong> determined using <strong>the</strong> new Tb <strong>an</strong>d DAV thresholds,<br />

compare very well with SSM/I for <strong>the</strong> Wheaton River b<strong>as</strong>in, differing by a maximum of two days.<br />

The enh<strong>an</strong>ced spatial resolution of <strong>the</strong> AMSR-E data provides <strong>an</strong> improvement over SSM/I in<br />

detecting snowmelt onset in are<strong>as</strong> of mixed terrain. Using this AMSR-E melt onset algorithm, <strong>the</strong><br />

snowpack dynamics of <strong>the</strong> Wheaton River b<strong>as</strong>in, <strong>as</strong> well <strong>as</strong> o<strong>the</strong>r are<strong>as</strong> of <strong>the</strong> upper Yukon River<br />

b<strong>as</strong>in, c<strong>an</strong> be fur<strong>the</strong>r investigated spatially <strong>an</strong>d temporally for <strong>the</strong> record of AMSR-E observations.<br />

The improvements by <strong>the</strong> AMSR-E sensor upon SSM/I allow for a more effective spatial <strong>an</strong>d<br />

temporal examination of snowmelt dynamics in <strong>the</strong> heterogeneous terrain of <strong>the</strong> upper Yukon<br />

River b<strong>as</strong>in. It will be a signific<strong>an</strong>t improvement to our ability to underst<strong>an</strong>d <strong>the</strong> melt processes<br />

<strong>an</strong>d timing in <strong>an</strong> import<strong>an</strong>t, remote, high latitude drainage b<strong>as</strong>in.<br />

ACKNOWLEDGEMENTS<br />

We would like to extend our th<strong>an</strong>ks to Mary Jo Brodzik, Am<strong>an</strong>da Leon, <strong>an</strong>d <strong>the</strong> National <strong>Snow</strong><br />

<strong>an</strong>d Ice Data Center for EASE-grid satellite data <strong>an</strong>d continued <strong>as</strong>sist<strong>an</strong>ce. Th<strong>an</strong>ks to<br />

Environment C<strong>an</strong>ada for ongoing support with field logistics <strong>an</strong>d additional data. We th<strong>an</strong>k Sarah<br />

Kopczynski <strong>an</strong>d Sh<strong>an</strong>non Haight for <strong>the</strong>ir <strong>as</strong>sist<strong>an</strong>ce collecting field data. We are also th<strong>an</strong>kful<br />

for helpful comments provided by Andrew Klein <strong>an</strong>d two <strong>an</strong>onymous reviewers. Additional<br />

th<strong>an</strong>ks to Michael Chupa. Funding w<strong>as</strong> provided by Lehigh University <strong>an</strong>d <strong>the</strong> National<br />

Aeronautics <strong>an</strong>d Space Administration’s Terrestrial Hydrology Program (gr<strong>an</strong>t # NNG04GR31G).<br />

REFERENCES<br />

Aagaard K, Carmack EC. 1989. The role of sea ice <strong>an</strong>d o<strong>the</strong>r freshwater in <strong>the</strong> Arctic circulation.<br />

Journal of Geophysical Research 94: 14485-14498.<br />

Abdalati W, Steffen K 1997. <strong>Snow</strong>melt on <strong>the</strong> Greenl<strong>an</strong>d Ice Sheet <strong>as</strong> derived from p<strong>as</strong>sive<br />

microwave satellite data. Journal of Climate 10: 165-175.<br />

Ashcroft P, Wentz F. 2004 updated daily. AMSR-E/Aqua L2A Global swath spatially-resampled<br />

brightness temperatures (Tb) V001. J<strong>an</strong>uary 2004 to December 2005. Boulder, CO, USA:<br />

National <strong>Snow</strong> <strong>an</strong>d Ice Data Center. Digital media.<br />

Brabets TP, W<strong>an</strong>g B, Meade RH. 2000. Environmental <strong>an</strong>d hydrologic overview of <strong>the</strong> Yukon<br />

River b<strong>as</strong>in, Al<strong>as</strong>ka <strong>an</strong>d C<strong>an</strong>ada. USGS Water Resources Investigations Report 99-4204,<br />

Anchorage.<br />

Denoth A. 1989. <strong>Snow</strong> dielectric me<strong>as</strong>urements. Adv<strong>an</strong>ces in Space Research 9: 233-243.<br />

Derksen C, Walker AE, Goodison BE, Strapp JW. 2005. Integrating In Situ <strong>an</strong>d multiscale p<strong>as</strong>sive<br />

microwave data for estimation of subgrid scale snow water equivalent distribution <strong>an</strong>d<br />

variability. IEEE Tr<strong>an</strong>sactions on Geoscience <strong>an</strong>d Remote Sensing 43: 960-972.<br />

Environment Yukon. 2000. 90 Meter Yukon Digital Elevation Model.<br />

http://www.environmentyukon.gov.yk.ca/geomatics/data/90m_dem.html [13 March 2005].<br />

Foster JL, Chaojiao S, Walker JP, Kelly RE, Ch<strong>an</strong>g A, Dong J, Powell H. 2005. Qu<strong>an</strong>tifying <strong>the</strong><br />

uncertainty in p<strong>as</strong>sive microwave snow water equivalent observations. Remote Sensing of<br />

Environment 94: 187-203.<br />

Goita K, Walker AE, Goodison BE. 2003. Algorithm development for <strong>the</strong> estimation of snow<br />

water equivalent in <strong>the</strong> boreal forest using p<strong>as</strong>sive microwave data. International Journal of<br />

Remote Sensing 24: 1097-1102.<br />

Josberger EG, Mognard NM. 2002. A p<strong>as</strong>sive microwave snow depth algorithm with a proxy for<br />

snow metamorphism. Hydrological Processes 16: 1557-1568.<br />

151


Kelly RE, Ch<strong>an</strong>g AT, Ts<strong>an</strong>g L, Foster JL. 2003. A prototype AMSR-E global snow area <strong>an</strong>d snow<br />

depth algorithm. IEEE Tr<strong>an</strong>sactions on Geoscience <strong>an</strong>d Remote Sensing 41: 230-242.<br />

Kim EJ, Engl<strong>an</strong>d AW. 2003. A yearlong comparison of plot-scale <strong>an</strong>d satellite footprint-scale 19<br />

<strong>an</strong>d 37 GHz brightness of <strong>the</strong> Al<strong>as</strong>k<strong>an</strong> North Slope. Journal of Geophysical Research 108:<br />

ACL 7.<br />

Mätzler C. 1994. P<strong>as</strong>sive microwave signatures of l<strong>an</strong>dscapes in winter. Meteorology <strong>an</strong>d<br />

Atmospheric Physics 54: 241-260.<br />

Ramage JM, Isacks BL. 2002. Determination of melt-onset <strong>an</strong>d refreeze timing on sou<strong>the</strong><strong>as</strong>t<br />

Al<strong>as</strong>k<strong>an</strong> icefields using SSM/I diurnal amplitude variations. Annals of Glaciology 34: 391-398.<br />

Ramage JM, Isacks BL. 2003. Inter<strong>an</strong>nual variations of snowmelt <strong>an</strong>d refreeze timing on<br />

sou<strong>the</strong><strong>as</strong>t-Al<strong>as</strong>k<strong>an</strong> icefields, U.S.A. Journal of Glaciology 49: 102-116.<br />

Ramage JM, McKenney RA, Thorson B, Maltais P, Kopczynski SE. 2006. Relationship between<br />

p<strong>as</strong>sive microwave-derived snowmelt <strong>an</strong>d surface-me<strong>as</strong>ured discharge, Wheaton River, Yukon<br />

Territory, C<strong>an</strong>ada. Hydrological Processes 20: 689-704.<br />

Ulaby FT, Moore RK, Fung AK. 1986. Microwave Remote Sensing: Active <strong>an</strong>d P<strong>as</strong>sive, Vol. III:<br />

From Theory to Applications. Artech House: Dedham.<br />

Vorosmarty CJ, Hinzm<strong>an</strong> LD, Peterson BJ, Bromwich DH, Hamilton LC, Morison J,<br />

Rom<strong>an</strong>ovsky VE, Sturm M, Webb RS. 2001. The Hydrologic Cycle <strong>an</strong>d its Role in Arctic <strong>an</strong>d<br />

Global Environmental Ch<strong>an</strong>ge: A Rationale <strong>an</strong>d Strategy for Syn<strong>the</strong>sis Study. Arctic Research<br />

Consortium of <strong>the</strong> U.S.: Fairb<strong>an</strong>ks, Al<strong>as</strong>ka; 84.<br />

Walker AE, Goodison BE. 1993. Discrimination of a wet snowcover using p<strong>as</strong>sive microwave<br />

satellite data. Annals of Glaciology 17: 307-311.<br />

W<strong>an</strong>g L, Sharp M, Brown R, Derksen C, Rivard B. 2005. Evaluation of spring snow covered area<br />

depletion in <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Arctic from NOAA snow charts. Remote Sensing of Environment 95:<br />

453-463.<br />

152


ABSTRACT<br />

153<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Combination of Active <strong>an</strong>d P<strong>as</strong>sive Microwave<br />

to Estimate <strong>Snow</strong>pack Properties in Great Lakes Area<br />

AMIR E. AZAR 1 , TARENDRA LAKHANKAR 1 ,<br />

NARGES SHAHROUDI 1 , AND REZA KHANBILVARDI 1<br />

In this research we examine active <strong>an</strong>d p<strong>as</strong>sive microwave to snow water equivalent (SWE) <strong>an</strong>d<br />

to investigate <strong>the</strong> potential of combining active <strong>an</strong>d p<strong>as</strong>sive microwaves to improve <strong>the</strong> estimation<br />

of SWE. The study area is located in Great Lakes area between <strong>the</strong> latitudes of 41N–49N <strong>an</strong>d <strong>the</strong><br />

longitudes of 87W–98W. P<strong>as</strong>sive microwave are obtained from DMSP SSM/I sensors provided by<br />

NSIDC. Active microwave were obtained from different sensors: 1) RADARSAT C-B<strong>an</strong>d SAR.<br />

2) QuikSCAT Ku-b<strong>an</strong>d (13.4GHz) for both vertical <strong>an</strong>d horizontal polarizations. The ground truth<br />

data w<strong>as</strong> obtained from SNODAS data set produced by NOHRSC. An Artificial Neural Network<br />

model w<strong>as</strong> defined to model various combinations of inputs to SWE. The results indicate that<br />

none of <strong>the</strong> active microwave ch<strong>an</strong>nels produce satisfactory results. However, when combined<br />

with p<strong>as</strong>sive microwave, <strong>the</strong>y improve <strong>the</strong> estimated SWE.<br />

INTRODUCTION<br />

Microwave remote sensing techniques have been effective for monitoring snowpack parameters<br />

(snow extend, depth, water equivalent, wet/dry state). <strong>Snow</strong> parameters are extremely import<strong>an</strong>t<br />

for input to hydrological models for underst<strong>an</strong>ding ch<strong>an</strong>ges in climate due to global warming.<br />

<strong>Snow</strong> parameters been investigated by numerous researchers using m<strong>an</strong>y sensors such <strong>as</strong> SMMR<br />

<strong>an</strong>d SSMI for p<strong>as</strong>sive microwave <strong>an</strong>d SAR <strong>an</strong>d QSCAT for active microwave. Space-borne<br />

microwave sensors c<strong>an</strong> monitor characteristics of se<strong>as</strong>onal snow cover at high latitudes regardless<br />

of lighting conditions, time of <strong>the</strong> day, <strong>an</strong>d vegetation.<br />

In p<strong>as</strong>sive microwave radiometer, microwave energy emitted from <strong>the</strong> ground surface is<br />

tr<strong>an</strong>smitted through <strong>the</strong> snow layer into <strong>the</strong> atmosphere <strong>an</strong>d recorded by <strong>the</strong> sensor. <strong>Snow</strong><br />

parameters c<strong>an</strong> be extracted from remote sensing data by empirical algorithms. Hallikainen (1984)<br />

introduced his algorithm for estimating SWE using p<strong>as</strong>sive microwave SMMR data. The process<br />

involved <strong>the</strong> subtraction vertical polarizations of 18 <strong>an</strong>d 37 GHz frequencies. The subtracted<br />

value, dT, w<strong>as</strong> used to define linear relationships between dT <strong>an</strong>d SWE. Ch<strong>an</strong>g et al. (1987)<br />

proposed using <strong>the</strong> difference between <strong>the</strong> horizontally polarized ch<strong>an</strong>nels SMMR 37 GHz <strong>an</strong>d 18<br />

GHz to derive snow depth – brightness temperature relationship for a uniform snow field (Ch<strong>an</strong>g<br />

et al 1987). Goodison <strong>an</strong>d Walker (1995) introduced <strong>the</strong> most widely used algorithm for North<br />

America. The algorithm w<strong>as</strong> originally for C<strong>an</strong>adi<strong>an</strong> prairies. It defines a linear relationship<br />

between GTV ([37V–19V]/18) <strong>an</strong>d SWE. They also suggested using 37H <strong>an</strong>d 37H polarization<br />

differences for identifying wet snow. Derksen et al. (2004) developed a new algorithm which<br />

derives SWE for open environments, deciduous, coniferous, <strong>an</strong>d spars forest cover [SWE =<br />

FDSWED + FC SWEC + FS SWES + FOSWEO]. The algorithm represents <strong>an</strong> improvement, however<br />

1 NOAA-CREST, City University of NY, 137thst & Convent Ave. New York, NY.


154<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

still underestimates SWE in densely forested are<strong>as</strong>. Tedesco et al. (2004) developed <strong>an</strong>d tested <strong>an</strong><br />

inversion technique for retrieval of SWE <strong>an</strong>d dry snow depths b<strong>as</strong>ed on artificial neural networks<br />

(ANN) by using 19- <strong>an</strong>d 37-GHz SSM/I me<strong>as</strong>ured brightness temperatures.<br />

Hallikainen et al 2003 combined active (QuikSCAT/SeaWinds) <strong>an</strong>d p<strong>as</strong>sive (SSMI/DMSP) data<br />

for monitoring key snow parameters in Finl<strong>an</strong>d. The results show that combined active <strong>an</strong>d<br />

p<strong>as</strong>sive microwave sensors provide useful diurnal <strong>an</strong>d se<strong>as</strong>onal information. These results are<br />

more accurate th<strong>an</strong> those obtained by only p<strong>as</strong>sive microwave. In <strong>an</strong>o<strong>the</strong>r research Hallikainen<br />

showed that using space borne scatterometer (QuikSCAT onboard SeaWinds) for dry snow<br />

conditions, <strong>the</strong> backscattering coefficient incre<strong>as</strong>es with incre<strong>as</strong>ing SWE. For wet snow condition<br />

backscattering coefficient decre<strong>as</strong>es with incre<strong>as</strong>ing SWE. Ku-b<strong>an</strong>d scatterometer were used<br />

successfully to determine <strong>the</strong> onset <strong>an</strong>d <strong>the</strong> end of snow melt, <strong>an</strong>d to derive time series for <strong>the</strong><br />

fraction of snow-free ground during <strong>the</strong> se<strong>as</strong>onal snow melt period (Hallikainen et al 2004).<br />

Syn<strong>the</strong>tic Aperture Radar (SAR) particularly C-b<strong>an</strong>d SAR h<strong>as</strong> shown <strong>the</strong> potential for<br />

monitoring snow <strong>an</strong>d ice for more th<strong>an</strong> two decades. The high spatial resolution <strong>an</strong>d <strong>the</strong><br />

independence of <strong>the</strong> sensors from sun illumination <strong>an</strong>d cloud cover make SAR <strong>an</strong> ideal tool for<br />

snow studies. Launched in 1995, Radarsat-1 offers spatial resolutions between 10m to 100m <strong>an</strong>d a<br />

swath up to 500km. To estimate SWE using C-b<strong>an</strong>d SAR, Bernier et al. (1998) introduced <strong>an</strong><br />

approach b<strong>as</strong>ed on <strong>the</strong> fact that snow cover characteristics influence <strong>the</strong> underlying soil. The snow<br />

influence on soil temperature affects <strong>the</strong> dielectric properties of <strong>the</strong> soil which h<strong>as</strong> a major role on<br />

<strong>the</strong> backscattered signal. To recover <strong>the</strong> SWE from SAR data <strong>an</strong> algorithm made of two equations<br />

w<strong>as</strong> used. The first equation defines a linear relationship between <strong>the</strong> snow <strong>the</strong>rmal resist<strong>an</strong>ce <strong>an</strong>d<br />

<strong>the</strong> backscattering ratio between a winter image <strong>an</strong>d a reference (snow-free) image in DB. The<br />

snow-free image helps to eliminate <strong>the</strong> radiometric distortion due to topography <strong>as</strong> well <strong>as</strong> to<br />

minimize <strong>the</strong> effect of soil roughness on <strong>the</strong> signal. The second equation is a linear relationship<br />

between <strong>the</strong>rmal resist<strong>an</strong>ce <strong>an</strong>d <strong>the</strong> SWE. To estimate SWE from <strong>the</strong>rmal resist<strong>an</strong>ce <strong>the</strong> me<strong>an</strong><br />

density of <strong>the</strong> snowpack h<strong>as</strong> to be derived. This approach h<strong>as</strong> been applied for cold winter<br />

conditions <strong>an</strong>d dry snow (Bernier et al. 1999). The critical variables influencing <strong>the</strong> algorithm are<br />

variety of l<strong>an</strong>d cover, specifically forest density, <strong>Snow</strong>pack properties (depth>2m), <strong>an</strong>d severe<br />

topography. In a research on p<strong>as</strong>sive <strong>an</strong>d active airborne microwave remote sensing of snow cover<br />

Sokol et al. (2003) showed that SAR sensors are highly sensitive to ch<strong>an</strong>ges in <strong>the</strong> dielectric<br />

const<strong>an</strong>t <strong>an</strong>d have better spatial resolution th<strong>an</strong> <strong>the</strong>ir p<strong>as</strong>sive counterparts. They concluded that<br />

p<strong>as</strong>sive techniques estimate SWE most accurately under dry snow conditions with minimal<br />

stratified snow structures (Sokol et al. 2003).<br />

The focus of this research is estimating <strong>Snow</strong> Water Equivalent (SWE) in Great Lakes area by<br />

using active <strong>an</strong>d p<strong>as</strong>sive microwaves. Different approaches were examined for SWE estimations<br />

by RADARSAT SAR <strong>an</strong>d also QuikSCAT-Ku along <strong>an</strong>d p<strong>as</strong>sive SSM/I.<br />

STUDY AREA<br />

It is also located on <strong>the</strong> tr<strong>an</strong>sitional zone for snow me<strong>an</strong>ing that <strong>the</strong> nor<strong>the</strong>rn part of <strong>the</strong> study<br />

area is covered by snow for <strong>the</strong> whole winter se<strong>as</strong>on however for <strong>the</strong> sou<strong>the</strong>rn part <strong>the</strong>re is a<br />

pattern of snow-fall <strong>an</strong>d snow melt within <strong>the</strong> se<strong>as</strong>on. In addition to snow pattern, <strong>the</strong> l<strong>an</strong>d cover<br />

type varies a wide r<strong>an</strong>ge including, Evergreen Needle leaf forest, Deciduous Broadleaf forest,<br />

cropl<strong>an</strong>d, woodl<strong>an</strong>d <strong>an</strong>d dry l<strong>an</strong>d (Figure 1).


DATA USED<br />

SSM/I &<br />

QuikSCAT<br />

Coverage<br />

LANDCOVER<br />

Figure 1. Study area <strong>an</strong>d SSM/I <strong>an</strong>d RADARSAT coverage<br />

155<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

RADARSAT<br />

Images<br />

SSM/I<br />

SSM/I p<strong>as</strong>sive microwave radiometer with seven ch<strong>an</strong>nels is operating at five frequencies (19,<br />

35, 22, 37, <strong>an</strong>d 85.5 GHz) <strong>an</strong>d dual- polarization (except at 22 GHz which is V-polarization<br />

only). The sensor’s spatial resolution varies for different ch<strong>an</strong>nel frequencies. In this study we<br />

used Scalable Equal Area Earth Grid EASE-Grid SSM/I products distributed by National <strong>Snow</strong><br />

<strong>an</strong>d Ice Data Center (NSIDC). EASE-Grid spatial resolution is slightly more th<strong>an</strong> 25km (25.06)<br />

for all <strong>the</strong> ch<strong>an</strong>nels (NSIDC) although <strong>the</strong> recorded resolution of <strong>the</strong> sensor for longer wavelengths<br />

is more th<strong>an</strong> 50km. The three EASE-Grid projections comprise two azimuthal equal-area<br />

projections for <strong>the</strong> Nor<strong>the</strong>rn or Sou<strong>the</strong>rn hemispheres, respectively <strong>an</strong>d a global cylindrical equal<br />

area projection. In our study we used a Nor<strong>the</strong>rn hemisphere azimuthal equal-area. The study area<br />

is covered by 980 (28 by 35) SSM/I EASE-Grid pixels.<br />

Ku-B<strong>an</strong>d<br />

The QuikSCAT/SeaWinds scatterometer provides normalized radar cross section me<strong>as</strong>urements<br />

of <strong>the</strong> Earth’s surface at unprecedented coverage <strong>an</strong>d resolution. The QuikSCAT sensor on <strong>the</strong><br />

SeaWinds satellite operates at 13.4 GHz vertical <strong>an</strong>d horizontal ch<strong>an</strong>nels. The sigma (0) browse<br />

product of QuikSCAT h<strong>as</strong> <strong>the</strong> grid size of 5 pixels per degree or about 22.5km at <strong>the</strong> equator. In<br />

order to match SSM/I grid size <strong>the</strong> QuikSCAT images were averaged to 25km resolution for <strong>the</strong><br />

study area.<br />

Normalized Difference Vegetation Index (NDVI)<br />

NDVI is used to represent <strong>the</strong> variety of l<strong>an</strong>d cover in <strong>the</strong> study area. The NDVI data obtained<br />

from <strong>the</strong> NOAA/NASA Pathfinder AVHRR is distributed at Goddard Space Flight Center<br />

(GSFC). The spatial resolution is 8km by 8km obtained within a 10 day period that h<strong>as</strong> <strong>the</strong> fewest<br />

cloud. To match <strong>the</strong> with RADARSAT images, NDVI image w<strong>as</strong> resampled <strong>an</strong>d projected to<br />

UTM.


156<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

RADARSAT Images<br />

RADARSAT Sc<strong>an</strong>SAR images were obtained for February 06, 2003 (winter image), February<br />

02, 2006, <strong>an</strong>d May 01, 2002 (snow-free image) <strong>as</strong> show in Table 1. Sc<strong>an</strong>SAR images (500km by<br />

500km) have <strong>the</strong> nominal spatial resolution of 100m. The Sc<strong>an</strong>SAR products currently offered by<br />

us do not come in a map-projected format. However, images have <strong>the</strong> geo-referencing information<br />

contained in <strong>the</strong> CEOS format. This information is derived from <strong>the</strong> satellite orbit (ephemeris) <strong>an</strong>d<br />

is typically accurate to 100–200 meters, depending on beam mode <strong>an</strong>d <strong>the</strong> topography. To reduce<br />

geometric distortions caused by radar sensor viewing geometry satellite movement, earth<br />

curvature <strong>an</strong>d rotation, both RADARSAT images were registered to a L<strong>an</strong>dsat image of <strong>the</strong> study<br />

area. More th<strong>an</strong> 15 Ground Control Points <strong>an</strong>d a second order model <strong>an</strong>d nearest neighbor<br />

resampling mode were used to register <strong>the</strong> RADARSAT images. The images were projected to a<br />

UTM projection <strong>an</strong>d subseted for <strong>the</strong> in-common area of coverage.<br />

Table 1. RADARSAT images for Great Lake area<br />

Image Date Time(UTC) Inc Angle NW NE SW SE Center<br />

Winter 2/2/2006 23:47 34.26<br />

48 51 N<br />

94 13 W<br />

49 46 N<br />

86 37 W<br />

43 43 N<br />

92 27 W<br />

44 37 N<br />

85 33 W<br />

46 44 N<br />

89 42 W<br />

Winter 2/6/2004 23:53 34.26<br />

48 26 N<br />

94 47 W<br />

49 17 N<br />

87 46 W<br />

43 50 N<br />

93 17 W<br />

44 40 N<br />

86 50 W<br />

46 53 N<br />

90 44 W<br />

<strong>Snow</strong>-free 5/1/2003 23:49 34.26<br />

48 37N<br />

93 47 W<br />

49 27 N<br />

86 45 W<br />

44 00 N<br />

92 17 W<br />

44 51 N<br />

85 49 W<br />

46 47 N<br />

89 40 W<br />

Ground Truth Data<br />

NOAA National Wea<strong>the</strong>r Service's National Operational Hydrologic Remote Sensing Center<br />

(NOHRSC) started producing SNOw Data Assimilation System (SNODAS), beginning 1 October<br />

2003. SNODAS includes <strong>an</strong>d procedures to <strong>as</strong>similate airborne gamma radiation <strong>an</strong>d groundb<strong>as</strong>ed<br />

observations of snow covered area <strong>an</strong>d snow water equivalent, downscaled output from<br />

Numerical Wea<strong>the</strong>r Prediction (NWP) models combined in a physically b<strong>as</strong>ed, spatially<br />

distributed energy- <strong>an</strong>d m<strong>as</strong>s-bal<strong>an</strong>ce model. The output product h<strong>as</strong> 1km spatial <strong>an</strong>d hourly<br />

temporal resolution.<br />

METHODOLOGY AND MODEL<br />

The objective of this study is to estimate snow depth <strong>an</strong>d SWE using active <strong>an</strong>d p<strong>as</strong>sive<br />

microwave images. However, active microwave RADARSAT SAR <strong>an</strong>d p<strong>as</strong>sive microwave SSM/I<br />

are totally different in <strong>the</strong>ir nature <strong>an</strong>d applications. RADARSAT images have high spatial<br />

resolution <strong>an</strong>d are suitable for regional studies. On <strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d, low spatial resolution <strong>an</strong>d high<br />

temporal resolution of SSM/I make it suitable for studies on a global scale. In order to consider<br />

above differences, each of <strong>the</strong> active <strong>an</strong>d p<strong>as</strong>sive data were <strong>an</strong>alyzed separately to investigate <strong>the</strong>ir<br />

potential to estimate snow parameters. This section focuses on: 1) Using high resolution active<br />

microwave RADARSAT SAR to estimate SWE. This section focuses on how to process<br />

RADARSAT images <strong>an</strong>d corresponding ground truth data along with suggesting modeling<br />

approaches to improve <strong>the</strong> estimations. 2) Using low resolution p<strong>as</strong>sive SSM/I <strong>an</strong>d active<br />

QuikSCAT with statistical b<strong>as</strong>ed models. Also, <strong>the</strong> improvement by combining various data types<br />

w<strong>as</strong> qu<strong>an</strong>tified.<br />

An adaptive network is a network structure that consists of a number of nodes (neurons)<br />

connected through directional links. Each node represents a process unit, <strong>an</strong>d <strong>the</strong> links specify<br />

c<strong>as</strong>ual relationship between <strong>the</strong> connected nodes. Nodes are adaptive me<strong>an</strong>ing that <strong>the</strong> outputs of<br />

<strong>the</strong>se nodes depend on modifiable parameters pertaining to <strong>the</strong>se nodes. The learning rule specifies<br />

how <strong>the</strong>se parameters should be updated to minimize error which is discrep<strong>an</strong>cy between <strong>the</strong><br />

networks actual output <strong>an</strong>d desired one. In our study we used a feed forward backpropagation


157<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

model. The network h<strong>as</strong> two hidden layers with ten nodes at each layer. To train <strong>the</strong> network data<br />

were divided into three sets (training, validation, <strong>an</strong>d test). The model testing is <strong>the</strong> process by<br />

which <strong>the</strong> input vectors from input/output data sets on which <strong>the</strong> network is not trained, are<br />

presented to <strong>the</strong> trained model, to see how well <strong>the</strong> ANN model predicts <strong>the</strong> corresponding data<br />

set output values. The o<strong>the</strong>r type of validation which is also referred <strong>as</strong> checking data set is used to<br />

control <strong>the</strong> potential for <strong>the</strong> model over fitting <strong>the</strong> data. In principle, <strong>the</strong> model error for <strong>the</strong><br />

checking data set tends to decre<strong>as</strong>e <strong>as</strong> <strong>the</strong> training takes place up to <strong>the</strong> point that overfitting<br />

begins, <strong>an</strong>d <strong>the</strong>n <strong>the</strong> model error for checking data suddenly incre<strong>as</strong>es.<br />

High Resolution Active Microwave RADARSAT SAR<br />

Four different approaches for ANN input data were considered: A. Input consists of <strong>the</strong> only<br />

backscattering ration at 25km resolution. B. Input includes NDVI data in addition to<br />

backscattering both in 25km resolution. C. Input includes NDVI data in addition to backscattering<br />

with modified ANN training. D. Input includes NDVI <strong>an</strong>d Backscattering in 5km averaged<br />

resolution.<br />

A. Input consists of <strong>the</strong> only backscattering ration at 1km resolution<br />

In <strong>the</strong> first approach, <strong>the</strong> backscattering ratio w<strong>as</strong> used <strong>as</strong> <strong>the</strong> input for <strong>the</strong> ANN model.<br />

Figure 5.8 shows <strong>the</strong> spatial variation of <strong>the</strong> backscattering ratio while <strong>the</strong> water bodies are filtered<br />

out of <strong>the</strong> image. The higher backscattering ratio is detected in high latitudes <strong>an</strong>d around <strong>the</strong> lake.<br />

The scatter plot of backscattering versus <strong>the</strong> SWE indicates low correlation.<br />

1.5<br />

0<br />

-2<br />

Figure 2. Spatial variation of backscattering ratio <strong>an</strong>d SWE in 1km resolution, February 06, 04<br />

R=0.22<br />

Figure 3. Backscattering vs. SWE, February 06, 04


158<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

The ANN model w<strong>as</strong> trained using backscattering ratio <strong>as</strong> input <strong>an</strong>d SWE <strong>the</strong> output of <strong>the</strong><br />

model. The simulated SWE b<strong>as</strong>ed on <strong>the</strong> backscattering ratio for February 06, 2004 is shown in<br />

Figure 4. Both <strong>the</strong> produced image <strong>an</strong>d <strong>the</strong> corresponding scatter plot show unsatisfactory results.<br />

The model output is highly underestimated <strong>an</strong>d show very low correlation coefficient. Introducing<br />

l<strong>an</strong>d cover characteristics c<strong>an</strong> be helpful to incre<strong>as</strong>e <strong>the</strong> accuracy of <strong>the</strong> model.<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

SWE(mm)<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Scatter Plot of ANN Output <strong>an</strong>d SNODAS<br />

R=0.34695<br />

RMSE=40mm<br />

0<br />

0 20 40 60 80 100 120 140 160 180 200<br />

SWE Estimated by ANN (mm)<br />

Figure 4. Spatial variation of SWE for February 06, 04 <strong>an</strong>d <strong>the</strong> scatter plot of model output vs. ground truth<br />

SWE<br />

B. Input includes NDVI data in addition to backscattering both in 1km resolution<br />

In order to introduce <strong>the</strong> l<strong>an</strong>d cover characteristics in a qu<strong>an</strong>titative way <strong>the</strong> NDVI image of<br />

<strong>the</strong> study area w<strong>as</strong> added <strong>as</strong> <strong>an</strong> input to <strong>the</strong> ANN model. The model w<strong>as</strong> trained <strong>an</strong>d validated<br />

using <strong>the</strong> data from two winter days of February 06, 2004 <strong>an</strong>d February 02, 2006. The simulated<br />

SWE for February 06, 2004 is illustrated (Fig 5).<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

SWE(mm)<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Scatter Plot of ANN Output <strong>an</strong>d SNODAS<br />

R=0.471<br />

RMSE=38mm<br />

0<br />

0 20 40 60 80 100 120 140 160 180 200<br />

SWE Estimated by ANN (mm)<br />

Figure 5. Simulated SWE b<strong>as</strong>ed on NDVI <strong>an</strong>d Backscattering for February 06, 04 <strong>an</strong>d <strong>the</strong> scatter plot of<br />

model output vs. ground truth SWE<br />

The above results show <strong>the</strong> incre<strong>as</strong>e of correlation by adding <strong>the</strong> NDVI to <strong>the</strong> ANN model. On<br />

<strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d, <strong>the</strong> problem with underestimation of SWE still exists. As shown in <strong>the</strong> scatter plot,<br />

for SWE varying between 0 to 250mm <strong>the</strong> model output mostly varies between 50 to 150mm. This<br />

problem might originate from <strong>the</strong> training process considering <strong>the</strong> fact that <strong>the</strong>re are more pixels<br />

with low values of snow th<strong>an</strong> high values. This forces <strong>the</strong> training towards <strong>the</strong> low values in order<br />

to minimize <strong>the</strong> RMSE of <strong>the</strong> total estimation. To minimize <strong>the</strong> influence of number of <strong>the</strong> pixels<br />

with different values for SWE, <strong>the</strong> network w<strong>as</strong> trained with same number of pixels from each<br />

cl<strong>as</strong>s <strong>as</strong> <strong>the</strong> third approach.


159<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

C. Input includes NDVI data in addition to backscattering with modified ANN training<br />

In <strong>the</strong> third approach <strong>the</strong> pixels were divided to four cl<strong>as</strong>ses b<strong>as</strong>ed on <strong>the</strong> SWE values<br />

(SWE


160<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

image. On <strong>the</strong> o<strong>the</strong>r h<strong>an</strong>d, <strong>the</strong> backscattering ratio trend is almost const<strong>an</strong>t. Also, <strong>the</strong> r<strong>an</strong>ge for <strong>the</strong><br />

ratio is highly different for years 2004 <strong>an</strong>d 2006. This limits <strong>an</strong>y kind of modeling for estimating<br />

SWE from backscattering. The comparison of <strong>the</strong> scatter plots of SWE versus backscattering is<br />

shown in Figure 9.<br />

Backscattering<br />

SWE (mm)<br />

2004 2006<br />

Backscattering<br />

2004 2006<br />

SWE (mm)<br />

Figure 8. Variation of backscattering ratio <strong>an</strong>d SWE for different parts of <strong>the</strong> study area<br />

SWE (mm) SWE (mm)<br />

2004<br />

Backscattering Ratio (DB) Backscattering Ratio (DB)<br />

Backscattering Ratio (DB) Backscattering Ratio (DB)<br />

SWE (mm) SWE (mm)<br />

2004<br />

2006 2006<br />

Figure 9. Scatter plots of Backscattering vs. SWE at 5km resolution for two images (Feb 06, 04 & Feb 02,<br />

06)


161<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

The left scatter plots representing <strong>the</strong> complete dat<strong>as</strong>et <strong>an</strong>d <strong>the</strong> right ones are modified by<br />

eliminating <strong>the</strong> backscattering ratio higher th<strong>an</strong> zero since only few percentage points are points<br />

have backscattering values higher th<strong>an</strong> zero. This could be <strong>the</strong> effect urb<strong>an</strong> are<strong>as</strong> that have high<br />

backscattering. The new scatter plots are shown in Figure 9. These scatter plots clearly show<br />

where SWE is in <strong>the</strong> r<strong>an</strong>ge of 10mm to 60mm <strong>the</strong> backscattering r<strong>an</strong>ge is very large. Also, <strong>the</strong><br />

backscattering r<strong>an</strong>ge between two years (2004 <strong>an</strong>d 2006) is very different. This indicates that<br />

developing a model to estimate SWE using backscattering ratio is very difficult. A solution for<br />

this problem is modifying <strong>the</strong> model for each year.<br />

Low Resolution Active QuikSCAT <strong>an</strong>d P<strong>as</strong>sive SSM/I<br />

A. Evaluation of using NDVI <strong>an</strong>d QuikSCAT in SWE estimation<br />

A feed forward backpropagation neural network model with 2 hidden layers, 20 neurons each<br />

layer, w<strong>as</strong> developed. The output of <strong>the</strong> model consists of ground truth SWE data from NOHRSC.<br />

In order to evaluate <strong>the</strong> effect SWE on various microwave ch<strong>an</strong>nels different combination of<br />

inputs were used. First, <strong>the</strong> input consisted four of SSM/I ch<strong>an</strong>nels (19V, 19H, 37V, <strong>an</strong>d 37H). Then,<br />

NDVI <strong>an</strong>d SSM/I were introduced <strong>as</strong> <strong>the</strong> input to <strong>the</strong> model. Finally, QuikSCAT-ku along with<br />

SSM/I <strong>an</strong>d NDVI were used <strong>as</strong> <strong>the</strong> input. For all above approaches <strong>the</strong> model w<strong>as</strong> validated by a<br />

dependent data. In o<strong>the</strong>r words, <strong>the</strong> training <strong>an</strong>d validation data were <strong>the</strong> same. Figure 10 shows<br />

<strong>the</strong> results for various approaches. SSM/I brightness temperatures have shown correlations with<br />

SWE to some extent. Adding NDVI to <strong>the</strong> input brings information about <strong>the</strong> l<strong>an</strong>d cover type for<br />

<strong>the</strong> ANN model <strong>an</strong>d incre<strong>as</strong>es <strong>the</strong> accuracy of <strong>the</strong> estimate. By adding active QuikSCAT to <strong>the</strong><br />

input we have <strong>an</strong> input of three independent data sets. The incre<strong>as</strong>e of correlation coefficients<br />

indicates that combining active <strong>an</strong>d p<strong>as</strong>sive using a neural network c<strong>an</strong> improve <strong>the</strong> SWE<br />

estimation. Figure 10 also indicates that combining SSM/I with QuikSCAT <strong>an</strong>d NDVI produces<br />

<strong>the</strong> best results. The sudden decre<strong>as</strong>e of correlation coefficients for days in February should be due<br />

to <strong>the</strong> existence of wet snow in parts of <strong>the</strong> study area which w<strong>as</strong> already explained in chapter for<br />

snow cover estimation.<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Dec<br />

6, 03<br />

SSM/I<br />

QSCAT<br />

SSM/I &NDVI<br />

SSM/I &QSCAT<br />

QSCAT & NDVI<br />

SSM/I & QSCAT & NDVI<br />

Dec<br />

13, 03<br />

Dec<br />

20, 03<br />

J<strong>an</strong><br />

4, 04<br />

J<strong>an</strong><br />

11, 04<br />

J<strong>an</strong><br />

18, 04<br />

J<strong>an</strong><br />

25, 4<br />

Feb<br />

1, 04<br />

Feb<br />

8, 04<br />

Feb<br />

23, 04<br />

Figure 10. Correlation coefficients (R 2 ) between estimated SWE <strong>an</strong>d <strong>the</strong> corresponding ground truth data<br />

(winter 2003–2004, model validated by dependent data)


162<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

B. Using Active <strong>an</strong>d P<strong>as</strong>sive to Estimate SWE<br />

The results above showed that adding NDVI <strong>an</strong>d QuikSCAT-ku incre<strong>as</strong>es <strong>the</strong> correlation<br />

between input of microwave ch<strong>an</strong>nels <strong>an</strong>d SWE. To investigate <strong>the</strong> capability of <strong>the</strong> model to<br />

estimate SWE, it w<strong>as</strong> examined to independent data. The approach consisted of training <strong>the</strong> model<br />

with <strong>the</strong> data from <strong>the</strong> days before <strong>the</strong> selected day for estimation. Table 2 describes <strong>the</strong> approach.<br />

The results for correlation coefficients <strong>an</strong>d RMSE in <strong>the</strong> table show <strong>an</strong> incre<strong>as</strong>ing trend for RMSE.<br />

This incre<strong>as</strong>e in <strong>the</strong> error originates from <strong>the</strong> incre<strong>as</strong>e in <strong>the</strong> average depth of snow during <strong>the</strong><br />

winter se<strong>as</strong>on. For correlation coefficients <strong>the</strong>re is <strong>an</strong> improvement in <strong>the</strong> beginning but it<br />

decre<strong>as</strong>es after J<strong>an</strong>uary 25, 2004. For <strong>the</strong> selected days in February especially Feb08, Feb16, <strong>an</strong>d<br />

Feb23, The error incre<strong>as</strong>es dramatically. This is due to <strong>the</strong> wet snow conditions for those days.<br />

Wet snow c<strong>an</strong> not be detected by p<strong>as</strong>sive SSM/I scattering ch<strong>an</strong>nels. This section is already<br />

discussed in snow cover section. Figures 11 <strong>an</strong>d 12 show <strong>the</strong> estimated snow for February 01,<br />

2004. It is observed that <strong>the</strong> model is incapable of detect <strong>an</strong>d estimating deep snow. The scatter<br />

plot of <strong>the</strong> ground truth versus <strong>the</strong> estimate (Fig 11) illustrates <strong>the</strong> results in a qu<strong>an</strong>titative way.<br />

The best fitted line (red line) is below <strong>the</strong> 1:1 line indicating underestimation of <strong>the</strong> estimate.<br />

Training Data<br />

(Days)<br />

Table 2: Estimating SWE by ANN model<br />

Validation<br />

Data (Day)<br />

Correlation<br />

Coe.<br />

RMSE Bi<strong>as</strong><br />

Dec06,Dec13 Dec 20 0.37 21 –14<br />

Dec06,<br />

Dec13,Dec20<br />

J<strong>an</strong>04 0.43 17 1<br />

Dec06,Dec13,<br />

Dec20, J<strong>an</strong>04<br />

J<strong>an</strong>11 0.47 19 5<br />

Dec13,Dec20,<br />

J<strong>an</strong>04,J<strong>an</strong>11<br />

J<strong>an</strong>18 0.44 30 –11<br />

Dec20,J<strong>an</strong>04,<br />

J<strong>an</strong>11,J<strong>an</strong>18<br />

J<strong>an</strong>25 0.53 45 –34<br />

J<strong>an</strong>04,J<strong>an</strong>11,<br />

J<strong>an</strong>18,J<strong>an</strong>25<br />

Feb01 0.47 44 –30<br />

J<strong>an</strong>18,J<strong>an</strong>25,<br />

Feb01<br />

Feb08 0.37 42 –31<br />

J<strong>an</strong>18,J<strong>an</strong>25<br />

Feb01,Feb08<br />

Feb16 0.12 75 –58<br />

J<strong>an</strong>25, Feb01,<br />

Feb08, Feb16<br />

Feb23 0 90 –83


180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

163<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Figure 11. Estimated SWE by ANN (left) <strong>an</strong>d SNODAS ground truth SWE (right), February 01, 2004<br />

CONCLUSION<br />

ANN output<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

R = 0.533<br />

0<br />

0 50 100<br />

Snocover(SNODAS)<br />

150 200<br />

Figure 12. Estimated SWE by ANN vs. SNODAS ground truth SWE, February 01, 2004<br />

Three RADARSAT images were obtained to investigate <strong>the</strong> potential of RADARSAT SAR in<br />

estimating SWE. The images were processed <strong>an</strong>d georefenced using PCIGeomatica. The Ground<br />

truth data were obtained from NOHRSC SNODAS dat<strong>as</strong>et through NSIDC. The backscattering<br />

ratio of RADARSAT images w<strong>as</strong> derived by subtracting <strong>the</strong>m from a reference image. The<br />

<strong>an</strong>alysis indicates that backscattering ratio h<strong>as</strong> limited correlation with SWE (20 percent). An<br />

ANN model w<strong>as</strong> used to explore non-linear relationships between backscattering ratio <strong>an</strong>d SWE.<br />

The results showed low correlation between estimated <strong>an</strong>d ground truth SWE. In order to<br />

introduce l<strong>an</strong>d cover characteristics, <strong>an</strong> NDVI image w<strong>as</strong> added to <strong>the</strong> input of <strong>the</strong> ANN model.<br />

The results showed a more th<strong>an</strong> 15 percent improvement in correlation coefficient. To improve <strong>the</strong><br />

estimation <strong>the</strong> input cl<strong>as</strong>sified b<strong>as</strong>ed on SWE values. This improved <strong>the</strong> r<strong>an</strong>ge of <strong>the</strong> estimated<br />

SWE although it did not ch<strong>an</strong>ge <strong>the</strong> correlation coefficients. Finally <strong>the</strong> resolution w<strong>as</strong> ch<strong>an</strong>ged to<br />

5km. It w<strong>as</strong> concluded that where SWE is in <strong>the</strong> r<strong>an</strong>ge of 10mm to 60mm <strong>the</strong> backscattering r<strong>an</strong>ge<br />

is very large. Also, <strong>the</strong> backscattering r<strong>an</strong>ge between two years (2004 <strong>an</strong>d 2006) is very different.<br />

This indicates that developing a model to estimate SWE using backscattering ratio is very<br />

difficult. A solution for this problem is modifying <strong>the</strong> model for each year.<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0


164<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

In c<strong>as</strong>e of low resolution p<strong>as</strong>sive SSM/I <strong>an</strong>d active QuikSCAT, <strong>the</strong> ANN model shows<br />

satisfactory result in dependent estimation of SWE. Also, adding QuikSCAT-Ku incre<strong>as</strong>ed <strong>the</strong><br />

accuracy of <strong>the</strong> estimated SWE by neural networks. It w<strong>as</strong> concluded that estimating SWE by<br />

neural networks is highly dependent on training data. This c<strong>an</strong> become a source of error on model<br />

development. In order to resolve this problem a large dat<strong>as</strong>et is necessary.<br />

ACKNOWLEDGEMENTS<br />

The Authors express <strong>the</strong>ir gratitude to Dr Jeff Hurley, a Senior Project M<strong>an</strong>ager in<br />

RADARSAT International, for his support in processing <strong>an</strong>d <strong>an</strong>alyzing RADARSAT images.<br />

REFERENCES<br />

Bernier M, Fortin J (1998).The Potential of Times Series pf C-B<strong>an</strong>d SAR Data to Monitor Dry<br />

<strong>an</strong>d Shallow <strong>Snow</strong> Cover. IEEE Tr<strong>an</strong>saction on Geoscinece <strong>an</strong>d remote sensing, 36(1): 226–<br />

242.<br />

Bernier M, Fortin J Y.Gauthier, R.Gauthier, R.Roy, <strong>an</strong>d P. Vincent (1999i).Determination of<br />

<strong>Snow</strong> Water Equivalent using RADARSAT SAR data in e<strong>as</strong>tern C<strong>an</strong>ada..Hydrological<br />

Processes,13:3041–3051.<br />

Ch<strong>an</strong>g, A. T. C, L. Foster, D. K. Hall. (1987). Nimbus-7 SMMR derived global snow cover<br />

parameters. Annals Glaciology, 9:39–44.<br />

Derksen, E. LeDrew, A.Walker, <strong>an</strong>d B. Goodison (2001). Evaluation of a Multi-Algorithm<br />

Approach to P<strong>as</strong>sive Microwave Monitoring of Central North Americ<strong>an</strong> <strong>Snow</strong> Water<br />

Equivalent. IEEE, 952–954.<br />

Hallikainen, M. T. (1984). Retrieval of snow water equivalent from Nimbus-7 SSMR data: effect<br />

of l<strong>an</strong>d cover categories <strong>an</strong>d wea<strong>the</strong>r conditions. IEEE Oce<strong>an</strong>ic Engineering, 9(5): 372–376.<br />

Hallikainene, M. T,P.Halme,M Takala, J.Pulliainen (2003). Combined Active <strong>an</strong>d P<strong>as</strong>sive<br />

Mirowave Remote Sensing of <strong>Snow</strong> in Finl<strong>an</strong>d. IEEE:830–832.<br />

Hallikainene, M. T,P.Halme,P.Lahtinen, M Takala, J.Pulliainen (2004). Retrieval of <strong>Snow</strong><br />

Characteristics from Scapeborne Scatteromter Data. IEEE:1849–1852.<br />

Sokol. J, T.J. Pultz, <strong>an</strong>d A.E. Walker (2003). P<strong>as</strong>sive <strong>an</strong>d Active microwave remote sensing of<br />

snow cover. INT. J. Remote Sensing, 24(24):5327–5344.<br />

Tedesco, J. Pulli<strong>an</strong>inen, M. Takala, M Hallikainen, P. Pampaloni. (2004). Artificial neural<br />

network-b<strong>as</strong>ed techniques for <strong>the</strong> retrieval of SWE <strong>an</strong>d snow depth from SSM/I data. Remote<br />

Sensing of Environment 90:76–85.<br />

Walker, G. B. E. a. A. E. (1995). C<strong>an</strong>adi<strong>an</strong> development <strong>an</strong>d use of snow cover information from<br />

p<strong>as</strong>sive microwave satellite data. P<strong>as</strong>sive microwave remote sensing of l<strong>an</strong>d–atmosphere<br />

interactions. B. J. Choudhury, Y. H. Kerr, E. G. Njoku <strong>an</strong>d P. Pampaloni. The Ne<strong>the</strong>rl<strong>an</strong>ds,<br />

VSP BV 245–262.


165<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Enh<strong>an</strong>cements <strong>an</strong>d Forthcoming Developments to <strong>the</strong> Interactive<br />

Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping System (IMS)<br />

ABSTRACT<br />

SEAN R. HELFRICH 1 , DONNA MCNAMARA 1 , BRUCE H. RAMSAY 2 ,<br />

THOMAS BALDWIN 1 , AND TIM KASHETA 3<br />

The national oce<strong>an</strong>ic <strong>an</strong>d atmospheric administration’s national environmental satellite data, <strong>an</strong>d<br />

information service (noaa/nesdis) interactive multisensor snow <strong>an</strong>d ice mapping system (ims) h<strong>as</strong><br />

undergone subst<strong>an</strong>tial ch<strong>an</strong>ges since <strong>the</strong> inception in 1997. These ch<strong>an</strong>ges include <strong>the</strong> data sources<br />

used to generate <strong>the</strong> product, methodology of product creation, <strong>an</strong>d even ch<strong>an</strong>ges in <strong>the</strong> output.<br />

Among <strong>the</strong> most notable of <strong>the</strong> p<strong>as</strong>t upgrades to <strong>the</strong> ims are a 4 kilometer resolution grid output,<br />

ingest of <strong>an</strong> automated snow detection algorithm, exp<strong>an</strong>sion to a global extent, <strong>an</strong>d a static digital<br />

elevation model for mapping b<strong>as</strong>ed on elevation. Fur<strong>the</strong>r developments to this dynamic system<br />

will continue <strong>as</strong> noaa strives to improve snow parameterization for wea<strong>the</strong>r forec<strong>as</strong>t modeling.<br />

Several future short term enh<strong>an</strong>cements will be evaluated for possible tr<strong>an</strong>sition to operations<br />

before <strong>the</strong> nor<strong>the</strong>rn hemisphere winter of 2006–07. Current <strong>an</strong>d historical data will be adobted to a<br />

geographic information systems (GIS) format before 2007, <strong>as</strong> well. Longer-term enh<strong>an</strong>cements are<br />

also pl<strong>an</strong>ned to account for new snow dat<strong>as</strong>ources, mapping methodologies <strong>an</strong>d user requirements.<br />

These modifications are being made with care to preserve <strong>the</strong> integrity of <strong>the</strong> long st<strong>an</strong>ding<br />

satellite derived snow record that is vital to global ch<strong>an</strong>ge detection.<br />

Keywords: satellite remote sensing; environmental data; snow cover; ice cover; geographic<br />

information systems; climate<br />

INTRODUCTION<br />

The interactive multisensor snow <strong>an</strong>d ice mapping system (IMS) h<strong>as</strong> been <strong>the</strong> main operational<br />

snow <strong>an</strong>d ice charting tool of <strong>the</strong> National Oce<strong>an</strong>ic <strong>an</strong>d Atmospheric Administration’s National<br />

Environmental Satellite Data, <strong>an</strong>d Information Service (NOAA/NESDIS) for almost a decade.<br />

This product w<strong>as</strong> primarily created to improve <strong>the</strong> quality <strong>an</strong>d timeliness of Nor<strong>the</strong>rn Hemisphere<br />

snow maps for National Centers for Environmental Prediction (NCEP) numerical forec<strong>as</strong>ting<br />

(Ramsay, 1998). Prior to <strong>the</strong> IMS’s operational inception in 1997, snow charts were constructed<br />

m<strong>an</strong>ually once per week. The IMS is produced daily using geographic information systems (GIS)<br />

technology. This system had subst<strong>an</strong>tial impacts on production speed, product spatial accuracy,<br />

<strong>an</strong>d time between observations. A comparison <strong>an</strong>d validation review of <strong>the</strong> product tr<strong>an</strong>sition from<br />

1<br />

NOAA/NESDIS/OSDPD – NOAA Science Center, 5200 Auth Rd., Camp Springs Maryl<strong>an</strong>d,<br />

20746 se<strong>an</strong>.helfrich@noaa.gov<br />

2<br />

NOAA/NESDIS/STAR/CORP – NOAA Science Center, 5200 Auth Rd., Camp Springs<br />

Maryl<strong>an</strong>d, 20746<br />

3<br />

Riverside Technology, Inc. – 2290 E<strong>as</strong>t Prospect Road, Suite 1, Fort Collins, Colorado 80525


m<strong>an</strong>ual weekly to IMS daily charts w<strong>as</strong> conducted between 1997 <strong>an</strong>d 1999. Preliminary<br />

examination of <strong>the</strong> data between <strong>the</strong>se periods suggests <strong>the</strong> IMS output to be superior to <strong>the</strong><br />

weekly m<strong>an</strong>ual snow charts (Ramsay, 2000). In June 1999, <strong>the</strong> m<strong>an</strong>ual charting of snow extent<br />

w<strong>as</strong> suppl<strong>an</strong>ted operationally with <strong>the</strong> daily IMS. Since <strong>the</strong> charts are now constructed digitally,<br />

<strong>the</strong>ir distribution h<strong>as</strong> incre<strong>as</strong>ed, with hundreds of known users viewing data each month from <strong>the</strong><br />

NESDIS site <strong>an</strong>d <strong>an</strong> unknown number of users obtaining <strong>the</strong> IMS data from o<strong>the</strong>r sources.<br />

While <strong>the</strong>re are potentially m<strong>an</strong>y uses, <strong>the</strong> primary function of <strong>the</strong> product is to provide<br />

cryospheric input for environmental modeling. There are two operational government customers<br />

for this product, <strong>the</strong> NCEP / Environmental Modeling Center (EMC) <strong>an</strong>d <strong>the</strong> NCEP / Climate<br />

Prediction Center (CPC). These customers help support <strong>an</strong>d influence <strong>the</strong> direction of <strong>the</strong> product.<br />

The feedback from <strong>the</strong> NCEP modeling agencies <strong>an</strong>d <strong>the</strong> preliminary NOAA Program<br />

Observational Requirements have led to adv<strong>an</strong>cements in <strong>the</strong> product <strong>an</strong>d point toward continued<br />

improvement. The EMC applies <strong>the</strong> models for each three hour modeling run for North America<br />

<strong>an</strong>d temporally coarser models for <strong>the</strong> <strong>entire</strong> pl<strong>an</strong>et. <strong>Snow</strong> plays <strong>an</strong> import<strong>an</strong>t role in model input<br />

<strong>an</strong>d c<strong>an</strong> lead to subst<strong>an</strong>tial error in forec<strong>as</strong>t results b<strong>as</strong>ed on incorrect representations of snow<br />

distribution, age, depth, snow water equivalent (SWE), <strong>an</strong>d snow density (Mitchell et al., 1993;<br />

Sheffeld et al., 2003).<br />

Along with serving <strong>as</strong> <strong>an</strong> initial state of <strong>the</strong> surface cryosphere for <strong>the</strong> Nor<strong>the</strong>rn Hemisphere for<br />

wea<strong>the</strong>r prediction, NOAA’s snow maps serve <strong>as</strong> a 40 year environmental monitoring record for<br />

hemispheric snow cover. This is considered <strong>the</strong> longest continuous satellite-derived record of <strong>an</strong>y<br />

environmental variable (Robinson et al., 1993). It is vital for climate ch<strong>an</strong>ge detection <strong>an</strong>d a key<br />

element in NOAA’s Mission Goals to “underst<strong>an</strong>d climate variability <strong>an</strong>d ch<strong>an</strong>ge to enh<strong>an</strong>ce<br />

society’s ability to pl<strong>an</strong> <strong>an</strong>d respond” (USDOC/NOAA, 2005). Given <strong>the</strong> import<strong>an</strong>ce of this<br />

record, ch<strong>an</strong>ges in <strong>the</strong> record should be considered with great care to preserve <strong>the</strong> integrity of <strong>the</strong><br />

product for climate monitoring. Consultation within <strong>the</strong> snow <strong>an</strong>d ice climate monitoring<br />

community h<strong>as</strong> been sought before <strong>the</strong> integration of ch<strong>an</strong>ges to safeguard <strong>the</strong> IMS’s<br />

environmental monitoring role.<br />

The IMS w<strong>as</strong> designed to allow meteorologists to chart snow cover interactively on a daily b<strong>as</strong>is<br />

using a variety of data sources within a common geographic system. Since first outlined by<br />

Ramsay (1998), <strong>the</strong>re h<strong>as</strong> been additional information discerned about <strong>the</strong> production<br />

methodology, <strong>an</strong>d <strong>the</strong>re have been noticeable ch<strong>an</strong>ges in <strong>the</strong> input data sources, production<br />

techniques, <strong>an</strong>d output format. This paper will cover ch<strong>an</strong>ges in <strong>the</strong> input, production techniques,<br />

<strong>an</strong>d output files since 1997, including statistics regarding <strong>the</strong> production methodology. The paper<br />

will also discuss <strong>the</strong> future enh<strong>an</strong>cements <strong>an</strong>d pending developments to <strong>the</strong> product, both short <strong>an</strong>d<br />

long term pl<strong>an</strong>s. The conclusion will summarize <strong>the</strong> present <strong>an</strong>d future of <strong>the</strong> product <strong>an</strong>d what<br />

this me<strong>an</strong>s to <strong>the</strong> user community.<br />

IMS PRODUCT EVOLUTION<br />

System architecture enh<strong>an</strong>cements<br />

A limiting factor of <strong>the</strong> original IMS system architecture w<strong>as</strong> <strong>the</strong> inability of <strong>an</strong>alysts to draw<br />

while looping imagery. This adversely affected <strong>the</strong> are<strong>as</strong> covered by geostationary satellites where<br />

imagery <strong>an</strong>imation distinguishes snow <strong>an</strong>d ice from fog <strong>an</strong>d clouds. This limitation w<strong>as</strong> ch<strong>an</strong>ged<br />

in February 2004 to allow IMS <strong>an</strong>alysts <strong>the</strong> freedom to loop imagery while drawing, er<strong>as</strong>ing, or<br />

using <strong>an</strong>y of <strong>the</strong> o<strong>the</strong>r IMS features. With this enh<strong>an</strong>cement <strong>an</strong>d o<strong>the</strong>r features such <strong>as</strong> image<br />

enh<strong>an</strong>cement, product overlays <strong>an</strong>d terrain mapping, <strong>the</strong> <strong>an</strong>alysts c<strong>an</strong> deduct snow <strong>an</strong>d ice without<br />

relying on looking at a nearby system with looping imagery.<br />

Ano<strong>the</strong>r feature modified within <strong>the</strong> system architecture in February 2004 w<strong>as</strong> enh<strong>an</strong>cing how<br />

<strong>the</strong> geographic area <strong>as</strong>signments of imagery were made within <strong>the</strong> system. The system before <strong>the</strong><br />

ch<strong>an</strong>ge applied fixed are<strong>as</strong> for each satellite. These fixed geographic are<strong>as</strong> often only covered <strong>the</strong><br />

best viewed regions. These regions also had set screen boundaries which did not allow <strong>an</strong>alysts to<br />

recenter. The current IMS allows <strong>an</strong>alysts to p<strong>an</strong> globally <strong>an</strong>d select different dat<strong>as</strong>ets/satellite data<br />

166


independent of regions. This greatly enh<strong>an</strong>ced <strong>the</strong> flexibility of <strong>the</strong> IMS to optimize <strong>the</strong> snow<br />

mapping display.<br />

Improved Resolution<br />

One of <strong>the</strong> largest ch<strong>an</strong>ges in <strong>the</strong> product since <strong>the</strong> creation of <strong>the</strong> IMS h<strong>as</strong> been <strong>the</strong><br />

improvement in output resolution introduced in February 2004. This incre<strong>as</strong>ed resolution w<strong>as</strong><br />

made to improve model input for <strong>the</strong> EMC, providing greater detail of snow <strong>an</strong>d ice cover. These<br />

improvements were possible due to adv<strong>an</strong>cements in computer speed <strong>an</strong>d imagery resolution that<br />

produced a higher resolution product at approximately 4km resolution (6144 × 6144 grid).<br />

Ch<strong>an</strong>ges were made to all <strong>an</strong>cillary data layers such <strong>as</strong> co<strong>as</strong>tlines, elevation, <strong>an</strong>d vegetation cover<br />

to support <strong>the</strong> improved resolution. Impacts of this ch<strong>an</strong>ge are still under evaluation but positive<br />

feedback h<strong>as</strong> been noted from National Wea<strong>the</strong>r Service (NWS) field stations in regard to snow<br />

se<strong>as</strong>on temperature predictions. This is in part due to <strong>the</strong> improvements in snow distribution<br />

mapping, <strong>as</strong> well <strong>as</strong> <strong>the</strong> resolving of <strong>the</strong> m<strong>an</strong>y water bodies, not possible on larger scale products.<br />

When not correctly mapped <strong>the</strong>re c<strong>an</strong> be signific<strong>an</strong>t forec<strong>as</strong>t errors where sea or lake ice cover<br />

affects heat <strong>an</strong>d moisture fluxes to <strong>the</strong> atmosphere (Grumbine, 2005). Figures 1A <strong>an</strong>d 1B<br />

demonstrate <strong>the</strong> difference in resolutions over nor<strong>the</strong>rn North America. This e<strong>as</strong>ily demonstrates<br />

<strong>the</strong> noticeable differences in <strong>the</strong> inclusion of interior lakes <strong>an</strong>d more detailed co<strong>as</strong>tlines. <strong>Snow</strong> on<br />

mountainous terrain is also better represented using <strong>the</strong> 4km versus lower resolution products. The<br />

4km product is also upscaled to <strong>the</strong> original previous resolution of approximately 24km resolution<br />

(1024 × 1024 grid). This is to maintain <strong>the</strong> satellite snow cover historical dat<strong>as</strong>et. As previously<br />

mentioned, <strong>the</strong> IMS record is <strong>an</strong> import<strong>an</strong>t climate monitoring element <strong>an</strong>d careful consideration<br />

must be taken to preserve <strong>the</strong> integrity of this snow cover record. Validation <strong>an</strong>d monitoring of <strong>the</strong><br />

IMS product at <strong>the</strong> 24km resolution is carried out under a joint effort by NESDIS <strong>an</strong>d Rutgers<br />

University (Robinson, 2003).<br />

Added input data sources<br />

The IMS w<strong>as</strong> designed to allow meteorologists to chart snow cover interactively on a daily b<strong>as</strong>is<br />

using a variety of data sources within a common geographic system. The original input satellite<br />

data sources were outlined <strong>as</strong> NOAA polar orbiters (POES), NOAA geostationary (GOES) data,<br />

Jap<strong>an</strong>ese geostationary meteorological satellites (GMS), Europe<strong>an</strong> geostationary meteorological<br />

satellites (METEOSAT), <strong>an</strong>d US Department of Defense (DOD) polar orbiters (DMSP). Indirect<br />

satellite sources also include a weekly National Ice Center (NIC) chart <strong>an</strong>d <strong>the</strong> US Air Force<br />

(USAF) daily snow depth & ice cover product (Ramsay, 1998). Several additional input products<br />

have been added to <strong>the</strong> IMS over <strong>the</strong> p<strong>as</strong>t decade. A few of <strong>the</strong>se enh<strong>an</strong>cements were outlined <strong>as</strong><br />

prospects before, but have since become operational input options (Ramsay, 2000). These<br />

products include <strong>the</strong> Adv<strong>an</strong>ced Very High Resolution Radiometer (AVHRR) ch<strong>an</strong>nel 3A, added in<br />

February 2001, <strong>the</strong> MOderate Resolution Imaging Spectroradiometer (MODIS) ch<strong>an</strong>nel 1 added<br />

in February 2004, <strong>an</strong>d <strong>an</strong> experimental automated snow mapping system over North America<br />

added in February 2004. O<strong>the</strong>r product enh<strong>an</strong>cements <strong>an</strong>d <strong>the</strong>ir impacts are outlined in <strong>the</strong><br />

following paragraphs.<br />

Meteosat 5 for INDOEX<br />

The original primary geostationary coverage left large portions of Siberia, central Asia,<br />

Himalay<strong>as</strong>, <strong>an</strong>d <strong>the</strong> Tibet<strong>an</strong> Plateau unobserved by looping imagery. This is <strong>an</strong> import<strong>an</strong>t <strong>an</strong>d<br />

difficult area for snow charting. <strong>Snow</strong> cover in this v<strong>as</strong>t area h<strong>as</strong> been identified <strong>as</strong> a signific<strong>an</strong>t<br />

influence on <strong>the</strong> Asi<strong>an</strong> Monsoon (Hahn <strong>an</strong>d Shukla, 1976; Huaqi<strong>an</strong>g et al, 2004), global<br />

circulation (Bamzai <strong>an</strong>d Marx, 2000; Clark <strong>an</strong>d Serreze, 2000; Gong et al., 2003), <strong>an</strong>d regional<br />

river discharge (Y<strong>an</strong>g et al., 2003; Sham<strong>an</strong> et al., 2005). While non-geostationary satellite data<br />

sources such <strong>as</strong> polar orbiting imagery <strong>an</strong>d microwave me<strong>as</strong>urements provide mapping snow<br />

input, <strong>the</strong>y are not <strong>the</strong> preferred data source by <strong>an</strong>alysts. Microwave retrievals over <strong>the</strong> area are<br />

often erroneous in <strong>the</strong> winter due to atmospheric signal distortion, high elevation bare ground low<br />

temperatures, <strong>an</strong>d/or soil grain scattering (B<strong>as</strong>ist et al., 1996; Armstrong <strong>an</strong>d Brodzik, 2001).<br />

Polar orbiters have limited over p<strong>as</strong>s times, thus providing only a limited number of observations<br />

167


per day. This amplifies <strong>the</strong> common problem of misidentified snow <strong>an</strong>d clouds in <strong>an</strong> already area<br />

known for its difficulty in visual identification from satellites (Clark <strong>an</strong>d Serreze, 2000).<br />

The placement of Meteosat-5 at <strong>the</strong> equator <strong>an</strong>d 63°E in 1998, helped fill <strong>the</strong> void of<br />

geostationary data. This satellite w<strong>as</strong> moved in support of <strong>the</strong> Indi<strong>an</strong> Oce<strong>an</strong> Experiment<br />

(INDOEX), <strong>an</strong>d w<strong>as</strong> first incorporated into IMS in March 2001. This platform w<strong>as</strong> preferred over<br />

o<strong>the</strong>r geostationary meteorological satellites such <strong>as</strong> Feng Yun 2 (FY-2) from China <strong>an</strong>d Indi<strong>an</strong><br />

National Satellite System 2 (INSAT 2) from India. The observation footprint is comparable for all<br />

<strong>the</strong>se satellites, but only a single geostationary satellite is required. Meteosat-5 w<strong>as</strong> chosen<br />

because FY2 is experimental <strong>an</strong>d near <strong>the</strong> end of its serviceable period, while data from INSAT<br />

are available for Indi<strong>an</strong> national use. The inclusion of Meteosat-5 h<strong>as</strong> provided a great boost to<br />

Asi<strong>an</strong> snow mapping during <strong>the</strong> winter se<strong>as</strong>on.<br />

Multi-functional Tr<strong>an</strong>sport Satellite (MTSAT)<br />

The Jap<strong>an</strong>ese Meteorological Agency (JMA) MTSAT series succeed <strong>the</strong> Geostationary<br />

Meteorological Satellite (GMS) series <strong>as</strong> <strong>the</strong> next generation satellite series covering E<strong>as</strong>t Asia<br />

<strong>an</strong>d <strong>the</strong> Western Pacific. While MTSAT’s offers only marginal improvements in visible resolution<br />

over GMS, <strong>the</strong> improved calibration <strong>an</strong>d correction algorithms provide improved detection of<br />

snow. The IMS beg<strong>an</strong> using MTSAT in November 2005.<br />

National Operational Hydrologic Remote Sensing Center (NOHRSC)<br />

The inclusion of NOHRSC maps into IMS <strong>an</strong>alysis beg<strong>an</strong> in February 2004. The national<br />

<strong>an</strong>alysis provided by NOHRSC, called <strong>the</strong> SNOw Data Assimilation System (SNODAS), provides<br />

a 1 km resolution estimation of snow cover for <strong>the</strong> conterminous United States. SNODAS is a<br />

system that amalgamates NCEP modeling, multisensor, station report, <strong>an</strong>d airborne information<br />

into a single daily or sub-daily product (Barrett, 2003). The output timing of this product<br />

corresponds to <strong>the</strong> IMS observation day <strong>an</strong>d serves <strong>as</strong> <strong>an</strong> import<strong>an</strong>t winter input source when<br />

clouds bl<strong>an</strong>ket <strong>the</strong> conterminous United States. While <strong>the</strong> fine resolution <strong>an</strong>d multi-source data of<br />

SNODAS provides reliable data, its spatial extent is limited compared to <strong>the</strong> IMS, so it provides<br />

only a small but nationally import<strong>an</strong>t area for snow mapping.<br />

MODIS Looping<br />

Soon after <strong>the</strong> inclusion of MODIS visible imagery <strong>as</strong> <strong>an</strong> IMS dat<strong>as</strong>ource, <strong>an</strong>alysts found <strong>the</strong><br />

utility of looping recent MODIS overp<strong>as</strong>ses for a given location. Looping of this polar orbiter is<br />

available due in part to <strong>the</strong> Aqua <strong>an</strong>d Terra satellites sharing <strong>the</strong> identical visible ch<strong>an</strong>nel at 1km<br />

<strong>an</strong>d <strong>the</strong> poles having multiple daily overp<strong>as</strong>ses. While <strong>the</strong> time sp<strong>an</strong> between images used in <strong>the</strong><br />

loop is somewhat coarse compared with geostationary observations, <strong>the</strong>se loops allow <strong>the</strong> <strong>an</strong>alyst<br />

greater ability to distinguish between <strong>the</strong> surface cyrosphere <strong>an</strong>d clouds. MODIS is <strong>an</strong><br />

experimental satellite <strong>an</strong>d will not be followed by a direct legacy product. The merger of <strong>the</strong> DoD<br />

<strong>an</strong>d NOAA satellites in <strong>the</strong> future, known <strong>as</strong> <strong>the</strong> National Polar-Orbiting Operational<br />

Environmental Satellite System (NPOESS), will provide a similar product <strong>as</strong> that of MODIS <strong>an</strong>d<br />

hopefully c<strong>an</strong> be exploited for image looping once NPOESS is launched <strong>an</strong>d declared operational.<br />

Marine Modeling <strong>an</strong>d Analysis Br<strong>an</strong>ch (MMAB) Sea Ice Analysis<br />

The tracking of sea ice cover presents m<strong>an</strong>y difficulties. The IMS relies primary on visible<br />

imagery but this method is contingent on clear sky or thin cloud during <strong>the</strong> observing periods.<br />

When wea<strong>the</strong>r or low illumination obscure visual interpretation of sea ice, microwaves play a<br />

greater role. IMS <strong>an</strong>alysts often apply a 1/16 mesh Nor<strong>the</strong>rn Hemispheric grid of sea ice<br />

concentrations from <strong>the</strong> MMAB to demarcate those locations with >50% ice cover. The MMAB<br />

sea ice <strong>an</strong>alysis is solely b<strong>as</strong>ed on Special Sensor Microwave Imager (SSM/I) <strong>an</strong>d applies a<br />

modified version of <strong>the</strong> National Aeronautics <strong>an</strong>d Space Administration (NASA) team algorithm<br />

to derive sea ice concentrations (Grumbine, 1996). All SSM/I b<strong>as</strong>ed products suffer from melt<br />

water attenuation, co<strong>as</strong>tal contamination, poor thin ice detection, <strong>an</strong>d difficulties in identifying<br />

concentration along <strong>the</strong> marginal ice zone. Analysts apply <strong>the</strong> MMAB product cautiously <strong>an</strong>d will<br />

168


slightly vary <strong>the</strong> IMS with this product when o<strong>the</strong>r external sources suggest <strong>the</strong> MMAB may be<br />

contaminated with a false emissivity.<br />

A B<br />

Figure 1. The 24km IMS product (A on <strong>the</strong> left) <strong>as</strong> remapped from <strong>the</strong> 4km IMS (B on <strong>the</strong> right) for May 19,<br />

2006. The 24km represents <strong>the</strong> same l<strong>an</strong>d/sea m<strong>as</strong>k used in previous 24km products prior to February 2004.<br />

Note <strong>the</strong> differences in representations for mountainous snow, inl<strong>an</strong>d lakes, <strong>an</strong>d sea ice leads.<br />

Daily NIC Ice Edge<br />

In its inception, <strong>the</strong> IMS w<strong>as</strong> designed to exploit <strong>the</strong> NIC weekly ice charts for <strong>the</strong> Nor<strong>the</strong>rn<br />

Hemisphere. The NIC produced detailed sea ice maps coded in <strong>an</strong> ice charting nomenclature<br />

known <strong>as</strong> egg code once per week, usually updating <strong>the</strong> product on Friday afternoons, until 2001.<br />

Since that time <strong>the</strong> NIC switched from weekly to biweekly hemispheric coverage. This decre<strong>as</strong>ed<br />

<strong>the</strong> utility of <strong>the</strong> charts for daily operational ice mapping. Fur<strong>the</strong>rmore, ice charts rele<strong>as</strong>ed on<br />

Fridays would use input data typically three to seven days old for <strong>the</strong> <strong>an</strong>alysis. While <strong>the</strong> NIC<br />

charts could still provide a general outlook of ice thickness <strong>an</strong>d ice distribution, <strong>the</strong> dynamic<br />

nature of ice made <strong>the</strong> charts too old for <strong>the</strong> IMS <strong>an</strong>alysis. Since February 2004, a daily vector sea<br />

ice edge h<strong>as</strong> been incorporated into <strong>the</strong> IMS. The NIC daily sea ice vector product applies visible,<br />

infrared, p<strong>as</strong>sive microwave, <strong>an</strong>d radar data to outline those are<strong>as</strong> with >10% ice cover. This<br />

product differs from <strong>the</strong> IMS product in three key ways. First, <strong>the</strong> NIC product is vector-b<strong>as</strong>ed <strong>an</strong>d<br />

attempts to enclose amorphous polygons, while <strong>the</strong> IMS defines ice cover within predefined<br />

pixels. The NIC ice cover h<strong>as</strong> no set size requirements on <strong>the</strong> polygon size or shape, thus <strong>the</strong> scale<br />

of what are<strong>as</strong> enclose >10% is at <strong>the</strong> NIC <strong>an</strong>alyst’s digression. This often leads to smoothing along<br />

<strong>the</strong> ice edge in some are<strong>as</strong>, while o<strong>the</strong>r are<strong>as</strong> may be more detailed th<strong>an</strong> <strong>the</strong> IMS. Experience h<strong>as</strong><br />

revealed more of <strong>the</strong> former th<strong>an</strong> <strong>the</strong> latter. A second difference in <strong>the</strong> products is <strong>the</strong> ice<br />

concentrations captured by each product. The NIC outlines are<strong>as</strong> with >10% ice cover, while <strong>the</strong><br />

IMS demarcates are<strong>as</strong> with >50% ice cover. IMS <strong>an</strong>alysts must adjust for this difference when<br />

using <strong>the</strong> NIC daily ice edge for IMS input. A third difference in <strong>the</strong> ice <strong>an</strong>alysis output is <strong>the</strong><br />

IMS’s inclusion of lake ice. The NIC vector ice edge only outlines sea ice <strong>an</strong>d lake ice in <strong>the</strong> Great<br />

169


Lakes. O<strong>the</strong>r signific<strong>an</strong>t lake bodies that freeze (Great Bear, Great Slave, C<strong>as</strong>pi<strong>an</strong> Sea, Lake<br />

Baikal, Aral Sea, among m<strong>an</strong>y o<strong>the</strong>rs) are not included within <strong>the</strong> NIC product.<br />

Despite <strong>the</strong> differences, IMS <strong>an</strong>alysts will bridge <strong>the</strong> <strong>an</strong>alysis outputs <strong>an</strong>d methodologies to use<br />

NIC data <strong>as</strong> <strong>an</strong> input source. NIC data is often applied with or taken in <strong>the</strong> context of <strong>the</strong> MMAB<br />

sea ice product to provide a best guess approach for <strong>the</strong> IMS. Figure 2 shows how <strong>the</strong> NIC ice<br />

cover product is able to suggest ice cover in <strong>the</strong> Hudson Bay even when imagery h<strong>as</strong> cloud<br />

contamination.<br />

Figure 2. AVHRR Ch<strong>an</strong>nel 1 with <strong>the</strong> NIC ice edge overlayed for May 29, 2006. The NIC ice edge is a<br />

vector file with shorelines. White lines represent shoreline <strong>an</strong>d 10% ice cover isopleth.<br />

One c<strong>an</strong> notice that each product suggests a different ice cover pattern, perhaps due to times of<br />

observations. But <strong>the</strong> NIC ice edge is able to provide IMS <strong>an</strong>alysts information regarding <strong>the</strong><br />

shape of <strong>the</strong> lead in James Bay, even through <strong>the</strong> sou<strong>the</strong>rn part of James Bay is cloud-covered.<br />

Pl<strong>an</strong>s are being developed to bring <strong>the</strong> radar data available to NIC <strong>an</strong>alysis into <strong>the</strong> IMS <strong>an</strong>d to<br />

have <strong>an</strong> NIC ice edge product that outlines using similar criteria <strong>as</strong> that of <strong>the</strong> IMS. This will be<br />

discussed later in this paper in greater detail.<br />

Automated <strong>Snow</strong> <strong>an</strong>d Ice Cover Products<br />

In August 1999, NOAA/NESDIS beg<strong>an</strong> <strong>the</strong> production of automated snow maps over North<br />

America. The product generates daily maps of 4km resolution b<strong>as</strong>ed on visible, near-infrared,<br />

middle-infrared, infrared <strong>an</strong>d microwave imagery. This imagery comes from both polar-orbiting<br />

<strong>an</strong>d geostationary satellites. The algorithm applies a series of decision-trees to bin pixels<br />

containing ei<strong>the</strong>r a majority of <strong>the</strong> area snow covered or having less <strong>the</strong>n a majority of area snow<br />

covered (Rom<strong>an</strong>ov et al., 2000). While <strong>the</strong> midlatitudes maps are generally mapped using GOES<br />

170


imagery, <strong>the</strong> higher latitudes rely on polar orbiter spectral differences <strong>an</strong>d microwave signals.<br />

Microwave retrievals are <strong>the</strong> default observation when shorter wavelength data is attenuated by<br />

clouds for several days. Since being declared operational <strong>the</strong> product h<strong>as</strong> exp<strong>an</strong>ded beyond North<br />

America to <strong>the</strong> Sou<strong>the</strong>rn Hemisphere <strong>an</strong>d <strong>the</strong> remainder of <strong>the</strong> Nor<strong>the</strong>rn Hemisphere, though <strong>the</strong><br />

spatially exp<strong>an</strong>ded versions of <strong>the</strong> automated snow maps remain experimental at <strong>the</strong> present time<br />

(Rom<strong>an</strong>ov <strong>an</strong>d Tarpley, 2003). Examples of <strong>the</strong> Nor<strong>the</strong>rn Hemisphere multisensor product are<br />

demonstrated on Figure 3. The pattern closely resembles that of <strong>the</strong> IMS for <strong>the</strong> same date (IMS<br />

not shown). Figure 4 displays <strong>an</strong> example of <strong>the</strong> experimental automated Sou<strong>the</strong>rn Hemisphere<br />

product. Validation efforts remain ongoing for both hemispheric products. O<strong>the</strong>r comparison<br />

studies using automated snow cover mapping versus IMS suggest that IMS may have<br />

underestimation problems in <strong>the</strong> tr<strong>an</strong>sition se<strong>as</strong>on but outperform automated products in cloudy<br />

are<strong>as</strong> with new snow cover <strong>an</strong>d during winter (Brubaker et al., 2005). The validation efforts will<br />

help determine how <strong>the</strong>se products will be incorporated with <strong>the</strong> IMS. Without a current Sou<strong>the</strong>rn<br />

Hemisphere IMS product, <strong>an</strong> automated product could serve <strong>as</strong> <strong>the</strong> NOAA NESDIS Sou<strong>the</strong>rn<br />

Hemisphere output in <strong>the</strong> future. However should <strong>the</strong> product be unable to provide a serviceable<br />

input for EMC or CPC modeling efforts, <strong>the</strong> output will merely serve <strong>as</strong> one input to a sou<strong>the</strong>rn<br />

hemispheric IMS <strong>an</strong>alysis that will need to be created. Likewise, <strong>the</strong> role <strong>the</strong> Nor<strong>the</strong>rn<br />

Hemispheric automated <strong>an</strong>alysis product plays in <strong>the</strong> IMS will be determined b<strong>as</strong>ed on <strong>the</strong><br />

validation results <strong>an</strong>d <strong>the</strong> customer requirements. Possible scenarios include serving <strong>as</strong> <strong>an</strong>o<strong>the</strong>r<br />

layer in <strong>the</strong> IMS (much like <strong>the</strong> North Americ<strong>an</strong> product does now), serving <strong>as</strong> <strong>the</strong> initial state of<br />

<strong>the</strong> IMS, or even replacing <strong>the</strong> current Nor<strong>the</strong>rn Hemispheric IMS product.<br />

Current Production Methodology for IMS<br />

Since <strong>the</strong> inception of <strong>the</strong> IMS, production methodologies <strong>an</strong>d image preferences have become<br />

more tr<strong>an</strong>sparent. The production of <strong>the</strong> IMS h<strong>as</strong> evolved over time, with inclination in which<br />

imagery types are applied at certain times of year <strong>an</strong>d how long <strong>the</strong> production process takes. This<br />

section will provide a greater insight into <strong>the</strong> production methodology used to make multisensor<br />

output. Production of <strong>the</strong> IMS products are not tremendously time consuming for <strong>an</strong>alysts, who<br />

spend <strong>an</strong>ywhere from one to five hours generating a daily product depending on <strong>the</strong> se<strong>as</strong>on,<br />

<strong>an</strong>alyst familiarity, <strong>an</strong>d satellite data available. The month of August requires <strong>the</strong> lowest average<br />

time to conduct <strong>an</strong> <strong>an</strong>alysis, averaging between 70 <strong>an</strong>d 75 minutes. Production time during <strong>the</strong><br />

accumulation se<strong>as</strong>on (Oct–Apr) averages about 120 minutes. The ablation se<strong>as</strong>on (May–Sep)<br />

averages approximately 90 minutes for production time. Products are due to <strong>the</strong> primary customer<br />

by 2300z. The IMS <strong>an</strong>alysis currently starts with <strong>the</strong> previous day’s <strong>an</strong>alysis <strong>as</strong> <strong>the</strong> initial state.<br />

The <strong>an</strong>alyst <strong>the</strong>n reconfigures <strong>the</strong> current IMS b<strong>as</strong>ed on <strong>the</strong> input data available at <strong>the</strong> time of <strong>the</strong><br />

<strong>an</strong>alysis.<br />

Se<strong>as</strong>ons determine not only just how much time is required to generate a chart, but also what<br />

data will be used <strong>as</strong> input into <strong>the</strong> IMS. Geostationary data looping is <strong>the</strong> primary tool for<br />

determination of snow cover (Ramsay, 1998).<br />

171


Figure 3. Examples of <strong>the</strong> experimental NOAA automated snow <strong>an</strong>d ice multisensor retrieval output for May<br />

30, 2006 over Eur<strong>as</strong>ia (above) <strong>an</strong>d North America (below).<br />

The geostationary looping represents <strong>an</strong> estimated 60% of <strong>the</strong> snow <strong>an</strong>alysis examination are<strong>as</strong><br />

during <strong>the</strong> winter (Dec–Feb). This decre<strong>as</strong>es to <strong>an</strong> estimated 30% of <strong>the</strong> snow examination area<br />

during <strong>the</strong> summer (Jun–Aug). During <strong>the</strong> summer months, polar orbiting satellites’ visible<br />

ch<strong>an</strong>nels (b<strong>an</strong>ds) characterize <strong>an</strong> estimated 65% of <strong>the</strong> snow <strong>an</strong>alysis. Analysts generally prefer<br />

using visible imagery for snow extent mapping, but will use microwave data in <strong>the</strong> event that light<br />

is unavailable, due to cloud occlusion or low solar illumination <strong>an</strong>gles. The combination of high<br />

albedo, static motion, <strong>an</strong>d meteorological conditions provides <strong>the</strong> <strong>an</strong>alyst with 80–90% of input<br />

data used in <strong>the</strong> <strong>an</strong>alysis of observed snow cover. Even during <strong>the</strong> winter, microwave derived<br />

snow data generally only represents 5% of <strong>an</strong> <strong>an</strong>alysis. Microwave snow extent data have welldocumented<br />

snow misidentification errors due to signal obstructions, snow grain size, l<strong>an</strong>d cover<br />

influences, <strong>an</strong>d algorithm limitations (Ch<strong>an</strong>g et al., 1996; Foster et al., 1999; Foster et al., 2004).<br />

Analysts rely more on snow climatology to estimate snow cover in <strong>the</strong> high latitudes during <strong>the</strong><br />

winter th<strong>an</strong> pure microwave data. Where <strong>an</strong>d when sources are available <strong>the</strong> IMS uses METAR<br />

<strong>an</strong>d cooperative observations, NOAA NWS’ NOHRSC, <strong>an</strong>d SNODAS data, <strong>an</strong>d automated snow<br />

cover maps from both NOAA’s Center for Satellite Applications <strong>an</strong>d Research (STAR) <strong>an</strong>d <strong>the</strong><br />

172


MODIS L<strong>an</strong>d Science Team, NASA Goddard Space Flight Center (GSFC). These external data<br />

sources are often used to validate <strong>an</strong> area having snow or snow obscured with clouds. Often, <strong>the</strong><br />

IMS <strong>an</strong>alyst must use a consensus of data sources to provide <strong>an</strong> optimal “best-guess” approach to<br />

determining <strong>the</strong> presence of snow.<br />

Figure 4. Examples of <strong>the</strong> experimental NOAA automated snow <strong>an</strong>d ice NOAA-17 retrieval output for May<br />

30, 2006 over Sou<strong>the</strong>rn Hemisphere. This product is at 4km resolution <strong>an</strong>d could exp<strong>an</strong>d IMS from <strong>the</strong><br />

Nor<strong>the</strong>rn Hemisphere to global coverage. Green are<strong>as</strong> represent are<strong>as</strong> sc<strong>an</strong>ned for snow cover, while grey<br />

are<strong>as</strong> represent l<strong>an</strong>d are<strong>as</strong> not sc<strong>an</strong>ned for snow cover.<br />

173


Ice cover <strong>an</strong>alysis relies on a different approach th<strong>an</strong> snow cover for charting. Ice cover<br />

determination must rely less on high albedo, stagnate cover, <strong>an</strong>d meteorological conditions. Sea<br />

ice in <strong>the</strong> Nor<strong>the</strong>rn Hemisphere winter is primarily located in are<strong>as</strong> with low solar illumination.<br />

New ice formation often h<strong>as</strong> a low albedo until <strong>the</strong> ice thickens, becomes more opaque, <strong>an</strong>d albedo<br />

incre<strong>as</strong>es (Wadhams, 2000). Fur<strong>the</strong>rmore, sea ice is a dynamic surface making it less discernible<br />

from clouds using image loops. The presence of lake <strong>an</strong>d sea ice c<strong>an</strong> be unrelated to current<br />

meteorological conditions due to ice tr<strong>an</strong>sport, ice thickness, water temperatures, among o<strong>the</strong>r<br />

factors. The prominence of low-level stratus clouds over polar <strong>an</strong>d sub-polar region also preclude<br />

<strong>the</strong> use of visible imagery <strong>as</strong> <strong>the</strong> primary source of ice observation. While this reduces <strong>the</strong><br />

efficiency of albedo-b<strong>as</strong>ed observation of ice, it is still a valuable input. On average, about 60% of<br />

ch<strong>an</strong>ges in winter ice cover are noted via geostationary, AVHRR or MODIS observations. Much<br />

of <strong>the</strong> higher latitude are<strong>as</strong> are verified <strong>as</strong> being ice-covered using microwave-b<strong>as</strong>ed retrievals.<br />

Microwave-b<strong>as</strong>ed observations often represent 30–35% of <strong>the</strong> winter <strong>an</strong>d autumn (Sep–Nov) ice<br />

cover input. Ice climatology is <strong>an</strong>o<strong>the</strong>r tool for estimating ice cover in places where observations<br />

are unavailable. Since ice cover often exists in remote <strong>an</strong>d d<strong>an</strong>gerous are<strong>as</strong>, no station data is<br />

currently incorporated into <strong>the</strong> <strong>an</strong>alysis. The NIC produces a sea ice edge vector file that provides<br />

<strong>the</strong> IMS with <strong>an</strong> external source for ice cover information. Currently, <strong>the</strong> NIC ice edge<br />

encomp<strong>as</strong>ses total polygons with greater th<strong>an</strong> 10% ice cover. The IMS attempts to identify<br />

whe<strong>the</strong>r each 4km × 4km pixel contains more th<strong>an</strong> 50% ice cover. These products do not<br />

correspond directly due to each product requiring different output specifications. Despite <strong>the</strong><br />

differences between products, <strong>the</strong> NIC ice edge is used when o<strong>the</strong>r sources of data fail to provide<br />

<strong>an</strong>y clear input on ice-covered oce<strong>an</strong> or Great Lakes waters. This represents about 2–10% of <strong>the</strong><br />

time, depending on <strong>the</strong> se<strong>as</strong>on.<br />

Mountainous snow mapping<br />

Elevation plays <strong>an</strong> import<strong>an</strong>t role in snow generation due mostly to orographic lifting <strong>an</strong>d<br />

temperature decre<strong>as</strong>ing with incre<strong>as</strong>ing height. <strong>Snow</strong> often outlines higher elevation are<strong>as</strong> during<br />

tr<strong>an</strong>sition se<strong>as</strong>on <strong>an</strong>d during <strong>the</strong> winter in semi-arid, mid-latitude regions. To mimic this effect in<br />

mapping snow, <strong>the</strong> IMS allows <strong>an</strong>alysts to chart are<strong>as</strong> dynamically <strong>as</strong> having snow b<strong>as</strong>ed on a<br />

digital elevation model (DEM). The DEM resolution is 4km <strong>an</strong>d matches <strong>the</strong> IMS, thus providing<br />

a direct relationship between elevation <strong>an</strong>d IMS pixels. This provides <strong>the</strong> <strong>an</strong>alyst with <strong>the</strong> ability<br />

to toggle <strong>the</strong> pixels within a given polygon to match <strong>the</strong> outline of snow revealed through imagery.<br />

The <strong>an</strong>alyst c<strong>an</strong> optimize <strong>the</strong> snow cover pattern b<strong>as</strong>ed on elevation, local geography, <strong>an</strong>d<br />

reflect<strong>an</strong>ce revealed through imagery. This h<strong>as</strong> become a frequently applied tool in IMS snow<br />

mapping.<br />

The strengths <strong>an</strong>d shortcomings of this DEM-b<strong>as</strong>ed mapping are considered by <strong>the</strong> <strong>an</strong>alyst while<br />

applying this tool. The strength is a more detailed <strong>an</strong>d realistic mapping technique th<strong>an</strong> previously<br />

available b<strong>as</strong>ed on a physiographic relationship of snow with elevation. A weakness is <strong>the</strong> DEM<br />

b<strong>as</strong>ed mapping does not account for o<strong>the</strong>r known state or physiographic factors that play a role in<br />

snow cover distribution, such <strong>as</strong> slope <strong>an</strong>d <strong>as</strong>pect. Nor does it take solar, vegetation, or<br />

climatologic wind <strong>an</strong>d storm patterns into account. Studies reveal that physiographic features such<br />

<strong>as</strong> radiation, elevation, slope, <strong>an</strong>d <strong>as</strong>pect account for between 50–80% of snow depth variability in<br />

<strong>the</strong> Rocky Mountains, Sierra Nevad<strong>as</strong>, <strong>an</strong>d Alps (March<strong>an</strong>d <strong>an</strong>d Killingtveit, 2001; Balk <strong>an</strong>d<br />

Elder, 2000). Elevation tends to be <strong>the</strong> second largest influence on snow cover distribution next to<br />

radiation, but <strong>the</strong> weight of this influence is dependent on scale (Balk <strong>an</strong>d Elder, 2000, Marks et<br />

al., 2002). Elevation <strong>an</strong>d radiation appear to be greater factors at incre<strong>as</strong>ing scales, likely playing a<br />

large role in distribution variability at <strong>the</strong> 4km scale in semi-arid <strong>an</strong>d mountainous environments.<br />

Despite <strong>the</strong> shortcoming of this tool, it is just a methodology for mapping, with <strong>an</strong>alysts b<strong>as</strong>ing<br />

snow distribution on numerous input data not merely elevation. Analysts c<strong>an</strong> compensate for<br />

inhomogeneous spatial patterns noted with regional elevation due to <strong>the</strong> o<strong>the</strong>r state factors that<br />

influence snow cover distribution.<br />

174


FUTURE OF IMS<br />

Enh<strong>an</strong>cements to <strong>the</strong> IMS will continue to push <strong>the</strong> bounds of cryospheric observation <strong>an</strong>d<br />

charting. Requirements for snow <strong>an</strong>d sea ice extent differ for climate verse numerical wea<strong>the</strong>r<br />

prediction. Since surface cyrospheric climate dat<strong>as</strong>ets are constrained by p<strong>as</strong>t data with <strong>the</strong><br />

previous scale limitations, data needs to be maintained at original resolutions to preserve <strong>the</strong><br />

climate record integrity (Robinson, 2003). Downscaling of older dat<strong>as</strong>ets would be required to<br />

blend <strong>the</strong> old <strong>an</strong>d new data into a common 4km grid with interpolated daily values from weekly<br />

values. Fox <strong>an</strong>d Gh<strong>an</strong> (2004) point out <strong>an</strong>o<strong>the</strong>r limitation to <strong>the</strong> current computational limits that<br />

fine resolution climate grids would need to overcome <strong>as</strong> well. While <strong>the</strong> IMS’s resolution for<br />

global climate outlooks <strong>an</strong>d monitoring exceeds <strong>the</strong> current requirements by using a 4km<br />

resolution at one observation per day, improvements in wea<strong>the</strong>r prediction are predicated on <strong>the</strong><br />

prediction grid resolutions <strong>an</strong>d prefer up to date observations. The need for global snow<br />

information at improved spatial <strong>an</strong>d temporal resolutions for numerical wea<strong>the</strong>r prediction models<br />

is driving <strong>the</strong> adv<strong>an</strong>cements in <strong>the</strong> IMS. A 4km resolution snow <strong>an</strong>d ice cover h<strong>as</strong> <strong>the</strong> spatial<br />

resolution to initialize models such <strong>as</strong> <strong>the</strong> NCEP North Americ<strong>an</strong> Mesoscale (NAM) model with<br />

12km resolution <strong>an</strong>d could even provide improved ice <strong>an</strong>d snow initialization for finer resolution<br />

models such <strong>as</strong> <strong>the</strong> Fifth-Generation Penn State/National Center for Atmospheric Research<br />

Mesoscale Model (MM5) (Dudhia et al., 1999; Rogers et al., 2001). The temporal resolution of<br />

once per day could improve model results since afternoon (E<strong>as</strong>tern St<strong>an</strong>dard Time, EST) model<br />

runs may be up to 21 hours removed from <strong>the</strong> l<strong>as</strong>t snow <strong>an</strong>d ice cover initialization. With daily<br />

snow depth depletions of over 12 inches reported at SNOTEL stations, spatial distribution of snow<br />

cover may ch<strong>an</strong>ge dr<strong>as</strong>tically over one day, given ideal wea<strong>the</strong>r conditions for ablation. IMS will<br />

attempt to respond to this need for more timely information by introducing a second IMS<br />

observation over North America at <strong>the</strong> 4km resolution. This will be ch<strong>an</strong>ging given time<br />

constraints of Satellite Analysis Br<strong>an</strong>ch (SAB) <strong>an</strong>alysts <strong>an</strong>d <strong>the</strong> window of visible imagery<br />

available for <strong>an</strong>alysis by <strong>the</strong> late afternoon EST model run, particularly in <strong>the</strong> western United<br />

States in <strong>the</strong> winter.<br />

In addition to a second IMS daily product, <strong>the</strong> IMS will be exp<strong>an</strong>ding to provide global<br />

coverage. While <strong>the</strong> IMS provides adequate coverage in <strong>the</strong> Nor<strong>the</strong>rn Hemisphere, <strong>the</strong> product<br />

failed to capture <strong>the</strong> snow <strong>an</strong>d ice extent in <strong>the</strong> Sou<strong>the</strong>rn Hemisphere. Like <strong>the</strong> improved temporal<br />

resolution North Americ<strong>an</strong> IMS product, this presents a challenge to <strong>the</strong> resources required to<br />

provide such data. As previously mentioned, <strong>the</strong> automated snow <strong>an</strong>d ice product is likely to play<br />

a large role in <strong>the</strong> production of sou<strong>the</strong>rn hemispheric <strong>an</strong>alysis. The completion of <strong>the</strong> Sou<strong>the</strong>rn<br />

Hemisphere IMS will complete <strong>the</strong> global snow <strong>an</strong>d ice coverage for model initialization.<br />

The IMS currently employs over 15 separate sources of data for input. This number c<strong>an</strong> seem<br />

daunting to navigate, but each source is expertly selected to provide <strong>an</strong> optimal snow <strong>an</strong>alysis.<br />

Still, NOAA is looking to exploit new technologies for underst<strong>an</strong>ding <strong>the</strong> current state of <strong>the</strong><br />

surface cryosphere. To improve <strong>the</strong> output <strong>an</strong>d to meet future product requirements, several new<br />

products are being tested for implementation into <strong>the</strong> IMS. In <strong>the</strong> short term, <strong>the</strong>se products<br />

include snow <strong>an</strong>d sea ice cover from <strong>the</strong> Adv<strong>an</strong>ced Microwave Sc<strong>an</strong>ning Radiometer-EOS<br />

(AMSR-E), <strong>the</strong> Nor<strong>the</strong>rn <strong>an</strong>d Sou<strong>the</strong>rn Hemisphere automated snow mapping systems, NASA’s<br />

Quick Scatterometer (QuikSCAT), <strong>an</strong>d ESA’s Environmental Satellite (Envisat) Adv<strong>an</strong>ced<br />

Syn<strong>the</strong>tic Aperture Radar (ASAR) Global Monitoring Mode (GMM), <strong>an</strong>d MetOp’s Adv<strong>an</strong>ced<br />

Scatterometer (ASCAT). MetOp’s impending launch will also offer <strong>an</strong> exp<strong>an</strong>sion in <strong>the</strong> platforms<br />

carrying <strong>the</strong> AVHRR <strong>an</strong>d Adv<strong>an</strong>ced Microwave Sounding Unit (AMSU) sensors. Recent<br />

improvements to <strong>the</strong> Air Force Wea<strong>the</strong>r Agency (AFWA) snow depth <strong>an</strong>d MMAB sea ice<br />

products will also be incorporated in <strong>the</strong> near future.<br />

The IMS output product h<strong>as</strong> been available to users for almost ten years now. Archival <strong>an</strong>d<br />

archived product dissemination h<strong>as</strong> been done through cooperation with <strong>the</strong> National <strong>Snow</strong> <strong>an</strong>d<br />

Ice Data Center (NSIDC). The NSIDC currently provides users with Americ<strong>an</strong> St<strong>an</strong>dard Code for<br />

Information Interch<strong>an</strong>ge (ASCII) output data at <strong>the</strong> original 24km product <strong>as</strong> well <strong>as</strong> <strong>the</strong> recently<br />

added 4km output. While <strong>the</strong>se products have been a popular data source, <strong>the</strong> formatting of <strong>the</strong><br />

output c<strong>an</strong> be complex to novice users. To promote a broader user community, NOAA NESDIS<br />

175


h<strong>as</strong> begun to generate GIS GeoTiff compatible output at <strong>the</strong> 4km resolution. This product will be<br />

archived <strong>an</strong>d disseminated at <strong>the</strong> NSIDC. The GeoTiff archive will sp<strong>an</strong> from February 2004 until<br />

<strong>the</strong> latest <strong>an</strong>alysis day when complete.<br />

<strong>Snow</strong> <strong>an</strong>d ice extent <strong>an</strong>d coverage h<strong>as</strong> been <strong>the</strong> primary output for <strong>the</strong> IMS. However, this is far<br />

from <strong>the</strong> lone variable needed for modeling snow <strong>an</strong>d ice behavior at regional <strong>an</strong>d global scales.<br />

NOAA/NESDIS is at <strong>the</strong> cusp of introducing new snow products that work in conjunction with <strong>the</strong><br />

IMS to improve initialization in atmospheric models. A common problem reported with IMS h<strong>as</strong><br />

been continuing to keep snow cover during cloud obscured periods. The IMS <strong>an</strong>alysts apply m<strong>an</strong>y<br />

tools <strong>an</strong>d images to produce a “best guess” approach to snow observing. However, IMS <strong>an</strong>alysts<br />

leave conditions <strong>as</strong> <strong>the</strong>y were since <strong>the</strong> l<strong>as</strong>t observation when clouds obscure visible observations,<br />

snow is too thin for microwave detection, <strong>an</strong>d <strong>the</strong>re are no station reports. This c<strong>an</strong> be problematic<br />

when snow h<strong>as</strong> actually melted away. Atmospheric models contain algorithms to estimate snow<br />

depth throughout day <strong>an</strong>d predict ablation of snow cover. However, snow ablation in atmospheric<br />

models is reinitialized with <strong>the</strong> current snow extent from IMS. If <strong>the</strong> snow in <strong>the</strong> IMS is merely<br />

<strong>the</strong> result of continu<strong>an</strong>ce <strong>an</strong>d not observed, this reinitialization c<strong>an</strong> lead to false snow observations<br />

<strong>an</strong>d propagate errors throughout <strong>the</strong> NCEP model. A file of l<strong>as</strong>t observation time for <strong>the</strong> IMS h<strong>as</strong><br />

been developed <strong>an</strong>d is undergoing testing. This will allow modelers to choose to use IMS for snow<br />

cover observation or to b<strong>as</strong>e snow cover on modeled estimates. <strong>Snow</strong> water equivalent (SWE) h<strong>as</strong><br />

been produced by SSM/I me<strong>as</strong>urements for m<strong>an</strong>y decades, <strong>an</strong>d is a valuable snow variable for<br />

atmospheric <strong>an</strong>d hydrologic modelers. NOAA NESDIS is currently testing a combined AMSU<br />

<strong>an</strong>d IMS SWE product that will merge <strong>the</strong> reliable IMS snow cover observations with AMSU’s<br />

capacity for estimating SWE (Kongoli et al., 2006). An example of <strong>the</strong> pre-merged <strong>an</strong>d merged<br />

products is demonstrated in Figure 5. Additional snow variables like snow depth <strong>an</strong>d fractional<br />

snow covered area are being experimented with to improve model initialization in combination<br />

with IMS output (Rom<strong>an</strong>ov et al., 2003). Future NOAA efforts for sea <strong>an</strong>d lake ice variables<br />

utilizing IMS ice cover such <strong>as</strong> ice concentration <strong>an</strong>d ice thickness are in <strong>the</strong> pl<strong>an</strong>ning stage.<br />

The enh<strong>an</strong>cements of input <strong>an</strong>d output are not <strong>the</strong> only short-term ch<strong>an</strong>ges pl<strong>an</strong>ned for <strong>the</strong><br />

product’s future. Current pl<strong>an</strong>s are to relocate <strong>the</strong> operational production of <strong>the</strong> IMS from its<br />

current location of SAB to <strong>the</strong> NIC. This tr<strong>an</strong>sition from one agency within NESDIS to <strong>an</strong>o<strong>the</strong>r is<br />

being done with <strong>the</strong> hope of producing <strong>an</strong> improved product at a reduced cost for NESDIS <strong>as</strong> a<br />

whole. This should lead to less duplication in sea ice monitoring within NESDIS by parallel<br />

offices, consolidate network systems <strong>an</strong>d imagery storage, free SAB personnel time for o<strong>the</strong>r<br />

products, <strong>an</strong>d allow time for NIC personnel trained for IMS <strong>an</strong>alysis to meet NCEP requirements<br />

of two IMS observations per day. All NIC personnel will undergo <strong>the</strong> same training <strong>as</strong> SAB<br />

personnel <strong>an</strong>d meet IMS qualifications before being <strong>as</strong>signed to IMS product generation. Parallel<br />

product generation will be performed at <strong>the</strong> SAB <strong>an</strong>d NIC until comparable output products are<br />

obtained between offices. After evaluation <strong>an</strong>d duplication of <strong>the</strong> IMS h<strong>as</strong> been achieved,<br />

production of <strong>the</strong> IMS will tr<strong>an</strong>sition fully to <strong>the</strong> NIC. No production methods o<strong>the</strong>r <strong>the</strong>n<br />

personnel will ch<strong>an</strong>ge during this process.<br />

Longer-term pl<strong>an</strong>s for enh<strong>an</strong>cements to <strong>the</strong> IMS input data revolve around <strong>the</strong> future<br />

deployment of NPOESS <strong>an</strong>d GOES-R. NPOESS will be a joint military <strong>an</strong>d civili<strong>an</strong> satellite<br />

replacing m<strong>an</strong>y of <strong>the</strong> existing U.S. polar orbiting sensors with improved sensors such <strong>as</strong> <strong>the</strong><br />

Visible Infrared Imager Radiometer Suite (VIIRS), Conical Sc<strong>an</strong>ning Microwave Imager/Sounder<br />

(CMIS), <strong>an</strong>d Adv<strong>an</strong>ced Technology Microwave Sounder (ATMS).<br />

GOES-R will be <strong>the</strong> next generation of NOAA geostationary satellites that will aide snow <strong>an</strong>d<br />

ice observations. Geostationary satellites are regarded by <strong>an</strong>alysts <strong>as</strong> <strong>the</strong> most valuable input<br />

source. IMS <strong>an</strong>alysts <strong>as</strong>sessing surface cryospheric coverage will benefit greatly by <strong>the</strong><br />

adv<strong>an</strong>cements in remote sensing provided on GOES-R, particularly <strong>the</strong> Adv<strong>an</strong>ced B<strong>as</strong>eline Imager<br />

(ABI). The ABI will have 16 spectral b<strong>an</strong>ds, compared with five on <strong>the</strong> current GOES imagers.<br />

The ABI will improve <strong>the</strong> spatial coverage from 1km to 0.5km at nadir for broadb<strong>an</strong>d visible <strong>an</strong>d<br />

from 4km to 2km for <strong>the</strong> infrared b<strong>an</strong>ds. For snow <strong>an</strong>d ice detection this will improve <strong>the</strong> ability<br />

of <strong>the</strong> <strong>an</strong>alyst to confirm <strong>the</strong> presence of ice on <strong>the</strong> surface through recognition of spatial patterns<br />

more discernable at higher resolutions, such <strong>as</strong> dendritic spatial patterns of snow on mountains, ice<br />

floe shapes that indicate certain ice thickness, or ice fractures. The ABI also includes spectral<br />

176


information never present before on GOES imagers. One b<strong>an</strong>d of particular relev<strong>an</strong>ce to snow <strong>an</strong>d<br />

ice detection will be centered at 1.61μm, which would exp<strong>an</strong>d <strong>the</strong> ice/cloud discrimination<br />

sampling beyond <strong>the</strong> temporally coarse polar orbiters (Schmit et al., 2005). This should improve<br />

<strong>the</strong> IMS accuracy <strong>an</strong>d reduce <strong>the</strong> amount of time required for detection. This ch<strong>an</strong>nel differencing<br />

would also improve automated snow <strong>an</strong>d ice detection. GOES-R will incre<strong>as</strong>e <strong>the</strong> coverage<br />

acquisition rate by nearly fivefold, allowing closer to real-time observations <strong>an</strong>d incre<strong>as</strong>ed<br />

discrimination of relatively static surface features from highly dynamic atmospheric features.<br />

Figure 5. Examples of <strong>the</strong> experimental NOAA merged AMSU SWE-IMS output (bottom) for February 1,<br />

2006 over <strong>the</strong> Nor<strong>the</strong>rn Hemisphere. The pre-merged AMSU SWE (top) h<strong>as</strong> erroneous or questionable<br />

signals m<strong>as</strong>ked by using <strong>the</strong> IMS.<br />

177


SUMMARY<br />

The IMS underwent enh<strong>an</strong>cements to <strong>the</strong> resolutions, input data, methodology, <strong>an</strong>d output<br />

formats that augmented <strong>the</strong> product’s utility. These ch<strong>an</strong>ges have been largely beneficial to <strong>the</strong><br />

NCEP EMC modeler (Mitchell, K., 2006, professional conversation). The ch<strong>an</strong>ges involve <strong>an</strong><br />

improved architecture, superior output resolution, exp<strong>an</strong>ded input sources, <strong>an</strong>d topographic<br />

mapping capabilities. The current system architecture allows IMS <strong>an</strong>alysts to quickly record snow<br />

<strong>an</strong>d ice, thus speeding <strong>the</strong> production time. The adv<strong>an</strong>ced resolution product begun in February<br />

2004 allows for greater detail in <strong>the</strong> snow <strong>an</strong>d ice information to be conveyed. While product<br />

evaluations are still ongoing at NCEP, improvements in <strong>the</strong> resolution are likely to have a positive<br />

impact on numerical wea<strong>the</strong>r prediction. The exp<strong>an</strong>ded imagery from which <strong>the</strong> user may derive<br />

data sources allows for <strong>an</strong> incre<strong>as</strong>e in <strong>the</strong> likelihood of correct snow <strong>an</strong>d ice identification. The<br />

more accurately <strong>the</strong> surface cryosphere is depicted, <strong>the</strong> better ch<strong>an</strong>ce a wea<strong>the</strong>r prediction model<br />

should have at accurate forec<strong>as</strong>ts.<br />

The enh<strong>an</strong>cements made to <strong>the</strong> IMS were mostly driven by <strong>the</strong> need for better NCEP EMC<br />

initializations. Even <strong>the</strong> pl<strong>an</strong>ned ch<strong>an</strong>ges in <strong>the</strong> product are geared to improving NCEP EMC<br />

forec<strong>as</strong>ts. Unfortunately, <strong>the</strong> impacts of <strong>the</strong> enh<strong>an</strong>cements to NCEP CPC climate evaluation<br />

remain unknown. The product h<strong>as</strong> likely become more accurate due to improved spatial<br />

resolution, enh<strong>an</strong>ced methodologies, <strong>an</strong>d improved input sources. However, <strong>the</strong> impact of <strong>an</strong><br />

improved product may cause heterogeneity in <strong>the</strong> snow <strong>an</strong>d ice data record is still under<br />

evaluation. While consultation of ch<strong>an</strong>ge impacts is sought by <strong>the</strong> CPC <strong>an</strong>d non-federal<br />

researchers, formal examinations are to date forthcoming. This concern is not without note <strong>an</strong>d is<br />

in need of fur<strong>the</strong>r investigation.<br />

Hopefully, providing <strong>the</strong> data in new formats <strong>an</strong>d with accomp<strong>an</strong>ying products will exp<strong>an</strong>d <strong>the</strong><br />

user community for <strong>the</strong> products. Fur<strong>the</strong>r adv<strong>an</strong>ces in <strong>the</strong> products are predicated on customer<br />

requirements, new sensors, <strong>an</strong>d adv<strong>an</strong>ced technologies. The adv<strong>an</strong>cements in <strong>the</strong> IMS, <strong>as</strong> well <strong>as</strong><br />

links to p<strong>as</strong>t snow mapping, should continue to make this a viable snow mapping system for years<br />

to come.<br />

ACKNOWLEDGEMENTS<br />

The authors th<strong>an</strong>k Dr. D<strong>an</strong> Tarpley <strong>an</strong>d Dr. Peter Rom<strong>an</strong>ov, Office of Research <strong>an</strong>d<br />

Applications, NOAA/NESDIS; <strong>an</strong>d Dr. Cezar Kongoli, QSS Group Inc. for <strong>the</strong>ir direct<br />

contributions to this work. The authors also wish to acknowledge <strong>the</strong> import<strong>an</strong>t contributions <strong>an</strong>d<br />

support of, Dr. Ken Mitchell, NOAA/NWS; Mr. Ricky Irving, PIB NOAA/NESDIS; <strong>an</strong>d Dr.<br />

David Robinson, Rutgers University.<br />

REFERENCES<br />

Armstrong, R. L. <strong>an</strong>d M. J. Brodzik, 2001. ‘Recent Nor<strong>the</strong>rn Hemisphere snow extent: a<br />

comparison of data derived from visible <strong>an</strong>d microwave satellite sensors.’ Geophysical Res.<br />

Letters, 28(19), 3673–3676.<br />

Balk, B.C., <strong>an</strong>d K. Elder. 2000. ‘Combining binary decision tree <strong>an</strong>d geostatistical methods to<br />

estimate snow distribution in a mountain watershed.’ Water Resources Research., 36, 13–26.<br />

Bamzai, A. S. <strong>an</strong>d L. Marx, 2000: ‘COLA AGCM simulation of <strong>the</strong> effect of <strong>an</strong>omalous spring<br />

snow over Eur<strong>as</strong>ia on <strong>the</strong> Indi<strong>an</strong> summer monsoon.’ Quart. J. Roy. Meteor. Soc., 126, 2575.<br />

Barrett, A., 2003. National Operational Hydrologic Remote Sensing Center SNOw Data<br />

Assimilation System (SNODAS) Products at NSIDC. NSIDC Special Report 11. Boulder, CO,<br />

USA: National <strong>Snow</strong> <strong>an</strong>d Ice Data Center. 19 pages<br />

B<strong>as</strong>ist, A., D. Garrett, R. Ferraro, N. Grody, <strong>an</strong>d K. Mitchell, 1996. ‘A comparison between snow<br />

cover products derived from visible <strong>an</strong>d microwave satellite observations.’ J. Appl. Meteorol.,<br />

35, 163–177.<br />

178


Brubaker, K.L., R. Pinker <strong>an</strong>d E. Deviatova, 2005. ‘Evaluation <strong>an</strong>d Comparison of MODIS <strong>an</strong>d<br />

IMS <strong>Snow</strong> Cover Estimates for <strong>the</strong> Continental United States Using Station Data’. J.<br />

Hydrometeorology, 6(6), 1002–1017.<br />

Ch<strong>an</strong>g, A.T.C., J.L. Foster <strong>an</strong>d D.K. Hall, 1996. ‘Effects of forest on <strong>the</strong> snow parameters derived<br />

from microwave me<strong>as</strong>urements during <strong>the</strong> BOREAS winter field campaign.’ Hydrol.<br />

Processes, 10, 1565–1574.<br />

Clark, M.P., <strong>an</strong>d M.C. Serreze, 2000. ‘Effects of variations in E<strong>as</strong>t Asi<strong>an</strong> snow cover on<br />

modulating atmospheric circulation over <strong>the</strong> North Pacific Oce<strong>an</strong>.’ J. Climate, 13, 3700–3710.<br />

Dudhia, J., D. Gill, Y-R. Guo, D. H<strong>an</strong>sen, K. M<strong>an</strong>ning, <strong>an</strong>d W. W<strong>an</strong>g, 1999. PSU/NCAR<br />

Mesoscale Modeling System Tutorial Cl<strong>as</strong>s Notes <strong>an</strong>d Users' Guide: MM5 Modeling System<br />

Version 2. Mesoscale <strong>an</strong>d Microscale Meteorology Division, NCAR, 182 pages.<br />

Foster, J.L., Hall, D.K., Ch<strong>an</strong>g, A.T.C., R<strong>an</strong>go, A., Wergin, W. <strong>an</strong>d Erbe, E. 1999. ‘Effects of<br />

snow crystal shape on <strong>the</strong> scattering of p<strong>as</strong>sive microwave radiation.’ IEEE Tr<strong>an</strong>s. Geosci.<br />

Remote Sens., 37, 1165–1168.<br />

Foster, J., C. Sun, J.P. Walker, R. Kelly, J. Dong, <strong>an</strong>d A. Ch<strong>an</strong>g, 2004. ‘Mapping r<strong>an</strong>dom <strong>an</strong>d<br />

systematic errors of satellite-derived snow water equivalent observations in Eur<strong>as</strong>ia’, Proc.<br />

SPIE Int. Soc. Opt. Eng.,, M. Owe, G. D'Urso, B.T. Gouweleeuw, <strong>an</strong>d A.M. Jochum, Eds.,<br />

5568, M<strong>as</strong>palom<strong>as</strong>, Gr<strong>an</strong> C<strong>an</strong>aria, Spain, September 14–16, 2004, 150–158.<br />

Gong, G, D. Entekhabi, <strong>an</strong>d J. Cohen. 2003. ‘Relative impacts of Siberi<strong>an</strong> <strong>an</strong>d North Americ<strong>an</strong><br />

snow <strong>an</strong>omalies on <strong>the</strong> winter Arctic Oscillation’, Geophysical Res. Letters, 30(16), 1848.<br />

doi:10.1029/2003GL017749.<br />

Grumbine, R. W., 1996. Automated P<strong>as</strong>sive Microwave Sea Ice Concentration Analysis at NCEP<br />

OMB Tech. Note 120, 13 pages.<br />

Grumbine, R. W., 2005. http://meted.ucar.edu/nwp/pcu2/avice1.htm.<br />

Hahn, D. J. <strong>an</strong>d J. Shukla, 1976. ‘An apparent relationship between Eur<strong>as</strong>i<strong>an</strong> snow cover <strong>an</strong>d<br />

Indi<strong>an</strong> monsoon rainfall.’ J. Atmos. Sci., 33, 2461–2462.<br />

Hall, D.K., J.L. Foster, V.V. Salomonson, A.G. Klein <strong>an</strong>d J.Y.L. Chien, 2001. "Development of a<br />

Technique to Assess <strong>Snow</strong>-Cover Mapping Errors from Space," IEEE Tr<strong>an</strong>sactions on<br />

Geoscience <strong>an</strong>d Remote Sensing, 39(2), 432–438.<br />

Hall, D.K., A.B. Tait, J.L. Foster, A.T.C. Ch<strong>an</strong>g <strong>an</strong>d M. Allen, 2000. ‘Intercomparison of satellitederived<br />

snow-cover maps.’ Annals of Glaciology, 31, 369–376.<br />

Huaqi<strong>an</strong>g, L., S. Zhaobo, W. Ju, <strong>an</strong>d M. Jinzhong. 2004 ‘A Modeling Study of <strong>the</strong> Effects of<br />

Anomalous <strong>Snow</strong> Cover over <strong>the</strong> Tibet<strong>an</strong> Plateau upon <strong>the</strong> South Asi<strong>an</strong> Summer Monsoon.’<br />

Adv<strong>an</strong>ces in Atmospheric Sciences, 21(6), 964–975.<br />

Kongoli, C, C.A.De<strong>an</strong>, S.R. Helfrich, <strong>an</strong>d R.R.Ferraro, 2006. ‘The retrievals of snow cover extent<br />

<strong>an</strong>d snow water equivalent from a blended p<strong>as</strong>sive microwave – interactive multi-sensor snow<br />

product’ Proceedings of <strong>the</strong> 63rd E<strong>as</strong>tern <strong>Snow</strong> Conference. 7–9 June, 2006. Newark,<br />

Delaware.<br />

March<strong>an</strong>d, W-D <strong>an</strong>d Å. Killingtveit, 2001 ‘Analyses of <strong>the</strong> Relation between Spatial <strong>Snow</strong><br />

Distribution <strong>an</strong>d Terrain Characteristics.’ Proceedings of <strong>the</strong> 58th E<strong>as</strong>tern <strong>Snow</strong> Conference.<br />

14–17 May, 2001. Ottawa, C<strong>an</strong>ada.<br />

Marks, D.G., Winstral, A.H., <strong>an</strong>d Seyfried, M.S. 2002. ‘Simulation of Terrain <strong>an</strong>d Forest Shelter<br />

Effects on Patterns of <strong>Snow</strong> Deposition, <strong>Snow</strong>melt <strong>an</strong>d Runoff over a Semi-Arid Mountain<br />

Catchment.’ Hydrol. Processes, 16, 3605–3626.<br />

Mitchell, K. E., G. J. DiMego, <strong>an</strong>d T. L. Black, 1993. ‘C<strong>as</strong>e study on <strong>the</strong> impact of improved snow<br />

cover <strong>an</strong>d sea ice on wintertime forec<strong>as</strong>ts of NMC’s regional NGM <strong>an</strong>d Eta models.’ Preprints,<br />

13th Conf. of Wea<strong>the</strong>r Analysis <strong>an</strong>d Forec<strong>as</strong>ting, Vienna, VA, Amer. Meteor. Soc., 516–517.<br />

Ramsay, B., 1998. ‘The interactive multisensor snow <strong>an</strong>d ice mapping system.’ Hydrol. Processes,<br />

12, 1537–1546.<br />

Ramsay, B., 2000. ‘Prospects for <strong>the</strong> Interactive Multisensor <strong>Snow</strong> <strong>an</strong>d Ice Mapping<br />

System.(IMS)’ Proceedings of <strong>the</strong> 57th E<strong>as</strong>tern <strong>Snow</strong> Conference. 18–19 May 2000, Syracuse,<br />

New York. 161–170.<br />

Robinson, D.A., K.F. Dewey <strong>an</strong>d R.R. Heim Jr., 1993. ‘Global snow cover monitoring: <strong>an</strong><br />

update.’ Bull. Am. Meteorol. Soc., 74, 1689–1696.<br />

179


Robinson, D.A., 2003. ‘Recent variability of nor<strong>the</strong>rn hemisphere snow cover.’ Preprints: Seventh<br />

Conference on Polar Meteorology <strong>an</strong>d Oce<strong>an</strong>ography, Hy<strong>an</strong>nis, MA, Americ<strong>an</strong><br />

Meteorological Society, paper 13.12, 6.<br />

Rogers, E., T. Black, B. Ferrier, Y. Lin, D. Parrish, <strong>an</strong>d G. DiMego, 2001. ‘Ch<strong>an</strong>ges to <strong>the</strong> NCEP<br />

Meso Eta Analysis <strong>an</strong>d Forec<strong>as</strong>t System: Incre<strong>as</strong>e in resolution, new cloud microphysics,<br />

modified precipitation <strong>as</strong>similation, modified 3DVAR <strong>an</strong>alysis.’ NWS Technical Procedures<br />

Bulletin 488, NOAA/NWS, W<strong>as</strong>hington DC.<br />

Rom<strong>an</strong>ov, P., Gutm<strong>an</strong> G. <strong>an</strong>d Csiszar I., 2000. ‘Automated monitoring of snow cover over North<br />

America with multispectral satellite data’ J. Appl. Meteorol., 39, 1866–1880.<br />

Rom<strong>an</strong>ov, P., D. Tarpley, G. Gutm<strong>an</strong>, <strong>an</strong>d T. Carroll, 2003. ‘Mapping <strong>an</strong>d monitoring of <strong>the</strong> snow<br />

cover fraction over North America.’ J. Geophys. Res., 108(D16), 8619.<br />

Rom<strong>an</strong>ov, P. <strong>an</strong>d D.Tarpley, 2003. ‘Automated monitoring of snow cover over South America<br />

using GOES Imager data. Int. J. Remote Sens., 24(5), 1119–1125.<br />

Schmit T.J., M. M. Gunshor, W.P. Menzel, J. Li, S. Bachmeier, J.J. Gurka, 2005. ‘Introducing <strong>the</strong><br />

Next-generation Adv<strong>an</strong>ced B<strong>as</strong>eline Imager (ABI) on GOES-R’. Bull. Amer. Meteor. Soc., 8,<br />

1079–109.<br />

Sham<strong>an</strong>, J., M.A.C<strong>an</strong>e, A.Kapl<strong>an</strong>, 2005. ‘The relationship between Tibet<strong>an</strong> snow depth, ENSO,<br />

river discharge <strong>an</strong>d <strong>the</strong> monsoons of B<strong>an</strong>gladesh.’ Int. J. Remote Sens., 26(17), 3735–3748.<br />

Sheffeild, J.; P<strong>an</strong>, M.; Wood, E.F.; Mitchell, K.E.; Houser, P.R.; Schaake, J.C.; Robock, A.;<br />

Lohm<strong>an</strong>n, D.; Cosgrove, B.; Du<strong>an</strong>, Q.; Luo, L.; Higgins, R.W.; Pinker, R.T.; Tarpley, J.D.;<br />

Ramsay, B.H., Journal of Geophysical Research-Atmospheres, November 21, 2003 (Vol. 108,<br />

No. D22, 8849, Paper No. 10.1029/2002JD003274)<br />

United States Department of Commerce, National Oce<strong>an</strong>ic <strong>an</strong>d Atmospheric Administration<br />

(USDOC / NOAA), 2005. New Priorities for <strong>the</strong> 21 st Century – NOAA’s Strategic Pl<strong>an</strong>,<br />

Updated for FY 2006 – FY 2011. NOAA's Strategic Pl<strong>an</strong>ning Office, Silver Spring, MD.<br />

Wadhams, P., 2000. Ice in <strong>the</strong> oce<strong>an</strong>. Gordon <strong>an</strong>d Breach Science Publishers. 351.<br />

180


181<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Retreat of Tropical Glaciers in Colombia <strong>an</strong>d Venezuela<br />

from 1984 to 2004 <strong>as</strong> Me<strong>as</strong>ured from ASTER <strong>an</strong>d L<strong>an</strong>dsat Images<br />

ABSTRACT<br />

JENNIFER N. MORRIS, 1 ALAN J. POOLE, 2 AND ANDREW G. KLEIN 3<br />

Like glaciers throughout <strong>the</strong> world, tropical glaciers in <strong>the</strong> nor<strong>the</strong>rn Andes have retreated during<br />

<strong>the</strong> l<strong>as</strong>t century. In this study, <strong>the</strong> retreat of Ande<strong>an</strong> glaciers in Colombi<strong>an</strong> <strong>an</strong>d Venezuela h<strong>as</strong> been<br />

mapped from L<strong>an</strong>dsat <strong>an</strong>d ASTER images acquired between 1984 <strong>an</strong>d 2004. Glacier retreat h<strong>as</strong><br />

been mapped for three glaciated regions in Columbia, <strong>the</strong> Sierra Nevada de S<strong>an</strong>ta Marta, <strong>the</strong> Sierra<br />

Nevada del Cocuy, <strong>an</strong>d <strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif <strong>as</strong> h<strong>as</strong> <strong>the</strong> retreat of <strong>the</strong> single remaining glacier<br />

on <strong>the</strong> Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif in Venezuela. From <strong>the</strong> L<strong>an</strong>dsat archives, several satellite images<br />

sp<strong>an</strong>ning <strong>the</strong> L<strong>an</strong>dsat record were selected b<strong>as</strong>ed on cloud cover over <strong>the</strong> glaciers <strong>an</strong>d which<br />

provided adequate temporal coverage. All selected satellite images of <strong>the</strong> study sites were coregistered<br />

to each o<strong>the</strong>r with a root me<strong>an</strong> square (RMS) error for <strong>the</strong> ground control points of less<br />

th<strong>an</strong> 20 meters. <strong>Snow</strong> <strong>an</strong>d ice extent in each image w<strong>as</strong> cl<strong>as</strong>sified using <strong>the</strong> Normalized Difference<br />

<strong>Snow</strong> Index (NDSI) method <strong>an</strong>d a density slice w<strong>as</strong> used to create a binary snow/ice map. The<br />

total glaciated area on each individual peak w<strong>as</strong> <strong>the</strong>n determined from <strong>the</strong> cl<strong>as</strong>sified snow <strong>an</strong>d ice<br />

are<strong>as</strong>. This approach w<strong>as</strong> found to be effective in determining glacier area. Incorporating historical<br />

data, a time series of glacier retreat from <strong>the</strong> early 1950s to 2003 w<strong>as</strong> established. All four studied<br />

regions studied showed glacier retreat over this period. In <strong>the</strong> 1950s, <strong>the</strong> overall area of glaciers<br />

studied in Colombia <strong>an</strong>d Venezuela glaciers w<strong>as</strong> 91.36 km 2 in <strong>the</strong> 1950s. By 2003, <strong>the</strong> glaciers<br />

had retreated to approximately 62.07 km 2 which represents a total ice loss of 32%.<br />

Keywords: Remote Sensing, Colombia, Venezuela, Tropical Glaciers<br />

INTRODUCTION<br />

As one indicator of global ch<strong>an</strong>ge, glaciers around <strong>the</strong> world have been studied. Topical glaciers<br />

are of particular interest to study because <strong>the</strong> distinct characteristics of tropical climates make<br />

glacier-climate interactions different from <strong>the</strong> mid- <strong>an</strong>d high-latitudes (K<strong>as</strong>er 1999). However, not<br />

all tropical glaciers have been comprehensively studied because <strong>the</strong>ir remoteness c<strong>an</strong> limit<br />

accessibility <strong>an</strong>d extensive cloud cover c<strong>an</strong> preclude extended temporal studies from remote<br />

sensing. The recent incre<strong>as</strong>e in availability of satellite images over glacier regions h<strong>as</strong> incre<strong>as</strong>ed<br />

<strong>the</strong> likelihood of studies throughout <strong>the</strong>se regions.<br />

The tropical glaciers of Colombia <strong>an</strong>d Venezuela, <strong>the</strong> focus of this study, were at a recent<br />

maximum extent during <strong>the</strong> Little Ice Age <strong>an</strong>d have receded since this time (K<strong>as</strong>er 1999).<br />

1<br />

Department of Geography, MS 3147, Tex<strong>as</strong> A&M University, College Station TX 77843-3147<br />

USA email: jenmorris@tamu.edu.<br />

2<br />

Department of Geography, MS 3147, Tex<strong>as</strong> A&M University, College Station TX 77843-3147<br />

USA email: al<strong>an</strong>poole@tamu.edu<br />

3<br />

Department of Geography, MS 3147, Tex<strong>as</strong> A&M University, College Station TX 77843-3147<br />

USA email: klein@geog.tamu.edu


Utilizing historical data <strong>an</strong>d satellite observations, glacier retreat in Columbia <strong>an</strong>d Venezuela is<br />

me<strong>as</strong>ured over <strong>the</strong> p<strong>as</strong>t four decades.<br />

The objective of this study is to qu<strong>an</strong>tify ch<strong>an</strong>ges in <strong>the</strong> area of glaciers in several are<strong>as</strong> of<br />

Colombia <strong>an</strong>d Venezuela using both historical data <strong>an</strong>d me<strong>as</strong>urements from satellite images to<br />

construct a time series starting in <strong>the</strong> early 1950’s <strong>an</strong>d ending in 2004. Glaciers in <strong>the</strong> following<br />

four regions have been studied: <strong>the</strong> Pico Bonplad M<strong>as</strong>sif in Venezuela, <strong>the</strong> Sierra Nevada de S<strong>an</strong>ta<br />

Marta, <strong>the</strong> Sierra Nevada de Cocuy <strong>an</strong>d <strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif in Columbia. The study locations<br />

are illustrated in Figure 1. Glacier retreat on <strong>the</strong> Nevado del Huila, <strong>the</strong> only o<strong>the</strong>r glaciated region<br />

in Colombia, w<strong>as</strong> not studied.<br />

Figure 1: Map of Study Area<br />

STUDY SITES AND THEIR GLACIAL HISTORY<br />

The glaciers of <strong>the</strong> nor<strong>the</strong>rn tropical Andes Mountains studied are located in Venezuela <strong>an</strong>d<br />

Colombia along <strong>the</strong> western edge of South America <strong>an</strong>d c<strong>an</strong> be broken into four distinct glacial<br />

regions (Figure 1).<br />

The Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif, of Venezuela, w<strong>as</strong> formally <strong>the</strong> site of numerous glaciers extending<br />

around <strong>the</strong> peak with <strong>an</strong> elevation of 4,983 meters; now it is site of <strong>the</strong> l<strong>as</strong>t me<strong>as</strong>urable Venezuel<strong>an</strong><br />

glacier, Sinigüis located at 8º32′ N <strong>an</strong>d 72º00′ W.<br />

The Sierra Nevada de S<strong>an</strong>ta Marta in northwestern Colombia is part of <strong>the</strong> Cordillera Central<br />

br<strong>an</strong>ch of <strong>the</strong> Andes Mountains extending into Colombia. Located near <strong>the</strong> Pacific co<strong>as</strong>t, <strong>the</strong><br />

Sierra Nevada de S<strong>an</strong>ta Marta is <strong>the</strong> site of numerous glaciers located on a single peak at 10º34´N<br />

<strong>an</strong>d 73º43´W.<br />

182


The Sierra Nevada del Cocuy region of <strong>the</strong> Cordillera Oriental in <strong>the</strong> Columbi<strong>an</strong> Andes is<br />

located in north-central Colombia, south of <strong>the</strong> Venezuel<strong>an</strong> border. The glaciated r<strong>an</strong>ge trends<br />

north–south, <strong>an</strong>d is located at 6º27′ N <strong>an</strong>d 72º18′ W.<br />

The Ruiz-Tolima M<strong>as</strong>sif is located in <strong>the</strong> Colombi<strong>an</strong> Parque Nacional de los Nevados on <strong>the</strong><br />

Cordillera Central chain of <strong>the</strong> Andes Mountains at 4º50´N <strong>an</strong>d 75º20´W. In 1959, <strong>the</strong> Parque<br />

Nacional de los Nevados contained 53 glaciers when <strong>the</strong> glaciers were me<strong>as</strong>ured by aerial<br />

photography (Hoy<strong>as</strong>-Patiño 1998).<br />

Work by Thouret et al. (1996) suggests that <strong>the</strong> glacier extent of <strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif, Sierra<br />

Nevada de S<strong>an</strong>ta Marta, <strong>an</strong>d Sierra Nevada del Cocuy of Colombia were approximately 1500 km 2 ,<br />

1500 km 2 , <strong>an</strong>d 2000 km 2 , respectively, during <strong>the</strong> l<strong>as</strong>t glacial maximum which occurred<br />

approximately 27,000–24,000 years BP.<br />

A later major glaciation occurred before 13,000–12,400 years BP at which time <strong>the</strong> glaciers of<br />

<strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif had <strong>an</strong> approximate area of 800 km 2 . During <strong>the</strong> late neoglacial period,<br />

during <strong>the</strong> Little Ice Age which occurred from <strong>the</strong> 1600’s to <strong>the</strong> 1900’s, <strong>the</strong> glacierized area w<strong>as</strong><br />

reduced to 100 km 2 . The Sierra Nevada de S<strong>an</strong>ta Marta had <strong>an</strong> approximate glacier area of 850<br />

km 2 around 13,000–12,400 years BP, <strong>an</strong>d 107 km 2 during <strong>the</strong> late neoglacial period. Glaciers in<br />

<strong>the</strong> Sierra Nevada del Cocuy showed similar retreat with <strong>an</strong> approximate glacier size of 1000 km 2<br />

around 13,000–12,400 years BP <strong>an</strong>d 150 km 2 during <strong>the</strong> Little Ice Age.<br />

Similar to <strong>the</strong> Colombi<strong>an</strong> glacier regions <strong>an</strong>d o<strong>the</strong>r glaciers in <strong>the</strong> tropical <strong>an</strong>d temperate zones,<br />

Venezuel<strong>an</strong> glaciers were affected by Quaternary glaciations. Between 20,000–13,000 years BP,<br />

glaciers in <strong>the</strong> area adv<strong>an</strong>ced. This period is denoted <strong>as</strong> <strong>the</strong> Mérida Glaciation. The Cordillera de<br />

Mérida br<strong>an</strong>ch of <strong>the</strong> nor<strong>the</strong>rn Andes Mountains is <strong>the</strong> site to high peaks tall enough to support<br />

glacier conditions. At <strong>the</strong> end of <strong>the</strong> Merida Glaciation period, <strong>the</strong> Cordillera de Mérida chain had<br />

<strong>an</strong> approximate glacier area of 600 km 2 . Since 13,000 years BP, signific<strong>an</strong>t decre<strong>as</strong>es in glacial<br />

size are apparent (Schubert <strong>an</strong>d Clapperton 1990).<br />

DATA AND METHODS<br />

Satellite Images<br />

ASTER <strong>an</strong>d L<strong>an</strong>dsat TM <strong>an</strong>d ETM+ images from 1984 to 2004 were used in this study. Images,<br />

listed in Table 1, were selected b<strong>as</strong>ed on minimum amounts of snow <strong>an</strong>d cloud cover <strong>an</strong>d on<br />

which produced <strong>the</strong> best time series for each region. All images available for each region were<br />

reviewed <strong>an</strong>d critiqued b<strong>as</strong>ed on <strong>the</strong> following criteria: se<strong>as</strong>onal precipitation, cloud cover, extent<br />

of snow cover, <strong>an</strong>d image acquisition date. Selection of images w<strong>as</strong> limited by se<strong>as</strong>onal patterns in<br />

precipitation <strong>an</strong>d temperature, causing incre<strong>as</strong>ed cloud cover in <strong>the</strong> area. Following visual <strong>an</strong>alysis<br />

of <strong>the</strong> initially screened images, only those images with minimal snow <strong>an</strong>d cloud cover were<br />

selected for fur<strong>the</strong>r <strong>an</strong>alysis.<br />

ASTER (Adv<strong>an</strong>ced Space-borne Thermal Emission <strong>an</strong>d Reflection Radiometer) is <strong>an</strong> imaging<br />

instrument aboard NASA’s Terra satellite which w<strong>as</strong> launched in 1999 <strong>as</strong> a part of <strong>the</strong> Earth<br />

Observing System (EOS). Each ASTER scene covers <strong>an</strong> approximate area of 60 km by 60 km<br />

with 15 spectral b<strong>an</strong>ds at 3 spatial resolutions. The visual near infrared (VNIR) b<strong>an</strong>ds, short-wave<br />

infrared (SWIR) b<strong>an</strong>ds, <strong>an</strong>d <strong>the</strong>rmal infrared (TIR) b<strong>an</strong>ds of ASTER have 15, 30, <strong>an</strong>d 90 meter<br />

resolution, respectively (Abrams 2000).<br />

Each L<strong>an</strong>dsat scene covers a larger area, approximately 185 km by 185 km. The L<strong>an</strong>dsat<br />

Thematic Mapper (TM) on L<strong>an</strong>dsat 4 <strong>an</strong>d 5 h<strong>as</strong> seven b<strong>an</strong>ds; six b<strong>an</strong>ds in <strong>the</strong> visible to short-wave<br />

infrared have a spatial resolution of 30 meters while <strong>the</strong> remaining <strong>the</strong>rmal b<strong>an</strong>ds have 120 meter<br />

spatial resolution. The Enh<strong>an</strong>ced Thematic Mapper Plus (ETM+), carried on <strong>the</strong> L<strong>an</strong>dsat 7<br />

satellite, h<strong>as</strong> eight multispectral b<strong>an</strong>ds with six b<strong>an</strong>ds at 30 meter spatial resolution, two b<strong>an</strong>ds<br />

with 60 meter resolution, <strong>an</strong>d one p<strong>an</strong>chromatic b<strong>an</strong>d at 15 meter resolution (L<strong>an</strong>dsat Project<br />

Science Office 2006).<br />

183


Colombia<br />

Venezuela<br />

Table 1. Optimum images of <strong>the</strong> glaciers of Venezuela <strong>an</strong>d Colombia<br />

Glacier Series Date Sensor Type Path Row<br />

Ruiz-Tolima<br />

M<strong>as</strong>sif<br />

Sierra Nevada<br />

de S<strong>an</strong>ta Marta<br />

Sierra Nevada<br />

del Cocuy<br />

Pico Bonpl<strong>an</strong>d<br />

M<strong>as</strong>sif<br />

(Siniguis)<br />

01/01/1959 * Aerial — —<br />

02/01/1976 * L<strong>an</strong>dsat 1 9 57<br />

10/24/1997 L<strong>an</strong>dsat 5 9 57<br />

01/28/2001 L<strong>an</strong>dsat 7 9 57<br />

01/01/1957 * Aerial — —<br />

01/01/1973 * L<strong>an</strong>dsat 1 8 53<br />

12/16/1984 L<strong>an</strong>dsat 5 8 53<br />

01/18/1991 L<strong>an</strong>dsat 4 8 53<br />

12/20/2000 ASTER 8 53<br />

01/14/2004 ASTER 8 53<br />

01/01/1959 * Aerial — —<br />

01/18/1973 * L<strong>an</strong>dsat 1 7 56<br />

01/13/1986 L<strong>an</strong>dsat 5 7 56<br />

12/23/1992 L<strong>an</strong>dsat 4 7 56<br />

05/25/1999 L<strong>an</strong>dsat 5 7 56<br />

01/14/2001 ASTER 7 56<br />

01/30/2001 ASTER 7 56<br />

03/06/2002 ASTER 7 56<br />

03/02/2003 ASTER 7 56<br />

01/01/1952 * Aerial — —<br />

03/24/1985 L<strong>an</strong>dsat 5 6 54<br />

01/20/1988 L<strong>an</strong>dsat 4 6 54<br />

01/29/2000 L<strong>an</strong>dsat 5 6 54<br />

02/11/2002 ASTER 6 54<br />

03/02/2003 ASTER 6 54<br />

02/01/2004 ASTER 6 54<br />

* Me<strong>as</strong>urements taken from Hoyos-Patiño (1998) <strong>an</strong>d Schubert (1998)I<br />

Image Preprocessing<br />

For ASTER scenes, <strong>the</strong> six SWIR b<strong>an</strong>ds were spatially oversampled to 15 meter resolution to<br />

match <strong>the</strong> VNIR b<strong>an</strong>ds using nearest neighbor resampling. For <strong>the</strong> L<strong>an</strong>dsat images, all b<strong>an</strong>ds were<br />

oversampled from 28.5 meter to 15 meter resolution. Resizing all images to 15 meters allowed for<br />

better comparisons of snow <strong>an</strong>d ice cl<strong>as</strong>sification between <strong>the</strong> ASTER <strong>an</strong>d L<strong>an</strong>dsat platforms <strong>an</strong>d<br />

facilitated visual interpretation of glacier area <strong>an</strong>d comparisons of <strong>the</strong> cl<strong>as</strong>sified glacier extent<br />

between platforms.<br />

The images for each glacier region were coregistered using ground control points (GCPs)<br />

spaced throughout each scene. A minimum of 9 GCPs were used with all images. Images for each<br />

glacier area were coregistered to a m<strong>as</strong>ter image (Table 2) selected for each glacier series. All<br />

images were projected into Universal Tr<strong>an</strong>sverse Mercator, Zone 19N with a WGS-84 datum.<br />

Once <strong>the</strong> images in a glacial series were coregistered, <strong>the</strong>y were overlaid to visualize glacial<br />

retreat over time for each area.<br />

184


Colombia<br />

Venezuela<br />

Glacier<br />

Series<br />

Ruiz-Tolima<br />

M<strong>as</strong>sif<br />

Sierra<br />

Nevada de<br />

S<strong>an</strong>ta Marta<br />

Sierra<br />

Nevada del<br />

Cocuy<br />

Pico<br />

Bonpl<strong>an</strong>d<br />

M<strong>as</strong>sif<br />

(Sinigüis)<br />

Table 2. M<strong>as</strong>ter Images for Image Coregistration<br />

Date<br />

185<br />

Sensor<br />

Type<br />

L<strong>an</strong>dsat<br />

WRS Path<br />

L<strong>an</strong>dsat<br />

WRS Row<br />

01/28/2001 L<strong>an</strong>dsat 7 9 57<br />

01/14/2004 ASTER 8 53<br />

03/06/2002 ASTER 7 56<br />

02/11/2002 ASTER 6 54<br />

<strong>Snow</strong> <strong>an</strong>d Ice Cl<strong>as</strong>sification<br />

M<strong>an</strong>y image processing techniques could have been used in this study including m<strong>an</strong>ual<br />

digitization methods, b<strong>an</strong>d ratios <strong>an</strong>d normalized difference ratios. Because this study focuses on<br />

digital area <strong>as</strong>sessments, m<strong>an</strong>ual techniques were not considered. Fur<strong>the</strong>r <strong>an</strong>alysis of <strong>the</strong> b<strong>an</strong>d ratio<br />

method <strong>an</strong>d normalized difference method were completed before selecting <strong>the</strong> best technique to<br />

use for all satellite images.<br />

The Normalized Difference <strong>Snow</strong> Index (NDSI) <strong>an</strong>d B<strong>an</strong>d Math Ratio were both considered <strong>an</strong>d<br />

computed for all ASTER images for comparison. The NDSI (equation 1) is a commonly employed<br />

snow detection method that helps to differentiate between cloud cover <strong>an</strong>d snow/ice (Hall et al.<br />

1995). The B<strong>an</strong>d Math Ratio (equation 2) is a simple ratio between two b<strong>an</strong>ds, a <strong>an</strong>d b<br />

( NIR − Red)<br />

( NIR + Red)<br />

NDSI = (1)<br />

B<strong>an</strong>d a<br />

B<strong>an</strong>d Math Ratio = (2)<br />

B<strong>an</strong>d b<br />

Both methods were <strong>an</strong>alyzed for accuracy using several ASTER b<strong>an</strong>d combinations: one <strong>an</strong>d<br />

five, two <strong>an</strong>d five, <strong>an</strong>d three <strong>an</strong>d five. After visually <strong>as</strong>sessing both methods <strong>an</strong>d all b<strong>an</strong>d<br />

combinations, <strong>the</strong> NDSI method using ASTER b<strong>an</strong>d two (0.63–0.69 µm) <strong>an</strong>d five (2.145–2.185<br />

µm) w<strong>as</strong> selected <strong>as</strong> best for determining glacial area for <strong>the</strong>se study are<strong>as</strong>. The NDSI method w<strong>as</strong><br />

<strong>the</strong>n applied to <strong>the</strong> L<strong>an</strong>dsat scenes using <strong>the</strong> L<strong>an</strong>dsat b<strong>an</strong>d combination, three (0.63–0.69 µm) <strong>an</strong>d<br />

five (1.55–1.75 µm), that best approximated <strong>the</strong> ASTER b<strong>an</strong>ds.<br />

To determine <strong>the</strong> glacial area in each image using <strong>the</strong> NDSI ratio, a density slice w<strong>as</strong> performed<br />

to cl<strong>as</strong>sify pixels containing snow <strong>an</strong>d ice. For both ASTER <strong>an</strong>d L<strong>an</strong>dsat, pixels with NDSI values<br />

r<strong>an</strong>ging between 0.4 <strong>an</strong>d 1.0 were cl<strong>as</strong>sified <strong>as</strong> snow <strong>an</strong>d ice. To compute <strong>the</strong> total glaciated area<br />

for each area, a region of interest (ROI) w<strong>as</strong> drawn around glaciated area to determine <strong>the</strong> number<br />

of pixels with NDSI values in <strong>the</strong> selected r<strong>an</strong>ge. This ROI approach helped eliminate clouds <strong>an</strong>d<br />

o<strong>the</strong>r features miscl<strong>as</strong>sified <strong>as</strong> snow/ice. In some inst<strong>an</strong>ces, tr<strong>an</strong>sient snow cover still may be<br />

misinterpreted <strong>as</strong> glacier.


RESULTS<br />

Glacier ch<strong>an</strong>ges between <strong>the</strong> 1950s <strong>an</strong>d 2003 for <strong>the</strong> study are<strong>as</strong> in Columbia <strong>an</strong>d Venezuela<br />

were determined using historical information <strong>an</strong>d satellite images. In <strong>the</strong> 1950s, <strong>the</strong> total glacier<br />

area for three study regions in Colombia w<strong>as</strong> 89.33 km 2 . By 2003 <strong>the</strong> total glacier area had been<br />

reduced to 45.77 km 2 . From <strong>the</strong> 1950s to 2003 <strong>the</strong> calculated total ice loss in Colombia w<strong>as</strong> 43.56<br />

km 2 . The Sierra Nevada del Cocuy contributed 52% of <strong>the</strong> total ice lost, <strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif<br />

contributed 42% of this loss <strong>an</strong>d <strong>the</strong> Sierra Nevada de S<strong>an</strong>ta Marta contributed 6%. These ice loss<br />

percentages are directly related to <strong>the</strong> total glacier area of each region.<br />

Out of <strong>the</strong> 10 Venezuela glaciers mapped in 1952 with a total area of 2.91 km 2 (Schubert 1998),<br />

only one glacier is still visible from ASTER <strong>an</strong>d L<strong>an</strong>dsat images <strong>as</strong> of 1985. In 2004 <strong>the</strong> l<strong>as</strong>t<br />

remaining glacier on <strong>the</strong> Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif-Sinigüis, decre<strong>as</strong>ed from 2.03 km 2 in 1952 to 0.29<br />

km 2 or 86% of its 1952 area. Unfortunately, several images <strong>an</strong>alyzed of this glaciated region had<br />

to be eliminated due to se<strong>as</strong>onal snow cover in <strong>the</strong> area.<br />

Colombia<br />

Table 3. Glacier Are<strong>as</strong><br />

Glacier Series Date Area (km 2 )<br />

Ruiz-Tolima<br />

M<strong>as</strong>sif<br />

1959 Historic *<br />

10/24/1997<br />

01/28/2001<br />

33.95<br />

17.47<br />

15.81<br />

Sierra Nevada<br />

de S<strong>an</strong>ta Marta<br />

1957 Historic *<br />

1973 Historic *<br />

12/20/2000<br />

16.26<br />

14.10<br />

13.66<br />

1959 Historic * 39.12<br />

Sierra Nevada<br />

del Cocuy<br />

1973 Historic *<br />

05/25/1999<br />

03/06/2002<br />

28.00<br />

20.39<br />

16.30<br />

03/02/2003 16.30<br />

1952 Historic * 2.03<br />

Pico Bonpl<strong>an</strong>d<br />

M<strong>as</strong>sif<br />

Sinigüis<br />

03/24/1985<br />

01/20/1988<br />

01/29/2000<br />

02/11/2002<br />

0.66<br />

0.58<br />

0.38<br />

0.33<br />

03/02/2003 0.29<br />

* Me<strong>as</strong>urements taken from Hoyos-Patiño (1998) <strong>an</strong>d Schubert (1998)<br />

Venezuela<br />

Ruiz-Tolima M<strong>as</strong>sif<br />

In 1959, five peaks with glaciers existed in <strong>the</strong> Parque Nacional de los Nevados, but <strong>as</strong> of 1976<br />

only three glaciated peaks are still visible, Ruiz, Tolima, <strong>an</strong>d S<strong>an</strong>ta Isabel (Hoy<strong>as</strong>-Patiño, 1998). In<br />

1959, glacier area on <strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif w<strong>as</strong> 33.95 km 2 <strong>an</strong>d w<strong>as</strong> reduced to 15.81 km 2 by<br />

2001 (Figure 2). This represents a 53% loss in ice over a 42-year time sp<strong>an</strong>. Individually, glaciers<br />

on <strong>the</strong> Nevado del Ruiz decre<strong>as</strong>ed from 21.4 km 2 to 10.92 km 2 , a total loss of 49%. The Nevado<br />

del S<strong>an</strong>ta Isabel glaciers decre<strong>as</strong>ed from 9.78 km 2 to 3.61 km 2 , losing 63% of <strong>the</strong>ir area. The<br />

Nevado del Tolima glaciers declined from 2.22 km 2 in 1959 to 1.26 km 2 in 2001, a total ice loss of<br />

43% (Tables 3 <strong>an</strong>d 4).<br />

186


Table 4. Glacier Are<strong>as</strong> of <strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif<br />

Glacier Date Area (km 2 )<br />

Nevado de 1959 Historic * 9.78<br />

S<strong>an</strong>ta Isabel 10/24/1997 4.19<br />

01/28/2001 3.61<br />

Nevado del Ruiz 1959 Historic * 21.40<br />

10/24/1997 12.13<br />

01/28/2001 10.92<br />

Nevado del 1959 Historic * 2.22<br />

Tolima 10/24/1997 1.15<br />

01/28/2001 1.26<br />

*Me<strong>as</strong>urements taken from Hoyos-Patiño(1998)<br />

Figure 2. Time Series for Ruiz-Tolima M<strong>as</strong>sif Glaciers<br />

Sierra Nevada de S<strong>an</strong>ta Marta<br />

The Sierra Nevada de S<strong>an</strong>ta Marta w<strong>as</strong> <strong>the</strong> second Colombi<strong>an</strong> Glacier region studied (10º34′ N<br />

<strong>an</strong>d 73º43′ W). As determined from aerial photography, <strong>the</strong> region had 88 glaciers in 1957 with a<br />

total glacial area of 16.26 km 2 (Hoy<strong>as</strong>-Patiño 1998). B<strong>as</strong>ed on <strong>the</strong> 2000 ASTER image <strong>the</strong> total<br />

glaciated area w<strong>as</strong> 13.66 km 2 . Using <strong>the</strong> 2000 ASTER image, <strong>the</strong> most recent available snow free<br />

image, <strong>the</strong> total ice loss for <strong>the</strong> Sierra Nevada de S<strong>an</strong>ta Marta is 2.86 km 2 or 16% (Table 3 <strong>an</strong>d<br />

Figure 3). However, <strong>the</strong> 2000 estimate of glacier area still may be impacted by some tr<strong>an</strong>sient<br />

snow cover.<br />

187


Figure 3. Time Series for Sierra Nevada de S<strong>an</strong>ta Marta<br />

Sierra Nevada del Cocuy<br />

The Sierra Nevada del Cocuy is also a glaciated area in Colombia (6º27′ N <strong>an</strong>d 72º18′ W). The<br />

total glaciated area in 1959 w<strong>as</strong> 39.12 km 2 (Hoy<strong>as</strong>-Patiño 1998) <strong>an</strong>d had retreated to 16.3 km 2 in<br />

2003. In 44 years <strong>the</strong> area had a total ice loss of 58% (Table 3 <strong>an</strong>d Figure 4).<br />

Figure 4. Time Series for Sierra Nevada del Cocuy Glaciers<br />

188


Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif–Sinigüis Glacier<br />

The Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif glacier region is located in Venezuela at 8º32′ N <strong>an</strong>d 71º00′ W. In<br />

1952 two glaciers, Siniguis <strong>an</strong>d Nuestra Senora, existed on Pico Bonpl<strong>an</strong>d/Humbolt. In a L<strong>an</strong>dsat<br />

5 image acquired for <strong>the</strong> Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif, on March, 24, 1985, <strong>the</strong> only visible glacier w<strong>as</strong><br />

<strong>the</strong> Sinigüis glacier. The total glacier area <strong>as</strong>sessment in 1952 w<strong>as</strong> 2.03 km 2 <strong>an</strong>d by 2003 <strong>the</strong><br />

glacier area had decre<strong>as</strong>ed to 0.29 km 2 . This is a total area loss of 1.77 km 2 or 87% of <strong>the</strong> 1952<br />

area (Table 3 <strong>an</strong>d Figure 5).<br />

DISCUSSION<br />

Figure 5. Time Series for Pico Bonpl<strong>an</strong>d M<strong>as</strong>sif – Sinigüis Glacier<br />

Aerial photography <strong>an</strong>d L<strong>an</strong>dsat Multispectral Sc<strong>an</strong>ner (MSS) images provide historical<br />

me<strong>as</strong>urements of <strong>the</strong> Colombi<strong>an</strong> <strong>an</strong>d Venezuel<strong>an</strong> glaciers. Although <strong>the</strong>se me<strong>as</strong>urements help to<br />

exp<strong>an</strong>d <strong>the</strong> time sp<strong>an</strong> used in this study, it is unknown exactly how <strong>the</strong> are<strong>as</strong> were calculated. In<br />

this study, <strong>the</strong> Normalized Difference <strong>Snow</strong> Index (NDSI) provided a strong foundation for<br />

estimating <strong>the</strong> are<strong>as</strong> of <strong>the</strong> glacier regions by discriminating <strong>the</strong> snow/ice in <strong>the</strong> images through<br />

<strong>the</strong> use of a 0.4 threshold. By providing a uniform criterion to delineate snow <strong>an</strong>d ice, it is possible<br />

to construct a glacier retreat time series. Examining <strong>the</strong> trend for each glacier area, it is evident<br />

that each region h<strong>as</strong> experienced a decre<strong>as</strong>e in glacier area over <strong>the</strong> p<strong>as</strong>t fifty years.<br />

The glaciated region of <strong>the</strong> Sierra Nevada de S<strong>an</strong>ta Marta illustrates <strong>the</strong> commission errors<br />

caused by incre<strong>as</strong>es in se<strong>as</strong>onal snow cover <strong>as</strong> several images could not be used due to tr<strong>an</strong>sient<br />

snowfall obscuring glacier boundaries. Never<strong>the</strong>less, <strong>the</strong> overall trend on this glacier series<br />

confirms a general decre<strong>as</strong>e in glacier area from 1957 through 2004.<br />

In <strong>the</strong> c<strong>as</strong>e of <strong>the</strong> glacier area Ruiz-Tolima M<strong>as</strong>sif, volc<strong>an</strong>ic activity in <strong>the</strong> area is a partial cause<br />

for <strong>the</strong> decline in glacial area. As part of a series of stratovolc<strong>an</strong>oes, extending along <strong>the</strong> Andes<br />

from <strong>the</strong> sou<strong>the</strong>rn tip of Chile to <strong>the</strong> nor<strong>the</strong>rn portion of Venezuela, Nevado del Ruiz is currently<br />

in <strong>an</strong> active eruptive state. Recent eruptions in <strong>the</strong> mid-1980s affected temperatures in <strong>the</strong> glacier<br />

area enough to trigger snow <strong>an</strong>d ice melt, which is supported by <strong>the</strong> decline in areal extent of <strong>the</strong><br />

Ruiz-Tolima M<strong>as</strong>sif glaciers between <strong>the</strong> mid-1980’s <strong>an</strong>d 2001 (Linder <strong>an</strong>d Jord<strong>an</strong> 1991, Linder et<br />

al. 1994).<br />

On November 13, 1985, <strong>an</strong> eruption on Nevado del Ruiz caused South America’s deadliest<br />

eruption, a result of dev<strong>as</strong>tating lahars. The l<strong>as</strong>t known eruption of Nevado Del Ruiz occurred in<br />

189


1991. Unlike Ruiz, Nevado del Tolima <strong>an</strong>d Nevado de S<strong>an</strong>ta Isabel, <strong>the</strong> o<strong>the</strong>r stratovolc<strong>an</strong>oes in<br />

<strong>the</strong> Ruiz-Tolima M<strong>as</strong>sif, are not currently in <strong>an</strong> eruptive state at this time (Hoy<strong>as</strong>-Patiño 1998).<br />

Figure 4 illustrates <strong>the</strong> time series of area me<strong>as</strong>urements, <strong>an</strong>d shows a general decline in glacier<br />

size from 1959 to 2001.<br />

The series of images from <strong>the</strong> glaciers of Sierra Nevada del Cocuy make up <strong>the</strong> largest time<br />

series in this study, 9 images over 44 years. Originally consisting of numerous distinct glaciers,<br />

Sierra Nevada del Cocuy h<strong>as</strong> continually shown a signific<strong>an</strong>t loss in glacier size over <strong>the</strong> study<br />

period, resulting in steep general decre<strong>as</strong>e in glacial area. The observed retreat of <strong>the</strong> Sierra<br />

Nevada del Cocuy glaciers confirm notions of tropical glacier loss in this region.<br />

Finally, <strong>the</strong> glaciers of Venezuela are also disappearing. The sole remaining glacier, Sinigüis,<br />

showed a steady reduction in size between 1988 <strong>an</strong>d 2003 after a dramatic decre<strong>as</strong>e in size from<br />

1952 to 1985. The computed loss rate from 1952 to 1985 <strong>an</strong>d from 1988 to 2003 is 41515 m 2 /year<br />

<strong>an</strong>d 19333 m 2 /year, respectively. Overall, a general decre<strong>as</strong>e in glacier size is shown in Figure 5<br />

during <strong>the</strong> 51-year period.<br />

CONCLUSION<br />

Overall a decline in glacial area c<strong>an</strong> be seen throughout all four studied glacier regions of <strong>the</strong><br />

tropical Colombi<strong>an</strong> <strong>an</strong>d Venezuel<strong>an</strong> Andes. Steady to extreme decre<strong>as</strong>es in area are captured by<br />

<strong>the</strong> Normalized Difference <strong>Snow</strong> Index (NDSI) calculations completed in this study. From <strong>the</strong><br />

1950’s to present, signific<strong>an</strong>t snow <strong>an</strong>d ice loss over <strong>the</strong> area portray proof of a regional glacier<br />

recession. Broadening <strong>the</strong> scale, glaciers throughout <strong>the</strong> tropics are showing similar retreating<br />

characteristics (K<strong>as</strong>er 1999, Kincaid <strong>an</strong>d Klein 2004), <strong>an</strong>d globally, glaciers around <strong>the</strong> world,<br />

from different regions, climates, <strong>an</strong>d hum<strong>an</strong> interferences, confirm <strong>the</strong> results of this survey (IPCC<br />

2001).<br />

ACKNOWLEDGEMENTS<br />

This work w<strong>as</strong> funded by NASA Young Investigator Program Gr<strong>an</strong>t 02-000-0115. This project<br />

w<strong>as</strong> completed <strong>as</strong> <strong>an</strong> undergraduate directed studies course in <strong>the</strong> Department of Geography at<br />

Tex<strong>as</strong> A&M University.<br />

REFERENCES<br />

Abrams, M. 2000. The Adv<strong>an</strong>ced Spaceborne Thermal Emission <strong>an</strong>d Reflection Radiometer<br />

(ASTER): data products for <strong>the</strong> high spatial resolution imager on NASA’s Terra platform.<br />

International Journal of Remote Sensing 21: 847–849.<br />

Hall DK, Riggs GA, Salomonson VV. 1995. Development of methods for mapping global snow<br />

cover using moderate resolution imaging spectroradiometer data. Remote Sensing of<br />

Environment 54: 127–140.<br />

Hoyos-Patiño F. 1998. Glaciers of Columbia. In Williams RS Jr. <strong>an</strong>d Ferrigno JG (eds) Satellite<br />

Image Atl<strong>as</strong> of Glaciers of <strong>the</strong> World – South America. USGS Professional Paper 1386-I.<br />

United States Government Printing Office: W<strong>as</strong>hington, DC: 11–30.<br />

IPCC, 2001: Climate Ch<strong>an</strong>ge 2001: The Scientific B<strong>as</strong>is. Contribution of Working Group I to <strong>the</strong><br />

Third Assessment Report of <strong>the</strong> Intergovernmental P<strong>an</strong>el on Climate Ch<strong>an</strong>ge. Cambridge<br />

University Press: Cambridge; 881pp.<br />

K<strong>as</strong>er G. 1999. A review of <strong>the</strong> modern fluctuations of tropical glaciers. Global <strong>an</strong>d Pl<strong>an</strong>etary<br />

Ch<strong>an</strong>ge 22: 93–103.<br />

Kincaid JL, Klein AG. 2004. Retreat of <strong>the</strong> Iri<strong>an</strong> Jaya glaciers from 2000 to 2002 <strong>as</strong> me<strong>as</strong>ured<br />

from IKONIS Satellite Images. In Proceedings of <strong>the</strong> 61 st E<strong>as</strong>tern <strong>Snow</strong> Conference held June<br />

9–11, Portl<strong>an</strong>d Maine, USA, 147–157.<br />

L<strong>an</strong>dsat Project Science Office 2006. The L<strong>an</strong>dsat-7 Science Data User's H<strong>an</strong>dbook.<br />

http://l<strong>an</strong>dsath<strong>an</strong>dbook.gsfc.n<strong>as</strong>a.gov/h<strong>an</strong>dbook.html<br />

190


Linder W, Jord<strong>an</strong> E, 1991. Ice-m<strong>as</strong>s losses at <strong>the</strong> Nevado del Ruiz, Columbia, under <strong>the</strong> effect of<br />

<strong>the</strong> volc<strong>an</strong>ic eruption of 1985; a study b<strong>as</strong>ed on digital elevation models. S<strong>an</strong>ta Fé de Bogotá<br />

Revista Cartográfica 59: 479–484.<br />

Linder W, Jord<strong>an</strong> E, Christke K. 1994. Post-eruptive ice losses on <strong>the</strong> Nevado del Ruiz, Columbia.<br />

Zentralblatt für Geologie und Paläontologie 1: 479–484.<br />

Schubert C. 1998. Glaciers of Venezuela. In Williams RS Jr. <strong>an</strong>d Ferrigno JG (eds) Satellite Image<br />

Atl<strong>as</strong> of Glaciers of <strong>the</strong> World – South America. USGS Professional Paper 1386-I. United<br />

States Government Printing Office: W<strong>as</strong>hington, DC: 1–10.<br />

Schubert C, Clapperton C. 1990. Quaternary Glaciations of <strong>the</strong> Nor<strong>the</strong>rn Andes (Venezuela,<br />

Colombia <strong>an</strong>d Ecuador). Quaternary Science Reviews 9: 123–135.<br />

Thouret JC, V<strong>an</strong> der Hammen T, Salomons B, Juvigné E. 1996, Palaeoenvironmental ch<strong>an</strong>ges <strong>an</strong>d<br />

glacial stades of <strong>the</strong> l<strong>as</strong>t 50,000 years in <strong>the</strong> Cordillera Central, Colombia. Quaternary<br />

Research 46: 1–18.<br />

191


This page is intentionally bl<strong>an</strong>k.<br />

192


<strong>Snow</strong>pack Processes<br />

193


This page is intentionally bl<strong>an</strong>k.<br />

194


63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

<strong>Snow</strong> Cover Patterns <strong>an</strong>d Evolution at B<strong>as</strong>in Scale: GEOtop Model<br />

Simulations <strong>an</strong>d Remote Sensing Observations<br />

STEFANO ENDRIZZI, 1 GIACOMO BERTOLDI, 2 MARKUS NETELER 3 AND RICCARDO<br />

RIGON 1<br />

ABSTRACT: Remote sensing data c<strong>an</strong> provide images of snow covered are<strong>as</strong> <strong>an</strong>d, <strong>the</strong>refore,<br />

it is possible to follow <strong>the</strong> time evolution of snow melting spatial patterns with incre<strong>as</strong>ing<br />

spatial <strong>an</strong>d temporal resolution. <strong>Snow</strong> cover patterns are dominated by <strong>the</strong> complex<br />

interplay of topography, radiation forcings <strong>an</strong>d atmospheric turbulent tr<strong>an</strong>sfer processes.<br />

The snow cover evolution in <strong>an</strong> alpine b<strong>as</strong>in in Trentino (Italy) is here studied, comparing<br />

<strong>the</strong> simulations of <strong>the</strong> distributed hydrological model GEOtop with remotely sensed data.<br />

GEOtop describes <strong>the</strong> soil-snow-atmosphere energy <strong>an</strong>d m<strong>as</strong>s exch<strong>an</strong>ges, taking into account<br />

<strong>the</strong> snow physics <strong>an</strong>d <strong>the</strong> topographic effects of elevation, slope <strong>an</strong>d <strong>as</strong>pect on solar radiation<br />

<strong>an</strong>d air temperature. <strong>the</strong> <strong>Snow</strong> cover extent h<strong>as</strong> been provided by MODIS 8-day composite<br />

maps with a resolution of 500 metres. The model reproduces <strong>the</strong> physical features of snow<br />

melting re<strong>as</strong>onably <strong>an</strong>d shows a fair agreement with <strong>the</strong> data. The relative import<strong>an</strong>ce of<br />

precipitation, solar radiation, <strong>an</strong>d temperature to control <strong>the</strong> snow accumulation <strong>an</strong>d<br />

melting processes is also investigated.<br />

Keywords: snow water equivalent, distributed modelling, remote sensing, topography<br />

INTRODUCTION<br />

To have information about <strong>the</strong> evolution of <strong>the</strong> snow cover extent <strong>an</strong>d <strong>the</strong> snow water equivalent<br />

is very useful for <strong>the</strong> water resource m<strong>an</strong>agement, <strong>an</strong>d in relation to climate ch<strong>an</strong>ge. So far<br />

temperature index models, like SRM (Martinec <strong>an</strong>d R<strong>an</strong>go, 1986), b<strong>as</strong>ed on a statistical relation<br />

between snow melting <strong>an</strong>d temperature, have been widely used. Although m<strong>an</strong>y of <strong>the</strong>se models<br />

have evolved so that topographic effects could be considered (Cazorzi <strong>an</strong>d Dalla Font<strong>an</strong>a, 1995),<br />

distributed physically b<strong>as</strong>ed models, like ISNOBAL (Marks et al., 1999) c<strong>an</strong> provide more<br />

detailed information about snow physics <strong>an</strong>d improved predictions of snow cover. Their<br />

application also allows to study what phenomena may be import<strong>an</strong>t in snow accumulation <strong>an</strong>d<br />

melting, <strong>an</strong>d what may be <strong>the</strong> sensitivity to ch<strong>an</strong>ges of atmospheric forcing.<br />

At <strong>the</strong> same time, remote sensing data are available with improved spatial <strong>an</strong>d temporal<br />

resolution (for example MODIS (Hall et al., 2002) <strong>an</strong>d NOHRSC (Hall et al., 2000)). Several<br />

models, using remote sensing data <strong>as</strong> input (Turpin et al., 1999) or to perform data <strong>as</strong>similation<br />

(Carroll et al., 2006), have been recently developed. However, in mountain b<strong>as</strong>ins, remote sensing<br />

data c<strong>an</strong> be affected by some errors, <strong>an</strong>d consistent physical modelling maintains great<br />

1 Department of Civil <strong>an</strong>d Environmental Engineering, University of Trento, Via Mesi<strong>an</strong>o 77,<br />

38050 Povo (TN), Italy, email: stef<strong>an</strong>o.endrizzi@ing.unitn.it<br />

2 Department of Civil <strong>an</strong>d Environmental Engineering, Duke University, Durham, NC, USA<br />

3 ITC-IRST, Center for Scientific <strong>an</strong>d Technological Research, 38050 Povo (TN), Italy<br />

195


import<strong>an</strong>ce, because snow cover patterns are dominated by <strong>the</strong> complex interplay of topography,<br />

radiation forcing <strong>an</strong>d atmospheric turbulent tr<strong>an</strong>sfer processes (Pomeroy et al., 2003).<br />

The GEOtop model (Rigon et al., 2006) is a distributed hydrological model that solves <strong>the</strong><br />

energy <strong>an</strong>d water bal<strong>an</strong>ce on a l<strong>an</strong>dscape whose topographical surface is described by a digital<br />

elevation model. The model h<strong>as</strong> been conceived to be applied to mountain b<strong>as</strong>ins characterized by<br />

complex topography, where snow accumulation <strong>an</strong>d melting have to be accounted. GEOtop<br />

includes a snow module, <strong>the</strong> first version (version 0.875) of which w<strong>as</strong> <strong>the</strong> object of a previous<br />

work (Z<strong>an</strong>otti et al., 2004), where its capability to predict <strong>the</strong> snow water equivalent (SWE)<br />

evolution in a point w<strong>as</strong> tested, <strong>an</strong>d <strong>the</strong> results were compared with local me<strong>as</strong>urements. In Z<strong>an</strong>otti<br />

et al. (2004) <strong>an</strong> application on a small mountain b<strong>as</strong>in with a surface area of a few square<br />

kilometres w<strong>as</strong> <strong>the</strong>n presented, but <strong>the</strong> results could not be checked in more th<strong>an</strong> one me<strong>as</strong>urement<br />

point. Since <strong>the</strong>n, <strong>the</strong> snow cover module in GEOtop h<strong>as</strong> been improved (version 0.9375) such<br />

that a multilayer representation h<strong>as</strong> been implemented, similar to <strong>the</strong> one of <strong>the</strong> CLM l<strong>an</strong>d surface<br />

model (Oleson et al., 2004). The GEOtop snow module is quite similar to ISNOBAL (Marks et al.,<br />

1999), <strong>as</strong> it solves <strong>the</strong> snow energy bal<strong>an</strong>ce, but it is part of a complete hydrological model that<br />

considers <strong>the</strong> whole soil-snow system. The next step is to predict <strong>the</strong> snow cover evolution <strong>an</strong>d its<br />

variability at <strong>the</strong> distributed scale.<br />

Some works have been published about <strong>the</strong> temporal distribution of snow cover, for example<br />

Alfnes et al. (2004) using statistical tools. In order to do this, a distributed field of <strong>the</strong><br />

meteorological forcing <strong>as</strong> input data is needed, or, if only a few me<strong>as</strong>urement stations are available<br />

throughout <strong>the</strong> b<strong>as</strong>in, we need to find some criteria to spatially extrapolate <strong>the</strong> me<strong>as</strong>urements,<br />

although <strong>the</strong>y might lead to some errors.<br />

An application in <strong>an</strong> Alpine b<strong>as</strong>in with surface area of about 250 square kilometres is shown<br />

here. The snow cover extension area is compared with corresponding maps provided by remote<br />

sensing techniques. In particular, MODIS maps (Hall et al., 2002; Riggs et al., 2003) have been<br />

used because <strong>the</strong> Aqua <strong>an</strong>d Terra satellites overp<strong>as</strong>s locations daily. <strong>Snow</strong> products are delivered<br />

at a spatial resolution of 500 metres, which is sufficient to keep <strong>the</strong> signature of <strong>the</strong> topographic<br />

features in <strong>an</strong> Alpine environment (Cline et al., 1998) where <strong>the</strong> snow cover spatial distribution is<br />

strongly dependent on <strong>as</strong>pect <strong>an</strong>d elevation. In addition, <strong>the</strong> MODIS data are e<strong>as</strong>ily available for<br />

applicative uses.<br />

<strong>Snow</strong> cover maps provided by remote sensing have already been used, mainly <strong>as</strong> <strong>an</strong> auxiliary to<br />

<strong>the</strong> models (Lee et al., 2005; Turpin et al., 1999), but here <strong>the</strong>y are used only to check <strong>the</strong> model<br />

results. The problem of <strong>the</strong> initial SWE distribution does not exist <strong>as</strong> <strong>the</strong> simulation starts from a<br />

late summer initial condition, when in <strong>the</strong> considered b<strong>as</strong>in no snow is present.<br />

Our aim in this paper is to test <strong>the</strong> capability of <strong>the</strong> model to reproduce a realistic snow cover<br />

evolution during a whole winter se<strong>as</strong>on <strong>an</strong>d to investigate <strong>the</strong> relative import<strong>an</strong>ce of precipitation,<br />

solar radiation, <strong>an</strong>d temperature to control <strong>the</strong> snow accumulation <strong>an</strong>d melting processes in <strong>an</strong><br />

Alpine environment, with reference to its time evolution <strong>an</strong>d its dependence on elevation <strong>an</strong>d<br />

<strong>as</strong>pect. We w<strong>an</strong>t to develop a modelling framework <strong>as</strong> physically b<strong>as</strong>ed <strong>as</strong> possible, though<br />

parsimonious in its input data requirements <strong>an</strong>d, <strong>the</strong>refore, e<strong>as</strong>ily applicable for operational use<br />

(Carroll et al, 2006).<br />

THE MODEL<br />

The GEOtop model h<strong>as</strong> been fully described in Rigon et al. (2006) <strong>an</strong>d in Bertoldi et al. (2006).<br />

This model h<strong>as</strong> been conceived to be <strong>an</strong> integration of a rainfall-runoff model <strong>an</strong>d a l<strong>an</strong>d surface<br />

model <strong>as</strong> it solves <strong>the</strong> three-dimensional soil water budget equation toge<strong>the</strong>r with <strong>the</strong> onedimensional<br />

energy budget equation, so that it c<strong>an</strong> calculate <strong>the</strong> spatial distribution of m<strong>an</strong>y hydrometeorological<br />

variables (such <strong>as</strong> soil moisture, surface temperature, convective <strong>an</strong>d radiative<br />

fluxes) without losing <strong>the</strong> main purpose of hydrological models, that is predicting <strong>the</strong> water<br />

discharge at a specific closure section. However, this paper is focused on <strong>the</strong> simulation of <strong>the</strong><br />

spatial patterns <strong>an</strong>d of <strong>the</strong> time evolution of <strong>the</strong> snow covered area <strong>an</strong>d of <strong>the</strong> SWE at b<strong>as</strong>in scale.<br />

196


In <strong>the</strong> following paragraph <strong>the</strong> main improvements of <strong>the</strong> snow module with reference to <strong>the</strong><br />

version presented in Rigon et al. (2006) are briefly discussed.<br />

In <strong>the</strong> version 0.9375 of <strong>the</strong> snow module in GEOtop, <strong>the</strong> snow cover is described with a<br />

multilayer scheme. For each snow layer <strong>the</strong> liquid <strong>an</strong>d solid water budget equations <strong>an</strong>d <strong>the</strong> energy<br />

budget equation are solved toge<strong>the</strong>r with <strong>the</strong> continuity equation <strong>an</strong>d a formula linking <strong>the</strong> ph<strong>as</strong>e<br />

ch<strong>an</strong>ge with <strong>the</strong> temperature. So five equations for <strong>the</strong> following five unknowns are available:<br />

liquid <strong>an</strong>d ice contents, porosity, snow temperature, <strong>an</strong>d amount of water ch<strong>an</strong>ging ph<strong>as</strong>e.<br />

The water budget equations, for liquid water <strong>an</strong>d ice, are expressed <strong>as</strong> follows:<br />

!<br />

!<br />

" w = 1<br />

# w<br />

" i = 1<br />

# i<br />

$W<br />

$t + $Q % ( w<br />

' *<br />

& $z )<br />

$W<br />

$t + $Q % ( i<br />

' *<br />

& $z )<br />

where θw <strong>an</strong>d θi are <strong>the</strong> nondimensional liquid water <strong>an</strong>d ice content, t is time, z <strong>the</strong> vertical<br />

coordinate (positive upwards), ρw <strong>an</strong>d ρI <strong>the</strong> density of liquid water <strong>an</strong>d ice, respectively, Qw <strong>an</strong>d<br />

Qi <strong>the</strong> vertical flux of liquid water <strong>an</strong>d ice (<strong>the</strong> latter is equal to 0 except at <strong>the</strong> snow-atmosphere<br />

interface where it becomes <strong>the</strong> snow precipitation), <strong>an</strong>d W is <strong>the</strong> amount of water ch<strong>an</strong>ging ph<strong>as</strong>e<br />

(positive in <strong>the</strong> c<strong>as</strong>e of melting, negative in <strong>the</strong> c<strong>as</strong>e of freezing). As <strong>the</strong> liquid water flow inside<br />

<strong>the</strong> snow cover is mainly due to gravity, <strong>the</strong> equations are solved only in <strong>the</strong> vertical direction,<br />

neglecting <strong>the</strong> lateral flow inside <strong>the</strong> snowpack, but not in <strong>the</strong> underlying soil. Qi depends on <strong>the</strong><br />

liquid water fraction in <strong>the</strong> snow exceeding <strong>the</strong> capillary retention <strong>as</strong> in Colbeck <strong>an</strong>d Anderson<br />

(1982).<br />

The energy budget equation is <strong>the</strong> following:<br />

!<br />

C "T<br />

"t + L f<br />

"W<br />

"T<br />

( )<br />

" # "T &<br />

= % k ( +<br />

"z $ "z '<br />

" QwU w<br />

"z<br />

where T is <strong>the</strong> temperature, Lf <strong>the</strong> latent heat of fusion, Uw <strong>the</strong> specific energy content of <strong>the</strong> liquid<br />

water, <strong>an</strong>d k <strong>an</strong>d C <strong>the</strong> <strong>the</strong>rmal conductivity <strong>an</strong>d <strong>the</strong> <strong>the</strong>rmal capacity, respectively, calculated <strong>as</strong><br />

averages of <strong>the</strong> values of <strong>the</strong> different ph<strong>as</strong>es present. The equation (3) is solved with <strong>the</strong><br />

following boundary condition at <strong>the</strong> snow-atmosphere interface, where <strong>the</strong> exch<strong>an</strong>ge fluxes occur:<br />

!<br />

#<br />

%<br />

$<br />

k "T<br />

"z<br />

(1)<br />

(2)<br />

(3)<br />

&<br />

( = )Rn + H + L (4)<br />

'<br />

where Rn is <strong>the</strong> net radiation, H <strong>the</strong> sensible heat flux <strong>an</strong>d L <strong>the</strong> latent heat flux (both positive<br />

upwards). The treatment of radiation is essentially <strong>the</strong> same <strong>as</strong> in Z<strong>an</strong>otti et al. (2004). The effects<br />

of shadowing, slope <strong>an</strong>d <strong>as</strong>pect, which are extremely import<strong>an</strong>t in a complex topography, are<br />

taken into account for direct shortwave radiation, <strong>an</strong>d <strong>the</strong> sky view factor is considered for <strong>the</strong><br />

diffuse shortwave radiation <strong>an</strong>d for <strong>the</strong> longwave radiation. Shortwave radiation is parameterized<br />

<strong>as</strong> in Iqbal (1983). Longwave radiation is calculated using Stef<strong>an</strong>-Boltzm<strong>an</strong>n formula. Where<strong>as</strong><br />

<strong>the</strong> outcoming longwave radiation is clearly dependent on <strong>the</strong> surface temperature, <strong>the</strong>re are some<br />

uncertainties in calculating <strong>the</strong> incoming longwave radiation, <strong>as</strong> <strong>the</strong> latter depends on <strong>the</strong> <strong>entire</strong><br />

profile of <strong>the</strong> air temperature above <strong>the</strong> snowpack. Different experimental formulae, <strong>as</strong> those of<br />

Brutsaert (1975), Satterlund (1979), <strong>an</strong>d Idso (1981), that parameterize this component of<br />

radiation in function of <strong>the</strong> skin temperature, have been used. The convective heat fluxes are<br />

calculated after Monin-Obukhov similarity <strong>the</strong>ory, using <strong>the</strong> stability function of Businger et al.<br />

(1981). As <strong>the</strong>y depend also on surface temperature, which is unknown <strong>as</strong> long <strong>as</strong> <strong>the</strong> energy<br />

budget is not solved, some iterations are needed.<br />

The continuity equation correlates <strong>the</strong> different ph<strong>as</strong>es in <strong>the</strong> snowpack:<br />

!<br />

" w + " i + " v =1 (5)<br />

197


!<br />

where θv is <strong>the</strong> nondimensional g<strong>as</strong> <strong>an</strong>d vapour content, that is <strong>the</strong> porosity.<br />

Finally, we need to find a relation between <strong>the</strong> snow temperature <strong>an</strong>d <strong>the</strong> ph<strong>as</strong>e ch<strong>an</strong>ge. We<br />

have chosen that <strong>the</strong> ph<strong>as</strong>e ch<strong>an</strong>ge should occur when <strong>the</strong> temperature in each snow layer is equal<br />

to 273.15 K: nei<strong>the</strong>r <strong>the</strong> temperature is allowed to incre<strong>as</strong>e <strong>as</strong> long <strong>as</strong> ice is present nor c<strong>an</strong> it<br />

decre<strong>as</strong>e until liquid water is <strong>entire</strong>ly frozen. Therefore,<br />

W " 0 if T = 273.15K ; W = 0 if T " 273.15K (6)<br />

Thus <strong>the</strong> system of equations (1)-(2)-(3)-(5)-(6) is solved with a finite difference method with<br />

respect to <strong>the</strong> unknowns T, W, θw, θi, θv.<br />

The snow layers have a varying thickness in accord<strong>an</strong>ce with <strong>the</strong> snow precipitation <strong>an</strong>d snow<br />

metamorphism (Anderson, 1976). The algorithms to combine <strong>an</strong>d subdivide snow layers have<br />

been taken from Oleson et al. (2004), which makes it possible to describe <strong>the</strong> density ch<strong>an</strong>ges of<br />

<strong>the</strong> snow pack.<br />

APPLICATION<br />

The model h<strong>as</strong> been applied to <strong>the</strong> upper part of Brenta B<strong>as</strong>in (Trentino, Italy), in <strong>the</strong> E<strong>as</strong>tern<br />

Itali<strong>an</strong> Alps, a b<strong>as</strong>in with a surface area of about 250 square kilometres. Its elevation r<strong>an</strong>ges from<br />

340 m a.s.l. to 2303 m a.s.l., <strong>an</strong>d <strong>the</strong> l<strong>an</strong>d cover is very variable. The bottom of <strong>the</strong> valley is<br />

urb<strong>an</strong>ized, <strong>an</strong>d m<strong>an</strong>y small villages are present, although <strong>the</strong>re is still a wide countryside area.<br />

Two lakes, with total area of about 6 square kilometres, are also present. They play <strong>an</strong> import<strong>an</strong>t<br />

role in <strong>the</strong> regulation of <strong>the</strong> discharge, but, <strong>as</strong> here our main interest is to follow <strong>the</strong> snow cover<br />

evolution, <strong>the</strong>y are neglected. As elevation incre<strong>as</strong>es, woodl<strong>an</strong>d prevails, even if <strong>the</strong>re are some<br />

small resort villages <strong>an</strong>d wide gr<strong>as</strong>sl<strong>an</strong>d <strong>an</strong>d p<strong>as</strong>ture are<strong>as</strong>. The vegetation limit r<strong>an</strong>ges from 1900<br />

m a.s.l. to 2000 m a.s.l.. The snow cover h<strong>as</strong> been simulated for <strong>the</strong> se<strong>as</strong>on 2003-2004, namely<br />

from 16 th October 2003 to 16 th June 2004, using a model grid resolution of 250 m.<br />

METEOROLOGICAL DATA<br />

For <strong>the</strong> time considered, hourly precipitation me<strong>as</strong>urement data were available for 15 stations<br />

located in <strong>the</strong> b<strong>as</strong>in or in <strong>the</strong> nearby are<strong>as</strong>. Precipitation w<strong>as</strong> <strong>the</strong>n distributed according to kriging<br />

techniques, <strong>an</strong>d it w<strong>as</strong> split into liquid <strong>an</strong>d solid form in conformity with <strong>the</strong> air temperature,<br />

defining a threshold air temperature TL above which <strong>the</strong> precipitation is always liquid <strong>an</strong>d <strong>an</strong>o<strong>the</strong>r<br />

threshold temperature TS below which <strong>the</strong> precipitation is always solid, <strong>as</strong> in U.S. Army Corps of<br />

Engineers (1956). These threshold temperatures are considered tuneable <strong>an</strong>d adjustable<br />

parameters. In fact, <strong>as</strong> discussed later, in order to catch <strong>the</strong> snow events with precision, it results<br />

very import<strong>an</strong>t to distribute <strong>the</strong> air temperature in <strong>the</strong> b<strong>as</strong>in accurately.<br />

Hourly air temperature observations performed in 17 meteorological stations in <strong>an</strong>d around <strong>the</strong><br />

b<strong>as</strong>in were used to find a linear dependence of <strong>the</strong> temperature on <strong>the</strong> elevation, which is<br />

considered variable in time, but valid all over <strong>the</strong> b<strong>as</strong>in, <strong>an</strong>d <strong>an</strong> hourly value of <strong>the</strong> lapse rate w<strong>as</strong><br />

found.<br />

Figure 1 shows <strong>the</strong> relation between <strong>the</strong> air temperature me<strong>as</strong>urements averaged on <strong>the</strong> whole<br />

se<strong>as</strong>on <strong>an</strong>d <strong>the</strong> station elevations. The temperature variation presents a clear linear trend with a<br />

lapse rate of 3.9 °C/km (regression const<strong>an</strong>t equal to 0.94), which is lower th<strong>an</strong> <strong>the</strong> lapse rate<br />

normally used (6.5 °C/km). This me<strong>an</strong>s that at <strong>the</strong> <strong>an</strong>nual scale <strong>the</strong> temperature me<strong>as</strong>urements do<br />

not exhibit evident <strong>an</strong>d regular variability with <strong>the</strong> horizontal coordinates. However, some stations<br />

located approximately at <strong>the</strong> same elevation c<strong>an</strong> record temperature differences of some degrees<br />

(at <strong>the</strong> most 2 °C) due to local factors (for example shadowing) <strong>an</strong>d to particular meteorological<br />

events. This could be a me<strong>as</strong>ure of <strong>the</strong> error in <strong>the</strong> temperature estimation.<br />

198


Figure 1: Variation of <strong>the</strong> se<strong>as</strong>on average of <strong>the</strong> air temperature me<strong>as</strong>urements with <strong>the</strong> elevations of <strong>the</strong><br />

meteorological stations (dots), <strong>an</strong>d regression line.<br />

Figure 2: Time evolution of <strong>the</strong> hourly <strong>an</strong>d daily average lapse rate inferred by <strong>the</strong> me<strong>as</strong>urements.<br />

But if <strong>the</strong> time evolution of <strong>the</strong> lapse rate, daily <strong>an</strong>d hourly averaged (Figure 2) is plotted, it c<strong>an</strong><br />

be noticed that during <strong>the</strong> winter <strong>the</strong>re are frequent periods of <strong>the</strong>rmal inversion, with a nearly<br />

iso<strong>the</strong>rm atmosphere, but during <strong>the</strong> spring time <strong>the</strong> lapse rate is more clearly defined <strong>an</strong>d reaches<br />

values around <strong>the</strong> adiabatic lapse rate. The lapse rate also shows a typical daily cycle, with a<br />

marked nocturnal inversion in winter. The definition of a correct value of <strong>the</strong> lapse rate is critical<br />

to capture <strong>the</strong> correct snow-rain limit. Therefore, we have chosen to use <strong>an</strong> hourly varying lapse<br />

rate <strong>as</strong> input data for <strong>the</strong> model.<br />

199


The relative humidity <strong>an</strong>d <strong>the</strong> wind velocity are kept const<strong>an</strong>t all over <strong>the</strong> b<strong>as</strong>in.<br />

PARAMETERS OF THE MODEL<br />

The parameters of <strong>the</strong> model were guessed in order to obtain <strong>the</strong> best agreement with <strong>the</strong> remote<br />

sensing data. By all me<strong>an</strong>s, only a few parameters are import<strong>an</strong>t for <strong>the</strong> snow cover description.<br />

They are <strong>the</strong> temperature roughness length, <strong>the</strong> wind velocity roughness length, <strong>the</strong> capillary water<br />

amount retained by <strong>the</strong> snow, <strong>an</strong>d <strong>the</strong> threshold temperatures aforesaid. The following table<br />

reports <strong>the</strong> value of <strong>the</strong>se parameters with <strong>the</strong> reference.<br />

Table 1. Main parameters of <strong>the</strong> snow module in GEOtop<br />

Value Parameter Description Reference<br />

0.05 z0T Temperature roughness length [m] Calibration<br />

0.5 z0V Wind velocity roughness length [m] Calibration<br />

0.04 Sr Liquid water in snow retained by capillarity<br />

forces, <strong>as</strong> fraction of porosity [-]<br />

Jord<strong>an</strong> (1991)<br />

2.0 TL Temperature above which all <strong>the</strong> precipitation is<br />

rain [°C]<br />

Calibration<br />

0.5 TS Temperature below which all <strong>the</strong> precipitation is<br />

snow [°C]<br />

Calibration<br />

MODIS<br />

The Moderate Resolution Imaging Spectroradiometer (MODIS) is a 36-ch<strong>an</strong>nel sensor<br />

me<strong>as</strong>uring <strong>the</strong> spectral r<strong>an</strong>ge from visible to <strong>the</strong>rmal-infrared. It w<strong>as</strong> launched <strong>as</strong> part of <strong>the</strong><br />

payload of <strong>the</strong> Terra (12/1999) <strong>an</strong>d Aqua (5/2002) satellites. Among o<strong>the</strong>r products, daily <strong>an</strong>d 8day<br />

composite snow related maps are produced at 500 m <strong>an</strong>d 1000 m resolution. The maps created<br />

over l<strong>an</strong>d include daily snow albedo <strong>an</strong>d 8-day maximum snow extent. For <strong>the</strong> latter multiple days<br />

of observations for a cell are examined. If snow cover is found for at le<strong>as</strong>t one day <strong>the</strong> cell is<br />

indicated <strong>as</strong> snow. O<strong>the</strong>r values are lake ice, cloud, oce<strong>an</strong>, inl<strong>an</strong>d water <strong>an</strong>d l<strong>an</strong>d. If no snow is<br />

found in <strong>the</strong> cell, <strong>the</strong> most frequent value of <strong>the</strong> o<strong>the</strong>r cl<strong>as</strong>ses is <strong>as</strong>signed. Due to <strong>the</strong> 8-day<br />

compositing technique, <strong>the</strong> impact of clouds is minimized (Riggs et al., 2003).<br />

In this study we have used <strong>the</strong> MOD10A2 8-day composite maximum snow extent data at level<br />

V004. The data were processed in GRASS GIS (Neteler, 2005).<br />

However, MODIS snow cover maps c<strong>an</strong> be affected by some errors, <strong>the</strong> most common of which<br />

are <strong>the</strong> errors of commission. This me<strong>an</strong>s mapping cells <strong>as</strong> snow covered when <strong>the</strong> occurrence of<br />

snow is extremely unlikely. These errors are usually <strong>as</strong>sociated with <strong>the</strong> le<strong>as</strong>t favorable conditions<br />

of illumination <strong>an</strong>d cloud type for optimal snow mapping (Riggs <strong>an</strong>d Hall, 2006).<br />

COMPARISON BETWEEN THE RESULTS OF THE MODEL AND MODIS DATA<br />

<strong>Snow</strong> covered area spatial distribution<br />

In this paragraph <strong>the</strong> MODIS 8-day maximum snow extent maps <strong>an</strong>d <strong>the</strong> corresponding 8-day<br />

GEOtop model output maps are compared. The snow free area is reported in green, <strong>an</strong>d <strong>the</strong> snow<br />

covered area is in white. Where<strong>as</strong> MODIS c<strong>an</strong> only provide information about <strong>the</strong> presence of<br />

snow, <strong>the</strong> model gives results about <strong>the</strong> amount of SWE, which is reported in <strong>the</strong> pictures in<br />

white-blue colour scale. Five pairs of maps have been chosen <strong>as</strong> representative of <strong>the</strong> main<br />

qualities <strong>an</strong>d problems of <strong>the</strong> models.<br />

200


Figure 3: 8-day composite maps (24 th October 2003) for snow cover obtained by MODIS (left) <strong>an</strong>d by <strong>the</strong><br />

model (right). The snow free area is reported in green, <strong>the</strong> snow covered area is in white in MODIS map, in<br />

white-blue colour scale (legend on <strong>the</strong> right) in <strong>the</strong> model map.<br />

Figure 4: 8-day composite maps (17 th November 2003) for snow cover obtained by MODIS (left) <strong>an</strong>d by <strong>the</strong><br />

model (right). The snow free area is reported in green, <strong>the</strong> snow covered area is in white in MODIS map, in<br />

white-blue colour scale (legend on <strong>the</strong> right) in <strong>the</strong> model map.<br />

The first pair of pictures (Figure 3) regards <strong>the</strong> autumn se<strong>as</strong>on: snow cover is present only at <strong>the</strong><br />

highest elevations, <strong>an</strong>d <strong>the</strong> model predicts re<strong>as</strong>onably <strong>the</strong> snow limit. However, in <strong>the</strong> model<br />

output map <strong>the</strong> snow limit strictly follows <strong>the</strong> elevation contour lines, because snow precipitation<br />

mainly depends on <strong>the</strong> air temperature, which, in turn, depends only on elevation. In <strong>the</strong> MODIS<br />

map <strong>the</strong> snow limit is close to <strong>the</strong> elevation contour, but <strong>the</strong>re are some deviations probably due to<br />

local variations of air temperature, or, simply, to <strong>the</strong> coarser resolution of <strong>the</strong> MODIS maps (500<br />

m grid cell) th<strong>an</strong> in <strong>the</strong> model maps (250 m grid cell).<br />

201


Figure 5: 8-day composite maps (17 th J<strong>an</strong>uary 2004) for snow cover obtained by MODIS (left) <strong>an</strong>d by <strong>the</strong><br />

model (right). The snow free area is reported in green, <strong>the</strong> snow covered area is in white in MODIS map, in<br />

white-blue colour scale (legend on <strong>the</strong> right) in <strong>the</strong> model map. The pixels for which it w<strong>as</strong> not possible to<br />

detect snow due to cloud cover are in yellow.<br />

Figure 6: 8-day composite maps (29 h March 2004) for snow cover obtained by MODIS (left) <strong>an</strong>d by <strong>the</strong><br />

model (right). The snow free area is reported in green, <strong>the</strong> snow covered area is in white in MODIS map, in<br />

white-blue colour scale (legend on <strong>the</strong> right) in <strong>the</strong> model map.<br />

The second pair (Figure 4) is <strong>an</strong> example of <strong>the</strong> late autumn se<strong>as</strong>on, after several days without<br />

solid precipitation. The snow patterns are controlled here also by slope <strong>an</strong>d <strong>as</strong>pect, <strong>an</strong>d <strong>the</strong>re are<br />

some white spots inside <strong>the</strong> snow free area in <strong>the</strong> pixels receiving a less amount of radiation. The<br />

agreement is fairly good, although not all <strong>the</strong> white spots agree in detail, but it h<strong>as</strong> to be taken into<br />

account that MODIS maps may be affected by some commission errors (for example <strong>the</strong> white<br />

spot located near <strong>the</strong> lake in <strong>the</strong> bottom of <strong>the</strong> valley is very unlikely).<br />

202


The third pair (Figure 5) represents <strong>the</strong> snow cover after a snowfall in winter. It is difficult for<br />

<strong>the</strong> model to reproduce <strong>the</strong> snow in <strong>the</strong> bottom of <strong>the</strong> valley, <strong>as</strong> <strong>the</strong> air temperature field c<strong>an</strong> be<br />

uneven <strong>an</strong>d snow precipitation occurs usually at low elevations when <strong>the</strong> air temperature is close<br />

or even slightly above 0 °C. For this re<strong>as</strong>on <strong>the</strong> threshold temperature above which <strong>the</strong><br />

precipitation is always liquid <strong>an</strong>d <strong>the</strong> threshold temperature below which <strong>the</strong> precipitation is<br />

always solid have been fixed to 2 °C <strong>an</strong>d 0.5 °C, respectively, after some calibrations. These values<br />

are a little different from <strong>the</strong> values proposed in US Army Corps of Engineers (1956), respectively<br />

3 °C <strong>an</strong>d -1 °C. The are<strong>as</strong> located at <strong>the</strong> bottom valley are quite large, but <strong>the</strong>y are usually covered<br />

by a small amount of snow, which melts quickly.<br />

The fourth pair (Figure 6) is relative to early springtime. The snow patterns are dominated by<br />

elevation, <strong>as</strong> a consequence of late snowfall events, <strong>as</strong> well <strong>as</strong> by slope <strong>an</strong>d <strong>as</strong>pect, <strong>as</strong> it is possible<br />

to appreciate <strong>the</strong> differential melting in north facing <strong>an</strong>d south spacing slopes, even though in <strong>the</strong><br />

two maps <strong>the</strong> snow limit c<strong>an</strong> be slightly different. In this time SWE reaches its maximum value at<br />

<strong>the</strong> highest elevations.<br />

Figure 7: 8-day composite maps (16 h May 2004) for snow cover obtained by MODIS (left) <strong>an</strong>d by <strong>the</strong> model<br />

(right). The snow free area is reported in green, <strong>the</strong> snow covered area is in white in MODIS map, in whiteblue<br />

colour scale (legend on <strong>the</strong> right) in <strong>the</strong> model map.<br />

Finally, <strong>the</strong> fifth pair (Figure 7) refers to late springtime: snow cover is present only at <strong>the</strong><br />

highest elevations, <strong>an</strong>d <strong>the</strong> snow patterns in <strong>the</strong> two maps are similar. However, <strong>the</strong> model result<br />

map shows a little f<strong>as</strong>ter melting on <strong>the</strong> south facing slopes th<strong>an</strong> <strong>the</strong> MODIS map, which, on <strong>the</strong><br />

o<strong>the</strong>r h<strong>an</strong>d, may also be affected by commission errors, because it shows a snow limit lower in <strong>the</strong><br />

south facing slopes th<strong>an</strong> in <strong>the</strong> north facing slopes.<br />

In conclusion, <strong>the</strong> model seems to catch quite well <strong>the</strong> main patterns of <strong>the</strong> snow cover extent,<br />

which are dominated firstly by elevation, <strong>an</strong>d secondly by slope <strong>an</strong>d <strong>as</strong>pect. In order to reproduce<br />

<strong>the</strong> snow cover in <strong>the</strong> best way, it seems crucial to catch <strong>the</strong> snowfall elevation limits, <strong>an</strong>d to<br />

calculate accurately <strong>the</strong> radiation fluxes <strong>an</strong>d <strong>the</strong> radiation topographic controls, which play a<br />

fundamental role in melting.<br />

<strong>Snow</strong> covered area <strong>as</strong> a fraction of total area<br />

The model simulates SWE <strong>as</strong> a continuous variable <strong>an</strong>d, <strong>the</strong>refore, a SWE threshold value<br />

should be chosen to consider <strong>the</strong> pixel <strong>as</strong> <strong>entire</strong>ly covered by snow. Figure 8 reports <strong>the</strong> evolution<br />

in time of <strong>the</strong> snow covered surface <strong>as</strong> fraction of <strong>the</strong> total b<strong>as</strong>in area, using threshold values of 1<br />

203


mm or 10 mm. This h<strong>as</strong> been done to draw a better comparison between <strong>the</strong> model <strong>an</strong>d <strong>the</strong> remote<br />

sensing data, <strong>as</strong> MODIS c<strong>an</strong>not detect snow if SWE is very small. Moreover, <strong>the</strong> MODIS data are<br />

sometimes affected by uncertainty due to <strong>the</strong> persistence of cloud cover, <strong>an</strong>d <strong>the</strong> related r<strong>an</strong>ge of<br />

variability in <strong>the</strong> snow cover area is shown in <strong>the</strong> figure.<br />

The agreement is good, in particular in <strong>the</strong> melting se<strong>as</strong>on. However, <strong>the</strong> limitations of <strong>the</strong><br />

MODIS 8 day product have to be considered. Large uncertainties during <strong>the</strong> precipitation periods<br />

due to <strong>the</strong> cloud cover c<strong>an</strong> occur, <strong>an</strong>d <strong>the</strong> product h<strong>as</strong> a resolution which does not always allow to<br />

appreciate <strong>the</strong> detailed elevation patterns of snow cover, <strong>as</strong> <strong>the</strong> mountain b<strong>as</strong>in considered is<br />

characterized by strong elevation gradients.<br />

The model underestimates <strong>the</strong> snow cover extent in <strong>the</strong> early snow events, probably because it<br />

does not predict well <strong>the</strong> snowfall in <strong>the</strong> bottom of <strong>the</strong> valley, which covers quite a wide area. As<br />

it h<strong>as</strong> been said in <strong>the</strong> previous paragraph, it is difficult to distinguish <strong>the</strong> rain <strong>an</strong>d snow events,<br />

because <strong>the</strong>y are not only dominated by <strong>the</strong> air temperature at <strong>the</strong> ground, but by <strong>the</strong> <strong>entire</strong><br />

temperature profile. The MODIS curve agrees more with <strong>the</strong> 1 mm threshold model result curve:<br />

this probably me<strong>an</strong>s that a thin layer of fresh snow in <strong>the</strong> bottom of <strong>the</strong> valley, where vegetation is<br />

not dense, is e<strong>as</strong>ily recognized by remote sensing techniques.<br />

The model snow cover extent is much less sensitive to <strong>the</strong> threshold value during <strong>the</strong> spring,<br />

when <strong>the</strong> averaged SWE distribution is higher <strong>an</strong>d more strongly dependent on elevation, <strong>as</strong><br />

discussed later.<br />

Figure 8: Fraction of <strong>the</strong> total b<strong>as</strong>in area covered by at le<strong>as</strong>t 1 mm (grey continuous line) <strong>an</strong>d 10 mm (black<br />

continuous line) of SWE, according to <strong>the</strong> model results, <strong>an</strong>d snow covered area according to MODIS<br />

(d<strong>as</strong>hed black lines) with uncertainty r<strong>an</strong>ge due to cloudiness against time.<br />

DEPENDENCE OF SNOW COVER ON ELEVATION AND ASPECT<br />

In this paragraph <strong>the</strong> model results are used to find some relations between SWE, elevation <strong>an</strong>d<br />

<strong>as</strong>pect for <strong>the</strong> b<strong>as</strong>in considered. The b<strong>as</strong>in is here divided into 10 elevation r<strong>an</strong>ks, each sp<strong>an</strong>ning<br />

200 metres. SWE is <strong>the</strong>n averaged in each r<strong>an</strong>k. Figure 9 shows <strong>the</strong> time evolution of SWE for 4<br />

different elevation r<strong>an</strong>ks, <strong>an</strong>d it c<strong>an</strong> be noticed that its maximum value shifts later into spring <strong>as</strong><br />

elevation gets higher. At <strong>the</strong> bottom of <strong>the</strong> valley a strong snowfall occurred at <strong>the</strong> beginning of<br />

March, but at higher elevations o<strong>the</strong>r strong snowfalls occurred later.<br />

204


Figure 9: Time evolution of SWE averaged in 4 elevation r<strong>an</strong>ks<br />

The slope of <strong>the</strong> SWE profile with respect to elevation ch<strong>an</strong>ges after every snowfall <strong>an</strong>d with<br />

<strong>the</strong> se<strong>as</strong>on. Here some examples after 3 different snowfall events - in autumn, in winter, <strong>an</strong>d in<br />

early springtime - are shown in Figures 10, 11, <strong>an</strong>d 12.<br />

Figure 10: Daily averaged SWE-elevation chart for 4 days after a strong autumn snowfall (occurred on 8 th<br />

November 2003)<br />

205


Figure 11: Daily averaged SWE-elevation chart for 4 days after a strong winter snowfall (occurred on 30 th<br />

December 2003)<br />

Figure 12: Daily averaged SWE-elevation chart for 5 days after a strong spring snowfall (occurred on 11 th<br />

March 2004)<br />

For <strong>the</strong> autumn snowfall (Figure 10), at <strong>the</strong> lower elevations SWE exhibits little dependence on<br />

elevation. In fact, snowfall h<strong>as</strong> more or less <strong>the</strong> same intensity at <strong>the</strong> elevations higher th<strong>an</strong> <strong>the</strong><br />

snow limit (r<strong>an</strong>ging from 700 m to 1000 m a.s.l.), but at <strong>the</strong> lower elevations snow w<strong>as</strong> not<br />

formerly present at <strong>the</strong> ground. On <strong>the</strong> contrary, at <strong>the</strong> higher elevations o<strong>the</strong>r snowfalls had<br />

already occurred earlier with different snow limits, <strong>an</strong>d this results in a linear relation. Some days<br />

after <strong>the</strong> snowfall, <strong>an</strong> incre<strong>as</strong>e in temperature causes rainfall events at <strong>the</strong> lower elevations, where<br />

206


melting occurs, where<strong>as</strong> at <strong>the</strong> higher elevations o<strong>the</strong>r snowfall events still occur, <strong>an</strong>d <strong>the</strong> curve<br />

becomes steeper. For <strong>the</strong> winter snowfall (Figure 11), <strong>the</strong> relation is almost linear for all <strong>the</strong><br />

elevations, <strong>as</strong> m<strong>an</strong>y o<strong>the</strong>r snowfall events had formerly occurred, <strong>an</strong>d <strong>the</strong>re is no melting.<br />

Concerning <strong>the</strong> early spring snowfall (Figure 12), just after <strong>the</strong> snowfall <strong>the</strong> relation is linear but<br />

after some days differential melting occurs at <strong>the</strong> lower elevations, this causes a steepening of <strong>the</strong><br />

curve <strong>an</strong>d <strong>an</strong> incre<strong>as</strong>e in its concavity, where<strong>as</strong> at <strong>the</strong> higher elevations <strong>the</strong>re is still snow<br />

accumulation.<br />

In conclusion, <strong>the</strong> relation between SWE <strong>an</strong>d elevation is more or less linear, but in autumn<br />

SWE is weekly dependent on elevation, where<strong>as</strong> <strong>the</strong> dependence is very strong in spring, due to<br />

differential melting <strong>an</strong>d to <strong>the</strong> succession of <strong>the</strong> snowfall events.<br />

Figure 13: Time evolution of <strong>the</strong> ratio between SWE averaged only in north facing pixels <strong>an</strong>d SWE<br />

averaged, for 2 elevation r<strong>an</strong>ks<br />

In Figure 13 <strong>the</strong> dependence of SWE on <strong>as</strong>pect is explored. The ratio between <strong>the</strong> SWE<br />

averaged in <strong>the</strong> north facing pixels <strong>an</strong>d <strong>the</strong> SWE averaged in all pixels for each of <strong>the</strong> elevation<br />

r<strong>an</strong>ks c<strong>an</strong> be <strong>an</strong> index to evaluate <strong>the</strong> import<strong>an</strong>ce of <strong>the</strong> effects exerted by radiation <strong>an</strong>d <strong>as</strong>pect. We<br />

consider <strong>as</strong> north facing pixels those characterized by <strong>as</strong>pect r<strong>an</strong>ging from -20° to +20°, being 0°<br />

<strong>the</strong> north <strong>as</strong>pect. Figure 13 shows <strong>the</strong> time evolution of this ratio for 2 different elevation r<strong>an</strong>ks,<br />

one characteristic of <strong>the</strong> bottom of <strong>the</strong> valley <strong>an</strong>d <strong>the</strong> o<strong>the</strong>r of <strong>the</strong> higher elevations. When <strong>the</strong> ratio<br />

is close to 1 no signific<strong>an</strong>t relation of SWE with <strong>as</strong>pect is present. When differential melting<br />

between N <strong>an</strong>d S facing slopes occurs, <strong>the</strong> ratio incre<strong>as</strong>es to very high values. It c<strong>an</strong> be noticed<br />

that in <strong>the</strong> bottom of <strong>the</strong> valley differential melting is always present all through <strong>the</strong> winter, except<br />

immediately after snowfalls. At <strong>the</strong> higher elevations, on <strong>the</strong> contrary, melting is <strong>as</strong>pect dominated<br />

only in autumn <strong>an</strong>d in spring.<br />

207


CONCLUSION<br />

The comparison of <strong>the</strong> GEOtop SWE results with MODIS snow extent maps shows a fairly<br />

good agreement. The problems of <strong>the</strong> model are mainly due to uncertainties about <strong>the</strong> distribution<br />

of <strong>the</strong> meteorological forcing in space <strong>an</strong>d time, <strong>as</strong> only me<strong>as</strong>urements in some points are<br />

available. Only total precipitation is me<strong>as</strong>ured, <strong>an</strong>d a criterion b<strong>as</strong>ed on two thresholds on <strong>the</strong> air<br />

temperature h<strong>as</strong> been used to distinguish solid <strong>an</strong>d liquid precipitation. However, threshold values<br />

a little higher th<strong>an</strong> <strong>the</strong> values found in literature were needed to reproduce <strong>the</strong> snow cover at <strong>the</strong><br />

bottom of <strong>the</strong> valley. An hourly varying lapse rate w<strong>as</strong> inferred by <strong>the</strong> temperature me<strong>as</strong>urements,<br />

<strong>an</strong>d w<strong>as</strong> used <strong>as</strong> input data for <strong>the</strong> model.<br />

Though characterized by some commission errors, MODIS maps c<strong>an</strong> be valuable tools to<br />

validate distributed models. However, images of finer resolution <strong>an</strong>d more frequent overp<strong>as</strong>sing<br />

time would be better tools.<br />

With <strong>the</strong> results of <strong>the</strong> model we investigated <strong>the</strong> relationship between SWE <strong>an</strong>d elevation <strong>an</strong>d<br />

<strong>as</strong>pect in different se<strong>as</strong>ons. It h<strong>as</strong> been shown that <strong>the</strong> dependence of SWE on elevation is more or<br />

less linear, but tends to be weaker in autumn <strong>an</strong>d stronger in spring. Moreover, differential melting<br />

in north <strong>an</strong>d south facing slopes occurs all through <strong>the</strong> winter at lower elevation (approximately<br />

below 800 m a.s.l.), where<strong>as</strong> it is present only in autumn <strong>an</strong>d in spring at <strong>the</strong> higher elevations.<br />

However, <strong>the</strong>se relations should be verified in o<strong>the</strong>r applications to different <strong>an</strong>d wider b<strong>as</strong>ins.<br />

ACKNOWLEDGEMENT<br />

We th<strong>an</strong>k prof. John Albertson who supported <strong>the</strong> visit of <strong>the</strong> first author at <strong>the</strong> Department of<br />

Civil <strong>an</strong>d Environmental Engineering of Duke University, <strong>an</strong>d Joseph Tom<strong>as</strong>i who corrected a<br />

first version of <strong>the</strong> m<strong>an</strong>uscript.<br />

REFERENCES<br />

Alfnes E, Andre<strong>as</strong>sen LM, Engeset RV, Skaugen T, Udnaes HC. 2004. Temporal variability in<br />

snow distribution. Annals of Glaciology 38: 101-105.<br />

Anderson EA. 1976. A point energy <strong>an</strong>d m<strong>as</strong>s bal<strong>an</strong>ce model of a snow cover. Office of<br />

Hydrology National Wea<strong>the</strong>r Service.<br />

Bertoldi G, Rigon R, Over TM. 2006. Impact of Watershed Geomorphic Characteristics on <strong>the</strong><br />

Energy <strong>an</strong>d Water Budgets. Accepted to Journal of Hydrometeorology.<br />

Brutsaert W. 2004. On a derivable formula for long-wave radiation from clear sky. Water<br />

Resources Research 11: 742-744.<br />

Businger JA, Wyngaard JC, Izumi Y, Bradley EF. 2004. Flux-profile relationship in <strong>the</strong><br />

atmospheric surface layer. Journal of Atmospheric Sciences 28: 181-189.<br />

Carroll T, Cline D, Olheiser C, Rost A, Nilsson A, Fall G, Bovitz C, Li L. 2006. NOAA’s national<br />

snow <strong>an</strong>alyses. Paper presented at <strong>the</strong> 74 th Annual Meeting of <strong>the</strong> Western <strong>Snow</strong> Conference<br />

2006.<br />

Cazorzi F, Dalla Font<strong>an</strong>a G. 1996. <strong>Snow</strong>melt modeling by combining air temperature <strong>an</strong>d a<br />

distributed radiation index. Journal of Hydrology 181: 169-187.<br />

Cline DW, Bales CB, Dozier J. Estimating <strong>the</strong> spatial distribution of snow in mountain b<strong>as</strong>ins<br />

using remote sensing <strong>an</strong>d energy bal<strong>an</strong>ce modeling. Water Resources Research 34(5): 1275-<br />

1285.<br />

Cline D, Elder K, Bales R. 1998. Scale effects in a distributed snow water equivalence <strong>an</strong>d snow<br />

melt model for mountain b<strong>as</strong>ins. Hydrological Processes 12: 1527-1536.<br />

208


Colbeck SC, Anderson EA. 1982. The permeability of a melting snow cover. Water Resources<br />

Research 18(4): 904-908.<br />

Hall DK, Andrew BT, James LF, Alfred TCC, Mil<strong>an</strong> A. 2000. Intercomparison of satellite-derived<br />

snow-cover maps. Annals of Glaciology 31: 369-376.<br />

Hall KH, Riggs GA, Salomonson VV, DiGirolamo NE, Bayr KJ. 2002. MODIS snow-cover<br />

products. Remote Sensing of Environment 83: 181-194.<br />

Idso SB. 1981. A set of equation for full spectrum <strong>an</strong>d 8 to 14 µm <strong>an</strong>d 10.5 to 12.5 µm <strong>the</strong>rmal<br />

radiation from cloudless skies. Water Resources Research 17(2): 295-304.<br />

Iqbal M. 1983. An introduction to solar radiation. Academic Press.<br />

Jord<strong>an</strong> R. 1991. A one dimensional temperature model for a snowcover: Technical documentation<br />

for SNTHERM 89, Spec. Rep. 657, User’s guide. US Army Cold Regions Research <strong>an</strong>d<br />

Engineering Laboratory, H<strong>an</strong>over, NH, USA.<br />

Lee S., Klein AG, Over TM. 2005. A comparison of MODIS <strong>an</strong>d NOHRSC snow-cover products<br />

for simulating streamflow using <strong>the</strong> <strong>Snow</strong>melt Runoff Model. Hydrological Processes 19:<br />

2951-2972.<br />

Marks D, Domingo J, Susong D, Link T, Garen D. 1999. A spatially distributed energy bal<strong>an</strong>ce<br />

snowmelt model for application in mountain b<strong>as</strong>ins. Hydrological Processes 13: 1935-1959.<br />

Martinec J, R<strong>an</strong>go A. 1986. Parameter values for snowmelt runoff modeling. Journal of<br />

Hydrology 84: 197-219.<br />

Neteler M. 2005. Time series processing of MODIS satellite data for l<strong>an</strong>dscape epidemiological<br />

applications. International Journal of Geoinformatics 1(1): 133-138.<br />

Oleson KW, Dai Y, Bon<strong>an</strong> G, Bosilovich M, Dickinson R, Dirmeyer P, Hoffm<strong>an</strong> F, Houser P,<br />

Levis S, Niu G, Thornton P, Vertenstein M, Y<strong>an</strong>g Z, Zeng X. 2004. Technical description of<br />

<strong>the</strong> Community L<strong>an</strong>d Model (CLM). NCAR, Boulder, CO, USA.<br />

Pomeroy JW, Toth B, Gr<strong>an</strong>ger RJ, Hedstrom NR, Essery RLH. 2003. Variation in Surface<br />

Energetics during <strong>Snow</strong>melt in a Subartic Mountain Catchment. Journal of Hydrometeorology<br />

4: 702-719.<br />

Riggs GA, Hall DH, Salomonson VV. 2003. MODIS <strong>Snow</strong> Products User Giude. Electronic<br />

document: http://modis-snow-ice.gsfc.n<strong>as</strong>a.gov/sugkc.html.<br />

Riggs GA, Hall DK. 2006. Evaluation of snow errors in <strong>the</strong> MODIS snow data products. Poster<br />

presented at <strong>the</strong> 63 rd Annual Meeting of <strong>the</strong> E<strong>as</strong>tern <strong>Snow</strong> Conference 2006.<br />

Rigon R, Bertoldi G, Over TM. 2006. GEOtop: A Distributed Hydrological Model with Coupled<br />

Water <strong>an</strong>d Energy Budgets. Accepted to Journal of Hydrometeorology.<br />

Satterlund DR. 1979. An improved equation for estimating longwave radiation from <strong>the</strong><br />

atmosphere. Water Resources Research 15(6): 1649-1650.<br />

Turpin O, Ferguson R, Joh<strong>an</strong>sson B. 1999. Use of remote sensing to test <strong>an</strong>d update simulated<br />

snow cover in hydrological models. Hydrological Processes 13: 2067-2077.<br />

US Army Corps of Engineers. 1956. <strong>Snow</strong> hydrology, Summary report of <strong>the</strong> snow investigations.<br />

US Army Corps of Engineers, North Pacific Division, Portl<strong>an</strong>d, OR, USA.<br />

Z<strong>an</strong>otti F, Endrizzi S, Bertoldi G, Rigon R. 2004. The GEOTOP snow module. Hydrological<br />

Processes 18: 3667-3679.<br />

209


This page is intentionally bl<strong>an</strong>k.<br />

210


ABSTRACT<br />

211<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Microstructural Characterization of Firn<br />

I. BAKER 1 , R. OBBARD 1 , D. ILIESCU 1 , AND D. MEESE 1,2<br />

In this paper, we use a sc<strong>an</strong>ning electron microscope (SEM) coupled with x-ray spectroscopy<br />

<strong>an</strong>d electron back-scattered diffraction patterns to examine firn in cores retrieved by <strong>the</strong> United<br />

States International Tr<strong>an</strong>s-Antarctic Scientific Expedition. From grain boundary grooves we were<br />

able to see where <strong>the</strong> previously-existing snow crystals were joined, <strong>an</strong>d c<strong>an</strong> determine grain sizes.<br />

From <strong>the</strong> SEM images, <strong>the</strong> porosity <strong>an</strong>d <strong>the</strong> surface area per unit volume of <strong>the</strong> pores were<br />

me<strong>as</strong>ured. Finally, we have shown that we c<strong>an</strong> determine <strong>the</strong> microchemistry of impurities in firn<br />

<strong>an</strong>d demonstrated that we c<strong>an</strong> determine <strong>the</strong> orientations of <strong>the</strong> firn crystals.<br />

Keywords: Firn, snow, porosity, surface area, sc<strong>an</strong>ning electron microscopy, x-ray micro<strong>an</strong>alysis<br />

INTRODUCTION<br />

In a number of recent papers, we have used a sc<strong>an</strong>ning electron microscope coupled with<br />

energy-dispersive x-ray micro<strong>an</strong>alysis (EDS) to determine <strong>the</strong> microstructural location of<br />

impurities in ice cores (Cullen <strong>an</strong>d Baker 2000; 2001; 2002a; 2002b; Cullen et al. 2002; Iliescu et<br />

al. 2002; Baker et al. 2003; Baker <strong>an</strong>d Cullen 2003a; 2003b; Obbard et al. 2003a; 2003b; Iliescu<br />

<strong>an</strong>d Baker 2004; Baker et al. 2006). A key feature of this work is that <strong>the</strong> ice specimens were<br />

uncoated, <strong>an</strong>d examined at -80 ± 10 °C so that slight sublimation from <strong>the</strong> ice prevented charge<br />

build-up. This is a technique that we have also used to examine snow (Iliescu <strong>an</strong>d Baker 2002) <strong>an</strong>d<br />

that Barnes <strong>an</strong>d co-workers used to examine both ice cores <strong>an</strong>d snow from Antarctica (Barnes et<br />

al. 2002a; 2002b; 2003). This approach h<strong>as</strong> <strong>an</strong> adv<strong>an</strong>tage over <strong>the</strong> use of aluminum or gold<br />

coating on ice to prevent charging in <strong>the</strong> SEM (Mulv<strong>an</strong>ey, Wolff <strong>an</strong>d Oates 1988; Wolff,<br />

Mulv<strong>an</strong>ey <strong>an</strong>d Oates 1988; Barnes et al. 2002b) since <strong>the</strong> metal coatings c<strong>an</strong> obscure weaker Xray<br />

fluorescence signals from <strong>the</strong> sample.<br />

In o<strong>the</strong>r work, we have demonstrated that Ram<strong>an</strong> spectroscopy coupled with a confocal<br />

sc<strong>an</strong>ning optical microscope c<strong>an</strong> be used to <strong>an</strong>alyze impurities in grain boundaries <strong>an</strong>d triple<br />

junctions that c<strong>an</strong>not be detected by EDS (Iliescu et al. 2004). MicroRam<strong>an</strong> spectroscopy w<strong>as</strong> first<br />

used to look at impurities in a triple junction in ice by Fukazawa <strong>an</strong>d co-workers (1998).<br />

More recently, we have demonstrated that we c<strong>an</strong> obtain <strong>the</strong> complete orientations of crystals in<br />

polycrystalline ice (not simply <strong>the</strong> c-axis direction) with high <strong>an</strong>gular (0.1 o ) <strong>an</strong>d spatial resolution<br />

(50 nm) by using electron back-scattered patterns (EBSPs) obtained in a SEM (Iliescu et al. 2004;<br />

2005; Obbard et al. 2006).<br />

In this paper, we show, for <strong>the</strong> first time, that we c<strong>an</strong> obtain SEM images, EDS <strong>an</strong>d EBSPs from<br />

uncoated firn using <strong>the</strong> same approach. Ultimately, such information c<strong>an</strong> provide <strong>an</strong> efficient <strong>an</strong>d<br />

accurate method of me<strong>as</strong>uring porosity, internal surface area <strong>an</strong>d grain size <strong>an</strong>d will allow us <strong>an</strong><br />

1<br />

Thayer School Of Engineering, Dartmouth College, H<strong>an</strong>over, NH 03755, Email: I<strong>an</strong> Baker<br />

(I<strong>an</strong>.Baker@Dartmouth.Edu)<br />

2<br />

Climate Ch<strong>an</strong>ge Institute, University Of Maine, Orono, ME 04469


underst<strong>an</strong>ding of how impurities are redistributed <strong>an</strong>d how fabric forms in <strong>the</strong> shallow layers of<br />

ice sheets.<br />

EXPERIMENTAL<br />

We examined six firn specimens from depths of 9.71 m, 19.545 m <strong>an</strong>d 34.465 m from core 02-<br />

SP, <strong>an</strong>d depths of 13.019 m, 23.028 m <strong>an</strong>d 38.533 m, from core 02-5 obtained on <strong>the</strong> United States<br />

International Tr<strong>an</strong>s-Antarctic Scientific Expedition (ITASE). Core 02-SP is at <strong>the</strong> South Pole<br />

while core 02-5 is <strong>the</strong> next core along <strong>the</strong> route, see <strong>the</strong> route map at<br />

http://www2.umaine.edu/USITASE/html/map.html for details.<br />

For examination in <strong>the</strong> SEM, ~25-mm- × 25-mm- × 10-mm-thick specimens were cut<br />

perpendicular to <strong>the</strong> core axis, shaved flat with a razor blade under a HEPA-filtered, laminar-flow<br />

hood at -10°C <strong>an</strong>d frozen onto a br<strong>as</strong>s plate. Specimens were <strong>the</strong>n ei<strong>the</strong>r sealed in a small<br />

container for later examination or mounted onto a cold stage for immediate observation. For<br />

imaging <strong>an</strong>d micro<strong>an</strong>alysis, <strong>the</strong> uncoated specimens were examined at -80°C ± 10°C using a JEOL<br />

5310 low-vacuum SEM operated at 10 kV equipped with a Princeton Gamma-Tech IMIX EDS<br />

system utilizing a lithium-drifted silicon thin-window detector. For orientation determination,<br />

specimens were frozen to a purpose-built pre-tilted br<strong>as</strong>s sled that w<strong>as</strong> <strong>the</strong>n mounted on a cold<br />

stage in a field emission gun FEI XL-30 environmental SEM. The SEM w<strong>as</strong> operated at 15 kV<br />

with a beam current of 0.15 nA, <strong>an</strong>d <strong>the</strong> ice w<strong>as</strong> again examined at temperatures between -80°C ±<br />

10°C at a pressure of approximately 5 × 10 –4 Pa. Secondary electron imaging w<strong>as</strong> used on both<br />

SEMs. EBSPs were captured using <strong>the</strong> techniques described by Iliescu <strong>an</strong>d o<strong>the</strong>rs (Iliescu et al.<br />

2004; 2005) <strong>an</strong>d indexed using HKL Technologies’ CHANNEL 5 software.<br />

RESULTS AND DISCUSSION<br />

Figure 1 shows low-magnification secondary electron images of firn from three different depths.<br />

The images were obtained <strong>as</strong> soon <strong>as</strong> possible after <strong>the</strong> specimens were inserted into <strong>the</strong> SEM to<br />

reduce <strong>the</strong> degree of sublimation, which c<strong>an</strong> affect <strong>the</strong> appear<strong>an</strong>ce of <strong>the</strong> specimens. Several<br />

features are worth noting. First, shaving <strong>the</strong> ice produced some debris, which fell into <strong>the</strong> pores in<br />

<strong>the</strong> ice. This w<strong>as</strong> much more noticeable in <strong>the</strong> higher porosity shallow ice, presumably because <strong>the</strong><br />

grains were less strongly bonded toge<strong>the</strong>r <strong>an</strong>d <strong>the</strong> pore volume is larger. Examples of such debris<br />

are labeled “D” in Figure 1(a).<br />

Second, grain boundaries grooves were observed in some images. These indicate where <strong>the</strong><br />

snow crystals are joined toge<strong>the</strong>r, see Figure 1 (a). By allowing <strong>the</strong> specimens to sublimate in <strong>the</strong><br />

SEM it is possible to see grain boundary grooves between all <strong>the</strong> individual ice crystals <strong>an</strong>d,<br />

hence, it should be possible to determine how <strong>the</strong> grain size ch<strong>an</strong>ges <strong>as</strong> a function of depth in <strong>the</strong><br />

firn. However, one needs to be aware that longer sublimation times, due to loss of m<strong>as</strong>s,<br />

dr<strong>as</strong>tically affect <strong>the</strong> appear<strong>an</strong>ce of <strong>the</strong> microstructure <strong>an</strong>d hence a delicate compromise needs to<br />

be achieved between <strong>the</strong> need to depict <strong>the</strong> microstructure in <strong>an</strong> accurate unadulterated m<strong>an</strong>ner<br />

<strong>an</strong>d <strong>the</strong> desired to better visualize <strong>the</strong> details within <strong>the</strong> grain boundary region.<br />

Third, it is straightforward to determine <strong>the</strong> % areal porosity from <strong>the</strong>se images by dividing <strong>the</strong><br />

area occupied by <strong>the</strong> shaved flat regions by <strong>the</strong> area of <strong>the</strong> image. The pore area <strong>an</strong>d length of<br />

boundary area were me<strong>as</strong>ured on <strong>the</strong> SEM images using Image SXM (Barrett 2005), a derivative<br />

of NIH Image (R<strong>as</strong>b<strong>an</strong>d 1997). The boundaries of pore are<strong>as</strong> were traced with <strong>the</strong> drawing tool,<br />

<strong>an</strong>d <strong>the</strong> perimeter <strong>an</strong>d enclosed area me<strong>as</strong>ured with <strong>the</strong> pixel-counting me<strong>as</strong>urement utility, see<br />

Figure 2. As is quite common in <strong>the</strong> SEM, <strong>the</strong> edges of <strong>the</strong> flat surface tended to charge. This<br />

<strong>an</strong>alysis w<strong>as</strong> performed from multiple images at each depth (except for <strong>the</strong> 9.71 m specimen,<br />

where only one image w<strong>as</strong> available). The porosity values obtained at each depth are shown in<br />

Table 1. As might be expected, (except for <strong>the</strong> 9.71 m specimen), <strong>the</strong> porosity at a particular site<br />

decre<strong>as</strong>es with incre<strong>as</strong>ing depth.<br />

212


A B<br />

Figure 1. Low-magnification secondary electron images of firn samples from depths of (a) 9.71 m, (b) 23.028<br />

m, <strong>an</strong>d (c) 34.465 m. The features labeled “D” in (a) are debris produced by shaving. Some grain boundaries<br />

where <strong>the</strong> original snow crystals joined toge<strong>the</strong>r, <strong>as</strong> indicated by grooves, are labeled “GB”. Some air bubbles<br />

are indicated by “A”.<br />

Table 1. Porosity, me<strong>as</strong>ured length of internal surface around <strong>the</strong> pores in 16 mm2 are<strong>as</strong>, length of<br />

(internal surface) line per unit area, LA, <strong>an</strong>d internal pore surface area per unit volume, SV, for<br />

various depths at two sites from <strong>the</strong> US ITASE me<strong>as</strong>ured from SEM images.<br />

Site Depth<br />

(m)<br />

%<br />

Porosity<br />

C<br />

Me<strong>as</strong>ured length<br />

of surface (mm)<br />

213<br />

LA (mm -1 ) SV (mm -1 )<br />

02-SP 9.71 80 33.2 2.08 2.65<br />

02-5 13.019 85 28.7 1.79 2.28<br />

02-SP 19.545 87 44.5 2.78 3.54<br />

02-5 23.028 58 30.4 1.98 2.52<br />

02-SP 34.465 36 40.7 2.54 3.23<br />

02-5 38.533 38 40.6 2.54 3.23


Fourth, by me<strong>as</strong>uring <strong>the</strong> length of <strong>the</strong> surface around <strong>the</strong> pores per unit area it is possible to<br />

obtain <strong>the</strong> internal surface area per unit volume, see Figure 2. Microstructural examinations are<br />

normally performed on firn by infiltrating <strong>the</strong> firn with a liquid, e.g. dimethyl Phthalate (Albert<br />

<strong>an</strong>d Shultz 2002; Rick <strong>an</strong>d Albert 2004), allowing it to set, <strong>an</strong>d <strong>the</strong>n sublimating <strong>the</strong> firn. The<br />

problems with this method are that <strong>the</strong> viscous liquid c<strong>an</strong>not e<strong>as</strong>ily penetrate small pores, which<br />

incre<strong>as</strong>e in frequency with depth; <strong>an</strong>d, of course, <strong>the</strong> liquid c<strong>an</strong>not infiltrate closed off pores,<br />

which also incre<strong>as</strong>e in frequency with depth. The lengths of projected surface around <strong>the</strong> pores<br />

me<strong>as</strong>ured from images of area 16 mm 2 are shown along with <strong>the</strong> length of (internal surface) line<br />

per unit area, LA, in Table 1. The internal surface area per unit volume, SV, w<strong>as</strong> calculated from<br />

LA using: SV = (4/π) LA (Underwood 1970). In contr<strong>as</strong>t to <strong>the</strong> volume porosity, <strong>the</strong> internal surface<br />

area per unit volume of <strong>the</strong> pores w<strong>as</strong> generally found to incre<strong>as</strong>e with incre<strong>as</strong>ing depth. This is <strong>an</strong><br />

unexpected result. It may indicate that <strong>the</strong> pores are more convoluted at greater depth even though<br />

<strong>the</strong>ir percentage of <strong>the</strong> volume is less. A more detailed SEM <strong>an</strong>alysis would involve taking both<br />

horizontal <strong>an</strong>d vertical sections. This becomes more import<strong>an</strong>t with incre<strong>as</strong>ing depth where <strong>the</strong><br />

pore <strong>an</strong>d grain structures become more inhomogeneous due to <strong>the</strong> flattening of <strong>the</strong> microstructure<br />

from <strong>the</strong> incre<strong>as</strong>ing overburden. Thus, it is possible that this sectioning issue may be at <strong>the</strong> root of<br />

<strong>the</strong> unexpected incre<strong>as</strong>e in internal surface area with incre<strong>as</strong>ing depth.<br />

Fifth, <strong>the</strong> edges of <strong>the</strong> flats shaved on <strong>the</strong> specimens preferentially etch before <strong>the</strong> rest of <strong>the</strong><br />

surface, see Figures 1(b) <strong>an</strong>d 1(c).<br />

And, sixth, some air bubbles c<strong>an</strong> be observed, which have been trapped in <strong>the</strong> ice, see Figures<br />

1(b) <strong>an</strong>d 1(c).<br />

Figure 2. Secondary electron image of firn sample from 13.019 m. The boundaries of <strong>the</strong> pore are<strong>as</strong> were<br />

traced with <strong>the</strong> drawing tool, <strong>an</strong>d <strong>the</strong> perimeter <strong>an</strong>d enclosed area me<strong>as</strong>ured with <strong>the</strong> pixel-counting<br />

me<strong>as</strong>urement utility of Image SXM (Barrett 2005).<br />

Figure 3 shows a higher magnification image of two grains in firn. In this c<strong>as</strong>e, ridges<br />

(indicated) appear to have formed on ei<strong>the</strong>r side of <strong>the</strong> grain boundary groove. Because <strong>the</strong> ridges<br />

were well below <strong>the</strong> surface of <strong>the</strong> ice, EDS <strong>an</strong>alysis w<strong>as</strong> not possible. Adams et al. (2001)<br />

observed a grain boundary ridge on snow crystals that were sintering toge<strong>the</strong>r, a feature that <strong>the</strong>y<br />

suggested w<strong>as</strong> indicative of direct evidence for grain boundary diffusion <strong>as</strong> a sintering mech<strong>an</strong>ism<br />

in ice. However, Barnes (2003) later pointed out that <strong>the</strong> ridge may have been a grain boundary<br />

impurity filament of <strong>the</strong> kind observed previously in ice (Baker <strong>an</strong>d Cullen 2003a; 2003b; Baker et<br />

al. 2003; Baker et al. 2006; Barnes et al. 2002a; 2002b; Cullen <strong>an</strong>d Baker 2000; 2001; 2002a;<br />

2002b; Cullen et al. 2002; Iliescu et al. 2002; Iliescu <strong>an</strong>d Baker 2004; Obbard et al. 2003a; 2003b).<br />

214


GB ridge<br />

GB groove<br />

Figure 3. Higher-magnification secondary electron images of firn samples showing ridging on ei<strong>the</strong>r side of a<br />

grain boundary groove (arrowed). The vertical lines are sc<strong>an</strong> faults.<br />

Figure 4 shows x-ray spectra taken from white spots formed during <strong>the</strong> sublimation of <strong>the</strong><br />

surface of firn specimens. The oxygen peak is from <strong>the</strong> ice, while <strong>the</strong> carbon peak arises from <strong>the</strong><br />

breakdown of diffusion pump oil under <strong>the</strong> electron beam in <strong>the</strong> SEM. In <strong>the</strong> shallow samples<br />

(9.71 m <strong>an</strong>d 13.019 m), for example Figure 4(a), only small, barely-detectable qu<strong>an</strong>tities of<br />

impurities were found in <strong>the</strong> white spots. In this c<strong>as</strong>e, S <strong>an</strong>d Cl were found. Spectra from white<br />

spots on deeper firn specimens showed more pronounced peaks. The example spectra shown in<br />

Figure 4(b) contain Cl, K, Na, S, Si <strong>an</strong>d possibly Ca. Such impurities have often been noted in<br />

white spots in ice specimens from Greenl<strong>an</strong>d <strong>an</strong>d Antarctica (Cullen <strong>an</strong>d Baker 2001; 2002b Baker<br />

<strong>an</strong>d Cullen 2003a; Barnes et al. 2002a; 2002b; Obbard et al. 2003a; 2003b; Baker et al. 2006). It is<br />

evident that impurities are spread throughout <strong>the</strong> firn. We did not find <strong>an</strong>y impurities segregated to<br />

<strong>the</strong> grain boundaries <strong>as</strong> h<strong>as</strong> been found in ice from Antarctica <strong>an</strong>d Greenl<strong>an</strong>d (Cullen <strong>an</strong>d Baker<br />

2000; 2001; 2002b; Cullen et al. 2002; Barnes et al. 2002a; 2002b; Baker et al. 2003; Baker <strong>an</strong>d<br />

Cullen 2003a; 2003b; Obbard et al. 2003a; 2003b; Baker et al. 2006).<br />

Finally, Figure 5 shows <strong>an</strong> EBSP from firn from 3 m. The reds lines indicate <strong>the</strong> centers of <strong>the</strong><br />

indexed Kikuchi b<strong>an</strong>ds. The pattern, while not <strong>as</strong> good quality <strong>as</strong> patterns previously presented for<br />

ice (Cullen <strong>an</strong>d Baker 2002a; Iliescu et al. 2004; 2005; Obbard et al. 2006) is of sufficient quality<br />

that it is indexable. Various poles are indicated on <strong>the</strong> pattern. Each EBSP contain <strong>the</strong> complete<br />

three-dimensional orientation information of <strong>the</strong> crystal from whence it came <strong>an</strong>d toge<strong>the</strong>r <strong>the</strong><br />

patterns enable <strong>the</strong> misorientations between grains to be determined.<br />

215<br />

GB ridge


A<br />

B<br />

Figure 4. X-ray spectra taken from white spots, such <strong>as</strong> that indicated in Figure 1(c), on <strong>the</strong> surface of firn<br />

samples from depths of: (a) 9.71 m; <strong>an</strong>d (b) 38.533 m. The full scale counts are 1130 for (a) <strong>an</strong>d 230 from <strong>the</strong><br />

three spectra in (b).<br />

216


CONCLUSION<br />

Figure 5. Indexed electron backscatter diffraction pattern from firn from a depth of 3 m.<br />

In conclusion, we have shown that it is possible <strong>an</strong>d useful, using various adv<strong>an</strong>ced electronoptical<br />

techniques <strong>an</strong>d x-ray spectroscopy me<strong>as</strong>urements in a SEM, to examine firn at high<br />

resolution. By observing grain boundary grooves, one c<strong>an</strong> determine where snow crystals were<br />

joined; <strong>the</strong>ir orientations c<strong>an</strong> be determined from EBSPs, allowing <strong>the</strong> misorientations between<br />

grains to be calculated; <strong>the</strong> porosity, approximate grain size <strong>an</strong>d internal pore surface area c<strong>an</strong> be<br />

me<strong>as</strong>ured; <strong>an</strong>d <strong>the</strong> microchemistry of impurities c<strong>an</strong> be determined by EDS.<br />

ACKNOWLEDMENTS<br />

The cores were collected during <strong>the</strong> 2005 US International Tr<strong>an</strong>s Antarctic Scientific Expedition<br />

<strong>an</strong>d samples provided from NSF Gr<strong>an</strong>t 9980434. Support by <strong>the</strong> U.S. National Science Foundation<br />

Office of Polar Programs Gr<strong>an</strong>t 0440523 <strong>an</strong>d U.S. Army Research Office under contract DAAD<br />

19-03-1-0110 are gratefully acknowledged. The views <strong>an</strong>d conclusions contained herein are those<br />

of <strong>the</strong> authors <strong>an</strong>d should not be interpreted <strong>as</strong> necessarily representing official policies, ei<strong>the</strong>r<br />

expressed or implied, of <strong>the</strong> National Institute of St<strong>an</strong>dards <strong>an</strong>d Technologies, or <strong>the</strong> U.S.<br />

Government.<br />

REFERENCES<br />

Adams EE, Miller DA, Brown RL. 2001. Grain boundary ridge on sintered bonds between ice<br />

crystals. Journal of Applied Physics 90: 5782–5785.<br />

Albert M, Shultz EF. 2002. <strong>Snow</strong> <strong>an</strong>d firn properties <strong>an</strong>d air–snow tr<strong>an</strong>sport processes at Summit,<br />

Greenl<strong>an</strong>d. Atmospheric Environment 36: 2789–2797.<br />

Baker I, Cullen D. 2003a. The Structure <strong>an</strong>d Chemistry of 94m GISP2 ice. Annals of Glaciology<br />

35: 224–230.<br />

Baker I, Cullen D. 2003b. SEM/EDS observations of impurities in polar ice: artifacts or not?<br />

Journal of Glaciology 49(16): 184–190.<br />

217


Baker I, Cullen D, Iliescu D. 2003. The microstructural location of impurities in ice. C<strong>an</strong>adi<strong>an</strong><br />

Journal of Physics 81: 1–9.<br />

Baker I, Iliescu D, Obbard R, Ch<strong>an</strong>g H, Bostick B, Daghli<strong>an</strong> CP. 2006. Microstructural<br />

Characterization of Ice Cores. Annals of Glaciology, in press.<br />

Barnes PRF, Mulv<strong>an</strong>ey R, Robinson K, Wolff EW. 2002a. Observations of polar ice from <strong>the</strong><br />

Holocene <strong>an</strong>d <strong>the</strong> glacial period using <strong>the</strong> sc<strong>an</strong>ning electron microscope. Annals of Glaciology<br />

35: 559–566.<br />

Barnes PRF, Mulv<strong>an</strong>ey R, Wolff EW, Robinson K. 2002b. A technique for <strong>the</strong> examination of<br />

polar ice using <strong>the</strong> sc<strong>an</strong>ning electron microscope. Journal of Microscopy 205: 118–124.<br />

Barnes PRJ. 2003. Comment on "Grain boundary ridge on sintered bonds between ice crystals" [J.<br />

Appl. Phys. 90, 5782 (2001)]. Journal of Applied Physics 93: 783–785.<br />

Barnes PRJ, Wolff EW, Mallard DC, Meider HM. 2003. SEM Studies of <strong>the</strong> Morphology <strong>an</strong>d<br />

Chemistry of Polar Ice. Microscopy Research <strong>an</strong>d Technique 62: 62–69.<br />

Barrett S. 2005. Image SXM, Surface Science Research Centre, University of Liverpool,<br />

Liverpool, UK.<br />

Cullen D, Baker I. 2000. The Chemistry of Grain Boundaries in Greenl<strong>an</strong>d Ice. Journal of<br />

Glaciology 46: 703–706.<br />

Cullen D, Baker I. 2001. Observation of Impurities in Ice. Microscopy Research <strong>an</strong>d Technique<br />

55: 198–207.<br />

Cullen D, Baker I. 2002a. Sulfate Crystallites in Vostok Accretion Ice. Materials Characterization<br />

48: 263–270.<br />

Cullen D, Baker I. 2002b. Sc<strong>an</strong>ning Electron Microscopy of Vostok Acrretion Ice. In Proceedings<br />

of Microscopy <strong>an</strong>d Micro<strong>an</strong>alysis; 1546-7CD.<br />

Cullen D, Iliescu D, Baker I. 2002. SEM/EDS Studies of Impurities in Natural Ice. In Proceedings<br />

of Microscopy <strong>an</strong>d Micro<strong>an</strong>alysis; 1398-9CD.<br />

Fukazawa H, Sugiyama K, Mae S, Narita H, Hondoh T. 1998. Acid ions at triple junction of<br />

Antarctic ice observed by Ram<strong>an</strong> scattering. Geophysical Research Letters, 25: 2845.<br />

Iliescu D, Baker I. 2002. Imaging of Uncoated <strong>Snow</strong> Crystals Using a Low-Vacuum Sc<strong>an</strong>ning<br />

Electron Microscope. Journal of Glaciology 48(162): 479–480.<br />

Iliescu D, Baker I, Cullen D. 2002. Preliminary Microstructural <strong>an</strong>d Microchemical Observations<br />

of Pond <strong>an</strong>d River Accretion Ice. Cold Regions Science <strong>an</strong>d Engineering 35: 81–99.<br />

Iliescu D, Baker I. 2004. Characterization of <strong>the</strong> microstructure <strong>an</strong>d mech<strong>an</strong>ical properties in<br />

se<strong>as</strong>onal lake <strong>an</strong>d river ice. In Proc. 24 th Army Science Conference, Orl<strong>an</strong>do, FL; OS-26.<br />

Iliescu D, Baker I, Ch<strong>an</strong>g H. 2004. Determining <strong>the</strong> Orientations of Ice Crystals using Electron<br />

Backscatter Patterns. Microscopy Research <strong>an</strong>d Technique 63: 183–187.<br />

Iliescu D, Baker I, Daghli<strong>an</strong> CP. 2005. Orientation Mapping in Polycrystalline Ice Using Electron<br />

Backscatter Patterns. In Proceedings of Microscopy & Micro<strong>an</strong>alysis, Honolulu, HI; 1500–<br />

1501.<br />

Mulv<strong>an</strong>ey R, Wolff EW, Oates K. 1988. Sulphuric acid at grain boundaries in Antarctic ice.<br />

Nature 331: 247.<br />

Obbard R, Iliescu D, Cullen D, Ch<strong>an</strong>g J, Baker I. 2003a. SEM/EDS Comparison of Polar <strong>an</strong>d<br />

Se<strong>as</strong>onal Temperate Ice. Microscopy Research <strong>an</strong>d Technique 62: 49–61.<br />

Obbard R, Iliescu D, Baker I, Cullen D. 2003b. A Technique for <strong>the</strong> Sc<strong>an</strong>ning Electron<br />

Microscopy <strong>an</strong>d Micro<strong>an</strong>alysis of Uncoated Ice Crystals. In Electron Microscopy: Its Role in<br />

Materials Science, The Mike Meshii Symposium, TMS, Warrendale, PA; 133–140.<br />

Obbard R, Baker I, Sieg K. 2006. Using Electron Backscatter Diffraction Patterns To Examine<br />

Recystallization in Polar Ice Sheets. Submitted to Journal of Glaciology.<br />

R<strong>as</strong>b<strong>an</strong>d W. S. 1997–2006. NIH ImageJ, http://rsb.info.nih.gov/ij/.<br />

Rick U, Albert, M. 2004. Microstructure of West Antarctic Firn <strong>an</strong>d its Effect on Air Permeability.<br />

ERDC/CREEL Report TR-04-16.<br />

Underwood EE. 1970. Qu<strong>an</strong>titative Stereology, Addision-Wesley Publishing Co., Reading, M<strong>as</strong>s,<br />

U.S.A., p24.<br />

Wolff EW, Mulv<strong>an</strong>ey R, Oates K. 1998. The location of impurities in Antarctic ice. Annals of<br />

Glaciology, 11: 194.<br />

218


219<br />

63rd EASTERN SNOW CONFERENCE<br />

Delaware, DE, USA 2005<br />

Computational Time Steps of Winter Water Bal<strong>an</strong>ce<br />

for <strong>Snow</strong> Losses at United States Meteorological Stations<br />

ABSTRACT<br />

S.R. FASSNACHT 1<br />

When estimating <strong>the</strong> water bal<strong>an</strong>ce for a cold region watershed, that is one that receive a<br />

subst<strong>an</strong>tial portion of its <strong>an</strong>nual precipitation <strong>as</strong> snow, accumulation <strong>an</strong>d o<strong>the</strong>r winter hydrological<br />

processes must be considered. For m<strong>an</strong>y of <strong>the</strong>ses watersheds, all but <strong>the</strong> most fundamental<br />

meteorological data (temperature <strong>an</strong>d precipitation), are ei<strong>the</strong>r not me<strong>as</strong>ured or not me<strong>as</strong>ured at a<br />

re<strong>as</strong>onable time step. Of particular import<strong>an</strong>ce are wind data, <strong>as</strong> wind influences underestimates of<br />

precipitation due to wind undercatch <strong>an</strong>d losses of snow from <strong>the</strong> snowpack, specifically,<br />

snowpack sublimation, <strong>an</strong>d <strong>the</strong> occurrence <strong>an</strong>d magnitude of blowing snow. Estimating snow<br />

accumulation to yield snowmelt amounts requires summing of gauged precipitation <strong>an</strong>d gauge<br />

undercatch, <strong>an</strong>d subtracting minus snowpack sublimation <strong>an</strong>d blowing snow tr<strong>an</strong>sport. The first<br />

two components are computed on a daily time step, while <strong>the</strong> latter two are computed on <strong>an</strong> hourly<br />

time step. From five National Wea<strong>the</strong>r Service meteorological stations, <strong>the</strong> variations in computed<br />

snowpack m<strong>as</strong>s losses minus undercatch using data at different time intervals show that at most<br />

sites it is difficult to use monthly time steps for computations derived using hourly or daily data.<br />

At <strong>the</strong> relative dry <strong>an</strong>d cold Leadville, Colorado site <strong>the</strong> computations were tr<strong>an</strong>sferable between<br />

time steps.<br />

Keywords: solid precipitation, meteorological data, undercatch, sublimation, blowing snow<br />

INTRODUCTION<br />

<strong>Snow</strong> is disturbed by wind, from snowfall to movement through redistribution to sublimation.<br />

<strong>Snow</strong>fall qu<strong>an</strong>tities are underestimated from precipitation gauges due to undercatch from wind,<br />

wetting of gauges, <strong>an</strong>d to a lesser degree evaporation (Goodison et al., 1998). Undercatch due to<br />

wind is caused by <strong>the</strong> deformation of <strong>the</strong> wind field around <strong>the</strong> gauge orifice. As well, falling<br />

snow crystals are more e<strong>as</strong>ily blown away from <strong>the</strong> gauge orifice th<strong>an</strong> rain drops. <strong>Snow</strong> on <strong>the</strong><br />

ground c<strong>an</strong> be redistributed b<strong>as</strong>ed on wind characteristics, upwind <strong>an</strong>d downwind fetch length, <strong>an</strong>d<br />

<strong>the</strong> history of <strong>the</strong> snowpack surface (Pomeroy et al., 1991). Wind across a snowpack (or across<br />

snow held by vegetation) c<strong>an</strong> sublimate snow away from or towards <strong>the</strong> surface depending upon<br />

temperature <strong>an</strong>d humidity profiles (Sverdrup, 1936).<br />

A watershed <strong>an</strong>alysis uses <strong>the</strong> water bal<strong>an</strong>ce to partition water storage <strong>an</strong>d movement into<br />

different components of <strong>the</strong> hydrological cycle. At <strong>the</strong> end of <strong>the</strong> accumulation period, <strong>the</strong><br />

remaining snow melts to contribute to runoff. Accumulation, given <strong>as</strong> snow water equivalent<br />

(SWE), is typically estimated <strong>as</strong> <strong>the</strong> cumulative precipitation occurring at air temperatures colder<br />

th<strong>an</strong> freezing (0 degrees Celsius), without correcting for precipitation underestimation due to<br />

gauge undercatch, nor snowpack losses due to sublimation or blowing snow.<br />

1<br />

Watershed Science Program, College of Natural Resources, Colorado State University, Fort<br />

Collins, Colorado 80523-1472 USA.


F<strong>as</strong>snacht (2004) used equations derived from field me<strong>as</strong>urements to estimate gauge undercatch<br />

<strong>an</strong>d compared it to sublimation <strong>an</strong>d blowing snow from six wea<strong>the</strong>r stations across <strong>the</strong> United<br />

States (U.S.) in order to adjust monthly <strong>an</strong>d se<strong>as</strong>onal accumulation. With <strong>the</strong>se considerations, <strong>the</strong><br />

amount of SWE that accumulates c<strong>an</strong> be computed <strong>as</strong><br />

SWE = P + P m F m q<br />

(1)<br />

g<br />

U<br />

E<br />

BS<br />

where Pg is <strong>the</strong> me<strong>as</strong>ured amount of precipitation, PU is <strong>the</strong> estimated amount of gauge<br />

underestimation (hereinafter <strong>as</strong>sumed to be mainly due to undercatch), FE is <strong>the</strong> amount of<br />

sublimation (away from or towards <strong>the</strong> snowpack), <strong>an</strong>d qBS is <strong>the</strong> amount of blowing snow<br />

redistributed (scoured away from or deposited at <strong>the</strong> snowpack). The precipitation (me<strong>as</strong>ured plus<br />

undercatch) is <strong>an</strong> accumulation of snow, <strong>as</strong> estimated from a gauge, where<strong>as</strong> sublimation <strong>an</strong>d<br />

blowing snow are losses from <strong>the</strong> snowpack which is h<strong>as</strong> accumulated beside a precipitation<br />

gauge. These components c<strong>an</strong> be computed for a point location using meteorological data.<br />

F<strong>as</strong>snacht (2004) compared <strong>the</strong>se components to see if me<strong>as</strong>ured precipitation, without<br />

consideration of undercatch or o<strong>the</strong>r bi<strong>as</strong>es, could be used <strong>as</strong> <strong>an</strong> estimate of snowpack<br />

accumulation after sublimation <strong>an</strong>d blowing snow had reduced accumulation.<br />

For U.S. National Wea<strong>the</strong>r Service (NWS) automated surface observation stations (ASOS),<br />

meteorological data are reported over <strong>an</strong> hourly interval (to be used to compute FE <strong>an</strong>d qBS in<br />

equation 1). The NWS cooperative (COOP) stations data are reported over a daily interval (to be<br />

used to compute Pg <strong>an</strong>d PU in equation 1). These data <strong>an</strong>d monthly summaries are available online<br />

via <strong>the</strong> NWS National Climate Data Center (NCDC, 2006). Data are presented <strong>as</strong> qu<strong>an</strong>tities, with<br />

<strong>the</strong> exception of precipitation events that are less th<strong>an</strong> 0.254 mm (0.01 inches), which are reported<br />

<strong>as</strong> trace events. F<strong>as</strong>snacht (2004) <strong>as</strong>sumed that <strong>the</strong>se trace events yielded precipitation at one half<br />

of <strong>the</strong> minimum detection (0.127 mm). Y<strong>an</strong>g et al. (1998a) stated that trace events c<strong>an</strong> be<br />

signific<strong>an</strong>t in drier environments, such <strong>as</strong> Al<strong>as</strong>ka.<br />

F<strong>as</strong>snacht (2004) scrutinized <strong>the</strong> validity of undercatch, sublimation <strong>an</strong>d redistribution estimates<br />

in trying to determine if <strong>an</strong>d where PU is approximately equal to FE plus qBS, so that SWE c<strong>an</strong> be<br />

set to Pg for equation 1. Considering that water bal<strong>an</strong>ce computations are typically made for<br />

monthly intervals, this paper compares <strong>the</strong> components of equation 1 for individual winter months<br />

<strong>an</strong>d <strong>the</strong> <strong>entire</strong> winter se<strong>as</strong>on <strong>as</strong> computed using different time steps. Specifically <strong>the</strong> objectives are<br />

1) to compare <strong>the</strong> tr<strong>an</strong>sferability of computed snow loss rates (precipitation undercatch, snowpack<br />

sublimation <strong>an</strong>d blowing snow tr<strong>an</strong>sport) over different time scales (hourly, daily, <strong>an</strong>d monthly);<br />

2) for monthly undercatch to determine if <strong>the</strong>re is a difference using monthly average (for<br />

temperature <strong>an</strong>d wind speed, with totals for precipitation) of <strong>the</strong> daily data (hereinafter called<br />

average monthly data) versus using monthly data adjusted for <strong>the</strong> monthly probability of each<br />

precipitation type (snow, mixed precipitation, or rain) toge<strong>the</strong>r with <strong>the</strong> average wind speed during<br />

each precipitation type (hereinafter called monthly ph<strong>as</strong>e partitioned data); <strong>an</strong>d 3) to determine if<br />

monthly or se<strong>as</strong>onal gauged precipitation c<strong>an</strong> be used to estimate discrep<strong>an</strong>cies in computations of<br />

Pg, PU, FE <strong>an</strong>d qBS from different time steps. Since <strong>the</strong> precipitation undercatch equations were<br />

derived from data at daily interval, undercatch w<strong>as</strong> not computed using hourly data.<br />

STUDY SITES<br />

Four of <strong>the</strong> six meteorological stations across <strong>the</strong> conterminous U.S. used by F<strong>as</strong>snacht (2004)<br />

were <strong>an</strong>alysed in this study (Table 1). Pullm<strong>an</strong> WA w<strong>as</strong> been substituted for <strong>the</strong> St<strong>an</strong>ley ID<br />

station, since <strong>the</strong>re were no observed trace events at St<strong>an</strong>ley during <strong>the</strong> study period. Pullm<strong>an</strong> WA<br />

h<strong>as</strong> a similar climate to St<strong>an</strong>ley (Table 2 <strong>an</strong>d F<strong>as</strong>snacht, 2004), receiving 6 mm more precipitation<br />

per winter month, being warmer (–0.6 degrees C average air temperature versus –5.9 degrees C),<br />

more humid (a vapour pressure of 4.9 mb versus 3.3 mb), <strong>an</strong>d more windy (4.1 m s –1 average wind<br />

speed versus 1.3 m s –1 ), but having <strong>the</strong> same vapour pressure deficit. The South Lake Tahoe<br />

station w<strong>as</strong> not used, <strong>as</strong> <strong>the</strong> no suitable undercatch equation h<strong>as</strong> been derived for <strong>the</strong> heating<br />

tipping bucket gauge used to estimate daily precipitation.<br />

220


Meteorological data for <strong>the</strong> winter (October–June) of three water years (2000–2002) were<br />

retrieved from <strong>the</strong> NCDC online datab<strong>as</strong>e (NCDC, 2006). Data were available only for water years<br />

2001 <strong>an</strong>d 2002 for <strong>the</strong> Rawlins WY station (Table 1).<br />

<strong>Snow</strong> depths were not recorded, thus, hourly temperature <strong>an</strong>d precipitation data were examined<br />

for each year for each station to determine when snow started to accumulate <strong>an</strong>d when it ablation<br />

w<strong>as</strong> likely complete. These dates were rounded to <strong>the</strong> nearest month (Table 1). While <strong>the</strong> ph<strong>as</strong>e of<br />

precipitation w<strong>as</strong> not known for <strong>the</strong> study sites, it w<strong>as</strong> observed that only in <strong>the</strong> later winter<br />

months did precipitation occ<strong>as</strong>ionally occur at air temperatures warmer th<strong>an</strong> 0 degrees C. This<br />

factor also helped determine <strong>the</strong> start <strong>an</strong>d end of ablation. The monthly average precipitation,<br />

temperature, wind speed <strong>an</strong>d vapour pressure for <strong>the</strong> winter months are summarized in Table 2, for<br />

<strong>the</strong> winter months given in Table 1.<br />

Table 1. Summary of stations used in <strong>an</strong>alyses, <strong>an</strong>d <strong>the</strong> periods considered winter for each of <strong>the</strong> three<br />

water years of interest (2000–2002). The precipitation gauge type is denoted <strong>as</strong> SRN for <strong>the</strong> NWS<br />

st<strong>an</strong>dard 8" rain gauge or BUG for <strong>the</strong> Belfort Universal Recording Rain Gauge.<br />

elevation latitude longitude<br />

winter period precipitation<br />

station state (m) (N) (W) 2000 2001 2002 gauge type<br />

Pullm<strong>an</strong> WA 778 46Ε45' 117Ε7' Dec–J<strong>an</strong> Nov–Feb Dec–Feb SRN<br />

Rawlins WY 2053 41Ε48' 107Ε12' no data Nov–Mar Nov–Mar SRN<br />

Leadville CO 3029 39Ε14' 106Ε19' Dec–Apr Nov–Apr Nov–Apr SRN<br />

Rhinel<strong>an</strong>der WI 487 45Ε38' 89Ε28' Dec–Feb Nov–Mar Dec–Apr SRN<br />

Syracuse NY 125 43Ε7' 76Ε6' J<strong>an</strong>–Feb Dec–Mar Dec–Feb BUG<br />

Table 2. The average (me<strong>an</strong>) <strong>an</strong>d coefficient of variation (COV) of <strong>the</strong> station meteorology for <strong>the</strong><br />

winter periods listed in Table 1. Note: † precipitation is corrected using daily data.<br />

precipitation temperature humidity<br />

vapour<br />

pressure<br />

(mm)† (ΕC) (mb) deficit (mb) wind (m/s)<br />

station me<strong>an</strong> COV me<strong>an</strong> COV me<strong>an</strong> COV me<strong>an</strong> COV me<strong>an</strong> COV<br />

Pullm<strong>an</strong>, WA 37.3 0.37 –0.63 –1.57 4.94 0.06 1.04 0.301 4.1 0.176<br />

Rawlins, WY 19.2 0.552 –4.71 –0.62 3.12 0.218 1.36 0.294 5.6 0.170<br />

Leadville, CO 27 0.552 –5.24 –0.74 2.44 0.250 1.76 0.449 3.6 0.111<br />

Rhinel<strong>an</strong>der, WI 26.7 0.749 –6.64 –0.51 3.19 0.251 0.97 0.278 3.4 0.097<br />

Syracuse, NY 89.9 0.473 –1.66 –1.69 4.32 0.183 1.48 0.250 4.3 0.086<br />

METHODOLOGY<br />

To address <strong>the</strong> objectives of this paper, precipitation gauge undercatch w<strong>as</strong> estimated using<br />

daily <strong>an</strong>d monthly data. The monthly me<strong>an</strong> of daily data produced <strong>the</strong> average monthly data.<br />

Precipitation w<strong>as</strong> summed for each month. To generate undercatch estimates from <strong>the</strong> monthly<br />

ph<strong>as</strong>e partitioned data, daily data were used to identify <strong>the</strong> form of <strong>the</strong> precipitation, <strong>an</strong>d yielded a<br />

fraction of <strong>the</strong> monthly precipitation. The average wind speed during each precipitation type w<strong>as</strong><br />

used toge<strong>the</strong>r with <strong>the</strong> fraction of <strong>the</strong> monthly precipitation type. <strong>Snow</strong>pack sublimation <strong>an</strong>d<br />

blowing snow tr<strong>an</strong>sport were estimated from data at hourly, daily, <strong>an</strong>d average monthly, <strong>as</strong><br />

detailed in F<strong>as</strong>snacht (2004).<br />

The amount of gauge undercatch due to wind w<strong>as</strong> computed <strong>as</strong> a function of me<strong>as</strong>ured<br />

precipitation <strong>an</strong>d wind speed (Uz) (Y<strong>an</strong>g et al., 1998b):<br />

U<br />

( P U )<br />

P = f , . (2)<br />

g<br />

z<br />

221


The height of <strong>the</strong> <strong>an</strong>emometer at each station w<strong>as</strong> <strong>as</strong>sumed to be at 6m, <strong>an</strong>d <strong>the</strong> height of <strong>the</strong><br />

gauge orifice w<strong>as</strong> <strong>as</strong>sumed to be at 2m. The wind speed w<strong>as</strong> converted from a 6m height to a 2 m<br />

height, <strong>as</strong> per Goodison et al. (1998) using <strong>the</strong> snowpack aerodynamic roughness of 0.005 m<br />

(F<strong>as</strong>snacht et al., 1999).<br />

For solid precipitation, wetting losses were <strong>as</strong>sumed to be small <strong>as</strong> compared to wind induced<br />

losses, <strong>an</strong>d for monthly computations were <strong>as</strong>sumed to be minimal. Similarly, evaporation losses<br />

were <strong>as</strong>sumed to be negligible (Goodison et al., 1998). Undercatch equations were derived from<br />

daily data for snowfall when <strong>the</strong> daily air temperature (Ta) w<strong>as</strong> colder th<strong>an</strong> freezing <strong>an</strong>d for mixed<br />

precipitation when Ta w<strong>as</strong> between 0 <strong>an</strong>d 3 degrees C (Goodison et al., 1998). The specific<br />

equations derived by Y<strong>an</strong>g et al. (1998b) for <strong>the</strong> unshielded 8" NWS St<strong>an</strong>dard precipitation gauge<br />

with respect to <strong>the</strong> Double Fence Intercomparison Reference gauge (DFIR) were used for all sites<br />

except Syracuse (Table 1). This station used a Belfort Universal gauge. As per Groism<strong>an</strong> et al.<br />

(1999), <strong>the</strong> Y<strong>an</strong>g et al. (1998b) equation w<strong>as</strong> used for snowfall undercatch of <strong>the</strong> Belfort Universal<br />

gauge. Mixed precipitation undercatch w<strong>as</strong> incre<strong>as</strong>ed by 7% for <strong>the</strong> Belfort Universal gauge used<br />

at Syracuse (Groism<strong>an</strong> et al., 1999). As per F<strong>as</strong>snacht (2004) <strong>an</strong>d Bogart et al. (2006), a maximum<br />

wind speed of 6.5 m/s w<strong>as</strong> used in <strong>the</strong> undercatch equations due to <strong>the</strong> incre<strong>as</strong>ed uncertainty at <strong>the</strong><br />

higher wind speeds.<br />

Sublimation <strong>an</strong>d <strong>the</strong> occurrence of blowing snow are episodic <strong>an</strong>d were thus computed using<br />

hourly data. Sublimation w<strong>as</strong> estimated using <strong>the</strong> bulk tr<strong>an</strong>sfer approach for <strong>the</strong> latent heat flux <strong>as</strong><br />

a function of humidity (RH), air temperature, wind speed, <strong>an</strong>d station pressure (PR):<br />

( RH,<br />

T , U , PR)<br />

FE = f a z , (3)<br />

<strong>as</strong> initially formulated by Sverdrup (1936). The occurrence of blowing snow (BSY/N) w<strong>as</strong><br />

initially estimated <strong>as</strong> a function of wind speed <strong>an</strong>d different temperature considerations, <strong>as</strong> per Li<br />

<strong>an</strong>d Pomeroy (1997):<br />

Y / N = . (4)<br />

( U T )<br />

BS f ,<br />

z<br />

a<br />

Once blowing snow w<strong>as</strong> determined to have initiated, <strong>the</strong> qu<strong>an</strong>tity of blowing snow w<strong>as</strong><br />

computed <strong>as</strong> a function of wind speed:<br />

BS<br />

( U )<br />

q = f , (5)<br />

z<br />

using <strong>the</strong> equation derived by Pomeroy et al. (1991). Sublimation <strong>an</strong>d blowing snow qu<strong>an</strong>tities<br />

were summed to yield net snowpack loss estimates.<br />

The NWS denotes trace events (PT) <strong>as</strong> precipitation amounts less th<strong>an</strong> 0.01 inches or 0.254 mm<br />

per hour or day (NWS, 2005), while Legates et al. (2005) called all me<strong>as</strong>urements less th<strong>an</strong> half<br />

<strong>the</strong> me<strong>as</strong>urable precipitation depth (0.005 inches or 0.127 mm) <strong>as</strong> a trace event. In this paper, daily<br />

trace events will be <strong>as</strong>signed a value of 0.127 mm.<br />

RESULTS<br />

Hourly, daily <strong>an</strong>d monthly meteorological data were used to compute PU (Figure 1) <strong>an</strong>d<br />

snowpack losses (FE plus qBS) (Figure 2). The estimated difference between losses <strong>an</strong>d undercatch<br />

using <strong>the</strong> time step specific to <strong>the</strong> derived equations versus using a monthly time step is presented<br />

in Figure 3 for monthly totals, <strong>an</strong>d Figure 4 for se<strong>as</strong>onal totals. For <strong>the</strong> difference comparison, <strong>the</strong><br />

same net result would appear along <strong>the</strong> 1:1 line, while data at <strong>the</strong> origin would indicate that<br />

me<strong>as</strong>ured precipitation could be used <strong>as</strong> <strong>an</strong> estimate of snow on <strong>the</strong> ground. Below <strong>the</strong> x-axis or to<br />

<strong>the</strong> left of <strong>the</strong> y-axis, more snow is actually accumulating th<strong>an</strong> estimated from <strong>the</strong> me<strong>as</strong>ured<br />

precipitation alone. The difference between <strong>the</strong> two time step estimates <strong>an</strong>d gauge precipitation for<br />

individual months is illustrated in Figure 5, <strong>an</strong>d for each se<strong>as</strong>on is illustrated in Figure 6.<br />

222


monthly probability<br />

adjusted data 40<br />

[% of δt<br />

cumulative]<br />

90<br />

80<br />

70<br />

60<br />

50<br />

30<br />

20<br />

10<br />

0<br />

100<br />

0<br />

0 10 20 30 40 50<br />

daily data<br />

60 70 80 90 100<br />

[% of δt<br />

cumulative]<br />

100<br />

223<br />

90<br />

80<br />

70<br />

60<br />

average monthly data<br />

[% of δt<br />

cumulative]<br />

50<br />

Pullm<strong>an</strong> WA<br />

Rawlins WY<br />

Leadville CO<br />

Rhinel<strong>an</strong>der WI<br />

Syracuse NY<br />

Figure 1. Comparison of monthly precipitation undercatch estimates using daily <strong>an</strong>d monthly data. Each<br />

estimate is presented <strong>as</strong> a percentage of <strong>the</strong> total of <strong>the</strong> three different dat<strong>as</strong>ets. The values for undercatch<br />

have been derived for daily data. The average monthly data were derived from <strong>the</strong> me<strong>an</strong> of <strong>the</strong> daily data,<br />

where<strong>as</strong> <strong>the</strong> monthly probability adjusted data were derived using <strong>the</strong> monthly precipitation distribution<br />

(snow, mixed precipitation, or rain) <strong>an</strong>d <strong>the</strong> average wind speed during each type of precipitation.<br />

average<br />

monthly data 40<br />

[% of δt<br />

cumulative]<br />

90<br />

80<br />

70<br />

60<br />

50<br />

30<br />

20<br />

10<br />

0<br />

100<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

hourly data<br />

[% of δt<br />

cumulative]<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

40<br />

30<br />

30<br />

20<br />

20<br />

10<br />

Pullm<strong>an</strong> WA<br />

Rawlins WY<br />

Leadville CO<br />

Rhinel<strong>an</strong>der WI<br />

Syracuse NY<br />

daily data<br />

[% of δt<br />

cumulative]<br />

Figure 2. Comparison of relative monthly snowpack loss (sublimation plus blowing snow) estimates using<br />

hourly, daily, <strong>an</strong>d monthly data. Each estimate is presented <strong>as</strong> a percentage of <strong>the</strong> total of <strong>the</strong> three different<br />

time step estimates. The values have been derived for hourly data.<br />

10


Pullm<strong>an</strong> WA<br />

Rawlins WY<br />

Leadville CO<br />

Rhinel<strong>an</strong>der WI<br />

Syracuse NY<br />

Syracuse<br />

(-94.4,-178.4)<br />

Figure 3. Monthly total comparison of <strong>the</strong> difference between snowpack losses (sublimation plus blowing<br />

snow) <strong>an</strong>d precipitation undercatch estimated using average monthly meteorological data versus <strong>the</strong> time step<br />

for which <strong>the</strong> values were derived (hourly for losses <strong>an</strong>d daily for undercatch). The d<strong>as</strong>hed line represents <strong>the</strong><br />

1:1 relationship.<br />

Pullm<strong>an</strong> WA<br />

Rawlins WY<br />

Leadville CO<br />

Rhinel<strong>an</strong>der WI<br />

Syracuse NY<br />

Syracuse<br />

(-191,-337)<br />

Figure 4. Se<strong>as</strong>onal total comparison of <strong>the</strong> difference between snowpack losses (sublimation plus blowing<br />

snow) <strong>an</strong>d precipitation undercatch estimated using average monthly meteorological data versus <strong>the</strong> time step<br />

for which <strong>the</strong> values were derived (hourly for losses <strong>an</strong>d daily for undercatch). The d<strong>as</strong>hed line represents <strong>the</strong><br />

1:1 relationship.<br />

224


225<br />

Pullm<strong>an</strong> WA<br />

Rawlins WY<br />

Leadville CO<br />

Rhinel<strong>an</strong>der WI<br />

Syracuse NY<br />

Figure 5. Monthly total difference between monthly derived <strong>an</strong>d equation appropriate time step losses minus<br />

undercatch versus monthly gauge precipitation.<br />

Pullm<strong>an</strong> WA<br />

Rawlins WY<br />

Leadville CO<br />

Rhinel<strong>an</strong>der WI<br />

Syracuse NY<br />

Figure 6. Se<strong>as</strong>onal total difference between monthly derived <strong>an</strong>d equation appropriate time step losses minus<br />

undercatch versus monthly gauge precipitation.<br />

With <strong>the</strong> exception of J<strong>an</strong>uary 2002 at Pullm<strong>an</strong> WA, March 2002 at Rawlins WY, <strong>an</strong>d<br />

December 2001 at Syracuse NY, undercatch estimates from daily data were at le<strong>as</strong>t comparable to<br />

those from average monthly <strong>an</strong>d from monthly partitioned data (Figure 1). Undercatch estimated<br />

from monthly average data w<strong>as</strong> representative of those estimated from daily data, while <strong>the</strong><br />

monthly partitioned data w<strong>as</strong> more representative. The results from both monthly dat<strong>as</strong>et were<br />

most similar for Leadville CO, Rhinel<strong>an</strong>der WI <strong>an</strong>d Syracuse NY, yet for Pullm<strong>an</strong> WA <strong>an</strong>d<br />

Rawlins WY <strong>the</strong> monthly partitioned data yielded undercatch estimates more similar to using <strong>the</strong><br />

daily data th<strong>an</strong> using <strong>the</strong> average monthly data. The me<strong>an</strong> monthly undercatch estimate for all<br />

stations <strong>an</strong>d months using daily data w<strong>as</strong> 29.3 mm (st<strong>an</strong>dard deviation, s of 27.4, <strong>an</strong>d a maximum<br />

of 127), using average monthly data w<strong>as</strong> 35.8 mm (s of 40.6 <strong>an</strong>d a maximum of 229), <strong>an</strong>d using<br />

<strong>the</strong> monthly partitioned data w<strong>as</strong> 31.5 mm (s of 31.5 <strong>an</strong>d a maximum of 168).<br />

<strong>Snow</strong>pack loss estimates were more similar for <strong>the</strong> different time steps (Figure 2), except for <strong>the</strong><br />

25% of <strong>the</strong> months when monthly data yielded no loss estimates. For 13 months, monthly derived


estimates were equal to <strong>the</strong> sum of hourly <strong>an</strong>d daily derived estimates, primarily due to very small<br />

hourly estimates. The average monthly snowpack losses were 22.7 mm (s of 12.3 <strong>an</strong>d a maximum<br />

of 54.9), 24.3 mm (s of 13.4 <strong>an</strong>d a maximum of 57.1), <strong>an</strong>d 27.5 mm (s of 22.8 <strong>an</strong>d a maximum of<br />

113) for hourly, daily <strong>an</strong>d monthly time steps.<br />

The net difference between losses <strong>an</strong>d undercatch w<strong>as</strong> greater or equal for Rhinel<strong>an</strong>der WI<br />

when <strong>the</strong> monthly time step w<strong>as</strong> used compared to <strong>the</strong> appropriate time step, where<strong>as</strong> <strong>the</strong>y tended<br />

to be equal or less for Syracuse NY (Figure 3 for monthly totals <strong>an</strong>d Figure 4 for se<strong>as</strong>onal totals).<br />

As shown by F<strong>as</strong>snacht (2004), losses tended to be larger th<strong>an</strong> undercatch, i.e., only 4 months of<br />

data are in <strong>the</strong> third quadr<strong>an</strong>t of Figure 3, <strong>an</strong>d only 1 se<strong>as</strong>on in <strong>the</strong> second quadr<strong>an</strong>t of Figure 4.<br />

Data for Leadville CO <strong>an</strong>d to a lesser extent Rawlins WY plotted along or close to <strong>the</strong> 1:1 line in<br />

Figures 3 <strong>an</strong>d 4 (estimates were similar for both time steps). At Rawlins WY, some months had<br />

greater differences derived from <strong>the</strong> appropriate time step in <strong>the</strong> first quadr<strong>an</strong>t <strong>an</strong>d below <strong>the</strong> 1:1<br />

line.<br />

The difference between <strong>the</strong> y-axis (monthly time step) <strong>an</strong>d x-axis (equation appropriate time<br />

step) (Figure 3 <strong>an</strong>d 4 for monthly <strong>an</strong>d se<strong>as</strong>onal totals) is presented <strong>as</strong> a function of gauged<br />

precipitation for monthly <strong>an</strong>d se<strong>as</strong>onal time steps (Figures 5 <strong>an</strong>d 6, respectively). Points along <strong>the</strong><br />

x-axis in Figures 5 <strong>an</strong>d 6 correspond with points along <strong>the</strong> 1:1 line in Figures 3 <strong>an</strong>d 4. For monthly<br />

totals (Figure 5), <strong>the</strong>re is no systematic trend, except that <strong>the</strong> monthly difference may decre<strong>as</strong>e <strong>as</strong><br />

gauged precipitation incre<strong>as</strong>es. The Leadville CO data are clustered around <strong>the</strong> x-axis more<br />

closely th<strong>an</strong> o<strong>the</strong>r stations.<br />

DISCUSSION<br />

For some months at some stations estimates of month undercatch <strong>an</strong>d snowpack losses<br />

(snowpack sublimation <strong>an</strong>d blowing snow tr<strong>an</strong>sport) are similar using data at <strong>an</strong> hourly, a daily, or<br />

a monthly time step. The appropriate time step is daily for undercatch <strong>an</strong>d hourly for sublimation<br />

<strong>an</strong>d blowing snow. To estimate accumulated SWE from gauge precipitation (equation 1), data at a<br />

monthly time step could be used for <strong>the</strong> Leadville CO station, which is a dry environment (low<br />

humidity) with moderate precipitation <strong>an</strong>d wind (Table 2). Computationally this occurs in part<br />

since snowpack losses are precipitation limited for some months, i.e., FE plus qBS are equal to Pg<br />

plus PU for some months. However, using <strong>the</strong> bulk tr<strong>an</strong>sfer method h<strong>as</strong> been shown to<br />

overestimate sublimation (Hood et al., 1999). In particular, <strong>the</strong> latent heat flux equation uses <strong>the</strong><br />

vapour pressure deficit to compute sublimation, which is drier environments c<strong>an</strong> be greater th<strong>an</strong><br />

<strong>the</strong> available energy flux. As well, blowing snow estimates are likely larger th<strong>an</strong> actual.<br />

Information on <strong>the</strong> snowpack history may <strong>as</strong>sist in improving blowing snow estimates. The<br />

occurrence of blowing snow equations from Li <strong>an</strong>d Pomeroy (1997) are b<strong>as</strong>ed on <strong>the</strong> initiation of<br />

blowing snow on <strong>an</strong> hourly b<strong>as</strong>is. Finer temporal resolution data (<strong>an</strong>d observations) could improve<br />

blowing snow estimates. However, archived meteorological data are usually not available at<br />

shorter time intervals th<strong>an</strong> hourly.<br />

Both me<strong>an</strong> monthly undercatch <strong>an</strong>d snowpack loss estimates incre<strong>as</strong>ed <strong>as</strong> <strong>the</strong> temporal<br />

resolution of <strong>the</strong> data decre<strong>as</strong>ed, since <strong>the</strong>re is more variation (s is larger) due to more large<br />

values. In particular, <strong>the</strong>re were subst<strong>an</strong>tially larger monthly estimates from average monthly data<br />

for Syracuse NY, which had a persistent wind <strong>an</strong>d various large precipitation events. Removal of<br />

<strong>the</strong> March 2001 Syracuse NY (six large events) estimate reduced <strong>the</strong> me<strong>an</strong> monthly average<br />

undercatch of <strong>the</strong> remaining 69 station-months by 1.6, 3.4, <strong>an</strong>d 2.3 mm using <strong>the</strong> daily, average<br />

monthly <strong>an</strong>d monthly partitioned data. This w<strong>as</strong> also <strong>the</strong> month with <strong>the</strong> most gauged precipitation<br />

(138 mm), <strong>as</strong> illustrated in Figure 5. The monthly me<strong>an</strong> undercatch estimates for <strong>the</strong> o<strong>the</strong>r four<br />

stations were 22.9, 24.9, <strong>an</strong>d 24.5 mm.<br />

Undercatch estimates for December 2000 at Syracuse NY were similar for <strong>the</strong> different time<br />

intervals, but it should be noted that snow occurred on 27 days of <strong>the</strong> month with <strong>an</strong> average<br />

monthly wind speed of 5.3 m s –1 . For five days with snow, <strong>the</strong> daily average wind speed w<strong>as</strong><br />

greater th<strong>an</strong> 7 m s –1 . This w<strong>as</strong> <strong>the</strong> only month where <strong>the</strong> monthly difference w<strong>as</strong> greater th<strong>an</strong> <strong>the</strong><br />

equation difference (Figure 5).<br />

226


The daily wind speed c<strong>an</strong> exceed 6.5 m s –1 (achieved at all stations), which would result in <strong>an</strong><br />

undercatch ratio in <strong>the</strong> order of 20% (collecting one-fifth of <strong>the</strong> actual snowfall). The undercatch<br />

equation is b<strong>as</strong>ed on data from a number of stations. To improve this <strong>an</strong>d <strong>the</strong> estimates of<br />

sublimation <strong>an</strong>d blowing snow requires field observations. The precipitation undercatch computed<br />

in this paper is for unshielded NWS gauges, except at Syracuse. The NWS unshielded gauge is<br />

typical, but results were similar for <strong>the</strong> Belfort Universal gauge (Groism<strong>an</strong> et al., 1999). Net<br />

precipitation incre<strong>as</strong>ed by 0.2 to 1% when <strong>the</strong> Groism<strong>an</strong> et al. (1999) incre<strong>as</strong>e to mixed<br />

precipitation w<strong>as</strong> considered for <strong>the</strong> Syracuse site.<br />

Using data at a monthly time step produced <strong>an</strong> average wind speed that w<strong>as</strong> more th<strong>an</strong> <strong>the</strong><br />

threshold for blowing snow for only one station for one month (J<strong>an</strong>uary 2002 at Rawlins WY with<br />

<strong>an</strong> average Uz of 7 m s –1 ). Using daily data, <strong>the</strong> blowing snow threshold wind speed w<strong>as</strong> only<br />

achieved 41% of <strong>the</strong> time. With hourly data, it w<strong>as</strong> achieved at le<strong>as</strong>t twice each month at each<br />

station. <strong>Snow</strong> blowing into <strong>the</strong> gauge is a consideration for certain gauge <strong>an</strong>d/or shield<br />

configuration, such <strong>as</strong> <strong>the</strong> Tretyakov (Goodison et al., 1998). However, this h<strong>as</strong> not been observed<br />

to be a problem for <strong>the</strong> NWS St<strong>an</strong>dard Rain gauge nor <strong>the</strong> Belfort Universal gauge (Y<strong>an</strong>g et al.,<br />

1998b).<br />

The appropriate time step versus monthly time step (Figures 3 <strong>an</strong>d 4) yielded similar estimates<br />

for most months at Leadville CO <strong>an</strong>d some months Pullm<strong>an</strong> WA, Rawlins WY, <strong>an</strong>d Rhinel<strong>an</strong>der<br />

WI. Additional winters of data should be examined at <strong>the</strong>se <strong>an</strong>d o<strong>the</strong>r stations with a variety of<br />

climates.<br />

At present, monthly or se<strong>as</strong>onal gauged precipitation c<strong>an</strong>not be used to estimate discrep<strong>an</strong>cies in<br />

computations from different time steps. With more months <strong>an</strong>d stations, it may be possible to<br />

identify a systematic trend, at le<strong>as</strong>t for se<strong>as</strong>onal data.<br />

Averaging of data to coarser temporal resolutions does not always produce smaller estimates of<br />

undercatch, while sublimation estimates tend to be smaller. This w<strong>as</strong> due in part to difference in<br />

wind speed during precipitation events compared to when no precipitation occurs, <strong>as</strong> well <strong>as</strong> <strong>the</strong><br />

nature of averaging, <strong>as</strong> indicated by comparing undercatch estimated with average monthly data<br />

versus monthly partitioned data.<br />

Gauge precipitation is not a useful indicator of <strong>the</strong> difference between monthly <strong>an</strong>d equation<br />

appropriate time steps losses minus undercatch, i.e., it c<strong>an</strong>not be used to systematically adjust<br />

monthly time step estimates of losses minus undercatch for monthly or se<strong>as</strong>onal totals (Figure 5<br />

<strong>an</strong>d 6). Thus, <strong>the</strong> appropriate time step should be used for most locations for <strong>the</strong> computations. At<br />

some stations for some months, <strong>the</strong>re is even a subst<strong>an</strong>tial difference between monthly totals<br />

estimated using hourly versus daily data (Figures 1 <strong>an</strong>d 2).<br />

Trace events have been illustrated to be import<strong>an</strong>t for determining monthly <strong>an</strong>d se<strong>as</strong>onal<br />

precipitation amounts (e.g., Y<strong>an</strong>g et al., 1998a). The amount of precipitation <strong>as</strong>signed to trace<br />

events is especially import<strong>an</strong>t for drier environments. Undercatch during trace events is a problem,<br />

<strong>as</strong> it may be uncertain how much snow is actually falling, <strong>an</strong>d thus what <strong>the</strong> actual undercatch<br />

ratio is. The frequency of occurrence of me<strong>as</strong>urable versus trace precipitation events could be used<br />

<strong>as</strong> a guide to highlight <strong>the</strong> import<strong>an</strong>ce of trace events.<br />

To qu<strong>an</strong>tify <strong>the</strong> monthly total precipitation, cumulative monthly precipitation should be<br />

me<strong>as</strong>ured at gauging sites, toge<strong>the</strong>r with daily (<strong>an</strong>d hourly) me<strong>as</strong>urements. While gauge<br />

evaporation w<strong>as</strong> not considered <strong>as</strong> it tends to be small for snow (Goodison et al., 1998), it c<strong>an</strong><br />

account for some non-wind undercatch. Finer resolution precipitation me<strong>as</strong>urements may need to<br />

qu<strong>an</strong>tify gauge evaporation <strong>an</strong>d wetting losses. The resolution of <strong>the</strong> non-recording gauge that is<br />

used at thous<strong>an</strong>ds of NWS Cooperative sites across <strong>the</strong> U.S. is 0.254 mm (0.01 inches) thus<br />

making <strong>the</strong> <strong>as</strong>signment of a value to trace events import<strong>an</strong>t. For automated sites, especially those<br />

at airport (presented in <strong>the</strong> paper), finer resolution gauges, such <strong>as</strong> <strong>the</strong> Geonor (2006) vibrating<br />

wire gauge, should be explored, especially for hourly me<strong>as</strong>urements.<br />

227


CONCLUSIONS<br />

Precipitation gauge undercatch of snowfall should be estimated using daily data, unless new<br />

relationships between undercatch, wind speed, etc., are developed at different temporal<br />

resolutions. The inaccuracies of using monthly averaged data are evident from adjusting <strong>the</strong><br />

average data for <strong>the</strong> monthly probability of each precipitation type <strong>an</strong>d considering wind speeds<br />

during each type. However, <strong>the</strong>se data must be derived from daily data. For snowpack losses<br />

(sublimation <strong>an</strong>d blowing snow), it is re<strong>as</strong>onable to use daily data in lieu of hourly data when<br />

computing monthly or se<strong>as</strong>onal values.<br />

Average monthly data yielded no blowing snow, expect for one month at Rawlins WY. No<br />

sublimation w<strong>as</strong> estimated 25% of <strong>the</strong> time using monthly data. For Leadville CO <strong>an</strong>d Rhinel<strong>an</strong>der<br />

WI, monthly sublimation plus blowing snow w<strong>as</strong> computed to be very similar using each of <strong>the</strong><br />

time steps, for most months.<br />

The net accumulation difference, i.e., sublimation plus blowing snow minus undercatch,<br />

computed from <strong>the</strong> time step of data for which <strong>the</strong> equations were formulated (hourly for<br />

sublimation <strong>an</strong>d blowing snow, daily for undercatch) c<strong>an</strong> be estimated from monthly time step<br />

data at Leadville CO <strong>an</strong>d for several months at Rhinel<strong>an</strong>der WI. Se<strong>as</strong>onally, monthly time step<br />

data overestimate for 2 of 3 se<strong>as</strong>ons at Rhinel<strong>an</strong>der WI. Underestimates occur at Rawlins WY, for<br />

1 se<strong>as</strong>on at Pullm<strong>an</strong> WA, <strong>an</strong>d 2 of 3 se<strong>as</strong>ons at Syracuse NY.<br />

Nei<strong>the</strong>r monthly nor se<strong>as</strong>onal gauged precipitation c<strong>an</strong> be used to estimate discrep<strong>an</strong>cies in<br />

computations from different time steps, <strong>as</strong> no systematic trend is obvious. More stations <strong>an</strong>d years<br />

are required, especially focusing on specific climates.<br />

ACKNOWLEDGEMENTS<br />

The insight <strong>an</strong>d thoughtful comments of two <strong>an</strong>onymous reviewers <strong>an</strong>d Special Editor Dr.<br />

Andrew Klein improved this paper <strong>an</strong>d are appreciated.<br />

REFERENCES<br />

Bogart TA, Legates DR, Y<strong>an</strong>g D. 2006. Sensitivity <strong>an</strong>d spatial variability of Arctic precipitation<br />

bi<strong>as</strong> adjustments. Oral presentation at <strong>the</strong> 63 rd Annual E<strong>as</strong>tern <strong>Snow</strong> Conference, Newark,<br />

Delaware, June 7–9, 2006.<br />

F<strong>as</strong>snacht SR, Soulis ED, Kouwen N. 1999. Shape characteristics of freshly fallen snowflakes <strong>an</strong>d<br />

<strong>the</strong>ir short-term ch<strong>an</strong>ges. Interactions between <strong>the</strong> Cryosphere, Climate <strong>an</strong>d Greenhouse G<strong>as</strong>es<br />

(Proceedings IUGG 99 Symposium HS2, Birmingham, July 1999), IAHS, 256: 111–122.<br />

F<strong>as</strong>snacht SR. 2004. Estimating Alter-shielded gauge snowfall undercatch, snowpack sublimation,<br />

<strong>an</strong>d blowing snow tr<strong>an</strong>sport at six sites in <strong>the</strong> coterminous United States. Hydrological<br />

Processes, 18(18): 3481–3492 [doi:10.1002/hyp.5806].<br />

Geonor. 2006. Geonor Meteorological Equipment. Geonor Industries, Millb<strong>an</strong>k PA, URL:<br />

, l<strong>as</strong>t accessed: June 20, 2006.<br />

Goodison BE, Louie PYT, Y<strong>an</strong>g D. 1998. WMO Solid Precipitation Me<strong>as</strong>urement<br />

Intercomparison Final Report. WMO Instruments <strong>an</strong>d Observing Methods Report No. 67,<br />

WMO/TD No. 872.<br />

Groism<strong>an</strong> PY, Peck EL, Quayle RG. 1999. Intercomparison of recording <strong>an</strong>d st<strong>an</strong>dard<br />

nonrecording U.S. gauges. Journal of Atmospheric <strong>an</strong>d Oce<strong>an</strong>ic Technology, 16: 602–609.<br />

Hood E, Williams M, Cline D. 1999. Sublimation from a se<strong>as</strong>onal snowpack at a continental, midlatitude<br />

alpine site. Hydrological Processes 13: 1781–1797.<br />

Legates DR, Y<strong>an</strong>g D, Quiring S, Freem<strong>an</strong> K, Bogart T. 2005. Bi<strong>as</strong> adjustment to arctic<br />

precipitation: a comparison of daily versus monthly bi<strong>as</strong> adjustments. 8th Conference on Polar<br />

Meteorology <strong>an</strong>d Oce<strong>an</strong>ography, S<strong>an</strong> Diego, CA, Americ<strong>an</strong> Meteorological Society, 6 pages.<br />

Li L, Pomeroy JW. 1997. Estimates of threshold wind speeds for snow tr<strong>an</strong>sport using<br />

meteorological data. Journal of Applied Meteorology 36: 205–213.<br />

228


NCDC. 2006. National Climate Data Center. National Climate Data Center, Ashville NC,<br />

, page l<strong>as</strong>t updated June 19, 2006.<br />

Pomeroy JW, Gray DM, L<strong>an</strong>dine PG. 1991. Modelling <strong>the</strong> tr<strong>an</strong>sport <strong>an</strong>d sublimation of blowing<br />

snow on <strong>the</strong> prairies. In Proceedings of <strong>the</strong> 48 th E<strong>as</strong>tern <strong>Snow</strong> Conference, Guelph, Ontario, p<br />

175–188.<br />

Sverdrup HU. 1936. The eddy conductivity of <strong>the</strong> air over a smooth snow field. Geofysiske<br />

Publik<strong>as</strong>joner 11(7): 69pp.<br />

Y<strong>an</strong>g D, Goodison BE, Ishida S. 1998a. Adjustment of daily precipitation data at 10 climate<br />

stations in Al<strong>as</strong>ka: Application of World Meteorological Org<strong>an</strong>ization intercomparison results.<br />

Water Resources Research 34(2): 241–256.<br />

Y<strong>an</strong>g D, Goodison BE, Metcalfe JR. 1998b. Accuracy of NWS 8" St<strong>an</strong>dard Nonrecording<br />

Precipitation Gauge: Results <strong>an</strong>d Application of WMO Intercomparison. Journal of<br />

Atmospheric & Oce<strong>an</strong>ic Technology 15: 54–68.<br />

229


This page is intentionally bl<strong>an</strong>k.<br />

230


ABSTRACT<br />

231<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Shaped Solution Domains for <strong>Snow</strong> Properties<br />

RAE A. MELLOH, 1 SALLY A. SHOOP, 2 AND BARRY A. COUTERMARSH 2<br />

The objective of this work w<strong>as</strong> to develop a method for distributing snow properties on a<br />

l<strong>an</strong>dscape that is less dependent on extensive m<strong>as</strong>s <strong>an</strong>d energy bal<strong>an</strong>ce modeling, yet provides a<br />

realistic distribution of snow depth <strong>an</strong>d density across <strong>the</strong> environmental gradients of elevation,<br />

slope, azimuth, <strong>an</strong>d forest type. In this paper, we present import<strong>an</strong>t progress toward development<br />

of such <strong>an</strong> approach. A shaped solution domain w<strong>as</strong> identified for snow depth, water equivalent,<br />

<strong>an</strong>d density. The pinched-cone shape describes <strong>the</strong> differentiation of <strong>the</strong> snow properties with<br />

slope <strong>an</strong>d azimuth <strong>an</strong>d is approximated by <strong>an</strong> <strong>an</strong>alytical equation with only two coefficients.<br />

Knowledge of <strong>the</strong> solution domain shape permits a few model runs or me<strong>as</strong>urements to be<br />

exploited to define a continuous solution for all slope–azimuth combinations. The shaped<br />

solutions morph over time <strong>an</strong>d environmental gradients.<br />

Keywords: Shaped solution domain, <strong>Snow</strong> cone, <strong>Snow</strong> distribution, <strong>Snow</strong> properties<br />

INTRODUCTION<br />

This research effort arose from a requirement for distributed snow depth <strong>an</strong>d density for use in<br />

high-resolution ground vehicle mobility <strong>as</strong>sessment models used by <strong>the</strong> U.S. Army (Shoop et al.<br />

2004, Richmond et al. 2005). The requirement w<strong>as</strong> for a realistic distribution of snow depth <strong>an</strong>d<br />

density across <strong>the</strong> environmental gradients of elevation, slope, azimuth, <strong>an</strong>d forest type. An<br />

approach that is widely applicable <strong>an</strong>d that decre<strong>as</strong>es <strong>the</strong> dependence on extensive snow property<br />

modeling is needed because <strong>the</strong> computational <strong>an</strong>d data storage <strong>as</strong>pects of <strong>the</strong> mobility model<br />

c<strong>an</strong>not be overburdened by snow process computations <strong>an</strong>d snow solution storage. Yet, <strong>the</strong><br />

distribution of snow properties across a l<strong>an</strong>dscape is a complex problem that requires distribution<br />

of m<strong>as</strong>s <strong>an</strong>d energy bal<strong>an</strong>ce in a high-resolution (30-m) l<strong>an</strong>dscape.<br />

One way to visualize this problem is to determine <strong>the</strong> likely snow depth at a remote point, given<br />

a snow depth at a point on a l<strong>an</strong>dscape (Fig. 1). The known point might be where we are st<strong>an</strong>ding,<br />

at a low elevation <strong>an</strong>d on a south-facing slope. The point in question may be across <strong>the</strong> next ridge<br />

on a north-facing slope <strong>an</strong>d at higher elevation. We know <strong>the</strong>re is likely to be more snow at <strong>the</strong><br />

remote point, but we do not know how much more. A method for making this extrapolation to <strong>an</strong><br />

unknown point (or every point in <strong>the</strong> l<strong>an</strong>dscape) without resorting to extensive calculations would<br />

be quite useful.<br />

1 Correspondence to Rae A. Melloh, U.S. Army Engineer Research <strong>an</strong>d Development Center,<br />

Cold Regions Research <strong>an</strong>d Engineering Laboratory, 72 Lyme Road, H<strong>an</strong>over, NH USA 03755-<br />

1290, e-mail rmelloh@crrel.usace.army.mil.<br />

2 U.S. Army Engineer Research <strong>an</strong>d Development Center, Cold Regions Research <strong>an</strong>d<br />

Engineering Laboratory, 72 Lyme Road, H<strong>an</strong>over, NH USA 03755-1290.


In this paper, we present import<strong>an</strong>t progress toward development of such <strong>an</strong> approach. We have<br />

found that <strong>the</strong> solution domains for snow depth, snow water equivalent, <strong>an</strong>d snow density have a<br />

predictable shape that c<strong>an</strong> be approximated by <strong>an</strong> <strong>an</strong>alytical equation with only two coefficients.<br />

Figure 1. <strong>Snow</strong> depth across a l<strong>an</strong>dscape showing shallow snow (dark) at south-facing point A<br />

<strong>an</strong>d deeper snow (light tone) on north-facing slope <strong>an</strong>d higher elevation point B.<br />

BACKGROUND AND APPROACH<br />

The approach taken w<strong>as</strong> to investigate snow properties <strong>as</strong> a function of slope, azimuth,<br />

elevation, <strong>an</strong>d forest c<strong>an</strong>opy tr<strong>an</strong>smissivity using a snow model that h<strong>as</strong> a proven record of<br />

accurately representing snowpack processes. The model solution domain w<strong>as</strong> <strong>an</strong>alyzed for<br />

structure in snow property differentiation without mapping to <strong>an</strong>y particular l<strong>an</strong>dscape.<br />

SNTHERM (Jord<strong>an</strong> 1991), SHAW (Flerchinger et al. 1994), UEB (Tarboton <strong>an</strong>d Luce 1996), <strong>an</strong>d<br />

ISNOBAL (Marks 1999) are examples of physically b<strong>as</strong>ed models that accurately simulate m<strong>as</strong>s<br />

<strong>an</strong>d energy tr<strong>an</strong>sfer in snowpacks. The complexity <strong>an</strong>d computational intensity of <strong>the</strong> models <strong>an</strong>d<br />

meteorological driver requirements usually restrict <strong>the</strong>ir use to point simulations or small research<br />

drainage b<strong>as</strong>ins (Marks 1999). With <strong>the</strong> exception of ISNOBAL, <strong>the</strong>se models are not intended for<br />

explicit mapping (pixel by pixel). We chose SNTHERM (Jord<strong>an</strong> 1991) because of p<strong>as</strong>t experience<br />

with <strong>the</strong> model (Melloh et al. 2004) <strong>an</strong>d used a recent update (SLTHERM) that includes m<strong>as</strong>s <strong>an</strong>d<br />

energy tr<strong>an</strong>sfer between <strong>the</strong> snow <strong>an</strong>d soil.<br />

We designed a method that would be broadly applicable to a hilly, forested l<strong>an</strong>dscape in New<br />

Hampshire or Vermont similar to <strong>the</strong> l<strong>an</strong>dscapes of <strong>the</strong> Sleepers River Research Watershed near<br />

D<strong>an</strong>ville, Vermont; Hubbard Brook Experimental Forest near Thornton, New Hampshire; <strong>an</strong>d <strong>the</strong><br />

Eth<strong>an</strong> Allen Firing R<strong>an</strong>ge near Jericho, Vermont. The first pl<strong>an</strong>ned application of <strong>the</strong> method<br />

developed here is for a high-resolution mobility model of <strong>the</strong> Eth<strong>an</strong> Allen Firing R<strong>an</strong>ge. Hubbard<br />

Brook, Sleepers River, <strong>an</strong>d Eth<strong>an</strong> Allen all have similar climate, topography, <strong>an</strong>d elevation r<strong>an</strong>ges<br />

from approximately 200 to 1000 m. The tree species are predomin<strong>an</strong>tly nor<strong>the</strong>rn deciduous<br />

hardwoods, including sugar maple (Acer sacharum), beech (Fagus gr<strong>an</strong>difoia) <strong>an</strong>d yellow birch<br />

(Betula allegheniensis). White <strong>as</strong>h (Fraximus americ<strong>an</strong>a) is found at middle <strong>an</strong>d lower elevations.<br />

Red spruce (Picea rubens), balsam fir (Abies balsamea), <strong>an</strong>d white birch (Betula papyrifera var.<br />

cordifolia) occur at <strong>the</strong> higher elevations <strong>an</strong>d on rock outcrops. Hemlock (Tsunga c<strong>an</strong>adensis) is<br />

found along <strong>the</strong> streams.<br />

Solar radiation, when calculated <strong>as</strong> a function of slope <strong>an</strong>d azimuth <strong>an</strong>d visualized in <strong>an</strong> array,<br />

displays a gradually ch<strong>an</strong>ging symmetrical pattern across <strong>the</strong> array <strong>an</strong>d with calendar progression<br />

(Fig. 2). It seemed plausible that snowmelt differentiation due to slope <strong>an</strong>d azimuth would also<br />

follow a tractable pattern. Slope–azimuth combinations were selected by examining plots of clearsky<br />

solar radiation over <strong>the</strong> snow se<strong>as</strong>on (Fig. 2). South- <strong>an</strong>d north-facing exposures were<br />

emph<strong>as</strong>ized over e<strong>as</strong>t- <strong>an</strong>d west-facing exposures in <strong>the</strong> selection of slope–azimuth combinations<br />

because <strong>the</strong> magnitude of daily solar radiation varies less with terrain slope for e<strong>as</strong>t <strong>an</strong>d west<br />

exposures. Modifications were made to <strong>the</strong> b<strong>as</strong>e meteorological data to drive <strong>the</strong> snowpack energy<br />

232


<strong>an</strong>d m<strong>as</strong>s bal<strong>an</strong>ces for 540 environments: <strong>the</strong> product of 27 slope–azimuth combinations, five<br />

forest c<strong>an</strong>opy tr<strong>an</strong>smissivities, <strong>an</strong>d four elevations.<br />

55o 55o 0o 0o Oct Oct<br />

Nov Nov<br />

Dec Dec<br />

N<br />

233<br />

J<strong>an</strong> J<strong>an</strong><br />

E<br />

Feb Feb<br />

Figure 2. (Top) Relative clear sky radiation for <strong>the</strong> months of October through February plotted for terrain<br />

slopes of 0° to 55° for each month, <strong>an</strong>d for all azimuths. (Bottom) SLTHERM solutions were obtained for <strong>the</strong><br />

center points of <strong>the</strong>se 27 slope–azimuth regions.<br />

METHODS<br />

Distributing <strong>the</strong> snowpack energy bal<strong>an</strong>ce<br />

Terrain elevation, slope, azimuth, <strong>an</strong>d forest c<strong>an</strong>opy greatly influence <strong>the</strong> m<strong>as</strong>s <strong>an</strong>d energy<br />

bal<strong>an</strong>ce of <strong>the</strong> snowpack <strong>an</strong>d <strong>the</strong> spatial distribution of snowpack properties. Our approach to<br />

distributing a m<strong>as</strong>s <strong>an</strong>d energy bal<strong>an</strong>ce in <strong>the</strong> hilly, forested terrain of New Engl<strong>an</strong>d w<strong>as</strong> to modify<br />

meteorological data me<strong>as</strong>ured at <strong>an</strong> open site so that it represents <strong>the</strong> meteorological gradients<br />

imposed by elevation, slope, azimuth, <strong>an</strong>d forest cover. The energy bal<strong>an</strong>ce of a snowpack may be<br />

expressed <strong>as</strong><br />

Q + Q = Q + Q + Q + Q + Q + Q<br />

(1)<br />

M Δ K L E H P G<br />

where<br />

QM = snowmelt,<br />

Q� = ch<strong>an</strong>ge in stored heat,<br />

QK = solar (or shortwave) radiation,<br />

QL = terrestrial (or longwave) radiation,<br />

QE = latent heat tr<strong>an</strong>sfer (evaporation <strong>an</strong>d condensation),<br />

S<br />

W<br />

N


QH = sensible heat tr<strong>an</strong>sfer,<br />

QP = heat advected by rainwater, <strong>an</strong>d<br />

QG = conduction of ground heat.<br />

South-facing slopes receive more solar radiation (Fig. 2). Forest c<strong>an</strong>opies tr<strong>an</strong>smit only a<br />

percentage of <strong>the</strong> above c<strong>an</strong>opy radiation to <strong>the</strong> snowpack; decre<strong>as</strong>e longwave losses under cold,<br />

clear sky conditions; emit longwave radiation; <strong>an</strong>d reduce wind speeds that reduce latent <strong>an</strong>d<br />

sensible heat tr<strong>an</strong>sfers. Precipitation <strong>an</strong>d temperature lapse rates combine to accumulate more<br />

snow at higher elevations.<br />

The meteorological data used were collected by <strong>the</strong> Forest Service at <strong>the</strong> Hubbard Brook<br />

Experimental Forest (HBEF) at 252-m elevation in <strong>an</strong> open field during <strong>the</strong> fall of 2002 <strong>an</strong>d<br />

winter of 2003 (USDA 2004). Meteorological variations were created from <strong>the</strong> original b<strong>as</strong>e<br />

meteorology for 540 combinations of 27 slope–azimuths, four elevations (300, 500, 700, <strong>an</strong>d 900<br />

m) <strong>an</strong>d five c<strong>an</strong>opy solar tr<strong>an</strong>smissivities (0.14, 0.3, 0.5, 0.75, <strong>an</strong>d 1.0). C<strong>an</strong>opy tr<strong>an</strong>smissivity<br />

(Tr) is a continuous variable ra<strong>the</strong>r th<strong>an</strong> discrete; however, most forest cover maps are categorical<br />

<strong>an</strong>d <strong>the</strong>se discrete values were chosen to represent conifer, mixed conifer-deciduous, deciduous,<br />

sparse c<strong>an</strong>opy, <strong>an</strong>d no c<strong>an</strong>opy, respectively.<br />

Solar <strong>an</strong>d terrestrial radiation<br />

Solar radiation on slopes w<strong>as</strong> calculated within SLTHERM using a cosine correction for<br />

illumination <strong>an</strong>gle. Subc<strong>an</strong>opy shortwave radiation (KSC) w<strong>as</strong> calculated from solar radiation<br />

me<strong>as</strong>ured in <strong>the</strong> open (KIN):<br />

KSC IN<br />

= K Tr . (2)<br />

Reflected solar radiation w<strong>as</strong> calculated using a const<strong>an</strong>t albedo of 0.78:<br />

K = 0.<br />

78 K . (3)<br />

SCout<br />

SC<br />

Terrestrial radiation under <strong>the</strong> forest c<strong>an</strong>opy (LINforest) w<strong>as</strong> calculated <strong>as</strong><br />

INforest<br />

4<br />

( − Tr)<br />

TempK ( Tr L )<br />

L = 1 ε +<br />

(4)<br />

IN<br />

where emissivity (ε) of <strong>the</strong> forest is taken <strong>as</strong> 0.96, <strong>the</strong> effective temperature of <strong>the</strong> forest is<br />

estimated by <strong>the</strong> air temperature (TempK), <strong>an</strong>d σ is <strong>the</strong> Steph<strong>an</strong>–Boltzm<strong>an</strong> const<strong>an</strong>t (5.67 ×<br />

10 –8 Wm –2 K –4 ). LIN is terrestrial radiation in <strong>the</strong> open (no c<strong>an</strong>opy) <strong>an</strong>d is calculated for three<br />

conditions: 100% clear (LINclear), 100% cloudy (LINcloud), <strong>an</strong>d partly cloudy (LINpart).<br />

L TempK<br />

4<br />

INcloud = εσ .<br />

LINclear INcloud<br />

= L − A + OFF . (6)<br />

0.5 0.5<br />

( VP VP )<br />

A=− 0.4842 [228 + 11.6 S − A ] . (7)<br />

LINpart INcloud<br />

( A − OFF)<br />

= L − CI<br />

. (8)<br />

Equations 5, 6, <strong>an</strong>d 7 were adapted from Anderson <strong>an</strong>d Baker (1967) with a locality offset<br />

(OFF) determined for a site near D<strong>an</strong>ville, Vermont (Melloh et al. 2004). SVP <strong>an</strong>d AVP are<br />

saturation <strong>an</strong>d actual vapor pressures. The clear-sky index (CI) is <strong>the</strong> ratio of observed solar<br />

radiation KIN to clear-sky radiation KCS calculated by methods presented by Dingm<strong>an</strong> (1993).<br />

234<br />

(5)


Temperature, precipitation, windspeed<br />

An average temperature lapse rate of –6.5 °C over 1000-m elevation rise w<strong>as</strong> adopted.<br />

Precipitation gage undercatch w<strong>as</strong> corrected in <strong>the</strong> b<strong>as</strong>e meteorology by adding between 0 to 41<br />

percent of <strong>the</strong> me<strong>as</strong>ured precipitation over <strong>the</strong> windspeed r<strong>an</strong>ge of 0 to greater th<strong>an</strong> 7 ms –1 for a<br />

Universal gage with <strong>an</strong> Alter shield (Goodison 1978). A precipitation multiplier (Pm) applied with<br />

incre<strong>as</strong>ed elevation (E)<br />

P<br />

M<br />

⎛E−252 ⎞<br />

= 1+ ⎜ ⎟0.75<br />

⎝ 1000 ⎠<br />

w<strong>as</strong> adapted from Dingm<strong>an</strong>’s (1988, 1993) study of me<strong>an</strong> <strong>an</strong>nual precipitation–elevation trends in<br />

New Hampshire <strong>an</strong>d Vermont. Precipitation interception loss (PI) w<strong>as</strong> expressed in terms of<br />

c<strong>an</strong>opy tr<strong>an</strong>smissivity (Tr) <strong>as</strong><br />

I<br />

( Tr)<br />

PM<br />

P = 1 −<br />

(10)<br />

<strong>an</strong>d PI w<strong>as</strong> limited to a maximum of<br />

PIm ax<br />

( ) 3<br />

1−<br />

= 0. 0005 Tr . (11)<br />

Precipitation type w<strong>as</strong> <strong>as</strong>sumed to be rain if <strong>the</strong> temperature w<strong>as</strong> above 275.5 K, snow below<br />

273.2 K <strong>an</strong>d a mix of snow <strong>an</strong>d rain between <strong>the</strong>se two temperatures (Jord<strong>an</strong> 1991). The initial<br />

snow-grain diameter w<strong>as</strong> set at 0.0000254 m. The results of using Equations 9, 10, <strong>an</strong>d 11 (Fig. 3)<br />

are incre<strong>as</strong>ed precipitation (rain <strong>an</strong>d snow) <strong>an</strong>d incre<strong>as</strong>ed snowfall with incre<strong>as</strong>ed elevation. There<br />

is also decre<strong>as</strong>ed precipitation <strong>an</strong>d snowfall under fuller c<strong>an</strong>opies. The larger difference between<br />

<strong>the</strong> 300- <strong>an</strong>d 500-m elevation snowfalls indicates a snowline occurs, at times, below <strong>the</strong> 500-m<br />

elevation (Fig. 3).<br />

A wind lapse rate w<strong>as</strong> set at 15% incre<strong>as</strong>e for each 200-m elevation gain. Wind modification by<br />

c<strong>an</strong>opy w<strong>as</strong> <strong>an</strong> adaptation of <strong>an</strong> equation suggested by Dunne <strong>an</strong>d Leopold (1978)<br />

( Tr)<br />

W = w [ 1−<br />

0.<br />

8 1−<br />

(12)<br />

where w is wind me<strong>as</strong>ured in <strong>the</strong> open <strong>an</strong>d W is wind in <strong>the</strong> subc<strong>an</strong>opy.<br />

Precipitation (m)<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

300m<br />

500m<br />

700m<br />

900m<br />

conifer mixed decid o<strong>the</strong>r open<br />

235<br />

<strong>Snow</strong>fall (m) (m)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

(9)<br />

conifer mixed decid o<strong>the</strong>r open<br />

Figure 3. Precipitation <strong>an</strong>d snowfall values are water equivalent summed over Juli<strong>an</strong> days 300 to 110. (Left)<br />

Incre<strong>as</strong>ed precipitation with elevation <strong>an</strong>d incre<strong>as</strong>ed interception with incre<strong>as</strong>ed c<strong>an</strong>opy fullness. (Right)<br />

Incre<strong>as</strong>ed snowfall with elevation, incre<strong>as</strong>ed interception with incre<strong>as</strong>ed c<strong>an</strong>opy fullness, <strong>an</strong>d snowline<br />

development below 500 m. Only snowfalls for temperatures below 273.2 K are included.


SLTHERM Model initiation<br />

<strong>Snow</strong>pack properties for <strong>the</strong> 540 meteorological sub-environments were modeled with<br />

SLTHERM. SLTHERM is a new version of SNTHERM (Jord<strong>an</strong> 1991) that h<strong>as</strong> enh<strong>an</strong>ced<br />

capability to model moisture <strong>an</strong>d energy tr<strong>an</strong>sfer between <strong>the</strong> snowpack <strong>an</strong>d soil. SNTHERM is a<br />

well-tested <strong>an</strong>d internationally known physically b<strong>as</strong>ed snow process model. The simulation<br />

period initiated on 27 October (Juli<strong>an</strong> day 300) <strong>an</strong>d continued through snow accumulation <strong>an</strong>d<br />

finally snowmelt. An initial snow temperature profile w<strong>as</strong> available from Sleepers River Research<br />

Watershed (SRRW) at 900-m elevation <strong>an</strong>d lapsed at a rate of 6.5 °C over 1000 m to <strong>the</strong> lower<br />

elevations.<br />

<strong>Snow</strong> depth (m)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

300 320 340 360 15 35 55 75 95 115<br />

Juli<strong>an</strong> day<br />

Figure 4. The modeled snow depth time-series for each of <strong>the</strong> 27 slope–azimuth c<strong>as</strong>es for sparse forest.<br />

The top line is <strong>the</strong> most north-facing, <strong>an</strong>d <strong>the</strong> bottom line <strong>the</strong> most south-facing modeled time series.<br />

RESULTS<br />

Shaped solution domains for snow depth<br />

The snow model solutions (Fig. 4) for sparse c<strong>an</strong>opy at 900-m elevation for each of <strong>the</strong> 27<br />

slope–azimuth c<strong>as</strong>es illustrate that <strong>the</strong> snow depths differentiate with time following snowfall<br />

events, <strong>an</strong>d that <strong>the</strong> differentiation becomes more pronounced in late winter <strong>an</strong>d spring <strong>as</strong> <strong>the</strong><br />

snowpack ablates. The slope–azimuth <strong>an</strong>d c<strong>an</strong>opy dependence of solar radiation is <strong>the</strong> driving<br />

force behind snow property differentiation. On day 60 <strong>the</strong> snowpack h<strong>as</strong> differentiated mildly with<br />

azimuth <strong>an</strong>d terrain slope (Figs. 4, 5), by day 82 additional snow h<strong>as</strong> fallen <strong>an</strong>d fur<strong>the</strong>r<br />

differentiation h<strong>as</strong> occurred, <strong>an</strong>d on day 110 <strong>the</strong> snowpack is shallow <strong>an</strong>d highly differentiated.<br />

The 540 model solutions generated output for 20 time-series plots such <strong>as</strong> Figure 4, one for each<br />

of <strong>the</strong> five forest <strong>an</strong>d four elevation combinations.<br />

A three-dimensional (3-D) solution domain in <strong>the</strong> shape of a pinched cone is suggested by <strong>the</strong><br />

terrain slope versus snow-depth plot (Fig. 5) if one visualizes <strong>the</strong> bold lines (e<strong>as</strong>t-facing azimuths)<br />

to be in <strong>the</strong> foreground <strong>an</strong>d <strong>the</strong> d<strong>as</strong>hed lines (west-facing azimuths) to be in <strong>the</strong> background. The<br />

cone vertex position along <strong>the</strong> x-axis (Fig. 5) represents <strong>the</strong> snow depth for <strong>the</strong> no-slope c<strong>as</strong>e <strong>an</strong>d<br />

<strong>the</strong> vertex tr<strong>an</strong>slates to <strong>the</strong> right along <strong>the</strong> x-axis in response to new snow (day 60 to day 82) or to<br />

236


<strong>the</strong> left in response to snow compaction or ablation (day 82 to day 110). The incre<strong>as</strong>ed radius of<br />

<strong>the</strong> cone indicates incre<strong>as</strong>ed snow depth differentiation by slope–azimuth with incre<strong>as</strong>ed terrain<br />

slope. Elongation along <strong>the</strong> north-south axis of <strong>the</strong> cone (Fig. 6) occurs because snow depth<br />

differentiates more for north- to south-facing (x-axis) th<strong>an</strong> for e<strong>as</strong>t- to west-facing slopes (y-axis).<br />

A pinched-cone shape that approximates <strong>the</strong> solution domain (Fig. 6) is given by <strong>the</strong> equation<br />

2 2 2<br />

x + y x<br />

z = −<br />

(13)<br />

2<br />

2<br />

a b<br />

where x <strong>an</strong>d y are Cartesi<strong>an</strong> coordinates that determine <strong>the</strong> radius<br />

r +<br />

2 2<br />

= x y<br />

(14)<br />

<strong>an</strong>d where <strong>the</strong> radius represents <strong>the</strong> difference in <strong>the</strong> snow depth from <strong>the</strong> 0° terrain slope c<strong>as</strong>e.<br />

The slope azimuth (θ) direction is given by a radial coordinate system about <strong>the</strong> z-axis (Fig. 6).<br />

Coefficients a <strong>an</strong>d b satisfy <strong>the</strong> equations<br />

a =<br />

y<br />

z<br />

when x = 0 <strong>an</strong>d y = minimum radius (15)<br />

−1<br />

⎛1z⎞ b = ⎜ − ⎟ when y = 0 <strong>an</strong>d x= maximum radius.<br />

⎝a x ⎠<br />

(16)<br />

Terrain slope (degrees)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

18 o<br />

342 o<br />

Day 110<br />

54 o<br />

126 o<br />

234 o<br />

306 o<br />

90 o<br />

270 o<br />

162 o 198 o<br />

237<br />

Day 60<br />

Day 82<br />

0.15 0.25 0.35 0.45 0.55 0.65 0.75 0.85<br />

<strong>Snow</strong> depth (m)<br />

Figure 5. Modeled snow depths (x-axis) for days 60, 82, <strong>an</strong>d 110 plotted against terrain slope. The circles<br />

on day 110 indicate <strong>the</strong> snow model solutions. Bold lines are e<strong>as</strong>t-facing azimuths <strong>an</strong>d d<strong>as</strong>hed lines are<br />

west-facing azimuths. The 18° azimuth is <strong>the</strong> most south-facing <strong>an</strong>d 198° <strong>the</strong> most north-facing.


Terrain slope (degrees)<br />

-15<br />

-10<br />

S<br />

-5<br />

E<br />

0<br />

θo θo θo θo θo r<br />

ΔΔ ΔΔ Δ snow snow snow snow snow depth depth depth depth depth (x) (x) (x) (x) (x)<br />

5<br />

z<br />

238<br />

10<br />

W<br />

15<br />

5<br />

N<br />

0<br />

5<br />

10<br />

ΔΔ ΔΔ Δ snow snow snow snow snow depth depth depth depth depth (y) (y) (y) (y) (y)<br />

Figure6. The squ<strong>as</strong>hed-cone shaped solution in Cartesi<strong>an</strong> <strong>an</strong>d cylindrical coordinate systems,<br />

representing <strong>the</strong> snow-depth differentiation with slope <strong>an</strong>d azimuth.<br />

The SLTHERM model solutions plot along <strong>the</strong> z = 8.5° <strong>an</strong>d z = 20° contours (Fig. 7) of <strong>the</strong><br />

pinched cone <strong>an</strong>d verify that equation 13 approximates <strong>the</strong> shape of <strong>the</strong> solution domain. Southfacing<br />

Δ snow depths are negative <strong>an</strong>d north-facing Δ snow depths are positive (Fig. 6 <strong>an</strong>d 7) <strong>an</strong>d,<br />

<strong>as</strong> a result, <strong>the</strong> shape given by equation 13 must be corrected to p<strong>as</strong>s through r = 0 on <strong>the</strong><br />

minimum axis. The available SLTHERM solutions suggest that <strong>an</strong> 18° rotation corrects for e<strong>as</strong>twest<br />

<strong>as</strong>ymmetry. The long axis is along 18° <strong>an</strong>d 198° ra<strong>the</strong>r th<strong>an</strong> directly north-south <strong>an</strong>d <strong>the</strong><br />

minimum axis is along 108 o <strong>an</strong>d 288 o azimuths (ra<strong>the</strong>r th<strong>an</strong> directly e<strong>as</strong>t-west). The <strong>as</strong>ymmetry<br />

likely arises because west-facing slopes receive <strong>the</strong> afternoon sun through a more tr<strong>an</strong>smissive<br />

atmosphere th<strong>an</strong> e<strong>as</strong>t-facing slopes that receive <strong>the</strong> morning sun. A gradual correction to r=0 at <strong>the</strong><br />

108° <strong>an</strong>d 288° azimuths modifies <strong>the</strong> cone shape within ±18 o of <strong>the</strong> minimum axis (between 90 o<br />

<strong>an</strong>d 126 o , <strong>an</strong>d between 270 o <strong>an</strong>d 306 o ) where SLTHERM solutions were not computed (Fig. 7).


Pinched-cone dimensions for each slope–azimuth–forest environment were determined by finding<br />

<strong>the</strong> coefficients (a <strong>an</strong>d b) needed to match <strong>the</strong> SLTHERM solutions for each respective<br />

environment. Some of <strong>the</strong> SLTHERM model solutions do not lie <strong>as</strong> close to <strong>the</strong> contours <strong>as</strong> o<strong>the</strong>rs<br />

(Fig. 7). Additional model studies at additional azimuths are needed to refine <strong>the</strong> rotation <strong>an</strong>d<br />

fundamental shape if necessary, <strong>an</strong>d shorter time steps (< 1hr) may be necessary to obtain more<br />

consistent results. Equation 13 (with <strong>the</strong> r=0 at <strong>the</strong> minimum axis correction) is a re<strong>as</strong>onable first<br />

approximation of <strong>the</strong> solution domain shape. An equation that predicts r (Δ snow depth) <strong>as</strong> a<br />

function of z <strong>an</strong>d θ o will be developed for future applications (Melloh et al. in review).<br />

E<br />

306o 306o 234o 234o 342o 342o 198o 198o S<br />

N<br />

Figure 7. Contour plot of <strong>the</strong> cone at terrain slopes of z = 20° (outer contour) <strong>an</strong>d 8.5° (inner contour)<br />

with day 110 sparse forest SLTHERM solutions (circles) shown. Solid circles are for z = 20° <strong>an</strong>d open circles<br />

for z = 8.5° terrain slopes. The facing azimuths are indicated on <strong>the</strong> comp<strong>as</strong>s. The Δ snow depths (cm) are<br />

positive beneath <strong>an</strong>d negative above <strong>the</strong> e<strong>as</strong>t-west axis. An r=0 correction is sketched in at <strong>the</strong> minimum axis.<br />

Shaped solution tr<strong>an</strong>sitions over time <strong>an</strong>d c<strong>an</strong>opy gradients<br />

The shaped solutions for deciduous forest for days 60, 82, <strong>an</strong>d 110 (Fig. 8 top) present <strong>an</strong><br />

exp<strong>an</strong>ding cone over time <strong>as</strong> <strong>the</strong> snowpack ablates. The radii of <strong>the</strong> cone-shaped solutions contract<br />

progressively across <strong>the</strong> decre<strong>as</strong>ing c<strong>an</strong>opy tr<strong>an</strong>smitt<strong>an</strong>ce gradient of open to conifer (Fig. 8<br />

bottom). The cone-shaped solutions over <strong>the</strong> c<strong>an</strong>opy tr<strong>an</strong>sition nest neatly inside one <strong>an</strong>o<strong>the</strong>r when<br />

<strong>the</strong> vertices are co-located.<br />

239<br />

18o 18o 162o 162o 54o 54o 126o 126o 0 5 10 15<br />

15<br />

Δ snow depth (cm)<br />

W


deciduous<br />

Terrain slope (degrees)<br />

Terrain slope (degrees)<br />

20<br />

2.8<br />

66-cm 82-cm 41-cm<br />

1<br />

6.3<br />

day 60 day 82<br />

day 110<br />

240<br />

1.2<br />

9<br />

1.6<br />

open open sparse<br />

sparse deciduous deciduous mixed<br />

mixed conifer<br />

conifer<br />

9<br />

14<br />

4<br />

9<br />

1.6<br />

17<br />

17 32<br />

32<br />

41<br />

41 41<br />

41 40<br />

40<br />

<strong>Snow</strong>depth <strong>Snow</strong>depth (cm)<br />

(cm)<br />

Figure 8. (Top) <strong>Snow</strong> depth shaped solutions over time in <strong>the</strong> deciduous forest at 900-m elevation. The snow<br />

depths for 0° terrain slopes are shown above each cone for days 60, 82, <strong>an</strong>d 110. In both top <strong>an</strong>d bottom, <strong>the</strong><br />

arrows at <strong>the</strong> vertices represent <strong>the</strong> maximum (north-south) <strong>an</strong>d minimum (e<strong>as</strong>t-west) Δ snow depths (cm)<br />

relative to <strong>the</strong> 0° slope c<strong>as</strong>e. (Bottom) <strong>Snow</strong> depth shaped solutions over <strong>the</strong> forest environmental r<strong>an</strong>ge on<br />

day 110 at 900-m elevation. <strong>Snow</strong> depths for 0° slopes are listed on <strong>the</strong> x-axis. Corrections on minimum axes<br />

not shown.<br />

Shaped solutions for snow water equivalent <strong>an</strong>d density<br />

<strong>Snow</strong> water equivalent (SWE) variation with terrain slope <strong>an</strong>d azimuth (Fig. 9) displays<br />

similarly to snow depth variation (Fig. 5). The SWE solution domain c<strong>an</strong> also be approximated by<br />

<strong>the</strong> pinched-cone equation (Fig. 10). A snow density shaped solution w<strong>as</strong> calculated by dividing<br />

<strong>the</strong> axes dimensions of <strong>the</strong> snow water equivalent cone by those of <strong>the</strong> snow depth cone (Fig. 10).<br />

DISCUSSION<br />

Shaped solution domains for snow properties are a way to extend what we know about snow<br />

property differentiation with solar radiation <strong>an</strong>d forest cover. The snow cone is <strong>an</strong> interpolation<br />

surface; a few model runs or me<strong>as</strong>urements are exploited to define a continuous solution. The<br />

pinched-cone shape c<strong>an</strong> be defined by three model runs: 1) <strong>the</strong> no-slope c<strong>as</strong>e, 2) a slope–azimuth<br />

combination that defines <strong>the</strong> maximum diameter of <strong>the</strong> cone, <strong>an</strong>d 3) a slope–azimuth combination<br />

5<br />

1.5<br />

2<br />

1


Terrain slope (degrees)<br />

Terrain slope (degrees)<br />

a.<br />

14-cm SWE<br />

at 0<br />

b.<br />

o a.<br />

14-cm SWE<br />

at 0 slope<br />

b.<br />

o slope<br />

-10<br />

-5<br />

25<br />

20<br />

15<br />

10<br />

5.5<br />

5<br />

0<br />

S<br />

ΔΔ SWE SWE (x) (x)<br />

1.4<br />

0<br />

5<br />

10<br />

162<br />

18 342<br />

198<br />

N<br />

5<br />

0<br />

5<br />

10<br />

ΔΔ SWE SWE (y) (y)<br />

54 306 126<br />

241<br />

90<br />

Terrain Terrain slope (degrees)<br />

100<br />

270<br />

40<br />

450-kg m-3 density<br />

at 0o 450-kg m<br />

slope<br />

-3 density<br />

at 0o slope<br />

S<br />

10<br />

0<br />

ΔΔ Density Density (x) (x)<br />

234<br />

0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

<strong>Snow</strong> water equivalent (m)<br />

Figure 9. Modeled snow water equivalent (x-axis) for day 110 plotted against terrain slope.<br />

The circles indicate <strong>the</strong> SLTHERM solution points.<br />

-100<br />

N<br />

0<br />

ΔΔ Density Density (y) (y)<br />

Figure 10. (Left) Shaped solution for snow water equivalent for day 110 in <strong>the</strong> sparse forest at 900-m<br />

elevation. (Right) Shaped solution for snow density derived from <strong>the</strong> depth <strong>an</strong>d water equivalent cones.<br />

The arrows at <strong>the</strong> vertices represent <strong>the</strong> maximum (north-south) <strong>an</strong>d minimum (e<strong>as</strong>t-west) Δ SWE (cm)<br />

<strong>an</strong>d Δ density (kg m –3 ), respectively, relative to <strong>the</strong> 0° slope c<strong>as</strong>e. The arrows are not to scale. Corrections to<br />

r=0 on minimum axes are not shown.


that defines a second point on <strong>the</strong> cone. The three points defining <strong>the</strong> shaped solution do not have<br />

to be <strong>the</strong>se specific ordinates because <strong>an</strong>y three points on <strong>the</strong> cone will define <strong>the</strong> shape; however,<br />

points near <strong>the</strong>se extremes are recommended. The fundamental shape of <strong>the</strong> solution domain c<strong>an</strong><br />

be used to help pl<strong>an</strong> snow me<strong>as</strong>urement networks. In <strong>the</strong> absence of model solutions,<br />

me<strong>as</strong>urements at a few key locations in a b<strong>as</strong>in might be used to define a cone shape <strong>an</strong>d provide<br />

<strong>an</strong> estimate of <strong>the</strong> snow properties for <strong>an</strong>y slope–azimuth combination of similar c<strong>an</strong>opy <strong>an</strong>d<br />

elevation. Locations in contr<strong>as</strong>ting c<strong>an</strong>opy or elevation c<strong>an</strong> define <strong>the</strong> ch<strong>an</strong>ge in cone dimension<br />

with <strong>the</strong>se environmental gradients.<br />

Ma<strong>the</strong>matics on <strong>the</strong> solution domains provides <strong>an</strong>o<strong>the</strong>r interpolation route. Our first application<br />

of <strong>the</strong> shaped solution domains will be to distribute snow depth <strong>an</strong>d density across <strong>the</strong> l<strong>an</strong>dscape<br />

of <strong>the</strong> Eth<strong>an</strong> Allen Firing R<strong>an</strong>ge near Jericho, Vermont, for a vehicle mobility model. Interpolating<br />

between <strong>the</strong> shaped solution domains of <strong>the</strong> 20 slope–azimuth–c<strong>an</strong>opy–elevation environments<br />

permits creation of a continuous snow property map across <strong>the</strong> l<strong>an</strong>dscape if continuous maps of<br />

each of <strong>the</strong> environmental variables are available (slope, azimuth, c<strong>an</strong>opy tr<strong>an</strong>smissivity, <strong>an</strong>d<br />

elevation). The forest map at Eth<strong>an</strong> Allen Firing R<strong>an</strong>ge is categorical, thus our forest component<br />

interpolation will also be categorical. O<strong>the</strong>r examples of solution domain ma<strong>the</strong>matics include<br />

snow density computations from snow water equivalent <strong>an</strong>d snow depth cones, <strong>an</strong>d melt rate<br />

calculations by differencing snow water equivalent cones over time. Equation 13 (with a shape<br />

correction to r=0 at <strong>the</strong> minimum axis) is a good approximation of <strong>the</strong> fundamental shape for our<br />

intended application. Additional model studies at shorter time steps to improve model consistency<br />

<strong>an</strong>d potentially refine <strong>the</strong> azimuths of <strong>the</strong> major axes are warr<strong>an</strong>ted. An equation that predicts r (Δ<br />

snow depth) <strong>as</strong> a function of z <strong>an</strong>d θ o will be developed for future applications (Melloh et al. in<br />

review).<br />

The shaped solution domains were developed to map snow properties in a hilly forested terrain<br />

typical of New Hampshire <strong>an</strong>d Vermont <strong>an</strong>d over a relatively small b<strong>as</strong>in. The method is likely to<br />

be useful in o<strong>the</strong>r situations, though adaptations are needed to deal with large b<strong>as</strong>ins, topographic<br />

shading, <strong>an</strong>d blowing snow. The presented method will not be especially useful on plains <strong>an</strong>d<br />

prairies where <strong>the</strong> topography is flat, <strong>an</strong>d where <strong>the</strong>re is signific<strong>an</strong>t blowing snow. Using a cone<br />

method over larger b<strong>as</strong>ins will require adaptations to account for latitudinal <strong>an</strong>d longitudinal<br />

meteorological gradients. The cone does not account for topographic shading, <strong>an</strong> import<strong>an</strong>t<br />

consideration in mountainous b<strong>as</strong>ins. Also of interest is how <strong>the</strong> cone shape ch<strong>an</strong>ges with<br />

ch<strong>an</strong>ging maximum sun <strong>an</strong>gle at lower <strong>an</strong>d higher latitudes. It is also import<strong>an</strong>t to recognize that<br />

<strong>the</strong> cone solutions represent me<strong>an</strong> snow properties, <strong>an</strong>d on <strong>an</strong>y given slope in <strong>the</strong> field <strong>the</strong>re will<br />

be variability about <strong>the</strong> me<strong>an</strong> due to microclimate <strong>an</strong>d local deposition.<br />

CONCLUSION<br />

A shaped solution domain w<strong>as</strong> identified for snow depth, water equivalent, <strong>an</strong>d density. The<br />

pinched-cone shape describes <strong>the</strong> differentiation of <strong>the</strong> snow properties with slope <strong>an</strong>d azimuth<br />

relative to <strong>the</strong> no-slope c<strong>as</strong>e represented by <strong>the</strong> vertex of <strong>the</strong> cone. The shaped solution domain is<br />

approximated by <strong>an</strong> <strong>an</strong>alytical equation with only two coefficients. A time series of snow<br />

properties over a snow se<strong>as</strong>on c<strong>an</strong> be described efficiently by a time series of three numbers: <strong>the</strong><br />

snow property for <strong>the</strong> no-slope c<strong>as</strong>e, <strong>an</strong>d <strong>the</strong> two coefficients. Interpolating between <strong>the</strong> shaped<br />

solution domains of slope–azimuth–c<strong>an</strong>opy–elevation environments permits creation of a<br />

continuous snow property map across <strong>the</strong> l<strong>an</strong>dscape if continuous maps of each of <strong>the</strong><br />

environmental variables are available (slope, azimuth, c<strong>an</strong>opy tr<strong>an</strong>smissivity, <strong>an</strong>d elevation).<br />

Knowledge of <strong>the</strong> shape of <strong>the</strong> solution domain c<strong>an</strong> be exploited to portray snow property<br />

differentiation across environments <strong>an</strong>d may lead to less model-intensive approaches of estimating<br />

snow properties across a l<strong>an</strong>dscape.<br />

242


ACKNOWLEDGEMENTS<br />

This work w<strong>as</strong> funded by <strong>the</strong> U.S. Army AT42 High-Fidelity Ground Platform <strong>an</strong>d Terrain<br />

Mech<strong>an</strong>ics Modeling Program. The meteorological data used in this publication were obtained by<br />

scientists of <strong>the</strong> Hubbard Brook Ecosystem Study; this publication h<strong>as</strong> not been reviewed by those<br />

scientists. The Hubbard Brook Experimental Forest is operated <strong>an</strong>d maintained by <strong>the</strong><br />

Nor<strong>the</strong><strong>as</strong>tern Forest Experiment Station, U.S. Department of Agriculture, Radnor, Pennsylv<strong>an</strong>ia.<br />

We th<strong>an</strong>k Sus<strong>an</strong> Fr<strong>an</strong>kenstein <strong>an</strong>d Paul Richmond for <strong>the</strong>ir reviews of this m<strong>an</strong>uscript.<br />

REFERENCES<br />

Anderson EA, Baker DR. 1967. Estimating incident terrestrial radiation under all atmospheric<br />

conditions. Water Resources Research 3(4): 975–988.<br />

Dingm<strong>an</strong> SL. 1988. Application of kriging to estimating me<strong>an</strong> <strong>an</strong>nual precipitation in a region of<br />

orographic influence. Water Resources Bulletin 24(2): 329–339.<br />

Dingm<strong>an</strong> SL. 1993. Physical Hydrology. Englewood Cliffs, NJ: Prentice Hall, 575 pp.<br />

Dunne T, Leopold LB. 1978. Water in Environmental Pl<strong>an</strong>ning. S<strong>an</strong> Fr<strong>an</strong>cisco, CA: W.H.<br />

Freem<strong>an</strong> <strong>an</strong>d Co.<br />

Flerchinger GN, Cooley KR, Deng Y. 1994. Impacts of spatially <strong>an</strong>d temporally varying snowmelt<br />

on subsurface flow in a mountainous watershed: 1. <strong>Snow</strong>melt simulation. Hydrological<br />

Sciences 39(5): 507–520.<br />

Goodison BE. 1978. Accuracy of C<strong>an</strong>adi<strong>an</strong> snow gage me<strong>as</strong>urements. Journal of Applied<br />

Meteorology 27: 1542–1548.<br />

Jord<strong>an</strong> R. 1991. A one-dimensional temperature model for a snow cover: Technical documentation<br />

for SNTHERM.89. U.S. Army Cold Regions Research <strong>an</strong>d Engineering Laboratory, Special<br />

Report 91-16.<br />

Marks D, Domingo J, Susong D, Link T, Garen D. 1999. A spatially distributed energy bal<strong>an</strong>ce<br />

snowmelt model for application in mountain b<strong>as</strong>ins. Hydrological Processes 13: 1925–1959.<br />

Melloh RA, Richmond P, Shoop S, Coutermarsh B, Affleck R. in review. Continuous snow<br />

property mapping for mobility models. Cold Regions Science <strong>an</strong>d Technology.<br />

Melloh RA, Hall TJ, Bailey RN. 2004. Radiation data corrections for snow-covered sensors: Are<br />

<strong>the</strong>y needed for snowmelt modeling? Hydrological Processes 18: 1113–1126.<br />

Richmond, P., Reid A, Shoop S, M<strong>as</strong>on G. 2005. Terrain Surface Codes for <strong>an</strong> All-Se<strong>as</strong>on, Off-<br />

Road Ride Motion Simulator, MSIAC M&S Journal On-Line.<br />

Shoop S, Coutermarsh B, Reid A. 2004. All-Se<strong>as</strong>on Virtual Test Site for a Real-Time Vehicle<br />

Simulator, Paper No. 2004-01-2644, SAE Tr<strong>an</strong>sactions 113(2): 333–343.<br />

Tarboton DG, Luce CH. 1996. Utah Energy Bal<strong>an</strong>ce <strong>Snow</strong> Accumulation <strong>an</strong>d Melt Model (UEB)<br />

Computer Model Technical Description <strong>an</strong>d Users Guide. Utah Water Research Laboratory<br />

<strong>an</strong>d USDA Forest Service Intermountain Research Station: 64 p.<br />

U.S. Department of Agriculture, Forest Service, Nor<strong>the</strong><strong>as</strong>tern Research Station, Hubbard Brook.<br />

2003. Web page (http://www.hubbardbrook.org), GIS coverages, <strong>an</strong>d overview (guidebook).<br />

243


This page is intentionally bl<strong>an</strong>k.<br />

244


Glacial <strong>an</strong>d Periglacial<br />

Processes<br />

245


This page is intentionally bl<strong>an</strong>k.<br />

246


247<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

Qu<strong>an</strong>tifying <strong>the</strong> Effect of Anisotropic Properties in <strong>Snow</strong><br />

for Modelling Meltwater Retention<br />

INTRODUCTION<br />

C. E. BØGGILD 1<br />

In cold snowpacks of <strong>the</strong> Arctic meltwater retention is a signific<strong>an</strong>t factor for timing <strong>an</strong>d<br />

magnitude of runoff. Despite being <strong>an</strong> import<strong>an</strong>t component in Arctic hydrology <strong>an</strong>d glacier m<strong>as</strong>s<br />

bal<strong>an</strong>ce <strong>the</strong> meltwater infiltration <strong>an</strong>d subsequent re-freezing in cold snow is ra<strong>the</strong>r little qu<strong>an</strong>tified<br />

in literature. The qu<strong>an</strong>tity of super-imposed ice (SI) h<strong>as</strong> previously been modeled in onedimensional<br />

vertical profiles. But, since <strong>the</strong>se approaches rely on few points <strong>the</strong> spatial<br />

distribution of SI still needs to be determined more precisely. The problem of moving interface<br />

<strong>as</strong>sociated with SI growth calls for high precision in numerical models, which again greatly<br />

enh<strong>an</strong>ce <strong>the</strong> computational dem<strong>an</strong>d. This computational dem<strong>an</strong>d c<strong>an</strong> likely be <strong>the</strong> re<strong>as</strong>on for <strong>the</strong><br />

lack of SI treatment in most cryospheric models.<br />

Bøggild (submitted) proposes <strong>an</strong> <strong>an</strong>alytical solution to <strong>the</strong> ice warming, from which <strong>the</strong><br />

temporal SI-qu<strong>an</strong>tity c<strong>an</strong> be derived directly without comprehensive numerical modeling.<br />

However, this solution is only valid for iso<strong>the</strong>rmal <strong>an</strong>d isotropic condition, which likely never<br />

occurs in nature. Here is presented <strong>the</strong> result from statistical fitting of <strong>the</strong>rmal gradients <strong>an</strong>d<br />

<strong>the</strong>rmal diffusivity to <strong>the</strong> resulting effect on meltwater refreezing by SI formation. The purpose is<br />

to extend <strong>the</strong> existing <strong>an</strong>alytical solution to <strong>an</strong>isotropic condition also. And, since <strong>the</strong> <strong>the</strong>ory<br />

behind <strong>an</strong> <strong>an</strong>alytical solution is only valid for isotropic condition <strong>the</strong> approach here h<strong>as</strong> to rely on<br />

statistical methods.<br />

METHODS<br />

Bøggild et al (2005) found that for iso<strong>the</strong>rmal <strong>an</strong>d isotropic condition <strong>the</strong> growth rate of SI is:<br />

H ( t)<br />

=<br />

2γ<br />

αt<br />

where H is SI, t is time α <strong>an</strong>d γ are const<strong>an</strong>ts. With a temperature gradient <strong>an</strong>d <strong>an</strong>isotropic<br />

conditions SI is modelled by temporal heat diffusion modelling followed by:<br />

dH<br />

dt<br />

[ ρsi<br />

− ρws(<br />

1−<br />

ω ) ]<br />

L<br />

= xm<br />

∂T<br />

ciρi<br />

∫ dx<br />

∂t<br />

0<br />

1<br />

The University Centre in Svalbard (UNIS), PO Box 156, N-9171 Longyearbyen, Norway<br />

e-mail: carl.egede.boggild@unis.no


where ρ is density, ω is water content, L is Latent heat of fusion, k i is <strong>the</strong>rmal conduction of ice<br />

<strong>an</strong>d c i is specific heat of ice.<br />

A numerical model h<strong>as</strong> been developed b<strong>as</strong>ed on <strong>the</strong> above equation <strong>an</strong>d description. The model<br />

is used for <strong>an</strong>alyzing <strong>the</strong> effect of variable temperature gradients <strong>an</strong>d <strong>the</strong>rmal properties on SI<br />

formation. Sets of results have been produced, from which gradients c<strong>an</strong> be derived using linear<br />

regression.<br />

Effect of variable temperature gradient<br />

In <strong>the</strong> <strong>an</strong>alysis of <strong>the</strong> effect of temperature gradient in snow on <strong>the</strong> resulting superimposed ice<br />

formation <strong>the</strong> aim h<strong>as</strong> been twofold, namely 1) <strong>an</strong> expression valid for different ‘average’<br />

temperatures in <strong>the</strong> snow <strong>an</strong>d 2) derivation b<strong>as</strong>ed on realistically temperature gradients <strong>as</strong><br />

expected to occur in nature. B<strong>as</strong>ed on 1) <strong>an</strong>d 2) <strong>the</strong> experiments performed have first been carried<br />

out with <strong>an</strong> average temperature of –5 °C <strong>an</strong>d <strong>the</strong>n with <strong>an</strong> average temperature of –10 °C. As for<br />

2) <strong>the</strong> experiments have been b<strong>as</strong>ed on <strong>the</strong> following temperature gradients from <strong>the</strong> ice surface<br />

<strong>an</strong>d downward of –1, –0.5, –0.25, 0, +0.25, +0.5 °C per meter. Beyond this r<strong>an</strong>ge of gradients<br />

values are considered to be highly rare or non-existing in nature for longer time sp<strong>an</strong>s.<br />

Effect of variable <strong>the</strong>rmal diffusivity<br />

Since <strong>the</strong> highest <strong>the</strong>rmal diffusivity in a combined ice/snow system occurs with ice only, <strong>the</strong><br />

<strong>the</strong>rmal diffusivity h<strong>as</strong> been set <strong>as</strong> a fraction of <strong>the</strong> diffusivity of ice. These values have been<br />

inserted into <strong>the</strong> numerical model <strong>an</strong>d resulting output are presented in <strong>the</strong> next section.<br />

RESULTS<br />

A set of results w<strong>as</strong> produced using <strong>the</strong> a –5 °C ice/snow interface temperature but ch<strong>an</strong>ging <strong>the</strong><br />

temperature gradient inside <strong>the</strong> ice. This Approach w<strong>as</strong> repeated for <strong>an</strong> interface temperature of –<br />

10 °C (Fig. 1). And, it w<strong>as</strong> found that const<strong>an</strong>ts did show invariable with temperature <strong>as</strong> long, <strong>as</strong> <strong>the</strong><br />

temperature gradient w<strong>as</strong> accounted for in a separate term.<br />

Fitting of gradients<br />

0,00035<br />

0,0003<br />

0,00025<br />

0,0002<br />

0,00015<br />

0,0001<br />

0,00005<br />

0<br />

-12 -10 -8 -6 -4 -2 0<br />

Temperature r<strong>an</strong>ge<br />

Figure 1. The ch<strong>an</strong>ge in SI formation <strong>as</strong> a function of snow/ice-interface temperature <strong>an</strong>d r<strong>an</strong>ges of<br />

temperature gradients inside <strong>the</strong> ice.<br />

248


From linear regression <strong>the</strong> const<strong>an</strong>ts could be determined. The resulting expression with a<br />

variable temperature gradient is:<br />

Accounting for <strong>the</strong>rmal diffusivity did prove more complex because variable diffusivity does<br />

not relate linearly to H. Instead <strong>the</strong> logarithm of <strong>the</strong> ratio between <strong>the</strong> adjusted <strong>the</strong>rmal diffusivity<br />

to <strong>the</strong> diffusivity of ice(Diff/Diff ice ) did prove invariable with H. Fig 2 shows this relation for<br />

T ice = –10 °C.<br />

For T(ice)=-10 °C<br />

0,0007<br />

0,00065<br />

0,0006<br />

0,00055<br />

0,0005<br />

0,00045<br />

0,0004<br />

0,00035<br />

0,0003<br />

-1,5 -1 -0,5 0<br />

Figure 2. The ch<strong>an</strong>ge in SI formation <strong>as</strong> a function of <strong>the</strong> property Diff/Diffice The resulting expression with<br />

variable diffusivity becomes.<br />

When combining <strong>the</strong> contribution from temperature gradient <strong>an</strong>d <strong>the</strong>rmal diffusivity,<br />

respectively, <strong>the</strong> final equation becomes:<br />

CONCLUSIONS<br />

⎛ dT ⎞<br />

H ( t)<br />

= ⎜k<br />

2 − Tsk1⎟<br />

⎝ dx ⎠<br />

⎛ ⎛ ⎜ ⎛ ⎛<br />

f<br />

⎞<br />

H t = ⎜Ts⎜<br />

⎜ ⎜κ<br />

⎟<br />

( ) k1−<br />

k<br />

⎜ ⎜<br />

⎜ 3ln⎜<br />

⎟<br />

⎜ ⎜ ⎜ κi<br />

⎟<br />

⎝ ⎝ ⎝ ⎝ ⎠<br />

t<br />

⎞<br />

⎞<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎠ ⎠<br />

⎞<br />

⎟<br />

⎠<br />

⎛ ⎛ dT ⎜ ⎛ ⎛<br />

f<br />

H t = ⎜k<br />

Ts<br />

⎜<br />

− ⎜<br />

⎜ ⎜κ<br />

( ) 2<br />

k1−<br />

k ⎜<br />

⎜ 3ln⎜<br />

i<br />

⎝ dx ⎜ ⎜ ⎜ κ<br />

⎝ ⎝ ⎝<br />

A parameterization <strong>as</strong> addition to <strong>the</strong> Neum<strong>an</strong> solution h<strong>as</strong> been derived valid for temperature<br />

gradients r<strong>an</strong>ging from –1 to 0.5 K/m. It is believed that gradients exceeding this r<strong>an</strong>ge do not<br />

occur over longer time in nature. The const<strong>an</strong>ts were found to be robust over a temperature r<strong>an</strong>ge<br />

of at le<strong>as</strong>t 10 K.<br />

As a second addition to <strong>the</strong> Neum<strong>an</strong>n solution <strong>the</strong> effect of variable <strong>the</strong>rmal diffusivity is<br />

examined <strong>an</strong>d parameterized. Here a logarithmic term did prove to be <strong>the</strong> best solution. The<br />

motivation for deriving <strong>the</strong> diffusivity term is because <strong>the</strong> parameterization <strong>the</strong>n also becomes<br />

t<br />

249<br />

⎞<br />

⎞ ⎞⎟<br />

⎟ ⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟ ⎟<br />

⎟<br />

⎠<br />

⎟<br />

⎠⎠<br />

⎞<br />

⎟<br />

⎠<br />

t


valid for formation of ice lenses where SI is formed on top of e.g. firn with lover diffusivity. This<br />

lower diffusivity result in slower SI formation rates, <strong>as</strong> long <strong>as</strong> <strong>the</strong> temperature gradient is<br />

remaining.<br />

REFERENCES<br />

Bøggild, C.E. 1991: En smeltende snepakkes m<strong>as</strong>se- og energifluxe – belyst ved<br />

beregningsmetoder. Unpublished Thesis. Univ. Copenhagen. 94 pp.<br />

Bøggild, C.E., 2000 Preferential flow <strong>an</strong>d meltwater retention in cold snow packs in West-<br />

Greenl<strong>an</strong>d, Nordic Hydrology. 31 (4/5), 287–300.<br />

Bøggild, C.E., Forsberg, R., Reeh, N. 2005. Meltwater retention in a tr<strong>an</strong>sect across <strong>the</strong> Greenl<strong>an</strong>d<br />

ice sheet. Ann. Glac., 40: 102–105.<br />

250


251<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

The Equilibrium Flow <strong>an</strong>d M<strong>as</strong>s Bal<strong>an</strong>ce of <strong>the</strong> Taku Glacier,<br />

Al<strong>as</strong>ka, 1950–2005<br />

ABSTRACT<br />

M.S. PELTO 1 , , G.W. ADEMA 2 , M.J. BEEDLE 2 , S.R. MCGEE 2 ,<br />

M.M. MILLER 2 , K.F. SPRENKE 2 , AND M. LANG 3<br />

The Taku Glacier, Al<strong>as</strong>ka, h<strong>as</strong> adv<strong>an</strong>ced 7.5 km since <strong>the</strong> late nineteenth century, while all o<strong>the</strong>r<br />

primary outlet glaciers of <strong>the</strong> Juneau Icefield are in retreat. The Juneau Icefield Research Program<br />

(JIRP) h<strong>as</strong> completed field work on <strong>the</strong> Taku Glacier <strong>an</strong>nually since 1946. The thickest known<br />

alpine temperate glacier, it h<strong>as</strong> a maximum me<strong>as</strong>ured depth of 1480 m. The Taku is a tidewater<br />

glacier that formerly calved, but is now adv<strong>an</strong>cing slowly over its outw<strong>as</strong>h delta. Velocity<br />

me<strong>as</strong>ured over a twelve-month sp<strong>an</strong> <strong>an</strong>d <strong>an</strong>nual summer velocity me<strong>as</strong>urements completed at<br />

Profile 4 slightly above <strong>the</strong> ELA from 1950–2004 indicate insignific<strong>an</strong>t variations in velocity<br />

se<strong>as</strong>onally or from year to year . The consistency of velocity over <strong>the</strong> 50-year period indicates that<br />

in <strong>the</strong> vicinity of <strong>the</strong> equilibrium line, <strong>the</strong> flow of <strong>the</strong> Taku Glacier h<strong>as</strong> been in <strong>an</strong> equilibrium<br />

state.<br />

Surface m<strong>as</strong>s bal<strong>an</strong>ce w<strong>as</strong> positive from 1946–1988 averaging +0.42 m/a. This led to glacier<br />

thickening. From 1988–2005 <strong>an</strong> import<strong>an</strong>t ch<strong>an</strong>ge h<strong>as</strong> occurred <strong>the</strong> <strong>an</strong>nual bal<strong>an</strong>ce h<strong>as</strong> been<br />

–0.17m/a, <strong>an</strong>d <strong>the</strong> glacier thickness h<strong>as</strong> stopped incre<strong>as</strong>ing.<br />

Field me<strong>as</strong>urements of ice depth <strong>an</strong>d surface velocity allow calculation of <strong>the</strong> volume flux.<br />

Volume flux is <strong>the</strong>n compared with <strong>the</strong> surface bal<strong>an</strong>ce flux from <strong>the</strong> accumulation zone<br />

determined <strong>an</strong>nually in <strong>the</strong> field. On each profile <strong>the</strong> me<strong>an</strong> surface bal<strong>an</strong>ce flux 1946–2005 is<br />

positive versus <strong>the</strong> <strong>an</strong>nually determined volume flux, leading to glacier thickening. From 1950–<br />

2000, Taku Glacier thickened by 20–30 m at Profile 4, 33 km above <strong>the</strong> terminus. At this profile,<br />

<strong>the</strong> expected surface flux for equilibrium is 5.50 × 10 8 m 3 /a (±10%), while <strong>the</strong> calculated volume<br />

flux r<strong>an</strong>ge is 5.00–5.47 × 10 8 m 3 /a. At Profile 7, 43 km above <strong>the</strong> terminus, <strong>the</strong> observed surface<br />

flux above Profile 7 is 1.90 × 10 8 m 3 /a, <strong>an</strong>d <strong>the</strong> volume flux r<strong>an</strong>ge is 1.72–1.74 × 10 8 m 3 /a<br />

INTRODUCTION<br />

Taku Glacier is a temperate, maritime valley glacier in <strong>the</strong> Co<strong>as</strong>t Mountains of Al<strong>as</strong>ka. With <strong>an</strong><br />

area of 671 km 2 , it is <strong>the</strong> principal outlet glacier of <strong>the</strong> Juneau Icefield. It attracts special attention<br />

because of its continuing, century-long adv<strong>an</strong>ce (Motyka <strong>an</strong>d Post, 1995), while every o<strong>the</strong>r outlet<br />

glaciers of <strong>the</strong> Juneau Icefield is retreating. Taku Glacier is also noteworthy for its positive m<strong>as</strong>s<br />

bal<strong>an</strong>ce from 1946–1988 (Pelto <strong>an</strong>d Miller, 1990), during a period when alpine glacier m<strong>as</strong>s<br />

bal<strong>an</strong>ces have been domin<strong>an</strong>tly negative (Dyugerov <strong>an</strong>d Meier, 1997). Finally, it is unique <strong>as</strong> <strong>the</strong><br />

1<br />

Nichols College Dudley, MA 01571<br />

2<br />

Glaciological <strong>an</strong>d Arctic Sciences Institute, University of Idaho, Moscow ID 83843<br />

3<br />

Institute of Geodesy, Universitat der Bundeswehr, Werner Heisenberg Weg, D-85577<br />

Neubiberg, Germ<strong>an</strong>y


thickest alpine glacier yet me<strong>as</strong>ured, with a fjord extending 38–48 km upglacier from its terminus<br />

(Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs, 1995).<br />

The Juneau Icefield Research Program (JIRP) h<strong>as</strong> completed field work on <strong>the</strong> Taku Glacier<br />

<strong>an</strong>nually since 1946 (Miller, 1963; Pelto <strong>an</strong>d Miller, 1990). In this paper we present a data set for<br />

<strong>the</strong> Taku Glacier that is unique in its temporal <strong>an</strong>d spatial extent containing:<br />

1) Multi-year surface tr<strong>an</strong>sverse velocity data from three profiles, sp<strong>an</strong>ning 50 years on one<br />

profile; 2) Seismic profiling depth data on <strong>the</strong> same profiles; 3) Centerline longitudinal velocity<br />

tr<strong>an</strong>sects from <strong>the</strong> glacier divide to <strong>the</strong> ablation zone; <strong>an</strong>d 4) Multi-year surface m<strong>as</strong>s bal<strong>an</strong>ce data.<br />

From this data we c<strong>an</strong> determine surface bal<strong>an</strong>ce <strong>an</strong>d volume bal<strong>an</strong>ce tr<strong>an</strong>sfers. This provides a<br />

field-b<strong>as</strong>ed qu<strong>an</strong>titative determination of <strong>the</strong> volume flux at multiple locations <strong>an</strong>d hence<br />

constraints on <strong>the</strong> future behavior of <strong>the</strong> Taku Glacier.<br />

The glacier is divided into three zones that describe both m<strong>as</strong>s bal<strong>an</strong>ce <strong>an</strong>d flow dynamics: (1)<br />

The ablation zone, below <strong>the</strong> me<strong>an</strong> <strong>an</strong>nual ELA of 925 m (113 km 2 ), descends <strong>the</strong> trunk valley<br />

with no tributaries joining <strong>the</strong> glacier, <strong>an</strong>d only <strong>the</strong> distributary tongue, Hole in <strong>the</strong> Wall, leaving<br />

<strong>the</strong> glacier 11 km above <strong>the</strong> terminus. (2) The lower neve zone, extending from <strong>the</strong> ELA at 925 m<br />

to 1350 m, is a zone where summer ablation is signific<strong>an</strong>t (178 km 2 ). All <strong>the</strong> main tributaries<br />

(Southwest, West, Mat<strong>the</strong>s, Demorest, <strong>an</strong>d Hades Highway) join in this zone. (3) The upper neve<br />

zone extends from 1350 m to <strong>the</strong> head of <strong>the</strong> glacier (380 km 2 ), comprising <strong>the</strong> principal<br />

accumulation region for each tributary except <strong>the</strong> Southwest Br<strong>an</strong>ch. Ablation is limited in this<br />

zone, with much of <strong>the</strong> summer meltwater refreezing within <strong>the</strong> firnpack. This results in a unique<br />

signature in SAR imagery (Ramage et al., 2000).<br />

The Taku Glacier h<strong>as</strong> been adv<strong>an</strong>cing since 1890: It adv<strong>an</strong>ced 5.3 km between 1890 <strong>an</strong>d 1948<br />

(Moytka <strong>an</strong>d Post, 199; Pelto <strong>an</strong>d Miller, 1990). The glacier adv<strong>an</strong>ced 1.8 km from 1948–1988,<br />

<strong>an</strong>d 0.4 km from 1988 to 2003. The adv<strong>an</strong>ce rate me<strong>as</strong>ured by dist<strong>an</strong>ce is slowing. The rate of<br />

adv<strong>an</strong>ce is best <strong>as</strong>sessed in terms of area <strong>as</strong> <strong>the</strong> terminus lobe is spreading out on a terminus shoal.<br />

Motyka <strong>an</strong>d Post (1995) noted that <strong>the</strong> rate from 1948–1963 w<strong>as</strong> 0.428 km 2 /year, 0.345 km 2 /year<br />

from 1963–1979 <strong>an</strong>d 0.11 km 2 /year from 1979–1988. The slowing of <strong>the</strong> adv<strong>an</strong>ce h<strong>as</strong> been<br />

attributed to <strong>the</strong> imped<strong>an</strong>ce of <strong>the</strong> terminus outw<strong>as</strong>h plain shoal (Motyka <strong>an</strong>d Post, 1995), but it<br />

h<strong>as</strong> also been conjectured <strong>as</strong> due to <strong>the</strong> inability of <strong>the</strong> m<strong>as</strong>s bal<strong>an</strong>ce to sustain this adv<strong>an</strong>ce. With<br />

<strong>an</strong> AAR of 82, Taku Glacier had a continuously positive m<strong>as</strong>s bal<strong>an</strong>ce from 1946–1994, that h<strong>as</strong><br />

driven <strong>the</strong> continued adv<strong>an</strong>ce (Pelto <strong>an</strong>d Miller, 1990). From 1988–2005 m<strong>as</strong>s bal<strong>an</strong>ce h<strong>as</strong> been<br />

slightly negative.<br />

252


DATA COLLECTION<br />

Figure 1. Location map for Taku Glacier.<br />

Surface M<strong>as</strong>s Bal<strong>an</strong>ce<br />

JIRP h<strong>as</strong> me<strong>as</strong>ured <strong>the</strong> <strong>an</strong>nual bal<strong>an</strong>ce of <strong>the</strong> Taku Glacier from 1946 to 2005 (Pelto <strong>an</strong>d Miller,<br />

1990; Miller <strong>an</strong>d Pelto, 1999). Glacier <strong>an</strong>nual m<strong>as</strong>s bal<strong>an</strong>ce is <strong>the</strong> difference between <strong>the</strong> net snow<br />

accumulation <strong>an</strong>d net ablation over one hydrologic year. On non-calving glaciers, such <strong>as</strong> <strong>the</strong><br />

present Taku Glacier, surface m<strong>as</strong>s bal<strong>an</strong>ce observations are used to identify ch<strong>an</strong>ges in glacier<br />

volume. JIRP h<strong>as</strong> relied on applying consistent methods at st<strong>an</strong>dard me<strong>as</strong>urement sites (Pelto <strong>an</strong>d<br />

Miller, 1990; Miller <strong>an</strong>d Pelto, 1999). Taku Glacier m<strong>as</strong>s bal<strong>an</strong>ce me<strong>as</strong>urements are similar to<br />

those used on <strong>the</strong> Lemon Creek Glacier. The primary difference between <strong>the</strong> two is <strong>the</strong> much<br />

larger extent of <strong>the</strong> Taku Glacier (671 km 2 ). On <strong>the</strong> Taku Glacier, JIRP digs 17 test pits at fixed<br />

sites, monitors <strong>the</strong> migration of <strong>the</strong> tr<strong>an</strong>sient snow line, <strong>an</strong>d locates <strong>the</strong> final ELA position at <strong>the</strong><br />

end of <strong>the</strong> bal<strong>an</strong>ce year (Pelto <strong>an</strong>d Miller, 1990). These me<strong>as</strong>urements of retained accumulation<br />

are taken during late July <strong>an</strong>d early August <strong>an</strong>d must be adjusted to end of <strong>the</strong> bal<strong>an</strong>ce year values.<br />

253


This is done via daily ablation rates derived through stake me<strong>as</strong>urements <strong>an</strong>d migration of <strong>the</strong><br />

tr<strong>an</strong>sient snow line (Miller <strong>an</strong>d Pelto, 1999).<br />

Possible errors for <strong>the</strong> Taku Glacier m<strong>as</strong>s bal<strong>an</strong>ce record include sparse density of me<strong>as</strong>urement<br />

points (1 per 13 km 2 ), extrapolation to <strong>the</strong> end of <strong>the</strong> bal<strong>an</strong>ce year, infrequent me<strong>as</strong>urements of<br />

melting in <strong>the</strong> ablation zone, <strong>an</strong>d me<strong>as</strong>urements carried out by m<strong>an</strong>y different investigators.<br />

However, Pelto <strong>an</strong>d Miller (1990), suggest that <strong>the</strong>se sources of error are mitigated by <strong>an</strong>nual<br />

(since 1946) me<strong>as</strong>urements at 17 fixed locations, using nine years of ablation data to extrapolate<br />

m<strong>as</strong>s bal<strong>an</strong>ce in <strong>the</strong> ablation zone, using a bal<strong>an</strong>ce gradient derived from <strong>the</strong> 17 fixed sites <strong>an</strong>d<br />

known values for <strong>the</strong> ablation zone that shifts in altitude from year to year b<strong>as</strong>ed on <strong>the</strong> ELA, <strong>an</strong>d<br />

through supervision of field work by at le<strong>as</strong>t one experienced researcher (i.e., Matt Beedle from<br />

2003–2005).<br />

The me<strong>as</strong>urement network consists of 17 locations where m<strong>as</strong>s bal<strong>an</strong>ce h<strong>as</strong> been <strong>as</strong>sessed in test<br />

pits <strong>an</strong>nually since 1946. The majority of <strong>the</strong> snowpits are in <strong>the</strong> region from 950–1400 m. In<br />

1984, 1998 <strong>an</strong>d 2004, JIRP me<strong>as</strong>ured <strong>the</strong> m<strong>as</strong>s bal<strong>an</strong>ce at <strong>an</strong> additional 100–500 points with<br />

probing tr<strong>an</strong>sects in <strong>the</strong> accumulation area to better determine <strong>the</strong> distribution of accumulation<br />

around <strong>the</strong> snowpit locations. Me<strong>as</strong>urements were taken along profiles at 100–250 m intervals.<br />

The st<strong>an</strong>dard deviation for me<strong>as</strong>urements sites within 3 km, with less th<strong>an</strong> a 100 m elevation<br />

ch<strong>an</strong>ge, w<strong>as</strong> ±0.09 m/a; this indicates <strong>the</strong> consistency of m<strong>as</strong>s bal<strong>an</strong>ce around <strong>the</strong> snowpit sites.<br />

Ano<strong>the</strong>r possible source of error is <strong>the</strong> <strong>as</strong>sumption that <strong>the</strong> density me<strong>as</strong>ured at test pits is<br />

representative of a larger area. However, a study at 40 points within 1 km 2 at different elevations<br />

in different years resulted in a st<strong>an</strong>dard deviation of ±0.07 meters of water equivalent (m w.e.) in a<br />

snow pack of 1 to 2 m, displaying <strong>the</strong> highly uniform density of snow on <strong>the</strong> Taku Glacier (Pelto<br />

<strong>an</strong>d Miller, 1990).<br />

An independent check of m<strong>as</strong>s bal<strong>an</strong>ce is now available in <strong>the</strong> form of direct me<strong>as</strong>urement of<br />

<strong>the</strong> surface elevation of <strong>the</strong> glacier at specific points. The elevation h<strong>as</strong> been determined <strong>an</strong>nually<br />

since 1993 at fixed locations along Profile 4 using differential GPS <strong>as</strong> part of <strong>the</strong> velocity<br />

surveying program. GPS <strong>an</strong>nual elevation ch<strong>an</strong>ge me<strong>as</strong>urements along Profile 4 at 1100 m show a<br />

strong correlation with <strong>an</strong>nual m<strong>as</strong>s bal<strong>an</strong>ce me<strong>as</strong>urements. This would be expected <strong>as</strong> elevation at<br />

<strong>the</strong> me<strong>an</strong> ELA is likely to rise with incre<strong>as</strong>ed accumulation during years of positive m<strong>as</strong>s bal<strong>an</strong>ce,<br />

<strong>an</strong>d fall with incre<strong>as</strong>ed ablation during years of negative m<strong>as</strong>s bal<strong>an</strong>ce. For <strong>the</strong> period 1993 to<br />

2004, correlation between <strong>the</strong> average surface elevation ch<strong>an</strong>ge of a 31-point profile across <strong>the</strong><br />

Taku Glacier <strong>an</strong>d net m<strong>as</strong>s bal<strong>an</strong>ce is 0.77 (95% signific<strong>an</strong>ce). This provides <strong>an</strong> independent<br />

validation for Taku Glacier record (Table 1).<br />

GPS Survey Methods<br />

St<strong>an</strong>dard rapid-static <strong>an</strong>d real-time differential GPS methods have been employed for all survey<br />

work from 1996–2004. A key objective of <strong>the</strong> surveying program is to collect data that allows<br />

qu<strong>an</strong>titative comparison of surface movements <strong>an</strong>d surface elevation ch<strong>an</strong>ge from year to year. In<br />

order to ensure <strong>the</strong> consistency of year-to-year movement <strong>an</strong>d elevation data, all survey flags are<br />

located within one meter of <strong>the</strong> st<strong>an</strong>dard point coordinates (L<strong>an</strong>g, 1997; McGee, 2000). After<br />

establishment of each survey profile is complete, each profile is surveyed two times, with <strong>the</strong> time<br />

differential between <strong>the</strong> surveys r<strong>an</strong>ging from 6 to 9 days. For all surveys, a reference receiver is<br />

centered <strong>an</strong>d leveled over <strong>an</strong> appropriate bedrock benchmark. A roving receiver is mounted on <strong>an</strong><br />

aluminum monopole inserted into <strong>the</strong> same hole that <strong>the</strong> survey flag is placed. The height above<br />

<strong>the</strong> snow surface of <strong>the</strong> <strong>an</strong>tenna is noted. For rapid-static work, <strong>the</strong> roving receiver collected<br />

readings at 15-second intervals for 10 to 20 minutes at each flag. Real-time methods require only<br />

enough time at each flag sufficient to obtain a position fix from <strong>the</strong> reference receiver.<br />

254


Table 1. Comparison of GPS me<strong>as</strong>ured height ch<strong>an</strong>ge on Profile 4 <strong>an</strong>d <strong>the</strong> me<strong>an</strong> <strong>an</strong>nual bal<strong>an</strong>ce of <strong>the</strong><br />

Taku Glacier. The GPS data is strictly in height, <strong>an</strong>d <strong>the</strong> field me<strong>as</strong>urement in meters of water<br />

equivalent.<br />

Year Taku bn Pro 4 Ht. Ch.<br />

1994 0.09 0.09<br />

1995 –0.76 –1.33<br />

1996 –0.96 –0.67<br />

1997 –1.34 –0.46<br />

1998 –0.98 –1.06<br />

1999 0.4 0.59<br />

2000 1.03 1.42<br />

2001 0.88 0.9<br />

2002 0.45 –0.7<br />

2003 –0.9 –1.64<br />

2004 –0.23 0.63<br />

2005 0.02 –0.29<br />

Correlation = 0.77<br />

The major focus of <strong>the</strong> survey program is to continue <strong>the</strong> <strong>an</strong>nual survey of st<strong>an</strong>dard movement<br />

profiles on <strong>the</strong> Taku Glacier <strong>an</strong>d its main tributaries. In this study we focus on tr<strong>an</strong>sects 4 <strong>an</strong>d 7.<br />

Profile 4 h<strong>as</strong> <strong>an</strong> upper line <strong>an</strong>d a lower line separated by 0.5 km. In 1999, 2000 <strong>an</strong>d 2004, a<br />

longitudinal profile down <strong>the</strong> centerline of <strong>the</strong> Mat<strong>the</strong>s Br<strong>an</strong>ch <strong>an</strong>d <strong>the</strong> Taku Glacier, from <strong>the</strong><br />

glacier divide (located 65 km from <strong>the</strong> terminus), down to 24 km above <strong>the</strong> terminus. The surface<br />

velocity w<strong>as</strong> observed at survey locations spaced 0.5 km apart. The surface slope w<strong>as</strong> determined<br />

between each survey location.<br />

Seismic Methods<br />

Seismic methods are required to determine ice depths on tr<strong>an</strong>sverse profiles because of <strong>the</strong><br />

thickness of <strong>the</strong> Taku Glacier (Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs, 1995). The seismic program completed<br />

me<strong>as</strong>urements of ice thickness along eight tr<strong>an</strong>sects across <strong>the</strong> glacier, each following <strong>the</strong> same<br />

tr<strong>an</strong>sects <strong>an</strong>d using <strong>the</strong> same points used in <strong>the</strong> GPS movement <strong>an</strong>d elevation surveys.<br />

The seismic methods used for determining ice thickness are typical. A Bison 9024 series<br />

seismograph w<strong>as</strong> used with 24 high-frequency (100 Hz) geophones to record <strong>the</strong> seismic signals<br />

produced by explosive charges. The geophones were spaced at 30 meter intervals along profile<br />

lines that are perpendicular to <strong>the</strong> direction of glacier flow, covering 690 meters with each<br />

geophone spread. Explosive detonations (shots) were generally made at 500 meter intervals from<br />

each end of <strong>the</strong> geophone spread, to a maximum dist<strong>an</strong>ce of 2000 meters. Shot <strong>an</strong>d geophone<br />

locations were surveyed using st<strong>an</strong>dard differential GPS surveying techniques, accurate to ±5 cm.<br />

Up to twelve shots were taken on each profile, with up to four reflectors evident on each shot's<br />

record. The seismograph w<strong>as</strong> normally set to record two seconds of data, recording at a 0.25 ms<br />

sampling rate. The energy for a shot w<strong>as</strong> produced by 4 to 20 sticks of Kinepak (1/3 stick)<br />

explosive (ammonium nitrate <strong>an</strong>d petroleum distillate combination), buried approximately 1 meter<br />

deep in <strong>the</strong> firn.<br />

Reflections from <strong>the</strong> glacier bed were generally clear <strong>an</strong>d e<strong>as</strong>y to recognize on <strong>the</strong> records by<br />

<strong>the</strong>ir frequency, character, <strong>an</strong>d distinct moveout times. Migrations were completed using <strong>the</strong><br />

common-depth-point technique described in Dobrin (1960) <strong>an</strong>d adapted by Sprenke <strong>an</strong>d o<strong>the</strong>rs<br />

(1997). Calculations were made using a const<strong>an</strong>t ice velocity of 3660 m/s, a value determined<br />

from P-wave first arrival times. The migration <strong>an</strong>d geomorphic profiling process is b<strong>as</strong>ed on <strong>the</strong><br />

simplifying <strong>as</strong>sumption that <strong>the</strong> glacier cross-sections are two-dimensional. The results on Profile<br />

4 match closely <strong>the</strong> results of Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs (1995) with a maximum depth of 1450 m in this<br />

study versus 1400 m.<br />

255


DATA OBSERVATIONS<br />

M<strong>as</strong>s Bal<strong>an</strong>ce<br />

The <strong>an</strong>nual bal<strong>an</strong>ce record shows a markedly positive trend from 1946–1988 period. The me<strong>an</strong><br />

<strong>an</strong>nual bal<strong>an</strong>ce for <strong>the</strong> 42-year period is 0.42 m/a, representing a total thickening of 17.5 m w.e or<br />

20 m in total ice thickness. From 1988–2005 m<strong>as</strong>s bal<strong>an</strong>ce ch<strong>an</strong>ged signific<strong>an</strong>tly <strong>an</strong>d h<strong>as</strong> been<br />

slightly negative averaging –0.17 m/a, a 3 m w.e. loss, or 3.3 meters of glacier thickness ch<strong>an</strong>ge<br />

(Figure 2). A comparison of surface height ch<strong>an</strong>ge along Profile 4 from 1993–2004 <strong>an</strong>d me<strong>an</strong><br />

surface bal<strong>an</strong>ce yield –3.5 m <strong>an</strong>d –2.7 m respectively. The next t<strong>as</strong>k in this research program is to<br />

determine <strong>the</strong> height ch<strong>an</strong>ges at survey location along additional profiles over <strong>an</strong> extended period,<br />

thus providing additional surface bal<strong>an</strong>ce data each year.<br />

m we<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

1946<br />

1949<br />

1952<br />

1955<br />

1958<br />

1961<br />

Juneau Icefield M<strong>as</strong>s Bal<strong>an</strong>ce<br />

Taku Cumulative bn Taku bn<br />

1964<br />

1967<br />

1970<br />

1973<br />

1976<br />

Year<br />

Figure 2. Annual <strong>an</strong>d cumulative m<strong>as</strong>s bal<strong>an</strong>ce of <strong>the</strong> Taku Glacier 1946–2005.<br />

With <strong>the</strong> distribution of <strong>an</strong>nual bal<strong>an</strong>ce more accurately mapped in <strong>the</strong> lower neve section of <strong>the</strong><br />

glacier b<strong>as</strong>ed on <strong>the</strong> 1998, 2003 <strong>an</strong>d 2004 combined snowpit <strong>an</strong>d probing me<strong>as</strong>urements, me<strong>an</strong><br />

<strong>an</strong>nual surface flux w<strong>as</strong> determined both for <strong>the</strong> glacier region above Profile 4 <strong>an</strong>d Profile 7,<br />

summing <strong>the</strong> products of <strong>the</strong> area observed between each 0.2 m m<strong>as</strong>s bal<strong>an</strong>ce interval <strong>an</strong>d <strong>the</strong><br />

<strong>an</strong>nual bal<strong>an</strong>ce for that interval. The surface flux accumulating above at Profile 4 w<strong>as</strong> calculated to<br />

be 5.5 × 10 8 m 3 a –1 , <strong>an</strong>d 1.90 × 10 8 m 3 a –1 above Profile 7.<br />

256<br />

1979<br />

1982<br />

1985<br />

1988<br />

1991<br />

1994<br />

1997<br />

2000<br />

2003<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5<br />

-1<br />

-1.5


Tr<strong>an</strong>sverse Velocity Profiles<br />

Surface velocity h<strong>as</strong> been const<strong>an</strong>t over a 50-year period on <strong>the</strong> Taku Glacier at Profile 4 (Table<br />

2) (Miller, 1963; Dallenbach <strong>an</strong>d Welsch, 1993; L<strong>an</strong>g, 1995 <strong>an</strong>d 1997 <strong>an</strong>d McGee, 1998 <strong>an</strong>d<br />

2000). In addition, velocity shows no signific<strong>an</strong>t variations se<strong>as</strong>onally, b<strong>as</strong>ed on year round<br />

me<strong>as</strong>urement of <strong>the</strong> movement of <strong>the</strong> top of a glacier borehole <strong>an</strong>d <strong>the</strong> <strong>as</strong>sociated semi-perm<strong>an</strong>ent<br />

camp from 1950–1953 along Profile 4 (Miller, 1963). Movement of a meteorologic station<br />

instrument that endured from 1997 to 1998 on Profile 4 provided a second me<strong>as</strong>ure of me<strong>an</strong><br />

<strong>an</strong>nual velocity. In <strong>the</strong> former c<strong>as</strong>e me<strong>an</strong> <strong>an</strong>nual velocity w<strong>as</strong> 0.60 m/day, <strong>an</strong>d in <strong>the</strong> latter c<strong>as</strong>e<br />

0.61 m/day. The me<strong>an</strong> observed velocity for <strong>the</strong>se same locations during <strong>the</strong> summer is 0.61<br />

m/day. A magnet w<strong>as</strong> buried in <strong>the</strong> glacier in July 2003 <strong>an</strong>d <strong>the</strong>n checked in July 2004 to identify<br />

<strong>an</strong>nual glacier velocity at Profile 4. The <strong>an</strong>nual velocity w<strong>as</strong> 0.587 m/day <strong>an</strong>d <strong>the</strong> summer velocity<br />

w<strong>as</strong> 0.588 m/day. Observations of velocity at specific stake locations reoccupied each summer<br />

along Profile 4 indicate remarkable uniformity of flow from year to year (Table 3; Figure 3).<br />

St<strong>an</strong>dard deviation r<strong>an</strong>ges from 0.010–0.020 m/day. Along Profile 7 <strong>the</strong> variation is slightly less,<br />

though sp<strong>an</strong>ning <strong>the</strong> same r<strong>an</strong>ge.<br />

Table 2 Me<strong>an</strong> summer surface velocity me<strong>as</strong>ured in <strong>the</strong> center of Profile 4 from 1950–2004. Survey<br />

locations are identical from 1996–2004. The velocity for <strong>the</strong> o<strong>the</strong>r years is <strong>the</strong> me<strong>an</strong> of locations<br />

between <strong>the</strong> current locations of Point 10 <strong>an</strong>d Point 24.<br />

Year<br />

Velocity<br />

(m/day)<br />

1950 0.58<br />

1952 0.51<br />

1953 0.51<br />

1960 0.53<br />

1967 0.52<br />

1986 0.52<br />

1987 0.51<br />

1996 0.53<br />

1997 0.53<br />

1998 0.52<br />

1999 0.53<br />

2000 0.55<br />

2001 0.55<br />

2004 0.52<br />

st dev 0.02<br />

257


Velocity m/day<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Taku Glacier Profile 4 Velocity<br />

2 4 6 8 10 12 14 16 18 20 22 24 26 28<br />

Point<br />

258<br />

1996<br />

1997<br />

1998<br />

1999<br />

2000<br />

2001<br />

2004<br />

Figure 3. Observed surface velocity along Taku Glacier Profile 4 Lower Line. The lack of signific<strong>an</strong>t ch<strong>an</strong>ge<br />

is evident.<br />

Most temperate glaciers have a subst<strong>an</strong>tial component of glacier sliding that depends on bed<br />

hydrology, hence displaying se<strong>as</strong>onal variations. Taku Glacier, however, h<strong>as</strong> exceptionally thick<br />

ice, <strong>an</strong>d a low b<strong>as</strong>al gradient. The flow law for internal deformation suggests that negligible b<strong>as</strong>al<br />

sliding is taking place in <strong>the</strong> accumulation zone (Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs, 1995). The lack of se<strong>as</strong>onal<br />

velocity ch<strong>an</strong>ges noted in this study <strong>an</strong>d <strong>the</strong> remarkable uniformity in velocity suggest that sliding<br />

is a minor part of <strong>the</strong> glacier velocity at Profile 4. It is not re<strong>as</strong>onable to expect a glacier of this<br />

size to slow down each fall or winter, <strong>an</strong>d <strong>the</strong>n accelerate to exactly <strong>the</strong> same speed <strong>the</strong> following<br />

summer. This tendency h<strong>as</strong> been noted on four o<strong>the</strong>r repeat profiles on <strong>the</strong> Taku Glacier (McGee,<br />

2000).<br />

Table 3. Observed velocity each summer at specific survey locations on Profile 4. Note <strong>the</strong> nearly<br />

identical velocity at each location. Me<strong>an</strong> st<strong>an</strong>dard deviation in velocity is below 0.02 m/day.<br />

Point 1996 1997 1998 1999 2000 2001 2004 st dev<br />

10 0.42 0.43 0.41 0.43 0.43 0.5 0.46 0.031<br />

12 0.53 0.53 0.5 0.52 0.53 0.57 0.54 0.021<br />

14 0.58 0.57 0.56 0.56 0.59 0.61 0.55 0.024<br />

16 0.59 0.59 0.58 0.58 0.6 0.62 0.57 0.016<br />

18 0.59 0.59 0.58 0.6 0.61 0.61 0.57 0.015<br />

20 0.57 0.59 0.56 0.59 0.6 0.59 0.55 0.019<br />

22 0.53 0.54 0.52 0.54 0.55 0.52 0.54 0.011<br />

24 0.43 0.43 0.43 0.45 0.45 0.4 0.4 0.021<br />

me<strong>an</strong> 0.53 0.53 0.52 0.53 0.55 0.55 0.52 0.013<br />

Longitudinal Profile<br />

The variation in velocity along <strong>the</strong> longitudinal profile from Goat Ridge at 800 m elevation, 12<br />

km from <strong>the</strong> terminus, up <strong>the</strong> main trunk of <strong>the</strong> glacier, along <strong>the</strong> Mat<strong>the</strong>s Br<strong>an</strong>ch, to <strong>the</strong> ice divide


is shown in Figure 4. This variation indicates incre<strong>as</strong>ing velocity with dist<strong>an</strong>ce down glacier from<br />

<strong>the</strong> divide at Point 94, to Goat Ridge just below <strong>the</strong> ELA at Point 13. The one notable deviation<br />

from this pattern occurs where <strong>the</strong> glacier steepens, causing longitudinal extension. At this point<br />

Taku Glacier leaves <strong>the</strong> high plateau <strong>an</strong>d enters <strong>the</strong> narrower valley of <strong>the</strong> Mat<strong>the</strong>s Br<strong>an</strong>ch. The<br />

surface velocity incre<strong>as</strong>e lags <strong>the</strong> ch<strong>an</strong>ge in surface slope by several kilometers. The glacier <strong>the</strong>n<br />

slows under longitudinal compression <strong>as</strong> <strong>the</strong> surface slope declines. The velocity along this<br />

longitudinal profile h<strong>as</strong> been repeated in 2001 <strong>an</strong>d 2004 <strong>the</strong> maximum velocity ch<strong>an</strong>ge w<strong>as</strong> 0.02<br />

m/day <strong>an</strong>d <strong>the</strong> me<strong>an</strong> ch<strong>an</strong>ge for each point w<strong>as</strong> 0.004 m/day. Again, indicating <strong>the</strong> <strong>an</strong>nual <strong>an</strong>d<br />

se<strong>as</strong>onal consistency of velocity along <strong>the</strong> glacier <strong>an</strong>d <strong>the</strong> equilibrium nature of its flow (Figure 4).<br />

m/day<br />

1.00<br />

0.90<br />

0.80<br />

0.70<br />

0.60<br />

0.50<br />

0.40<br />

0.30<br />

0.20<br />

0.10<br />

0.00<br />

13<br />

17<br />

21<br />

25<br />

29<br />

33<br />

Taku Glacier Longitudinal Velocity Profile<br />

37<br />

41<br />

45<br />

49<br />

259<br />

53<br />

57<br />

61<br />

Point Location<br />

Figure 4. Comparison of gradient <strong>an</strong>d velocity along a longitudinal profile, Taku Glacier, Al<strong>as</strong>ka.<br />

Thickness<br />

The greatest thickness of <strong>the</strong> Taku Glacier w<strong>as</strong> noted to be 1477 m at Goat Ridge, 22 km above<br />

<strong>the</strong> terminus (Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs, 1995). The centerline depth of <strong>the</strong> glacier remains thicker th<strong>an</strong><br />

1400 m at Profile 4, 33 km up glacier, <strong>an</strong>d 1100 m at Profile 7, 44 km upglacier. It is likely that<br />

<strong>the</strong> Taku Glacier centerline depth is greater th<strong>an</strong> 1100 m in thickness <strong>the</strong> <strong>entire</strong> dist<strong>an</strong>ce between<br />

<strong>the</strong>se points, b<strong>as</strong>ed on <strong>the</strong> consistency in <strong>the</strong> velocity incre<strong>as</strong>e from Profile 7 to Goat Ridge. The<br />

minimum b<strong>as</strong>al elevation at Profile 4 is approximately –350 m, <strong>an</strong>d is 300 m at Profile 7. Given<br />

<strong>the</strong> relatively uniform ch<strong>an</strong>ges in slope of <strong>the</strong> glacier <strong>an</strong>d velocity between <strong>the</strong> profiles it is likely<br />

that <strong>the</strong> fjord threshold is near <strong>the</strong> mid-point between <strong>the</strong> locations.<br />

The incre<strong>as</strong>e in slope from Point 54 to Point 50 (Figure 3) <strong>an</strong>d <strong>the</strong>n decre<strong>as</strong>e is <strong>the</strong> most likely<br />

location of <strong>the</strong> sea level threshold. Point 49–52 are 39–40 km from <strong>the</strong> terminus slope, supporting<br />

<strong>the</strong> hypo<strong>the</strong>sis of Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs (1995) that <strong>the</strong> threshold w<strong>as</strong> from 38–48 km above <strong>the</strong><br />

terminus. The steady incre<strong>as</strong>e in velocity with dist<strong>an</strong>ce below this point <strong>an</strong>d <strong>the</strong> consistency of<br />

velocity with time both argue for <strong>an</strong> equilibrium flow of <strong>the</strong> Taku Glacier.<br />

65<br />

69<br />

73<br />

77<br />

81<br />

85<br />

89<br />

93<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

velocity<br />

Velocity<br />

gradient


The tr<strong>an</strong>sverse bed profile at Profile 4 indicates benches on both <strong>the</strong> e<strong>as</strong>t <strong>an</strong>d west sides of <strong>the</strong><br />

glacier (Figure 5). The bench on <strong>the</strong> e<strong>as</strong>t side in <strong>an</strong> extension of <strong>the</strong> North B<strong>as</strong>in that is at <strong>the</strong> b<strong>as</strong>e<br />

of Taku B <strong>an</strong>d just north of Camp 10. The bench on <strong>the</strong> west side lacks a clear surface topographic<br />

connection. Profile 7 lacks <strong>an</strong>y benches <strong>an</strong>d h<strong>as</strong> a much more u-shaped profile.<br />

Figure 5. Surface elevation of stations along Profile IV <strong>an</strong>d seismically determined bottom topography along<br />

<strong>the</strong> profile, Taku Glacier, Al<strong>as</strong>ka.<br />

CALCULATION OF VOLUME FLUX<br />

Profile IV<br />

Figure 5. Surface elevation of stations along Profile IV <strong>an</strong>d seismically determined bottom topography along<br />

1400<br />

<strong>the</strong> profile, Taku Glacier, Al<strong>as</strong>ka.<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

-200<br />

-400<br />

-600<br />

Elevation (m)<br />

0 1000 2000 3000 4000 5000 6000<br />

Dist<strong>an</strong>ce, looking upglacier (m)<br />

With direct me<strong>as</strong>urement of surface velocity, ice thickness <strong>an</strong>d width for each increment of<br />

glacier width on <strong>the</strong> profiles, <strong>the</strong> only unknown in determining volume flux is determination of<br />

depth average velocity. Several points led Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs (1995) to conclude that b<strong>as</strong>al sliding is<br />

minimal; most import<strong>an</strong>tly, calculation of b<strong>as</strong>al shear stresses yielded values of 125 kPa. We<br />

determined b<strong>as</strong>al shear stress to be 120–180 kPa along Profile 4, <strong>an</strong>d 75 to 100 kPa along Profile<br />

7. These values are beyond that at which b<strong>as</strong>al sliding would be <strong>an</strong>ticipated. In addition, <strong>the</strong><br />

consistency in velocity from year to year at each point indicates that <strong>the</strong>re is probably negligible<br />

se<strong>as</strong>onal fluctuation in velocity in <strong>the</strong> accumulation zone. This h<strong>as</strong> been confirmed by <strong>an</strong>nual<br />

velocity observations. It is not re<strong>as</strong>onable to expect velocity each summer to be within ±5% if<br />

se<strong>as</strong>onal variations in velocity were signific<strong>an</strong>t. Following <strong>the</strong> lead of Nol<strong>an</strong> <strong>an</strong>d o<strong>the</strong>rs (1995) <strong>an</strong>d<br />

Nye (1965) we have <strong>as</strong>sumed <strong>the</strong> depth averaged velocity is 0.8 <strong>the</strong> observed surface velocity.<br />

The me<strong>an</strong> velocity between each two survey flags is used to represent <strong>the</strong> average velocity for<br />

that width increment of <strong>the</strong> glacier. The me<strong>an</strong> depth for that width increment from <strong>the</strong> seismic<br />

profile is <strong>the</strong>n determined. The product of <strong>the</strong> width of <strong>the</strong> increment <strong>an</strong>d depth of <strong>the</strong> increment<br />

provide <strong>the</strong> me<strong>an</strong> cross-sectional area. The me<strong>an</strong> surface velocity for each increment is converted<br />

to a me<strong>an</strong> depth averaged velocity by multiplying by 0.8.<br />

260


Table 4. The calculated volume flux at Profile 4. Volume flux is in m 3 /year. Annual values are followed<br />

by a comparison of <strong>the</strong> me<strong>an</strong> volume flux, me<strong>an</strong> observed surface flux <strong>an</strong>d <strong>the</strong> difference between<br />

<strong>the</strong>m. The difference in this c<strong>as</strong>e is a positive surface flux. The volume flux is determined from <strong>an</strong>nual<br />

me<strong>as</strong>urements. The surface flux is <strong>the</strong> me<strong>an</strong> for <strong>the</strong> 1946–2004 period.<br />

1996 1997 1998 1999 2000 2001 2004 Me<strong>an</strong> Surface Difference<br />

4LL 5.25 5.34 5.07 5.38 5.42 5.41 5.20 5.27 5.5 0.20<br />

4UL 5.33 5.22 5.19 5.23 5.33 5.39 5.17 5.25 5.5 0.23<br />

The volume flux w<strong>as</strong> determined separately for parallel survey lines along Profiles 4 (upper line<br />

<strong>an</strong>d lower line) <strong>an</strong>d for <strong>the</strong> Profile 7 lower line. For profile 4, 33 km above <strong>the</strong> terminus, <strong>the</strong><br />

expected surface flux w<strong>as</strong> 5.50 × 10 8 m 3 a –1 (+10%), <strong>the</strong> volume flux r<strong>an</strong>ge w<strong>as</strong> 5.00–5.47 × 10 8<br />

m 3 a –1 , with a me<strong>an</strong> of 5.25 × 10 8 m 3 a –1 for <strong>the</strong> upper line <strong>an</strong>d 5.27 × 10 8 m 3 a –1 for <strong>the</strong> lower line.<br />

This indicates a slightly positive bal<strong>an</strong>ce <strong>an</strong>d glacier thickening above Profile 4. Glacier<br />

thickening in <strong>the</strong> r<strong>an</strong>ge of 30 m h<strong>as</strong> been suggested for this section of <strong>the</strong> glacier (Motyka,<br />

personal communication). For Profile 7, 43 km above <strong>the</strong> terminus <strong>the</strong> observed surface flux w<strong>as</strong><br />

1.90 × 10 8 m 3 a –1 <strong>an</strong>d <strong>the</strong> volume flux r<strong>an</strong>ge w<strong>as</strong> 1.72–1.74 × 10 8 m 3 a –1 , again indicating a slight<br />

glacier thickening above this profile due to a positive m<strong>as</strong>s bal<strong>an</strong>ce. The calculated volume flux is<br />

b<strong>as</strong>ed on <strong>the</strong> 1946–2004 average m<strong>as</strong>s bal<strong>an</strong>ce profile for <strong>the</strong> glacier <strong>an</strong>d not for a given year. The<br />

surface flux during recent negative bal<strong>an</strong>ce years would obviously give a lower surface flux value.<br />

CONCLUSION<br />

The results indicate that Taku Glacier h<strong>as</strong> had <strong>an</strong> equilibrium flow, with no signific<strong>an</strong>t <strong>an</strong>nual<br />

velocity ch<strong>an</strong>ges in <strong>the</strong> l<strong>as</strong>t 50 years. Fur<strong>the</strong>rmore, although se<strong>as</strong>onal variations had been expected<br />

(Miller, 1963), observations of velocity throughout <strong>the</strong> year indicate no se<strong>as</strong>onal variations,<br />

probably due to high b<strong>as</strong>al shear stress which prevents sliding. The surface m<strong>as</strong>s bal<strong>an</strong>ce<br />

accumulating above each profile in <strong>the</strong> l<strong>as</strong>t half century is in excess of <strong>the</strong> volume flux through<br />

each profile. The result, supported by both survey results of JIRP <strong>an</strong>d radio echo sounding by <strong>the</strong><br />

University of Al<strong>as</strong>ka–Fairb<strong>an</strong>ks (Nol<strong>an</strong> et al., 1995), is glacier thickening. The sustained<br />

thickening, positive bal<strong>an</strong>ce, <strong>an</strong>d consistent flux of <strong>the</strong> 1946–1988 period suggested that <strong>the</strong><br />

glacier terminus would continue to adv<strong>an</strong>ce (Pelto <strong>an</strong>d Miller, 1990). From 1988–2005 <strong>the</strong> m<strong>as</strong>s<br />

bal<strong>an</strong>ce h<strong>as</strong> been negative, though <strong>the</strong> volume flux at Profile 4 h<strong>as</strong> not declined appreciably. The<br />

reduced m<strong>as</strong>s bal<strong>an</strong>ce, if it continues, along with <strong>the</strong> proglacial delta <strong>an</strong>d exp<strong>an</strong>ding front of <strong>the</strong><br />

glacier Post <strong>an</strong>d Motyka (1995), should lead to a reduction in <strong>the</strong> adv<strong>an</strong>ce rate. The signific<strong>an</strong>t<br />

ch<strong>an</strong>ge in glacier m<strong>as</strong>s bal<strong>an</strong>ce beginning in 1988 is expected to influence <strong>the</strong> glacier velocity,<br />

volume flux <strong>an</strong>d eventually <strong>the</strong> terminus, if it is sustained. Given <strong>the</strong> slow response time of this<br />

glacier to climate ch<strong>an</strong>ge, if sustained it would not for sometime. The glacier velocity did not<br />

ch<strong>an</strong>ge appreciably <strong>as</strong> <strong>the</strong> glacier thickened by 10–20 m at Profile 4 <strong>an</strong>d it is expected that it<br />

would take a thinning of more th<strong>an</strong> this to subst<strong>an</strong>tially alter glacier velocity.<br />

261


REFERENCES<br />

Daellenbach, K. <strong>an</strong>d W. Welsch. 1993. Determination of Surface Velocities, Strain Rates, <strong>an</strong>d<br />

M<strong>as</strong>s Flow Rates on <strong>the</strong> Taku Glacier, Juneau Icefield, Al<strong>as</strong>ka. Zeitschrift fur Gletscherkunde<br />

und Glazialgeologie, B<strong>an</strong>d 26, Heft 2, p. 169–177.<br />

Dobrin, M.B. 1960. Introduction to Geophysical Prospecting, 3rd Edition, McGraw-Hill, New<br />

York, p 201–254, 630 pages.<br />

Dyugerov, M.B. <strong>an</strong>d M.F. Meier, 1997. Year to year fluctuations of global m<strong>as</strong>s bal<strong>an</strong>ce of small<br />

glaciers <strong>an</strong>d <strong>the</strong>ir contributions to sea level. Arctic <strong>an</strong>d Alpine Res., 29: 392–402.<br />

Echelmeyer, K., M. Nol<strong>an</strong>, R. Motyka, <strong>an</strong>d D. Trab<strong>an</strong>t. 1995. Ice thickness Me<strong>as</strong>urements of Taku<br />

Glacier, Al<strong>as</strong>ka, USA, <strong>an</strong>d <strong>the</strong>ir Relev<strong>an</strong>ce to it Recent Behavior. J. Glaciol., 41 (139) 541–<br />

552.<br />

L<strong>an</strong>g, M. 1997. Geodetic Activities of <strong>the</strong> 1997 Juneau Icefield Research Program Field Se<strong>as</strong>on.<br />

Open File Survey Report. Juneau Icefield Research Program, Foundation for Glacier <strong>an</strong>d<br />

Environmental Research, Moscow, Idaho. 110 pp.<br />

McGee, S. 1996. Geodetic Activities of <strong>the</strong> 1996 Juneau Icefield Research Program Field Se<strong>as</strong>on.<br />

Open File Survey Report. Juneau Icefield Research Program, Foundation for Glacier <strong>an</strong>d<br />

Environmental Research, Moscow, Idaho. 79 pp.<br />

McGee, S. 1998. Geodetic Activities of <strong>the</strong> 1998 Juneau Icefield Research Program Field Se<strong>as</strong>on.<br />

Open File Survey Report. Juneau Icefield Research Program, Foundation for Glacier <strong>an</strong>d<br />

Environmental Research, Moscow, Idaho.<br />

McGee, S. 2000. Juneau Icefield GPS Movement Profile Coordinates. JIRP Open File Survey<br />

Report. Juneau Icefield Research Program, Foundation for Glacier <strong>an</strong>d Environmental<br />

Research. Moscow, Idaho. September, 2000. 54 pp.<br />

Miller, M.M. 1963. Taku Glacier evaluation report. State of Al<strong>as</strong>ka, Dept. of Highways <strong>an</strong>d<br />

Bureau of Public Roads, U.S. Dept. of Commerce.<br />

Miller, M. M. 1997. The Juneau Icefield Research Program <strong>an</strong>d its Surveying Mission. In:<br />

Geodetic Activities Juneau Icefield, Al<strong>as</strong>ka 1981–1996. Schriftenreihe des Studieng<strong>an</strong>gs<br />

Vermessungswesen, Universität der Bundeswehr München, Heft 50. p 45.<br />

Miller, M.M. <strong>an</strong>d M.S. Pelto. 1999. M<strong>as</strong>s Bal<strong>an</strong>ce me<strong>as</strong>urements on <strong>the</strong> Lemon Creek Glacier,<br />

Juneau Icefield, AK 1953–1998. Geogr. Ann. 81A, 671–681.<br />

Nye, J.F. 1965. The flow of a glacier in a ch<strong>an</strong>nel of rect<strong>an</strong>gular, elliptic or parabolic cross<br />

section. J. Glaciol., 5(41) 661–690.<br />

Pelto, M. <strong>an</strong>d M.M. Miller. 1990. M<strong>as</strong>s Bal<strong>an</strong>ce of <strong>the</strong> Taku Glacier, Al<strong>as</strong>ka from 1946 to 1986.<br />

Northwest Science. 64 (3)121–130.<br />

Post A. <strong>an</strong>d R.J. Motyka, 1995. Taku <strong>an</strong>d Le Conte Glaciers, Al<strong>as</strong>ka: Calving speed control of<br />

late-Holocene <strong>as</strong>ynchronous adv<strong>an</strong>ces <strong>an</strong>d retreats. Phys. Geogr., 16; 59–82.<br />

Poulter, Thom<strong>as</strong> (1950) The Poulter Seismic Method of Geophysical Exploration. Geophysics, 15(<br />

2) 181–207.<br />

Ramage, J.M., B.L. Isacks, <strong>an</strong>d M.M. Miller. 2000. Radar Glacier zones in sou<strong>the</strong><strong>as</strong>t Al<strong>as</strong>ka,<br />

USA: Field <strong>an</strong>d satellite observations. J. Glaciol. 46(153), 287–296.<br />

Sprenke, K.F., M.M. Miller, F. McDonald, C. Haagen, G. Adema, M. Kelly, S. Barbour, <strong>an</strong>d B.<br />

Caceres. 1997. Geophysical investigation of <strong>the</strong> Taku-Llewellyn Divide: A NASA Earth<br />

Systems Field Research Project, Juneau Icefield, Al<strong>as</strong>ka. Juneau Icefield Research Program<br />

Geophysics Open File Report 97-1. (Moscow, Idaho: Glaciological <strong>an</strong>d Arctic Science<br />

Institute, University of Idaho).<br />

Welsch Walter, Martin L<strong>an</strong>g <strong>an</strong>d M.M. Miller, 1996. Geodetic Activities—Juneau Icefield, Al<strong>as</strong>ka<br />

1981–1996. Schriftenreihe des Studieng<strong>an</strong>gs Vermessungskunde. Universität der Bundeswehr,<br />

Heft 50.<br />

262


263<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

An Embedded Sensor Network for Me<strong>as</strong>uring<br />

Hydrometeorological Variability Within <strong>an</strong> Alpine Valley<br />

ABSTRACT<br />

ROBERT Å. HELLSTRÖM 1 AND BRYAN G. MARK 2<br />

Conditions of glacier recession in <strong>the</strong> se<strong>as</strong>onally dry tropical Peruvi<strong>an</strong> Andes motivate research<br />

to better constrain <strong>the</strong> hydrological bal<strong>an</strong>ce in alpine valleys. Studies suggest that glacial m<strong>as</strong>s<br />

bal<strong>an</strong>ce in <strong>the</strong> outer tropics of <strong>the</strong> Andes is particularly sensitive to variations between <strong>the</strong> dry <strong>an</strong>d<br />

wet se<strong>as</strong>on humidity flux. In this context, we introduce a novel embedded network of low-cost,<br />

discrete temperature microloggers <strong>an</strong>d <strong>an</strong> automatic wea<strong>the</strong>r station installed in <strong>the</strong> Ll<strong>an</strong>g<strong>an</strong>uco<br />

valley of <strong>the</strong> Cordillera Bl<strong>an</strong>ca. This paper presents data for distinct dry <strong>an</strong>d wet periods sampled<br />

from a full <strong>an</strong>nual cycle (2004-2005) <strong>an</strong>d reports on modeled estimations of evapotr<strong>an</strong>spiration<br />

(ET). The tr<strong>an</strong>sect of temperature sensors r<strong>an</strong>ging from about 3500 to 4700 m revealed se<strong>as</strong>onally<br />

characteristic diurnal fluctuations in up-valley lapse rates that promote up-slope warm air<br />

convection that will affect <strong>the</strong> energy bal<strong>an</strong>ce of <strong>the</strong> glacier tongue. Nocturnal rainfall dominated<br />

<strong>the</strong> wet se<strong>as</strong>on. Strong solar forcing dominated during both dry <strong>an</strong>d wet periods, but extreme<br />

se<strong>as</strong>onal variations in soil water content <strong>an</strong>d cooler wet se<strong>as</strong>on near-surface air temperature<br />

suggests <strong>the</strong> import<strong>an</strong>ce of considering <strong>the</strong> process of ET. Estimates of potential ET using <strong>the</strong><br />

widely applied Penm<strong>an</strong>-Monteith FAO model suggest nearly twice <strong>as</strong> much for <strong>the</strong> dry period, <strong>an</strong>d<br />

we attribute this primarily to <strong>the</strong> five times higher dry period vapor pressure deficit. We r<strong>an</strong> a<br />

process-b<strong>as</strong>ed water bal<strong>an</strong>ce model, BROOK90, to estimate actual ET, which w<strong>as</strong> nearly 100<br />

times greater for <strong>the</strong> wet se<strong>as</strong>on. These results reinforce <strong>the</strong> import<strong>an</strong>ce of diurnal cloud cover<br />

variability in regulating ET in <strong>the</strong> Peruvi<strong>an</strong> Andes.<br />

Keywords: tropical, alpine, embedded sensors, evapotr<strong>an</strong>spiration, diurnal, se<strong>as</strong>onal modeling<br />

INTRODUCTION<br />

Tropical Ande<strong>an</strong> glacier recession over <strong>the</strong> p<strong>as</strong>t century h<strong>as</strong> profound local consequences for<br />

water resources, <strong>an</strong>d motivates fur<strong>the</strong>r hydroclimatic research (Fr<strong>an</strong>cou et al., 1997; H<strong>as</strong>tenrath<br />

<strong>an</strong>d Kruss, 1992; Mark <strong>an</strong>d Seltzer, 2003, 2005).<br />

While observations of continued glacier recession exist throughout <strong>the</strong> Andes, only a few efforts<br />

have been made to qu<strong>an</strong>tify <strong>the</strong> hydroclimatic ch<strong>an</strong>ges on a scale most relev<strong>an</strong>t to hum<strong>an</strong> impact.<br />

Previous studies primarily focus on <strong>the</strong> rates, controls, <strong>an</strong>d flux of glacier melt water directly from<br />

<strong>an</strong>d within <strong>the</strong> glaciers (Fr<strong>an</strong>cou et al., 1995; Ribstein et al., 1995; Wagnon et al., 1999; Wagnon<br />

et al., 1998), but few studies look at <strong>the</strong> fate of <strong>the</strong> water once it leaves <strong>the</strong> glacier system.<br />

Moreover, lack of accomp<strong>an</strong>ying precipitation <strong>an</strong>d stream discharge data also preclude <strong>an</strong>alyses in<br />

1<br />

Bridgewater State College, Department of Geography, Con<strong>an</strong>t Science Building, Bridgewater,<br />

MA 02325, rhellstrom@bridgew.edu<br />

2<br />

The Ohio State University, Department of Geography, 1036 Derby Hall, 154 N Oval Mall,<br />

Columbus, OH 43210, mark.9@osu.edu


all but a h<strong>an</strong>dful of sites. There is <strong>an</strong> outst<strong>an</strong>ding need to better qu<strong>an</strong>tify <strong>the</strong> volume of<br />

contributions from all hydrologic components in this dynamic climate on <strong>the</strong> regional scale where<br />

hum<strong>an</strong>s utilize water resources, <strong>an</strong>d where future m<strong>an</strong>agement decisions must be made.<br />

Only initial attempts have been made to account for <strong>the</strong> regional flux of hydrologic components<br />

in <strong>the</strong> valleys of <strong>the</strong> tropical Andes, utilizing major simplifying <strong>as</strong>sumptions for lack of empirical<br />

data. It h<strong>as</strong> proven very challenging to maintain regular meteorological observations due to hum<strong>an</strong><br />

disturb<strong>an</strong>ce <strong>an</strong>d technical issues in <strong>the</strong> extreme environment (Georges <strong>an</strong>d K<strong>as</strong>er, 2002).<br />

Groundwater <strong>an</strong>d evaporation have been <strong>as</strong>sumed negligible in glacierized valleys, given high<br />

relief <strong>an</strong>d steep slopes (Caballero et al., 2004; K<strong>as</strong>er et al., 2003). The hydrological processes in<br />

<strong>the</strong>se high tropical mountain valley deposits remain understudied, <strong>an</strong>d <strong>the</strong>re is a need for<br />

systematic research to better underst<strong>an</strong>d <strong>the</strong> impact of runoff down valley by different<br />

morphological zones (Caballero et al., 2002).<br />

The hydrologic bal<strong>an</strong>ce of tropical glaciers at various time scales (H<strong>as</strong>tenrath <strong>an</strong>d Ames, 1995;<br />

K<strong>as</strong>er et al, 2003; Fr<strong>an</strong>cou et al., 2000; Fr<strong>an</strong>cou et al., 2003) h<strong>as</strong> been <strong>the</strong> focus of recent research,<br />

but me<strong>as</strong>urements <strong>an</strong>d model simulations of <strong>the</strong> role of evapotr<strong>an</strong>spiration (ET) in <strong>the</strong> pro-glacial<br />

valleys are lacking. The low barometric pressure in high elevations promotes ET, but incident<br />

solar radiation, wind, humidity, precipitation <strong>an</strong>d temperature are primary regulating components<br />

of meteorological forcing. Studies of mid-latitude pro-glacial valleys suggest signific<strong>an</strong>t<br />

contributions of ET to <strong>the</strong> hydrologic bal<strong>an</strong>ce. For example, Konzelm<strong>an</strong>n et al. (1996) found that<br />

<strong>the</strong> type of vegetation <strong>an</strong>d surface texture strongly control <strong>the</strong> rate of ET at different elevations in<br />

<strong>the</strong> Dischma valley, near Davos, Switzerl<strong>an</strong>d. Konzelm<strong>an</strong>n et al. concluded that ET in this alpine<br />

valley w<strong>as</strong> regulated by soil moisture <strong>an</strong>d <strong>the</strong> physiological behavior of <strong>the</strong> vegetation. However,<br />

<strong>the</strong> influence of soil, vegetation, <strong>an</strong>d local meteorological forcing on ET in tropical pro-glacial<br />

valleys is largely unverified <strong>an</strong>d <strong>as</strong>sumed negligible. Current ground-b<strong>as</strong>ed meteorological<br />

observations in <strong>the</strong> Peruvi<strong>an</strong> Andes, such <strong>as</strong> Hardy et al. (1998) <strong>an</strong>d Vuille et al. (2003) do not<br />

extend to pro-glacial valleys, so it is necessary to develop <strong>an</strong>d deploy a new me<strong>as</strong>urement<br />

network. ET is largely controlled by a combination of available surface moisture, atmospheric<br />

vapor pressure, air temperature <strong>an</strong>d wind speed. Vuille et al. (2003) report signific<strong>an</strong>t incre<strong>as</strong>es in<br />

near-surface air temperature throughout most of <strong>the</strong> tropical Andes. Interestingly, <strong>the</strong> temperature<br />

incre<strong>as</strong>e varies markedly between <strong>the</strong> e<strong>as</strong>tern <strong>an</strong>d western Ande<strong>an</strong> slopes with a much larger<br />

temperature incre<strong>as</strong>e to <strong>the</strong> west.<br />

In this paper, we introduce a new network of monitoring instruments from <strong>the</strong> nor<strong>the</strong><strong>as</strong>tsouthwest<br />

trending Ll<strong>an</strong>g<strong>an</strong>uco pro-glacial valley of <strong>the</strong> western Cordillera Bl<strong>an</strong>ca, <strong>an</strong>d report on<br />

model estimations of ET. We present data for dry <strong>an</strong>d wet periods from a full <strong>an</strong>nual cycle (2004-<br />

2005) <strong>an</strong>d report on <strong>the</strong> sensor network design, including a tr<strong>an</strong>sect of automatic temperature<br />

sensors r<strong>an</strong>ging from about 3500 to 4700 m. We applied <strong>the</strong> Penm<strong>an</strong>-Monteith FAO method of<br />

estimating hourly potential ET (ET0) during <strong>the</strong> wet <strong>an</strong>d dry periods. We r<strong>an</strong> a process-b<strong>as</strong>ed<br />

water bal<strong>an</strong>ce model (Brook90: Federer, 2003) to examine <strong>the</strong> influence of meteorological forcing<br />

on ET rates <strong>an</strong>d compare <strong>the</strong> contributing sources of ET, <strong>an</strong>d made comparisons with <strong>the</strong> ET0<br />

estimations. Unlike most research on processes affecting ET, which targets mid-latitude <strong>an</strong>d<br />

subtropical sites, we present methods <strong>an</strong>d results for a hydroclimatologically sensitive region at<br />

<strong>the</strong> tropical-sub-tropical interface, <strong>the</strong> outer tropics.<br />

Study area<br />

The Ande<strong>an</strong> Cordillera Bl<strong>an</strong>ca is home to <strong>the</strong> greatest concentration of tropical glaciers on earth<br />

(Fig. 1). It is <strong>the</strong> largest <strong>an</strong>d most nor<strong>the</strong>rly mountain r<strong>an</strong>ge in Peru, trending NW-SE over 130 km<br />

between 8º–10º S latitude (Ames, 1998) along <strong>the</strong> Ande<strong>an</strong> continental divide. Most of <strong>the</strong><br />

glacierized area in <strong>the</strong> Cordillera Bl<strong>an</strong>ca discharges northwest via <strong>the</strong> S<strong>an</strong>ta River, that h<strong>as</strong> <strong>the</strong><br />

le<strong>as</strong>t variable monthly runoff of all Pacific draining rivers.<br />

264


Figure 1. Cordillera Bl<strong>an</strong>ca regional map. The Ll<strong>an</strong>g<strong>an</strong>uco is one of several pro-glacial valleys in <strong>the</strong> region.<br />

The Ll<strong>an</strong>g<strong>an</strong>uco valley is a cl<strong>as</strong>sic U-shaped h<strong>an</strong>ging valley tributary draining SW to <strong>the</strong> S<strong>an</strong>ta<br />

River (Figs. 2 <strong>an</strong>d 3). Its mouth is fl<strong>an</strong>ked by steep walls of <strong>the</strong> glacially sculpted gr<strong>an</strong>odiorite<br />

bedrock that comprises <strong>the</strong> Upper Miocene batholithic core of <strong>the</strong> r<strong>an</strong>ge (McNulty et al., 1998).<br />

Summits that border <strong>the</strong> catchment include Hu<strong>as</strong>carán (6768 m) to <strong>the</strong> south, highest in Peru, <strong>the</strong><br />

Hu<strong>an</strong>doy m<strong>as</strong>sif (6395 m) to <strong>the</strong> north, <strong>an</strong>d Chacraraju (6113 m) at <strong>the</strong> NE valley head. The valley<br />

contains two lakes, <strong>an</strong>d is a well-visited tourist attraction within <strong>the</strong> Hu<strong>as</strong>carán National Park <strong>an</strong>d<br />

International Biodiversity Reserve. The road along <strong>the</strong> long axis of <strong>the</strong> valley forms one of three<br />

principal tr<strong>an</strong>sect routes over <strong>the</strong> Cordillera Bl<strong>an</strong>ca <strong>an</strong>d reaches a high p<strong>as</strong>s at <strong>the</strong> Portachuelo de<br />

Ll<strong>an</strong>g<strong>an</strong>uco (4767 m) <strong>an</strong>d it continues down <strong>the</strong> e<strong>as</strong>t side.<br />

265


3469<br />

3862<br />

3850<br />

3871<br />

Figure 2. Ll<strong>an</strong>g<strong>an</strong>uco valley <strong>an</strong>d location of wea<strong>the</strong>r instruments: note <strong>the</strong> glaciers north <strong>an</strong>d south of SW-NE<br />

trending valley.<br />

Figure 3. SW-looking view of Ll<strong>an</strong>g<strong>an</strong>uco valley from Portachuelo site near <strong>the</strong> highest elevation iButton<br />

logger. Note <strong>the</strong> highl<strong>an</strong>d gr<strong>as</strong>ses, steep rocky walls, terminus of glacier above <strong>the</strong> north wall, <strong>an</strong>d S<strong>an</strong>ta<br />

River valley in <strong>the</strong> dist<strong>an</strong>ce.<br />

266<br />

3948<br />

4148<br />

4559<br />

4344<br />

4742


MATERIALS AND METHODS<br />

Field Me<strong>as</strong>urements<br />

In July 2004, we installed <strong>an</strong> automatic wea<strong>the</strong>r station (Onset HOBO®) in <strong>the</strong> Ll<strong>an</strong>g<strong>an</strong>uco<br />

tributary valley (3850 m.a.s.l.). Instrumentation specifications are available:<br />

http://www.onsetcomp.com/Products/Product_Pages/wea<strong>the</strong>rstation/wea<strong>the</strong>r_station_logger.html.<br />

Hourly data are being collected in collaboration with <strong>the</strong> University of Innsbruck including soil<br />

moisture, soil temperature, air temperature, wind speed <strong>an</strong>d direction, relative humidity, solar<br />

radiation, <strong>an</strong>d precipitation.<br />

In June 2005, we installed a set of 9 temperature loggers (iButton Thermochron®) at different<br />

elevations r<strong>an</strong>ging from 3469 to 4742 m above me<strong>an</strong> sea-level (Fig. 2). Each logger is powered by<br />

<strong>an</strong> internal battery (1 year lifetime) <strong>an</strong>d w<strong>as</strong> programmed to record hourly intervals of air<br />

temperature. The reported iButton resolution is 0.5°C, accuracy is ±1°C with a r<strong>an</strong>ge of -40°C to<br />

+85°C (http://www.maxim-ic.com). Each iButton logger w<strong>as</strong> placed inside a specially designed<br />

¾” PVC shields that allowed us to embed <strong>the</strong> sensors in small trees <strong>an</strong>d o<strong>the</strong>r inconspicuous<br />

locations (Figs. 4 <strong>an</strong>d 5).<br />

Figure 4. HOBO wea<strong>the</strong>r station at b<strong>as</strong>e of lower lake, looking SW. Note <strong>the</strong> Polylipus (Queñual) trees <strong>an</strong>d<br />

gr<strong>as</strong>s vegetation cover. The precipitation logger to <strong>the</strong> right w<strong>as</strong> set up by K<strong>as</strong>er, U. Innsbruck.<br />

267


Figure 5. (a) ¾” PVC ventilated radiation shield for iButton Thermochron logger. Note that <strong>the</strong> elbow<br />

sections are detachable for installation in vegetation. (b) Typical installation of iButton temperature loggers,<br />

one atop 1.5 m PVC post (top) <strong>an</strong>d one embedded at 1.5 m inside a Queñual tree.<br />

One iButton w<strong>as</strong> calibrated under actual field conditions against <strong>the</strong> air temperature from <strong>the</strong><br />

HOBO wea<strong>the</strong>r station (Fig. 6). The calibration is within <strong>the</strong> toler<strong>an</strong>ce of <strong>the</strong> iButton sensors<br />

(approximately 3%). Every two months, each iButton w<strong>as</strong> checked <strong>an</strong>d <strong>the</strong> data were downloaded<br />

to a laptop computer using <strong>the</strong> iButton USB port adaptor (Jesus Gomez, INRENA, Perú).<br />

iButton Temperature (°C)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

(a)<br />

iButton Temperature Callibration<br />

iButton = 1.0211(HOBO) - 0.0561<br />

R 2 = 0.988<br />

2 4 6 8 10 12 14 16<br />

HOBO Temperature (°C)<br />

Figure 6. Calibration for iButton sensor inside PVC radiation shield.<br />

268<br />

(b)<br />

(c)


Data <strong>an</strong>alysis techniques<br />

We syn<strong>the</strong>sized all data from <strong>the</strong> 8 iButton temperature loggers on <strong>the</strong> western slope, <strong>the</strong> HOBO<br />

wea<strong>the</strong>r station, <strong>an</strong>d <strong>the</strong> precipitation logger into hourly values to create 24-day periods during <strong>the</strong><br />

2005 dry, 17 June to 11 July, <strong>an</strong>d wet, 7 to 31 December, se<strong>as</strong>ons. These sample periods were<br />

chosen b<strong>as</strong>ed on data availability <strong>an</strong>d central proximity to <strong>the</strong> dry <strong>an</strong>d wet se<strong>as</strong>ons. We created<br />

composite averages of 24 hourly me<strong>as</strong>urements for all 24 days making up <strong>the</strong> wet <strong>an</strong>d dry<br />

periods—<strong>the</strong> diurnal cycle. Hence, composites are averages for each hour over <strong>the</strong> 24-day periods.<br />

The vector components of wind speed were calculated to create composite averages of speed <strong>an</strong>d<br />

direction. Because we were interested in microscale variability, particularly elevation effects, we<br />

concentrated on <strong>an</strong>alyzing <strong>an</strong>d comparing <strong>the</strong> composite diurnal cycles of meteorological<br />

components for <strong>the</strong> wet <strong>an</strong>d dry periods.<br />

Estimating Evapotr<strong>an</strong>spiration<br />

It is difficult to me<strong>as</strong>ure ET accurately, particularly in remote mountainous regions with steep<br />

topography; most me<strong>as</strong>urements require expensive equipment <strong>an</strong>d frequent mainten<strong>an</strong>ce (refs).<br />

Consequently, <strong>the</strong>re are a multitude of methods for estimating ET, most of which require site<br />

specific parameters <strong>an</strong>d b<strong>as</strong>ic meteorological me<strong>as</strong>urements. Evaporation rates, ignoring<br />

tr<strong>an</strong>spiration, are commonly estimated by energy bal<strong>an</strong>ce, aerodynamic, <strong>an</strong>d a combination of both<br />

methods, such <strong>as</strong> Penm<strong>an</strong> (1945; 1963) <strong>an</strong>d Penm<strong>an</strong>-Monteith. Xu <strong>an</strong>d Singh (1998; 2002) report<br />

on meteorological forcing of <strong>an</strong>d compares methods of estimating ET b<strong>as</strong>ed on input data from<br />

Ch<strong>an</strong>gines station in Switzerl<strong>an</strong>d. We used a similar modeling approach <strong>an</strong>d applied it to <strong>the</strong><br />

Ll<strong>an</strong>g<strong>an</strong>uco valley. Allen et al. (1989) reports on <strong>the</strong> FAO methods for estimating potential ET<br />

(ET0) b<strong>as</strong>ed on Penm<strong>an</strong> combination equations for reference surfaces, such <strong>as</strong> uniform 120 mm<br />

tall gr<strong>as</strong>s used herein. Because of it’s inclusion of several commonly used models for estimating<br />

ET0, we elected to use <strong>the</strong> REF-ET software developed by Allen et al. (1998) <strong>an</strong>d selected <strong>the</strong><br />

Penm<strong>an</strong>-Monteith FAO method for estimating hourly ET0 during <strong>the</strong> dry <strong>an</strong>d wet periods. This<br />

will serve <strong>as</strong> a b<strong>as</strong>e-line for comparison to a more realistic physics-b<strong>as</strong>ed model for ET. The<br />

equations for <strong>the</strong> FAO methods are explained in detail by Allen et al. (1998).<br />

We selected <strong>the</strong> BROOK90 (v.4.4e) model (Federer et al., 2003; Federer, 1995) to estimate <strong>the</strong><br />

actual ET during <strong>the</strong> dry <strong>an</strong>d wet se<strong>as</strong>on composite days. BROOK90 includes strong physicallyb<strong>as</strong>ed<br />

determination of ET, a graphic user interface, <strong>an</strong>d <strong>the</strong> visual b<strong>as</strong>ic code is available for<br />

modification. The model simulates deposition <strong>an</strong>d sublimation of frozen water <strong>an</strong>d <strong>as</strong>sumes<br />

snowfall for near-surface air temperature below –1.5°C. The model meteorological input includes<br />

solar radiation, air temperature, vapor pressure, wind speed, <strong>an</strong>d precipitation. Soil <strong>an</strong>d vegetation<br />

parameters were me<strong>as</strong>ured in <strong>the</strong> field. Additional parameters were taken from <strong>the</strong> literature <strong>as</strong><br />

reported by Federer et al. (2003). Model output includes hydrological bal<strong>an</strong>ce components of <strong>the</strong><br />

vegetation c<strong>an</strong>opy <strong>an</strong>d soil surface; we here report on <strong>the</strong> ET components.<br />

BROOK90 simulates evaporation <strong>an</strong>d soil-water movement using a process-oriented approach<br />

for sparse c<strong>an</strong>opies at a single location within a watershed. The model estimates interception <strong>an</strong>d<br />

tr<strong>an</strong>spiration from a single-layer pl<strong>an</strong>t c<strong>an</strong>opy, soil <strong>an</strong>d snow evaporation/sublimation, snow<br />

accumulation <strong>an</strong>d melt, <strong>an</strong>d soil water movement through multiple soil layers. Potential<br />

evaporation rates are obtained using <strong>the</strong> Shuttleworth <strong>an</strong>d Wallace (1985) modification of <strong>the</strong><br />

Penm<strong>an</strong>-Monteith combination equation.<br />

Actual tr<strong>an</strong>spiration is <strong>the</strong> lesser of potential tr<strong>an</strong>spiration <strong>an</strong>d a soil water supply rate<br />

determined by <strong>the</strong> resist<strong>an</strong>ce to liquid water flow in <strong>the</strong> pl<strong>an</strong>ts <strong>an</strong>d on root distribution <strong>an</strong>d soil<br />

water potential in <strong>the</strong> soil layers (Federer, 1979). For potential tr<strong>an</strong>spiration, c<strong>an</strong>opy resist<strong>an</strong>ce<br />

depends on maximum leaf conduct<strong>an</strong>ce, reduced for humidity, temperature, <strong>an</strong>d light penetration.<br />

Each soil layer <strong>an</strong>d <strong>the</strong> roots of vegetation have a resist<strong>an</strong>ce to water flow b<strong>as</strong>ed on field<br />

observations <strong>an</strong>d published results. Aerodynamic resist<strong>an</strong>ces are modified from Shuttleworth <strong>an</strong>d<br />

Gurney (1990); <strong>the</strong>y depend on leaf area index (LAI), which c<strong>an</strong> vary se<strong>as</strong>onally, <strong>an</strong>d on c<strong>an</strong>opy<br />

height, which determines stem area index (SAI). Potential tr<strong>an</strong>spiration or potential interception<br />

are obtained using <strong>the</strong> actual or existing soil surface wetness in <strong>the</strong> Shuttleworth-Wallace<br />

equations. The equations <strong>the</strong>n provide <strong>the</strong> total soil or ground evaporation. The total ET is <strong>the</strong> sum<br />

of precipitation evaporated from <strong>the</strong> c<strong>an</strong>opy, soil evaporation, <strong>an</strong>d tr<strong>an</strong>spiration from <strong>the</strong> c<strong>an</strong>opy.<br />

269


In contr<strong>as</strong>t to <strong>the</strong> six ET0 models, <strong>the</strong> BROOK90 model provided more realistic representation<br />

of vegetation <strong>an</strong>d soil parameters <strong>an</strong>d allowed evaluation of <strong>the</strong> hydrologic components affecting<br />

ET. The components included infiltration, surface evaporation, tr<strong>an</strong>spiration, <strong>an</strong>d c<strong>an</strong>opy<br />

intercepted evaporation.<br />

RESULTS<br />

Field Me<strong>as</strong>urements<br />

We report on <strong>the</strong> results of intercomparisons for dry <strong>an</strong>d wet composite diurnal cycles of air<br />

temperature, relative humidity, insolation, precipitation, wind speed <strong>an</strong>d direction, <strong>an</strong>d iButton<br />

temperatures.<br />

HOBO automated wea<strong>the</strong>r station<br />

We were particularly careful with our data syn<strong>the</strong>ses from <strong>the</strong> humidity sensor (naturally<br />

<strong>as</strong>pirated), which w<strong>as</strong> affected by supersaturation during portions of <strong>the</strong> wet se<strong>as</strong>on. We computed<br />

<strong>the</strong> dew point temperatures from humidity <strong>an</strong>d relative humidity me<strong>as</strong>urements <strong>an</strong>d adjusted<br />

erroneous humidity values according to our corrections. Table 1 summarizes <strong>the</strong> results from <strong>the</strong><br />

HOBO wea<strong>the</strong>r station. The ratio of dry to wet w<strong>as</strong> applied to demonstrate differences.<br />

Table 1. Comparison of average daily statistics for variables me<strong>as</strong>ured by HOBO wea<strong>the</strong>r station<br />

during <strong>the</strong> dry <strong>an</strong>d wet periods.<br />

HOBO Wea<strong>the</strong>r Station<br />

(daily statistics from hourly data)<br />

Dry Wet<br />

Dry/<br />

Wet<br />

Air Temp. (°C) 7.6 6.6 1.15<br />

Max. Temp. (°C) 16.0 12.7 1.26<br />

Min. Temp. (°C) 1.5 3.6 0.42<br />

Temp. R<strong>an</strong>ge. (°C) 14.5 9.1 1.59<br />

RH (%) 59.5 92.4 0.64<br />

Vap. Press. (kPa) 0.62 0.90 0.69<br />

VP Deficit (kPa) 0.43 0.08 5.38<br />

Insolation (MJ/m 2 ) 16.82 13.21 1.27<br />

Wind (m/s) 1.42 1.18 1.20<br />

Wind Direction (°) 39 233 N/A<br />

Wind Const<strong>an</strong>cy 0.54 0.72 0.75<br />

0.10 m Soil Moist. (m 3 /m 3 ) 0.005 0.119 0.04<br />

0.10 m Soil Temp. (°C) 13.8 12.8 1.08<br />

Precip. (mm/day) 0.08 7.36 0.01<br />

Comparisons of <strong>the</strong> HOBO wea<strong>the</strong>r station data for <strong>the</strong> composite dry <strong>an</strong>d wet periods suggest<br />

signific<strong>an</strong>t differences between all variables of meteorological forcing. The average day during <strong>the</strong><br />

wet period is 1°C cooler th<strong>an</strong> <strong>the</strong> dry period, given <strong>the</strong> lack of rainfall <strong>an</strong>d greater insolation<br />

during <strong>the</strong> dry period. Because of incre<strong>as</strong>ed cloud cover <strong>as</strong>sociated with higher rainfall, <strong>the</strong> diurnal<br />

temperature r<strong>an</strong>ge is 5.4°C less during <strong>the</strong> wet period. The average relative humidity of 59.5% for<br />

<strong>the</strong> dry period is 33% less th<strong>an</strong> that of <strong>the</strong> wet se<strong>as</strong>on. The vapor pressure of 0.62 kPa for <strong>the</strong> dry<br />

period w<strong>as</strong> 69% of that for <strong>the</strong> wet period, 0.90 kPa.<br />

Figs. 7 through 12 illustrate <strong>the</strong> diurnal cycles for <strong>the</strong> dry <strong>an</strong>d wet periods. The persistent valley<br />

winds (Figs. 10b, 11b <strong>an</strong>d 12b) during <strong>the</strong> daylight hours of <strong>the</strong> wet period controls <strong>the</strong> efficiency<br />

of <strong>the</strong> impact of turbulent sensible <strong>an</strong>d latent heat flux on ET rate. Assuming a saturated<br />

evaporating surface, <strong>the</strong> aerodynamic term of <strong>the</strong> Penm<strong>an</strong>-Monteith method maximizes ET0 with<br />

concurrently high wind speed <strong>an</strong>d vapor pressure deficit. This <strong>as</strong>sumption is good for <strong>the</strong> wet<br />

270


se<strong>as</strong>on, but unacceptable for <strong>the</strong> dry se<strong>as</strong>on when <strong>the</strong> surface is extremely dry (Fig. 8). The<br />

process of near-surface water vapor removal depends to a large extent on wind speed <strong>an</strong>d air<br />

turbulence. When vaporizing water, <strong>the</strong> air above <strong>the</strong> evaporating surface becomes gradually<br />

saturated. If this air is not continuously replaced with drier air, <strong>the</strong> driving force for water vapor<br />

removal <strong>an</strong>d <strong>the</strong> ET rate decre<strong>as</strong>es. Hence, <strong>the</strong> domin<strong>an</strong>ce of daytime valley winds during <strong>the</strong> wet<br />

se<strong>as</strong>on plays a critical role in controlling ET within <strong>the</strong> pro-glacial valley, particularly given <strong>the</strong><br />

steep confining walls. Fur<strong>the</strong>rmore, drier surface conditions in <strong>the</strong> lower valley will enh<strong>an</strong>ce <strong>the</strong><br />

potential for ET throughout <strong>the</strong> valley, which is plausible given <strong>the</strong> predomin<strong>an</strong>tly e<strong>as</strong>terly<br />

synoptic flow <strong>an</strong>d <strong>the</strong> rain shadow effect to <strong>the</strong> west of <strong>the</strong> mountain r<strong>an</strong>ge.<br />

(a)<br />

Air Temperature (°C)<br />

Dry Period Diurnal Variation of Average Air Temperature<br />

<strong>an</strong>d Relative Humidity<br />

Relative Humidity Air Temperature<br />

20<br />

100<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

Time of Day (hours)<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Relative Humidity (%)<br />

Air Temperature (°C)<br />

271<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0:00<br />

2:00<br />

Wet Period Diurnal Variation of Average Air<br />

Temperature <strong>an</strong>d Relative Humidity<br />

Relative Humidity Air Temperature<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

Time of Day (hours)<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

Figure 7. Composite diurnal cycle of air temperature <strong>an</strong>d relative humidity for dry (a) <strong>an</strong>d wet (b) periods.<br />

Soil Temperature (°C)<br />

30<br />

26<br />

22<br />

18<br />

14<br />

10<br />

6<br />

0:00<br />

2:00<br />

Dry Period Diurnal Variation of Average Soil<br />

Temperature <strong>an</strong>d Soil Water Content<br />

Soil Water Content Soil Temperature<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

Time of Day (hours)<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

0.11<br />

0.09<br />

0.07<br />

0.05<br />

0.03<br />

0.01<br />

-0.01<br />

Water Content (m 3 /m 3 )<br />

Soil Temperature (°C)<br />

30<br />

26<br />

22<br />

18<br />

14<br />

10<br />

6<br />

0:00<br />

2:00<br />

Wet Period Diurnal Variation of Average Soil<br />

Temperature <strong>an</strong>d Soil Water Content<br />

Soil Water Content Soil Temperature<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

Time of Day (hours)<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

Figure 8. Composite diurnal cycle of incoming soil moisture <strong>an</strong>d temperature for dry (a) <strong>an</strong>d wet (b) periods.<br />

Precipitation, swe (mm)<br />

(a)<br />

(a)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Dry Period Diurnal Variation of Average Precipitation<br />

<strong>an</strong>d Insolation<br />

Precipitation Insolation<br />

1000<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

Time of Day (hours)<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Incident Solar (W/m 2 )<br />

Precipitation, swe (mm)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0.11<br />

0.09<br />

0.07<br />

0.05<br />

0.03<br />

0.01<br />

-0.01<br />

Wet Period Diurnal Variation of Average Precipitation<br />

<strong>an</strong>d Insolation<br />

Precipitation Insolation<br />

1000<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

Time of Day (hours)<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

Figure 9. Composite diurnal cycle of incoming solar radiation <strong>an</strong>d precipitation for dry (a) <strong>an</strong>d wet (b)<br />

periods.<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Water Content (m 3 /m 3 )<br />

Relative Humidity (%)<br />

Incident Solar (W/m 2 )<br />

(a)<br />

(b)


(a)<br />

Wind Direction (°)<br />

Wind Speed (m/s)<br />

Air Temperature (°C)<br />

360<br />

300<br />

240<br />

180<br />

120<br />

60<br />

(a)<br />

(a)<br />

0<br />

3.5<br />

2.5<br />

1.5<br />

0.5<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

3<br />

2<br />

1<br />

0<br />

Dry Period Diurnal Variation of Wind Speed <strong>an</strong>d<br />

Direction<br />

0:00<br />

2:00<br />

Wind Dir.<br />

Wind Speed<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

12:00<br />

Time (hours)<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Wind Speed (m/s)<br />

Wind Direction (°)<br />

272<br />

360<br />

300<br />

240<br />

180<br />

120<br />

60<br />

0<br />

Wet Period Diurnal Variation of Wind Speed <strong>an</strong>d<br />

Direction<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

12:00<br />

Time (hours)<br />

Wind Dir.<br />

Wind Speed<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

Figure 10. Composite diurnal cycle of wind speed <strong>an</strong>d direction for dry (a) <strong>an</strong>d wet (b) periods.<br />

Dry Period Wind Speed versus Direction<br />

Valley Orientation<br />

55° => 235°<br />

0 45 90 135 180 225 270 315 360<br />

Wind Direction (°)<br />

Wind Speed (m/s)<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Wet Period Wind Speed versus Direction<br />

Valley Orientation<br />

55° => 235°<br />

0 45 90 135 180 225 270 315 360<br />

Wind Direction (°)<br />

Figure 11: Composite wind speed versus direction for dry (a) <strong>an</strong>d wet (b) periods.<br />

Dry Period Air Temperature versus Wind Direction<br />

Valley Orientation<br />

55° => 235°<br />

0 45 90 135 180 225 270 315 360<br />

Wind Direction (°)<br />

Air Temperature (°C)<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

Wet Period Air Temperature versus Wind Direction<br />

Valley Orientation<br />

55° => 235°<br />

0 45 90 135 180 225 270 315 360<br />

Wind Direction (°)<br />

Figure 12: Composite wind speed versus air temperature for dry (a) <strong>an</strong>d wet (b) periods.<br />

iButton temperatures<br />

Table 2 summarizes <strong>the</strong> iButton temperature profile observed during <strong>the</strong> wet <strong>an</strong>d dry periods.<br />

With exception of <strong>the</strong> iButton at 3948 m, which is cooler during <strong>the</strong> wet se<strong>as</strong>on due to shading<br />

from <strong>the</strong> south wall off <strong>the</strong> valley, <strong>the</strong> general trend is for stronger se<strong>as</strong>onal contr<strong>as</strong>ts at higher<br />

elevation. Although <strong>the</strong> up-valley “lapse rate” is not <strong>the</strong> conventional atmospheric lapse rate, it is a<br />

strong indicator of sensible heat flux within <strong>the</strong> surface boundary layer, <strong>the</strong>reby affecting <strong>the</strong> rates<br />

of ice <strong>an</strong>d snow ablation. Figs. 13 provides a unique perspective on <strong>the</strong> diurnal variability of<br />

temperature <strong>as</strong> function of elevation. The strong nocturnal inversion below <strong>the</strong> level of <strong>the</strong><br />

meltwater lakes during <strong>the</strong> dry period is not present during <strong>the</strong> wet period. In general, <strong>the</strong> lapse<br />

rate is greater below <strong>the</strong> lakes th<strong>an</strong> above for both se<strong>as</strong>ons, with greatest diurnal variability during<br />

<strong>the</strong> dry period. Fig. 14 shows <strong>the</strong> composite lapse rate for both se<strong>as</strong>ons. The average dry period<br />

zero degree iso<strong>the</strong>rm is at 5377 m above sea-level, 344 m higher th<strong>an</strong> <strong>the</strong> wet period freezing level<br />

(Fig. 14). Hence, <strong>the</strong> local valley impact on ablation rate of <strong>the</strong> glacial tongue is possibly<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Wind Speed (m/s)


signific<strong>an</strong>t, although more information on <strong>the</strong> global <strong>an</strong>d synoptic forcing is necessary to make<br />

conclusions at this time.<br />

Table 2. Average temperature recorded by iButtons during <strong>the</strong> dry <strong>an</strong>d wet periods. Note <strong>the</strong> se<strong>as</strong>onal<br />

impact of “lapse rate” on <strong>the</strong> elevation of <strong>the</strong> freezing line (0°C).<br />

Elevation (m)<br />

(a)<br />

4800<br />

4600<br />

4400<br />

4200<br />

4000<br />

3800<br />

3600<br />

3400<br />

Dry Period Diurnal Ch<strong>an</strong>ge of Temperature Profiles<br />

iButton Dry Wet Dry/<br />

Elevation (m) (°C) (°C) Wet<br />

4742 2.84 1.77 1.60<br />

4559 4.39 3.06 1.43<br />

4344 5.31 3.56 1.49<br />

4148 6.68 5.66 1.18<br />

3948 8.12 5.48 1.48<br />

3871 8.59 7.45 1.15<br />

3862 7.49 6.24 1.20<br />

3469 9.49 9.48 1.00<br />

Lapse Rate<br />

(°C/km)<br />

5.3 5.9 0.90<br />

0°C Elevation<br />

(m)<br />

5377 5033 1.07<br />

6:00 7:00 8:00 9:00 12:00 15:00<br />

16:00 17:00 18:00 21:00 0:00 3:00<br />

-5 0 5 10 15 20 25<br />

Temperature (°C)<br />

Elevation (m)<br />

273<br />

4800<br />

4600<br />

4400<br />

4200<br />

4000<br />

3800<br />

3600<br />

3400<br />

Wet Period Diurnal Ch<strong>an</strong>ge of Temperature Profiles<br />

6:00 7:00 8:00 9:00 12:00 15:00<br />

16:00 17:00 18:00 21:00 0:00 3:00<br />

-5 0 5 10 15 20 25<br />

Temperature (°C)<br />

Figure 13. Diurnal variability of near-surface air temperature profiles, “lapse rate,” for dry (a) <strong>an</strong>d wet (b) periods.<br />

Elevation (m)<br />

5400<br />

5200<br />

5000<br />

4800<br />

4600<br />

4400<br />

4200<br />

4000<br />

3800<br />

3600<br />

3400<br />

Average Temperature Profiles: Dry <strong>an</strong>d Wet Periods<br />

Dry Wet Linear (Wet) Linear (Dry)<br />

Tdry = -0.0053(Z) + 28.5<br />

R 2 = 0.95<br />

Twet = -0.0059(Z) + 29.7<br />

R 2 = 0.97<br />

0.0 2.0 4.0 6.0 8.0 10.0<br />

Air Temperature (°C)<br />

Figure 14. Composite iButton temperature profiles for dry (a) <strong>an</strong>d wet (b) periods.


Model Results<br />

We report on <strong>the</strong> (preliminary) results from <strong>the</strong> FAO <strong>an</strong>d BROOK90 model simulations of ET0<br />

<strong>an</strong>d actual ET, respectively, during <strong>the</strong> same dry <strong>an</strong>d wet composites created for <strong>the</strong><br />

meteorological data <strong>an</strong>alysis. Input from <strong>the</strong> HOBO wea<strong>the</strong>r station <strong>an</strong>d me<strong>as</strong>urements of site<br />

parameters in <strong>the</strong> field were made at one central location within <strong>the</strong> valley. The resulting<br />

hydrological components for <strong>the</strong> dry <strong>an</strong>d wet composite days are compared in Table 3 <strong>an</strong>d Figs. 15<br />

<strong>an</strong>d 16. The magnitude of ET0 w<strong>as</strong> much greater, about two times, for <strong>the</strong> dry period, which agrees<br />

with <strong>the</strong> lack of precipitation during <strong>the</strong> dry period (Table 1).<br />

Table 3 breaks down <strong>the</strong> hydrological components of ET b<strong>as</strong>ed on <strong>the</strong> BROOK90 model runs<br />

for <strong>the</strong> composite hourly dry <strong>an</strong>d wet periods. The BROOK90 estimate of actual ET from <strong>the</strong> wet<br />

period is 2.63 mm d -1 , which is 88 times that of <strong>the</strong> dry period, 0.03 mm d -1 . ET is 33% of daily<br />

precipitation for <strong>the</strong> dry period <strong>an</strong>d 37% of <strong>the</strong> daily precipitation for <strong>the</strong> wet period, which are<br />

both signific<strong>an</strong>t portions of <strong>the</strong> hydrologic bal<strong>an</strong>ce in <strong>the</strong> Ll<strong>an</strong>g<strong>an</strong>uco Valley. Tr<strong>an</strong>spiration is <strong>the</strong><br />

greatest contribution to ET, 33% for <strong>the</strong> dry <strong>an</strong>d 78% for <strong>the</strong> wet period.<br />

Table 3. BROOK90 Estimated Moisture Fluxes Dry & Wet Periods (<strong>as</strong>suming 50% true veg. cover)<br />

Period<br />

Pre*<br />

(mm)<br />

Inf<br />

(mm)<br />

SEv<br />

(mm)<br />

274<br />

Trs<br />

(mm)<br />

IEv<br />

(mm)<br />

ET<br />

(mm)<br />

Dry 0.09 0.09 0.02 0.01 0.00 0.03<br />

% of<br />

Precip.<br />

100 99 22 11 0 33<br />

Wet 7.06 6.82 0.35 2.06 0.22 2.63<br />

% of<br />

Precip.<br />

100 97 5 29 3 37<br />

Wet/Dry 78 76 18 206 n/a 88<br />

*Pre = precipitation; Inf = infiltration; SEv = surface evaporation; Trs = tr<strong>an</strong>spiration;<br />

IEv = intercepted evaporation; ET = evapotr<strong>an</strong>spiration<br />

Liquid Water (mm)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

ET: Dry <strong>an</strong>d Wet Period (+ = loss from surface)<br />

Dry Wet DryRef WetRef<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

8:00<br />

10:00<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

Time of Day (hour)<br />

Figure 15. Penm<strong>an</strong>-Monteith FAO <strong>an</strong>d BROOK90 modeled ET for composite diurnal cycle for dry (a) <strong>an</strong>d<br />

wet (b) periods.


(a)<br />

Liquid Water (mm)<br />

Dry Period: Modeled Hydrological Components<br />

Precip. Soil Evap. Tr<strong>an</strong>spiration Intercepted Evap. Infiltration<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

Figure 16. BROOK90 modeled components of ET for composite diurnal cycle for dry (a) <strong>an</strong>d wet<br />

(b) periods.<br />

DISCUSSION<br />

Time of Day (hour)<br />

8:00<br />

10:00<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00<br />

Field <strong>an</strong>d model results were interpreted <strong>an</strong>d compared to explain <strong>the</strong> processes controlling ET<br />

within a pro-glacial valley.<br />

Meteorological Me<strong>as</strong>urements<br />

The new sensor network initiated by this project permitted high resolution, discrete monitoring<br />

of air temperature at different elevations within a pro-glacial valley. Calibration of iButton<br />

temperature logger with <strong>the</strong> shielded temperature sensor on <strong>the</strong> HOBO wea<strong>the</strong>r station suggested a<br />

strong correlation, r 2 =0.988, <strong>an</strong>d this gave us high confidence in our interpretation of <strong>the</strong><br />

me<strong>as</strong>urements. As expected, <strong>an</strong>alysis of hourly data from 2005 revealed evidence for a strong<br />

diurnal <strong>an</strong>d wet versus dry se<strong>as</strong>onal dependence of meteorological forcing within <strong>the</strong> valley. The<br />

dry/wet ratio of 1.27 for daily insolation is smaller th<strong>an</strong> we expected given <strong>the</strong> absence of<br />

precipitation during <strong>the</strong> dry period <strong>an</strong>d much high precipitation total for <strong>the</strong> wet period. The<br />

hourly composite of precipitation for <strong>the</strong> wet period (Fig. 9) clearly shows <strong>the</strong> nocturnal tendency<br />

for precipitation <strong>an</strong>d hence convective cloud formation at night, <strong>an</strong>d strong solar forcing during<br />

daylight hours.<br />

However, solar forcing w<strong>as</strong> surprisingly similar for both <strong>the</strong> dry <strong>an</strong>d wet periods, since most<br />

convection (rainfall) during <strong>the</strong> wet se<strong>as</strong>on occurs <strong>an</strong> hour or two prior to sunset <strong>an</strong>d dissipates<br />

prior to sunrise. This strong diurnal convective cycle coupled with a persistent up-valley wind <strong>an</strong>d<br />

a strong daytime lapse rate during <strong>the</strong> wet se<strong>as</strong>on suggests <strong>the</strong> plausible influence of daytime<br />

surface heating west <strong>an</strong>d down slope of <strong>the</strong> pro-glacial valley. Fur<strong>the</strong>rmore, since clouds<br />

accomp<strong>an</strong>y precipitation, we concluded that nocturnal cloud cover strongly influences<br />

interse<strong>as</strong>onal <strong>an</strong>d diurnal cycles of net radiation. Fur<strong>the</strong>rmore, this may establish a connection<br />

between cloud cover (<strong>an</strong>d precipitation) <strong>an</strong>d <strong>an</strong>thropogenic development in <strong>the</strong> S<strong>an</strong>ta River valley,<br />

which receives most of its water from pro-glacial valleys along <strong>the</strong> western Cordillera Bl<strong>an</strong>ca,<br />

such <strong>as</strong> Ll<strong>an</strong>g<strong>an</strong>uco (Fig. 1). As demonstrated by Vuille et al. (2003), it is <strong>the</strong> western slope of <strong>the</strong><br />

Cordillera Bl<strong>an</strong>ca that is experiencing <strong>the</strong> highest rate of warming b<strong>as</strong>ed on wea<strong>the</strong>r station<br />

records. If this warming trend continues, we would expect to find drier <strong>an</strong>d stronger up valley<br />

winds during both se<strong>as</strong>ons. Forced by orographic uplift of <strong>the</strong> western slope, <strong>the</strong>se near-surface<br />

winds would promote stronger nocturnal convection during <strong>the</strong> wet se<strong>as</strong>on.<br />

This is one of our arguments for continued exp<strong>an</strong>sion of ground-b<strong>as</strong>ed meteorological networks<br />

within tropical proglacial valleys. We are also interested in more accurately estimating <strong>the</strong><br />

evapotr<strong>an</strong>spiration rate within <strong>the</strong> valleys. Our model results are preliminary <strong>an</strong>d we will require<br />

additional field me<strong>as</strong>urements to verify <strong>an</strong>d more accurately initiate <strong>the</strong> BROOK90 model.<br />

Liquid Water (mm)<br />

275<br />

Wet Period: Modeled Hydrological Components<br />

Precip.<br />

1<br />

Soil Evap. Tr<strong>an</strong>spiration Intercepted Evap. Infiltration<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

0:00<br />

2:00<br />

4:00<br />

6:00<br />

Time of Day (hour)<br />

8:00<br />

10:00<br />

12:00<br />

14:00<br />

16:00<br />

18:00<br />

20:00<br />

22:00


Import<strong>an</strong>ce of diurnal cycles in <strong>the</strong> Tropics<br />

Diurnal variation is related to <strong>the</strong> large <strong>an</strong>d well-defined cycle in solar heating during a 24-hour<br />

period, <strong>an</strong>d it represents one of <strong>the</strong> most fundamental components accounting for <strong>the</strong> variability of<br />

<strong>the</strong> climate system. Numerous observation studies have documented diurnal variation of deep<br />

convection, precipitation, cloudiness, <strong>an</strong>d outgoing longwave radiation over <strong>the</strong> tropics (e.g., Gray<br />

<strong>an</strong>d Jacobson, 1977; Duvel <strong>an</strong>d K<strong>an</strong>del, 1985; Hendon <strong>an</strong>d Woodberry, 1993; Chen <strong>an</strong>d Houze,<br />

1997; Y<strong>an</strong>g <strong>an</strong>d Slingo, 2001; Nesbitt <strong>an</strong>d Zipser, 2003; among m<strong>an</strong>y o<strong>the</strong>rs). Mapes et al. (2003)<br />

observed <strong>an</strong> afternoon maximum rainfall over most of South <strong>an</strong>d Central America that is typically<br />

composed of relatively small convective cloud systems. Fur<strong>the</strong>rmore, Mapes et al. found a<br />

nocturnal maximum of rainfall over some large valleys in <strong>the</strong> Andes, <strong>an</strong>d this would include <strong>the</strong><br />

larger h<strong>an</strong>ging valleys, such <strong>as</strong> Ll<strong>an</strong>g<strong>an</strong>uco. Poveda et al. (2005) use hourly records from 51 rain<br />

gages in <strong>the</strong> Tropical Andes of Columbia to find clear diurnal cycles in precipitation, with minima<br />

between 0900 <strong>an</strong>d 1100 local time, <strong>an</strong>d nocturnal maxima on <strong>the</strong> western fl<strong>an</strong>k of <strong>the</strong> Central<br />

Andes. In addition, Porveda concluded no relation between <strong>the</strong> timing of <strong>the</strong> strong se<strong>as</strong>onal<br />

variability of rainfall maxima <strong>an</strong>d elevation.<br />

Bendix et al. (2006) used K-b<strong>an</strong>d rain-radar to study convection in a valley with e<strong>as</strong>t-west<br />

orientation between <strong>the</strong> sou<strong>the</strong>rn Equadore<strong>an</strong> Andes <strong>an</strong>d <strong>the</strong> Amazon b<strong>as</strong>in. Results revealed that a<br />

great portion of rainfall is of stratiform character, <strong>an</strong>d discovered <strong>the</strong> existence of embedded<br />

convection <strong>an</strong>d/or showers produced by local heating for <strong>the</strong> overall amount of rainfall. Bendix et<br />

al. (2006) fur<strong>the</strong>r suggests that cold air drainage flow from <strong>the</strong> Andes <strong>an</strong>d low-level confluence<br />

due to <strong>the</strong> concavity of <strong>the</strong> Ande<strong>an</strong> chain in this area leads to convective instability in <strong>the</strong><br />

nocturnal Amazoni<strong>an</strong> boundary layer, which is extended to <strong>the</strong> e<strong>as</strong>t-west oriented Ande<strong>an</strong> valley<br />

<strong>the</strong> predomin<strong>an</strong>t e<strong>as</strong>terlies in <strong>the</strong> mid-troposphere. Rain clouds with at le<strong>as</strong>t embedded shallow<br />

convection c<strong>an</strong> overflow <strong>the</strong> bordering ridges of <strong>the</strong> S<strong>an</strong> Fr<strong>an</strong>cisco valley providing rains of higher<br />

intensity at <strong>the</strong> ECSF research station. On <strong>the</strong> contrary, Bendix et al. found that afternoon<br />

convective precipitation c<strong>an</strong> be caused by locally induced <strong>the</strong>rmal convection at <strong>the</strong> bordering<br />

slopes (up-slope breeze system) where <strong>the</strong> ECSF station profits from precipitation off <strong>the</strong> edge of<br />

<strong>the</strong>se local cells due to <strong>the</strong> narrow valley. Recent remote sensing studies demonstrate pronounced<br />

diurnal variability of tropical rainfall intensity at synoptic <strong>an</strong>d regional scales (Sorooshi<strong>an</strong> et al.,<br />

2002; Bowm<strong>an</strong> et al., 2005). Hence, evidence from various scales of observation suggest a strong<br />

influence of diurnal cycles in tropical regions, including <strong>the</strong> Andes Mountains.<br />

CONCLUSION<br />

From our meteorological <strong>an</strong>d model results, we make <strong>the</strong> following conclusions for this<br />

proglacial valley. Additional years <strong>an</strong>d observations of nearby valleys are <strong>an</strong>ticipated. It is<br />

import<strong>an</strong>t to note that we are using <strong>the</strong> term lapse rate in <strong>an</strong> unconventional sense, not <strong>the</strong> free<br />

atmosphere, but ra<strong>the</strong>r <strong>the</strong> effects within <strong>the</strong> surface boundary layer up <strong>the</strong> valley floor. First, <strong>the</strong><br />

steepest lapse rates occur below <strong>the</strong> lakes for both wet <strong>an</strong>d dry se<strong>as</strong>ons, <strong>an</strong>d <strong>the</strong> lapse rate is<br />

signific<strong>an</strong>tly smaller above <strong>the</strong> glacial lakes. We partially attribute this elevation effect to <strong>the</strong><br />

heterogeneous vegetation <strong>an</strong>d topographic characteristics of pro-glacial valleys. Fru<strong>the</strong>rmroe, <strong>the</strong><br />

dry se<strong>as</strong>on nocturnal inversion below <strong>the</strong> lakes is not evident during wet se<strong>as</strong>on. Most<br />

precipitation (<strong>an</strong>d cloud cover) occurs between sunset <strong>an</strong>d sunrise during <strong>the</strong> wet se<strong>as</strong>on, hence<br />

insolation at <strong>the</strong> ground is strong during both se<strong>as</strong>ons. Up-valley winds dominate during sunlight<br />

hours in both wet <strong>an</strong>d dry se<strong>as</strong>ons, but abruptly shift to katabatic winds during <strong>the</strong> dry se<strong>as</strong>on at<br />

sunset. The combination of valley winds <strong>an</strong>d steep lapse rate below <strong>the</strong> lakes suggest local warm<br />

air advection within <strong>the</strong> surface boundary layer will contribute to evapotr<strong>an</strong>spiration <strong>an</strong>d lower<br />

glacial melt. Fur<strong>the</strong>rmore, <strong>the</strong> up-slope winds may enh<strong>an</strong>ce convective precipitation, particularly<br />

during <strong>the</strong> wet se<strong>as</strong>on after sunset. The BROOK90 output suggests that tr<strong>an</strong>spiration is a<br />

signific<strong>an</strong>t source of ET when soil moisture is available during <strong>the</strong> wet se<strong>as</strong>on. The model results<br />

suggest that <strong>the</strong> predomin<strong>an</strong>ce of cloud-free daylight conditions <strong>an</strong>d relatively high solar input<br />

enh<strong>an</strong>ce ET during <strong>the</strong> wet se<strong>as</strong>on. ET w<strong>as</strong> insignific<strong>an</strong>t throughout <strong>the</strong> dry se<strong>as</strong>on.<br />

276


Future Work<br />

We are installing additional humidity, wind, solar, <strong>an</strong>d soil moisture sensors to evaluate spatial<br />

variation of ET. We pl<strong>an</strong> to make direct me<strong>as</strong>urements of ET using lysimeters <strong>an</strong>d evaporation<br />

p<strong>an</strong>s, <strong>an</strong>d to use BROOK90 to perform a formal sensitivity <strong>an</strong>alysis of ET to meteorological<br />

forcing. In addition to site-specific parameters, <strong>the</strong> meteorological input includes solar radiation,<br />

air temperature, vapor pressure, wind speed, <strong>an</strong>d precipitation. Model output includes various<br />

hydrological bal<strong>an</strong>ce components of <strong>the</strong> vegetation c<strong>an</strong>opy <strong>an</strong>d soil surface. We pl<strong>an</strong> to evaluate<br />

<strong>the</strong> reliability <strong>an</strong>d realism of estimated ET <strong>an</strong>d to determine <strong>the</strong> sensitivity of ET to variables<br />

me<strong>as</strong>ured at <strong>the</strong> HOBO site in <strong>the</strong> Ll<strong>an</strong>g<strong>an</strong>uco valley. Future modeling will focus on more<br />

accurate me<strong>as</strong>urements of parameters for BROOK90 <strong>an</strong>d <strong>the</strong> elevation effects on ET <strong>an</strong>d total ET<br />

for <strong>the</strong> valley.<br />

ACKNOWLEDGEMENTS<br />

Funding for this project w<strong>as</strong> provided by The Ohio State University, Department of Geography<br />

<strong>an</strong>d Office of International Affairs. We are grateful for <strong>the</strong> collaboration with Peruvi<strong>an</strong> Institute of<br />

Natural Resources (INRENA), especially <strong>the</strong> timely collection <strong>an</strong>d tr<strong>an</strong>smission of data by Jesús<br />

Gómez, INRENA-Huaraz, Perú.<br />

REFERENCES<br />

Allen, R.G., Jensen, M.E., Wright, J.L., <strong>an</strong>d Burm<strong>an</strong>, R.D. 1989. Operational Estimates of<br />

Reference Evapotr<strong>an</strong>spiration. Agronomy Journal, 81: 650–662.<br />

Allen, R.G., Pereira, L.S., Raes, D., <strong>an</strong>d Smith, M. 1998. Crop evapotr<strong>an</strong>spiration-Guidelines for<br />

computing crop water requirements, FAO irrigation <strong>an</strong>d drainage paper 56, FAO. ISBN 92-5-<br />

104219-5.<br />

Ames, A., 1998. A documentation of glacier tongue variations <strong>an</strong>d lake developments in <strong>the</strong><br />

Cordillera Bl<strong>an</strong>ca. Zeitschrift fur Gletscherkunde und Glazialgeologie, 34(1): 1–26.<br />

Barry, R. <strong>an</strong>d Seimon, A., 2000. Research for mountain area development: Climatic fluctuations in<br />

<strong>the</strong> mountains of <strong>the</strong> Americ<strong>as</strong> <strong>an</strong>d <strong>the</strong>ir signific<strong>an</strong>ce. Ambio, 29(7): 364–370.<br />

Bendix, J., Rollenbeck, R., <strong>an</strong>d Reudenbach, C., 2006. Diurnal patterns of rainfall in a tropical<br />

Ande<strong>an</strong> valley of sou<strong>the</strong>rn Ecuador <strong>as</strong> seen by a vertically pointing K-b<strong>an</strong>d Doppler radar.<br />

International Journal of Climatology, Volume 26, Issue 6, 829–846.<br />

Bowm<strong>an</strong>, KP, Collier, JC, North, GR, Wu, Q, Ha, E, Hardin, J, 2005. Diurnal cycle of tropical<br />

precipitation in Tropical Rainfall Me<strong>as</strong>uring Mission (TRMM) satellite <strong>an</strong>d oce<strong>an</strong> buoy rain<br />

gauge data. Journal of Geophysical Research, 110(d21): D21104-.<br />

Caballero, Y., Chevallier, P., Gallaire, R. <strong>an</strong>d Pillco, R., 2004. Flow modelling in a high mountain<br />

valley equipped with hydropower pl<strong>an</strong>ts: Rio zongo valley, Cordillera Real, Bolivia.<br />

Hydrological Processes, 18(5): 939–957.<br />

Caballero, Y., Jomelli, V., Chevallier, P. <strong>an</strong>d Ribstein, P., 2002. Hydrological characteristics of<br />

slope deposits in high tropical mountains (Cordillera Real, Bolivia). Catena, 47(2): 101–116.<br />

Chen, S. S., <strong>an</strong>d R. A. Houze Jr., 1997: Diurnal variation <strong>an</strong>d life-cycle of deep convective<br />

systems over <strong>the</strong> tropical Pacific warm pool. Quart. J. Roy. Meteor. Soc., 123, 357–388.<br />

Duvel, J. P., <strong>an</strong>d R. S. K<strong>an</strong>del, 1985: Regional-scale diurnal variations of outgoing infrared<br />

radiation observed by METEOSAT. J. Climate Appl. Meteor., 24, 335–349.<br />

Federer, C.A. 1979. A soil–pl<strong>an</strong>t–atmosphere model for tr<strong>an</strong>spiration <strong>an</strong>d availability of soil<br />

water. Water Resour Res., 15:555–562.<br />

Federer, C.A. 1995. BROOK90. A Simulation Model for Evaporation, Soil Water, <strong>an</strong>d<br />

Streamflow. Version 3.1 Computer Freeware <strong>an</strong>d Documentation. USDA Forest Service, P.O.<br />

Box 640, Durham, N.H. 03824.<br />

Federer, C.A., C. Vörösmarty, <strong>an</strong>d B. Fekete. 1996. Intercomparison of methods for calculating<br />

potential evaporation in regional <strong>an</strong>d global water bal<strong>an</strong>ce models. Water Resour Res., 32:<br />

2315–2321.<br />

277


Federer, C.A., C. Vörösmarty, <strong>an</strong>d B. Fekete. 2003. Sensitivity of <strong>an</strong>nual evaporation to soil <strong>an</strong>d<br />

root properties in two models of contr<strong>as</strong>ting complexity. J Hydrometeorology, 4:1276–1290.<br />

Fr<strong>an</strong>cou, B., Ribstein, P., Saravia, R. <strong>an</strong>d Tiriau, E., 1995. Monthly bal<strong>an</strong>ce <strong>an</strong>d water discharge of<br />

<strong>an</strong> inter-tropical glacier: Zongo Glacier, Cordillera Real, Bolivia, 16S. Journal of Glaciology,<br />

41(137): 61–67.<br />

Georges, C. <strong>an</strong>d K<strong>as</strong>er, G., 2002. Ventilated <strong>an</strong>d unventilated air temperature me<strong>as</strong>urements for<br />

glacier-climate studies on a tropical high mountain site. Journal of Geophysical Research-<br />

Atmospheres, 107(D24): art. no.-4775.<br />

Gray, W. M., <strong>an</strong>d R. W. Jacobson, 1977: The diurnal march of convective cloudiness over <strong>the</strong><br />

Americ<strong>as</strong>. Mon. Wea. Rev., 105, 1171–1188.Fr<strong>an</strong>cou, B., Ribstein, P. <strong>an</strong>d Pouyard, B., 1997.<br />

The melting of tropical glaciers. Recherche(302): 34–37.<br />

Hardy, D. R., M. Vuille, C. Braun, F. Keimig, <strong>an</strong>d R. S. Bradley, Annual <strong>an</strong>d daily meteorological<br />

cycles at high altitude on a tropical mountain, Bull. Am. Meteorol. Soc., 79(9), 1899–1913,<br />

1998.<br />

H<strong>as</strong>tenrath, S. <strong>an</strong>d Kruss, P.D., 1992. The dramatic retreat of Mount Kenya's glaciers between<br />

1963 <strong>an</strong>d 1987: greenhouse forcing. Annals of Glaciology, 16: 127–133.Hendon, H. H., <strong>an</strong>d K.<br />

Woodberry, 1993: The diurnal cycle of tropical convection. J. Geophys. Res., 98, 16,623–<br />

16,637.<br />

Hinckley, T.M., J.P. L<strong>as</strong>soie, <strong>an</strong>d S.W. Running. 1978. Temporal <strong>an</strong>d spatial variations in <strong>the</strong><br />

water status of forest trees. Forest Science Monograph, 20, 72 p.<br />

Jackson, R.B. J. C<strong>an</strong>adell, J.R. Ehleringer, H.A. Mooney, O.E. Sala, <strong>an</strong>d E.-D. Schulze. 1996. A<br />

global <strong>an</strong>alysis of root distributions for terrestrial biomes. Oecologia, 108:389–411.<br />

Juen, I., K<strong>as</strong>er, G. <strong>an</strong>d Georges, C., 2005. Modelling observed <strong>an</strong>d future runoff from a glacierized<br />

tropical catchment (Cordillera Bl<strong>an</strong>ca, Peru). Global & Pl<strong>an</strong>etary Ch<strong>an</strong>ge, submitted.<br />

K<strong>as</strong>er, G., Juen, I., Georges, C., Gomez, J. <strong>an</strong>d Tamayo, W., 2003. The impact of glaciers on <strong>the</strong><br />

runoff <strong>an</strong>d <strong>the</strong> reconstruction of m<strong>as</strong>s bal<strong>an</strong>ce history from hydrological data in <strong>the</strong> tropical<br />

Cordillera Bl<strong>an</strong>ca, Peru. Journal of Hydrology, 282(1–4): 130–144.<br />

Mapes, Bri<strong>an</strong> E., Warner, Thom<strong>as</strong> T., Xu, Mei, Negri, Andrew J., 2003. Diurnal Patterns of<br />

Rainfall in Northwestern South America. Part I: Observations <strong>an</strong>d Context. Monthly Wea<strong>the</strong>r<br />

Review, 131: 799–812.<br />

Mark, B.G., McKenzie, J.M. <strong>an</strong>d Gomez, J., 2005. Hydrochemical evaluation of ch<strong>an</strong>ging glacier<br />

meltwater contribution to stream discharge: Callejon de Huayl<strong>as</strong>, Peru. Hydrological Sciences<br />

Journal = Journal des Sciences Hydrologiques, 50(6).<br />

Mark, B.G. <strong>an</strong>d Seltzer, G.O., 2003. Tropical glacier meltwater contribution to stream discharge: a<br />

c<strong>as</strong>e study in <strong>the</strong> Cordillera Bl<strong>an</strong>ca, Peru. Journal of Glaciology, 49(165): 271–281.<br />

Mark, B.G. <strong>an</strong>d Seltzer, G.O., 2005. Recent glacial recession in <strong>the</strong> Cordillera Bl<strong>an</strong>ca, Peru (AD<br />

1962–1999). Quaternary Science Reviews, 24: 2265–2280.<br />

McNulty, B.A., Farber, D.L., Wallace, G.S., Lopez, R. <strong>an</strong>d Palacios, O., 1998. Role of plate<br />

kinematics <strong>an</strong>d plate-slip-vector partitioning in continental magmatic arcs: Evidence from <strong>the</strong><br />

Cordillera Bl<strong>an</strong>ca, Peru. Geology, 26(9): 827–830.<br />

Nesbitt, S. W., <strong>an</strong>d E. J. Zipser, 2003: The diurnal cycle of rainfall <strong>an</strong>d convective intensity<br />

according to three years of TRMM me<strong>as</strong>urements. J. Climate, 16, 1456–1475.<br />

Penm<strong>an</strong>, H.L., 1948. Natural evaporation from open water, bare soil <strong>an</strong>d gr<strong>as</strong>s. Proc. Roy. Soc.<br />

London. A193, 120–146.<br />

Penm<strong>an</strong>, H.L., 1963. Vegetation <strong>an</strong>d hydrology. Tech. Comm. No. 53. Commonwealth Bureau of<br />

Soils, Harpenden, Engl<strong>an</strong>d. 125 pp.<br />

Poveda, Germán, Mesa, Oscar J., Salazar, Luis F., Ari<strong>as</strong>, Paola A., Moreno, Hernán A., Vieira,<br />

Sara C., Agudelo, Paula A., Toro, Vladimir G., Alvarez, J. Felipe, 2005. The Diurnal Cycle of<br />

Precipitation in <strong>the</strong> Tropical Andes of Colombia. Monthly Wea<strong>the</strong>r Review, 133: 228–240.<br />

Ribstein, P., Tiriau, E., Fr<strong>an</strong>cou, B. <strong>an</strong>d Saravia, R., 1995. Tropical Climate <strong>an</strong>d Glacier<br />

Hydrology: A C<strong>as</strong>e-Study in Bolivia. Journal of Hydrology, 165(1–4): 221–234.<br />

Shuttleworth, W.J., <strong>an</strong>d R.J. Gurney. 1990. The <strong>the</strong>oretical relationship between foliage<br />

temperature <strong>an</strong>d c<strong>an</strong>opy resist<strong>an</strong>ce in sparse crops. Quart J Royal Meteorol Soc., 116: 497–<br />

519.<br />

278


Shuttleworth, W.J., <strong>an</strong>d J.S. Wallace. 1985. Evaporation from sparse crops: An energy<br />

combination <strong>the</strong>ory. Quart J Royal Meteorol Soc., 111: 839–855.<br />

Sorooshi<strong>an</strong>, S., Gao, X., Hsu, K., Maddox, R. A., Hong, Y., Gupta, H. V., Imam, B. Diurnal<br />

Variability of Tropical Rainfall Retrieved from Combined GOES <strong>an</strong>d TRMM Satellite<br />

Information. Journal of Climate, 2002 15: 983–1001.<br />

Vuille, M., Bradley, R.S., Werner, M. <strong>an</strong>d Keimig, F., 2003. 20th century climate ch<strong>an</strong>ge in <strong>the</strong><br />

tropical Andes: Observations <strong>an</strong>d model results. Climatic Ch<strong>an</strong>ge, 59(1–2): 75–99.<br />

Wagnon, P., Ribstein, P., K<strong>as</strong>er, G. <strong>an</strong>d Berton, P., 1999. Energy bal<strong>an</strong>ce <strong>an</strong>d runoff se<strong>as</strong>onality of<br />

a Bolivi<strong>an</strong> glacier. Global <strong>an</strong>d Pl<strong>an</strong>etary Ch<strong>an</strong>ge, 22(1–4): 49–58.<br />

Wagnon, P., Ribstein, P., Schuler, T. <strong>an</strong>d Fr<strong>an</strong>cou, B., 1998. Flow Separation On Zongo Glacier,<br />

Cordillera Real, Bolivia. Hydrological Processes, 12(12): 1911–1926.<br />

Xu, C.-Y. <strong>an</strong>d Singh, V.P. 2002. Cross comparison of empirical equations for calculating potential<br />

evapotr<strong>an</strong>spiration with data from Switzerl<strong>an</strong>d. Wat. Resour. M<strong>an</strong>ag., 16, 197–219.<br />

Xu, C.-Y. <strong>an</strong>d Singh, V.P. 1998. Dependence of evaporation on meteorological variables at<br />

different time-scales <strong>an</strong>d intercomparison of estimation methods. Hydrol. Process., 12, 429–<br />

442.<br />

Y<strong>an</strong>g, Gui-Ying, Slingo, Julia, 2001. The Diurnal Cycle in <strong>the</strong> Tropics. Monthly Wea<strong>the</strong>r Review,<br />

129: 784–801<br />

279


This page is intentionally bl<strong>an</strong>k.<br />

280


<strong>Snow</strong> <strong>an</strong>d Periglacial<br />

Processes Poster<br />

281


This page is intentionally bl<strong>an</strong>k.<br />

282


283<br />

63 rd EASTERN SNOW CONFERENCE<br />

Newark, Delaware USA 2006<br />

A Treatise on <strong>the</strong> Preponder<strong>an</strong>ce of Designs Over Historic <strong>an</strong>d<br />

Me<strong>as</strong>ured <strong>Snow</strong>falls, or No Two <strong>Snow</strong>flakes Are Alike:<br />

Considerations About <strong>the</strong> Formation of <strong>Snow</strong>flakes<br />

<strong>an</strong>d <strong>the</strong> Possible Numbers <strong>an</strong>d Shapes of <strong>Snow</strong>flakes<br />

MARK PILIPSKI AND JACOB DIESSNER PILIPSKI<br />

"The re<strong>as</strong>on for this macroscopic six-fold symmetry, which so exercised<br />

Kepler, is now readily explained in terms of <strong>the</strong> atomic architecture of ice, but <strong>the</strong><br />

variety <strong>an</strong>d ch<strong>an</strong>ge of <strong>the</strong> crystal shape h<strong>as</strong> always been a puzzle <strong>an</strong>d still awaits<br />

a complete expl<strong>an</strong>ation."<br />

— B.J.M<strong>as</strong>on in A Commentary on Kepler's essay ON THE SIX-CORNERED SNOWFLAKE<br />

ABSTRACT<br />

We use a generalized geometric model for <strong>the</strong> shape of snowflakes to approximate <strong>the</strong> finite<br />

number of possible unique snowflake configurations. Using historic geologic <strong>an</strong>d meteorological<br />

data we approximate <strong>the</strong> number of snowflakes created throughout history. Assuming <strong>the</strong><br />

development of snowflakes to be subject to guided r<strong>an</strong>dom molecular movements, we calculate<br />

<strong>the</strong> probability of duplicate configurations. The factorial estimate of possible snowflake<br />

configurations overwhelms <strong>the</strong> countable linear qu<strong>an</strong>tity of snowflakes. We show, even with very<br />

conservative estimates, that except for extremely small snowflake fragments, no two snowflakes<br />

are identical. Thus, at <strong>the</strong> macro-optical scale, no two snowflakes are alike.<br />

PREMISE<br />

Simply stated, we determine from a general model for snowflakes how m<strong>an</strong>y possible<br />

snowflake configurations may exist. Using geologic <strong>an</strong>d meteorological data we approximate <strong>the</strong><br />

number of snowflakes created throughout history. If <strong>the</strong> total number of actual snowflakes<br />

exceeds <strong>the</strong> possible number of snowflake configurations, we deduce that at le<strong>as</strong>t two snowflakes<br />

throughout history have been identical. If this is so, we will be able to determine how often<br />

identical snowflakes fall. If <strong>the</strong> total number of possible snowflake configurations greatly exceeds<br />

<strong>the</strong> total number of snowflakes that have fallen throughout history, we must concede that it is<br />

improbable that two identical snowflakes have ever fallen.<br />

To evaluate this premise, we sought to <strong>an</strong>swer <strong>the</strong>se two questions:<br />

1. How m<strong>an</strong>y possible snowflake configurations are <strong>the</strong>re?<br />

2. How m<strong>an</strong>y snowflakes have fallen?<br />

The <strong>an</strong>swer to <strong>the</strong> first question will be <strong>the</strong> result of our modeling. We develop a model that<br />

approximates <strong>the</strong> shape of a snowflake. We limit our efforts to one form of snowflake; <strong>the</strong><br />

dendritic plate.


The <strong>an</strong>swer to <strong>the</strong> second question requires approximations. We chose such to maximize our<br />

result. Thus, for <strong>the</strong> total number of snowflakes that have fallen, our approximation will be higher<br />

th<strong>an</strong> <strong>the</strong> historical value.<br />

To determine how m<strong>an</strong>y snowflakes have fallen we need to know, how much snow h<strong>as</strong> fallen<br />

<strong>an</strong>d <strong>the</strong> average size of a snowflake. Me<strong>as</strong>urements of snowfall are made locally not globally. Any<br />

amount of snow that we project to have fallen on <strong>the</strong> earth will be, at best, a guess.<br />

SNOWFLAKE MODEL<br />

Combinations <strong>an</strong>d configurations<br />

This is a model for a flat hexagonal snowflake <strong>an</strong>d incorporates dendritic plates. Several<br />

simplifications make this model workable <strong>an</strong>d yet remain consistent with what is known about<br />

real snowflakes. The model begins, <strong>as</strong> <strong>an</strong>y good model, with actual snowflakes. (3)<br />

Our model applies to dendritic plates:<br />

We <strong>as</strong>sume snowflakes have:<br />

Hexagonal radial symmetry (six-sided snowflakes),<br />

Two-dimensional configuration (flat snowflakes), <strong>an</strong>d<br />

The simplest unit of <strong>an</strong> ice crystal is a hexagonal ring composed of six water molecules.<br />

We know that real snowflakes have several documented forms <strong>an</strong>d, although, <strong>the</strong>se may be<br />

hexagonal in nature, <strong>the</strong>y are not strictly two-dimensional. They have thickness <strong>an</strong>d <strong>the</strong> b<strong>as</strong>ic<br />

building blocks of <strong>an</strong>y ice crystal may be far more complex th<strong>an</strong> a ring of only six water<br />

molecules. Thus, <strong>an</strong>y model we develop will underestimate <strong>the</strong> number of possible snowflake<br />

configurations.<br />

Each of <strong>the</strong> above <strong>as</strong>sumptions carries several corollaries that shape our model. Hexagonal<br />

radial symmetry implies that:<br />

a. Each of <strong>the</strong> six arms are identical to each o<strong>the</strong>r.<br />

i. One sixth of <strong>the</strong> total number of units is used to compose each of <strong>the</strong><br />

six arms.<br />

b. Each arm is bilaterally symmetric.<br />

(It is <strong>the</strong> same on each side of its radial axis.)<br />

i. There is a centerline of units. Assume <strong>the</strong> centerline (radial axis) to be<br />

<strong>the</strong> longest line of each arm.<br />

ii. The sidelines (parallel to <strong>the</strong> centerline) are paired, one on ei<strong>the</strong>r side<br />

of <strong>the</strong> centerline.<br />

iii. The sidelines are made up of m<strong>an</strong>y segments of lines each shorter th<strong>an</strong><br />

<strong>the</strong> centerline.<br />

iv. The length of each sideline must be at le<strong>as</strong>t three units shorter th<strong>an</strong> <strong>the</strong><br />

next sideline closer to <strong>the</strong> centerline <strong>an</strong>d at le<strong>as</strong>t three units longer th<strong>an</strong><br />

<strong>the</strong> next sideline far<strong>the</strong>r from <strong>the</strong> centerline.<br />

284


----- --------------<br />

------- ------------ ---------- sideline (3 segments)<br />

------- ----- - ----------- --------- sideline<br />

--------------------------------------------------- sideline<br />

Radial axis --------------------------------------------------------------- one of six arms<br />

---------------------------------------------------<br />

------- ----- - ----------- ---------<br />

------- ------------ ----------<br />

----- --------------<br />

We know that <strong>the</strong> most common form of ice is a tetrahedral structure. Idealized hexagonal rings<br />

are found within <strong>the</strong> tetrahedral form. Close examination of <strong>the</strong> b<strong>as</strong>al pl<strong>an</strong>e of <strong>an</strong> ice crystal<br />

reveals it is composed of a tiling of hexagonal plates each with a diameter of 5.5 <strong>an</strong>gstroms, 4.5<br />

<strong>an</strong>gstroms across <strong>the</strong> tiling diameter or minor chord <strong>an</strong>d 2.75 <strong>an</strong>gstroms between adjacent oxygen<br />

atoms. (1)(5)<br />

Using a simple idealized hexagonal ice crystal <strong>as</strong> a building block, we c<strong>an</strong> approximate how<br />

m<strong>an</strong>y such blocks or hexagonal tiles must accumulate to form a single axis arm of a snowflake.<br />

Ordinary snowflakes vary in size from 0.05 cm. to 1.8 cm. in diameter. A re<strong>as</strong>onable midr<strong>an</strong>ge<br />

value for <strong>the</strong> average radius of a snowflake is 0.66 cm. Thus, a radius axis of <strong>an</strong> average<br />

snowflake is<br />

285


6.6 × 10 –1 cm. (radius axis length) = 1.5 × 10 7 hex units<br />

4.5 × 10 –8 cm. (hexagonal unit length)<br />

1.5 × 10 7 units is <strong>the</strong> number of hexagonal units along <strong>an</strong>y one axis.<br />

The model calls for <strong>the</strong> axis <strong>an</strong>d <strong>an</strong>y of <strong>the</strong> sidearms to be at most filled with hexagonal<br />

units <strong>an</strong>d in most c<strong>as</strong>es present with gaps in <strong>the</strong> linear tiling. We may think of <strong>an</strong>y axis or sidearm<br />

of actual hex units to be represented by a binary number (hex unit present = 1, hex unit absent =<br />

0.)<br />

A sidearm three units in length would have<br />

2 3 – 1 = 7 possible configurations of ice hex units<br />

111, 110, 101, 100, 011, 010, 001<br />

The null configuration of no hex units (000) is not possible because <strong>the</strong> resulting<br />

snowflake would lack <strong>an</strong>y cohesion along this null line.<br />

A radius axis or sidearm with 1.5 × 10 7 units would have<br />

1.5 × 10 7<br />

(2 – 1) possible permutations.<br />

For a configuration consistent with our model we have, where N= 2<br />

286<br />

1.5 × 10 7<br />

2 (N–3) –1 is <strong>the</strong> number of possible permutations for <strong>the</strong> next smaller sidearm(s). Then it follows<br />

N/3<br />

2 (N) –1 . 2 (N–3) –1 . 2 (N–6) –1 . 2 (N–9) –1 …= Π 2 (N–3X) –1<br />

X=0<br />

Equals <strong>the</strong> number of possible permutations <strong>an</strong>d combinations of all sidearms, i.e. <strong>the</strong> number<br />

of possible unique snowflakes. The multiplication of exponential qu<strong>an</strong>tities may be expressed <strong>as</strong><br />

<strong>the</strong> sum of <strong>the</strong>se exponents. We may simplify this series <strong>as</strong><br />

N/3 N/3 N/3<br />

Σ (N–3X) Σ (3X+1) 3 Σ (X) + Σ (1)<br />

0 0 0<br />

2 –1 = 2 –1 = 2 – 1 =<br />

N/3 3.8 × 10 13 10 13<br />

Π 2 (N–3X) –1 or ≈ 2 or ≈ 10<br />

X=0<br />

That’s a one followed by ten trillion zeros, <strong>as</strong> <strong>the</strong> possible number of unique snowflake<br />

configurations.


No one to date h<strong>as</strong> shown why snowflakes (here we are discussing only <strong>the</strong> formation of plates,<br />

specifically dendritic plates) demonstrate a remarkably precise hexagonal radial symmetry. The<br />

currently accepted expl<strong>an</strong>ation for this is that <strong>as</strong> snowflakes form <strong>the</strong>y travel along unique r<strong>an</strong>dom<br />

paths that expose <strong>the</strong> developing snowflake to micro environmental fluctuations. Crystalization<br />

occurs in response to <strong>the</strong>se local conditions. The nature of water molecules leads to <strong>the</strong> familiar<br />

hexagonal crystals. Thus, <strong>the</strong> unique path taken by a developing snowflake determines its final<br />

configuration. (4)(9)<br />

Ano<strong>the</strong>r possible contributing factor for <strong>the</strong> r<strong>an</strong>domness we see in snowflake formation is <strong>the</strong><br />

presence of a qu<strong>as</strong>i-liquid ph<strong>as</strong>e <strong>as</strong> part of a developing snowflake. (7)<br />

We postulate <strong>an</strong>o<strong>the</strong>r possible expl<strong>an</strong>ation for <strong>the</strong>se guided r<strong>an</strong>dom configurations of<br />

snowflakes; <strong>the</strong> development of <strong>an</strong> emerging pi-obital like electron cloud. This elaborate pi-orbital<br />

develops <strong>as</strong> <strong>the</strong> crystal grows creating sites favored for <strong>the</strong> attachment of <strong>the</strong> next water<br />

molecules. Such pi-orbitals are found in large org<strong>an</strong>ic molecules (i.e. pl<strong>an</strong>ar benzene constructs)<br />

<strong>as</strong> well <strong>as</strong> several pl<strong>an</strong>ar inorg<strong>an</strong>ic crystalline forms. We consider <strong>the</strong> formation of ice pi-orbitals,<br />

much like those of benzene molecules. The individual water (ice) hex plates may each have such a<br />

pi-orbital <strong>an</strong>d <strong>as</strong> <strong>the</strong>y are joined to form a larger plate this orbital exp<strong>an</strong>ds to ‘guide’ <strong>the</strong> future<br />

formation of <strong>the</strong> snowflake.<br />

Regardless of <strong>the</strong> exact cause for <strong>the</strong> observed symmetry, we find it to be guided (following a<br />

symmetric pattern) <strong>an</strong>d r<strong>an</strong>dom (each crystalline growth point may or may not affix a water<br />

molecule or molecular complex.<br />

The physical implications of this model lead us to claim that <strong>the</strong>re are<br />

1.5 × 10 7 hex units in a single radial axis; 3.8 × 10 13 hex units in a complete radial arm (a radial<br />

axis <strong>an</strong>d all sidearms.) There are six radial arms per each snowflake <strong>an</strong>d six water molecules for<br />

each ice hex unit. These figures give us<br />

(3.8 × 10 13 hex units per arm) × (6 arms per snowflake) × (6 H2O per hex unit)<br />

= 1.4 × 10 15 H2O molecules in a model snowflake.<br />

This is b<strong>as</strong>ed upon a two-dimensional model. Actual snowflakes, <strong>as</strong> far <strong>as</strong> <strong>the</strong>y might inform<br />

this model, have thickness. By adding several layers of parallel pl<strong>an</strong>es of ice above <strong>an</strong>d below <strong>the</strong><br />

central pl<strong>an</strong>e of this model, we see that we approach <strong>the</strong> me<strong>as</strong>ured m<strong>as</strong>s of observed snowflakes.<br />

We also appreciate that each added pl<strong>an</strong>e of ice multiples our estimate of possible snowflake<br />

configurations to beyond <strong>the</strong> already predicted <strong>as</strong>tronomical numbers.<br />

Throughout <strong>the</strong> development of this model we’ve <strong>as</strong>sumed that <strong>the</strong>re is <strong>an</strong> equivalence of <strong>the</strong><br />

probability of <strong>an</strong>y possible snowflake configuration. In o<strong>the</strong>r words, <strong>the</strong> probability of <strong>an</strong>y one<br />

snowflake configuration is <strong>the</strong> same <strong>as</strong> <strong>an</strong>y o<strong>the</strong>r snowflake configuration. This may not be a valid<br />

<strong>as</strong>sumption. Indeed, <strong>the</strong>re may be favored snowflake configurations. Thus, <strong>the</strong> final form of <strong>an</strong>y<br />

actual snowflake may not be <strong>the</strong> product of strictly r<strong>an</strong>dom development. If this is true, our<br />

estimate of all possible snowflake configurations is <strong>an</strong> over-estimate of <strong>an</strong>y real qu<strong>an</strong>tity<br />

Photographs show that some real snowflakes are <strong>as</strong>ymmetric. Whe<strong>the</strong>r this <strong>as</strong>ymmetry is<br />

evident while a snowflake develops or induced owing <strong>the</strong> capture processes, symmetry prevails.<br />

We c<strong>an</strong> argue ad nauseum about <strong>the</strong> definition of ‘alike’ or ‘identical.’ We may also accept <strong>the</strong><br />

concept of ‘for all practical purposes.’<br />

SNOWFALL<br />

There are 1.4 × 10 15 H2O molecules or 4.2 × 10 –8 gms of H2O in a model snowflake. If we<br />

approximate <strong>the</strong> average water density of fresh snow to be 12%, this is equivalent to saying that<br />

1.0 cm 3 of snow is equal to 0.12 gms of H2O.<br />

Using <strong>an</strong> estimate of total snowfall over <strong>the</strong> earth’s surface per year of 3.3 × 10 21 cm 3 (see<br />

Appendix A) <strong>an</strong>d our model, we obtain 9.6 × 10 31 model snowflakes produced worldwide each<br />

287


year.(8) If we consider data for actual snowflakes we obtain 6.6 × 10 28 actual snowflakes<br />

produced worldwide each year. (2)(6)<br />

The earth is by currently available me<strong>as</strong>ures about 4.5 × 10 9 years old. Thus, for our model we<br />

obtain 4.3 × 10 39 snowflakes produced throughout history <strong>an</strong>d for actual snowflakes we obtain 3 ×<br />

10 38 <strong>as</strong> <strong>the</strong> number of historic snowflakes. Both numbers are close to each o<strong>the</strong>r <strong>an</strong>d we use this<br />

closeness <strong>as</strong> a validation of <strong>the</strong> accuracy of <strong>the</strong> model.<br />

From our model we obtain a value of 10 13 for <strong>the</strong> number of possible snowflake configurations.<br />

(We have no evidence that <strong>the</strong>re are preferred configurations.) We also estimate that only 10 40<br />

actual snowflakes have been produced throughout <strong>the</strong> history of <strong>the</strong> earth. (Of course <strong>the</strong>re may<br />

have been great variations in se<strong>as</strong>onal snowfall throughout history. By maintaining a consistent<br />

yearly snowfall, we may over-estimate or under-estimate this value by a few orders of magnitude.<br />

With <strong>the</strong> current rate of snow production on earth it will be m<strong>an</strong>y trillions upon trillions of years<br />

before <strong>the</strong> production of two identical snowflakes may be considered a possibility. For now we<br />

may <strong>as</strong>sume that <strong>the</strong>re simply h<strong>as</strong> not been enough snow to produce two identical snowflake<br />

configurations.<br />

288


APPENDIX A<br />

Nor<strong>the</strong>rn Hemisphere Yearly <strong>Snow</strong>fall Totals (8)<br />

B<strong>an</strong>d Area Avg <strong>Snow</strong>fall <strong>Snow</strong>fall<br />

(cm 2 )(x10 16 ) (cm) (cm 3 ) (x 10 19 )<br />

90 o – 86 o<br />

86 o – 82 o<br />

82 o – 78 o<br />

78 o – 74 o<br />

74 o – 70 o<br />

70 o – 66 o<br />

66 o – 62 o<br />

62 o – 58 o<br />

58 o – 54 o<br />

54 o – 50 o<br />

50 o – 46 o<br />

46 o – 42 o<br />

42 o – 38 o<br />

38 o – 34 o<br />

34 o – 30 o<br />

30 o – 26 o<br />

0.627 2888 1.81<br />

1.866 3142 5.86<br />

3.068 3584 11.0<br />

4.212 3975 16.7<br />

5.277 3800 20.1<br />

6.233 3546 22.1<br />

7.081 3618 25.6<br />

7.766 5207 40.4<br />

8.328 3698 30.8<br />

8.708 2029 17.7<br />

8.932 589 5.26<br />

8.959 490 4.39<br />

8.740 140 1.22<br />

8.640 28 0.24<br />

8.011 30 0.24<br />

7.194 56 0.40<br />

289


REFERENCES<br />

Fr<strong>an</strong>ks, F. The Properties of Ice, pp. 115–148 in WATER: A COMPREHENSIVE TREATISE<br />

Volume 1: The Physics <strong>an</strong>d Physical Chemistry of Water. Plenum Press, New York, 1972<br />

Fujiyoshi, Y. <strong>an</strong>d K.Muramoto. The Effect of Breakup of Melting <strong>Snow</strong>flakes on <strong>the</strong> Resulting<br />

Size Distribution of Raindrops. Journal of <strong>the</strong> Meteorological Society of Jap<strong>an</strong>, 74(3), 1996<br />

Kepler, J.. THE SIX-CORNERED SNOWFLAKE<br />

(STRENA SEUDE NIVE SEXANGULA). Oxford University Press, Ely House, London W.1,<br />

1966<br />

Lynch, DK. Atmospheric Halos. SCIENTIFIC AMERICAN, 1978<br />

Nakaya, U.. The Formation of Ice Crystals. COMPENDIUM OF METEOROLOGY. Americ<strong>an</strong><br />

Meteorological Society, Boston pp. 207–220, 1951<br />

R<strong>as</strong>mussen, R., J.Vivek<strong>an</strong><strong>an</strong>d<strong>an</strong>, J.Cole <strong>an</strong>d E.Karplus. Theoretical Considerations in <strong>the</strong><br />

Estimation of <strong>Snow</strong>fall Rate Using Visibility. The National Center for Atmospheric Research,<br />

1998.<br />

Sato, K. Instability of Qu<strong>as</strong>i-liquid on <strong>the</strong> Edges <strong>an</strong>d Vertices of <strong>Snow</strong> Crystals. arXiv:condmat/0208036<br />

v1 2 Aug 2002.<br />

Schultz, C., <strong>an</strong>d L.D.Bregm<strong>an</strong>. GLOBAL SNOW ACCUMULATION BY MONTHS. The R<strong>an</strong>d<br />

Corporation, 1988.<br />

Yosida, Z. ICE AND SNOW PROPERTIES, PROCESSES AND APPLICATIONS. Editor,<br />

W.D.Kingery. M<strong>as</strong>sachusetts Institute of Technology Press, Cambridge 1962<br />

290


62nd ESC Papers<br />

291


This page is intentionally bl<strong>an</strong>k.<br />

292


62 nd EASTERN SNOW CONFERENCE<br />

Waterloo, ON, C<strong>an</strong>ada 2005<br />

Extended Abstract: Simulations of North Americ<strong>an</strong> <strong>Snow</strong> Cover<br />

by AGCMs <strong>an</strong>d AOGCMs<br />

Keywords: snow, climate modeling<br />

A. FREI, 1 G. GONG, 2 AND R. BROWN 3 :<br />

We report on some highlights of recent evaluations of General Circulation Model (GCM) snow<br />

simulations in which <strong>the</strong> authors have participated. Two recent reports focus on evaluations of<br />

snow cover extent (SCE) (Frei et al. 2003) <strong>an</strong>d snow water equivalent (SWE) (Frei et al. 2005)<br />

simulations for <strong>the</strong> period 1979-1995 by Atmospheric GCMs (AGCMs) participating in <strong>the</strong><br />

Second Ph<strong>as</strong>e of <strong>the</strong> Atmospheric Model Intercomparison Project (AMIP-2). We also report on<br />

results of a recent preliminary evaluation of SCE simulations by coupled atmosphere-oce<strong>an</strong> GCMs<br />

(AOGCMs) participating in <strong>the</strong> upcoming Fourth Assessment Report of <strong>the</strong> Intergovernmental<br />

P<strong>an</strong>el on Climate Ch<strong>an</strong>ge (IPCC-AR4) (Frei <strong>an</strong>d Gong 2005).<br />

Data Satellite observations of Nor<strong>the</strong>rn Hemisphere SCE are available back to around 1967<br />

(Robinson 1993). Reconstructions of large scale SCE variations back to <strong>the</strong> early twentieth<br />

century over North America have been performed in two studies (Frei et al. 1999; Brown 2000),<br />

which utilized station observations of snow depth. The primary data set used for model evaluation<br />

of SWE is <strong>the</strong> gridded SWE data set produced specifically for <strong>the</strong> AMIP-2 project by Brown et al.<br />

(2003).<br />

Models AMIP-2 modeling groups run experiments for <strong>the</strong> years 1979-1996 with identically<br />

specified boundary conditions, including observed sea surface temperatures, so that discrep<strong>an</strong>cies<br />

in model results are attributable to internal differences between atmospheric models. All IPCC-<br />

AR4 AOGCM simulations are forced with a set of boundary conditions determined by scenarios<br />

of <strong>an</strong>thropogenic emissions of carbon dioxide (CO2) <strong>an</strong>d o<strong>the</strong>r g<strong>as</strong>es that influence <strong>the</strong> global<br />

radiation budget. Here we consider twentieth century simulations (20C3M), which use a best<br />

estimate of historical emissions between roughly 1850 <strong>an</strong>d 2000; <strong>an</strong>d three twenty-first century<br />

climatic ch<strong>an</strong>ge scenarios (COMMIT, SRESA1B, SRESA2), which represent a r<strong>an</strong>ge of<br />

socioeconomic developments <strong>an</strong>d <strong>as</strong>sociated emission rates.<br />

Twentieth century results Me<strong>an</strong> monthly North Americ<strong>an</strong> SWE (Figure 1a) <strong>an</strong>d SCE (Figure<br />

1b) for observations <strong>an</strong>d AMIP-2 AGCMs demonstrate that models capture <strong>the</strong> me<strong>an</strong> se<strong>as</strong>onal<br />

cycle, <strong>the</strong>re is signific<strong>an</strong>t between model variability, <strong>an</strong>d models have a tendency to overestimate<br />

<strong>the</strong> ablation rate during spring. The me<strong>an</strong> spatial pattern of SWE is re<strong>as</strong>onably well captured by<br />

<strong>the</strong> medi<strong>an</strong> value of AMIP-2 models, except for <strong>an</strong> overly smoo<strong>the</strong>d representation of SWE over<br />

<strong>the</strong> western cordillera related to model orography, underestimation of SWE over e<strong>as</strong>tern NA<br />

(figure 2), <strong>an</strong>d <strong>the</strong> widespread underestimation of SWE during spring.(not shown).<br />

Between-model variability in IPCC-AR4 AOGCMs (Figure 3) is comparable to that found in<br />

AMIP-2 AGCMs. Figure 3, which shows ensemble me<strong>an</strong> time series, also indicates that <strong>the</strong><br />

models disagree with each o<strong>the</strong>r, <strong>an</strong>d with observations, on <strong>the</strong> timing <strong>an</strong>d magnitude of decadal<br />

scale variations. With regards to <strong>the</strong> magnitude of decadal-scale variability, <strong>the</strong> disagreements<br />

between models <strong>an</strong>d observations are perhaps not <strong>as</strong> great <strong>as</strong> <strong>the</strong>y appear. Those models which<br />

1 Dept. of Geography, Hunter College, New York, NY, USA<br />

2 Dept. of Earth <strong>an</strong>d Environmental Engineering, Columbia University, New York, NY, USA<br />

3 Meteorological Service of C<strong>an</strong>ada, Dorval, Quebec, C<strong>an</strong>ada<br />

293


appear to exhibit smaller decadal scale variability tend to include more ensemble members, <strong>an</strong>d<br />

those models which exhibit greater decadal scale variability tend to include only one ensemble<br />

member. In fact, <strong>the</strong> decadal scale variability of ensemble me<strong>an</strong>s are consistently smaller th<strong>an</strong> <strong>the</strong><br />

that of <strong>the</strong> individual ensemble members, <strong>an</strong>d that of individual members tend to be closer to<br />

observed values. This indicates that decadal scale variability in <strong>the</strong>se models is due to internal<br />

dynamics, <strong>an</strong>d not due to external forcings.<br />

a b<br />

Figure 1. North Americ<strong>an</strong> SWE (a) <strong>an</strong>d SCE (b). Observations from Brown et al (2003) indicated with<br />

<strong>as</strong>terisks; from Robinson (1993) with crosses. Box <strong>an</strong>d whisker plots indicate model results from 18<br />

AMIP-2 AGCMs, <strong>an</strong>d are interpreted <strong>as</strong> follows: middle line shows <strong>the</strong> medi<strong>an</strong> value; top <strong>an</strong>d bottom of<br />

box show <strong>the</strong> upper <strong>an</strong>d lower quartiles (i.e. 75 th <strong>an</strong>d 25 th percentile values); <strong>an</strong>d whiskers show <strong>the</strong><br />

minimum <strong>an</strong>d maximum model values. Figure adapted from Frei et al (2005).<br />

a b<br />

Figure 2. Se<strong>as</strong>onal (October-June) me<strong>an</strong> SWE (mm) regridded to 2.5° x 2.5° latitude -<br />

longitude resolution. (a) Observed, (b) model <strong>an</strong>omaly. Regridding w<strong>as</strong> done using linear<br />

interpolation, <strong>an</strong>d results are plotted on <strong>an</strong> Albers equal area projection using a logarithmic<br />

scale for contour lines. In (b), blue are<strong>as</strong> indicate model underestimation of SWE. Figure<br />

adapted from Frei et al (2005).<br />

Twenty first century results At <strong>the</strong> time of this writing, output from five modeling groups are<br />

available to evaluate North Americ<strong>an</strong> SCE variations under <strong>the</strong> historical emission scenario<br />

(20C3M) <strong>an</strong>d all three twenty first century scenarios (COMMIT, SRESA1B, <strong>an</strong>d SRESA2). Under<br />

<strong>the</strong> COMMIT scenario, greenhouse g<strong>as</strong> emission rates remain const<strong>an</strong>t at year 2000 values. Under<br />

<strong>the</strong> SRES scenarios, emission rates incre<strong>as</strong>e ei<strong>the</strong>r moderately (SRESA1B) or severely (SRESA2).<br />

As <strong>an</strong> example, Figure 4 shows <strong>the</strong> responses of SCE to each emission scenario for one model.<br />

Although trends vary considerably between models, in all c<strong>as</strong>es <strong>the</strong> decre<strong>as</strong>ing trends for <strong>the</strong><br />

SRESA1B <strong>an</strong>d SRESA2 scenarios are statistically signific<strong>an</strong>t <strong>an</strong>d comparable in magnitude to<br />

294


each o<strong>the</strong>r; <strong>an</strong>d, <strong>the</strong>y both decre<strong>as</strong>e at greater rates th<strong>an</strong> under <strong>the</strong> COMMIT scenario. For all<br />

models under <strong>the</strong> SRESA1B <strong>an</strong>d SRESA2 scenarios, SCE decre<strong>as</strong>es at a greater rate during <strong>the</strong><br />

twenty first century th<strong>an</strong> during <strong>the</strong> twentieth century. In contr<strong>as</strong>t, under <strong>the</strong> COMMIT scenario,<br />

while some models do have weak but signific<strong>an</strong>t decre<strong>as</strong>ing trends in SCE, in no model does SCE<br />

decre<strong>as</strong>e at a greater rate during <strong>the</strong> twenty first century th<strong>an</strong> during <strong>the</strong> twentieth century. The<br />

differences in responses are not proportional to <strong>the</strong> differences in forcing under <strong>the</strong>se scenarios,<br />

indicating that non-linear dynamics are influencing <strong>the</strong> snow cover.<br />

Discussion <strong>an</strong>d conclusions Signific<strong>an</strong>t between-model variability is found in all comparisons<br />

of GCM snow simulations: e.g. <strong>the</strong> r<strong>an</strong>ge of simulated snow m<strong>as</strong>s of North America by AMIP-2<br />

AGCMs is ±50% of <strong>the</strong> estimated value. When using a GCM to evaluate potential ch<strong>an</strong>ges in<br />

regional to continental scale hydrological variations one must exercise caution. However, <strong>the</strong><br />

medi<strong>an</strong> result from all models tends to do a re<strong>as</strong>onably good job compared to observations. Thus,<br />

perhaps a “superensemble” of models, when numerous simulations from different models are<br />

combined, may be <strong>an</strong> effective method. Preliminary evaluations of AOGCMs suggest that:<br />

decadal scale variability is <strong>as</strong>sociated with internal climatic variations <strong>an</strong>d not with external<br />

forcings in <strong>the</strong>se models (whe<strong>the</strong>r that is true in <strong>the</strong> real climate system is unknown); <strong>an</strong>d,<br />

decre<strong>as</strong>es in snow extent are expected under realistic scenarios of future emissions, although <strong>the</strong><br />

precise response of <strong>the</strong> snow cover to possible future climate variations may be non linear.<br />

Figure 3. Nine-year running me<strong>an</strong> of twentieth century J<strong>an</strong>uary NA-SCE for 11 IPCC-AR4 ensembleme<strong>an</strong><br />

model simulations <strong>an</strong>d for reconstructions of observed variations. NA-SCE is defined <strong>as</strong> <strong>the</strong><br />

fraction of <strong>the</strong> l<strong>an</strong>d area from 20° N - 90°N <strong>an</strong>d 190° E - 340° E covered with snow. The legend shows<br />

<strong>the</strong> model number, which corresponds to model numbers <strong>an</strong>d ensemble sizes shown in Table 2; “B” <strong>an</strong>d<br />

“F” correspond to Brown (2000) <strong>an</strong>d Frei et al. (1999), respectively. Figure adapted from Frei <strong>an</strong>d Gong<br />

(2005).<br />

Fractional <strong>Snow</strong> Covered Area<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

MRI-CGCM2.3.2<br />

0.5<br />

1900 1950 2000<br />

Year<br />

2050 2100<br />

Figure 4. Annual time series (thin line), overlaid with nine-year running me<strong>an</strong>s (thick line), of ensembleme<strong>an</strong><br />

J<strong>an</strong>uary NA-SCE for 20 th <strong>an</strong>d 21 st century scenarios from one AOGCM. 20C3M, COMMIT,<br />

SRESA1B <strong>an</strong>d SRESA2 scenarios denoted by black, red, blue <strong>an</strong>d green lines, respectively. Figure<br />

adapted from Frei <strong>an</strong>d Gong (2005).<br />

295


References<br />

Brown, R. D. (2000). Nor<strong>the</strong>rn hemisphere snow cover variability <strong>an</strong>d ch<strong>an</strong>ge, 1915-1997.<br />

Journal of Climate 13(13): 2339-2355.<br />

Brown, R. D., B. Br<strong>as</strong>nett <strong>an</strong>d D. A. Robinson (2003). Gridded North Americ<strong>an</strong> monthly<br />

snow depth <strong>an</strong>d snow water equivalent for GCM evaluation. Atmosphere-Oce<strong>an</strong> 41(1): 1-14.<br />

Frei, A., R. Brown, J. A. Miller <strong>an</strong>d D. A. Robinson (2005). <strong>Snow</strong> m<strong>as</strong>s over North America:<br />

observations <strong>an</strong>d results from <strong>the</strong> second ph<strong>as</strong>e of <strong>the</strong> Atmospheric Model Intercomparison<br />

Project (AMIP-2). Journal of Hydrometeorology: in press.<br />

Frei, A. <strong>an</strong>d G. Gong (2005). Decadal to century scale trends in North Americ<strong>an</strong> snow extent<br />

in coupled Atmosphere-Oce<strong>an</strong> General Circulation Models. Geophysical Research Letters: in<br />

review.<br />

Frei, A., J. A. Miller <strong>an</strong>d D. A. Robinson (2003). Improved simulations of snow extent in <strong>the</strong><br />

second ph<strong>as</strong>e of <strong>the</strong> Atmospheric Model Intercomparison Project (AMIP-2). Journal of<br />

Geophysical Research - Atmospheres 108(D12): 4369, doi:4310.1029/2002JD003030.<br />

Frei, A., D. A. Robinson <strong>an</strong>d M. G. Hughes (1999). North Americ<strong>an</strong> snow extent: 1900-1994.<br />

International Journal of Climatology 19: 1517-1534.<br />

Robinson, D. A. (1993). Hemispheric snow cover from satellites. Annals of Glaciology 17:<br />

367-371.<br />

296


297<br />

62 nd EASTERN SNOW CONFERENCE<br />

Waterloo, ON, C<strong>an</strong>ada 2005<br />

The Impact of Patchy <strong>Snow</strong> Cover on <strong>Snow</strong> Water Equivalent<br />

Estimates Derived from P<strong>as</strong>sive Microwave Brightness<br />

Temperatures Over a Prairie Environment<br />

ABSTRACT<br />

KIM R. TURCHENEK 1 , JOSEPH M. PIWOWAR1, AND CHRIS DERKSEN 2<br />

Considerable se<strong>as</strong>onal <strong>an</strong>d inter-<strong>an</strong>nual variation in <strong>the</strong> physical properties <strong>an</strong>d extent of snow<br />

cover pose problems for obtaining reliable estimates of qu<strong>an</strong>tities <strong>an</strong>d characteristics of snow<br />

cover both from conventional <strong>an</strong>d satellite me<strong>as</strong>urements (Goodison <strong>an</strong>d Walker, 1994; Goita et<br />

al., 2003). In spite of <strong>the</strong>se challenges, <strong>the</strong> Climate Research Br<strong>an</strong>ch of <strong>the</strong> Meteorological Service<br />

of C<strong>an</strong>ada (MSC) h<strong>as</strong> developed a suite of algorithms to derive snow water equivalent (SWE)<br />

estimates from remotely sensed p<strong>as</strong>sive microwave imagery (Goodison <strong>an</strong>d Walker, 1994;<br />

Derksen et al., 2002; Goita et al., 2003). These algorithms work particularly well over open prairie<br />

environments under <strong>the</strong> <strong>as</strong>sumption of large are<strong>as</strong> of consistent snow cover (Derksen et al., 2002).<br />

While studies have documented underestimation in p<strong>as</strong>sive microwave estimates of snow extent in<br />

marginal are<strong>as</strong> when compared to optical satellite data (Derksen et al., 2003b), <strong>the</strong> accuracy in<br />

SWE retrievals under variable <strong>an</strong>d patchy snow conditions is not well understood.<br />

In <strong>an</strong> effort to better underst<strong>an</strong>d how a variable <strong>an</strong>d patchy snow cover impacts remotely sensed<br />

SWE retrievals, a field-b<strong>as</strong>ed experiment w<strong>as</strong> conducted over a patchy snow covered area in<br />

February 2005. A systematic sampling strategy w<strong>as</strong> developed over a 1600 km 2 area in sou<strong>the</strong>rn<br />

S<strong>as</strong>katchew<strong>an</strong> near a calibration/validation flight line used for algorithm development in <strong>the</strong> 1980s<br />

(Goodison <strong>an</strong>d Walker, 1994). L<strong>an</strong>d cover at <strong>the</strong> sampling sites included fallow <strong>an</strong>d stubble fields,<br />

p<strong>as</strong>tures, <strong>an</strong>d shelter belts. A large number of sampling sites contained snow pack layers that<br />

included one or more ice lenses.<br />

We verify that <strong>the</strong> continuous snow cover <strong>as</strong>sumption embedded in <strong>the</strong> MSC p<strong>as</strong>sive microwave<br />

SWE algorithm does not produce acceptable results over a patchy snow cover. Several in-situ<br />

observations that appear to play <strong>an</strong> import<strong>an</strong>t role in affecting <strong>the</strong> satellite p<strong>as</strong>sive microwave data<br />

over a variable snow cover include <strong>the</strong> presence or absence of <strong>an</strong> ice lens, <strong>the</strong> fractional snow<br />

covered area, snow depth, <strong>an</strong>d <strong>the</strong> ground temperature.<br />

Keywords: snow cover, snow water equivalent, p<strong>as</strong>sive microwave, remote sensing<br />

INTRODUCTION<br />

Microwave radiation is naturally emitted everywhere on <strong>the</strong> Earth. Its me<strong>as</strong>urable intensity<br />

varies from place to place b<strong>as</strong>ed on soil types, l<strong>an</strong>d covers, snow pack characteristics, <strong>an</strong>d o<strong>the</strong>r<br />

variables (Goita et al., 1997; Sokol et al., 1999). At microwave frequencies above 15 GHz, <strong>the</strong><br />

emitted radiation is scattered by snow particles <strong>as</strong> it p<strong>as</strong>ses through <strong>the</strong> snow pack (Goita et al.,<br />

1997). Incre<strong>as</strong>ing <strong>the</strong> snow pack depth or grain size results in <strong>an</strong> incre<strong>as</strong>e in scattering <strong>an</strong>d<br />

1 Department of Geography, University of Regina, Regina, S<strong>as</strong>katchew<strong>an</strong>, S4S 0A2<br />

2 Climate Research Br<strong>an</strong>ch, Meteorological Service of C<strong>an</strong>ada, Downsview, Ontario M3H 5T4


subsequent lower microwave brightness temperatures when me<strong>as</strong>ured above <strong>the</strong> surface (Goita et<br />

al., 1997). Wet <strong>an</strong>d/or dense snow packs however, decre<strong>as</strong>e <strong>the</strong> amount of scattering, <strong>an</strong>d produce<br />

near blackbody emissions (Sokol et al., 1999). Complex snow packs, containing numerous layers<br />

at different densities, selectively influence microwave radiation, producing inconsistent<br />

me<strong>as</strong>urements.<br />

Radiation recorded by a microwave sensor is expressed <strong>as</strong> a brightness temperature (TB) in<br />

Kelvin units. One parameter commonly derived from remotely sensed brightness temperatures is<br />

snow water equivalent (SWE), which is <strong>the</strong> amount of water stored in a snow pack that is available<br />

upon melt.<br />

Intensive research with p<strong>as</strong>sive microwave TB data h<strong>as</strong> focused on empirically derived<br />

algorithms used to estimate SWE for validation of ground-truth observations (Goodison <strong>an</strong>d<br />

Walker, 1994; Goita et al., 1997; Derksen et al., 2002; 2003a). Currently, <strong>the</strong> Meteorological<br />

Service of C<strong>an</strong>ada (MSC) employs a suite of linear algorithms to retrieve SWE estimates from<br />

p<strong>as</strong>sive microwave sensors. The MSC algorithms vary according to l<strong>an</strong>d cover, with different<br />

coefficients used for open prairie, deciduous forest, coniferous forest, <strong>an</strong>d sparse forest (Goita et<br />

al., 1997; Singh <strong>an</strong>d G<strong>an</strong>, 2000; Derksen et al, 2003a; 2003b). For example, <strong>the</strong> algorithm used to<br />

derive SWE over <strong>the</strong> open prairie is a vertically polarized TB gradient ratio (Goodison <strong>an</strong>d Walker,<br />

1994) defined <strong>as</strong>:<br />

SWE (mm) = –20.7 – (37v – 19v) * 2.59 (1)<br />

Variables 37v <strong>an</strong>d 19v are <strong>the</strong> brightness temperatures acquired from vertically polarized<br />

frequencies of 37 <strong>an</strong>d 19 GHz, while <strong>the</strong> coefficients 2.59 <strong>an</strong>d –20.7 are <strong>the</strong> slope <strong>an</strong>d intercept of<br />

<strong>the</strong> best-fit regression line found between ground <strong>an</strong>d airborne brightness temperatures (Goodison,<br />

1989). Numerous studies have found SWE estimates derived from this algorithm to be within<br />

±10–20 mm of in-situ observations (Goodison <strong>an</strong>d Walker, 1994; Derksen et al., 2002; 2003b),<br />

however, <strong>the</strong> MSC algorithm <strong>as</strong>sumes a complete snow cover <strong>an</strong>d <strong>the</strong> effects of patchy snow cover<br />

on SWE estimates are not well understood.<br />

To help underst<strong>an</strong>d <strong>the</strong> relationship between patchy snow covers <strong>an</strong>d remotely sensed SWE<br />

estimates, a field campaign to me<strong>as</strong>ure SWE in a highly variable snow pack w<strong>as</strong> conducted from<br />

February 21 st to 23 rd , 2005 in sou<strong>the</strong>rn S<strong>as</strong>katchew<strong>an</strong>. The objectives of this study were to:<br />

i) Compare remotely sensed SWE estimates against ground-truth me<strong>as</strong>urements;<br />

ii) Determine how well <strong>the</strong> MSC prairie SWE algorithm performs over a patchy snow cover;<br />

<strong>an</strong>d<br />

iii) Identify in-situ variables that signific<strong>an</strong>tly influence p<strong>as</strong>sive microwave SWE estimates<br />

over a patchy snow cover.<br />

STUDY AREA<br />

The study area is located approximately 100 km south of Regina, S<strong>as</strong>katchew<strong>an</strong> around <strong>the</strong><br />

town of Radville <strong>an</strong>d <strong>the</strong> villages of P<strong>an</strong>gm<strong>an</strong> <strong>an</strong>d Ceylon (Figure 1). This area w<strong>as</strong> selected<br />

because previous research had been performed in this area by Environment C<strong>an</strong>ada. In addition,<br />

<strong>the</strong> snow cover typical of this area h<strong>as</strong> been found to have regions of both patchy <strong>an</strong>d complete<br />

snow-cover (Goodison <strong>an</strong>d Walker, 1994; Turchenek, 2004).<br />

The Missouri Coteau, a remn<strong>an</strong>t glacial moraine, cuts across <strong>the</strong> southwest portion of <strong>the</strong> study<br />

area, forming a low, rolling topography. Although <strong>the</strong> natural vegetation in this area consists of<br />

short gr<strong>as</strong>ses, a large portion of <strong>the</strong> l<strong>an</strong>d is under agricultural production, where wheat <strong>an</strong>d o<strong>the</strong>r<br />

grains are farmed. Pockets of trees <strong>an</strong>d shrubs create shelter belts in are<strong>as</strong> with higher moisture<br />

supply.<br />

Relatively short warm summers <strong>an</strong>d long cold winters are characteristic of S<strong>as</strong>katchew<strong>an</strong>’s<br />

prairies (Hare <strong>an</strong>d Thom<strong>as</strong>, 1979). With few perennial streams, much of <strong>the</strong> region’s water supply<br />

298


comes in <strong>the</strong> form of precipitation. However, <strong>the</strong> <strong>an</strong>nual precipitation in <strong>the</strong> region is relatively<br />

low, <strong>an</strong>d evapotr<strong>an</strong>spiration usually exceeds <strong>the</strong> <strong>an</strong>nual precipitation, creating <strong>an</strong> average water<br />

deficit by middle to late summer (Laycock, 1972). Most <strong>an</strong>nual precipitation falls in <strong>the</strong> summer,<br />

while February is usually <strong>the</strong> driest month (Hare <strong>an</strong>d Thom<strong>as</strong>, 1979).<br />

Figure 1. Study area<br />

The winter se<strong>as</strong>on provides relatively low amounts of snow. Extended periods of cold, clear<br />

wea<strong>the</strong>r are interrupted by occ<strong>as</strong>ional blizzards with gusting winds. Warming periods are frequent<br />

in <strong>the</strong> early <strong>an</strong>d late winter (Laycock, 1972; Hare <strong>an</strong>d Thom<strong>as</strong>, 1979; Walker et al., 1995). Wind<br />

re-distributes <strong>the</strong> snow cover by removing snow from one area <strong>an</strong>d depositing it in <strong>an</strong>o<strong>the</strong>r.<br />

Similarly, warming periods also impact <strong>the</strong> snow cover through freeze–thaw processes (Laycock,<br />

1972; Walker et al., 1995). As <strong>the</strong> air <strong>an</strong>d ground temperatures rise, <strong>the</strong> snow pack melts. When<br />

<strong>the</strong> snow pack re-freezes, it becomes denser <strong>an</strong>d shallower. Thus, along with topographic effects<br />

<strong>an</strong>d ch<strong>an</strong>ges in vegetation, wea<strong>the</strong>r systems c<strong>an</strong> impart a considerable variability in snow pack<br />

depths <strong>an</strong>d densities.<br />

REMOTE SENSING DATA<br />

Four sets of coincident remote sensing data were <strong>an</strong>alyzed (Table 1). Three data sets were<br />

derived from <strong>the</strong> brightness temperatures collected from <strong>the</strong> Adv<strong>an</strong>ced Microwave Sc<strong>an</strong>ning<br />

Radiometer for NASA’s Earth Observing System (AMSR-E). The first AMSR-E data set includes<br />

TB re-sampled to <strong>the</strong> 12.5 km Equal Area Scalable Earth Grid (EASE-Grid) (Armstrong <strong>an</strong>d<br />

Brodzik, 1995). The second includes AMSR-E TB re-sampled to <strong>the</strong> 25 km EASE-Grid <strong>an</strong>d <strong>the</strong><br />

third AMSR-E data set includes non-gridded TB swath data. For comparison, a fourth data set of<br />

SWE estimates derived from Special Sensor Microwave/Imager (SSM/I) brightness temperatures<br />

re-sampled to <strong>the</strong> 25 km EASE-Grid w<strong>as</strong> included. Although data were acquired for each day of<br />

<strong>the</strong> field campaign, <strong>the</strong> only remote sensing data <strong>an</strong>alyzed here were those collected on <strong>the</strong> first<br />

day of <strong>the</strong> campaign (Feb 21 st , 2005). Future research will incorporate <strong>the</strong> coincident remote<br />

sensing data from each date of <strong>the</strong> field campaign.<br />

299


Table 1: Remote sensing data used within this study<br />

Data Set Label Spatial Resolution Description<br />

SWE estimates derived from<br />

AMSR-E TB re-sampled to 12.5 km<br />

1) 12.5k_AMSR-E SWE 12.5 km<br />

EASE-Grid<br />

18.7v TB obtained from AMSR-E<br />

12.5k_AMSR-E 18.7v 12.5 km<br />

re-sampled to 12.5 km EASE-Grid<br />

36.5v TB obtained from AMSR-E<br />

12.5k_AMSR-E 36.5v 12.5 km<br />

re-sampled to 12.5 km EASE-Grid<br />

SWE estimates derived from<br />

AMSR-E TB re-sampled to 25 km<br />

2) 25k_AMSR-E SWE<br />

25 km<br />

EASE-Grid<br />

18.7v TB obtained from AMSR-E<br />

25k_AMSR-E 18.7v<br />

25 km<br />

re-sampled to 25 km EASE-Grid<br />

36.5v TB obtained from AMSR-E<br />

25k_AMSR-E 36.5v<br />

25 km<br />

re-sampled to 25 km EASE-Grid<br />

SWE estimates derived from<br />

AMSR-E TB that have not been re-<br />

3) Swath_AMSR-E SWE 24 km x 12 km sampled<br />

18.7v TB obtained from AMSR-E<br />

Swath_AMSR-E 18.7v 24 km<br />

that have not been re-sampled<br />

36.5v TB obtained from AMSR-E<br />

Swath_AMSR-E 36.5v 12 km<br />

that have not been re-sampled<br />

SWE estimates derived from<br />

SSM/I TB re-sampled to 25 km<br />

4) 25k_SSM/I SWE<br />

25 km<br />

EASE-Grid<br />

METHODS<br />

The <strong>an</strong>alysis procedure w<strong>as</strong> divided into four steps: i) collection of <strong>the</strong> ground-truth<br />

observations, ii) processing of <strong>the</strong> data sets, iii) comparison with <strong>the</strong> remotely sensed SWE<br />

estimates, <strong>an</strong>d iv) statistical tests. These are discussed in <strong>the</strong> following subsections.<br />

In-situ Data Collection<br />

The study area w<strong>as</strong> systematically divided into 25 km square grid cells, with <strong>the</strong> centre of each<br />

cell located at 5 km intervals. Field me<strong>as</strong>urements were made nearest <strong>the</strong> centre of each grid cell<br />

<strong>as</strong> possible. The field campaign w<strong>as</strong> concentrated into a three-day period to minimize ch<strong>an</strong>ges in<br />

snow pack conditions due to melt or fresh snowfall.<br />

Two teams, of 3–4 surveyors each, collected a total of 88 ground observations from 84 sampling<br />

sites that covered <strong>an</strong> area of 1600 km 2 . As a control, four of <strong>the</strong> sites were sampled on consecutive<br />

days to ensure data consistency. Included in <strong>the</strong> 84 sampling sites were 20 sites coincident with <strong>an</strong><br />

established MSC validation/calibration snow course data archive (Flight Line 603). Comparisons<br />

among <strong>the</strong>se data will be <strong>the</strong> subject of a future communication. The l<strong>an</strong>d coverage of <strong>the</strong><br />

sampling sites included p<strong>as</strong>tures <strong>an</strong>d shelter belts, <strong>as</strong> well <strong>as</strong> fallow <strong>an</strong>d stubble fields. The<br />

ground-truth data collected from each sampling site included: geographic locations, snow pack<br />

profiles, depth me<strong>as</strong>urements, core samples, air <strong>an</strong>d ground temperatures, site photographs,<br />

sampling dates <strong>an</strong>d times, <strong>an</strong>d l<strong>an</strong>d cover <strong>an</strong>d wea<strong>the</strong>r observations. Sampling site locations were<br />

recorded using global positioning system (GPS) h<strong>an</strong>dsets. Since data for differential corrections<br />

were not available, <strong>the</strong> positions have <strong>an</strong> approximate accuracy of 10 metres.<br />

Approximately 90% of <strong>the</strong> sampling sites were found to have patchy snow coverage. As such,<br />

one of <strong>the</strong> first observations made at each site w<strong>as</strong> a visual <strong>as</strong>sessment of <strong>the</strong> percentage of snow<br />

cover. Each team member made this <strong>as</strong>sessment independently <strong>an</strong>d <strong>the</strong>se values were <strong>the</strong>n<br />

300


averaged to reduce bi<strong>as</strong> in this variable. For sites with complete snow cover, a total of 4 snow core<br />

samples were collected. This number w<strong>as</strong> reduced proportionally for partially covered sites. For<br />

example, only 2 cores were collected from sites determined <strong>as</strong> having 50% snow coverage, <strong>an</strong>d<br />

just 1 core w<strong>as</strong> collected from sites with 25% snow cover. This rationale w<strong>as</strong> used to satisfy <strong>the</strong><br />

requirement that <strong>the</strong> cores be taken r<strong>an</strong>domly within each site. Thus, if a site w<strong>as</strong> found to have<br />

50% snow cover, <strong>the</strong>n <strong>the</strong> probability of r<strong>an</strong>domly selecting a sampling location containing snow<br />

is only 50%. In this situation, <strong>the</strong> 2 core samples that were not actually collected were simply<br />

given zero values for <strong>the</strong>ir core lengths, depths, weights, <strong>an</strong>d densities.<br />

The core samples were obtained using E<strong>as</strong>tern <strong>Snow</strong> Conference (ESC-30) snow core tubes.<br />

Me<strong>as</strong>urements from <strong>the</strong> core samples include <strong>the</strong> actual depths of <strong>the</strong> snow packs from where <strong>the</strong><br />

cores were removed, <strong>the</strong> lengths of <strong>the</strong> cores, <strong>an</strong>d <strong>the</strong>ir weights. The lengths <strong>an</strong>d weights of <strong>the</strong><br />

cores were used to calculate <strong>the</strong> core densities <strong>an</strong>d ground SWE me<strong>as</strong>urements. <strong>Snow</strong> densities,<br />

represented <strong>as</strong> g/cm 3 , were b<strong>as</strong>ed on <strong>the</strong> average of <strong>the</strong> 0 to 4 core samples.<br />

A snow pit w<strong>as</strong> dug at each sampling site <strong>an</strong>d a detailed snow pack profile w<strong>as</strong> made that<br />

included <strong>the</strong> snow pack’s total depth, <strong>the</strong> number of layers <strong>an</strong>d ice lenses within <strong>the</strong> snow pack,<br />

<strong>the</strong> depth <strong>an</strong>d snow grain size of each layer (using Sears snow crystal screens, labeled with 1–3<br />

mm grids), a qualitative description of each layer, <strong>an</strong>d <strong>the</strong> air <strong>an</strong>d snow/ground interface<br />

temperatures. A total of 16 depth me<strong>as</strong>urements were made around each snow pit using 15-metre<br />

long ropes <strong>as</strong> guides for <strong>the</strong> purpose of consistently collecting <strong>the</strong> depth me<strong>as</strong>urements from 30metre<br />

diameter circles. Depth me<strong>as</strong>urements to <strong>the</strong> nearest one-half centimetre were made using 1metre<br />

long depth probes. The depth me<strong>as</strong>urements from each site were used to calculate <strong>the</strong><br />

average depths within <strong>the</strong> sites, which were <strong>the</strong>n used <strong>as</strong> representative values for <strong>the</strong> sampling<br />

sites. The average depths are b<strong>as</strong>ed on <strong>the</strong> 16 r<strong>an</strong>dom depth me<strong>as</strong>urements recorded from <strong>the</strong><br />

circle around <strong>the</strong> snow pit along with <strong>the</strong> 0 to 4 depth me<strong>as</strong>urements recorded from <strong>the</strong> snow core<br />

samples.<br />

O<strong>the</strong>r data recorded included <strong>the</strong> sampling dates <strong>an</strong>d times, wea<strong>the</strong>r observations, <strong>an</strong>d l<strong>an</strong>d<br />

cover types.<br />

In-situ Data Processing<br />

Two data sets were created from <strong>the</strong> ground sampled data in order to better underst<strong>an</strong>d how<br />

snow properties over a partial snow cover are m<strong>an</strong>ifested in <strong>the</strong> remotely sensed SWE estimates.<br />

In <strong>the</strong> “<strong>Snow</strong>-Only” data set, only those snow depth me<strong>as</strong>urements that were greater th<strong>an</strong> zero<br />

were included in <strong>the</strong> average depth, density, <strong>an</strong>d ground SWE calculations. For example, if a site<br />

w<strong>as</strong> found to have snow depth readings of 3, 4, 0, 1, <strong>an</strong>d 4 cm, <strong>the</strong>n <strong>the</strong> average depth for that site<br />

w<strong>as</strong> recorded <strong>as</strong> 3.0 cm ( (3 + 4 + 1 + 4) / 4 ), excluding <strong>the</strong> zero value. Conversely, <strong>the</strong> same site<br />

in <strong>the</strong> “Actual-Conditions” data set would have a me<strong>an</strong> depth of 2.4 cm ( (3 + 4 + 0 + 1 + 4 / 5) ).<br />

From each of <strong>the</strong>se data sets, four SWE estimates were calculated. The first ground SWE value,<br />

“Core_SWE,” represents <strong>the</strong> SWE calculated by using <strong>the</strong> me<strong>an</strong> SWE from <strong>the</strong> snow cores only.<br />

The second SWE value, “Derived_SWE,” is representative of <strong>the</strong> me<strong>an</strong> value for <strong>the</strong> Core_SWE<br />

plus <strong>the</strong> SWE derived from <strong>the</strong> 16 depth me<strong>as</strong>urements using <strong>the</strong> average density for each<br />

sampling site. The remaining SWE values, “Fractional_Core_SWE” <strong>an</strong>d<br />

“Fractional_Derived_SWE” are represented by <strong>the</strong> previous SWE values weighted by <strong>the</strong><br />

<strong>Snow</strong>_Cover_Percent, respectively. Four SWE values were deemed necessary to investigate <strong>the</strong><br />

most accurate way of representing SWE over a patchy snow cover.<br />

Remote Sensing Data<br />

The remote sensing images were re-projected to a UTM projection for <strong>an</strong>alysis in ArcGIS. Each<br />

pixel centroid w<strong>as</strong> <strong>as</strong>sumed to be a point, <strong>an</strong>d pixel footprints were created using Thiessen<br />

polygons. The Thiessen polygon algorithm segments <strong>the</strong> me<strong>as</strong>urement space into polygons such<br />

that every polygon encloses <strong>the</strong> region closest to each pixel centroid (O’Sulliv<strong>an</strong> <strong>an</strong>d Unwin,<br />

2003). Although <strong>the</strong> algorithm does not produce perfectly squared pixels, <strong>the</strong> fact that <strong>the</strong><br />

algorithm me<strong>as</strong>ures <strong>the</strong> mid-points between <strong>the</strong> pixel centroids ensures that <strong>the</strong> spatial resolutions<br />

of <strong>the</strong> remote sensing data sets are preserved.<br />

301


Following previous research by Goodison <strong>an</strong>d Walker (1995) <strong>an</strong>d Derksen et al. (2002, 2003b),<br />

remotely sensed SWE estimates were compared to ground SWE me<strong>as</strong>urements. The remotely<br />

sensed estimates found to be within ±20 mm (<strong>the</strong> previously determined accuracy of <strong>the</strong> MSC<br />

SWE algorithm) of ground me<strong>as</strong>urements were considered <strong>as</strong> equivalent. SWE estimates found<br />

not to be within <strong>the</strong> ±20 mm threshold were considered <strong>as</strong> <strong>an</strong>omalous.<br />

Statistical Comparison Tests<br />

Z-tests were performed between <strong>the</strong> results of <strong>the</strong> <strong>Snow</strong>-Only <strong>an</strong>d Actual-Conditions data sets to<br />

determine if <strong>the</strong>re were signific<strong>an</strong>t differences in algorithm perform<strong>an</strong>ce between <strong>the</strong> two groundtruth<br />

representations. Linear regression models were developed between <strong>the</strong> remote sensing data<br />

(<strong>as</strong> <strong>the</strong> depend<strong>an</strong>t variables) <strong>an</strong>d <strong>the</strong> ground observations from <strong>the</strong> coincident sampling sites (<strong>as</strong><br />

<strong>the</strong> independent variables).<br />

The linear regression outputs were interpreted following <strong>the</strong> systematic procedure proposed by<br />

Gupta (2000) (Figure 2). The signific<strong>an</strong>ce of <strong>the</strong> model fit is <strong>an</strong>alyzed. The model signific<strong>an</strong>ce<br />

explains <strong>the</strong> deviations of <strong>the</strong> dependent variables (eg. <strong>the</strong> SWE estimates <strong>an</strong>d TB). We used a<br />

model signific<strong>an</strong>ce of 0.10 (90% confidence level) <strong>as</strong> <strong>the</strong> cutoff for model accept<strong>an</strong>ce. Models<br />

with levels below <strong>the</strong> 90% confidence level were removed from fur<strong>the</strong>r <strong>an</strong>alysis.<br />

The next step in interpreting <strong>the</strong> regression output is to <strong>an</strong>alyze <strong>the</strong> Adjusted R 2 value from <strong>the</strong><br />

model summary. This value is sensitive to <strong>the</strong> addition of irrelev<strong>an</strong>t variables, <strong>an</strong>d is a me<strong>as</strong>ure of<br />

<strong>the</strong> proportion of <strong>the</strong> vari<strong>an</strong>ce in <strong>the</strong> dependent variables that are explained by <strong>the</strong> variations of <strong>the</strong><br />

independent variables. For example, <strong>an</strong> Adjusted R 2 value of 0.500 suggests that 50% of <strong>the</strong><br />

vari<strong>an</strong>ce in a SWE estimate is explained by <strong>the</strong> variation in <strong>the</strong> ground-truth me<strong>as</strong>urements.<br />

The third interpretation step involves identifying <strong>the</strong> reliability of <strong>the</strong> individual coefficients for<br />

<strong>the</strong> independent variables. The Beta values included in <strong>the</strong> coefficients output indicate <strong>the</strong><br />

predicted coefficients for <strong>the</strong> model along with <strong>the</strong>ir st<strong>an</strong>dard errors <strong>an</strong>d signific<strong>an</strong>ces. Similar to<br />

<strong>the</strong> signific<strong>an</strong>ce of <strong>the</strong> model fit, if a coefficient results in a signific<strong>an</strong>ce value above 0.10, <strong>the</strong>n it<br />

c<strong>an</strong> be concluded that <strong>the</strong> independent variable is not signific<strong>an</strong>t at a 90% level of confidence.<br />

Table 2 provides <strong>an</strong> instructive example of how <strong>an</strong> output coefficient table is <strong>an</strong>alyzed. The<br />

SWE predicted by <strong>the</strong> model is specified by <strong>the</strong> Const<strong>an</strong>t’s Beta coefficient, 30.5 mm. The<br />

st<strong>an</strong>dard error of this prediction is 1.9 mm of SWE. The next step is to identify <strong>the</strong> signific<strong>an</strong>ce of<br />

<strong>the</strong> independent variables’ Beta coefficients. In this example, a coefficient of .149 for <strong>the</strong> average<br />

snow pack depth is found to be signific<strong>an</strong>t at a 95% level of confidence (sig. = .043), but <strong>the</strong><br />

coefficient of –2.899 for <strong>the</strong> average density is irrelev<strong>an</strong>t (sig. = .776) towards <strong>the</strong> predicted SWE<br />

value.<br />

Table 2: Example coefficient table output<br />

Unst<strong>an</strong>dardized Coefficients<br />

Model<br />

Beta<br />

St. Error Signific<strong>an</strong>ce<br />

Const<strong>an</strong>t (predicted SWE value) 30.5 1.9 —<br />

Depth 0.149 0.111 0.043<br />

Density –2.899 10.148 0.776<br />

302


RESULTS<br />

Figure 2. Flow chart of regression <strong>an</strong>alyses<br />

The results are presented in two parts. First, <strong>the</strong> remotely sensed SWE estimates are compared<br />

with <strong>the</strong> SWE me<strong>as</strong>urements obtained from <strong>the</strong> coincident sampling sites. This is followed by <strong>an</strong><br />

<strong>an</strong>alysis of <strong>the</strong> linear regression models.<br />

In-situ vs. Remotely Sensed SWE Estimates<br />

The SSM/I <strong>an</strong>d swath AMSR-E SWE data provided <strong>the</strong> closest estimates to both in-situ data<br />

sets. This w<strong>as</strong> expected, because: i) <strong>the</strong> MSC algorithm used to derive SWE estimates w<strong>as</strong> actually<br />

developed for SSM/I TB, <strong>an</strong>d ii) <strong>the</strong> swath AMSR-E TB have not been re-sampled, thus, <strong>the</strong>y are<br />

truer representations of <strong>the</strong> interaction between <strong>the</strong> sensor <strong>an</strong>d <strong>the</strong> ground surface. Tables 3 <strong>an</strong>d 4<br />

illustrate <strong>the</strong> number (<strong>an</strong>d percentage in brackets) of sampling sites that were found to be<br />

equivalent (i.e. within ±20 mm SWE) to <strong>the</strong> remotely sensed SWE estimates. Table 3 shows <strong>the</strong><br />

results of <strong>the</strong> <strong>Snow</strong>-Only data set, while Table 4 presents <strong>the</strong> results of <strong>the</strong> Actual-Conditions data<br />

set. These results also show a slight incre<strong>as</strong>e in algorithm agreement when only <strong>the</strong> amount of<br />

snow found at each sampling site is included in <strong>the</strong> ground-truth observations (i.e. Core_SWE<br />

from Table 3 vs. Table 4).<br />

303


Table 3: Number of equivalent <strong>Snow</strong>-Only coincident sites (n=88)<br />

12.5k<br />

25k<br />

Swath<br />

SWE Calculation SSM/I<br />

55<br />

AMSR-E AMSR-E AMSR-E<br />

Core_SWE<br />

(63%)<br />

50<br />

42 (48%) 20 (23%) 56 (64%)<br />

Fractional_Core_SWE (57%)<br />

58<br />

29 (33%) 11 (13%) 49 (56%)<br />

Derived_SWE<br />

(66%)<br />

45<br />

32 (36%) 15 (17%) 54 (61%)<br />

Fractional_Derived_SWE (51%) 24 (27%) 7 (8%) 45 (51%)<br />

Table 4: Number of equivalent Actual-Conditions coincident sites (n=88)<br />

SWE Calculation SSM/I<br />

51<br />

Core_SWE<br />

(58%)<br />

44<br />

Fractional_Core_SWE (50%)<br />

45<br />

Derived_SWE<br />

(51%)<br />

33<br />

Fractional_Derived_SWE (38%)<br />

12.5k<br />

AMSR-E<br />

34<br />

(39%)<br />

27<br />

(31%)<br />

26<br />

(30%)<br />

23<br />

(26%)<br />

304<br />

25k<br />

AMSR-E<br />

13<br />

(15%)<br />

10<br />

(11%)<br />

9<br />

(10%)<br />

7<br />

(8%)<br />

Swath<br />

AMSR-E<br />

52<br />

(59%)<br />

45<br />

(51%)<br />

49<br />

(56%)<br />

40<br />

(45%)<br />

By weighting <strong>the</strong> in-situ SWE values by <strong>the</strong> percentage of snow cover found at <strong>the</strong> sites (i.e.<br />

reading down each column) <strong>the</strong> agreement with <strong>the</strong> remote sensing estimates decre<strong>as</strong>es by <strong>an</strong><br />

average of approximately 10% in <strong>the</strong> <strong>Snow</strong>-Only data set, <strong>an</strong>d by <strong>an</strong> average of approximately 7%<br />

in <strong>the</strong> Actual-Conditions data set. Fur<strong>the</strong>r, <strong>the</strong> remote sensing SWE algorithm generally had a<br />

higher level of agreement with Core_SWE me<strong>as</strong>urements th<strong>an</strong> with Derived_SWE values.<br />

Therefore, for a patchy snow cover, it appears that <strong>the</strong> MSC SWE algorithm had <strong>the</strong> closest<br />

agreement with ground SWE me<strong>as</strong>urements b<strong>as</strong>ed only on <strong>the</strong> core samples<br />

Fur<strong>the</strong>r <strong>an</strong>alyses using only <strong>the</strong>se Core_SWE me<strong>as</strong>urements found that, on average, <strong>the</strong> remote<br />

sensing algorithm tended to overestimate <strong>the</strong> patchy in-situ SWE me<strong>as</strong>urements in all c<strong>as</strong>es (Table<br />

5). This w<strong>as</strong> not surprising since <strong>the</strong> remote sensing algorithm w<strong>as</strong> originally derived for a<br />

complete snow cover.<br />

Table 5: Me<strong>an</strong> differences in SWE values between remote sensing estimates <strong>an</strong>d in-situ me<strong>as</strong>urements<br />

(for core samples only)<br />

12.5k<br />

25k<br />

Swath<br />

SSM/I AMSR-E AMSR-E AMSR-E<br />

<strong>Snow</strong>-Only Core_SWE 4.8 17.7 32.1 5.4<br />

Actual-Conditions Core_SWE 8.7 21.6 35.9 9.3<br />

We also w<strong>an</strong>ted to examine <strong>the</strong> effect that varying l<strong>an</strong>d covers had on <strong>the</strong> spaceborne SWE<br />

estimates. The r<strong>an</strong>ges in ground SWE me<strong>as</strong>urements were very high, particularly when <strong>the</strong><br />

sampling sites included shelter belts <strong>an</strong>d fallow fields, which were found to have dr<strong>as</strong>tically<br />

different snow conditions th<strong>an</strong> stubble fields <strong>an</strong>d p<strong>as</strong>tures. Table 6 shows that <strong>the</strong>re is little<br />

difference in <strong>the</strong> me<strong>an</strong> SWE values representative of stubble fields (26.1 mm) <strong>an</strong>d p<strong>as</strong>tures (23.0<br />

mm), but great disparity between <strong>the</strong>se values <strong>an</strong>d fallow fields (1.7 mm) <strong>an</strong>d shelter belts (94.8<br />

mm).


Table 6: Actual-Conditions SWE, depth, <strong>an</strong>d density me<strong>as</strong>urements by l<strong>an</strong>d cover type<br />

L<strong>an</strong>d Cover n SWE (mm) Depth (cm) Density (g/cm 3 )<br />

Stubble 54<br />

Fallow 15<br />

P<strong>as</strong>ture 12<br />

Shelter Belt 2<br />

me<strong>an</strong> = 26.1<br />

min = 1.3<br />

max = 66.3<br />

me<strong>an</strong> = 1.7<br />

min = 0<br />

max = 8.8<br />

me<strong>an</strong> = 23.0<br />

min = 5.1<br />

max = 44.1<br />

me<strong>an</strong> = 94.8<br />

min = 75.6<br />

max = 114.0<br />

305<br />

me<strong>an</strong> = 7.9<br />

min = 1.0<br />

max = 19.0<br />

me<strong>an</strong> = 0.8<br />

min = 0<br />

max = 4.7<br />

me<strong>an</strong> = 6.7<br />

min = 1.8<br />

max = 16.4<br />

me<strong>an</strong> = 26.8<br />

min = 22.7<br />

max = 30.9<br />

me<strong>an</strong> = 0.226<br />

min = 0.026<br />

max = 0.511<br />

me<strong>an</strong> = 0.054<br />

min = 0<br />

max = 0.265<br />

me<strong>an</strong> = 0.219<br />

min = 0.110<br />

max = 0.364<br />

me<strong>an</strong> = 0.292<br />

min = 0.267<br />

max = 0.316<br />

Linear Regression Models<br />

Linear regressions were performed using <strong>the</strong> Statistical Package for <strong>the</strong> Social Sciences (SPSS)<br />

software. The regressions were performed using <strong>the</strong> remotely sensed SWE estimates <strong>an</strong>d<br />

brightness temperatures <strong>as</strong> <strong>the</strong> dependent variables <strong>an</strong>d <strong>the</strong> in-situ observations <strong>as</strong> <strong>the</strong> independent<br />

variables. Analyses were again performed between <strong>the</strong> <strong>Snow</strong>-Only <strong>an</strong>d Actual-Conditions data<br />

sets. Table 7 lists <strong>an</strong>d describes <strong>the</strong> in-situ observations used in all of <strong>the</strong> regression models.<br />

Table 7: Me<strong>as</strong>ured snow properties used in linear regression models<br />

Independent variable Variable Type Description<br />

Percentage of snow cover found at<br />

Percent_<strong>Snow</strong>_Cover Ratio<br />

each sampling site.<br />

Depth Ordinal Me<strong>an</strong> snow pack depth of each site.<br />

Density Ratio Me<strong>an</strong> snow pack density of each site.<br />

Air_Temp Interval Air temperature recorded at each site.<br />

<strong>Snow</strong>/ground interface temperature<br />

Ground_Temp Interval recorded from <strong>the</strong> snow pit.<br />

Number of snow pack layers found in<br />

Num_Layers Ordinal <strong>the</strong> snow pit.<br />

L<strong>an</strong>d_Cover Nominal Type of l<strong>an</strong>d cover.<br />

Indicates whe<strong>the</strong>r or not one or more<br />

ice lenses were found in <strong>the</strong> snow pack<br />

Ice_Lens Binary/Nominal of <strong>the</strong> snow pit.<br />

Number of ice lenses found in <strong>the</strong><br />

Num_Lenses Ordinal snow pack of <strong>the</strong> snow pit.<br />

Total ice lens thickness found within<br />

Total_Lens_Thickness Ordinal <strong>the</strong> snow pack of <strong>the</strong> snow pit.<br />

The linear regression <strong>an</strong>alyses found that snow pack densities from <strong>the</strong> Actual-Conditions data<br />

set were signific<strong>an</strong>tly positively correlated with SWE estimates derived from <strong>the</strong> Swath AMSR-E<br />

data. This w<strong>as</strong> expected since dense <strong>an</strong>d complex snow packs have been shown to amplify<br />

scattering <strong>an</strong>d tend to produce remotely sensed SWE overestimates (Sokol et al., 1999). The<br />

Swath AMSR-E imagery w<strong>as</strong> <strong>the</strong> only remote sensing data set to show such a statistically<br />

signific<strong>an</strong>t correlation, likely because — since <strong>the</strong>y had not been re-sampled to <strong>the</strong> EASE-Grid —<br />

<strong>the</strong>se data were closer representations of <strong>the</strong> original Earth radi<strong>an</strong>ces originally detected by <strong>the</strong><br />

sensor.


Table 8 shows <strong>the</strong> regression results between <strong>the</strong> Swath AMSR-E SWE estimates <strong>an</strong>d both <strong>the</strong><br />

<strong>Snow</strong>-Only <strong>an</strong>d Actual-Conditions data sets. The Model Fit shows that both data sets match <strong>the</strong><br />

satellite estimates at 99% levels of confidence. From <strong>the</strong> Unst<strong>an</strong>dardized Beta Coefficients we see<br />

that <strong>the</strong> only signific<strong>an</strong>t variables are <strong>the</strong> percentage of snow cover, <strong>an</strong>d whe<strong>the</strong>r or not one or<br />

more ice lenses were found in <strong>the</strong> snow pit. Interestingly, density in <strong>the</strong> <strong>Snow</strong>-Only data set<br />

appears not to make a signific<strong>an</strong>t contribution, while it is found to be signific<strong>an</strong>t at a 90% level of<br />

confidence in <strong>the</strong> Actual-Conditions data set. The Model Summaries indicate that <strong>the</strong> proportion<br />

of <strong>the</strong> vari<strong>an</strong>ce in <strong>the</strong> satellite estimates that is explained by <strong>the</strong> ground observations are just<br />

15.8% <strong>an</strong>d 17.9% for <strong>the</strong> <strong>Snow</strong>-Only <strong>an</strong>d Actual-Conditions data sets, respectively.<br />

Table 8: Regression results between Swath AMSR-E SWE estimates <strong>an</strong>d in-situ observations<br />

(signific<strong>an</strong>t coefficients are shown in bold italics)<br />

Swath_AMSR-E SWE <strong>Snow</strong>-Only Actual-Conditions<br />

1) Model Fit<br />

2) Model Summary (Adjusted<br />

0.008 0.004<br />

R 2 ) 0.158 0.179<br />

St.<br />

St.<br />

3) Coefficients Beta Error Sig. Beta Error Sig.<br />

� Const<strong>an</strong>t (Predicted SWE) 32.646 2.280 — 32.552 2.140 ⎯<br />

� <strong>Snow</strong>_Cover_Percent 8.508 4.276 0.050 10.079 4.310 0.022<br />

� Depth 0.143 0.220 0.518 0.187 0.216 0.390<br />

� Density –6.009 6.646 0.369 19.131 11.357 0.096<br />

� Air_Temp –0.050 0.306 0.870 0.026 0.302 0.932<br />

� Ground_Temp –0.162 0.492 0.743 –0.257 0.489 0.601<br />

� Num_Layers 0.371 1.500 0.805 0.691 1.498 0.646<br />

� L<strong>an</strong>d_Cover –1.555 1.376 0.262 –1.490 1.355 0.275<br />

� Ice_Lens –7.354 3.369 0.032 –6.738 3.267 0.043<br />

� Num_Lenses 0.563 3.205 0.861 –0.151 3.204 0.963<br />

� Total_Lens_Thickness 0.108 0.647 0.868 0.214 0.643 0.740<br />

Similar regressions were run between all of <strong>the</strong> remote sensing <strong>an</strong>d in-situ data sets. The results<br />

are summarized in Table 9. Regression models derived for <strong>the</strong> 12.5k AMSR-E 18.7v <strong>an</strong>d 25k<br />

AMSR-E SWE data were not statistically signific<strong>an</strong>t. Regressions from <strong>the</strong> AMSR-E TB found that<br />

<strong>the</strong> 18.7v TB resulted in having more signific<strong>an</strong>t variables th<strong>an</strong> those collected from <strong>the</strong> 36.5v TB.<br />

While <strong>the</strong> <strong>Snow</strong>_Cover_Percent, <strong>an</strong>d Ice_Lens variables were found to be signific<strong>an</strong>t in both TB<br />

regressions, Depth <strong>an</strong>d Ground_Temp were found to also be signific<strong>an</strong>t in <strong>the</strong> 18.7v TB regression.<br />

With <strong>the</strong> 12.5k AMSR-E data it is interesting to note that in comparison to <strong>the</strong> non-gridded<br />

Swath_AMSR-E <strong>an</strong>alyses a completely different set of variables, except for <strong>the</strong> binary variable<br />

Ice_Lens, w<strong>as</strong> found to be signific<strong>an</strong>t. The signific<strong>an</strong>t variables in this data set include: Depth,<br />

Air_Temp, L<strong>an</strong>d_Cover, <strong>an</strong>d Ice_Lens. However, <strong>as</strong> <strong>the</strong>se data have been re-sampled, <strong>the</strong>re is less<br />

confidence in <strong>the</strong>se regression results compared to those of <strong>the</strong> swath results. Unlike <strong>the</strong> 12.5k<br />

AMSR-E regression results, <strong>the</strong> 25k AMSR-E 18.7v TB were found to be signific<strong>an</strong>t, <strong>an</strong>d <strong>the</strong> 36.5v<br />

TB were marginally signific<strong>an</strong>t. The 25k SSM/I SWE regressions produced nearly identical results<br />

between <strong>the</strong> <strong>Snow</strong>-Only <strong>an</strong>d Actual-Conditions data sets.<br />

306


307<br />

Table 9: Regression results between remotely sensed SWE estimates <strong>an</strong>d in-situ observations (· denotes a statistically signific<strong>an</strong>t correlation)<br />

Swath<br />

AMSR<br />

-E SWE<br />

S A<br />

-O -C<br />

Swath<br />

AMSR<br />

-E 18.7v<br />

S A<br />

-O -C<br />

Swath<br />

AMSR<br />

-E 36.5v<br />

S A<br />

-O -C<br />

12.5k<br />

AMSR-E<br />

SWE<br />

S A<br />

-O -C<br />

12.5k<br />

AMSR-E<br />

18.7v<br />

S A<br />

-O -C<br />

12.5k<br />

AMSR-E<br />

36.5v<br />

S A<br />

-O -C<br />

25k<br />

AMSR-E<br />

SWE<br />

S A<br />

-O -C<br />

25k<br />

AMSR-E<br />

18.7v<br />

S A<br />

-O -C<br />

25k<br />

AMSR-E<br />

36.5v<br />

S A<br />

-O -C<br />

25k<br />

SSM/I<br />

SWE<br />

S A<br />

-O -C<br />

1) Model Fit · · · · · · · · · · · · · · · ·<br />

2) Model Summary<br />

(Adjusted R 2 0 0 0 0 0 0 0 0<br />

0 0<br />

0 0 0 0 0 0<br />

)<br />

.2 .2 .3 .3 .2 .3 .3 .3<br />

.3 .3<br />

.3 .3 .1 .1 .1 .1<br />

3) Coefficients<br />

� <strong>Snow</strong>_Cover_Percent<br />

� Depth<br />

� Density<br />

� Air_Temp<br />

� Ground_Temp<br />

� Num_Layers<br />

� L<strong>an</strong>d_Cover<br />

� Ice_Lens<br />

� Num_Lenses<br />

� Total_Lens_Thickness<br />

· ·<br />

·<br />

· ·<br />

· ·<br />

·<br />

· ·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

· ·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

· ·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

S-O: <strong>Snow</strong>-Only<br />

A-C: Actual Conditions


CONCLUSIONS AND DISCUSSION<br />

Although statistically signific<strong>an</strong>t models were established between m<strong>an</strong>y of <strong>the</strong> remote sensing<br />

<strong>an</strong>d in-situ data sets, <strong>the</strong> proportion of <strong>the</strong> vari<strong>an</strong>ce in <strong>the</strong> satellite estimates that could be<br />

explained by <strong>the</strong> ground observations (i.e. <strong>the</strong> Model Summaries) w<strong>as</strong>, at most, 0.31. This<br />

suggests that ei<strong>the</strong>r <strong>the</strong> ground data are insufficient for deriving SWE from spaceborne p<strong>as</strong>sive<br />

microwave observations or that <strong>the</strong> remote sensing data were inappropriate. We know from<br />

previous research (reviewed earlier) that it is possible to obtain reliable SWE estimates through<br />

remote sensing, so we must conclude that <strong>the</strong>re were problems with remote sensing data we used<br />

in this experiment. Specifically, <strong>the</strong> continuous snow cover <strong>as</strong>sumption embedded in <strong>the</strong> MSC<br />

p<strong>as</strong>sive microwave SWE algorithm does not produce acceptable results over a patchy snow cover.<br />

The poor perform<strong>an</strong>ce of <strong>the</strong> MSC SWE algorithm for each remote sensing data set evaluated<br />

confirms that <strong>the</strong> algorithm fails under patchy <strong>an</strong>d variable snow conditions.<br />

In spite of <strong>the</strong> poorly articulated regression models, <strong>the</strong>re were several in-situ observations that<br />

appear to play <strong>an</strong> import<strong>an</strong>t role in affecting <strong>the</strong> satellite p<strong>as</strong>sive microwave data. The presence or<br />

absence of <strong>an</strong> ice lens in <strong>the</strong> snow pack w<strong>as</strong> consistently identified <strong>as</strong> a signific<strong>an</strong>t coefficient in<br />

<strong>the</strong> regression <strong>an</strong>alyses. O<strong>the</strong>r observations that may prove to be useful include <strong>the</strong> percent snow<br />

cover, snow depth, <strong>an</strong>d <strong>the</strong> ground temperature. These will need to be investigated fur<strong>the</strong>r.<br />

Consideration of patchy snow cover is challenging from a ground sampling perspective,<br />

however this study shows that <strong>the</strong> actual conditions found at each sampling site must be<br />

incorporated in ground-truth data sets when collecting observations over a partial snow cover.<br />

Subsequent <strong>an</strong>alysis will focus on using optical data to determine snow cover fraction within a<br />

p<strong>as</strong>sive microwave grid cell to greater qu<strong>an</strong>tify <strong>the</strong> impact of patchy snow cover.<br />

ACKNOWLEDGEMENTS<br />

Support from Environment C<strong>an</strong>ada’s CRYSYS (Cryosphere System in C<strong>an</strong>ada) research<br />

initiative is greatly appreciated. Special th<strong>an</strong>ks are extended to Nat<strong>as</strong>ha Neum<strong>an</strong>n, Arvids Silis,<br />

<strong>an</strong>d Peter Toose (all from <strong>the</strong> Meteorological Service of C<strong>an</strong>ada) for equipment <strong>an</strong>d data support.<br />

The EASE-Grid brightness temperatures were obtained from MSC through <strong>the</strong> EOSDIS National<br />

<strong>Snow</strong> <strong>an</strong>d Ice Data Center Distributed Active Archive Center (NSIDC DAAC), University of<br />

Colorado at Boulder. Acknowledgements are also extended to Aaron Fedje, Kari Geller, Kathie<br />

Legault, Mark Otterson, Greg Peterson, Sus<strong>an</strong> Rever, <strong>an</strong>d Mauricio Jimenez Salazar (all of <strong>the</strong><br />

University of Regina), <strong>an</strong>d Michelle Y<strong>as</strong>kowich (Nature S<strong>as</strong>katchew<strong>an</strong>) for collection of in-situ<br />

me<strong>as</strong>urements <strong>an</strong>d observations.<br />

REFERENCES<br />

Armstrong, R. L. <strong>an</strong>d M. J. Brodzik, 1995. An Earth-gridded SSM/I data set for cryospheric<br />

studies <strong>an</strong>d global ch<strong>an</strong>ge monitoring, Adv<strong>an</strong>ces in Space Research 16(10): 155-163.<br />

Derksen, C., A. Walker, E. LeDrew, B. Goodison, 2002. Time-series <strong>an</strong>alysis of p<strong>as</strong>sivemicrowave-derived<br />

central North Americ<strong>an</strong> snow water equivalent imagery, Annals of<br />

Glaciology, 34: 1-7.<br />

Derksen, C., A. Walker, B. Goodison, 2003a. A comparison of 18 winter se<strong>as</strong>ons of in situ <strong>an</strong>d<br />

p<strong>as</strong>sive microwave-derived snow water equivalent estimates in Western C<strong>an</strong>ada. Remote<br />

Sensing of Environment, 88: 271-282.<br />

Derksen, C., <strong>an</strong>d A. Walker, E. LeDrew, B. Goodison, 2003b. Combining SMMR <strong>an</strong>d SSM/I Data<br />

for Time Series Analysis of Central North Americ<strong>an</strong> <strong>Snow</strong> Water Equivalent. Journal of<br />

Hydrometeorology, 4: 304-316.<br />

Goita, K., A.E. Walker, B.E. Goodison, <strong>an</strong>d A.T.C. Ch<strong>an</strong>g, 1997. Estimation of <strong>Snow</strong> Water<br />

Equivalent in <strong>the</strong> Boreal Forest Using P<strong>as</strong>sive Microwave Data. Proceedings, GER '97<br />

(International Symposium: Geomatics in <strong>the</strong> Era of Radarsat), Ottawa, C<strong>an</strong>ada, May 25-30,<br />

1997. CD-ROM published by Natural Resources C<strong>an</strong>ada <strong>an</strong>d National Defence.<br />

308


Goita, K., A. Walker, B. Goodison, 2003. Algorithm development for <strong>the</strong> estimation of snow<br />

water equivalent in <strong>the</strong> boreal forest using p<strong>as</strong>sive microwave data. International Journal of<br />

Remote Sensing 24(5): 1097-1102.<br />

Goodison, B., 1989. Determination of areal snow water equivalent on <strong>the</strong> C<strong>an</strong>adi<strong>an</strong> Prairies using<br />

p<strong>as</strong>sive microwave satellite data. Proceedings, 1989 International Geoscience <strong>an</strong>d Remote<br />

Sensing Symposium, V<strong>an</strong>couver. 3:1243-1246.<br />

Goodison, B.E. <strong>an</strong>d A.E. Walker, 1994. C<strong>an</strong>adi<strong>an</strong> development <strong>an</strong>d use of snow cover information<br />

from p<strong>as</strong>sive microwave satellite data. In P<strong>as</strong>sive Microwave Remote Sensing of L<strong>an</strong>d-<br />

Atmosphere Interactions, Choudhury, B.J., Y.H. Kerr, E.G. Njoku, <strong>an</strong>d P. Pampaloni (eds.),<br />

VSP, Utrecht, The Ne<strong>the</strong>rl<strong>an</strong>ds, 245-262.<br />

Gupta, V., 2000. Regression explained in simple terms. Obtained online on April 24, 2005, from<br />

http://www.telecom.csuhayward.edu/~esuess/Links/Software/RegressionExplained/regression_<br />

explained.doc<br />

Hare, F.K. <strong>an</strong>d M.K. Thom<strong>as</strong>, 1979. Climate C<strong>an</strong>ada, 2 nd Ed., C<strong>an</strong>ada: Wiley.<br />

Laycock, Arleigh H., 1972. The Diversity of <strong>the</strong> Physical L<strong>an</strong>dscape. In The Prairie Provinces,<br />

Smith, P.J. (ed.), University of Toronto Press.<br />

O’Sulliv<strong>an</strong>, David, <strong>an</strong>d David J. Unwin, 2003. Geographic Information Analysis, Hoboken, NJ:<br />

Wiley.<br />

Singh, Purushottam Raj <strong>an</strong>d Thi<strong>an</strong> Yew G<strong>an</strong>, 2000. Retrieval of <strong>Snow</strong> Water Equivalent Using<br />

P<strong>as</strong>sive Microwave Brightness Temperature Data. Remote Sensing of Environment, 74: 275-<br />

286.<br />

Sokol, J., T.J. Pultz, <strong>an</strong>d A.E.Walker, 1999. P<strong>as</strong>sive <strong>an</strong>d Active Spaceborne Microwave Remote<br />

Sensing of <strong>Snow</strong> Cover. Proceedings, 4 th International Spaceborne Remote Sensing<br />

Conference/21 st C<strong>an</strong>adi<strong>an</strong> Symposium on Remote Sensing.<br />

Turchenek, Kim R., 2004. Validation of <strong>Snow</strong> Water Equivalents Derived from P<strong>as</strong>sive<br />

Microwave Brightness Temperatures over a Prairie Environment. Unpublished Undergraduate<br />

Thesis, submitted to <strong>an</strong>d accepted by <strong>the</strong> Department of Geography at <strong>the</strong> University of Regina,<br />

S<strong>as</strong>katchew<strong>an</strong>.<br />

Walker, Anne, Barry Goodison, Michael Davey <strong>an</strong>d David Olson, 1995. Atl<strong>as</strong> of Sou<strong>the</strong>rn<br />

C<strong>an</strong>adi<strong>an</strong> Prairies Winter <strong>Snow</strong> Cover from Satellite P<strong>as</strong>sive Microwave Data: November<br />

1978 to March 1986, Downsview, C<strong>an</strong>ada: Atmospheric Environment Service, Environment<br />

C<strong>an</strong>ada.<br />

309


This page is intentionally bl<strong>an</strong>k.<br />

310


TO ERR IS HUMAN, TO FORGET IT WOULD BE DIVINE.<br />

SNO-FOO 2006–Andrew Klein<br />

The E<strong>as</strong>tern <strong>Snow</strong> Conference <strong>an</strong>nually bestows this award on some poor hapless snow <strong>an</strong>d ice researcher, who,<br />

while striving to push back <strong>the</strong> frontiers of science, is overcome by a bout of “bone-headedness,” ill pl<strong>an</strong>ning, or just<br />

plain bad luck—in military parl<strong>an</strong>ce, referred to <strong>as</strong> SNAFU.<br />

This year’s winner h<strong>as</strong> made countless questionable decisions when it comes to snow research. Although he<br />

professes to know something about mapping snow <strong>an</strong>d ice from satellites, his geographic knowledge of snow must<br />

be questioned. He seemed blissfully unaware, when moving to Tex<strong>as</strong>, that <strong>the</strong> state isn’t all that well known for<br />

having snow to study. He also seems to think that <strong>the</strong> tropics are <strong>the</strong> world’s hotbed for glaciers, which he appears to<br />

strenuously study while hiding in his air-conditioned office from <strong>the</strong> Tex<strong>as</strong> summer heat.<br />

This individual claims that he actually conducts fieldwork in Antarctica. However, it is rumored (though never<br />

proven) that when one of his sampling sites actually is snow-covered, he moves it to a snow-free location so he c<strong>an</strong><br />

more e<strong>as</strong>ily sample <strong>the</strong> soil—<strong>an</strong>d he claims to consider himself to be a snow scientist!<br />

None of this, however, comes close to <strong>the</strong> egregious mistakes this year’s winner h<strong>as</strong> made while conducting<br />

ESC business. The ESC, for re<strong>as</strong>ons still not fully understood, placed in <strong>the</strong> h<strong>an</strong>ds of this individual its most public<br />

face to <strong>the</strong> world, its Web site. On numerous occ<strong>as</strong>ions, ESC individuals have fr<strong>an</strong>tically pointed out mistakes in <strong>the</strong><br />

“content” this individual h<strong>as</strong> chosen to post on <strong>the</strong> org<strong>an</strong>ization’s Web site for <strong>the</strong> <strong>entire</strong> world to see. Yet knowing<br />

full well this individual’s org<strong>an</strong>izational abilities, <strong>the</strong> ESC executive actually placed <strong>the</strong> org<strong>an</strong>ization of its <strong>an</strong>nual<br />

meeting in his h<strong>an</strong>ds…!<br />

It probably is not surprising, <strong>the</strong>refore, that in <strong>the</strong> course of producing <strong>the</strong> meeting materials, this year’s winner<br />

h<strong>as</strong> not once, but multiple times, shown complete <strong>an</strong>d utter disrespect for <strong>the</strong> holder of <strong>the</strong> highest office in <strong>the</strong><br />

ESC—its President. On numerous occ<strong>as</strong>ions, <strong>the</strong> ESC president h<strong>as</strong> shaken his head at <strong>the</strong> “interesting” ways his<br />

name h<strong>as</strong> been spelled in <strong>the</strong> program <strong>an</strong>d on <strong>the</strong> Web site by this year’s award winner. Is it Claude Dugay, Claude<br />

Dugaey, or maybe Claude Duguay?<br />

In fact, this year’s winner probably c<strong>an</strong>’t even be counted on to spell <strong>the</strong> name of this award correctly on <strong>the</strong><br />

Web site. So, Andrew Klein, is it <strong>the</strong> SNO-FOO or SNOWFOO Award that h<strong>as</strong> been bestowed upon you?<br />

311

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!