4
GEOMORPHOLOGY Erosion by an Alpine glacier Frédéric Herman, 1 * Olivier Beyssac, 2 Mattia Brughelli, 1 Stuart N. Lane, 1 Sébastien Leprince, 3 Thierry Adatte, 4 Jiao Y. Y. Lin, 5 Jean-Philippe Avouac, 3 Simon C. Cox 6 Assessing the impact of glaciation on Earths surface requires understanding glacial erosion processes. Developing erosion theories is challenging because of the complex nature of the erosion processes and the difficulty of examining the ice/bedrock interface of contemporary glaciers. We demonstrate that the glacial erosion rate is proportional to the ice-sliding velocity squared, by quantifying spatial variations in ice-sliding velocity and the erosion rate of a fast-flowing Alpine glacier. The nonlinear behavior implies a high erosion sensitivity to small variations in topographic slope and precipitation. A nonlinear rate law suggests that abrasion may dominate over other erosion processes in fast-flowing glaciers. It may also explain the wide range of observed glacial erosion rates and, in part, the impact of glaciation on mountainous landscapes during the past few million years. G laciers and icecaps played a major role in shaping the morphology of mid- to high- latitude mountain ranges during the Quater- nary period, spanning the past 2.6 million years of Earths history. Observations sug- gest that they have also played a fundamental role in the evolution of Earths climate through a system of positive feedbacks that involves climate, tec- tonics, and erosion (14). Glaciers erode their underlying bedrock mainly through abrasion and quarrying, which theories predict to be propor- tional to ice-sliding velocity raised to some power (57). Numerical models reproduce typical glacial landscape features, such as U-shaped valleys (3, 8), hanging valleys (9, 10), glacial cirques (10, 11), or fjords (12, 13), by implementing these relation- ships. Despite great advances in the sophistica- tion of these models through the inclusion of high-order ice dynamics ( 10), subglacial hydrology ( 10, 11 , 1315 ), or thermodynamics of water flow (14, 16), they also include poorly constrained parameters. Erosion lawsproportionality con- stants and velocity exponents are particularly uncertain (58, 11, 13). Estimates of glacial erosion rates ranging from annual to million-year time scales come from monitoring the sediment yield from glacial streams (1721) and using geochronometric methods (4, 18), respectively. Despite providing key information about the pace at which glaciers may shape moun- tainous landscapes, these studies have not es- tablished an accurate law for glacial erosion. Furthermore, estimates of glacial erosion rates vary by four orders of magnitude from polar to tem- perate regions on Earth (4, 17, 18). Existing theories do not reproduce such variations. Therefore, our current understanding of the link between climate and glacial erosion suffers from poor constraints on what controls spatial and temporal erosion var- iability in response to global changes in precipita- tion and temperature. We designed this study to specifically constrain how glacial erosion relates to ice-sliding velocity. We simultaneously quantified erosion rates and sliding velocity during a 5-month period, from November 2013 to April 2014, over the entire Franz Josef Glacier, New Zealand. This glacier exhibits surface velocities that are largely dominated by high sliding velocities on the bedrock (22), up to about 3 m/day. We measured these high velocities accurately from remote sensing and expected to find large erosion rates. The analysis of continu- ous suspended sediment load indicated very high erosion rates (about 10 mm/year), whereas glacial sediment production remained lower than the transport capacity of the glacial system (23). We also found that the glacial sediments come pre- dominantly from under the glacier, based on the mineralogy, fossil organic carbon, and the very low fraction of modern organic carbon found in the glacial stream (23). These observations imply that sediments collected at the glacier front can be used to constrain the glacial erosion law. We introduce here a method to measure sur- face displacement in three dimensions at a 1-m ground resolution and centimetric accuracy, using DigitalGlobe Worldview stereopair images (23). The results confirm fast velocities for most parts of the glacier (Fig. 1) dominated by sliding (22, 23). In addition, we observed similar velocity patterns during the austral summers 20122013 and 20132014, indicating steady spatial patterns of sliding. Extremely high snow accumulation rates of 4 to 8 m/year (water equivalent) (24) and steep topog- raphy account for such high velocities (22). We exploited the geology of the Southern Alps of New Zealand to determine how erosion varies spatially. This small mountain range resulted from the continental collision between the Australian and Pacific Plates, along a major plate boundary named the Alpine Fault (25), which led to a sharp metamorphic gradient within a 15-km distance (Fig. 2). Rocks adjacent to the Alpine Fault have experienced peak metamorphic temperatures up to about 650°C, whereas rocks about 15 km far- ther southeast have only experienced 300°C. The Franz Josef Glacier flows almost parallel to this temperature gradient. The rocks are highly frac- tured but have uniform, steep bedding and foliation (60° to 80°) without kilometer-scale variations in strength or erodability across the catchment. The rocks also contain fossil organic carbon (26), which can quantify the peak metamorphic temperature conditions based on Raman spectroscopy of car- bonaceous material (RSCM) (27). By comparing RSCM temperature data in samples collected from SCIENCE sciencemag.org 9 OCTOBER 2015 VOL 350 ISSUE 6257 193 1 Institute of Earth Surface Dynamics, University of Lausanne, CH-1015 Lausanne, Switzerland. 2 Institut de Minéralogie, de Physique des Matériaux, et de Cosmochimie, Sorbonne Universités, UMR CNRS 7590, Université Pierre et Marie Curie Paris 06, Muséum National dHistoire Naturelle, Institut de Recherche pour le Développement UMR 206, 4 place Jussieu, 75005 Paris, France. 3 Division of Geological and Planetary Science, California Institute of Technology, Pasadena, CA, USA. 4 Institute of Earth Sciences, University of Lausanne, CH-1015 Lausanne, Switzerland. 5 Division of Engineering and Applied Science, California Institute of Technology, Pasadena, CA, USA. 6 Institute of Geological and Nuclear Sciences, Private Bag 1930, Dunedin, New Zealand. *Corresponding author. E-mail: [email protected] Fig. 1. Franz Josef Glacier surface velocity. The Franz Josef Glacier (Ka Roimata o Hinehukatere in Māori) is located in Westland Tai Poutini National Park on the west coast of New Zealand's South Island. (A) Surface velocity measured in summer 2013 (integrated over 10 days). (B) Surface veloc- ity measured in summer 2014 (integrated over 12 days). The three-dimensional (3D) velocities were derived from the measurements of the 3D displace- ment derived from Worldview stereo images (23). Velocity (m/day) -43.53° -43.49° -43.46° 170.16° 170.21° 170.26° Austral summer 2013 0.5 1.0 1.5 2.0 2.5 3.0 170.16° 170.21° 170.26° Austral summer 2014 4 km RESEARCH | REPORTS on October 26, 2015 www.sciencemag.org Downloaded from on October 26, 2015 www.sciencemag.org Downloaded from on October 26, 2015 www.sciencemag.org Downloaded from

Erosion by an Alpine glacier

  • Upload
    ledung

  • View
    221

  • Download
    2

Embed Size (px)

Citation preview

Page 1: Erosion by an Alpine glacier

GEOMORPHOLOGY

Erosion by an Alpine glacierFrédéric Herman,1* Olivier Beyssac,2 Mattia Brughelli,1 Stuart N. Lane,1

Sébastien Leprince,3 Thierry Adatte,4 Jiao Y. Y. Lin,5

Jean-Philippe Avouac,3 Simon C. Cox6

Assessing the impact of glaciation on Earth’s surface requires understanding glacialerosion processes. Developing erosion theories is challenging because of the complexnature of the erosion processes and the difficulty of examining the ice/bedrockinterface of contemporary glaciers. We demonstrate that the glacial erosion rate isproportional to the ice-sliding velocity squared, by quantifying spatial variations inice-sliding velocity and the erosion rate of a fast-flowing Alpine glacier. The nonlinearbehavior implies a high erosion sensitivity to small variations in topographic slope andprecipitation. A nonlinear rate law suggests that abrasion may dominate over other erosionprocesses in fast-flowing glaciers. It may also explain the wide range of observed glacialerosion rates and, in part, the impact of glaciation on mountainous landscapes during thepast few million years.

Glaciers and icecaps played a major role inshaping the morphology of mid- to high-latitudemountain rangesduring theQuater-nary period, spanning the past 2.6 millionyears of Earth’s history. Observations sug-

gest that they have also played a fundamental rolein the evolution of Earth’s climate through a systemof positive feedbacks that involves climate, tec-tonics, and erosion (1–4). Glaciers erode theirunderlying bedrockmainly through abrasion andquarrying, which theories predict to be propor-tional to ice-sliding velocity raised to some power(5–7). Numerical models reproduce typical glaciallandscape features, such asU-shaped valleys (3,8),hanging valleys (9, 10), glacial cirques (10, 11), orfjords (12, 13), by implementing these relation-

ships. Despite great advances in the sophistica-tion of these models through the inclusion ofhigh-order ice dynamics (10), subglacial hydrology(10, 11, 13–15), or thermodynamics of water flow(14, 16), they also include poorly constrainedparameters. Erosion laws’ proportionality con-stants and velocity exponents are particularlyuncertain (5–8, 11, 13).Estimates of glacial erosion rates ranging from

annual to million-year time scales come frommonitoring the sediment yield fromglacial streams(17–21) andusing geochronometricmethods (4, 18),respectively. Despite providing key informationabout the pace at which glaciersmay shapemoun-tainous landscapes, these studies have not es-tablished an accurate law for glacial erosion.Furthermore, estimates of glacial erosion rates varyby four orders of magnitude from polar to tem-perate regions onEarth (4, 17, 18). Existing theoriesdo not reproduce such variations. Therefore, ourcurrent understanding of the link between climateand glacial erosion suffers from poor constraintson what controls spatial and temporal erosion var-iability in response to global changes in precipita-tion and temperature.We designed this study to specifically constrain

how glacial erosion relates to ice-sliding velocity.We simultaneously quantified erosion rates andsliding velocity during a 5-month period, from

November 2013 toApril 2014, over the entire FranzJosef Glacier, New Zealand. This glacier exhibitssurface velocities that are largely dominated byhigh sliding velocities on the bedrock (22), up toabout 3m/day.Wemeasured these high velocitiesaccurately from remote sensing and expected tofind large erosion rates. The analysis of continu-ous suspended sediment load indicated very higherosion rates (about 10mm/year), whereas glacialsediment production remained lower than thetransport capacity of the glacial system (23). Wealso found that the glacial sediments come pre-dominantly from under the glacier, based on themineralogy, fossil organic carbon, and the verylow fraction of modern organic carbon found inthe glacial stream (23). These observations implythat sediments collected at the glacier front canbe used to constrain the glacial erosion law.We introduce here a method to measure sur-

face displacement in three dimensions at a 1-mground resolution and centimetric accuracy, usingDigitalGlobe Worldview stereopair images (23).The results confirm fast velocities for most partsof the glacier (Fig. 1) dominated by sliding (22, 23).In addition, we observed similar velocity patternsduring the austral summers 2012–2013 and 2013–2014, indicating steady spatial patterns of sliding.Extremely high snow accumulation rates of 4 to8 m/year (water equivalent) (24) and steep topog-raphy account for such high velocities (22).We exploited the geology of the Southern Alps

of New Zealand to determine how erosion variesspatially. This smallmountain range resulted fromthe continental collision between the Australianand Pacific Plates, along a major plate boundarynamed the Alpine Fault (25), which led to a sharpmetamorphic gradient within a 15-km distance(Fig. 2). Rocks adjacent to the Alpine Fault haveexperienced peakmetamorphic temperatures upto about 650°C, whereas rocks about 15 km far-ther southeast have only experienced 300°C. TheFranz Josef Glacier flows almost parallel to thistemperature gradient. The rocks are highly frac-turedbuthaveuniform, steepbedding and foliation(60° to 80°) without kilometer-scale variations instrength or erodability across the catchment. Therocks also contain fossil organic carbon (26), whichcan quantify the peakmetamorphic temperatureconditions based on Raman spectroscopy of car-bonaceous material (RSCM) (27). By comparingRSCM temperature data in samples collected from

SCIENCE sciencemag.org 9 OCTOBER 2015 • VOL 350 ISSUE 6257 193

1Institute of Earth Surface Dynamics, University of Lausanne,CH-1015 Lausanne, Switzerland. 2Institut de Minéralogie, dePhysique des Matériaux, et de Cosmochimie, SorbonneUniversités, UMR CNRS 7590, Université Pierre et MarieCurie Paris 06, Muséum National d’Histoire Naturelle, Institutde Recherche pour le Développement UMR 206, 4 placeJussieu, 75005 Paris, France. 3Division of Geological andPlanetary Science, California Institute of Technology,Pasadena, CA, USA. 4Institute of Earth Sciences, Universityof Lausanne, CH-1015 Lausanne, Switzerland. 5Division ofEngineering and Applied Science, California Institute ofTechnology, Pasadena, CA, USA. 6Institute of Geological andNuclear Sciences, Private Bag 1930, Dunedin, New Zealand.*Corresponding author. E-mail: [email protected]

Fig. 1. Franz Josef Glacier surface velocity. TheFranz Josef Glacier (Ka Roimata o Hinehukatere inMāori) is located in Westland Tai Poutini NationalPark on the west coast of New Zealand's SouthIsland. (A) Surface velocity measured in summer2013 (integrated over 10 days). (B) Surface veloc-ity measured in summer 2014 (integrated over12 days).The three-dimensional (3D) velocities werederived from the measurements of the 3D displace-ment derived from Worldview stereo images (23).

Velocity (m

/day)

-43.53°

-43.49°

-43.46°

170.16° 170.21° 170.26°

Austral summer 2013

0.5

1.0

1.5

2.0

2.5

3.0

170.16° 170.21° 170.26°

Austral summer 2014

4 km

RESEARCH | REPORTS

on

Oct

ober

26,

201

5w

ww

.sci

ence

mag

.org

Dow

nloa

ded

from

o

n O

ctob

er 2

6, 2

015

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

26,

201

5w

ww

.sci

ence

mag

.org

Dow

nloa

ded

from

Page 2: Erosion by an Alpine glacier

the glacial stream, we determined the provenanceof the particulate load found in the glacial stream(23). We were then able to reconstruct the mag-nitude and patterns of erosion.Our results showerosionpatterns closelymimick-

ing the velocity patterns over the 5-month period(fig. S2).Maximum erosion rates occurred aroundthe steeper, faster parts of the glacier. Low erosionrates correspond either to the slowly moving ac-cumulation area or the glacier front. Analysis ofthe temporal evolution of glacial erosion revealsthat instantaneous erosion rates varied by severalorders of magnitude over the 5-month period(from about 1 to 500mm/year; movie S1). Erosionrates were highest in response to large rain events.During that time, water easily reached the ice/bedrock interface to induce an increase of waterpressure, glacier sliding velocity, and erosion rates(13, 15, 20, 28, 29). The water discharge data wecollected confirm a study showing that the sub-glacial drainage system does not evolve from acavity-driven to a channelized system throughtime (23, 29). These observations explain why wecontinuously observed such a high sensitivity towater inputs. Furthermore, the integrated erosionrates over the 5-month period compare well torates integrated over geological time scales (30).This suggests the potential to extrapolate the pro-cesses driving erosion during our observations tolonger time scales.We combined the integrated erosion rates with

the provenance and remote sensing data to con-strain the parameters in an erosion law (Fig. 3A).

We assume that the erosion rate is proportionalto the sliding velocity raised to some power (i.e.,e ¼ Kgjusjl , where e is the erosion rate, Kg is anerodability constant, us is the sliding velocity, andl is an exponent). We constrained Kg and l usingtwo independent methods. The first one is basedon the nonlinear least-squares method, and thesecond is based on Bayesian inversion that en-ables us to construct the probability density func-tion of the constrained parameters (23). Bothapproaches lead to a nonlinear relationship, withan exponent l close to 2 (Fig. 3, A, B, and D). Thisrelationship agrees with theoretical predictionsfor glacial abrasion (5). Abrasion is propor-tional to the product of the viscous drag forceof the ice as it moves on the bedrock and therate atwhichdebris contained in the ice is draggedagainst the bed, which both depend on the slid-ing velocity.Theoreticalmodels for abrasion and quarrying

assume basal sliding to be the primary driver oferosion (5–7). Discriminating which of these twoprocesses dominates is known to be difficult. Thisis in part because other variables than sliding playa role, including lithological variations or sub-glacial fluvial activity. One quarrying model (7)accounts for the effect of bedrock strength heter-ogeneities and variations inwater pressure at theice/bedrock interface. It predicts an erosion expo-nent l < 1 for weak bedrock strength such as thatin the Southern Alps. The model also impliesdecreasing or relatively steady erosion rates withincreasing water pressure (7), in contrast to our

observations of a different relationship (movieS1). This relationship, along with our sliding ve-locity exponent of about 2, may imply an abrasion-dominated process in the Franz Josef Glacier,although quarrying is required to produce rockfragments that abrade the bedrock.Existing field estimates in Alaska (19), the Eu-

ropean Alps (19), and the Patagonian Andes (18)all suggested an exponent l of 1 and a dimension-less Kg around 10−4. Most landscape evolutionmodels use these values. We observed similarvalues when spatially integrating erosion ratesand sliding velocities for the Franz Josef Glacier.Unfortunately, spatially integrated erosion ratesand velocities (19, 20) cannot rigorously constrainKg and l independently because of the trade-offbetween the erosion constant Kg and the expo-nent l (Fig. 3B).Our observations establish that glacial abrasion

is a nonlinear function of ice-sliding velocity. Ice-sliding velocity is to a first order nonlinearly pro-portional to ice thickness and ice surface slope[eq. 9 in (23, 31)], which are both set by the ba-lance between the glacier mass balance and thedivergence of the ice flux [eq. 8 in (23, 31)]. As aresult, an increase of ice flux induced by climatechange, or increased surface slope, will be accom-modated by faster sliding velocities in a nonlinearway. An additional nonlinearity on the sliding ex-ponent in the erosion law (i.e., l > 1) thus implies avery high sensitivity of erosion rates to changes inice flux (23). Therefore, fast-flowing glaciers arelikely to be particularly effective at erosion, and

194 9 OCTOBER 2015 • VOL 350 ISSUE 6257 sciencemag.org SCIENCE

Fig. 2. Bedrock metamorphism and RSCM tempera-ture data in the Southern Alps of New Zealand (tableS2). A metamorphic map (25) and RSCM temperaturedata (23) are shown. Both metamorphism and RSCMtemperature data show a sharp gradient in the hangingwall of the Alpine Fault. The inset is the location of theFranz Josef Glacier shown in fig. S5, which shows thebedrock temperature model used for the provenanceanalysis.

Franz Josef Glacier

Aoraki/Mt Cook

ALPINE FAULT

Fig.S5A

170.5°170°

-43°

-43.

25°

-43.

5°-4

3.75

°

0 10 20 30 405km

RSCM-T bedrock>640˚C600 - 640˚C550 - 600˚C500 - 550˚C450 - 500˚C400 - 450˚C350 - 400˚C<350˚C

Metamorphic mineral zoneprehnite-pumpellyite

pumpellyite-actinolite

chlorite greenschist

biotite greenschist

garnet greenschist

garnet amphibolite

oligoclase amphibolite

K-feldspar amphibolite

ice

RESEARCH | REPORTS

Page 3: Erosion by an Alpine glacier

abrasion might become the dominating processover quarrying for fast-flowing glaciers over weakrocks, such as the Franz Josef Glacier.The nonlinear glacial erosion law may explain

why glacial erosion rates span several orders ofmagnitude, from polar dry regions to temperatealpine glaciers and from soil-mantled hillslopelandscapes to steep, tectonically active mountainranges (4, 17, 18). Atmospheric circulation controlsglobal precipitation, with precipitation increasingfrom the poles to the equator. In addition, thepolar jet stream and its associated westerly windshave a major influence on glacial access to pre-cipitation as they bringmoisture onto continents.Several observations suggest that they migratetoward the equator during glacial periods (32).These effects, combined with the nonlinear re-sponse of glacial erosion to precipitation changes,would provide an appealing explanation for whythe impact of glaciation was more pronounced inmid-latitude regions with steep topography dur-ing the Quaternary (4).

REFERENCES AND NOTES

1. P. Molnar, P. England, Nature 346, 29–34 (1990).2. J. H. Tomkin, G. H. Roe, Earth Planet. Sci. Lett. 262, 385–397

(2007).3. J. Braun, D. Zwartz, J. H. Tomkin, Ann. Glaciol. 28, 282–290

(1999).4. F. Herman et al., Nature 504, 423–426 (2013).5. B. Hallet, J. Glaciol. 17, 209–222 (1979).6. B. Hallet, Ann. Glaciol. 22, 1–8 (1996).7. N. R. Iverson, Geology 40, 679–682 (2012).

8. J. Harbor, B. Hallet, C. F. Raymond, Nature 333, 347–349(1988).

9. K. R. MacGregor, R. S. Anderson, S. P. Anderson,E. D. Waddington, Geology 28, 1031–1034 (2000).

10. D. L. Egholm, S. B. Nielsen, V. K. Pedersen, J. E. Lesemann,Nature 460, 884–887 (2009).

11. K. R. MacGregor, R. S. Anderson, E. D. Waddington,Geomorphology 103, 189–204 (2009).

12. M. A. Kessler, R. S. Anderson, J. P. Briner, Nat. Geosci. 1,365–369 (2008).

13. F. Herman, F. Beaud, J.-D. Champagnac,J.-M. Lemieux, P. Sternai, Earth Planet. Sci. Lett. 310,498–508 (2011).

14. T. T. Creyts, G. K. C. Clarke, M. Church, J. Geophys. Res. 118,423–446 (2013).

15. F. Beaud, G. E. Flowers, S. Pimentel, Geomorphology 219,176–191 (2014).

16. R. B. Alley, D. E. Lawson, G. J. Larson, E. B. Evenson,G. S. Baker, Nature 424, 758–760 (2003).

17. B. Hallet, L. Hunter, J. Bogen, Global Planet. Change 12,213–235 (1996).

18. M. N. Koppes, D. R. Montgomery, Nat. Geosci. 2, 644–647(2009).

19. N. F. Humphrey, C. F. Raymond, J. Glaciol. 40, 539–552(1994).

20. C. A. Riihimaki, K. R. MacGregor, R. S. Anderson,S. P. Anderson, M. G. Loso, J. Geophys. Res. 110, F03003(2005).

21. D. A. Swift, P. W. Nienow, T. B. Hoey, Earth Surf. Process. Landf.30, 867–883 (2005).

22. J. Oerlemans, Arct. Alp. Res. 29, 233–239 (1997).23. Materials and methods are available as supplementary

materials on Science Online.24. B. Anderson, W. Lawson, I. Owens, B. Goodsell, J. Glaciol. 52,

597–607 (2006).25. S. C. Cox, D. J. A. Barrell, Geology of the Aoraki Area

(1:250,000 geological map 15, Institute of Geological andNuclear Sciences, GNS Science, Lower Hutt, New Zealand,2007), 71 pp. + 1 folded map.

26. R. G. Hilton, A. Galy, N. Hovius, Global Biogeochem. Cycles 22,GB1017 (2008).

27. O. Beyssac, B. Goffé, C. Chopin, J. N. Rouzaud, J. Metamorph.Geol. 20, 859–871 (2002).

28. R. L. Hooke, Geol. Soc. Am. Bull. 103, 1104–1108 (1991).29. B. Anderson et al., Arct. Antarct. Alp. Res. 46, 919–932

(2014).30. T. A. Little, S. Cox, J. K. Vry, G. Batt, Geol. Soc. Am. Bull. 117,

707 (2005).31. K. M. Cuffey, W. S. B. Paterson, The Physics

of Glaciers (Butterworth-Heinemann/Elsevier, Oxford,ed. 4, 2010).

32. G. H. Denton et al., Science 328, 1652–1656 (2010).

ACKNOWLEDGMENTS

The authors thank B. Anderson for field support. F.H., M.B.,and S.L thank the Fonds d'Investissement from the Faculty ofGeosciences and Environnement at the University of Lausanne.F.H. was funded by Swiss National Science Foundation grantPP00P2_138956. S.L, J.Y.Y.L., and J.-P.A. were funded by theGordon and Betty Moore Foundation through grant GBM 2808and NSF Earth Sciences grant 1348704. O.B. was funded by theCity of Paris (Emergence Program), Institut National desSciences de l’Univers, and Sorbonne Universités (Per-SU program).Three anonymous reviewers are thanked for their constructivereviews. K. Cuffey is thanked for his feedback. Data are availablein the supplementary materials.

SUPPLEMENTARY MATERIALS

www.sciencemag.org/content/350/6257/193/suppl/DC1Materials and MethodsFigs. S1 to S10Tables S1 and S2References (33–53)Movies S1 and S2

31 March 2015; accepted 3 September 201510.1126/science.aab2386

SCIENCE sciencemag.org 9 OCTOBER 2015 • VOL 350 ISSUE 6257 195

Fig. 3. Constraints on abra-sion law. (A) Erosion rateversus sliding velocity. Theblue dots represent measuredvelocities and integrated ero-sion rates using a 1-km binsize. Red and black lines cor-respond to the erosion ratepredictions with l = 2.02 andKg= 2.7 10−7 (m1-l/year1-l)and l = 1 and Kg = 10−4,respectively. The magnitudeof the error bars comes fromthe variability of erosion ratesthrough time. (B) Erosionexponent l versus the naturallogarithm of erosion constantKg. Each dot representssampling of the maximum-likelihood solution, with dotsbeing colored according to theirlikelihood (from blue to red,with red being most likely) (23).The black star is the estimatedvalue when integrating erosionand velocity over the entireglacier. The white star indicatesvalues obtained with the non-linear least-squares fit (23).Thequality of fit to data for eachdot in (A) is shown in fig. S2.(C and D) Probability densityfunctions for Kg and l, respec-tively (23). Black and red bars indicate 90 and 60% confidence intervals.

ln(K

g) Linear erosion law

Non-linear erosion law

0 50 100 150 200 250 300 3500

5

10

15

20

25

30

35

40

Ice velocity (m/yr)

Ero

sion

rat

e (m

m/y

r)

0.060.030Likelihood

0 1 2 3 4 5−35

−30

−25

−20

−15

−10

−5

Erosion parameter l

Non-linear least-squares solution

Solution with averaged velocity and erosion

1 2 3 4 50

0.02

0.04

0.06

0.08

0.1

0.12

Pro

babi

lity

Den

sity

Fun

ctio

nErosion parameter l

−30 −25 −20 −15 −100

0.02

0.04

0.06

0.08

0.1

0.12

Pro

babi

lity

Den

sity

Fun

ctio

n

ln(Kg)

RESEARCH | REPORTS

Page 4: Erosion by an Alpine glacier

DOI: 10.1126/science.aab2386, 193 (2015);350 Science

et al.Frédéric HermanErosion by an Alpine glacier

This copy is for your personal, non-commercial use only.

clicking here.colleagues, clients, or customers by , you can order high-quality copies for yourIf you wish to distribute this article to others

  here.following the guidelines

can be obtained byPermission to republish or repurpose articles or portions of articles

  ): October 26, 2015 www.sciencemag.org (this information is current as of

The following resources related to this article are available online at

http://www.sciencemag.org/content/350/6257/193.full.htmlversion of this article at:

including high-resolution figures, can be found in the onlineUpdated information and services,

http://www.sciencemag.org/content/suppl/2015/10/07/350.6257.193.DC1.html can be found at: Supporting Online Material

http://www.sciencemag.org/content/350/6257/193.full.html#ref-list-1, 9 of which can be accessed free:cites 45 articlesThis article

http://www.sciencemag.org/cgi/collection/geochem_physGeochemistry, Geophysics

subject collections:This article appears in the following

registered trademark of AAAS. is aScience2015 by the American Association for the Advancement of Science; all rights reserved. The title

CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005. (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience

on

Oct

ober

26,

201

5w

ww

.sci

ence

mag

.org

Dow

nloa

ded

from