rethinking low-temperature thermochronology data sampling strategies for quantification of...

Upload: devngri

Post on 07-Apr-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    1/14

    Rethinking low-temperature thermochronology data sampling strategies forquantification of denudation and relief histories: A case study in the Frenchwestern Alps

    Pierre G. Valla , Peter A. van der Beek, Jean Braun

    Institut des Sciences de la Terre, Universit Joseph Fourier, CNRS, BP 53, F-38041 Grenoble, France

    a b s t r a c ta r t i c l e i n f o

    Article history:Received 10 December 2010

    Received in revised form 29 April 2011

    Accepted 2 May 2011

    Available online 31 May 2011

    Editor: T.M. Harrison

    Keywords:

    low-temperature thermochronology

    numerical modeling

    inversion

    exhumation

    relief development

    sampling and modeling strategies

    We assess the importance of thermochronometric data sampling and modeling strategies for correctlyestimating mountain belt exhumation. Thermochronological age-elevation profiles have been widely used to

    infer orogenic exhumation histories; however, recent studies have shown that this sampling strategy may not

    be the most pertinent for quantifying both denudation and relief history. Here, we investigate the ability of

    combining different thermochronology data sampling schemes with numerical modeling to better constrain

    denudation rates and relief changes. We produce synthetic thermochronology datasets for real Alpine

    topography under a specific exhumation and relief scenario using the thermal-kinematic model Pecube. We

    then adopt an inverse approach based on the Neighborhood Algorithm to quantitatively assess the resolution of

    different thermochronology datasets collected following elevation profiles, long transects and valley bottom

    sampling. We also test the effect of the modeling approach on denudation and relief predictions, in particular

    the influence of the topographic grid resolution and of potential constraints on the geothermal gradient. Our

    results show that sampling along a single elevation profile does not allow to quantitatively constrain both

    denudation and relief histories. Numerical outputs clearly evidence tradeoffs that limit the capacity of

    simultaneously resolving denudation rates and relief change. Quantitative predictions are only slightly

    different when combining elevation profiles along different valleys, but are highly improved when using long

    transects or valley-bottom samples combined with an elevation profi

    le. The resolution with which reliefevolution can be predicted may be increased by a factor of 2 by using spatially distributed datasets. Results of

    thermal parameter inversions suggest that the geothermal gradient may be better estimated using elevation

    profiles or long-transect sampling rather than using valley bottom samples. Simulations with different model

    topography resolutionsshow that degrading the resolution for computational efficiency may result ina loss of

    quantitative information on denudation rates and relief history. In summary, we highlight that both

    thermochronological sampling strategies and the choice of thermal parameters or model topography resolution

    have a significant influence on predicted denudation and relief histories. Ideally,the samplingstrategy should be

    designed using preliminary modeling of expected denudation and relief histories, and a sensitivity study on

    assumed thermalparameters and model resolution should be performed whenmodeling the data. Although our

    modelingis basedon a particular case study of reliefevolution in theFrench westernAlps,we believe that these

    inferences have general relevance for thermochronological studies within mountain belts.

    2011 Elsevier B.V. All rights reserved.

    1. Introduction

    Low-temperature thermochronology has been widely used to

    quantify the exhumation history of mountain belts (e.g., Braun, 2005;

    Gallagher et al., 1998; Reiners and Brandon, 2006). Numerous studies in

    the last 20 yr or so have used apatite fission-track (AFT) (e.g., Fitzgerald

    et al., 1995; Gallagher et al., 1998; Hurford, 1991) and (U-Th)/He (AHe)

    (Ehlers and Farley, 2003; Farley, 2002) datato infer exhumation histories

    in specific areas. Moreover, low-temperature thermochronology may

    provide useful constraints for assessing paleo-topography and relief

    development (e.g., Clark et al., 2005; House et al., 1998; Reiners, 2007).

    The effect of steady-state topography on thermochronology data is well

    understood (e.g., Mancktelow and Grasemann, 1997; Stwe et al., 1994).

    However, potential effects of time-varying topography, leading to

    spatially and temporally varying exhumation rates as well as a transient

    thermal structure, are more difficult to assess (Braun, 2002a; Valla et al.,

    2010).

    Thermochronometric sampling strategies have often focused on

    obtaining age-elevation profiles, i.e., datasets of AFT and/or AHe ages

    collected along a steep elevation transect (e.g., Fitzgerald et al., 1995;

    Earth and Planetary Science Letters 307 (2011) 309322

    Corresponding author.

    E-mail address: [email protected] (P.G. Valla).

    0012-821X/$ see front matter 2011 Elsevier B.V. All rights reserved.

    doi:10.1016/j.epsl.2011.05.003

    Contents lists available at ScienceDirect

    Earth and Planetary Science Letters

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / e p s l

    http://dx.doi.org/10.1016/j.epsl.2011.05.003http://dx.doi.org/10.1016/j.epsl.2011.05.003http://dx.doi.org/10.1016/j.epsl.2011.05.003mailto:[email protected]://dx.doi.org/10.1016/j.epsl.2011.05.003http://www.sciencedirect.com/science/journal/0012821Xhttp://www.sciencedirect.com/science/journal/0012821Xhttp://dx.doi.org/10.1016/j.epsl.2011.05.003mailto:[email protected]://dx.doi.org/10.1016/j.epsl.2011.05.003
  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    2/14

    Gallagher et al., 2005; Wagnerand Reimer, 1972). Thermochronologic

    age-elevation relationships allow inferring apparent exhumation

    rates over a specific time window and, in theory, do not require

    knowledge of the geothermal gradient. One-dimensional consider-

    ation of age-elevation relationships thus remains the most direct way

    to interpret thermochronology data, either by simple linear regression

    of the data (e.g., Fitzgerald et al., 1995; Wagner and Reimer, 1972) or

    by using 1D thermal evolution models (e.g., Gallagher et al., 2005).

    However, topographic effects on thermochronology data are noteasily extracted from age-elevation relationships (Braun, 2002a) and

    recent modeling studies suggest that constraining mountain belt

    paleo-relief from elevation profiles is difficult, even when combining

    multiple thermochronometers (Valla et al., 2010; van der Beek et al.,

    2010; Vernon et al., 2009).

    Recent developments in numerical modeling offer tools for inter-

    preting low-temperature thermochronology data, either considering

    each sample individually (e.g., Ketcham, 2005) or exploiting the spatial

    correlation between samples (Braun et al., 2006; Gallagher et al., 2005;

    Stephenson et al., 2006a), to derive thermal histories and eventually

    constrain bothdenudation and relief histories. Braun (2002b) proposed

    an alternative approach to assess paleo-relief by sampling along a

    horizontal transect and using the spatial coherence between the

    topographic and thermochronological signals to determine both

    regional denudation rate and relief change (Beucher et al., submitted

    for publication; Herman et al., 2007, 2010a). However, the main

    limitation of this spectral method is that it assumes a spatially and

    temporally constant denudation history, and only provides an upper

    bound on the timing of denudation or the age of relief change.

    Numerical modeling offers the unique possibility to exploit dense

    spatial datasets either with or without any specific spatial relationship

    (Braun and van der Beek, 2004; Gallagher et al., 2005; Stephenson et al.,

    2006a). In this study, we aim to assess under what sampling and

    modeling conditionsquantitative constraintson the exhumationhistory

    can be extracted from low-temperature thermochronology data. We

    define different sampling strategies for a specific topography and test

    whether each sampling schemepermits deriving both denudation rates

    and relief history. Moreover, we assess to what extent the modeling

    approach, especially the spatial resolution of the model grid or anyimposed constraint on the geothermal gradient, can bias predicted

    denudation and relief histories. Our approach, although being specific

    with respect to the study area and the relief scenario we simulated, has

    been designed to more generally assess the potential influence of

    specific sampling schemes for constraining orogen exhumation. Our

    study thus follows an experimental design approach that has been

    widely usedin geophysics(e.g.,Curtis, 2004a,b; Wood andCurtis, 2004),

    although for computational simplicity we do not perform a complete

    inverse analysis but rather a priori define specific 3D sampling schemes

    that are traditionally used in thermochronological studies.

    Several studies (Glotzbach et al., 2011; Valla et al., 2010; van der

    Beek et al., 2010) have recently attempted to extract independent

    information on denudation and relief development in the western

    European Alps from thermochronology data. This was achieved bycombining the 3D thermal-kinematic model Pecube (Braun, 2003;

    Braun et al., submitted for publication), which predicts thermal

    histories and thermochronologicalages from an input denudation and

    relief history, with an inversion scheme based on the Neighborhood

    Algorithm (Sambridge, 1999a,b) to search the parameter space,

    extract best-fitting scenarios from the data and derive quantitative

    estimates for the parameters (rates, timing and relief ratio) describing

    denudation and relief histories.

    Weadopt a similarnumericalapproachas Valla et al.(2010); however,

    their study specifically focused on the problem of interpreting age-

    elevation profiles whereas here we consider various sampling strategies.

    Moreover, instead of building a synthetic topography for inverse

    modeling (Valla et al., 2010), we use real topography for predicting

    synthetic thermochronology data (Fig. 1a). We focus on the Ecrins

    Pelvoux massif (van der Beek et al., 2010), one of the External Crystalline

    Massifs (ECMs) within the western European Alps (e.g., Dumont et al.,

    2008). The Neogene exhumation history of the ECMs has been widely

    studied in the last few years, leading to contrasting views regarding

    timing and rates of exhumation, and the impact of Pliocene-Quaternary

    climate change on late-stage exhumation (Glotzbach et al., 2008, 2010;

    Reinecker et al., 2008; van der Beek et al., 2010; Vernon et al., 2009). Our

    synthetic approach is motivated by recent predictions forthe exhumation

    history of the Ecrins

    Pelvoux massif from thermochronology datacollected along either an elevation profile (van der Beek et al., 2010) or

    a long transect (Beucher et al., submitted for publication); both studies

    providedpreciseconstraints ondenudationhistoriesbut could notresolve

    relief evolution quantitatively. More generally, our study has been

    motivated by several attempts to quantify both timing and magnitude

    of topographicevolution in mountain belts using thermochronology (e.g.,

    Braun, 2005; Ehlers and Farley, 2003; Reiners, 2007; Valla et al., 2010).

    In the following, we first outline our modeling approach, present

    the synthetic data used for the simulations and discuss the inverse

    method. We then show inversion results and present to what extent

    predictions of denudation rates, relief change and near-surface

    thermal structure can be quantitatively determined by various

    thermochronological sampling strategies. The potential effects of

    varying the model gridresolution are reported in the Supplementary

    data. We finally discuss the implications of our findings for

    thermochronological sampling strategies to constrain regional

    denudation rates and/or topographic evolution of mountain belts.

    2. Modeling approach

    2.1. Thermal-kinematic modeling and synthetic thermochronology data

    The EcrinsPelvoux massif topography is characterized by a

    dominant NorthSouth topographic wavelength of ~10 km (Figs. 1a

    and 2a) organized around four major valleys (for details and locations

    see van der Beek et al., 2010). The relief amplitude is ~3 km, with

    summits around 4000 m and valley bottoms at ~1000 m (Fig. 1a). The

    EcrinsPelvoux massif has been chosen for our study as it forms a

    homogeneous area for the purpose of understanding Neogeneexhumation and relief development, i.e., no major active fault crosses

    the massif and consequently its exhumation history can be assumed

    to be spatiallyuniform (Delunel et al., 2010; van der Beek et al., 2010).

    Based on this topography and an assumed exhumation history, we

    predicted synthetic thermochronology data using the Pecube model

    (Braun, 2003). Pecube is a 3D thermal-kinematic model that calculates

    the thermal structure through time within an exhuming crustal block

    and provides thermal histories for rock particles that end up at the

    surface of the model, taking into account the effects of evolving

    topography (see Valla et al., 2010; Braun et al., submitted for

    publication for a detailed discussion). The Pecube finite-element

    code permits defining exhumation/relief scenarios that are described

    as discrete phases with an imposed regional denudation rate and

    paleo-relief. Relief evolution is quantified using the relief ratio R: forR =0, the paleo-relief is a plateau at maximum elevation; for R= 1,

    the paleo-relief is equal to present-day conditions; and for R=2, the

    paleo-relief is twice as large as the present-day relief. Note that relief

    increase operates through preferential valley incision in our model

    (Fig. 2a); i.e., ridges remain at a constant elevation but valley bottoms

    are lowered (see Valla et al., 2010; Braun et al., submitted for

    publication for a complete description), as would be expected for

    relief increase by glacial valley carving for instance. Isostatic rebound

    in response to this erosional unloading is included. Thus, local

    exhumation rates are the sum of: (1) the imposed regional

    (background) exhumation; (2) local relief change through valley

    incision; (3) spatially varying rock uplift induced by isostatic rebound.

    Thermal, mechanical and kinematic parameters used in the forward

    model are given in Table 1. The basal crustal temperature (Temp) and

    310 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    3/14

    the crustal heat production (H_prod; here normalized to heat capacity

    and density) are fixed to produce a surface geothermal gradient of

    ~25 C km1.

    Predicted thermal histories are then used to calculate different

    thermochronometric ages. In this study, we focus on late-stage

    exhumation histories and thus only compute low-temperature apatite

    fission-track (AFT) and (U-Th)/He (AHe) ages, as well as mean

    confined track lengths (MTL; see Valla et al., 2010). Thermal histories

    are translated into AHe ages using a simple model for He production

    diffusionejection (Farley, 2000); AFT ages are calculated using a

    model forAFT annealing(Greenet al., 1989; Stephenson et al., 2006b).

    We acknowledge that using these simple algorithms does not

    completely reflect the complex annealing and diffusion processes

    included in low-temperature thermochronology (Flowers et al., 2009;

    Gautheron et al., 2009; Ketcham, 2005). However, as discussed in

    Valla et al. (2010), we use the same forward models for AFT and AHe

    600' 612' 624'

    4448'

    4500'

    Longitude (E)

    La

    titude(N)

    AHe age (Ma)3 5 7 9 11 13 15

    c

    600' 612' 624'

    4448'

    4500'

    Longitude (E)

    La

    titude(N)

    AFT age (Ma)3 5 7 9 11 13 15

    b

    4506

    600

    612

    624

    4500

    4448500

    900

    1300

    1700

    2100

    2500

    2900

    3300

    3700

    4100

    Elevation(m)

    Latitude(N)

    Longitu

    de(

    E)

    a

    1_Profile

    3_Profiles

    Valleys

    Transect

    Profile + Valleys

    VeneonVeneon

    Bonn

    e

    Bonne

    MontontPelvouxelvoux

    Barre desarre desEcrinscrinsLa Meijea Meije

    Severais

    se

    Sever

    aisse

    Rom

    anche

    Romanc

    he

    Fig. 1. a) 3D map view of the modeling area (90-m resolution SRTM DEM of the Ecrins Pelvoux massif, Western Alps, France) used to produce AFT and AHe ages maps (b, c). Names

    of principal valleys and peaks are indicated for orientation. bc) Apatite fission-track (b) and (U-Th)/He (c) age map produced using Pecube (Braun, 2003; Table 1) and simulating a

    scenario of 50% relief increase (Rinput=0.5) during the last 3 Ma (Tinput). The constant background denudation rate is 0.3 km Myr1 (E1input= E2input) and resulting maximum rate of

    relief increase over the last 3 Ma is ~0.5 km Myr1 (see text for discussion). Sample locations for the 5 different sampling schemes are shown: (b) Sampling schemes 1_Profile

    (yellow circles, based on van der Beek et al., 2010), 3_Profiles (black circles) and Valleys (white circles). (c) Sampling schemes Transect(red circles, partly based on Beucher et al.,

    submitted for publication) and Profile+Valleys (yellow circles, partly based on van der Beek et al., 2010).

    311P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    4/14

    age predictions when generating the synthetic data and resolving the

    inverse problem; the results shown here are thus largely independent

    of the specific age-prediction models.

    To produce the synthetic datasets that we will use in the inversion,

    wedefine a 2-phase exhumation scenario with steady-statedenudation

    (0.3 km Myr1) over 15 Myr and a gradual 50% relief increase (R=0.5,

    Fig. 2a) since 3 Ma (T=3 Ma). This scenario simulates moderate relief

    increase (~1.5 km over the last 3 Myr, i.e., a maximum relief growth

    rate of ~0.5 km Myr1) associated with a relatively low and constant

    background denudation rate (0.3 km Myr1). Although this choicefor a

    specific denudation and relief scenario renders our analysis somewhat

    contextual, we focus on the type of scenario that has been inferred for

    the western Alpine massifs (e.g., Glotzbach et al., 2011; van der Beek et

    al., 2010; Vernon et al., 2009). Thermochronology data are predicted

    following this input scenario on a very high-resolution grid (90-m,

    Fig. 1b,c) to ensure that the numerical model properly resolves finite-

    amplitude topography effects on the crustal thermal structure. AFT and

    AHe age maps(Fig. 1b,c) show a significant age contrast between valley

    bottoms (AFT ~7 Ma and AHe ~34 Ma) and ridges (AFT ~15 Ma and

    AHe ~1213 Ma). Note that our synthetic AFT and AHe data are similarto measured thermochronological ages in the area (Beucher et al.,

    submittedfor publication; van der Beek et al., 2010), suggestingthat the

    imposed tectono-geomorphic scenario we define is reasonable.

    We sampled both the AFT (and associated MTL) and AHe maps

    following five different strategies: (1) a single elevation profile

    (Fig. 1b, 1_Profile), partly based on sample locations reported by van

    der Beek et al. (2010); (2) three elevation profiles sampling the three

    main valleys of the massif (Figs. 1b and 3b, 3_Profiles); (3) valley

    bottom samples along the major valleys (Figs. 1b and 3a, Valleys); (4)

    an elevation profile combined with valley bottom samples (Figs. 1c

    and 3c, Profile+Valleys); and finally (5) a long transect (partly based

    on sample locations in Beucher et al., submitted for publication)

    across the massif (Figs. 1c and 2, Transect). Each sampling scheme

    includes exactly 20 samples for which AFT, AHe and MTL data are

    1

    2

    3

    4

    Elevation(km)

    Distance (km)

    Age (Ma) MTL (m)

    North South

    a

    Elevation(km)

    Elevation(km)

    0 5 10 15 20 25 30 35

    4 6 80.5

    1.5

    2.5

    3.5

    10 11 12 13 140.5

    1.5

    2.5

    3.5

    cb

    AFT

    AHe

    -1

    -1

    10 12 14

    Fig. 2. a) Topographic cross-profile along the sampling scheme Transect(see Fig. 1b for location within the study area). Solid black and dashed red lines represent present-day and

    paleo-topography (R =0.5, see text for discussion), respectively. Black circles show sample locations for the sampling scheme Transect, and arrows represent relief evolution along

    the topographic profile (background denudation rate is 0.3 km Myr1). (b, c) Synthetic thermochronological ages versuselevation for Transectsamples, including AFT and AHe ages

    (b) and mean fission-track length data (MTL, c). Apparent exhumation rates (AER) deduced from linear regression of elevation vs age (b) are 0.34 and 0.33 km Myr 1 for AHe and

    AFT data, respectively (r=0.980.99). Note that the MTL pattern shows a normal correlation between MTL and elevation (c).

    Table 1

    Thermal, kinematic and elastic parameters used in Pecube. Crustal thickness, basal

    temperature (Temp) and heat production parameter (H_prod) are set to obtain a

    geothermal gradient of ~25 C km1 and a crustal heat production of ~7.5 CMyr1.

    Poissonratio,Young's modulusand equivalentelasticthickness areusedfor calculatingthe

    isostatic reboundin responseto relief change. Equivalent elasticthickness is set to a value

    that simulates moderate isostatic rebound (Valla et al., 2010; van der Beek et al., 2010).

    Parameter

    (unit)

    Inversions value

    Crustal thickness (km) 40

    Crustal density (kg m3) 2700

    Sublithospheric mantle density (kg m3) 3200

    Young's modulus (Pa) 1.1011

    Poisson ratio 0.25

    Equivalent elastic thickness (km) 25

    Thermal diffusivity (km2 Myr1) 25

    Basal crustal temperature (C) 600

    Crustal heat production parameter 0.8

    Sea-level temperature (C) 15

    Atmospheric lapse rate (C km1) 6

    312 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    5/14

    available. We thus compare the predictions of different sampling

    strategies while maintaining identical dataset size. Figs. 2 and 3 show

    synthetic AFT and AHe data plotted against elevation for the different

    sampling schemes. All AFT and AHe age-elevation relationships suggest

    nearly constant apparent exhumation rates (AER) of 0.330.39 and

    0.330.47 km Myr1 (see Figs. 2 and 3, correlation coefficients

    r0.950.99), respectively. The AER slightly overestimate the input

    background denudation rate due to additional local exhumation

    following relief growth (most strongly recorded by valley-bottomsamples, Fig. 3a).

    2.2. Inverse thermal-kinematic modeling

    For each sampling scheme, we run inversions to recover the four

    parameters used to describe the imposed tectono-geomorphic

    scenario, defining a range in which the inversion procedure will

    search for each: (1) denudation rate of the first phase (E1: 0

    2 km Myr1); (2) denudation rate of the second phase (E2: 0

    2 km Myr1); (3) transition time (T: 0.514.5 Ma); and (4) relief

    factor (R: 02). Each inverse simulation is run over 15 Myr. We run

    inversions constrained by the full dataset (AFT+AHe+MTL) as well

    as subsets of the input thermochronology data (AFT, AHe, AFT+AHe,

    AFT+MTL, and AFT+AHe+MTL data) to test what combination of

    thermochronometers is most useful to quantify relief and denudation

    histories in different sampling schemes.

    In a second set of inversions, we test the ability of the different

    sampling schemes to constrain not only relief and denudation

    histories, but also to retrieve the geothermal gradient (Temp and

    H_prod parameters, Table 3). To achieve this, we run five additional

    inverse simulations with the full dataset and 6 parameters to invert

    (E1, E2, T, and R as described previously; Temp: 500800 C, and

    H_prod: 03). We acknowledge that other thermo-physical parame-

    ters such as diffusivity or thermal conductivity, as well as the surface

    temperature history, are largely unknown and may also vary; butgiven the relatively homogenous lithology in our study area, we

    suspect they will do so within much narrower limits than the thermal

    parameters we test for and will therefore exert only a second-order

    influence on model predictions. Finally, we test the influence of the

    grid resolution on relief and denudation predictions (see Supplemen-

    tary data for results and discussion).

    To perform the inversions, we use the Neighborhood Algorithm

    (NA) of Sambridge (1999a,b). NA is a two-stage numerical procedure

    for non-linear inverse problems that allows efficient exploration of a

    multidimensional parameter space to find best-fitting models (NA

    sampling stage) as well as deriving estimates of posterior probability

    distribution for inverted parameters from the ensemble of models run

    during the sampling stage (NA appraisal stage).

    The sampling stage iteratively explores the multi-dimensional

    parameter space and progressively concentrates on regions where

    parameter values minimize the misfit to the data. NA makes use of

    1

    2

    3

    4

    1

    2

    3

    4

    4 6 8 14

    4 6 8 10 12 14 4 6 8 14

    1

    2

    3

    4

    Elevation(km)

    Elevation(km)

    Age (Ma) Age (Ma)

    Elevatio

    n(km)

    Age (Ma)

    a b

    c Profile + Valleys

    3_ProfilesValleys

    AHe

    AFTNorthCenterSouth

    ProfileValleys

    -1-1

    -1

    -1

    -1

    -1

    10 12

    10 12

    Fig. 3. Synthetic thermochronology data (AFT and AHe ages) versus elevation for different sampling schemes; see Fig. 1 for sample locations. a) Synthetic data for the sampling

    scheme Valleys. AER deduced from linear regression of elevation vs age are 0.47 and 0.39 km Myr 1 for AHe (white circles) and AFT (black circles) data, respectively (r=0.98). b)

    Synthetic data for the sampling scheme 3_Profiles (individual profiles are represented by diamonds, squares and triangles, see Fig. 1b for locations). AER are 0.33 and 0.35 km Myr1

    for AHe (white) and AFT (black) data, respectively (r=0.950.99). c) Synthetic data for the sampling scheme Profile+Valleys (profile and valleys samples are represented by

    diamonds and circles, respectively). AER are 0.37 and 0.36 km Myr1

    for AHe (white) and AFT (black) data, respectively ( r=0.98

    0.99).

    313P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    6/14

    simple geometrical concepts by dividing the entire parameter space

    into a set ofn-dimensional Voronoi cells, which represent the nearest

    neighborhood around each parameter combination. At each step, NA

    samples a forward model centered on each Voronoi cell andcalculates

    a misfit surface. Subsequent iterations make use of this misfit surface

    to sample in regions of the parameter space where the misfit function

    is minimized. The performance of the sampling stage is controlled by

    tuning parameters determining the number of iterations, the number

    of forward models generated at each iteration, and the number ofVoronoi cells to re-sample after each iteration. Varying these

    parameters will make the NA method either more explorative of the

    multi-dimensional parameter space, or more exploitative and faster in

    converging toward potential misfit minima (Sambridge, 1999a).

    Based on previous studies (Glotzbach et al., 2011; Valla et al., 2010),

    we choose to run ~5000 models distributed in 35 iterations for each

    inversion, with 50% of models re-sampled at each iteration.

    In our approach, we use a weighted least-squares misfit function

    for the NA appraisal stage:

    =

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    N

    i =1

    M

    j =1

    i

    j; modi

    j;dat

    2

    ij2

    vuuut 1

    where Nisthe numberof datasets(13;AFT,AHe,AFT+ AHe,AFT+ MTL,

    or AFT+AHe+MTL data), Mis the number of samples in each dataset

    (20 samples in total for each sampling scheme), j,modi and j,dat

    i are

    predicted and observedvalues for AFT/AHeagesor MTLrespectively, and

    ji is the uncertainty on the data. Here, we use constant synthetic

    uncertainties (ji) of 0.5 Ma and 0.5 m for AFT/AHe ages and MTL,

    respectively, to put an equal weight on all samples for misfit calculations

    (Valla et al., 2010). The above approach allows us to quantitatively infer

    the potential resolution ofperfect thermochronology data in resolving

    tectono-geomorphicscenarios(and potentially thermalparameters). We

    note, however, that adding random noise in the data could potentially

    lead to different model sensitivity in constraining numerical parameters.

    In the NA appraisal stage (Sambridge, 1999b), the model ensemblegenerated during the sampling stage is iteratively resampled to

    provide optimal estimates in the form of marginal probability-density

    functions (1D or 2D PDFs) of parameter values. For all parameters, we

    assumea uniform prior distribution withineach parameterrange. The

    posterior 1D (or 2D) PDFs are given based on the log-likelihood

    function (ln(L)), following the equation:

    ln L =0:5 2 =0:5 N

    i =1

    M

    j =1

    ij; modi

    j;dat

    2

    ij2

    0@

    1A 2

    3. Inversion results

    We first present inversion results for each sampling scheme to

    assess the effect of using various combinations of low-temperature

    thermochronometers. We then compare the different sampling

    schemes to discuss how best to predict denudation and relief

    histories. Finally, we show modeling results for inverting the

    geothermal gradient. We present inversion results as 1D marginal

    posterior PDF's as well as 2D joint PDF's of parameter values and

    discuss the reasons for the observed differences in the modeling

    outputs (resolution and errors on parameter estimates, and potential

    tradeoffs between model parameters) in the discussion section.

    Optimal parameter values (mode of the 1D posterior PDF) and 1

    uncertainties for the different sampling schemes and data combina-

    tions are given in Table 2.

    3.1. Sampling strategy

    The first sampling scheme (1_Profile, Table 2) considers 20 samples

    combined as a single elevation profile. Inversion results yield accurate

    predictions for E1 (E1=0.230.28 km Myr1; E1input=0.3 km Myr

    1) for

    all input thermochronological datasets with uncertainties ~2040%

    except when using AHe data alone. Estimates on E2 are less precise

    (uncertainties ~3080%) andslightlyoverestimated whenusingAFTdata

    alone or AFT+MTL data (E2 ~0.4 km Myr

    1

    ; E2input=0.3 km Myr

    1

    ).AHe data are required (alone or combined with AFT ages) to provide

    accurate predictions for E2 (respectively E2=0.330.11 and 0.28

    0.18 km Myr1). Adding MTL data, in contrast, does not improve

    inversion predictions (Table 2). Timing predictions are underestimated

    when using AFT alone, AFT+MTL or AFT+AHe+MTL data (T=1

    1.4 Ma) and clearly overestimated for AHealone (T=12.2 Ma). Only the

    AFT+ AHe inversion (T=2.8 Ma) gives predicted value of Tclose to the

    input value of 3 Ma. However, in all cases uncertainties on Tpredictions

    are very high (N100% for all inversions). Relief predictions are in

    agreement with theinput value (Rinput=0.5)when using AHedata alone

    or combined with AFT ages (R=0.530.58 and 0.63 0.55 respective-

    ly). In contrast, using AFT ages alone or combined with MTL

    measurements respectively under- (R=0.930.59) and over-predicts

    (R=0.18 +0.93/0.18) relief increase. Combining both AFT and AHe

    ages with MTL does not lead to an accurate estimate on R (R=1.08

    0.55; i.e., between ~50% increasing and 50% decreasing topographic

    relief).Moreover, as for Tpredictions, uncertainties on R are higher than

    100% for all inversions. In summary, inverse simulations using

    thermochronology data along a singleelevation profile provide accurate

    estimates on denudation histories (E1 and E2) with moderate to high

    uncertainties dependingon thedataset (~30 to 80%). In contrast, Tand R

    estimates are not in agreement with input values and systematically

    present uncertainties N100%. Best estimates are generally obtained

    when using AHe ages alone or combined with AFT data; adding MTL

    measurements does not improve denudation and relief predictions.

    Sampling along three different elevation profiles (3_Profiles,

    Table 2) provides similar predictions for E1, although the mean

    uncertainty is slightly reduced (~2040%, except for AFT+AHe data

    for which the uncertainty rises to 100%). E2 estimates are accurate forall inversions, but uncertainties remain quite high (~3080%). Similar

    to the previous inversions, T is systematically under-predicted

    (T=1.02.1 Ma) except for the AHe inversion (T=6.9 3.8 Ma);

    uncertainties remain N100% for all inversions. Relief predictions are

    much better than in the simulation using a single age-elevation

    profile. All inversions predict relief increase (R =0.180.58), but the

    inverse simulation combining all thermochronometric systems clearly

    overestimates the R input value (R=0.18+ 0.30/0.18). Uncertainties

    on R predictions are also reduced (~6590%, Table 2). As for a single

    elevation profile, best estimates on E2 and R (recent exhumation

    history) are provided using AHe data; however, for both Tand E1, most

    accurate predictions are obtained when using the full dataset (AFT+

    AHe+MTL data).

    Sampling along valley bottoms (Valleys, Table 2) provides accuratepredictions for E1 and E2. Uncertainties on E1 are low (~320%) only

    when using at least two thermochronometers, but are much higher

    when using AFT or AHe data alone (~100200%). In contrast,

    uncertainties on E2 are fairly constant, whatever the dataset used

    (~4060%). Similarly, constraints on Tare very poor using AFT or AHe

    data alone (T=11.9 +2.6/5.1 and 6.63.0 Ma, respectively) and

    are improved when combining thermochronometers (best estimate

    T=3.5 2.3 Ma for AFT+ AHe+MTL data). Uncertainties on T are

    high and variable depending on the dataset used: N100% for all

    inversions except for thefull dataset. Finally, relief increase is strongly

    overestimated when using AFTor AHe data alone (R =0.030.08) and

    predictions are not precise (uncertainties ~100%); however, combin-

    ing the thermochronometers provides more accurate values of R and

    reduced uncertainties (R=0.33

    0.48 and uncertainties ~ 50%). In

    314 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    7/14

    summary, this specific sampling scheme provides reasonably accurate

    and precise estimates of denudation history (E1 and E2) and accurate

    but not very well resolved predictions for T and R; however this

    requires combining at least two different thermochronometers (best

    estimates for the full AFT+ AHe+ MTL dataset).

    Combining valley bottom samples with an elevation profile(Profile+Valleys, Table 2) provides accurate and precise predictions

    for both E1 and E2 (E1=0.230.28 0.010.12 km Myr1 and

    E2=0.230.33 0.110.14 km Myr1). T estimates are similar for

    allinversions andmostly under-predict thetiming of change(T=0.7

    3.8 Ma); uncertainties on Tare around or above 100%. Similarly to the

    Valleys sampling scheme, relief change R is strongly overestimated

    when using AFT or AHe data alone (R =0.03 +0.55/0.03).

    Combining thermochronometers still over-predicts relief change

    (R =0.280.38) but uncertainties on R are reduced to ~50%.

    Combining valley bottom samples with an elevation profile thus

    provides reasonably precise and accurate estimates on E1, E2 and Tfor

    nearly all thermochronometer combinations, even though best

    estimates on inverted parameters, especially on relief change, are

    obtained when using two or more thermochronometers.

    The last inversion simulates sampling along a transect across the

    massif (Transect, Table 2). This sampling scheme leads to accurate

    estimates ofE1 withuncertainties around1040%(except forAFT+ AHe

    data, for which the uncertainty rises to ~100%); however, E2 is slightly

    over-predicted for AHe, AFT+ AHe and AFT+ MTL datasets (E2

    ~0.4 km Myr1

    , associated uncertainties ~60%). The timing parameteris also not well resolved (T=0.76.3 Ma) with prediction uncertainties

    of 100200%. Finally, relief estimates are quite accurate (R=0.180.58)

    but associated uncertainties are between 75 and 130%. In summary,

    sampling a long transect provides accurate predictions on denudation

    rates (E1 and E2), however, timing is not well constrained and relief

    estimates are accurate but associatedwith highuncertainties.Moreover,

    our inversions results show thatin this sampling configuration, using at

    least two thermochronometers strongly improves predictions on

    denudation rates and relief evolution.

    To better explore the potential differences between the sampling

    strategies, Fig. 4 presents 1D marginal PDFs of inverted parameters

    (E1, E2, Tand R) for each sampling strategy (using the full dataset). All

    sampling schemes provide accurate predictions for E1 (Fig. 4a), even

    though more precise estimates (uncertaintiesb

    10%) are obtained

    Table 2

    Output estimates after the NA appraisal stage for inputparameters E1, E2, Tand R (see text fordescriptionand discussion). Thesecond line reports parameterranges andinputvalues

    to calculate the synthetic thermochronology data (Fig. 1b,c). Optimal parameter values and associated uncertainties (mode 1of the 1D posterior PDF) are given in the following

    lines for various sampling schemes and different input thermochronology datasets. Best parameter estimates (and associated input thermochronology dataset) are also reported as

    percentage errors and uncertainties compared to input values (%Input).

    Parameter (unit) E 1 (km Myr1) E2 (km Myr

    1) T (Ma) R

    Range/input 02/0.3 02/0.3 0.514.5/3 02/0.5

    1_Profile

    AFT 0.28 0.13 0.38 0.26 1.4 + 5.3/0.9 0.93 0.59

    AHe 0.28 0.23 0.33 0.11 12.2 + 2.3/5.2 0.53 0.58AFT+AHe 0.28 0.06 0.28 0.18 2.8 + 3.9/2.3 0.63 0.55

    AFT+MTL 0.23 0.1 0.43 0.24 1.0 + 8/0.5 0.18 +0.93/0.18

    AFT+AHe+MTL 0.28 0.06 0.28 0.27 1.0 + 6.4/0.6 1.08 0.55

    Best estimate (%Input) 7% 20% +10% 37% 7% 130% +6% 116%

    Data AFT + AHe AHe AFT + AHe AHe

    3_Profiles

    AFT 0.28 0.12 0.28 0.21 2.1 + 5.5/1.6 0.38 0.41

    AHe 0.23 0.1 0.28 0.13 6.9 3.8 0.53 0.32

    AFT+AHe 0.28 0.3 0.28 0.14 1.0 + 3.9/0.5 0.58 0.29

    AFT+MTL 0.28 0.08 0.33 0.19 1.4 + 5.9/0.9 0.58 0.40

    AFT+AHe+MTL 0.28 0.05 0.28 0.15 1.7 + 3.8/1.2 0.18 +0.30/0.18

    Best estimate (%Input) 7% 17% 7% 43% 43%127% +6% 64%

    Data AFT + AHe + MTL AHe AFT + AHe + MTL AHe

    Valleys

    AFT 0.33 0.29 0.28 0.11 11.9 + 2.6/

    5.1 0.08 + 0.37/

    0.08AHe 0.33+0.56/0.33 0.23 0.12 6.6 3.0 0.03 + 0.50/0.03

    AFT+AHe 0.28 0.05 0.23 0.11 0.7 + 4.8/0.2 0.48 0.25

    AFT+MTL 0.28 0.06 0.33 0.10 2.1 + 5.3/1.6 0.33 0.21

    AFT+AHe+MTL 0.28 0.01 0.28 0.11 3.5 2.3 0.38 0.22

    Best Estimate (%Input) 7%3% 7%37% 17%77% 4%50%

    Data AFT + AHe + MTL AFT + AHe+ MTL AFT + AHe + MTL AFT + AHe

    Profile+Valleys

    AFT 0.23 0.12 0.28 0.13 3.8 + 5.0/3.3 0.03 +0.55/0.03

    AHe 0.23 0.11 0.28 0.14 2.8 + 3.7/2.3 0.03 +0.56/0.03

    AFT+AHe 0.28 0.04 0.28 0.14 0.7 + 3.6/0.2 0.38 0.24

    AFT+MTL 0.28 0.08 0.33 0.12 1.0 + 6.5/0.5 0.28 0.23

    AFT+AHe+MTL 0.28 0.01 0.23 0.11 2.1 + 2.6/1.6 0.28 0.21

    Best Estimate (%Input) 7%3% 7% 47% 30%87% 24% 48%

    Data AFT + AHe + MTL AFT + AHe AFT + AHe + MTL AFT + AHe

    Transect

    AFT 0.23 0.12 0.28 0.17 2.1 + 7.0/1.6 0.18 + 0.44/0.18AHe 0.28 0.15 0.43 0.18 1.7 + 6.4/1.2 0.48 + 0.65/0.48

    AFT+AHe 0.28 0.32 0.43 0.22 1.0 + 3.4/0.5 0.58 0.39

    AFT+MTL 0.28 0.07 0.43 0.21 6.3 3.8 0.38 0.62

    AFT+AHe+MTL 0.28 0.03 0.28 0.2 0.7 + 4.71/0.2 0.48 0.37

    Best estimate (%Input) 7% 10% 7% 57% 67%113% 4%74%

    Data AFT + AHe + MTL AFT AFT + AHe AFT + AHe + MTL

    315P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    8/14

    when sampling a transect, valley bottoms or an elevation profile

    combined with valley bottoms; uncertainties from sampling along

    elevationprofiles aretwiceas high. TheaccuracyofE2 estimates is also

    similar for all sampling schemes, with only the resolution differing

    (Fig. 4b); lowest uncertainties on E2 predictions are obtained when

    sampling along valley bottoms (~40% for Valleys and Profile+Valleys),

    the 3_Profiles sampling scheme yields an uncertainty of ~50% while

    the Transect and 1_Profile sampling strategies yield uncertainties of

    ~70

    90%. Fig. 4c confi

    rms that timing of change (T) is the leastresolved parameter, with uncertainties of 8090% for the Valleys and

    Profile+Valleys sampling schemes, and uncertainties N100% for the

    other schemes. Finally, the predicted R (Fig. 4d) is reasonably accurate

    (slightly overestimated) for the 3_Profiles, Valleys and Profile+Valleys

    sampling schemes and is associated with uncertainties of ~5060%.

    Transect sampling also leads to an accurate estimate for R but the

    associated uncertainty is high (~ 75%) whereas 1_Profile does not

    provide any constraint on R (i.e., it predicts either 50% increasing or

    decreasing topographic relief).

    3.2. Modeling approach

    In all previous inverse simulations, our modeling approach assumed

    the geothermal gradient was known. We now present results inverting

    for both denudation/relief histories (E1, E2, T and R) and thermal

    parameters (Temp and H_prod). We run inversions for allfive sampling

    schemes and use the full (AFT+AHe+MTL) dataset to compare

    output predictions with previous outcomes (Table 3). Basal temper-

    ature (Temp) is slightly under-predicted for most sampling schemes

    (Temp ~520565 C; Tempinput=600 C), except for 3_Profiles sampling

    that leads to an overestimate (Temp=714 80 C). Associated un-

    certainties are low and similar for all inversions (~1116%, Table 3).

    Constraints on the crustal heat production parameter (H_prod,

    Table 3) are less tight. The Profile+Valleys, 1_Profile and 3_Profiles

    sampling strategies lead to accurate predictions (H_prod =0.56

    0.94; H_prodinput=0.8) but estimated uncertainties are relativelyhigh (~80110%). Valleys sampling clearly does not resolve the heat-

    production parameter (H_prod =0.04 +0.65/0.04) whereas best

    predictions for H_prod are provided by the Transect inversion

    (H_prod =1.16 0.44).

    Predictions for E1 are slightly underestimated for the Transect,

    3_Profiles and 1_Profile sampling schemes (E1=0.23 0.1 km Myr1)

    and all results provide higher uncertainties on E1 (~30%) compared to

    inversions that imposed the geothermal gradient (~320%). Tightest

    constraints are placed on E1 using the Valleys and Profile+Valleys

    sampling schemes. Results for E2 are variable with best predictions for

    the3_Profiles inversion (E2=0.280.14 km Myr1). E2 is slightly over-

    predictedin the Transectsamplingscheme(E2=0.430.21 km Myr1)

    and under-predicted for Valleys and Profile+Valleys sampling schemes

    (E2=0.230.19 km Myr1). However, uncertainties on E2 are very

    similar to those obtained when imposing thegeothermal gradient (~50

    100%). Timing predictions are all underestimated (T ~0.71.7 Ma) and

    associated uncertainties are similar to those found during imposed

    0 0.4 0.8 1.2 1.6 2

    0.05

    0.15

    0.25

    0.35

    0

    0.04

    0.08

    0.12

    0.16

    Probabi

    lity

    Probability

    Probability

    Probabi

    lity

    a

    c

    b

    d

    E1(km Myr-1) E

    2(km Myr-1)

    T(Ma) R

    1_Profile3_ProfilesValleysProfile + valleysTransect

    Sampling scheme:

    2 6 10 14

    0.01

    0.03

    0.05

    0.07

    0.09

    0 2

    0 20

    0.02

    0.04

    0.06

    0.08

    0.1

    0.4 0.8 1.2 1.6

    0.4 0.8 1.2 1.6

    Fig. 4. 1D posterior parameter PDF's obtained after the NA appraisal stage for the five different sampling schemes (AFT+AHe+MTL dataset): (a) denudation rate of the first

    exhumation phase (E1); (b) denudation rate of the second exhumation phase (E2); (c) transition time between the two exhumation phases (T); and (d) relief factor between the

    paleo-relief and the final relief (R). Each line defines an inversion experiment with a sampling scheme (see figure for color legend and Fig. 1 for sampling locations). Black stars

    represent the input parameter values.

    316 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    9/14

    geothermal gradient inversions (~70230%). Moreover, best predictions

    for Tare provided by the Valleys inversion (T=1.7 + 2.0/1.2 Ma), in

    agreement with previous results. Finally, setting no constraint on the

    geothermal gradient does not have any effect on relief evolution

    predictions as R estimates are very close to values obtained with an

    imposed geothermal gradient, and optimal predictions are obtained for

    3_Profiles, Valleys and Profile+Valleys sampling schemes.

    To summarize, numerical inversions without imposing constraints

    on the geothermal gradient provide less accurate and precise

    estimates for exhumation rates (E1 and E2), but output results for

    timingand reliefchange (Tand R) areverycloseto thosefoundwith an

    imposed geothermal gradient. Best estimates for thermal parameters

    (H_prod and Temp) are given by Transectsampling, while inversions

    including at least one elevation profile (Profile+Valleys, 1_Profile and3_Profiles) provide accurate estimates but associated with relatively

    high uncertainties, and valley bottom sampling (Valleys) does not

    resolve thermal parameters.

    4. Discussion

    4.1. Limits to the numerical modeling approach

    Before interpreting our results, we should highlight potential

    shortcomings associated with our numerical modeling approach.

    First of all, the exhumation scenario we choose is fairly simple,

    with only two exhumation phases, a constant and spatially uniform

    regional denudation rate and a single episode of relief growth. This

    choice could be considered specific and restrictive, although weconsciously chose uniform background exhumation within the study

    area to focus on relief evolution. A similar inverse approach could,

    however, be developed to specifically explore spatially varying

    exhumation (e.g., Stephenson et al., 2006a). We have tested one

    specific geothermal gradient that we also assume to be spatially

    uniform, as we apply our inversions to a region of spatially relatively

    homogeneous lithology. However, we note that localfluid circulations

    along faults can induce large spatial variations in thermal gradient

    (Whipp and Ehlers, 2007).

    Despite several recent attempts (e.g., Ehlers et al., 2006; Glotzbach

    et al., 2008, 2010; van der Beek et al., 2010; Vernon et al., 2009 ),

    quantifying both timing and amplitude of topographic relief changes

    in mountain belts using low-temperature thermochronology has

    proven elusive, motivating the focus of our study on this specific

    exhumation scenario. Although our study design might be somewhat

    restrictive, we have consciously chosen quite a general synthetic

    study regarding the exhumation history and the near-surface thermal

    structure. The reliefevolution scenario should also be quite general, as

    moderate relief increase since the mid-Pliocene is expected or has

    been encountered in many mountain belts (e.g., Molnar and England,

    1990; Zhang et al., 2001). We thus are confident that our inferences

    regarding thermochronology sampling strategies for quantifying both

    exhumation and relief histories, despite the above limitations, might

    be of more general interest for studying at least the late-Neogene

    exhumation history of mountain belts.

    Second, the way we define relief changes in Pecube does not

    capture the complexities of landscape evolution processes. Some

    limitations include the constant planform drainage pattern assumedin our numerical approach, and a geometric approach to relief

    development (Fig. 2a). An alternative approach would be to combine

    Pecube with a landscape evolution model to fully simulate landscape

    changes with surface process modeling (Herman et al., 2010b).

    However, our synthetic approach allows us to address at first order

    the effect of transient topography on thermochronology data and

    their intrinsic resolution in quantitatively retrieving potential relief

    changes.

    We finally note that the different sampling scenarios explored

    here, although they represent the most commonly used and simplest

    ways to sample a given topography, might not be the most adequate

    to independently quantify exhumation and relief histories, i.e., we do

    not perform a complete experimental design study.

    4.2. Parameter resolution and tradeoffs

    Our numerical results show that some parameters(mainly Tand R,

    Fig. 4) have been poorly constrained by the NA inversion. The lack of

    convergence toward the input parameters may result from the poor

    resolution of thermochronology data to constrain timing or relief

    parameters (Valla et al., 2010), from some random elements, or from

    potential tradeoffs between model parameters. To illustrate this

    alternative explanation, Fig. 5 shows 2D scatter plots and PDFs for a

    single elevation profile (1_Profile). There is no apparent tradeoff

    between the initial and final denudation rates (E1 and E2, Fig. 5a), but

    significant tradeoffs appear between the recent denudation rate ( E2)

    and both the relief change (R, Fig. 5b) and the timing of change (T,

    Table 3

    Output estimates after the NA appraisal stage for input parameters E1, E2, T, R, Temp and H_prod. The second line reports parameter ranges and input values for the synthetic

    thermochronology data (Fig. 1b,c). Optimal parameter values and associated uncertainties (mode 1) are given in the following lines for various sampling schemes (AFT+AHe+

    MTL dataset). Inversion results are also reported as percentage errors and uncertainties compared to input values (%Input).

    Pa ra meter (u ni t) E 1 (k m Myr1) E2 (km Myr 1) T (Ma) R Temp (C) H_prod

    Range/input 02/0.3 02/0.3 0.514.5/3 02/0.5 500800/600 03/0.8

    1_Profile

    Mode1 0.23 0.1 0.33 0.29 0.7 + 3.1/0.2 1.08 0.52 549 + 98/49 0.64 + 0.82/0.64

    %Input (23% 33%) (+10% 97%) (77% 103%) (+ 116% 104%) (9% 16%) (20%103%)

    3_Profiles

    Mode 1 0.23 0.1 0.28 0.14 0.7 + 6.9/0.3 0.48 0.27 714 80 0.94 0.66

    %Input (23% 33%) (7% 47%) (77% 230%) (4% 54%) (+19% 13%) (+18% 83%)

    Valleys

    Mode 1 0.33 0.1 0.23 0.19 1.7 + 2.0/1.2 0.33 0.21 519 + 74/19 0.04 + 0.65/0.04

    %Input (+10% 33%) (23% 63%) (43% 67%) (34% 42%) (14% 12%) (95%81%)

    Profile+Valleys

    Mode 1 0.28 0.1 0.23 0.19 0.7 + 5.1/0.3 0.48 0.26 534 + 95/34 0.56 + 0.85/0.56

    %Input (7% 33%) (23% 63%) (77% 170%) (4% 52%) (11% 16%) (30%106%)

    Transect

    Mode 1 0.23 0.09 0.43 0.21 1.4 + 4.4/0.9 0.48 0.4 564 66 1.16 0.44

    %Input (23% 30%) (+43% 70%) (53% 147%) (4% 80%) (6% 11%) (+ 45% 55%)

    317P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    10/14

    Fig. 5c). These tradeoffs may explain the relatively poor resolution

    observed in the 1D marginal PDFs for these parameters ( Fig. 4c,d).

    The potential parameters tradeoffs for the other sampling schemes

    are shown in Figs. 6 and 7. Fig. 6 reveals that the tradeoff between E2and R is reduced for the 3_Profiles, Valleys, and Profile+Valleys

    schemes (Fig. 6ac), but confirms their tendency to over-estimate the

    relief increase. In contrast, there is an important tradeoff in the

    Transect sampling scheme (Fig. 6d), which explains the high uncer-

    tainties associated with R. Fig. 7 confi

    rms that the timing parameterT is poorly constrained by our inverse approach; 2D scatter plots

    reveal a significant tradeoff between E2 and T, even though the

    covariance appears to be slightly reduced for sampling schemes

    Valleys and Profile+Valleys (Fig. 7b,c). Interestingly, both Figs. 6 and 7

    suggest that parameter tradeoffs, similar to 1D marginal PDFs, are

    highly dependent on the sampling scheme and thus might also be

    considered when designing sampling strategies.

    4.3. Sampling strategy

    Inversionresultsfor a singleelevation profile (1_Profile) show that,

    although such a sampling approach allows accurately estimating

    denudation rates, it does not constrain relief evolution satisfactorily

    (uncertainties N100%, Table 2). This finding is similar to the

    conclusion of Valla et al. (2010), who also performed a sensitivity

    analysis and showed that relief evolution can only be precisely

    predicted from thermochronological age-elevation profi

    les if the rateof relief change is at least two to three times higher than the regional

    background denudation rate. In our case, the regional denudation rate

    is 0.3 km Myr1 and the relief scenario induces a maximum relief-

    growth rate of ~0.5 km Myr1 for the last 3 Myr (i.e., ~60% higher

    than the background denudation rate); under this configuration, our

    inverse results are in good agreement with Valla et al. (2010). We

    suggest that the inability of a single age-elevation profile to constrain

    0 20 40 60 80 100

    0

    0.5

    1

    1.5

    2

    0 0.5 1 1.5 2

    0

    0.5

    1

    1.5

    2

    0 0.5 1 1.5 2

    R

    0 0.4 0.8 1.2 1.6 2

    10

    6

    2

    14

    a b

    c

    E1(km Myr-1) E

    2(km Myr-1)

    E2(km Myr-1)

    E2

    (kmM

    yr-

    1)

    T(Ma)

    Fig. 5. Scatter plots (each dot represents a forward model color coded by its misfit value ) and 2D posterior marginal probability density functions with 1confidence contours

    (black lines)for the sampling scheme 1_Profile: a)denudation rates of thefirst(E1) vsthesecond(E2) exhumation phases;b) denudation rate of thesecond exhumation phase (E2) vs

    relief factor (R); and c) denudation rate of the second exhumation phase ( E2) vs transition time between the two exhumation phases ( T). Black stars represent the input parameter

    values.

    318 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    11/14

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    12/14

    E2) and relief change (R); however, the uncertainties on these

    parameters are higher than those obtained with Valleys and Profile+

    Valleys sampling strategies, due to higher parameters tradeoffs (Figs. 6

    and 7). These outcomes are consistent with recent analysis of AFT data

    from a transect across the Pelvoux massif (Beucher et al., submitted for

    publication): these data constrain exhumation rates well but provide

    little insight into relief change. These differences appear to result from

    the spatial configuration of the Transectsampling scheme. Although it

    covers a larger area than the 1_Profile and 3_Profiles sampling schemes,

    Transectdata only sampleone spatial direction(Fig. 2a) andmay notfully

    record the 2D topographic perturbations of isotherms compared to data

    that are spatially distributed along different valleys (Valleys or Profile+

    Valleys sampling scheme). Combining a long transect with at least one

    elevationprofile would have allowed to best capture both short and long

    topographic wavelength effects on the thermal structure (Braun, 2002a,

    b), simultaneously providing accurate and precise information on both

    denudation and relief predictions.

    4.4. Modeling approach

    Inversionsfor which the thermal parameters are supposed unknown

    lead to comparable accuracy but somewhat higher uncertainties for

    2

    6

    10

    14

    0 0.4 0.8 1.2 1.6 20 0.4 0.8 1.2 1.6 2

    00.4 0.8 1.2 1.6 2

    00.4 0.8 1.2 1.6 2

    2

    6

    10

    14

    2

    6

    10

    14

    2

    6

    10

    14

    a b

    c d

    3_Profiles Valleys

    Profile+Valleys Transect

    E2(km Myr-1)E

    2(km Myr-1)

    E2(km Myr-1) E

    2(km Myr-1)

    T(Ma)

    T

    (Ma)

    T

    (Ma)

    T(Ma)

    Fig. 7. Scatter plots (see Fig. 5 for legend) and 2D posterior marginal probability density functions with 1 confidence contours (black lines) of denudation rate of the secondexhumation phase (E2) vs transition time between the two exhumation phases (T) for different sampling schemes: a) 3_Profiles; b) Valleys; c) Profile+Valleys; and d) Transect. Black

    stars represent the input parameter values.

    320 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    13/14

    denudation and relief predictions than when imposing a geothermal

    gradient (Table 3). These differences in predicteduncertainties highlight

    the classic problem of determining an appropriate dimensionality for a

    numericalproblem,i.e.,the number of inverted parameters compared to

    the intrinsic data resolution (see Gallagher et al., 2005, for a discussion

    on this subject). Moreover, our results show that sampling along valley

    bottoms does not allow resolving accurately the thermal parameters.

    Sampling along elevation profiles helps constraining the geothermal

    gradient, as suggested by Gallagher et al. (2005), pointing out that eventhough numerical modeling allows us to infer exhumation histories

    from complex spatial sample distributions, age-elevation relationships

    remain useful for interpreting thermochronology data. The ability of

    age-elevation profiles to constrain both the exhumation history and

    thermal parameters comes from the altitudinal distribution of samples;

    the slope of the age-elevation relationship providing an estimate of the

    apparent exhumation rate (e.g., Hurford, 1991), while the age-window

    of the dataset may help to constrain the near-surface geothermal

    gradient (Gallagher et al., 2005). Surprisingly, the Transect sampling

    provided the best estimates on thermal parameters, which can be

    explained by the simultaneous spatial and altitudinal data coverage

    allowed by this sampling scheme(Fig.2a). It thusmightbe usednot only

    for spectral analysis (Braun, 2002b) but also, like elevation profiles, to

    constrain the geothermal gradient.

    Finally, we have tested the effect of the grid resolution on the

    predicted estimates of the various parameters. As reported in the

    supplementary material section, we show that a minimum grid

    spacing of 1 km (or ~1/10th of the valley spacing) is necessary to

    properly resolve exhumation rates and relief change in the typical

    mountainous terrain we have used here.

    5. Conclusions

    Our study was initially designed to determine to which extent

    different sampling and modeling strategies could allow constraining

    relief evolution from low-temperature thermochronology datasets.

    Our results confirm the findings ofValla et al. (2010), who suggested

    that age-elevation relationships do not resolve relief changes if these

    do not induce an additional spatially variable exhumation rate that isat least two times higher than the regional background denudation

    rate. Although our approach is slightly different from Valla et al.

    (2010), we have shown here that age-elevation relationships do not

    allow constraining relief changes. We have demonstrated that

    spatially distributed sampling along a transect, several elevation

    profiles or along valley bottoms, preferably combined with an

    elevation profile, leads to better predictions of relief evolution. The

    main drawback of only sampling along valley bottoms is the loss of

    information on the near-surface thermal structure. However, using

    higher-temperature thermochronometers may provide supplemen-

    tary constraints on thermal parameters. Glotzbach et al. (2011)

    recently used a dense thermochronology dataset combining both low-

    and high-temperature thermochronometers along elevation profiles

    and a tunnel transect to constrain both the near-surface thermalstructure and the exhumation history of the Mont-Blanc massif

    (western European Alps). Their modeling results resolved a recent

    relief increase whereas other recent studies using only data along a

    single age-elevation profile were not able to resolve the recent relief

    evolution (van der Beek et al., 2010; Vernon et al., 2009).

    Based on synthetic data producedunder an imposed exhumation and

    relief scenario, we have tested different sampling strategies to explore

    how to best constrain denudation rates, timing and relief change. Our

    results show that dense spatial sampling combining valley bottom

    samples and elevation profiles, as well as long transects may help to

    provide quantitative constraints on complex exhumation histories.

    Moreover, we have highlighted that the modeling approach (choice of

    thermalparametersor model resolution)potentiallyinfluencespredicted

    denudation and relief histories. Although age-elevationprofiles may bea

    useful and direct way to interpret thermochronology data, sampling

    strategies should be rethought in the light of recent numerical modeling

    developments. For a region where relief changeis expected to have been

    minimal but the emphasis is on deciphering changes in denudation rate

    through time, an appropriate sampling scheme would be using one or

    several elevation profiles. However, when one has reason to believe that

    exhumation rates may have been relatively constant (e.g., in stable

    post-orogenic mountain belts) but the aim is to constrain the timing and

    amount of relief change, our study would suggest a better choice is tosample valley bottoms, possibly combined with an elevation profile.

    Therefore, ideallythesamplingstrategy should be designeda priori,using

    preliminary modeling of expected denudation and relief histories and

    preferably using multiple working hypotheses; and a sensitivitystudy on

    assumed thermal parameters and model resolution should be performed

    when modeling the data.

    Acknowledgments

    This study is supported by the INSU-CNRS through the European

    Science Foundation Topo-Europe program 07-TOPO-EUROPE-FP-023Coupled climatic/tectonic forcing of European topography revealed

    through thermochronometry (Thermo-Europe) and the Agence

    Nationale de la Recherche project No. ANR-08-BLAN-0303-01 Erosionand Relief Development in the Western Alps. It forms part of PV's PhD

    project at Universit Joseph Fourier, supported by the French Ministry

    for Research and Higher Education. We thank Malcolm Sambridge and

    two anonymous reviewers for insightful comments that helped to

    improve the clarity of this manuscript. Computations were performed

    on Brutus, the high performance computing facilitiesat ETH Zurich. The

    codes are available at http://svn-geo.ethz.ch after registering at this

    site.

    Appendix A. Supplementary data

    Supplementary data to this article can be found online at

    doi:10.1016/j.epsl.2011.05.003.

    References

    Beucher, R., van der Beek, P.A., Braun, J. and Batt, G. (submitted for publication).Assessing exhumation and relief development in the Pelvoux and Dora-Maira massifs(western Alps) from spectral analysis and inversion of thermochronological agetransects. Journal of Geophysical Research.

    Braun, J., 2002a. Quantifying the effect of recent relief changes on age-elevationrelationships. Earth Planet. Sci. Lett. 200, 331343.

    Braun, J., 2002b. Estimating exhumation rate and relief evolution by spectral analysis ofage-elevation profiles. Terra Nova 14, 210214.

    Braun, J., 2003. Pecube: a new finite element code to solve the heat transport equationin three dimensions in the Earth's crust including the effects of a time-varying,finite amplitude surface topography. Comput. Geosci. 29, 787794.

    Braun, J., 2005. Quantitative constraints on the rate of landform evolution derived fromlow-temperature thermochronology. Rev. Mineral. Geochem. 58, 351374.

    Braun, J., van der Beek, P.A., 2004. Evolution of passive margin escarpments: what canwe learn from low-temperature thermochronology? J. Geophys. Res. 109, F04009.doi:10.1029/02004JF000147.

    Braun, J., van der Beek, P.A., Batt, G., 2006. Quantitative Thermochronology: NumericalMethods for the Interpretation of Thermochronological Data.Cambridge UniversityPress, Cambridge, New York. 258 pp.

    Braun,J., vander Beek, P.A., Valla,P.G.,Robert, X.,Herman, F.,Glotzbach, C.,Pedersen,V.,Perry, C., Simon-Labric, T. and Prigent, C. (submitted for publication). Quantifyingrates of landscape evolution and tectonic processes by thermochronology andnumerical modeling of heat transport in the crust using PECUBE. Tectonophysics.

    Clark, M.K., Maheo, G., Saleeby, J., Farley, K.A., 2005. The non-equilibrium landscape ofthe southern Sierra Nevada, California. GSA Today 15, 410.

    Curtis, A., 2004a. Theory of model-based geophysical survey and experimental designpart Alinear problems. Lead. Edge 23, 9971004.

    Curtis, A., 2004b. Theory of model-based geophysical survey and experimental designpart Bnonlinear problems. Lead. Edge 23, 11121117.

    Delunel, R.,van derBeek,P.A.,Carcaillet,J., Bourls, D.L., Valla,P.G.,2010.Frost-crackingcontrol on catchment denudation rates: Insights from in situ produced 10Beconcentrations in stream sediments (EcrinsPelvoux massif, FrenchWestern Alps).

    Earth Planet. Sci. Lett. 293, 72

    83.

    321P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

    http://svn-geo.ethz.ch/http://dx.doi.org/10.1029/02004JF000147http://dx.doi.org/10.1029/02004JF000147http://svn-geo.ethz.ch/
  • 8/6/2019 Rethinking Low-temperature Thermochronology Data Sampling Strategies for Quantification of Denudation and Reli

    14/14

    Dumont, T., Champagnac, J.-D., Crouzet, C., Rochat, P., 2008. Multistage Alpine shorteningin Central Dauphin (French Western Alps): implications for pre-Alpine restoration.Swiss J. Geosci. 101, 89110.

    Ehlers, T.A., Farley, K.A., 2003. Apatite (U-Th)/He thermochronometry: methods andapplications to problems in tectonic and surface processes. Earth Planet. Sci. Lett.206 (12), 114.

    Ehlers, T.A., Farley, K.A., Rusmore, M.E., Woodsworth, G.J., 2006. Apatite (U-Th)/Hesignal of large-magnitude accelerated glacial erosion, southwest British Columbia.Geology 34, 765768.

    Farley, K.A., 2000. Helium diffusion from apatite: general behavior as illustrated byDurango fluorapatite. J. Geophys. Res. 105, 29032914.

    Farley,K.A., 2002.(U-Th)/He dating:techniques, calibrations,and applications.In: Porcelli,D., Ballentine, C.J., Wieler, R. (Eds.), Reviews in Mineralogy and Geochemistry: NobleGases in Geochemistry and Cosmochemistry, 47. Mineralogical Society of America,Washington, DC, pp. 819844. 2002.

    Fitzgerald, P.G., Sorkhabi, R.B., Redfield, T.F., Stump, E., 1995. Uplift and denudation of thecentralAlaska Range:a casestudyin theuse of apatitefission trackthermochronologyto determine absolute uplift parameters. J. Geophys. Res. 100, 2017520191.

    Flowers, R.M., Ketcham, R.A., Shuster, D.L., Farley, K.A., 2009. Apatite (U-Th)/Hethermochronometry using a radiation damage accumulation and annealing model.Geochim. Cosmochim. Acta 73, 23472365.

    Gallagher, K., Brown, R., Johnson, C., 1998. Fission track analysis and its applications togeological problems. Annu. Rev. Earth Planet. Sci. 26, 519572.

    Gallagher, K., Stephenson, J., Brown, R., Holmes, C., Fitzgerald, P., 2005. Lowtemperature thermochronology and modeling strategies for multiple samples 1:vertical profiles. Earth Planet. Sci. Lett. 237 (12), 193208.

    Gautheron, C., Tassan-Got, L., Barbarand, J., Pagel, M., 2009. Effect of alpha-damageannealing on apatite (U-Th)/He thermochronology. Chem. Geol. 266, 157170.

    Glotzbach, C., Reinecker, J., Danisik, M., Rahn, M., Frisch, W., Spiegel, C., 2008. Neogeneexhumation history of the Mont Blanc massif, western Alps. Tectonics 27, TC4011.

    doi:10.1029/2008TC002257.Glotzbach, C., Reinecker, J., Danisik, M., Rahn, M., Frisch, W., Spiegel, C., 2010.

    Thermal history of the central Gotthard and Aar massifs (European Alps):evidence for steady state, long-term exhumation. J. Geophys. Res. doi:10.1029/2009JF001304.

    Glotzbach, C., van der Beek, P.A., Spiegel,C., 2011. Episodicexhumation of the MontBlancmassif, Western Alps: constraints from numerical modelling of thermochronologydata. Earth Planet. Sci. Lett. 304, 417430. doi:10.1016/j.epsl.2011.02.020.

    Green, P.F., Duddy, I.R., Laslett, G.M., Hegarty, K.A., Gleadow, A.J.W., Lovering, J.F., 1989.Thermal annealing offission tracks in apatite IV: quantitative modelling techniquesand extension to geological timescales. Chem. Geol. 79, 155182.

    Herman, F.,Braun, J.,Dunlap, W.,2007. Tectono-morphicscenarios in theSouthernAlpsof New Zealand. J. Geophys. Res. 112, B04201. doi:10.1029/2004JB003.

    Herman, F., Rhodes, E.J., Braun, J., Heiniger, L., 2010a. Uniform erosion rates and reliefamplitude during glacial cycles in the Southern Alps of New Zealand, as revealedfrom OSL-thermochronology. Earth Planet. Sci. Lett. 297, 183189.

    Herman, F., Copeland, P., Avouac, J.-P., Bollinger,L., Maho, G., Le Fort, P., Rai, S., Foster, D.,Pecher, A., Stuwe, K., Henry, P., 2010b. Exhumation, crustal deformation and thermalstructure of the Nepal Himalaya derived from the inversion of thermochronologicaland thermobarometric data and modeling of the topography. J. Geophys. Res. 115,B06407. doi:10.1029/2008JB006126.

    House, M.A., Wernicke, B.P., Farley, K.A., 1998. Dating topography of the Sierra Nevada,California, using apatite (U-Th)/He ages. Nature 396, 6669.

    Hurford, A.J., 1991. Uplift and cooling pathways derived from fission track analysis andmica dating: a review. Geol. Rundsch. 80, 349368.

    Ketcham,R.A., 2005.Forwardand inversemodelingof low-temperaturethermochronometrydata. Rev. Mineral. Geochem. 58, 275314.

    Mancktelow, N.S., Grasemann, B., 1997. Time-dependent effects of heat advection andtopography on cooling histories during erosion. Tectonophysics 270, 167195.

    Molnar, P.,England, P.,1990.LateCenozoicuplift ofmountainranges andglobal climatechange: chicken or egg? Nature 346, 2934.

    Reinecker, J., Danik,M., Schmid, C.,Glotzbach, C.,Rahn,M., Frisch, W.,Spiegel,C., 2008.

    Tectonic control on the late stage exhumation of the Aar Massif (Switzerland):constraints from apatite fission track and (U-Th)/He data. Tectonics 27, TC6009.doi:10.1029/2007TC002247.

    Reiners, P.W., 2007. Thermochronologic approaches to paleotopography. Rev. Mineral.Geochem. 66, 243267.

    Reiners, P.W., Brandon, M.T., 2006. Using thermochronology to understand orogenicerosion. Annu. Rev. Earth Planet. Sci. 34, 419466.

    Sambridge, M., 1999a. Geophysical inversion with a neighbourhood algorithmI.Searching a parameter space. Geophys. J. Int. 138, 479494.

    Sambridge, M., 1999b. Geophysical inversion with a neighbourhood algorithmII.Appraising the ensemble. Geophys. J. Int. 138, 727746.

    Stephenson, J., Gallagher, K., Holmes, C.C., 2006a. Low temperature thermochronologyand strategies for multiple samples: 2: partition modelling for 2D/3D distributionswith discontinuities. Earth Planet. Sci. Lett. 241, 557570.

    Stephenson, J., Gallagher, K., Holmes, C.C., 2006b. A Bayesian approach to calibratingapatite fission track annealing models for laboratory and geological timescales.Geochim. Cosmochim. Acta 70, 51835200.

    Stwe,K., White,L.,Brown,R.W.,1994.Theinfluenceof erodingtopography on steady-stateisotherms. Application to fission track analysis. Earth Planet. Sci. Lett. 124, 6374.

    Valla, P.G., Herman,F., van derBeek, P.A., Braun,J., 2010. Inversion of thermochronologicalage-elevation profiles to extract independent estimates of denudation and reliefhistoryI: theory and conceptual model. Earth Planet. Sci. Lett. 295, 511522.

    vander Beek, P.A., Valla, P.G., Herman,F., Braun,J., Persano,C., Dobson, K.J., Labrin,E., 2010.Inversion of thermochronological age-elevation profiles to extract independentestimates of denudation andreliefhistoryII:application to theFrenchWestern Alps.Earth Planet. Sci. Lett. 296, 922.

    Vernon, A.J., van der Beek, P.A., Sinclair, H.D., Persano, C., Foeken, J., Stuart, F.M., 2009.Variable late Neogene exhumation of the central European Alps: Low-temperaturethermochronology from the Aar Massif, Switzerland, and the Lepontine Dome,Italy. Tectonics 28, TC5004. doi:10.1029/2008TC002387.

    Wagner, G.A., Reimer, G.M., 1972. Fission track tectonics: the tectonic interpretation offission track apatite ages. Earth Planet. Sci. Lett. 14, 263268.

    Whipp, D.M., Ehlers, T.A.,2007. Influence of groundwater flow on thermochronometer-derived exhumation rates in the central Nepalese Himalaya. Geology 35, 851854.

    Wood, R., Curtis, A., 2004. Geological prior information, and its applications to solvinggeoscientific problems. In: Curtis, A., Wood, R. (Eds.), Geological Prior Information:Geological Society of London Special Publication, 239, pp. 114.

    Zhang, P., Molnar, P., Downs, W.R., 2001. Increased sedimentation rates and grain sizes24 Myr ago due to the influence of climate change on erosion rates. Nature 410,891897.

    322 P.G. Valla et al. / Earth and Planetary Science Letters 307 (2011) 309322

    http://dx.doi.org/10.1029/2008TC002257http://dx.doi.org/10.1029/2009JF001304http://dx.doi.org/10.1029/2009JF001304http://dx.doi.org/10.1029/2009JF001304http://dx.doi.org/10.1016/j.epsl.2011.02.020http://dx.doi.org/10.1029/2004JB003http://dx.doi.org/10.1029/2008JB006126http://dx.doi.org/10.1029/2007TC002247http://dx.doi.org/10.1029/2008TC002387http://dx.doi.org/10.1029/2008TC002387http://dx.doi.org/10.1029/2008TC002387http://dx.doi.org/10.1029/2007TC002247http://dx.doi.org/10.1029/2008JB006126http://dx.doi.org/10.1029/2004JB003http://dx.doi.org/10.1016/j.epsl.2011.02.020http://dx.doi.org/10.1029/2009JF001304http://dx.doi.org/10.1029/2009JF001304http://dx.doi.org/10.1029/2008TC002257