Hostname: page-component-848d4c4894-x24gv Total loading time: 0 Render date: 2024-05-01T00:48:46.229Z Has data issue: false hasContentIssue false

A synthetic study of acoustic full waveform inversion to improve seismic modelling of firn

Published online by Cambridge University Press:  20 March 2023

Emma Pearce*
Affiliation:
School of Earth and Environment, University of Leeds, Leeds, UK
Adam D. Booth
Affiliation:
School of Earth and Environment, University of Leeds, Leeds, UK
Sebastian Rost
Affiliation:
School of Earth and Environment, University of Leeds, Leeds, UK
Paul Sava
Affiliation:
Department of geophysics, Colorado School of Mines, Golden, USA
Tuğrul Konuk
Affiliation:
Department of geophysics, Colorado School of Mines, Golden, USA
Alex Brisbourne
Affiliation:
British Antarctica Survey, Natural Environmental Research Council, Cambridge, UK
Bryn Hubbard
Affiliation:
Department of Geography and Earth Sciences Aberystwyth University, Aberystwyth, UK
Ian Jones
Affiliation:
BrightSkies Geosciense, Maadi, Egypt
*
Author for correspondence: Emma Pearce, E-mail: epearce@unistra.fr
Rights & Permissions [Opens in a new window]

Abstract

The density structure of firn has implications for hydrological and climate modelling and for ice shelf stability. The firn structure can be evaluated from depth models of seismic velocity, widely obtained with Herglotz-Wiechert inversion (HWI), an approach that considers the slowness of refracted seismic arrivals. However, HWI is appropriate only for steady-state firn profiles and the inversion accuracy can be compromised where firn contains ice layers. In these cases, Full Waveform Inversion (FWI) can be more successful than HWI. FWI extends HWI capabilities by considering the full seismic waveform and incorporates reflected arrivals, thus offering a more accurate estimate of a velocity profile. We show the FWI characterisation of the velocity model has an error of only 1.7% for regions (vs. 4.2% with HWI) with an ice slab (20 m thick, 40 m deep) in an otherwise steady-state firn profile.

Type
Letter
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The International Glaciological Society

1. Introduction and motivation

The structure and properties of firn are important to understand for numerous glaciological applications (e.g. Kinar and Pomeroy, Reference Kinar and Pomeroy2007; Diez and others, Reference Diez2014; Schlegel and others, Reference Schlegel2019). For ice shelves, the loss of firn air content is linked to changes in surface elevation that could be misdiagnosed as evidence of basal melting (Holland and others, Reference Holland2011), and firn densification has been implicated in catastrophic ice-shelf collapse (Glasser and others, Reference Glasser and Scambos2008). Effective geophysical characterisation of the subglacial environment of ice masses requires firn effects to be removed, e.g., to correct observations of basal seismic and/or radar reflectivity for energy loss through firn layers (e.g., Kulessa and others, Reference Kulessa2017; Zechmann and others, Reference Zechmann2018).

Ice slabs form in firn where surface melt occurs, including the coastal regions of Antarctica and Greenland. With a lateral extent of tens-to-hundreds of metres, ice slabs make the shallow firn column impermeable (Benson, Reference Benson1962; MacFerrin and others, Reference MacFerrin2019, Miller and others, Reference Miller2022) and can increase its local density from between 400–800 kg m–3 (typical of firn) to that of pure glacier ice, 830–917 kg m−3 (e.g. Hubbard and others, Reference Hubbard2016; MacFerrin and others, Reference MacFerrin2019; Culberg and others, Reference Culberg, Schroeder and Chu2021). Without explicitly accounting for the effects of ice slabs, meltwater runoff can be underestimated: in Greenland, for example, regional climate models suggest runoff to be underestimated by almost 60% if ice slab formation is excluded (MacFerrin and others, Reference MacFerrin2019).

Borehole sampling can determine the thickness of firn layers (e.g., Hubbard and others, Reference Hubbard2016) but geophysical techniques facilitate measurement away from borehole control and can constrain other physical properties of the firn column. Several seismic techniques have been developed for firn characterisation (Kirchner and Bentley, Reference Kirchner and Bentley1979; King and Jarvis, Reference King and Jarvis1991, Reference King and Jarvis2007; Booth and others, Reference Booth2013; Hollmann and others, Reference Hollmann, Treverrow, Peters, Reading and Kulessa2021) but these make assumptions (e.g., that firn density always increases with depth) that limit their applicability. In this study, we report on the application of acoustic Full Waveform Inversion (FWI) techniques (Virieux and Operto, Reference Virieux and Operto2009) as a means of improving seismic surveying for constraining the physical properties of firn. We use synthetic seismic data to highlight promising results for the FWI approach and establish the next steps that would broaden the applicability of the method for firn studies.

2. Seismic methods to model FIRN and ice

Firn structures are detectable with seismic refraction techniques since, as firn is compacted and densified, its seismic velocity increases with depth. Seismic velocity models can therefore be linked, via empirical models, to firn density. A commonly used approach for characterising the seismic velocity through firn from controlled- source seismic Data are Herglotz-Wiechert inversion (HWI) (Herglotz, Reference Herglotz1907; Wiechert, Reference Wiechert1910; Slichter, Reference Slichter1932; Nowack, Reference Nowack1990). HWI obtains a velocity-depth model by considering the slowness (the reciprocal of velocity) of refracted seismic arrivals (e.g., Thiel and Ostenso, Reference Thiel and Ostenso1961, Rege and Godio, Reference Rege and Godio2011; Diez and others, Reference Diez, Eisen, Hofstede, Bohleber and Polom2013). HWI is practical for many firn applications since its key assumption – that seismic velocity increases gradually with depth – is fulfilled by many steady-state firn profiles. However, this assumption is violated for more complex firn profiles, including those which include so-called ‘ice slabs’ (locations in which water has infiltrated and refrozen). In such cases, the ice slab represents a high velocity anomaly, with a velocity reduction likely at its lower interface on returning to unmodified firn. Consequently, when HWI is applied to firn profiles containing infiltration ice, the boundaries of ice slabs are improperly represented, leading to errors in the final velocity model and any derivative density estimate.

2.1 FWI motivation

The limitations of HWI can potentially be overcome with FWI. Where HWI only considers the slowness of first-arrivals, FWI considers both slowness, amplitude and phase information for refracted arrivals (beyond just the first arrival) and reflected wavelets. In so doing, it iteratively improves the match between the recorded and model seismic data predicted by an evolving distribution of subsurface seismic properties. The incorporation of reflected arrivals into FWI also means that it is, in principle, sensitive to more complex velocity structures than can be detected by HWI. For the case of firn structures, this means that FWI should be able to reconstruct the extent of ice slabs by considering reflections from their bounding interfaces.

2.2 Progress on FWI development

We have been exploring the use of FWI to recover the extent of an ice slab located within a firn column. Synthetic firn velocity profiles were generated using the Herron-Langway (HL) (Herron and Langway, Reference Herron and Langway1980) accumulation model with the critical, pure ice and surface snow densities set as 550, 917 and 400 kg m–3 respectively, a 10 m depth temperature of −30°C and an accumulation rate of 0.2 m w.e. a−1. These parameters predict that firn density will increase from 400 to 830 kg m–3; and continues to increase to a maximum ice density, here considered as 917 kg m–3 at which the firn-ice transition would occur at a predicted depth of 200 m. This density profile is used to model seismic velocity with the Kohnen (Reference Kohnen1972) empirical expression. To represent the ice slab, velocities in the depth interval 40–60 m are increased to 3800 m s−1. Throughout this study we consider the synthetic velocity model with ice as the ‘True’ model; models implied by HWI and FWI are benchmarked against this reference model. The FWI algorithm models the acoustic wave equation with the finite-difference (FD) method, due to its simplicity and efficiency compared to other techniques available to solve partial differential equations (Virieux and Operto, Reference Virieux and Operto2009; Zhang and Yao, Reference Zhang and Yao2013). To simplify the source characterisation in the FD approach the source wavelet is modelled using a Ricker wavelet with a peak frequency of 60 Hz. To ensure modelling stability, data are recorded with a time sampling of 0.001 s, for 1 s of propagation (Courant and others, Reference Courant, Friedrichs and Lewy1967). In our FWI, we use a gradient approach to define the fit between recorded and modelled data. The objective function (OF) used is a least-squares (LS) formulation that minimises the sum of the square of the difference between the observed (d) and predicted (p’) datasets:

(1)$$f = \displaystyle{1 \over 2}\vert {\boldsymbol {\,p}^{\prime}}-{\boldsymbol d}\vert _2^2 $$

We use the Madagascar framework (Fomel and others, Reference Fomel, Sava, Vlad, Liu and Bashkardin2013), provided by Center for Wave Phenomena at Colorado School of Mines, to implement this approach (Aragao and Sava, Reference Aragao and Sava2020). To minimise the impact of cycle skipping, we start iterations for the low frequency wavefield, filtered between 3 and 10 Hz, until the objective function updates suggest model convergence. Thereafter, the frequency content is progressively increased in 10 Hz bands, to a maximum of 60 Hz.

3. Results

Synthetic seismic data are forward modelled (Pearce, Reference Pearce2022) from the ‘True’ velocity profile. These data are used to initiate the HWI and FWI, and their output velocity models are compared to the ‘True’ model (Fig. 1). Given its limiting assumptions, the model derived from HWI (blue) cannot resolve the base of the ice slab from the seismic data. Instead, it reaches its maximum velocity (~3750 m s−1) at 40 m depth. This corresponds to the top of the ice slab, but this velocity is incorrectly propagated to the base of the model (Fig. 1a). By contrast, the FWI velocity model (red) detects both contacts of the ice slab: the resolution of its upper contact is lower than HWI, but the increase and deeper decrease of velocity correctly indicates the ice slab's presence. By 80 m depth, the FWI velocity model is comparable to that of the ‘True’ model, again indicating that the ice slab extends from depths 40–80 m. The relative performance of HWI and FWI is benchmarked by defining a percentage error from the ‘True’ velocity model (Fig. 1b), with FWI providing the more accurate representation throughout. Our current models assume a maximum wavelet frequency of 60 Hz; extending the bandwidth at the high frequency end would facilitate improved resolution but at the potential cost of inversion instabilities such as cycle skipping (Hu and others, Reference Hu, Chen, Liu and Abubakar2018).

Fig. 1. (a) Velocity model outputs from HWI (Red) and FWI (Blue) compared to the True model (Black). The HWI generates a velocity profile that stops increasing in velocity at the depth of the ice slab top (40 m). This velocity is then extrapolated to the base of the model. FWI produces a velocity increase at 30 m depth and a decrease at 70 m depth, indicating the ice slab's location. (b) The absolute percentage error between the inversion models for HWI (Red) and (FWI) relative to the True model. The average difference for the total depth shows FWI reduces the error by 2.5%. A maximum difference of 20% for the HWI velocity model is observed in the top 40 m.

Figure 2 shows examples of the seismic traces generated by HWI (red) and FWI (blue), compared to the True model (black). Refractions in near-offset traces (Fig. 2a) are well-represented by both methods. However, at 500 m offset (Fig. 2b), the FWI trace is much closer to that of the reference model, while the over-estimate of velocity in the HWI case predicts a first-arrival that precedes that in the reference trace. Furthermore, FWI is able to model reflected arrivals, and Figure 2c shows an enlarged section of the near-offset traces, showing the insensitivity of HWI to a reflected arrival.

Fig. 2. Seismic data forward modelled from the True (Black), HWI (Red) and FWI (Blue). (a) All data with LMO applied with a velocity of 3800 m s−1. (b) Comparison of trace from an offset of 5 m, the first arrival is modelled well by both the HWI and FWI, the green panel indicates the zoomed in section of figure (d). (c) Comparison of trace from an offset of 500 m. FWI closely reproduces the true data, while HWI poorly represents the true data. (d) Zoom of the reflection at 5 m offset produced by the ice slab visible at near offsets.

4. Future research priorities

The presence of ice slabs influences drainage and meltwater runoff across glaciers and ice masses (MacFerrin and others, Reference MacFerrin2019). In the case of ice shelves, this meltwater increase and decrease in permeability can lead to a reduction in ice shelf stability (Munneke and others, Reference Munneke, Ligtenberg, Van Den Broeke and Vaughan2014). This process was observed on Larsen B Ice Shelf, where firn compaction, meltwater ponding and hydrofracturing were strongly implicated in the shelf's rapid disintegration in 2002 (Scambos and others, Reference Scambos, Bohlander, Shuman and Skvarca2004). Hubbard and others (Reference Hubbard2016) detected a 40 m thick ice slab in the upstream reaches of Larsen C Ice Shelf (LCIS) in borehole Optical Televiewer (OPTV) data, interpreted as being the accumulation of episodic refreezing of meltwater ponds. Here, the firn zone is 10°C warmer and 170 kg m–3 denser than undisturbed firn in the surrounding area. Regional geophysical surveys suggested that the ice slab is at least sixteen kilometres across and several kilometres long; while GPR surveys could constrain thickness, and seismic velocity models (Kulessa and others, Reference Kulessa2019) were consistent with pure ice, neither method could establish both the full depth extent and velocity anomaly of the slab. FWI methods show promise for this application and offer a means of monitoring the development of similar processes at other sites. This potential therefore motivates the future application of FWI to similarly stratigraphically-complex areas of firn, such as interior areas of the Greenland Ice Sheet (MacFerrin and others, Reference MacFerrin2019) and LCIS.

FWI was applied to legacy field datasets from Antarctica's Pine Island Glacier, but none are currently compliant with a stable inversion. Hence, the promising results from synthetic trials reported herein have not yet been validated for field data. The high frequency components of explosive sources used within glaciology lead to cycle skipping when the starting model is not close enough to the true model. Field data have additional difficulties vs synthetic data for FWI, due to the assumption of an acoustic wavefield. The effect of neglecting the elastic component in field data leads to a loss in resolution and accuracy in the recovered velocity model (Agudo and others, Reference Agudo, Da Silva, Warner and Morgan2018). Attenuation was not accounted for, even though firn is known to be highly attenuative and has an effect on the wavelet with offset (e.g. Eisen and others (Reference Eisen2010), King and Jarvis (Reference King and Jarvis1991)).

To obtain a compatible field dataset, particular care must be taken to ensure that the source wavelet is sufficiently consistent and rich in low-frequency content, to facilitate stable FWI. Many examples of terrestrial field data examples exist that satisfy these criteria, allowing for successful FWI and thus improved near-surface velocity models (e.g. Adamczyk and others (Reference Adamczyk, Malinowski and Malehmir2014), Borisov and others (Reference Borisov, Gao, Williamson and Tromp2020), Irnaka and others (Reference Irnaka, Brossier, Métivier, Bohlen and Pan2022)). The extension of this practice into glaciology motivates the acquisition of FWI-compliant data over an ice slab target, and LCIS offers a promising opportunity for this given its apparent complexity and the wealth of data already available for validation. FWI over an area such as this would provide greater constraints on the structure of the firn column, allowing for a more comprehensive reconstruction of seismic properties than currently feasible with existing seismic techniques. If our acoustic FWI algorithm is deemed to be effective, it could in principle undergo further development to recover the anisotropic and/or elastic properties of the firn and deeper ice column (Li and Alkhalifah (Reference Li and Alkhalifah2022), Kan and others (Reference Kan, Chevrot and Monteiller2023)).

Data availability

Synthetic firn seismic velocity profiles are available to download from the Figshare repository, https://doi.org/10.6084/m9.figshare.20765350.v1 (Pearce, Reference Pearce2022).

Acknowledgements

This research was funded by the Natural Environment Research Council of the UK, under Industrial CASE Studentship NE/P009429/1 with CASE funding from ION-GXT. Additional support was obtained from The International Association of Mathematical Geoscientists. Comments from two anonymous reviewers greatly benefitted the preparation of the manuscript.

Footnotes

*

Now at: Ecole et Observatoire des Sciences de la Terre, Institut Terre et Environnement de Strasbourg.

References

Adamczyk, A, Malinowski, M and Malehmir, A (2014) High-resolution near-surface velocity model building using full-waveform inversion, A case study from southwest Sweden. Geophysical Journal International 197(3), 16931704.CrossRefGoogle Scholar
Agudo, ÒC, Da Silva, NV, Warner, M and Morgan, J (2018) Acoustic full-waveform inversion in an elastic world. Geophysics 83(3), R257R271.Google Scholar
Aragao, O and Sava, P (2020) Elastic full-waveform inversion with probabilistic petrophysical model constraints. Geophysics 85(2), R101R111.Google Scholar
Benson, CS (1962) Stratigraphic studies in the snow and firn of the Green- land ice sheet. Technical report, Cold Regions Research And Engineering Lab HANOVER NH.Google Scholar
Booth, AD and 5 others (2013) A comparison of seismic and radar methods to establish the thickness and density of glacier snow cover. Annals of Glaciology 54(64), 7382.Google Scholar
Borisov, D, Gao, F, Williamson, P and Tromp, J (2020) Application of 2D full-waveform inversion on exploration land data: application of 2D FWI on land data. Geophysics 85(2), R75R86.Google Scholar
Courant, R, Friedrichs, K and Lewy, H (1967) On the partial difference equations of mathematical physics. IBM journal of Research and Development 11(2), 215234.Google Scholar
Culberg, R, Schroeder, DM and Chu, W (2021) Extreme melt season ice layers reduce firn permeability across Greenland. Nature Communications 12(1), 19.CrossRefGoogle ScholarPubMed
Diez, A and 7 others (2014) Influence of ice crystal anisotropy on seismic velocity analysis. Annals of Glaciology 55(67), 97106.Google Scholar
Diez, A, Eisen, O, Hofstede, C, Bohleber, P and Polom, U (2013) Joint interpretation of explosive and vibroseismic surveys on cold firn for the investigation of ice properties. Annals of Glaciology 54(64), 201210.Google Scholar
Eisen, O and 6 others (2010) A new approach for exploring ice sheets and sub-ice geology. EOS. Transactions American Geophysical Union 91(46), 429430, 260.Google Scholar
Fomel, S, Sava, PC, Vlad, I, Liu, Y and Bashkardin, V (2013) Madagascar: open-source software project for multidimensional data analysis and reproducible computational experiments. Journal of Open Research Software, 1(1).Google Scholar
Glasser, NF and Scambos, TA (2008) A structural glaciological analysis of the 2002 Larsen B ice-shelf collapse. Journal of Glaciology 54(184), 316.Google Scholar
Herglotz, G (1907) Über das benndorfsche problem der fortpflanzungsgeschwindigkeit der erdbebenstrahlen. Zeitschr für Geophys 8, 145147.Google Scholar
Herron, MM and Langway, CC (1980) Firn densification: an empirical model. Journal of Glaciology 25(93), 373385, ISSN 00221430.Google Scholar
Holland, PR and 6 others (2011) The air content of Larsen ice shelf. Geophysical Research Letters 38(10).Google Scholar
Hollmann, H, Treverrow, A, Peters, LE, Reading, AM and Kulessa, B (2021) Seismic observations of a complex firn structure across the Amery ice shelf, East Antarctica. Journal of Glaciology 67(265), 777787.Google Scholar
Hu, W, Chen, J, Liu, J and Abubakar, A (2018) Retrieving low wavenumber information in FWI: an overview of the cycle-skipping phenomenon and solutions. IEEE Signal Processing Magazine 35(2), 132141.Google Scholar
Hubbard, B and 10 others (2016) Massive subsurface ice formed by refreezing of ice-shelf melt ponds. Nature Communications 7(1), 16.Google Scholar
Irnaka, TM, Brossier, R, Métivier, L, Bohlen, T and Pan, Y (2022) 3-D Multicomponent full waveform inversion for shallow-seismic target: Ettlingen line case study. Geophysical Journal International 229(2), 10171040.Google Scholar
Kan, LY, Chevrot, S and Monteiller, V (2023) A consistent multiparameter Bayesian full waveform inversion scheme for imaging heterogeneous isotropic elastic media. Geophysical Journal International 232(2), 864883.Google Scholar
Kinar, N and Pomeroy, J (2007) Determining snow water equivalent by acoustic sounding. Hydrological Processes: An International Journal 21(19), 26232640. 18.Google Scholar
King, EC and Jarvis, EP (1991) Effectiveness of different shooting techniques in Antarctic firn. First Break 9(6), 281288.Google Scholar
King, EC and Jarvis, EP (2007) Use of shear waves to measure Poisson's ratio in polar firn. Journal of Environmental and Engineering Geophysics 12(1), 1521. 23, 26.Google Scholar
Kirchner, JF and Bentley, CR (1979) Seismic short-refraction studies on the Ross ice shelf, Antarctica. Journal of Glaciology 24(90), 313319.Google Scholar
Kohnen, H (1972) Über die Beziehung zwischen seismischen Geschwindigkeiten und der Dichte in Firn und Eis. Geophysics 38(5), 925935.Google Scholar
Kulessa, B and 10 others (2017) Seismic evidence for complex sedimentary control of Greenland ice sheet flow. Science Advances 3(8), e1603071.Google Scholar
Kulessa, B and 10 others (2019) Seawater softening of suture zones inhibits fracture propagation in Antarctic ice shelves. Nature Communications 10(1), p.5491.Google Scholar
Li, Y and Alkhalifah, T (2022) Target-oriented high-resolution elastic full waveform inversion with an elastic redatuming method. Geophysics 87(5), 177.Google Scholar
MacFerrin, M and 9 others and others (2019) Rapid expansion of Greenland's low permeability ice slabs. Nature 573(7774), 403407.Google Scholar
Miller, JZ and 5 others (2022) An empirical algorithm to map perennial firn aquifers and ice slabs within the Greenland ice sheet using satellite L-band microwave radiometry. The Cryosphere 16(1), 103125.Google Scholar
Munneke, PK, Ligtenberg, SR, Van Den Broeke, MR and Vaughan, DG (2014) Firn air depletion as a precursor of Antarctic ice-shelf collapse. Journal of Glaciology 60(220), 205214.Google Scholar
Nowack, RL (1990) Tomography and the Herglotz-Wiechert inverse formulation. Pure and Applied Geophysics 133(2), 305315.Google Scholar
Pearce, E (2022) Synthetic velocity models of firn. Figshare. Dataset. https://doi.org/10.6084/m9.figshare.20764423.v1.CrossRefGoogle Scholar
Rege, R and Godio, A (2011) Geophysical investigation for mechanical properties of a glacier. In Geophys. Res. Abstr, volume 13.Google Scholar
Scambos, TA, Bohlander, J, Shuman, CA and Skvarca, P (2004) Glacier acceleration and thinning after ice shelf collapse in the Larsen B embayment, Antarctica. Geophysical Research Letters 31(18).Google Scholar
Schlegel, R and 8 others (2019) Comparison of elastic moduli from seismic diving-wave and ice-core microstructure analysis in Antarctic polar firn. Annals of Glaciology 60(79), 220230.CrossRefGoogle Scholar
Slichter, LB (1932) The theory of the interpretation of seismic travel-time curves in horizontal structures. Physics 3(6), 273295.Google Scholar
Thiel, E and Ostenso, NA (1961) Seismic studies on Antarctic ice shelves. Geophysics 26(6), 706715.Google Scholar
Virieux, J and Operto, S (2009) An overview of full-waveform inversion in exploration geophysics. Geophysics 74(6), WCC1WCC26.Google Scholar
Wiechert, E (1910) Bestimmung des Weges der Erdbebenwellen im Erdinneren, I. Theoretisches. Physikalische Zeitschrift 11, 294304.Google Scholar
Zechmann, JM and 5 others (2018) Active seismic studies in valley glacier settings: strategies and limitations. Journal of Glaciology 64(247), 796810.Google Scholar
Zhang, JH and Yao, ZX (2013) Optimized finite-difference operator for broadband seismic wave modelling. Geophysics 78(1), A13A18.Google Scholar
Figure 0

Fig. 1. (a) Velocity model outputs from HWI (Red) and FWI (Blue) compared to the True model (Black). The HWI generates a velocity profile that stops increasing in velocity at the depth of the ice slab top (40 m). This velocity is then extrapolated to the base of the model. FWI produces a velocity increase at 30 m depth and a decrease at 70 m depth, indicating the ice slab's location. (b) The absolute percentage error between the inversion models for HWI (Red) and (FWI) relative to the True model. The average difference for the total depth shows FWI reduces the error by 2.5%. A maximum difference of 20% for the HWI velocity model is observed in the top 40 m.

Figure 1

Fig. 2. Seismic data forward modelled from the True (Black), HWI (Red) and FWI (Blue). (a) All data with LMO applied with a velocity of 3800 m s−1. (b) Comparison of trace from an offset of 5 m, the first arrival is modelled well by both the HWI and FWI, the green panel indicates the zoomed in section of figure (d). (c) Comparison of trace from an offset of 500 m. FWI closely reproduces the true data, while HWI poorly represents the true data. (d) Zoom of the reflection at 5 m offset produced by the ice slab visible at near offsets.