Hostname: page-component-8448b6f56d-tj2md Total loading time: 0 Render date: 2024-04-23T22:57:11.999Z Has data issue: false hasContentIssue false

Five decades of radioglaciology

Published online by Cambridge University Press:  09 March 2020

Dustin M. Schroeder*
Affiliation:
Department of Geophysics, Stanford University, Stanford, USA Department of Electrical Engineering, Stanford University, Stanford, USA
Robert G. Bingham
Affiliation:
School of GeoSciences, University of Edinburgh, Edinburgh, UK
Donald D. Blankenship
Affiliation:
Institute for Geophysics, University of Texas, Austin, USA
Knut Christianson
Affiliation:
Department of Earth and Space Sciences, University of Washington, Seattle, USA
Olaf Eisen
Affiliation:
Alfred-Wegener-Institut, Helmholtz-Zentrum für Polar- und Meeresforschung, Bremerhaven, Germany University of Bremen, Bremen, Germany
Gwenn E. Flowers
Affiliation:
Department of Earth Sciences, Simon Fraser University, Vancouver, Canada
Nanna B. Karlsson
Affiliation:
Geological Survey of Denmark and Greenland, Copenhagen, Denmark
Michelle R. Koutnik
Affiliation:
Department of Earth and Space Sciences, University of Washington, Seattle, USA
John D. Paden
Affiliation:
Center for the Remote Sensing of Ice Sheets, University of Kansas, Lawrence, USA
Martin J. Siegert
Affiliation:
Grantham Institute, and Department of Earth Science and Engineering, Imperial College London, London, UK
*
Author for correspondence: Dustin M. Schroeder, E-mail: dustin.m.schroeder@stanford.edu
Rights & Permissions [Opens in a new window]

Abstract

Radar sounding is a powerful geophysical approach for characterizing the subsurface conditions of terrestrial and planetary ice masses at local to global scales. As a result, a wide array of orbital, airborne, ground-based, and in situ instruments, platforms and data analysis approaches for radioglaciology have been developed, applied or proposed. Terrestrially, airborne radar sounding has been used in glaciology to observe ice thickness, basal topography and englacial layers for five decades. More recently, radar sounding data have also been exploited to estimate the extent and configuration of subglacial water, the geometry of subglacial bedforms and the subglacial and englacial thermal states of ice sheets. Planetary radar sounders have observed, or are planned to observe, the subsurfaces and near-surfaces of Mars, Earth's Moon, comets and the icy moons of Jupiter. In this review paper, and the thematic issue of the Annals of Glaciology on ‘Five decades of radioglaciology’ to which it belongs, we present recent advances in the fields of radar systems, missions, signal processing, data analysis, modeling and scientific interpretation. Our review presents progress in these fields since the last radio-glaciological Annals of Glaciology issue of 2014, the context of their history and future prospects.

Type
Article
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 in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s) 2020

Introduction

Five decades of radioglaciology (the use of radio waves to investigate ice masses of all types) since the first data were published have seen a progression of instruments and platforms, as well as data processing and analysis approaches applied to a growing data archive (e.g. Stern, Reference Stern1930; Steenson, Reference Steenson1951; Robin, Reference Robin1975; Gogineni and others, Reference Gogineni, Chuah, Allen, Jezek and Moore1998; Dowdeswell and Evans, Reference Dowdeswell and Evans2004; Allen, Reference Allen2008; Turchetti and others, Reference Turchetti, Dean, Naylor and Siegert2008). Radar-sounding (also known as ice-penetrating radar) data have been used to observe ice thickness, basal topography and englacial layers across Antarctica and Greenland, as well as many ice caps and glaciers. Major data-collection efforts started in the late 1960s and early 1970s, including a collaboration between the Technical University of Denmark, Scott Polar Research Institute, and National Science Foundation (TUD-SPRI-NSF) to map the bed of Antarctica. Other early surveys were also led by Russia, Germany, Iceland, Italy, China, and Canada (among others) across Antarctica and Greenland, as well as Iceland, Arctic Ice Caps, and mountain glaciers (e.g. Drewry, Reference Drewry1983; Bingham and Siegert, Reference Bingham and Siegert2007; Björnsson, Reference Björnsson and Pálsson2020; Popov, Reference Popov2020). Planetary radar sounders have also been used, or are planned, to observe the subsurface and near-surface conditions of Mars, Earth's moon, comets and the icy moons of Jupiter (e.g. Seu and others, Reference Seu2007; Jordan and others, Reference Jordan2009; Kofman and others, Reference Kofman, Orosei and Pettinelli2010; Bruzzone and others, Reference Bruzzone2013; Kofman and others, Reference Kofman2015; Patterson and others, Reference Patterson2017; Blankenship and others, Reference Blankenship2018). Fully exploiting the valuable information from these data, such as ice-sheet bed topography, the distribution of subglacial water, the spatial variation of basal melt, the transition between frozen and thawed bed conditions, englacial temperature, histories of accumulation, flow, and the distribution of age in ice masses remains an active area of international research. In this review paper, and the thematic issue of the Annals of Glaciology on ‘Five decades of radioglaciology’ to which it belongs, we present recent advances in the field in the context of their history and future prospects. We include papers published in this issue, topics presented at an International Glaciological Society Symposium on the same theme hosted at Stanford University during the summer of 2019, and work added to the published literature since the last thematic Symposium and Annals issue focused on radioglaciology in 2014.

Data

The data collected by radar surveys in the last five decades have transformed our appreciation of glacier and ice-sheet beds and how ice flows over them. Prior to this era, such information was gained from seismic data, taking orders of magnitude longer to acquire. Early radar surveys witnessed significant improvements in survey design, instrument capability (e.g. the Technical University of Denmark System), platforms (e.g. the US Hercules LC-130) and coastal airstrips, leading to systematic surveys of the Greenland and Antarcticice sheets (Sorge, Reference Sorge1933; Robin, Reference Robin1958; Gudmandsen, Reference Gudmandsen1975; Drewry, Reference Drewry1983). In the case of Antarctica, the TUD-SPRI-NSF collaboration collected over 400 000 line-km of data during the 1970s and, in some regions, these data provide the only measurements ever taken. By the early 1980s, those first long-range airborne radar surveys had ended, giving way to regional studies collected by, for example, Russian and German programs and the US Support Office for Aerogeophysical Research (SOAR) in Antarctica (Blankenship and others, Reference Blankenship1993; Behrendt and others, Reference Behrendt1994; Bell and others, Reference Bell1998; Hempel and others, Reference Hempel, Thyssen, Gundestrup, Clausen and Miller2000; Masolov and others, Reference Masolov, Popov, Lukin, Sheremetyev and Popkov2006; Dean and others, Reference Dean, Naylor, Turchetti and Siegert2008; Turchetti and others, Reference Turchetti, Dean, Naylor and Siegert2008). Other examples include surveys of glaciers and ice caps in Iceland, Alpine glaciers, Svalbard and the Russian and Canadian Arctic (e.g. Dowdeswell and others, Reference Dowdeswell1986, Reference Dowdeswell2002, Reference Dowdeswell, Benham, Gorman, Burgess and Sharp2004; Björnsson and others, Reference Björnsson1996; Fürst and others, Reference Fürst2018; Pritchard and others, Reference Pritchard, King, McCarthy and Mayer2020).

In the early 2000s, the Bedmap Consortium produced a new compilation of radar data from Antarctica for which the TUD-NSF-SPRI data still formed by far the most significant contribution, with dozens of other regional-scale surveys that form a patchwork coverage of parts of the ice sheet while other regions remained completely free of data (Lythe and Vaughan, Reference Lythe and Vaughan2001). Bedmap2 followed a decade later, including additional regional surveys as well as long- and medium-range airborne studies returned in 2008 by the US–UK–Australia–China–French ICECAP project and NASA's Operation IceBridge (OIB) (Holt and others, Reference Holt2006; Vaughan and others, Reference Vaughan2006; Bell and others, Reference Bell2011; Young and others, Reference Young2011; Ross and others, Reference Ross2012; Fretwell and others, Reference Fretwell2013). However, several regions remained free of data (Pritchard, Reference Pritchard2014). Other compilations are now due that will incorporate new data that have been acquired to fill many of Bedmap2's gaps, including, for example, across Marie Byrd Land, West Antarctica, the Recovery Basin/South Pole, the Dome F region and Princess Elisabeth Land, as well as newly remastered TUD-NSF-SPRI film data and updated thickness measurements for the Ross Ice Shelf (Tang and others, Reference Tang, Guo, Sun, Wang and Cui2016; Young and others, Reference Young, Schroeder, Blankenship, Kempf and Quartini2016; Popov, Reference Popov2017; Humbert and others, Reference Humbert, Steinhage, Helm, Beyer and Kleiner2018; Jordan and others, Reference Jordan2018a; Karlsson and others, Reference Karlsson2018; Morlighem and others, Reference Morlighem2019; Paxman and others, Reference Paxman2019; Schroeder and others, Reference Schroeder2019; Tinto and others, Reference Tinto2019). Compared to Antarctica, surveys of Greenland starting in the 1990s by the University of Kansas as part of NASA's Program for Arctic Regional Climate Assessment (PARCA) and later OIB have led to relatively abundant and mutually interpretable observations of the ice-sheet bed and englacial properties (Bamber and others, Reference Bamber2013; Gogineni and others, Reference Gogineni2014; MacGregor and others, Reference MacGregor2015a; Morlighem and others, Reference Morlighem2017).

In addition to the collection of radar sounding profiles, interpolation is a critical component of producing bed topography maps. Previous approaches focused on grid interpolation techniques such as spline interpolation or kriging (e.g. Fretwell and others, Reference Fretwell2013). However, in many regions, this gridded topography falls short of capturing topography at the scales most critical for resolving ice-flow processes (Durand and others, Reference Durand, Gagliardini, Favier, Zwinger and Le Meur2011; King and others, Reference King, Pritchard and Smith2016; Bingham and others, Reference Bingham2017; Kyrke-Smith and others, Reference Kyrke-Smith, Gudmundsson and Farrell2018). For this reason, other approaches, such as mass-conservation modeling or geostatistical approaches, which can provide multiple observation-consistent realizations, provide improved interpolations of bed topography (e.g. Rasmussen, Reference Rasmussen1988; Warner and Budd, Reference Warner and Budd2000; Goff and others, Reference Goff, Powell, Young and Blankenship2014; Morlighem and others, Reference Morlighem2017; MacKie and others, Reference MacKie and Schroeder2019).

Future surveys are unlikely to resemble those conducted previously since ice-sheet models require that data are collected with strategies optimized for their purpose, including flowlines for process interpretation, ground-based time-series for local process monitoring, and repeat flights (Kingslake and others, Reference Kingslake2014; Nicholls and others, Reference Nicholls2015; Chu and others, Reference Chu2016; Khazendar and others, Reference Khazendar2016; Holschuh and others, Reference Holschuh, Parizek, Alley and Anandakrishnan2017; Davies and others, Reference Davies2018; Schroeder and others, Reference Schroeder, Hilger, Paden, Young and Corr2018; Young and others, Reference Young2018; Bartlett and others, Reference Bartlett2020). These process- and site-specific surveys can also take advantage of systems with wider bandwidths and larger antenna arrays that provide enhanced performance, but with more limited range (e.g. Rodriguez-Morales and others, Reference Rodriguez-Morales2013; Kjær and others, Reference Kjær2018). Ultimately, new platforms, such as rovers, drones and satellites stand to transform the way radar-sounding observations are made (Jezek and others, Reference Jezek2006; Koh and others, Reference Koh, Lever, Arcone, Marshall and Ray2010; Arcone and others, Reference Arcone2016; Freeman and others, Reference Freeman, Pi and Heggy2017; Dall and others, Reference Dall, Corr, Walker, Rommen and Lin2018; Carrer and others, Reference Carrer, Gerekos and Bruzzone2018; Gogineni and others, Reference Gogineni2018; Culberg and Schroeder, Reference Culberg and Schroeder2019; Arnold and others, Reference Arnold, Leuschen, Paden, Hale and Keshmiri2020).

Systems

Early radar sounding systems spanned a range of frequency, bandwidth, power and array configurations including both short mono-pulse and chirped-waveform systems (Allen, Reference Allen2008; Gärtner-Roer and others, Reference Gärtner-Roer2014). However, until the 1990s (and the availability of faster and lower-cost electronics) the data recorded remained ‘incoherent’, limiting the azimuth resolution and processing gain below that achievable with phase-coherent stacking and Synthetic Aperture Radar (SAR) processing (Musil and Doake, Reference Musil and Doake1987; Hamran and Aarholt, Reference Hamran and Aarholt1993; Leuschen and others, Reference Leuschen, Gogineni and Tammana2000; Legarsky and others, Reference Legarsky, Gogineni and Akins2001; Hélière and others, Reference Hélière, Lin, Corr and Vaughan2007; Peters and others, Reference Peters2007). For stationary ground-based systems, a similar gain in the achievable post-processing signal tonoise ratio (SNR) and range-estimate precision has been achieved by coherent ‘phase-sensitive’ frequency-modulated continuous wave (FMCW) radars (Nicholls and others, Reference Nicholls2015).

Just as coherent radar sounders enabled improved along-track resolution and processing gain, the development of systems with multi-channel cross-track arrays improved cross-track resolution, processing gain, clutter discrimination and swath mapping (Gogineni and others, Reference Gogineni, Chuah, Allen, Jezek and Moore1998; Paden and others, Reference Paden, Akins, Dunson, Allen and Gogineni2010; Wu and others, Reference Wu2011; Rodriguez-Morales and others, Reference Rodriguez-Morales2013; Castelletti and others, Reference Castelletti2017; Holschuh and others, Reference Holschuh, Christianson, Paden, Alley and Anandakrishnan2020; Scanlan and others, Reference Scanlan, Rutishauser, Young and Blankenship2020). This is also true for ground-based multiple input, multipleoutputimplementations of the ‘phase-sensitive’ FMCW radars mentioned above (Young and others, Reference Young2018). While these multi-channel sounders do achieve some diversity in viewing angle and englacial propagation, true bistatic observations and tomographic inversions can be exploited to provide much richer constraints on subsurface properties including, for example, using commercial pulsed ground-penetrating radar (GPR) systems in common mid-pointor borehole configurations to achieve wider (though coherence-limited) offsets (e.g. Kofman and others, Reference Kofman2015; Holschuhand others, Reference Holschuh, Christianson, Anandakrishnan, Alley and Jacobel2016; Patterson and others, Reference Patterson2017; Church and others, Reference Church2019).

The evolution of distinct radar sounding systems has resulted in a diversity of frequencies, spanning HF (3–30 MHz), VHF (30–300 MHz), UHF (300 MHz–3 GHz) and higher frequency bands (Gudmandsen, Reference Gudmandsen1975; Paden and others, Reference Paden2005; Hélière and others, Reference Hélière, Lin, Corr and Vaughan2007; Peters and others, Reference Peters2007; Allen, Reference Allen2008; Shi and others, Reference Shi2010; Hindmarsh and others, Reference Hindmarsh2011; Rignot and others, Reference Rignot, Mouginot, Larsen, Gim and Kirchner2013; Rodriguez-Morales and others, Reference Rodriguez-Morales2013; Dall and others, Reference Dall, Corr, Walker, Rommen and Lin2018; Yan and others, Reference Yan, Gogineni and O'Neill2018). Although, this diversity can make it challenging to compare or combine distinct datasets, it also offers the opportunity to probe the radio-frequency responseof the ice sheet to constrain conditions and processes with greater fidelity (e.g. Carrer and Bruzzone, Reference Carrer and Bruzzone2017; Winter and others, Reference Winter2017).

In addition to systems capable of recording amplitude, phase and channel information, radar sounder development has also included systems that record multiple polarizations (e.g. Vaughan and others, Reference Vaughan2006; Dall and others, Reference Dall2010). These systems allow for the analysis of crystal-fabric orientation from polarization information (Doake and others, Reference Doake, Corr and Jenkins2002; Fujita and others, Reference Fujita, Matsuoka, Maeno and Furukawa2003; Matsuoka and others, Reference Matsuoka2003; Eisen and others, Reference Eisen, Hamann, Kipfstuhl, Steinhage and Wilhelms2007; Drews and others, Reference Drews2012; Li and others, Reference Li2018; Wang and others, Reference Wang2018; Jordan and others, Reference Jordan, Schroeder, Castelletti, Li and Dall2019). This information can be used to constrain the depth distribution of the crystal orientation fabric along survey lines, enabling the investigation of processes occurring in ice masses, comparison to ice-dynamic models, and interpretation of particle-astrophysical observations (e.g. Jordan and others, Reference Jordan, Schroeder, Castelletti, Li and Dall2019, Reference Jordan, Schroeder, Elsworth and Siegfried2020a,b; Shoemaker and others, Reference Shoemaker2020).

Technical advances in available hardware have also allowed the development of stationary systems designed for long-term (several months to years) autonomous operation with repeated observations over cycles ranging from minutes to days, targeting the temporal evolution of a particular site (Nicholls and others, Reference Nicholls2015; Kendrick and others, Reference Kendrick2018; Mingo and others, Reference Mingo, Flowers, Crawford, Mueller and Bigelow2020; Vankova and others, Reference Vankova2020). To further address the power demands of generating an active radar signal (particularly in the extreme resource constraints of planetary missions) passive radar sounding is also being developed as a new radioglaciological technique to exploit Jovian radio noise, or that from the Sun, as sources for radio echo detection, with the promise to enable pervasive monitoring of subsurface conditions by low-cost, low-power sensor networks (Romero-Wolf and others, Reference Romero-Wolf2015, Reference Romero-Wolf2016; Schroeder and others, Reference Schroeder2016b; Peters and others, Reference Peters, Schroeder, Castelletti, Haynes and Romero-Wolf2018).

Processing

Processing radar-sounding data turns low-SNR, low-resolution, high-clutter raw data into usable radargrams. With the exception of a subset of short-pulse systems such as commercial GPRs and some legacy sounders still in use today nearly all radar sounder processing begins with pulse-compression of a chirped waveform using some windowing function for range-sidelobe suppression, some amount of on-board pre-summing to increase SNR and moderate data-rates, and filtering (e.g. Peters and others, Reference Peters2007; Booth and others, Reference Booth, Clark and Murray2010; Lilien and others, Reference Lilien, Hills, Driscol, Jacobel and Christianson2020; Wang and others, Reference Wang2020). For coherent radar-sounding data, along-track SAR focusing is also nearly ubiquitous to improve the SNR, signal to clutter ratio and azimuth resolution (Legarsky and others, Reference Legarsky, Gogineni and Akins2001; Hélière and others, Reference Hélière, Lin, Corr and Vaughan2007; Peters and others, Reference Peters2007). Additionally, azimuth processing that evaluates along-track coherence, multiple apertures, large coherent apertures, layer-specific phase histories, or squinted processing, enhance layer resolution or provide information about the scattering function and fine-scale geometry of the bed (e.g. Oswald and Gogineni, Reference Oswald and Gogineni2008; Schroeder and others, Reference Schroeder, Blankenship, Raney and Grima2014a; Heister and Scheiber, Reference Heister and Scheiber2018; Castelletti and others, Reference Castelletti, Schroeder, Mantelli and Hilger2019; Ferro, Reference Ferro2019). For multi-channel sounders, cross-track processing can considerably increase the level of resolution, with large benefits for tomographic swath imaging of the ice bottom and internal structure, in particular of irregular disturbances of basal ice, such as folds or entrained matter (e.g. Paden and others, Reference Paden, Akins, Dunson, Allen and Gogineni2010; Wu and others, Reference Wu2011; Rodriguez-Morales and others, Reference Rodriguez-Morales2013; Castelletti and others, Reference Castelletti2017; Young and others, Reference Young2018).

In addition to the instantaneous or single-survey coherence required for focusing and array processing, modern high-stability and low-noise systems make it feasible to perform repeat-pass interferometric analysis on sounding data from ground-based platforms (Kingslake and others, Reference Kingslake2014; Nicholls and others, Reference Nicholls2015). While point-based observations of phase changes over time periods, ranging from months to years, are now widespread, for example, to deduce basal melt rates of ice shelves and vertical velocities in ice sheets, its spatial application to large airborne surveys is relatively recent and, as yet, rarely applied (e.g. Corr and others, Reference Corr, Jenkins, Nicholls and Doake2002; Castelletti and Schroeder, Reference Castelletti and Schroeder2017; Stewart and others, Reference Stewart, Christoffersen, Nicholls, Williams and Dowdeswell2019).

Another critical area of innovation in radioglaciological data processing and analysis is automatic methods for radargram image interpretation. These include algorithms for layer tracking, bed and surface mapping and basal feature categorization (Sime and others, Reference Sime, Hindmarsh and Corr2011; Crandall and others, Reference Crandall, Fox and Paden2012; Ferro and Bruzzone, Reference Ferro and Bruzzone2012; Ilisei and Bruzzone, Reference Ilisei and Bruzzone2015; Panton and Karlsson, Reference Panton and Karlsson2015; Carrer and Bruzzone, Reference Carrer and Bruzzone2016; Rahnemoonfar and others, Reference Rahnemoonfar, Fox, Yari and Paden2017; Berger and others, Reference Berger2018; Donini and others, Reference Donini, Thakur, Bovolo and Bruzzone2019). Success of these approaches is a prerequisite to be able to cope efficiently with the data volume of future surveys and effectively exploit their information content.

Ice sheet and glacier bed conditions

Five decades of radioglaciology have produced a diverse array of information pertaining to subglacial conditions. The vast majority of surveys have been motivated by the primary imperative of locating the bed reflector either to estimate the total volume and sea-level potential of the major ice sheets or to map basal topography (e.g. Bailey and others, Reference Bailey, Evans and Robin1964; Gudmandsen, Reference Gudmandsen1969; Bamber and others, Reference Bamber2013; Fretwell and others, Reference Fretwell2013). In the last two decades, the emphasis has expanded to the investigation of the geometric, thermal and material properties of the basal interface, by using the sounder-appropriate radar equation to solve for either basal reflectivity or echo character (Peters and others, Reference Peters, Blankenship and Morse2005; Oswald and Gogineni, Reference Oswald and Gogineni2008; Schroeder and others, Reference Schroeder, Blankenship and Young2013; Grima and others, Reference Grima, Schroeder, Blankenship and Young2014b; Haynes and others, Reference Haynes, Chapin and Schroeder2018b; Haynes, Reference Haynes2020).

Radar sounding data encode a range of information about the roughness of the basal interface. The most common glaciological definition of roughness is the extent to which terrain varies vertically over a given horizontal distance (Rippin and others, Reference Rippin2014). As mapped across a number of regions of Antarctica and Greenland, roughness variations at the multi-kilometer scale inform us about present and past ice-stream and ice-stream tributary locations (Siegert and others, Reference Siegert2004; Bingham and Siegert, Reference Bingham and Siegert2007, Reference Bingham and Siegert2009; Rippin and others, Reference Rippin2014; Frank and others, Reference Franke2020). Additionally, basal roughness at the wavelength-scale can affect the character of the reflected echo including its specularity (or spread in Doppler), waveform abruptness, statistical distribution of echo amplitudes, as well as the radar-derived topography itself (Goff and others, Reference Goff, Powell, Young and Blankenship2014; Grima and others, Reference Grima, Blankenship, Young and Schroeder2014a; Rippin and others, Reference Rippin2014; Schroeder and others, Reference Schroeder, Blankenship, Raney and Grima2014a; Jordan and others, Reference Jordan2017; Heister and Scheiber, Reference Heister and Scheiber2018; Eisen and others, Reference Eisen, Winter, Steinhage, Kleiner and Humbert2020; Franke and others, Reference Franke2020; King, Reference King2020). Principles from these studies have also been translated to paleoglacial landscapes and have also been compared to contemporary bed morphology and lithology (Gudlaugsson and others, Reference Gudlaugsson, Humbert, Winsborrow and Andreassen2013; Schroeder and others, Reference Schroeder, Blankenship, Young, Witus and Anderson2014c; Falcini and others, Reference Falcini, Rippin, Krabbendam and Selby2018; Cooper and others, Reference Cooper2019; Muto and others, Reference Muto, Alley, Parizek and Anandakrishnan2019; Holschuh and others, Reference Holschuh, Christianson, Paden, Alley and Anandakrishnan2020).

In radioglaciology, although reflectivity is used as an umbrella term encompassing all methods used to interrogate variations in the magnitude of the bed echo, it most commonly and appropriately refers to changes in the material properties (and therefore Fresnel reflection coefficient) of the ice–bed interface (Peters and others, Reference Peters, Blankenship and Morse2005). While there are challenges in correcting or constraining attenuation or surface roughness losses, the basal thermal state (frozen or thawed and the presence or absence of water) fundamentally affects the reflection coefficient (Peters and others, Reference Peters, Blankenship and Morse2005; Matsuoka, Reference Matsuoka2011; Schroeder and others, Reference Schroeder, Grima and Blankenship2016a). The reflection coefficient can provide a constraint on where the bed is frozen or thawed, the reach and character of ocean water at the grounding line, and basal conditions of ice streams (Peters and others, Reference Peters, Blankenship and Morse2005; Jacobel and others, Reference Jacobel, Welch, Osterhouse, Pettersson and MacGregor2009; Ashmore and others, Reference Ashmore, Bingham, Hindmarsh, Corr and Joughin2014; Christianson and others, Reference Christianson2016). The presence and volume of inferred basal water bodies have also been used to place constraints on the basal thermal state and/or geothermal flux, while layer drawdown has also been used to constrain basal melt rates and geothermal flux (Fahnestock and others, Reference Fahnestock, Abdalati, Joughin, Brozena and Gogineni2001; Catania and others, Reference Catania, Conway, Raymond and Scambos2006; Buchardt and Dahl-Jensen, Reference Buchardt and Dahl-Jensen2007; Schroeder and others, Reference Schroeder, Blankenship, Young and Quartini2014b; Rezvanbehbahani and others, Reference Rezvanbehbahani, Stearns, Kadivar, Walker and van der Veen2017, Reference Rezvanbehbahani, Stearns, van der Veen, Oswald and Greve2019; Seroussi and others, Reference Seroussi, Ivins, Wiens and Bondzio2017; Jordan and others, Reference Jordan2018a,Reference Jordanb).

Perhaps the most widely and successfully studied basal feature with radar sounding data has been subglacial water bodies, particularly subglacial lakes in Antarctica using the principle that subglacial water results in reflections brighter than surrounding bed echoes in radar data (Oswald and Robin, Reference Oswald and Robin1973; Peters and others, Reference Peters, Blankenship and Morse2005; Wright and Siegert, Reference Wright and Siegert2012). Because of the coherent specular character of subglacial water, small fractional areas can dominate the echo both in terms of reflectivity and geometric spreading (Haynes and others, Reference Haynes, Chapin and Schroeder2018b). This has been exploited to automatically detect lakes in radar sounding data (Carter and others, Reference Carter2007; Ilisei and others, Reference Ilisei, Khodadadzadeh, Ferro and Bruzzone2018). Additionally, lake-bottom echoes have been used to probe water thickness and conductivity (Gorman and Siegert, Reference Gorman and Siegert1999). Surface altimetry data have also been used to infer active lakes around Antarctica where the ice surface has been observed to rise and fall, yet, surprisingly, these lakes typically do not have higher reflectivities than their surroundings in radar data, showing that we still have much to learn about Antarctic subglacial lakes (Carter and others, Reference Carter2007; Smith and others, Reference Smith, Fricker, Joughin and Tulaczyk2009; Siegfried and others, Reference Siegfried, Fricker, Carter and Tulaczyk2016; Carter and others, Reference Carter, Fricker and Siegfried2017; Siegert, Reference Siegert2018). This is also emphasized by different observations with different systems of the same regions, leading to contrasting interpretations (e.g. Bell and others, Reference Bell, Studinger, Shuman, Fahnestock and Joughin2007; Humbert and others, Reference Humbert, Steinhage, Helm, Beyer and Kleiner2018). Recent advances in the analysis of subglacial hydrology from radar sounding data has focused on subglacial water systems beyond Antarctic subglacial lakes (e.g. Young and others, Reference Young, Schroeder, Blankenship, Kempf and Quartini2016). This includes utilizing bed-echo strength and character to investigate water body geometry and dynamic configuration, catchment-scale drainage systems and grounding zones (Schroeder and others, Reference Schroeder, Blankenship and Young2013, Reference Schroeder, Blankenship, Raney and Grima2014a; Ashmore and Bingham, Reference Ashmore and Bingham2014; Christianson and others, Reference Christianson2016). In Greenland, a range of studies has investigated the distribution of subglacial water, including lakes, topographically controlled seasonal storage and gradients in water near the onset of fast flow (Oswald and Gogineni, Reference Oswald and Gogineni2008; Palmer and others, Reference Palmer2013; Chu and others, Reference Chu2016, Reference Chu, Schroeder, Seroussi, Creyts and Bell2018b; Jordan and others, Reference Jordan2018b; Oswald and others, Reference Oswald, Rezvanbehbahani and Stearns2018; Bowling and others, Reference Bowling, Livingstone, Sole and Chu2019). Hypersaline lakes have also been identified beneath Devon Ice Cap in Arctic Canada (Rutishauser and others, Reference Rutishauser2018).

Radio-wave attenuation

Laboratory analyses of radio-wave absorption in ice, as well as radar sounding data from the field, have revealed that while relatively homogeneous ice is a very low-loss medium for radio-waves at VHF frequencies, there is a loss of returned power englacially due to dielectric absorption of radiowaves in ice. Dielectric absorption is proportional to the electrical conductivity of the ice, which is related to ice temperature and the presence of impurities (Glen and Paren, Reference Glen and Paren1975; Johari and Charette, Reference Johari and Charette1975; Moore and Fujita, Reference Moore and Fujita1993; Stillman and others, Reference Stillman, MacGregor and Grimm2013; Pettinelli and others, Reference Pettinelli2015). Without sufficiently distinct basal echo signals (e.g. relative changes that delineate sharp boundaries, such as ice stream shear margins) or sufficiently effective corrections, uncertainty in englacial attenuation can obfuscate the interpretation of basal reflectivities (Matsuoka, Reference Matsuoka2011; Siegert and others, Reference Siegert2016; Schroeder and others, Reference Schroeder, Grima and Blankenship2016a).

Empirical methods for estimating englacial attenuation using bed echoes range from simple linear fitting to adaptive or model-informed fitting (Jacobel and others, Reference Jacobel, Welch, Osterhouse, Pettersson and MacGregor2009; Wolovick and others, Reference Wolovick, Bell, Creyts and Frearson2013; Ashmore and others, Reference Ashmore, Bingham, Hindmarsh, Corr and Joughin2014; Jordan and others, Reference Jordan2016; Schroeder and others, Reference Schroeder, Seroussi, Chu and Young2016c). Englacial layers themselves have also been used to derive attenuation (Matsuoka and others, Reference Matsuoka, Morse and Raymond2010; MacGregor and others, Reference MacGregor2015b). These approaches can also be intercompared or combined (e.g. Hills and others, Reference Hills, Christianson and Holshuh2020; Jeofry and others, 2020). Additionally, investigating attenuation with variable offset can constrain englacial, attenuation, though there is a limit on the maximum offset achievable with commercial GPR systems (Holschuhand others, Reference Holschuh, Christianson, Anandakrishnan, Alley and Jacobel2016). These empirical attenuation values can either be used to correct losses to enable reflectivity interpretation or interpreted themselves as a proxy for englacial temperature.

In addition to applying empirical methods that estimate and correct for attenuation, attenuation rate can also be modeled (Matsuoka and others, Reference Matsuoka, MacGregor and Pattyn2012). This approach can be used when correcting attenuation effects or constraining the bed conditions using layer power (MacGregor and others, Reference MacGregor2015b; Chu and others, Reference Chu, Schroeder, Seroussi, Creyts and Bell2018b). Modeled attenuation can be compared to observations to constrain englacial temperature, parameterize basal conditions to match surface velocities or to quantify englacial water from persistent firn aquifers (Forster and others, Reference Forster2014; Schroeder and others, Reference Schroeder, Seroussi, Chu and Young2016c; Chu and others, Reference Chu, Schroeder and Siegfried2018a; Holschuh and others, Reference Holschuh, Lilien and Christianson2019).

Englacial structure

The study of radar-derived englacial properties dates back almost to the beginning of radioglaciology (e.g. Harrison, Reference Harrison1973; Gudmandsen, Reference Gudmandsen1975; Paren and Robin, Reference Paren and Robin1975). The englacial information that radar data contain has the potential to provide insights into ice-flow processes as well as climatic forcings. The layers have thus been widely used with models for ice-core site selection, stratigraphic control and inferring accumulation histories (see below) (e.g. Jacobel and Hodge, Reference Jacobel and Hodge1995; Cavitte and others, Reference Cavitte2016; Parrenin and others, Reference Parrenin2017). In recent years, the radioglaciological community has seen an increase in the retrieval of such information from radar data although barriers remain to the widespread usage of englacial stratigraphy. This is due to the fact that a substantial amount of manual work is generally needed to convert the stratigraphic information into, for example, dated isochrone surfaces that can readily be used by ice-flow models. Attempts to overcome this obstacle include methodologies focusing on quantifying the slope of the stratigraphy and extracting information from slopes instead (Panton and Karlsson, Reference Panton and Karlsson2015; Holschuh and others, Reference Holschuh, Parizek, Alley and Anandakrishnan2017; Castelletti and others, Reference Castelletti, Schroeder, Mantelli and Hilger2019). Studies focusing on the reorganization of ice flow often avoid tracing isochrones and take a qualitative approach. For example, imprints of shear margin migration or change in flow direction are typically identified based on the amount of stratigraphic disruption. Examples include studies showing changes in ice-flow structure or folded stratigraphy in Greenland and entrained debris in a glacier in Patriot Hills, West Antarctica (Catania and others, Reference Catania, Conway, Raymond and Scambos2006; Martín and others, Reference Martín, Gudmundsson, Pritchard and Gagliardini2009; Dahl-Jensen and others, Reference Dahl-Jensen2013; Bell and others, Reference Bell2014; Bingham and others, Reference Bingham2015; Kingslake and others, Reference Kingslake, Martín, Arthern, Corr and King2016; Winter and others, Reference Winter2019; Ross and Siegert, Reference Ross and Siegert2020). Advances in processing radargrams to extract ice-sheet structure make it possible to interpret these features in regions of complex flow (Elsworth and others, Reference Elsworth, Schroeder and Siegfried2020).

The tracing of englacial isochrones in the radar data acquired over Greenland between 1993 and 2013 by the University of Kansas Center for Remote Sensing of Ice Sheets and OIBis a vital step forward in the efforts to make englacial stratigraphic information readily available (Gogineni and others, Reference Gogineni, Chuah, Allen, Jezek and Moore1998, Reference Gogineni2001; MacGregor and others, Reference MacGregor2015a; Arnold and others, Reference Arnold2018). The resulting data archive has increased the availability of traced isochrones by orders of magnitude. Derived results include evidence of Holocene deceleration of the Greenland ice sheet, and improved constraints on its internal temperature (MacGregor and others, Reference MacGregor2015b, Reference MacGregor2016). In Antarctica, no such large-scale synthesis has been undertaken, but the SCAR AntArchitecture project has the potential to address this critical gap. Several studies have successfully linked isochrones between deep ice-core sites: the interior Antarctic ice-core sites are now linked from Dome Concordia through Vostok to Dome Argus, and Dome Fuji has been linked to the EPICA-DML (European Project for Ice Coring in Antarctica Dronning Maud Land) ice-core site (Cavitte and others, Reference Cavitte2016; Winter and others, Reference Winter, Steinhage, Creyts, Kleiner and Eisen2019). These efforts will play a key role in identifying optimal drill sites for the Oldest Ice (ice older than 1.5 million years, Fischer and others, Reference Fischer2013).

Other important derived products from traced isochrones are the past accumulation rates and patterns (e.g. Eisen, Reference Eisen2008). Recent work in this area has been carried out on time-scales ranging from annual to centennial to millennial (Eisen and others, Reference Eisen2008; Medley and others, Reference Medley2014; Nielsen and others, Reference Nielsen, Karlsson and Hvidberg2015; Grima and others, Reference Grima2016; Karlsson and others, Reference Karlsson2016; Koenig and others, Reference Koenig2016; Koutnik and others, Reference Koutnik2016; MacGregor and others, Reference MacGregor2016; Lewis and others, Reference Lewis2017; Cavitte and others, Reference Cavitte2018; Karlsson and others, Reference Karlsson2020; Montgomery and others, Reference Montgomery, Koenig, Lenaerts and Kuipers Munneke2020). Efforts to automate layer tracing continue, which include methodologies that use seed points to initiate semi-automatic tracing routines as well as fully automatic schemes. In parallel, the extra-terrestrial radar community has been working toward automatically extracting layer information from the Martian orbital radar sounders (Ferro and Bruzzone, Reference Ferro and Bruzzone2012; Onana and others, Reference Onana, Koenig, Ruth, Studinger and Harbeck2015; Xiong and others, Reference Xiong, Muller and Carretero2018; Xiong and Muller, Reference Xiong and Muller2019). Delf and others (Reference Delf, Schroeder, Bingham and Giannopoulos2020) (this issue) present some strategies for assessing automated algorithms inherited from both terrestrial and planetary work.

Interpretation

The history and dynamics of glaciers and ice sheets are written into radar-sensitive properties of these ice masses. Interpretation of radar data may be qualitative or quantitative, with the latter facilitated by process-based models in particular. In its most common form, however, interaction between radioglaciology and models is often limited and one-directional: radio-echo sounding of ice depth furnishes the basal boundary condition for ice-flow models (e.g. Fretwell and others, Reference Fretwell2013). While gaps in our knowledge of basal topography have spurred model development, radar studies have produced a trove of other data and discoveries, including, for example, evidence of retreat, past flow, basal accretion, firn-aquifers and ice-shelf conduits, that remain under-exploited by theory and models (Conway and others, Reference Conway, Hall, Denton, Gades and Waddington1999; Siegert and others, Reference Siegert2004; Bingham and Siegert, Reference Bingham and Siegert2007; Bell and others, Reference Bell2011; Morlighem and others, Reference Morlighem2011; Forster and others, Reference Forster2014; Bons and others, Reference Bons2016; Drews and others, Reference Drews2017; Jordan and others, Reference Jordan2018a; Leysinger Vieli and others, Reference Leysinger Vieli, Martín and Hindmarsh2018; Holschuh and others, Reference Holschuh, Lilien and Christianson2019; Langhammer and others, Reference Langhammer, Grab, Bauder and Maurer2019).

Theoretical work has established relationships between the architecture of internal layers and ice-sheet accumulation, topography, rheology and dynamics (e.g. Nereson and Waddington, Reference Nereson and Waddington2002; Siegert, Reference Siegert2003; Hindmarsh and others, Reference Hindmarsh, Leysinger Vieli, Raymond and Gudmundsson2006; Parrenin and others, Reference Parrenin, Hindmarsh and Rémy2006; Martín and others, Reference Martín, Gudmundsson, Pritchard and Gagliardini2009; Felix and King, Reference Felix and King2011). Internal layers have been integrated with models to determine ice rheology and to understand flow history, including migration of ice streams, divides and domes (e.g. Nereson and Raymond, Reference Nereson and Raymond2001; Ng and Conway, Reference Ng and Conway2004; Catania and others, Reference Catania, Conway, Raymond and Scambos2006; Gillet-Chaulet and others, Reference Gillet-Chaulet, Hindmarsh, Corr, King and Jenkins2011; Pettit and others, Reference Pettit2011; Drews and others, Reference Drews2015; MacGregor and others, Reference MacGregor2016). The discovery of deep internal structures that do not conform to the bed has prompted new model exploration of englacial and basal processes including interpretation of their radar scattering character, with implications for interpreting ice-sheet dynamics and the climate archive (e.g. Bell and others, Reference Bell2011, Reference Bell2014; Dahl-Jensen and others, Reference Dahl-Jensen2013; Wolovick and others, Reference Wolovick, Bell, Creyts and Frearson2013; Wrona and others, Reference Wrona, M, S and D2017; Kjær and others, Reference Kjær2018; Goldberg and others, Reference Goldberg2020).

In addition to englacial layers, radar sounding data have been used to detect channels under ice shelves that have also been the focus of a suite of model investigations (e.g. Jenkins, Reference Jenkins2011; Le Brocq and others, Reference Le Brocq2013; Sergienko, Reference Sergienko2013; Drews, Reference Drews2015; Alley and others, Reference Alley, Scambos, Siegfried and Fricker2016). Theory and observation are yielding new insight into ice–ocean interactions and real-time geomorphic processes in grounding zones, the influence of topography on channel position and formation, and the uncertain relationship between channels and ice-shelf stability (e.g. Gladish and others, Reference Gladish, Holland, Holland and Price2012; Greenbaum and others, Reference Greenbaum2015; Khazendar and others, Reference Khazendar2016; Drews and others, Reference Drews2017; Gourmelen and others, Reference Gourmelen2017; Jeofry and others, Reference Jeofry2018).

With so much radioglaciological data, the advent of resources such as ice-sheet-wide radiostratigraphic archives should help operationalize data–model integration (MacGregor and others, Reference MacGregor2015a). But how are such archives best exploited? Inverse methods present a natural approach, although the persistent problem of non-uniqueness demands care in defining the problem, choosing the tools and incorporating constraints (e.g. Waddington and others, Reference Waddington, Neumann, Koutnik, Marshall and Morse2007; Eisen, Reference Eisen2008; Gudmundsson, Reference Gudmundsson, VP, P and UK2011; Koutnik and Waddington, Reference Koutnik and Waddington2012; Nielsen and others, Reference Nielsen, Karlsson and Hvidberg2015; Koutnik and others, Reference Koutnik2016). Computational costs of large-scale models further demand attention to efficiency, for example, by the use of adjoint methods (e.g. Hascoët and Morlighem, Reference Hascoët and Morlighem2018). Consideration should also be given to the information content of different variables, including those sensitive to basal processes, as well as to the limitations of rendering 3-D effects in 2-D data (Leysinger-Vieli and others, Reference Leysinger-Vieli, Hindmarsh and Siegert2007; Holschuh and others, Reference Holschuh, Parizek, Alley and Anandakrishnan2017; Young and others, Reference Young2018). While we must devise modeling strategies to make best use of the data, this is far from a case of models simply lagging observations. Challenges remain in combining disparate datasets, conditioning data for comparison with modeling and utilizing radiometric, interferometric and polarimetric information in modeling (e.g. Hindmarsh and others, Reference Hindmarsh, Leysinger Vieli and Parrenin2009; Schroeder and others, Reference Schroeder, Seroussi, Chu and Young2016c; Castelletti and others, Reference Castelletti2017, Reference Castelletti, Schroeder, Mantelli and Hilger2019; Winter and others, Reference Winter2017, Reference Winter, Steinhage, Creyts, Kleiner and Eisen2019; Chu and others, Reference Chu, Schroeder, Seroussi, Creyts and Bell2018b; Jordan and others, Reference Jordan, Schroeder, Castelletti, Li and Dall2019). Finally, data–model interaction is a two-way street: testable hypotheses produced by theory and models may suggest new observational targets or provide new reasons to tap the rich radioglaciological archive (e.g. Raymond, Reference Raymond1983; Arthern and others, Reference Arthern, Hindmarsh and Williams2015).

Planetary radioglaciology

The bulk of extra-terrestrial ice-sounding data stems from the planet Mars, specifically from the two orbital radar sounders: MARSIS (Mars Advanced Radar for Subsurface and Ionospheric Sounding) onboard the European Space Agency's Mars Express, and SHARAD onboard the Mars Reconnaissance Orbiter launched by NASA (National Aeronautics and Space Administration, USA) (SHAllow RADar, Seu and others, Reference Seu2007; Jordan and others, Reference Jordan2009). The difference in frequency between the two sounders allowed for different penetration depths and thereby different insights into the planet's ice bodies (MARSIS operated at 1.3–5.5 MHz in its subsurface sounding mode while SHARAD used 15–25 MHz). Although both instruments are now inactive, analysis of the data is ongoing and continues to contribute to our understanding of water ice on Mars. The results from the radar sounders documented the high water content of the Martian water-ice reservoirs (e.g. Grima and others, Reference Grima2009). These have now been supplemented by more detailed studies of the composition of the polar ice bodies, the immediate subsurface of the north pole, and the mid-latitude water ice reservoirs (Guallini and others, Reference Guallini2018; Mirino and others, Reference Mirino, Frigeri, Orosei, Rossi and Cantini2018; Petersen and others, Reference Petersen, Holt and Levy2018; Putzig and others, Reference Putzig2018; Nerozzi and Holt, Reference Nerozzi and Holt2019). In addition, the radar sounding has confirmed areas on the planet also contains significant volumes of buried water ice (Bramson and others, Reference Bramson2015; Stuurman and others, Reference Stuurman2016). One of the most prominent findings is the discovery of a signal that shares similarities with those of a liquid water body (Orosei and others, Reference Orosei2018). In the MARSIS data, this proposed ‘subglacial lake’ has characteristically bright and specular reflections and was found 1.8 km below the South Polar Layered Deposits. The salt content and/or heat flux necessary to form and sustain such a lake is, however, still debated (Sori and Bramson, Reference Sori and Bramson2019). In addition to these findings, the radar data have successfully been utilized to gain insights into the glaciological and climatological processes on the planet, including the deformational properties of Martian water ice, and the past climate history and accumulation patterns of both the North Polar Layered Deposits and the South Polar Layered Deposits (Karlsson and others, Reference Karlsson, Schmidt and Hvidberg2015; Parsons and Holt, Reference Parsons and Holt2016; Smith and others, Reference Smith, Putzig, Holt and Phillips2016; Whitten and others, Reference Whitten, Campbell and Morgan2017; Nerozzi and Holt, Reference Nerozzi and Holt2018; Lalich and others, Reference Lalich, Holt and Smith2019; Schmidt and others, Reference Schmidt, Hvidberg, Kim and Karlsson2019). The radar data have also been used to reconcile observations from visual imagery with the radar-imaged englacial stratigraphy (Christian and others, Reference Christian, Holt, Byrne and Fishbaugh2013; Lalich and Holt, Reference Lalich and Holt2017).

Moving further afield, two radar sounders are now under preparation to probe the subsurface of the Jovian system. Two instruments have been selected for upcoming missions to Ganymede and Europa: the 9 MHz frequency Radar for Icy Moons Exploration (RIME) instrument on board the European Space Agency's Jupiter Icy Moons Explorer (JUICE) and the 9 and 60 MHz frequency Radar for Europa Assessment and Sounding: Ocean to Near-surface (REASON) instrument on board NASA's Europa Clipper (Bruzzone and others, Reference Bruzzone2013; Pappalardo and others, Reference Pappalardo2015; Lorente and others, Reference Lorente2017; Blankenship and others, Reference Blankenship2018). These sounders are designed to probe the moons' interiors and have penetration depths which are functions of surface roughness, volume scattering, ice-shell thermal structure, chemistry and the character of the ice/water interface (Moore, Reference Moore2000; McKinnon, Reference McKinnon2005; Blankenship and others, Reference Blankenship, Young, Moore and Moore2009; Bruzzone and others, Reference Bruzzone2011; Schmidt and others, Reference Schmidt, Blankenship, Patterson and Schenk2011; Berquin and others, Reference Berquin, Kofman, Herique, Alberti and Beck2013; Grima and others, Reference Grima, Schroeder, Blankenship and Young2014b; Pettinelli and others, Reference Pettinelli2015; Di Paolo and others, Reference Di Paolo2016; Grima and others, Reference Grima2016; Aglyamov and others, Reference Aglyamov, Schroeder and Vance2017; Heggy and others, Reference Heggy, Scabbia, Bruzzone and Pappalardo2017; Kalousová and others, Reference Kalousová, Schroeder and Soderlund2017; Campbell and others, Reference Campbell, Schroeder and Whitten2018; Gerekos and others, Reference Gerekos2018; Michaelides and Schroeder, Reference Michaelides and Schroeder2019; Culha and others, Reference Culha, Schroeder, Jordan and Haynes2020). The addition of a dual-channel VHF band on REASON also allows for characterization of the European ionosphere, altimetric investigation of Europa's shell and tides, and dual-frequency or interferometric clutter discrimination (Grima and others, Reference Grima, Blankenship and Schroeder2015; Carrer and Bruzzone, Reference Carrer and Bruzzone2017; Castelletti and others, Reference Castelletti2017; Haynes and others, Reference Haynes, Chapin, Moussessian and Madsen2018a; Steinbrügge and others, Reference Steinbrügge2018; Scanlan and others, Reference Scanlan2019). Finally, the ability of both instruments to record strong Jovian emissions raises the possibility of using those emissions to probe the ice shell using passive radio sounding (Romero-Wolf and others, Reference Romero-Wolf2015; Schroeder and others, Reference Schroeder2016b; Peters and others, Reference Peters, Schroeder, Castelletti, Haynes and Romero-Wolf2018).

In addition to Mars and the icy Jovian Moons, radar sounding is also being deployed to investigate ice on other planetary bodies. For example, NASA's Lunar Reconnaissance Orbiter was equipped with a radar sounder in the gigahertz frequency range in order to search for water ice on Earth's moon (Nozette and others, Reference Nozette2010). The data reveal the existence of large deposits of relatively clean ice in the polar regions (Spudis and others, Reference Spudis2013). Unfortunately, measurements temporarily discontinued after an instrument failure in 2011, but have resumed in a bi-static configuration (Patterson and others, Reference Patterson2017). Additionally ESA's Rosetta mission included the bistatic CONSERT experiment (COmet Nucleus Sounding Experiment by Radiowave Transmission), which performed the first tomographic imaging of the interior of a comet (Glassmeier and others, Reference Glassmeier, Boehnhardt, Koschny, Kührt and Richter2007; Kofman and others, Reference Kofman2015).

Conclusions

More than 50 years after the first collection of radioglaciological observations, radar-sounding data are being acquired over ice sheets, glaciers, ice shelves and ice shells across the solar system at unprecedented scales and rates. Terrestrially, this ever growing data volume, along with re-mastery of archival data, is enabling multi-temporal investigations of subglacial and englacial processes at the spatial and temporal scales relevant to ice-sheet and sea-level change. Recent advances in radar-sounder systems now allow for the acquisition of multi-frequency, multi-offset, polarimetric and interferometric data that can provide rich new information about conditions within and beneath the ice. At the same time, advances in data analysis, interpretation and modeling have paved the way for using that rich new information to investigate the fundamental physical processes that control the past, present and future evolution of ice masses. Additionally, recent progress in sensor and platform technologies is making it possible to move from mapping to monitoring approaches in radar-sounding surveys by exploiting low-cost radar-sounder sensor networks, autonomous rovers and drones, or even orbital sounding. Finally, planetary ice/water systems are only growing in their appeal and feasibility as targets of radio-echo sounding. After half a century, radioglaciology may just be entering its golden age.

Acknowledgments

We would like to thank Ala Khazendar for serving as the Scientific Editor and Hester Jiskoot for serving as Associate Chief Editor for this manuscript. We would also like to thank Joe MacGregor and two anonymous reviewers for their thoughtful feedback on the manuscript.

References

Aglyamov, Y, Schroeder, DM and Vance, SD (2017) Bright prospects for radar detection of Europa's ocean. Icarus 281, 334337. doi: 10.1016/j.icarus.2016.08.014.CrossRefGoogle Scholar
Allen, C (2008) A brief history of radio-echo sounding of ice. Earthzine. Available at: https://earthzine.org/a-brief-history-of-radio-echo-sounding-of-ice-2/.Google Scholar
Alley, KE, Scambos, TA, Siegfried, MR and Fricker, HA (2016) Impacts of warm water on Antarctic ice shelf stability through basal channel formation. Nature Geoscience 9(4), 290294. doi: 10.1038/ngeo2675.CrossRefGoogle Scholar
Arcone, SAand 5 others (2016) Ground-penetrating radar profiles of the McMurdo Shear Zone, Antarctica, acquired with an unmanned rover: interpretation of crevasses, fractures, and folds within firn and marine ice GPR profiles of the McMurdo shear zone. Geophysics 81(1), WA21WA34. doi: 10.1190/geo2015-0132.1.CrossRefGoogle Scholar
Arnold, Eand 9 others (2018) HF/VHF radar sounding of ice from manned and unmanned airborne platforms. Geosciences 8(5), 182. doi: 10.3390/geosciences8050182.CrossRefGoogle Scholar
Arnold, E, Leuschen, C, Paden, J, Hale, R and Keshmiri, S (2020) CReSIS airborne radars and platforms for ice and snow sounding. Annals of Glaciology 61(81), 5867. doi: 10.1017/aog.2019.37.Google Scholar
Arthern, RJ, Hindmarsh, RC and Williams, CR (2015) Flow speed within the Antarctic ice sheet and its controls inferred from satellite observations. Journal of Geophysical Research: Earth Surface 120(7), 11711188. doi: 10.1002/2014JF003239.Google Scholar
Ashmore, DW and Bingham, RG (2014) Antarctic subglacial hydrology: current knowledge and future challenges. Antarctic Science 26(6), 758773. doi: 10.1017/S0954102014000546.CrossRefGoogle Scholar
Ashmore, DW, Bingham, RG, Hindmarsh, RC, Corr, HF and Joughin, IR (2014) The relationship between sticky spots and radar reflectivity beneath an active West Antarctic ice stream. Annals of Glaciology 55(67), 2938. doi: 10.3189/2014AoG67A052.CrossRefGoogle Scholar
Bailey, J, Evans, S and Robin, GdQ (1964) Radio echo sounding of polar ice sheets. Nature 204(4957), 420421. doi: 10.1038/204420a0.CrossRefGoogle Scholar
Bamber, Jand 10 others (2013) A new bed elevation dataset for Greenland. The Cryosphere 7, 499510. doi: 10.5194/tc-7-499-2013.CrossRefGoogle Scholar
Bartlett, OTand 5 others (2020) Geospatial simulations of airborne ice-penetrating radar surveying reveal elevation under-measurement bias for ice sheet bed topography. Annals of Glaciology 61(81), 4657. doi: 10.1017/aog.2020.35.CrossRefGoogle Scholar
Behrendt, JCand 6 others (1994) CASERTZ aeromagnetic data reveal late Cenozoic flood basalts (?) in the West Antarctic rift system. Geology 22(6), 527530. doi: 10.1130/0091-7613(1994)022<0527:CADRLC>2.3.CO;2.2.3.CO;2>CrossRefGoogle Scholar
Bell, REand 6 others (1998) Influence of subglacial geology on the onset of a West Antarctic ice stream from aerogeophysical observations. Nature 394(6688), 5862. doi: 10.1038/27883.CrossRefGoogle Scholar
Bell, REand 9 others (2011) Widespread persistent thickening of the east Antarctic ice sheet by freezing from the base. Science (New York, N.Y.) 331(6024), 15921595. doi: 10.1126/science.1200109.CrossRefGoogle ScholarPubMed
Bell, REand 8 others (2014) Deformation, warming and softening of Greenland's ice by refreezing meltwater. Nature Geoscience 7, 497502. doi: 10.1038/ngeo2179.CrossRefGoogle Scholar
Bell, RE, Studinger, M, Shuman, CA, Fahnestock, MA and Joughin, I (2007) Large subglacial lakes in East Antarctica at the onset of fast-flowing ice streams. Nature 445(7130), 904907. doi: doi.org/10.1038/nature05554.CrossRefGoogle ScholarPubMed
Berger, Vand 5 others (2018) Automated tracking of 2D and 3D ice radar imagery using viterbi and TRW-S. In IGARSS 2018–2018 IEEE International Geoscience and Remote Sensing Symposium. IEEE, pp. 41624165. doi: 10.1109/IGARSS.2018.8519411.CrossRefGoogle Scholar
Berquin, Y, Kofman, W, Herique, A, Alberti, G and Beck, P (2013) A study on Ganymede's surface topography: perspectives for radar sounding. Planetary and Space Science 77, 4044. doi: 10.1016/j.pss.2012.07.004.CrossRefGoogle Scholar
Bingham, RGand 9 others (2015) Ice-flow structure and ice dynamic changes in the Weddell Sea sector of West Antarctica from radar-imaged internal layering. Journal of Geophysical Research: Earth Surface 120(4), 655670. doi: 10.1002/2014JF003291.Google Scholar
Bingham, RGand 9 others (2017) Diverse landscapes beneath Pine Island Glacier influence ice flow. Nature Communications 8(1), 1618. doi: 10.1038/s41467-017-01597-y.CrossRefGoogle ScholarPubMed
Bingham, RG and Siegert, MJ (2007) Radio-echo sounding over polar ice masses. Journal of Environmental and Engineering Geophysics 12(1), 4762. doi: 10.2113/JEEG12.1.47.CrossRefGoogle Scholar
Bingham, RG and Siegert, MJ (2009) Quantifying subglacial bed roughness in Antarctica: implications for ice-sheet dynamics and history. Quaternary Science Reviews 28(3–4), 223236. doi: 10.1016/j.quascirev.2008.10.014.CrossRefGoogle Scholar
Björnsson, Hand 6 others (1996) The thermal regime of sub-polar glaciers mapped by multi-frequency radio-echo sounding. Journal of Glaciology 42(140), 2332. doi: 10.3189/S0022143000030495.CrossRefGoogle Scholar
Björnsson, H and Pálsson, F (2020) Radio-echo soundings on Icelandic temperate glaciers: history of techniques and findings. Annals of Glaciology 61(81), 2534. doi: 10.1017/aog.2020.10.Google Scholar
Blankenship, DDand 5 others (1993) Active volcanism beneath the West Antarctic ice sheet and implications for ice-sheet stability. Nature 361(6412), 526529. doi: 10.1038/361526a0.CrossRefGoogle Scholar
Blankenship, DDand 5 others (2018) Reason for Europa. In 42nd COSPAR Scientific Assembly, Vol. 42, Pasadena, CA.Google Scholar
Blankenship, DD, Young, DA, Moore, WB and Moore, JC (2009) Radar sounding of Europa's subsurface properties and processes: the view from Earth. In Europa. Tucson: University of Arizona Press, pp. 631654.Google Scholar
Bons, PDand 10 others (2016) Converging flow and anisotropy cause large-scale folding in Greenland's ice sheet. Nature Communications 7. doi: 10.1038/ncomms11427.CrossRefGoogle ScholarPubMed
Booth, AD, Clark, R and Murray, T (2010) Semblance response to a ground-penetrating radar wavelet and resulting errors in velocity analysis. Near Surface Geophysics 8(3), 235246. doi: 10.3997/1873-0604.2010008.CrossRefGoogle Scholar
Bowling, J, Livingstone, S, Sole, A and Chu, W (2019) Distribution and dynamics of Greenland subglacial lakes. Nature Communications 10(1), 111. doi: 10.1038/s41467-019-10821-w.CrossRefGoogle ScholarPubMed
Bramson, AMand 6 others (2015) Widespread excess ice in Arcadia Planitia, Mars. Geophysical Research Letters 42(16), 65666574. doi: 10.1002/2015GL064844.CrossRefGoogle Scholar
Bruzzone, Land 5 others (2011) Subsurface radar sounding of the Jovian moon Ganymede. Proceedings of the IEEE 99(5), 837857. doi: 10.1109/JPROC.2011.2108990.CrossRefGoogle Scholar
Bruzzone, Land 9 others (2013) RIME: radar for icy moon exploration. In 2013 IEEE International Geoscience and Remote Sensing Symposium-IGARSS. IEEE, Melbourne, Australia, pp. 39073910.CrossRefGoogle Scholar
Buchardt, SL and Dahl-Jensen, D (2007) Estimating the basal melt rate at NorthGRIP using a Monte Carlo technique. Annals of Glaciology 45, 137142. doi: 10.3189/172756407782282435.CrossRefGoogle Scholar
Campbell, BA, Schroeder, DM and Whitten, JL (2018) Mars radar clutter and surface roughness characteristics from MARSIS data. Icarus 299, 2230. doi: 10.1016/j.icarus.2017.07.011.CrossRefGoogle Scholar
Carrer, L and Bruzzone, L (2016) Automatic enhancement and detection of layering in radar sounder data based on a local scale hidden Markov model and the Viterbi algorithm. IEEE Transactions on Geoscience and Remote Sensing 55(2), 962977. doi: 10.1109/TGRS.2016.2616949.CrossRefGoogle Scholar
Carrer, L and Bruzzone, L (2017) Solving for ambiguities in radar geophysical exploration of planetary bodies by mimicking bats echolocation. Nature Communications 8(1), 112. doi: 10.1038/s41467-017-02334-1.CrossRefGoogle ScholarPubMed
Carrer, L, Gerekos, C and Bruzzone, L (2018) Distributed radar sounder system: a novel approach to across-track resolution enhancement and clutter reduction. In IGARSS 2018-2018 IEEE International Geoscience and Remote Sensing Symposium. IEEE, Valencia, Spain, pp. 67656768. doi: 10.1109/IGARSS.2018.8519462.CrossRefGoogle Scholar
Carter, SPand 5 others (2007) Radar-based subglacial lake classification in Antarctica. Geochemistry, Geophysics, Geosystems 8(3). doi: 10.1029/2006GC001408.CrossRefGoogle Scholar
Carter, SP, Fricker, HA and Siegfried, MR (2017) Antarctic subglacial lakes drain through sediment-floored canals: theory and model testing on real and idealized domains. The Cryosphere 11(1), 381. doi: 10.5194/tc-11-381-2017.CrossRefGoogle Scholar
Castelletti, Dand 9 others (2017) An interferometric approach to cross-track clutter detection in two-channel VHF radar sounders. IEEE Transactions on Geoscience and Remote Sensing 55(11), 61286140. doi: 10.1109/TGRS.2017.2721433.CrossRefGoogle Scholar
Castelletti, D and Schroeder, D (2017) Estimating englacial vertical velocity from airborne radar sounding data. In AGU Fall Meeting Abstracts, San Francisco, CA.Google Scholar
Castelletti, D, Schroeder, DM, Mantelli, E and Hilger, A (2019) Layer optimized SAR processing and slope estimation in radar sounder data. Journal of Glaciology 65(254), 983988. doi: 10.1017/jog.2019.72.CrossRefGoogle Scholar
Catania, GA, Conway, H, Raymond, CF and Scambos, TA (2006) Evidence for floatation or near floatation in the mouth of Kamb Ice Stream, West Antarctica, prior to stagnation. Journal of Geophysical Research: Earth Surface 111, F01005. doi: 10.1029/2005JF000355.CrossRefGoogle Scholar
Cavitte, MGPand 7 others (2016) Deep radiostratigraphy of the East Antarctic plateau: connecting the Dome C and Vostok ice core sites. Journal of Glaciology 62(232), 323334. doi: 10.1017/jog.2016.11.CrossRefGoogle Scholar
Cavitte, MGand 7 others (2018) Accumulation patterns around Dome C, East Antarctica, in the last 73 kyr. The Cryosphere 12(4), 14011414. doi: 10.5194/tc-12-1401-2018.CrossRefGoogle Scholar
Christian, S, Holt, J, Byrne, S and Fishbaugh, K (2013) Integrating radar stratigraphy with high resolution visible stratigraphy of the North Polar Layered Deposits, Mars. Icarus 226(2), 12411251. doi: https://doi.org/10.1016/j.icarus.2013.07.003.CrossRefGoogle Scholar
Christianson, Kand 6 others (2016) Basal conditions at the grounding zone of Whillans Ice Stream, West Antarctica, from ice-penetrating radar. Journal of Geophysical Research: Earth Surface 121(11), 19541983. doi: 10.1002/2015JF003806.Google Scholar
Chu, Wand 5 others (2016) Extensive winter subglacial water storage beneath the Greenland ice sheet. Geophysical Research Letters 43(24), 12484. doi: 10.1002/2016GL071538.CrossRefGoogle Scholar
Chu, W, Schroeder, DM, Seroussi, H, Creyts, TT and Bell, RE (2018b) Complex basal thermal transition near the onset of Petermann Glacier, Greenland. Journal of Geophysical Research: Earth Surface 123(5), 985995. doi: 10.1029/2017JF004561.Google Scholar
Chu, W, Schroeder, D and Siegfried, M (2018a) Retrieval of englacial firn aquifer thickness from ice-penetrating radar sounding in Southeastern Greenland. Geophysical Research Letters 45(21), 1177011778. doi: 10.1029/2018GL079751.CrossRefGoogle Scholar
Church, Gand 5 others (2019) Detecting and characterising an englacial conduit network within a temperate Swiss glacier using active seismic, ground penetrating radar and borehole analysis. Annals of Glaciology 60(79), 193205. doi: 10.1017/aog.2019.19.CrossRefGoogle Scholar
Conway, H, Hall, BL, Denton, GH, Gades, AM and Waddington, ED (1999) Past and future grounding-line retreat of the West Antarctic ice sheet. Science (New York, N.Y.) 286(5438), 280283. doi: 10.1126/science.286.5438.280.CrossRefGoogle ScholarPubMed
Cooper, MAand 5 others (2019) Subglacial roughness of the Greenland ice sheet: relationship with contemporary ice velocity and geology. The Cryosphere 13(11), 30933115. doi: 10.5194/tc-2019-73.CrossRefGoogle Scholar
Corr, HF, Jenkins, A, Nicholls, KW and Doake, C (2002) Precise measurement of changes in ice-shelf thickness by phase-sensitive radar to determine basal melt rates. Geophysical Research Letters 29(8), 73. doi: 10.1029/2001GL014618.CrossRefGoogle Scholar
Crandall, DJ, Fox, GC and Paden, JD (2012) Layer-finding in radar echograms using probabilistic graphical models. In Proceedings of the 21st International Conference on Pattern Recognition (ICPR2012). IEEE, Sukuba Science City, Japan, pp. 15301533.Google Scholar
Culberg, R and Schroeder, DM (2019) Radar Scattering in Firn and its Implications for VHF/UHF Orbital Ice Sounding. In IGARSS 2019-2019 IEEE International Geoscience and Remote Sensing Symposium. IEEE, Yokohama, Japan, pp. 41374140. doi: 10.1109/IGARSS.2019.8898991.CrossRefGoogle Scholar
Culha, C, Schroeder, DM, Jordan, TM and Haynes, MS (2020) Assessing the detectability of Europa's eutectic zone using radar sounding. Icarus 339, 113578. doi: 10.1016/j.icarus.2019.113578.CrossRefGoogle Scholar
Dahl-Jensen, Dand 9 others (2013) Eemian interglacial reconstructed from a Greenland folded ice core. Nature 493(7433), 489494. doi: 10.1038/nature11789.Google Scholar
Dall, Jand 9 others (2010) ESA's polarimetric airborne radar ice sounder (POLARIS): design and first results. IET Radar, Sonar & Navigation 4(3), 488496. doi: 10.1049/iet-rsn.2009.0035.CrossRefGoogle Scholar
Dall, J, Corr, HF, Walker, N, Rommen, B and Lin, CC (2018) Sounding the Antarctic ice sheet from space: a feasibility study based on airborne P-band radar data. In IGARSS 2018-2018 IEEE International Geoscience and Remote Sensing Symposium. IEEE, Valencia, Spain, pp. 41424145. doi: 10.1109/IGARSS.2018.8518826.CrossRefGoogle Scholar
Davies, Dand 8 others (2018) How dynamic are ice-stream beds? The Cryosphere 12, 16151625. doi: 10.5194/tc-12-1615-2018.CrossRefGoogle Scholar
Dean, K, Naylor, S, Turchetti, S and Siegert, M (2008) Data in Antarctic science and politics. Social Studies of Science 38(4), 571604. doi: 10.1177/0306312708090693.CrossRefGoogle Scholar
Delf, R, Schroeder, DM, Bingham, RG and Giannopoulos, A (2020) A comparison of automated approaches to extracting englacial-layer geometry across ice sheets. Annals of Glaciology 61(81), 234241. doi: 10.1017/aog.2020.42.CrossRefGoogle Scholar
Di Paolo, Fand 9 others (2016) Radar signal penetration and horizons detection on Europa through numerical simulations. IEEE Journal of Selected Topics in Applied Earth Observations and Remote Sensing 10(1), 118129. doi: 10.1109/JSTARS.2016.2544103.CrossRefGoogle Scholar
Doake, C, Corr, H and Jenkins, A (2002) Polarization of radio waves transmitted through Antarctic ice shelves. Annals of Glaciology 34, 165170. doi: 10.3189/172756402781817572.CrossRefGoogle Scholar
Donini, E, Thakur, S, Bovolo, F and Bruzzone, L (2019) An automatic approach to map refreezing ice in radar sounder data. In Image and Signal Processing for Remote Sensing XXV, Vol. 11155. International Society for Optics and Photonics, Strasbourg, France, p. 111551B. doi: 10.1117/12.2533169.CrossRefGoogle Scholar
Dowdeswell, Jand 5 others (1986) Digital mapping of the Nordaustlandet ice caps from airborne geophysical investigations. Annals of Glaciology 8, 5158. doi: 10.3189/S0260305500001130.CrossRefGoogle Scholar
Dowdeswell, Jand 9 others (2002) Form and flow of the Academy of Sciences Ice Cap, Severnaya Zemlya, Russian High Arctic. Journal of Geophysical Research: Solid Earth 107(B4), 5-15-10. doi: 10.1029/2000JB000129.CrossRefGoogle Scholar
Dowdeswell, J, Benham, T, Gorman, M, Burgess, D and Sharp, M (2004) Form and flow of the Devon Island ice cap, Canadian Arctic. Journal of Geophysical Research: Earth Surface 109(F2), F02002. doi: 10.1029/2003JF000095.CrossRefGoogle Scholar
Dowdeswell, JA and Evans, S (2004) Investigations of the form and flow of ice sheets and glaciers using radio-echo sounding. Reports on Progress in Physics 67(10), 18211861. doi: 10.1088/0034-4885/67/10/R03.CrossRefGoogle Scholar
Drewry, DJ (1983) Antarctica: Glaciological and Geophysical Folio, Vol. 2. UK: University of Cambridge, Scott Polar Research Institute Cambridge.Google Scholar
Drews, Rand 5 others (2012) Potential mechanisms for anisotropy in ice-penetrating radar data. Journal of Glaciology 58(209), 613624. doi: 10.3189/2012JoG11J114.CrossRefGoogle Scholar
Drews, Rand 5 others (2015) Evolution of Derwael ice rise in Dronning Maud Land, Antarctica, over the last millennia. Journal of Geophysical Research: Earth Surface 120(3), 564579. doi: 10.1002/2014JF003246.Google Scholar
Drews, R (2015) Evolution of ice-shelf channels in Antarctic ice shelves. The Cryosphere 9(3), 11691181. doi: 10.5194/tc-9-1169-2015.CrossRefGoogle Scholar
Drews, Rand 9 others (2017) Actively evolving subglacial conduits and eskers initiate ice shelf channels at an Antarctic grounding line. Nature Communications 8(15228), 110. doi: 10.1038/ncomms15228.CrossRefGoogle ScholarPubMed
Durand, G, Gagliardini, O, Favier, L, Zwinger, T and Le Meur, E (2011) Impact of bedrock description on modeling ice sheet dynamics. Geophysical Research Letters 38(20). doi: 10.1029/2011GL048892.CrossRefGoogle Scholar
Eisen, Oand 13 others (2008) Ground-based measurements of spatial and temporal variability of snow accumulation in East Antarctica. Reviews of Geophysics 46, RG2001RG2008. doi: 10.1029/2006RG000218.CrossRefGoogle Scholar
Eisen, O (2008) Inference of velocity pattern from isochronous layers in firn, using an inverse method. Journal of Glaciology 54(187), 613630. doi: 10.3189/002214308786570818.CrossRefGoogle Scholar
Eisen, O, Hamann, I, Kipfstuhl, S, Steinhage, D and Wilhelms, F (2007) Direct evidence for continuous radar reflector originating from changes in crystal-orientation fabric. The Cryosphere 1, 110. doi: 10.5194/tc-1-1-2007.CrossRefGoogle Scholar
Eisen, O, Winter, A, Steinhage, D, Kleiner, T and Humbert, A (2020) Basal roughness of the East Antarctic Ice Sheet in relation to flow speed and basal thermal state. Annals of Glaciology 61(81), 162175. doi: 10.1017/aog.2020.47.Google Scholar
Elsworth, CW, Schroeder, DM and Siegfried, MR (2020) Interpreting englacial layer deformation in the presence of complex ice flow history with synthetic radargrams. Annals of Glaciology 61(81), 206213. doi: 10.1017/aog.2019.41.CrossRefGoogle Scholar
Fahnestock, M, Abdalati, W, Joughin, I, Brozena, J and Gogineni, P (2001) High geothermal heat flow, basal melt, and the origin of rapid ice flow in central Greenland. Science (New York, N.Y.) 294(5550), 23382342. doi: 10.1126/science.1065370.CrossRefGoogle ScholarPubMed
Falcini, FA, Rippin, DM, Krabbendam, M and Selby, KA (2018) Quantifying bed roughness beneath contemporary and palaeo-ice streams. Journal of Glaciology 64(247), 822834. doi: 10.1017/jog.2018.71.CrossRefGoogle Scholar
Felix, N and King, EC (2011) Kinematic waves in polar firn stratigraphy. Journal of Glaciology 57(206), 11191134. doi: 10.3189/002214311798843340.Google Scholar
Ferro, A (2019) Squinted SAR focusing for improving automatic radar sounder data analysis and enhancement. International Journal of Remote Sensing 40(12), 47624786. doi: 10.1080/01431161.2019.1573339.CrossRefGoogle Scholar
Ferro, A and Bruzzone, L (2012) Automatic extraction and analysis of ice layering in radar sounder data. IEEE Transactions on Geoscience and Remote Sensing 51(3), 16221634. doi: 10.1109/TGRS.2012.2206078.CrossRefGoogle Scholar
Fischer, Hand 5 others (2013) Where to find 1.5 million yr old ice for the IPICS Oldest-Ice ice core. Climate of the Past 9(6), 24892505. doi: 10.5194/cp-9-2489-2013.CrossRefGoogle Scholar
Forster, RRand 9 others (2014) Extensive liquid meltwater storage in firn within the Greenland ice sheet. Nature Geoscience 7(2), 9598. doi: 10.1038/ngeo2043.CrossRefGoogle Scholar
Franke, Sand 7 others (2020) Bed topography and subglacial landforms of the North East Greenland Ice Stream. Annals of Glaciology 61(81), 143153. doi: 10.1017/aog.2020.12.CrossRefGoogle Scholar
Freeman, A, Pi, X and Heggy, E (2017) Radar sounding through the Earth's ionosphere at 45 MHz. IEEE Transactions on Geoscience and Remote Sensing 55(10), 58335842. doi: 10.1109/TGRS.2017.2715838.CrossRefGoogle Scholar
Fretwell, Pand 5 others (2013) Bedmap2: improved ice bed, surface and thickness datasets for Antarctica. The Cryosphere 7(1), 375393. doi: 10.5194/tc-7-499-2013.CrossRefGoogle Scholar
Fujita, S, Matsuoka, K, Maeno, H and Furukawa, T (2003) Scattering of VHF radio waves from within an ice sheet containing the vertical-girdle-type ice fabric and anisotropic reflection boundaries. Annals of Glaciology 37, 305316. doi: 10.3189/172756403781815979.CrossRefGoogle Scholar
Fürst, JJand 9 others (2018) The ice-free topography of Svalbard. Geophysical Research Letters 45(21), 11760. doi: 10.1029/2018GL079734.CrossRefGoogle Scholar
Gärtner-Roer, Iand 5 others (2014) A database of worldwide glacier thickness observations. Global and Planetary Change 122, 330344. doi: 10.1016/j.gloplacha.2014.09.003.CrossRefGoogle Scholar
Gerekos, Cand 5 others (2018) A coherent multilayer simulator of radargrams acquired by radar sounder instruments. IEEE Transactions on Geoscience and Remote Sensing 56(12), 73887404. doi: 10.1109/TGRS.2018.2851020.CrossRefGoogle Scholar
Gillet-Chaulet, F, Hindmarsh, RC, Corr, HF, King, EC and Jenkins, A (2011) In situ quantification of ice rheology and direct measurement of the Raymond effect at Summit, Greenland using a phase-sensitive radar. Geophysical Research Letters 38(24), L24503. doi: 10.1029/2011GL049843.CrossRefGoogle Scholar
Gladish, CV, Holland, DM, Holland, PR and Price, SF (2012) Ice-shelf basal channels in a coupled ice/ocean model. Journal of Glaciology 58(212), 12271244. doi: 10.3189/2012JoG12J003.CrossRefGoogle Scholar
Glassmeier, KH, Boehnhardt, H, Koschny, D, Kührt, E and Richter, I (2007) The Rosetta mission: flying towards the origin of the solar system. Space Science Reviews 128(1-4), 121. doi: 10.1007/s11214-006-9140-8.CrossRefGoogle Scholar
Glen, J and Paren, J (1975) The electrical properties of snow and ice. Journal of Glaciology 15(73), 1538. doi: 10.3189/S0022143000034249.CrossRefGoogle Scholar
Goff, JA, Powell, EM, Young, DA and Blankenship, DD (2014) Instruments and methods conditional simulation of Thwaites Glacier (Antarctica) bed topography for flow models: incorporating inhomogeneous statistics and channelized morphology. Journal of Glaciology 60(22), 635646. doi: 10.3189/2014JoG13J200.CrossRefGoogle Scholar
Gogineni, SPand 9 others (2001) Coherent radar ice thickness measurements over the Greenland ice sheet. Journal of Geophysical Research: Atmospheres 106(D24), 3376133772. doi: 10.1029/2001JD900183.CrossRefGoogle Scholar
Gogineni, SPand 9 others (2014) Bed topography of Jakobshavn Isbræ, Greenland, and Byrd Glacier, Antarctica. Journal of Glaciology 60(233), 813833. doi: 10.3189/2014JoG14J129.CrossRefGoogle Scholar
Gogineni, SPand 6 others (2018) A CubeSat train for radar sounding and imaging of Antarctic ice sheet. In IGARSS 2018-2018 IEEE International Geoscience and Remote Sensing Symposium. IEEE, Valencia, Spain, pp. 41384141. doi: 10.1109/IGARSS.2018.8519162.CrossRefGoogle Scholar
Gogineni, S, Chuah, T, Allen, C, Jezek, K and Moore, RK (1998) An improved coherent radar sounder. Journal of Glaciology 44(148), 659669. doi: 10.3189/S0022143000002161.CrossRefGoogle Scholar
Goldberg, MLand 5 others (2020) Automated detection and characterization of Antarctic basal units using radar sounding data: demonstration in Institute Ice Stream, West Antarctica. Annals of Glaciology 61(81), 242248. doi: 10.1017/aog.2020.27.CrossRefGoogle Scholar
Gorman, MR and Siegert, MJ (1999) Penetration of Antarctic subglacial lakes by vhf electromagnetic pulses: information on the depth and electrical conductivity of basal water bodies. Journal of Geophysical Research: Solid Earth 104(B12), 2931129320. doi: 10.1029/1999JB900271.CrossRefGoogle Scholar
Gourmelen, Nand 9 others (2017) Channelized melting drives thinning under a rapidly melting Antarctic ice shelf. Geophysical Research Letters 44(19), 97969804. doi: 10.1002/2017GL074929.CrossRefGoogle Scholar
Greenbaum, Jand 9 others (2015) Ocean access to a cavity beneath Totten Glacier in East Antarctica. Nature Geoscience 8(4), 294298. doi: 10.1038/ngeo2388.CrossRefGoogle Scholar
Grima, Cand 7 others (2009) North polar deposits of Mars: extreme purity of the water ice. Geophysical Research Letters 36(3), L03203. doi:10.1029/2008GL036326.CrossRefGoogle Scholar
Grima, Cand 6 others (2016) Radar detection of the brine extent at McMurdo Ice Shelf, Antarctica, and its control by snow accumulation. Geophysical Research Letters 43(13), 70117018. doi: 10.1002/2016GL069524.CrossRefGoogle Scholar
Grima, C, Blankenship, DD and Schroeder, DM (2015) Radar signal propagation through the ionosphere of Europa. Planetary and Space Science 117, 421428. doi:10.1016/j.pss.2015.08.017.CrossRefGoogle Scholar
Grima, C, Blankenship, DD, Young, DA and Schroeder, DM (2014a) Surface slope control on firn density at Thwaites Glacier, West Antarctica: results from airborne radar sounding. Geophysical Research Letters 41(19), 67876794. doi: 10.1002/2014GL061635.CrossRefGoogle Scholar
Grima, C, Schroeder, DM, Blankenship, DD and Young, DA (2014b) Planetary landing-zone reconnaissance using ice-penetrating radar data: concept validation in Antarctica. Planetary and Space Science 103, 191204. doi: 10.1016/j.pss.2014.07.018.CrossRefGoogle Scholar
Guallini, Land 7 others (2018) Regional stratigraphy of the south polar layered deposits (Promethei Lingula, Mars): discontinuity-bounded units in images and radargrams. Icarus 308, 76107. doi: https://doi.org/10.1016/j.icarus.2017.08.030, mars Polar Science VI.CrossRefGoogle Scholar
Gudlaugsson, E, Humbert, A, Winsborrow, M and Andreassen, K (2013) Subglacial roughness of the former Barents Sea ice sheet. Journal of Geophysical Research: Earth Surface 118(4), 25462556. doi: 10.1002/2013JF002714.Google Scholar
Gudmandsen, P (1969) Airborne radio echo sounding of the Greenland ice sheet. The Geographical Journal 135(4), 548551. doi: 10.2307/1795099.CrossRefGoogle Scholar
Gudmandsen, P (1975) Layer echoes in polar ice sheets. Journal of Glaciology 15(73), 95101. doi: 10.3189/S0022143000034304.Google Scholar
Gudmundsson, GH (2011) Inverse methods in glaciology. In VP, Singh, P, Singh and UK, Haritashya (eds), Encyclopedia of Snow, Ice and Glaciers. Springer, pp. 653656.CrossRefGoogle Scholar
Hamran, SE and Aarholt, E (1993) Glacier study using wavenumber domain synthetic aperture radar. Radio Science 28(04), 559570. doi: 10.1029/92RS03022.CrossRefGoogle Scholar
Harrison, CH (1973) Radio echo sounding of horizontal layers in ice. Journal of Glaciology 12(66), 383397. doi: 10.3189/S0022143000031804.CrossRefGoogle Scholar
Hascoët, L and Morlighem, M (2018) Source-to-source adjoint algorithmic differentiation of an ice sheet model written in C. Optimization Methods and Software 33(4–6), 829843. doi: 10.1080/10556788.2017.1396600.CrossRefGoogle Scholar
Haynes, M (2020) Surface and subsurface radar equations for radar sounders. Annals of Glaciology 61(81), 135142. doi: 10.1017/aog.2020.16.CrossRefGoogle Scholar
Haynes, MS, Chapin, E, Moussessian, A and Madsen, SN (2018a) Surface clutter discrimination analysis for radar sounding interferometry. IEEE Transactions on Aerospace and Electronic Systems 55(2), 9891003. doi: 10.1109/TAES.2018.2867689.CrossRefGoogle Scholar
Haynes, MS, Chapin, E and Schroeder, DM (2018b) Geometric power fall-off in radar sounding. IEEE Transactions on Geoscience and Remote Sensing 56(11), 65716585. doi: 10.1109/TGRS.2018.2840511.CrossRefGoogle Scholar
Heggy, E, Scabbia, G, Bruzzone, L and Pappalardo, RT (2017) Radar probing of Jovian icy moons: understanding subsurface water and structure detectability in the JUICE and Europa missions. Icarus 285, 237251. doi: 10.1016/j.icarus.2016.11.039.CrossRefGoogle Scholar
Heister, A and Scheiber, R (2018) Coherent large beamwidth processing of radio-echo sounding data. Cryosphere 12(9), 29692979. doi: 10.5194/tc-12-2969-2018.CrossRefGoogle Scholar
Hélière, F, Lin, CC, Corr, H and Vaughan, D (2007) Radio echo sounding of Pine Island Glacier, West Antarctica: aperture synthesis processing and analysis of feasibility from space. IEEE Transactions on Geoscience and Remote Sensing 45(8), 25732582. doi: 10.1109/TGRS.2007.897433.CrossRefGoogle Scholar
Hempel, L, Thyssen, F, Gundestrup, N, Clausen, HB and Miller, H (2000) A comparison of radio-echo sounding data and electrical conductivity of the GRIP ice core. Journal of Glaciology 46(154), 369374. doi: 10.3189/172756500781833070.CrossRefGoogle Scholar
Hills, BH, Christianson, K and Holshuh, N (2020) A framework for attenuation method selection evaluated with ice-penetrating radar data at south pole lake. Annals of Glaciology 61(81), 176187. doi: 10.1017/aog.2020.32.CrossRefGoogle Scholar
Hindmarsh, RCand 5 others (2011) Flow at ice-divide triple junctions: 2. three-dimensional views of isochrone architecture from ice-penetrating radar surveys. Journal of Geophysical Research: Earth Surface 116(F2), F02024. doi: 10.1029/2009JF001611.CrossRefGoogle Scholar
Hindmarsh, RCA, Leysinger Vieli, GJMC and Parrenin, F (2009) A large-scale numerical model for computing isochrone geometry. Annals of Glaciology 50(51), 130140. doi: 10.3189/172756409789097450.CrossRefGoogle Scholar
Hindmarsh, RCA, Leysinger Vieli, GJMC, Raymond, MJ and Gudmundsson, GH (2006) Draping or overriding: the effect of horizontal stress gradients on internal layer architecture in ice sheets. Journal of Geophysical Research: Earth Surface 111(F02018). doi: 10.1029/2005JF000309.CrossRefGoogle Scholar
Holschuh, N, Christianson, K, Anandakrishnan, S, Alley, RB and Jacobel, RW (2016) Constraining attenuation uncertainty in common midpoint radar surveys of ice sheets. Journal of Geophysical Research: Earth Surface 121(10), 18761890. doi: 10.1002/2016JF003942.Google Scholar
Holschuh, N, Christianson, K, Paden, J, Alley, R and Anandakrishnan, S (2020) Linking postglacial landscapes to glacier dynamics using swath radar at Thwaites Glacier, Antarctica. Geology 48. doi: 10.1130/G46772.1.CrossRefGoogle Scholar
Holschuh, N, Lilien, D and Christianson, K (2019) Thermal weakening, convergent flow, and vertical heat transport in the Northeast Greenland Ice Stream shear margins. Geophysical Research Letters 46, 81848193. doi: 10.1029/2019GL083436.CrossRefGoogle Scholar
Holschuh, N, Parizek, BR, Alley, RB and Anandakrishnan, S (2017) Decoding ice sheet behavior using englacial layer slopes. Geophysical Research Letters 44(11), 55615570. doi: 10.1002/2017GL073417.CrossRefGoogle Scholar
Holt, JWand 8 others (2006) New boundary conditions for the West Antarctic ice sheet: subglacial topography of the Thwaites and Smith Glacier catchments. Geophysical Research Letters 33(9), GL025561. doi: 10.1029/2005GL025561.CrossRefGoogle Scholar
Humbert, A, Steinhage, D, Helm, V, Beyer, S and Kleiner, T (2018) Missing evidence of widespread subglacial lakes at Recovery Glacier, Antarctica. Journal of Geophysical Research: Earth Surface 123(11), 28022826. doi: 10.1029/2017JF004591.Google Scholar
Ilisei, AM and Bruzzone, L (2015) A system for the automatic classification of ice sheet subsurface targets in radar sounder data. IEEE Transactions on Geoscience and Remote Sensing 53(6), 32603277. doi: 10.1109/TGRS.2014.2372818.CrossRefGoogle Scholar
Ilisei, AM, Khodadadzadeh, M, Ferro, A and Bruzzone, L (2018) An automatic method for subglacial lake detection in ice sheet radar sounder data. IEEE Transactions on Geoscience and Remote Sensing 57(6), 32523270. doi: 10.1109/TGRS.2018.2882911.CrossRefGoogle Scholar
Jacobel, RW and Hodge, SM (1995) Radar internal layers from the Greenland summit. Geophysical Research Letters 22(5), 587590. doi: 10.1029/95GL00110.CrossRefGoogle Scholar
Jacobel, RW, Welch, BC, Osterhouse, D, Pettersson, R and MacGregor, JA (2009) Spatial variation of radar-derived basal conditions on Kamb Ice Stream, West Antarctica. Annals of Glaciology 50(51), 1016. doi: 10.3189/172756409789097504.CrossRefGoogle Scholar
Jenkins, A (2011) Convection-driven melting near the grounding lines of ice shelves and tidewater glaciers. Journal of Physical Oceanography 41(12), 22792294. doi: 10.1175/JPO-D-11-03.1.CrossRefGoogle Scholar
Jeofry, Hand 8 others (2018) Hard rock landforms generate 130 km ice shelf channels through water focusing in basal corrugations. Nature Communications 9(1), 4576. doi: 10.1038/s41467-018-06679-z.CrossRefGoogle ScholarPubMed
Jeofry, H, Ross, N and Siegert, MJ (2019) Comparing numerical ice-sheet model output with radio-echo sounding measurements in the Weddell Sea sector of West Antarctica. Annals of Glaciology 61(81), 188197. doi: 10.1017/aog.2019.39.Google Scholar
Jezek, Kand 7 others (2006) Glaciers and ice sheets mapping orbiter concept. Journal of Geophysical Research: Planets 111(E6), E06S20. doi: 10.1029/2005JE002572.CrossRefGoogle Scholar
Johari, GP and Charette, P (1975) The permittivity and attenuation in polycrystalline and single-crystal ice Ih at 35 and 60 MHz. Journal of Glaciology 14(71), 293303. doi: 10.3189/S002214300002178.CrossRefGoogle Scholar
Jordan, Rand 18 others (2009) The Mars express MARSIS sounder instrument. Planetary and Space Science 57, 19751986. doi: 10.1016/j.pss.2009.09.016.CrossRefGoogle Scholar
Jordan, TMand 7 others (2016) An ice-sheet-wide framework for englacial attenuation from ice-penetrating radar data. The Cryosphere 10(4), 15471570. doi: 10.5194/tc-10-1547-2016.CrossRefGoogle Scholar
Jordan, TMand 6 others (2017) Self-affine subglacial roughness: consequences for radar scattering and basal water discrimination in northern Greenland. The Cryosphere 11(3), 12471264. doi: 10.5194/tc-11-1247-2017.CrossRefGoogle Scholar
Jordan, TAand 7 others (2018a) Anomalously high geothermal flux near the south pole. Scientific Reports 8(1), 16785. doi: 10.1038/s41598-018-35182-0.CrossRefGoogle Scholar
Jordan, TMand 8 others (2018b) A constraint upon the basal water distribution and thermal state of the Greenland ice sheet from radar bed echoes. The Cryosphere 12(9), 28312854. doi: 10.5194/tc-12-2831-2018.CrossRefGoogle Scholar
Jordan, TM, Schroeder, DM, Castelletti, D, Li, J and Dall, J (2019) A polarimetric coherence method to determine ice crystal orientation fabric from radar sounding: application to the NEEM Ice Core Region. IEEE Transactions on Geoscience and Remote Sensing 57(11), 86418657. doi: 10.1109/TGRS.2019.2921980.CrossRefGoogle Scholar
Jordan, TM, Schroeder, DM, Elsworth, CW and Siegfried, MR (2020a) Estimation of ice fabric within Whillans Ice Stream using polarimetric phase-sensitive radar sounding. Annals of Glaciology 61(81), 7483. doi: 10.1017/aog.2020.6.CrossRefGoogle Scholar
Jordan, TMand 6 others (2020b) Modeling ice birefringence and oblique radio wave propagation for neutrino detection at the South Pole. Annals of Glaciology 61(81), 8491. doi: 10.1017/aog.2020.18.Google Scholar
Kalousová, K, Schroeder, DM and Soderlund, KM (2017) Radar attenuation in Europa's ice shell: obstacles and opportunities for constraining the shell thickness and its thermal structure. Journal of Geophysical Research: Planets 122(3), 524545. doi: 10.1002/2016JE005110.Google Scholar
Karlsson, NBand 9 others (2016) Accumulation rates during 1311-2011 CE in North-Central Greenland derived from air-borne radar data. Frontiers in Earth Science 4, 97. doi: 10.3389/feart.2016.00097.CrossRefGoogle Scholar
Karlsson, NBand 6 others (2018) Glaciological characteristics in the dome Fuji region and new assessment for ‘oldest ice’. The Cryosphere 12(7), 24132424. doi: 10.5194/tc-12-2413-2018.CrossRefGoogle Scholar
Karlsson, NBand 6 others (2020) Surface accumulation in Northern Central Greenland during the last 300 years. Annals of Glaciology 61(81), 214224. doi: 10.1017/aog.2020.30.CrossRefGoogle Scholar
Karlsson, NB, Schmidt, LS and Hvidberg, CS (2015) Volume of Martian midlatitude glaciers from radar observations and ice flow modeling. Geophysical Research Letters 42(8), 26272633. doi: 10.1002/2015GL063219.CrossRefGoogle Scholar
Kendrick, Aand 9 others (2018) Surface meltwater impounded by seasonal englacial storage in West Greenland. Geophysical Research Letters 45(19), 1047410481. doi: 10.1029/2018GL079787.CrossRefGoogle Scholar
Khazendar, Aand 8 others (2016) Rapid submarine ice melting in the grounding zones of ice shelves in West Antarctica. Nature Communications 7, 18. doi: 10.1038/ncomms13243.CrossRefGoogle ScholarPubMed
King, EC (2020) The precision of radar-derived subglacial bed topography, a case study from Pine Island Glacier, Antarctica. Annals of Glaciology 61(81), 154161. doi: 10.1017/aog.2020.33.CrossRefGoogle Scholar
King, EC, Pritchard, HD and Smith, AM (2016) Subglacial landforms beneath Rutford Ice Stream, Antarctica: detailed bed topography from ice-penetrating radar. Earth System Science Data 8(1), 151158. doi: 10.5194/essd-8-151-2016.CrossRefGoogle Scholar
Kingslake, Jand 9 others (2014) Full-depth englacial vertical ice sheet velocities measured using phase-sensitive radar. Journal of Geophysical Research: Earth Surface 119(12), 26042618. doi: 10.1002/2014JF003275.Google Scholar
Kingslake, J, Martín, C, Arthern, RJ, Corr, HFJ and King, EC (2016) Ice-flow reorganization in West Antarctica 2.5 kyr ago dated using radar-derived englacial flow velocities. Geophysical Research Letters 43(17), 91039112. doi: 10.1002/2016GL070278.CrossRefGoogle Scholar
Kjær, KHand 9 others (2018) A large impact crater beneath Hiawatha Glacier in northwest Greenland. Science Advances 4(11), eaar8173. doi: 10.1126/sciadv.aar8173.CrossRefGoogle ScholarPubMed
Koenig, LSand 12 others (2016) Annual Greenland accumulation rates (2009–2012) from airborne snow radar. The Cryosphere 10(4), 17391752. doi: 10.5194/tc-10-1739-2016.CrossRefGoogle Scholar
Kofman, Wand 9 others (2015) Properties of the 67P/Churyumov-Gerasimenko interior revealed by CONSERT radar. Science (New York, N.Y.) 349(6247), aab0639. doi: 10.1126/science.aab0639.CrossRefGoogle ScholarPubMed
Kofman, W, Orosei, R and Pettinelli, E (2010) Radar signal propagation and detection through ice. Space Science Reviews 153(1-4), 249271. doi: 10.1007/s11214-010-9642-.CrossRefGoogle Scholar
Koh, G, Lever, JH, Arcone, SA, Marshall, HP and Ray, LE (2010) Autonomous FMCW radar survey of Antarctic shear zone. In Proceedings of the XIII International Conference on Ground Penetrating Radar. IEEE, Lecce, Italy, pp. 15. doi: 10.1109/ICGPR.2010.5550174.CrossRefGoogle Scholar
Koutnik, MRand 7 others (2016) Holocene accumulation and ice flow near the West Antarctic ice sheet divide ice core site. Journal of Geophysical Research: Earth Surface 121, 907924. doi: 10.1002/2015JF003668.Google Scholar
Koutnik, MR and Waddington, ED (2012) Well-posed boundary conditions for limited-domain models of transient ice flow near an ice divide. Journal of Glaciology 58(211), 10081020. doi: 10.3189/2012JoG11J212.CrossRefGoogle Scholar
Kyrke-Smith, TM, Gudmundsson, GH and Farrell, PE (2018) Relevance of detail in basal topography for basal slipperiness inversions: a case study on Pine Island Glacier, Antarctica. Frontiers in Earth Science 6, 33. doi: 10.3389/feart.2018.00033.CrossRefGoogle Scholar
Lalich, DE and Holt, JW (2017) New Martian climate constraints from radar reflectivity within the North Polar Layered Deposits. Geophysical Research Letters 44(2), 657664. doi: 10.1002/2016GL071323.CrossRefGoogle Scholar
Lalich, DE, Holt, JW and Smith, IB (2019) Radar reflectivity as a proxy for the dust content of individual layers in the Martian North Polar Layered Deposits. Journal of Geophysical Research: Planets 124(7), 16901703. doi: 10.1029/2018JE005787.Google Scholar
Langhammer, L, Grab, M, Bauder, A and Maurer, H (2019) Glacier thickness estimations of alpine glaciers using data and modeling constraints. The Cryosphere 2019, 126. doi: 10.5194/tc-2019-55.Google Scholar
Le Brocq, AMand 9 others (2013) Evidence from ice shelves for channelized meltwater flow beneath the Antarctic ice sheet. Nature Geoscience 6(11), 945948. doi: 10.1038/ngeo1977.CrossRefGoogle Scholar
Legarsky, JJ, Gogineni, SP and Akins, TL (2001) Focused synthetic aperture radar processing of ice-sounder data collected over the Greenland ice sheet. IEEE Transactions on Geoscience and Remote Sensing 39(10), 21092117. doi: 10.1109/36.957274.CrossRefGoogle Scholar
Leuschen, C, Gogineni, S and Tammana, D (2000) SAR processing of radar echo sounder data. In IGARSS 2000. IEEE 2000 International Geoscience and Remote Sensing Symposium. Taking the Pulse of the Planet: The Role of Remote Sensing in Managing the Environment. Proceedings (Cat. No. 00CH37120), Vol. 6. IEEE, Honolulu Hawaii, pp. 25702572. doi: 10.1109/IGARSS.2000.859643.CrossRefGoogle Scholar
Lewis, Gand 5 others (2017) Regional Greenland accumulation variability from Operation IceBridge airborne accumulation radar. The Cryosphere 11(2), 773788. doi: 10.5194/tc-11-773-2017.CrossRefGoogle Scholar
Leysinger-Vieli, G, Hindmarsh, R and Siegert, M (2007) Three-dimensional flow influences on radar layer stratigraphy. Annals of Glaciology 46(1), 2228. doi: 10.3189/172756407782871729.CrossRefGoogle Scholar
Leysinger Vieli, GJMC, Martín, C and Hindmarsh, RCA (2018) Basal freeze-on generates complex ice-sheet stratigraphy. Nature Communications 9(1), 4669. doi: 10.1038/s41467-018-07083-3.CrossRefGoogle ScholarPubMed
Li, Jand 8 others (2018) Multi-channel and multi-polarization radar measurements around the NEEM site. The Cryosphere 12(8), 26892705. doi: 10.5194/tc-12-2689-2018.CrossRefGoogle Scholar
Lilien, DA, Hills, B, Driscol, J, Jacobel, R and Christianson, K (2020) ImpDAR: an open-source impulse radar processor. Annals of Glaciology 61(81), 114123. doi: 10.1017/aog.2020.44.CrossRefGoogle Scholar
Lorente, Rand 7 others (2017) The ESA JUICE mission: the Science and the Science Operations. In EGU General Assembly Conference Abstracts, Vol. 19, Vienna, Austria, p. 14611.Google Scholar
Lythe, MB and Vaughan, DG (2001) BEDMAP: a new ice thickness and subglacial topographic model of Antarctica. Journal of Geophysical Research: Solid Earth 106(B6), 1133511351. doi: 10.1029/2000JB900449.CrossRefGoogle Scholar
MacGregor, JAand 9 others (2015a) Radiostratigraphy and age structure of the Greenland ice sheet. Journal of Geophysical Research: Earth Surface 120, 130. doi: 10.1002/2014JF003215.Google Scholar
MacGregor, JAand 9 others (2015b) Radar attenuation and temperature within the Greenland ice sheet. Journal of Geophysical Research: Earth Surface 120(6), 9831008. doi: 10.1002/2014JF003418.Google Scholar
MacGregor, JAand 6 others (2016) Holocene deceleration of the Greenland ice sheet. Science (New York, N.Y.) 351(6273), 590593. doi: 10.1126/science.aab1702.CrossRefGoogle ScholarPubMed
MacKie, EJ and Schroeder, DM (2019) Paleo Observations Used to Geostatistically Simulate the Subglacial Geology of Thwaites Glacier. AGU Fall Meeting, August 2019, San Francisco.Google Scholar
Martín, C, Gudmundsson, GH, Pritchard, HD and Gagliardini, O (2009) On the effects of anisotropic rheology on ice flow, internal structure, and the age-depth relationship at ice divides. Journal of Geophysical Research: Earth Surface 114(F4), F04001. doi: 10.1029/2008JF001204.CrossRefGoogle Scholar
Masolov, VN, Popov, SV, Lukin, VV, Sheremetyev, AN and Popkov, AM (2006) Russian geophysical studies of Lake Vostok, Central East Antarctica. In Antarctica. Springer, Berlin, Heidelberg, pp. 135140. doi: 10.1007/3-540-32934-X-16.CrossRefGoogle Scholar
Matsuoka, Kand 6 others (2003) Crystal orientation fabrics within the Antarctic ice sheet revealed by a multipolarization plane and dual-frequency radar survey. Journal of Geophysical Research: Solid Earth 108(B10), 110. doi: 10.1029/2003JB002425.CrossRefGoogle Scholar
Matsuoka, K (2011) Pitfalls in radar diagnosis of ice-sheet bed conditions: lessons from englacial attenuation models. Geophysical Research Letters 38(5), L05505. doi: 10.1029/2010GL046205.CrossRefGoogle Scholar
Matsuoka, K, MacGregor, JA and Pattyn, F (2012) Predicting radar attenuation within the Antarctic ice sheet. Earth and Planetary Science Letters 359, 173183. doi: 10.1016/j.epsl.2012.10.018.CrossRefGoogle Scholar
Matsuoka, K, Morse, D and Raymond, C (2010) Estimating englacial radar attenuation using depth profiles of the returned power, central West Antarctica. Journal of Geophysical Research: Earth Surface 115(F2), F02012. doi: 10.1029/2009JF001496.CrossRefGoogle Scholar
McKinnon, W (2005) Radar sounding of convecting ice shells in the presence of convection: application to Europa, Ganymede, and Callisto. In Workshop on Radar Investigations of Planetary and Terrestrial Environments, Houston, TX.Google Scholar
Medley, Band 5 others (2014) Constraining the recent mass balance of Pine Island and Thwaites glaciers, West Antarctica, with airborne observations of snow accumulation. The Cryosphere 8(4), 13751392. doi: 10.5194/tc-8-1375-2014.CrossRefGoogle Scholar
Michaelides, RJ and Schroeder, DM (2019) Doppler-based discrimination of radar sounder target scattering properties: a case study of subsurface water geometry in Europa's ice shell. Icarus 326, 2936. doi: 10.1016/j.icarus.2019.02.037.CrossRefGoogle Scholar
Mingo, L, Flowers, GE, Crawford, AJ, Mueller, DR and Bigelow, DG (2020) A stationary impulse-radar system for autonomous deployment in cold and temperate environments. Annals of Glaciology 61(81), 99107. doi: 10.1017/aog.2020.2.CrossRefGoogle Scholar
Mirino, M, Frigeri, A, Orosei, R, Rossi, AP and Cantini, F (2018) MARSIS radar data interpretation to characterize the deeper layers in the North Polar Cap on Mars. Advances in Astronautics Science and Technology 1(1), 3137. doi: 10.1007/s42423-018-0008-2.CrossRefGoogle Scholar
Montgomery, L, Koenig, L, Lenaerts, JTM and Kuipers Munneke, P (2020) Accumulation rates (2009–2017) in Southeast Greenland derived from airborne snow radar and comparison with regional climate models. Annals of Glaciology 61(81), 225233. doi: doi:10.1017/aog.2020.8.Google Scholar
Moore, JC (2000) Models of radar absorption in Europan ice. Icarus 147(1), 292300. doi: 10.1006/icar.2000.6425.CrossRefGoogle Scholar
Moore, JC and Fujita, S (1993) Dielectric properties of ice containing acid and salt impurity at microwave and low frequencies. Journal of Geophysical Research: Solid Earth 98(B6), 97699780. doi: 10.1029/93JB00710.CrossRefGoogle Scholar
Morlighem, Mand 5 others (2011) A mass conservation approach for mapping glacier ice thickness. Geophysical Research Letters 38, L19503. doi: 10.1029/2011GL048659.CrossRefGoogle Scholar
Morlighem, Mand 5 others (2017) BedMachine v3: complete bed topography and ocean bathymetry mapping of Greenland from multibeam echo sounding combined with mass conservation. Geophysical Research Letters 44(21), 11051. doi: 10.1002/2017GL074954.CrossRefGoogle ScholarPubMed
Morlighem, Mand 9 others (2019) Deep glacial troughs and stabilizing ridges unveiled beneath the margins of the Antarctic ice sheet. Nature Geoscience 13, 132137. doi: 10.1038/s41561-019-0510-8.CrossRefGoogle Scholar
Musil, GJ and Doake, C (1987) Imaging subglacial topography by a synthetic aperture radar technique. Annals of Glaciology 9, 170175. doi: 10.3189/S0260305500000562.CrossRefGoogle Scholar
Muto, A, Alley, RB, Parizek, BR and Anandakrishnan, S (2019) Bed-type variability and till (dis)continuity beneath Thwaites Glacier, West Antarctica. Annals of Glaciology 19. doi: 10.1017/aog.2019.32.Google Scholar
Nereson, N and Raymond, C (2001) The elevation history of ice streams and the spatial accumulation pattern along the Siple Coast of West Antarctica inferred from ground-based radar data from three inter-ice-stream ridges. Journal of Glaciology 47(157), 303313. doi: 10.3189/172756501781832197.CrossRefGoogle Scholar
Nereson, NA and Waddington, ED (2002) Isochrones and isotherms beneath migrating ice divides. Journal of Glaciology 48(160), 95108. doi: 10.3189/172756502781831647.CrossRefGoogle Scholar
Nerozzi, S and Holt, JW (2018) Earliest accumulation history of the North Polar Layered Deposits, Mars from SHARAD. Icarus 308, 128137. doi: 10.1016/j.icarus.2017.05.027.CrossRefGoogle Scholar
Nerozzi, S and Holt, JW (2019) Buried ice and sand caps at the North Pole of Mars: revealing a record of climate change in the Cavi Unit with SHARAD. Geophysical Research Letters 46(13), 72787286. doi: 10.1029/2019GL082114.CrossRefGoogle Scholar
Ng, F and Conway, H (2004) Fast-flow signature in the stagnated Kamb Ice Stream, West Antarctica. Geology 32(6), 481484. doi: 10.1130/G20317.1.CrossRefGoogle Scholar
Nicholls, KWand 5 others (2015) A ground-based radar for measuring vertical strain rates and time-varying basal melt rates in ice sheets and shelves. Journal of Glaciology 61(230), 10791087. doi: 10.3189/2015JoG15J073.CrossRefGoogle Scholar
Nielsen, LT, Karlsson, NB and Hvidberg, CS (2015) Large-scale reconstruction of accumulation rates in northern Greenland from radar data. Annals of Glaciology 56(70), 7078. doi: 10.3189/2015AoG70A062.CrossRefGoogle Scholar
Nozette, Sand 10 others (2010) The lunar Reconnaissance Orbiter miniature radio frequency (Mini-RF) technology demonstration. Space Science Reviews 150, 285302. doi: 10.1007/s11214-009-9607-5.CrossRefGoogle Scholar
Onana, V, Koenig, LS, Ruth, J, Studinger, M and Harbeck, JP (2015) A semiautomated multilayer picking algorithm for ice-sheet radar echograms applied to ground-based near-surface data. IEEE Transactions on Geoscience and Remote Sensing 53(1), 5169. doi: 10.1109/TGRS.2014.2318208.CrossRefGoogle Scholar
Orosei, Rand 21 others (2018) Radar evidence of subglacial liquid water on Mars. Science (New York, N.Y.) 361(6401), 490493. doi: 10.1126/science.aar7268.Google ScholarPubMed
Oswald, G and Gogineni, SP (2008) Recovery of subglacial water extent from Greenland radar survey data. Journal of Glaciology 54(184), 94106. doi: 10.3189/002214308784409107.CrossRefGoogle Scholar
Oswald, GK, Rezvanbehbahani, S and Stearns, LA (2018) Radar evidence of ponded subglacial water in Greenland. Journal of Glaciology 64(247), 711729. doi: 10.1017/jog.2018.60.CrossRefGoogle Scholar
Oswald, G and Robin, GdQ (1973) Lakes beneath the Antarctic ice sheet. Nature 245(5423), 251254. doi: 10.1038/245251a0.CrossRefGoogle Scholar
Paden, JDand 5 others (2005) Wideband measurements of ice sheet attenuation and basal scattering. IEEE Geoscience and Remote Sensing Letters 2(2), 164168. doi: 10.1109/LGRS.2004.842474.CrossRefGoogle Scholar
Paden, JD, Akins, T, Dunson, D, Allen, C and Gogineni, P (2010) Ice-sheet bed 3-D tomography. Journal of Glaciology 56(195), 311. doi: 10.3189/002214310791190811.CrossRefGoogle Scholar
Palmer, SJand 8 others (2013) Greenland subglacial lakes detected by radar. Geophysical Research Letters 40(23), 61546159. doi: 10.1002/2013GL058383.CrossRefGoogle Scholar
Panton, C and Karlsson, NB (2015) Automated mapping of near bed radio-echo layer disruptions in the Greenland ice sheet. Earth and Planetary Science Letters 432, 323331. doi: 10.1016/j.epsl.2015.10.024.CrossRefGoogle Scholar
Pappalardo, Rand 7 others (2015) Science and reconnaissance from the Europa clipper mission concept: exploring Europa's habitability. In EGU General Assembly Conference Abstracts, Vol. 17, Vienna, Austria.Google Scholar
Paren, JG and Robin, G (1975) Internal reflections in polar ice sheets. Journal of Glaciology 14, 251259. doi: 10.3189/S0022143000021730.CrossRefGoogle Scholar
Parrenin, Fand 11 others (2017) Is there 1.5-million-year-old ice near Dome C, Antarctica? The Cryosphere 11(6), 24272437. doi: 10.5194/tc-11-2427-2017CrossRefGoogle Scholar
Parrenin, F, Hindmarsh, R and Rémy, F (2006) Analytical solutions for the effect of topography, accumulation rate and lateral flow divergence on isochrone layer geometry. Journal of Glaciology 52(177), 191202. doi: 10.3189/172756506781828728.CrossRefGoogle Scholar
Parsons, R and Holt, J (2016) Constraints on the formation and properties of a Martian Lobate debris apron: Insights from high-resolution topography, SHARAD radar data, and a numerical ice flow model. Journal of Geophysical Research: Planets 121(3), 2015JE004927. doi: 10.1002/2015JE004927.Google Scholar
Patterson, GWand 9 others (2017) Bistatic radar observations of the Moon using Mini-RF on LRO and the Arecibo Observatory. Icarus 283, 219. doi: 10.1016/j.icarus.2016.05.017.CrossRefGoogle Scholar
Paxman, GJand 5 others (2019) Subglacial geology and geomorphology of the Pensacola-Pole Basin, East Antarctica. Geochemistry, Geophysics, Geosystems 20, 27862807. doi: 10.1029/2018GC008126.CrossRefGoogle Scholar
Peters, MEand 5 others (2007) Along-track focusing of airborne radar sounding data from West Antarctica for improving basal reflection analysis and layer detection. IEEE Transactions on Geoscience and Remote Sensing 45(9), 27252736. doi: 10.1109/TGRS.2007.897416.CrossRefGoogle Scholar
Peters, ME, Blankenship, DD and Morse, DL (2005) Analysis techniques for coherent airborne radar sounding: application to West Antarctic ice streams. Journal of Geophysical Research: Solid Earth 110(B6), B06303. doi: 10.1029/2004JB003222.CrossRefGoogle Scholar
Peters, ST, Schroeder, DM, Castelletti, D, Haynes, M and Romero-Wolf, A (2018) In situ demonstration of a passive radio sounding approach using the sun for echo detection. IEEE Transactions on Geoscience and Remote Sensing 56(12), 73387349. doi: 10.1109/TGRS.2018.2850662.CrossRefGoogle Scholar
Petersen, EI, Holt, JW and Levy, JS (2018) High ice purity of Martian Lobate Debris Aprons at the regional scale: evidence from an orbital radar sounding survey in Deuteronilus and Protonilus Mensae. Geophysical Research Letters 45(21), 1159511604. doi: 10.1029/2018GL079759.CrossRefGoogle Scholar
Pettinelli, Eand 6 others (2015) Dielectric properties of Jovian satellite ice analogs for subsurface radar exploration: a review. Reviews of Geophysics 53(3), 593641. doi: 10.1002/2014RG000463.CrossRefGoogle Scholar
Pettit, ECand 6 others (2011) The crossover stress, anisotropy and the ice flow law at Siple Dome, West Antarctica. Journal of Glaciology 57(201), 3952. doi: 10.3189/002214311795306619.CrossRefGoogle Scholar
Popov, S (2017) Flow-lines computation and their use in subglacial geomorphology and glacial erosion modeling: the Princess Elizabeth land (East Antarctica) case study. Geomorfologiya 2017(1), 4654. doi: 10.15356/0435-4281-2017-1-46-54.Google Scholar
Popov, S (2020) Fifty-five years of Russian radio-echo sounding investigations in Antarctica. Annals of Glaciology 61(81), 1424. doi: 10.1017/aog.2020.4.CrossRefGoogle Scholar
Pritchard, HD (2014) Bedgap: where next for Antarctic subglacial mapping? Antarctic Science 26(6), 742757. doi: 10.1017/S095410201400025X.CrossRefGoogle Scholar
Pritchard, HD, King, EC, McCarthy, M and Mayer, C (2020) Bedmap Himalayas: development of an airborne ice-sounding radar for glacier thickness surveys in High-Mountain Asia. Annals of Glaciology 61(81), 3535. doi: 10.1017/aog.2020.29.CrossRefGoogle Scholar
Putzig, NEand 6 others (2018) Three-dimensional radar imaging of structures and craters in the Martian polar caps. Icarus 308, 138147. doi: 10.1016/j.icarus.2017.09.023.CrossRefGoogle ScholarPubMed
Rahnemoonfar, M, Fox, GC, Yari, M and Paden, J (2017) Automatic ice surface and bottom boundaries estimation in radar imagery based on level-set approach. IEEE Transactions on Geoscience and Remote Sensing 55(9), 51155122. doi: 10.1109/TGRS.2017.2702200.CrossRefGoogle Scholar
Rasmussen, L (1988) Bed topography and mass-balance distribution of Columbia Glacier, Alaska, USA, determined from 1048 sequential aerial photography. Journal of Glaciology 34(117), 208216. doi: 10.3189/S0022143000032251.CrossRefGoogle Scholar
Raymond, CF (1983) Deformation in the vicinity of ice divides. Journal of Glaciology 29(103), 357373. doi: 10.3189/S0022143000030288.CrossRefGoogle Scholar
Rezvanbehbahani, S, Stearns, LA, Kadivar, A, Walker, JD and van der Veen, CJ, 2017) Predicting the geothermal heat flux in Greenland: a machine learning approach. Geophysical Research Letters 44(24), 12271. doi: 10.1002/2017GL075661.CrossRefGoogle Scholar
Rezvanbehbahani, S, Stearns, LA, van der Veen, CJ, Oswald, GKA and Greve, R (2019) Constraining the geothermal heat flux in Greenland at regions of radar-detected basal water. Journal of Glaciology 65(254), 10231034. doi: 10.1017/jog.2019.79.CrossRefGoogle Scholar
Rignot, E, Mouginot, J, Larsen, C, Gim, Y and Kirchner, D (2013) Low-frequency radar sounding of temperate ice masses in Southern Alaska. Geophysical Research Letters 40(20), 53995405. doi: 10.1002/2013GL057452.CrossRefGoogle Scholar
Rippin, Dand 9 others (2014) Basal roughness of the institute and Möller ice streams, west Antarctica: process determination and landscape interpretation. Geomorphology 214, 139147. doi: 10.1016/j.geomorph.2014.01.021.CrossRefGoogle Scholar
Robin, GdQ (1958) Glaciology III: Seismic Shooting and Related Investigations. Norsk Polarinstitutt, Tromso, Norway.Google Scholar
Robin, GdQ (1975) Radio-echo sounding: glaciological interpretations and applications. Journal of Glaciology 15(73), 4964. doi: 10.3189/s0022143000034262.CrossRefGoogle Scholar
Rodriguez-Morales, Fand 9 others (2013) Advanced multifrequency radar instrumentation for polar research. IEEE Transactions on Geoscience and Remote Sensing 52(5), 28242842. doi: 10.1109/TGRS.2013.2266415.CrossRefGoogle Scholar
Romero-Wolf, Aand 5 others (2015) A passive probe for subsurface oceans and liquid water in Jupiter's icy moons. Icarus 248, 463477. doi: 10.1016/j.icarus.2014.10.043.CrossRefGoogle Scholar
Romero-Wolf, Aand 7 others (2016) Prospects of passive radio detection of a subsurface ocean on Europa with a lander. Planetary and Space Science 129, 118121. doi: 10.1016/j.pss.2016.06.010.CrossRefGoogle Scholar
Ross, N and Siegert, M (2020) Basal melting over Subglacial Lake Ellsworth and its catchment: insights from englacial layering. Annals of Glaciology 61(81), 198205. doi: 10.1017/aog.2020.50.Google Scholar
Ross, Nand 9 others (2012) Steep reverse bed slope at the grounding line of the Weddell Sea sector in West Antarctica. Nature Geoscience 5(6), 393396. doi: 10.1038/ngeo1468.CrossRefGoogle Scholar
Rutishauser, Aand 8 others (2018) Discovery of a hypersaline subglacial lake complex beneath Devon Ice Cap, Canadian Arctic. Science Advances 4(4), eaar4353. doi: 10.1126/sciadv.aar4353.CrossRefGoogle ScholarPubMed
Scanlan, KMand 5 others (2019) Geometric determination of ionospheric total electron content from dual frequency radar sounding measurements. Planetary and Space Science 178, 104696. doi: 10.1016/j.pss.2019.07.010.CrossRefGoogle Scholar
Scanlan, KM, Rutishauser, A, Young, DA and Blankenship, DD (2020) Interferometric discrimination of cross-track bed clutter in ice-penetrating radar sounding data. Annals of Glaciology 61(81), 6873. doi: 10.1017/aog.2020.20.CrossRefGoogle Scholar
Schmidt, B, Blankenship, DD, Patterson, G and Schenk, P (2011) Active formation of chaos terrain over shallow subsurface water on Europa. Nature 479(7374), 502505. doi: 10.1038/nature10608.CrossRefGoogle ScholarPubMed
Schmidt, LS, Hvidberg, CS, Kim, JR and Karlsson, NB (2019) Non-linear flow modelling of a Martian Lobate Debris Apron. Journal of Glaciology 65(254), 889899. doi: 10.1017/jog.2019.54.CrossRefGoogle Scholar
Schroeder, DMand 7 others (2016b) Assessing the potential for passive radio sounding of Europa and Ganymede with RIME and REASON. Planetary and Space Science 134, 5260. doi: 10.1016/j.pss.2016.10.007.CrossRefGoogle Scholar
Schroeder, DMand 9 others (2019) Multidecadal observations of the Antarctic ice sheet from restored analog radar records. Proceedings of the National Academy of Sciences 116(38), 1886718873. doi: 10.1073/pnas.1821646116.CrossRefGoogle ScholarPubMed
Schroeder, DM, Blankenship, DD, Raney, RK and Grima, C (2014a) Estimating subglacial water geometry using radar bed echo specularity: application to Thwaites Glacier, West Antarctica. IEEE Geoscience and Remote Sensing Letters 12(3), 443447. doi: 10.1109/LGRS.2014.2337878.CrossRefGoogle Scholar
Schroeder, DM, Blankenship, DD and Young, DA (2013) Evidence for a water system transition beneath Thwaites Glacier, West Antarctica. Proceedings of the National Academy of Sciences 110(30), 1222512228. doi: 10.1073/pnas.1302828110.CrossRefGoogle ScholarPubMed
Schroeder, DM, Blankenship, DD, Young, DA and Quartini, E (2014b) Evidence for elevated and spatially variable geothermal flux beneath the West Antarctic ice sheet. Proceedings of the National Academy of Sciences 111(25), 90709072. doi: 10.1073/pnas.1405184111.CrossRefGoogle Scholar
Schroeder, DM, Blankenship, DD, Young, DA, Witus, AE and Anderson, JB (2014c) Airborne radar sounding evidence for deformable sediments and outcropping bedrock beneath Thwaites Glacier, West Antarctica. Geophysical Research Letters 41(20), 72007208. doi: 10.1002/2014GL061645.CrossRefGoogle Scholar
Schroeder, DM, Grima, C and Blankenship, DD (2016a) Evidence for variable grounding-zone and shear-margin basal conditions across Thwaites Glacier, West Antarctica. Geophysics 81(1), WA35WA43. doi: 10.1190/geo2015-0122.1.CrossRefGoogle Scholar
Schroeder, DM, Hilger, AM, Paden, JD, Young, DA and Corr, HF (2018) Ocean access beneath the southwest tributary of Pine Island Glacier, West Antarctica. Annals of Glaciology 59(76), 1015. doi: 10.1017/aog.2017.45.CrossRefGoogle Scholar
Schroeder, DM, Seroussi, H, Chu, W and Young, DA (2016c) Adaptively constraining radar attenuation and temperature across the Thwaites Glacier catchment using bed echoes. Journal of Glaciology 62(236), 10751082. doi: 10.1017/jog.2016.100.CrossRefGoogle Scholar
Sergienko, O (2013) Basal channels on ice shelves. Journal of Geophysical Research: Earth Surface 118(3), 13421355. doi: 10.1002/jgrf.20105.Google Scholar
Seroussi, H, Ivins, ER, Wiens, DA and Bondzio, J (2017) Influence of a West Antarctic mantle plume on ice sheet basal conditions. Journal of Geophysical Research: Solid Earth 122(9), 71277155. doi: 10.1002/2017JB014423.Google Scholar
Seu, Rand 11 others (2007) SHARAD sounding radar on the Mars Reconnaissance Orbiter. Journal of Geophysical Research E: Planets 112(5). doi: 10.1029/2006JE002745.Google Scholar
Shi, Land 8 others (2010) Multichannel coherent radar depth sounder for NASA operation ice bridge. In 2010 IEEE International Geoscience and Remote Sensing Symposium. IEEE, Honolulu, Hawaii, pp. 17291732. doi: 10.1109/IGARSS.2010.5649518.CrossRefGoogle Scholar
Shoemaker, IMand 5 others (2020) Reflections on the anomalous ANITA Events: the Antarctic subsurface as a possible explanation. Annals of Glaciology 61(81), 9298. doi: 10.1017/aog.2020.19.CrossRefGoogle Scholar
Siegert, MJ (2003) Glacial–interglacial variations in central east Antarctic ice accumulation rates. Quaternary Science Reviews 22(5-7), 741750. doi: 10.1016/S0277-3791(02)00191-9.CrossRefGoogle Scholar
Siegert, MJand 9 others (2004) Ice flow direction change in interior West Antarctica. Science (New York, N.Y.) 305, 19481951. doi: 10.1126/science.1101072.CrossRefGoogle ScholarPubMed
Siegert, MJand 7 others (2016) Subglacial controls on the flow of Institute Ice Stream, West Antarctica. Annals of Glaciology 57(73), 1924. doi: 10.1017/aog.2016.17.CrossRefGoogle Scholar
Siegert, MJ (2018) A 60-year international history of Antarctic subglacial lake exploration. Geological Society, London, Special Publications 461(1), 721. doi: 10.1144/SP461.5.CrossRefGoogle Scholar
Siegfried, MR, Fricker, HA, Carter, SP and Tulaczyk, S (2016) Episodic ice velocity fluctuations triggered by a subglacial flood in West Antarctica. Geophysical Research Letters 43(6), 26402648. doi: 10.1002/2016GL067758.CrossRefGoogle Scholar
Sime, LC, Hindmarsh, RC and Corr, H (2011) Automated processing to derive dip angles of englacial radar reflectors in ice sheets. Journal of Glaciology 57(202), 260266. doi: 10.3189/002214311796405870.CrossRefGoogle Scholar
Smith, BE, Fricker, HA, Joughin, IR and Tulaczyk, S (2009) An inventory of active subglacial lakes in Antarctica detected by ICESat (2003–2008). Journal of Glaciology 55(192), 573595. doi: 10.3189/002214309789470879.CrossRefGoogle Scholar
Smith, IB, Putzig, NE, Holt, JW and Phillips, RJ (2016) An ice age recorded in the polar deposits of Mars. Science (New York, N.Y.) 352, 10751078. doi: 10.1126/science.aad6968.CrossRefGoogle ScholarPubMed
Sorge, E (1933) The scientific results of the Wegener expeditions to Greenland. The Geographical Journal 81(4), 333344. doi: 10.2307/1785439.CrossRefGoogle Scholar
Sori, MM and Bramson, AM (2019) Water on Mars, with a grain of salt: local heat anomalies are required for basal melting of ice at the South Pole today. Geophysical Research Letters 46(3), 12221231. doi: 10.1029/2018GL080985.CrossRefGoogle Scholar
Spudis, PDand 9 others (2013) Evidence for water ice on the Moon: results for anomalous polar craters from the LRO Mini-RF imaging radar. Journal of Geophysical Research: Planets 118(10), 20162029. doi: 10.1002/jgre.20156.Google Scholar
Steenson, BO (1951) Radar Methods for the Exploration of Glaciers (PhD thesis). California Institute of Technology.Google Scholar
Steinbrügge, Gand 5 others (2018) Assessing the potential for measuring Europa's tidal love number h2 using radar sounder and topographic imager data. Earth and Planetary Science Letters 482, 334341. doi: 10.1016/j.epsl.2017.11.028.CrossRefGoogle Scholar
Stern, W (1930) Principles, methods and results of electrodynamic thickness measurement of glacier ice. Zeitschrift fur Gletscherkunde 18, 24.Google Scholar
Stewart, CL, Christoffersen, P, Nicholls, KW, Williams, MJ and Dowdeswell, JA (2019) Basal melting of Ross Ice Shelf from solar heat absorption in an ice-front polynya. Nature Geoscience 12(6), 435440. doi: 10.1038/s41561-019-0356-0.CrossRefGoogle Scholar
Stillman, DE, MacGregor, JA and Grimm, RE (2013) The role of acids in electrical conduction through ice. Journal of Geophysical Research: Earth Surface 118(1), 116. doi: 10.1029/2012JF002603.Google Scholar
Stuurman, CMand 6 others (2016) SHARAD detection and characterization of subsurface water ice deposits in Utopia Planitia, Mars. Geophysical Research Letters 43(18), 94849491. doi: 10.1002/2016GL070138.CrossRefGoogle Scholar
Tang, XY, Guo, JX, Sun, B, Wang, TT and Cui, XB (2016) Ice thickness, internal layers, and surface and subglacial topography in the vicinity of Chinese Antarctic Taishan station in Princess Elizabeth Land, East Antarctica. Applied Geophysics 13(1), 203208. doi: 10.1007/s11770-016-0540-6.CrossRefGoogle Scholar
Tinto, Kand 9 others (2019) Ross Ice Shelf response to climate driven by the tectonic imprint on seafloor bathymetry. Nature Geoscience 12, 441449. doi: 10.1038/s41561-019-0370-2.CrossRefGoogle Scholar
Turchetti, S, Dean, K, Naylor, S and Siegert, M (2008) Accidents and opportunities: a history of the radio echo-sounding of Antarctica, 1958–79. The British Journal for the History of Science 41(3), 417444. doi: 10.1017/S0007087408000903.CrossRefGoogle Scholar
Vankova, Iand 5 others (2020) Depth-dependent artifacts resulting from ApRES signal clipping. Annals of Glaciology 61(81), 108113. doi: 10.1017/aog.2020.56.Google Scholar
Vaughan, DGand 9 others (2006) New boundary conditions for the West Antarctic ice sheet: Subglacial topography beneath Pine Island Glacier. Geophysical Research Letters 33(9), GL025561. doi: 10.1029/2005GL025588.CrossRefGoogle Scholar
Waddington, ED, Neumann, TA, Koutnik, MR, Marshall, HP and Morse, DL (2007) Inference of accumulation-rate patterns from deep layers in glaciers and ice sheets. Journal of Glaciology 53(183), 694712.CrossRefGoogle Scholar
Wang, Band 6 others (2018) Summit of the East Antarctic ice sheet underlain by thick ice-crystal fabric layers linked to glacial–interglacial environmental change. Geological Society, London, Special Publications 461(1), 131143. doi: 10.1144/SP461.1.CrossRefGoogle Scholar
Wang, Band 9 others (2020) Removal of ‘strip noise’ in radio-echo sounding data using combined wavelet and 2-D DFT filtering. Annals of Glaciology 61(81), 124134. doi: 10.1017/aog.2019.4.Google Scholar
Warner, RC and Budd, W (2000) Derivation of ice thickness and bedrock topography in data-gap regions over Antarctica. Annals of Glaciology 31, 191197. doi: 10.3189/172756400781820011.CrossRefGoogle Scholar
Whitten, JL, Campbell, BA and Morgan, GA (2017) A subsurface depocenter in the South Polar Layered Deposits of Mars. Geophysical Research Letters 44(16), 81888195. doi: 10.1002/2017GL074069.CrossRefGoogle Scholar
Winter, Aand 9 others (2017) Comparison of measurements from different radio-echo sounding systems and synchronization with the ice core at Dome C, Antarctica. The Cryosphere 11(1), 653668. doi: 10.5194/tc-11-653-2017.CrossRefGoogle Scholar
Winter, Kand 10 others (2019) Radar-detected englacial debris in the West Antarctic ice sheet. Geophysical Research Letters 46(10), 1045410462. doi: 10.1029/2019GL084012.CrossRefGoogle Scholar
Winter, A, Steinhage, D, Creyts, TT, Kleiner, T and Eisen, O (2019) Age stratigraphy in the East Antarctic ice sheet inferred from radio-echo sounding horizons. Earth System Science Data 11(3), 10691081. doi: 10.5194/essd-11-1069-2019.CrossRefGoogle Scholar
Wolovick, MJ, Bell, RE, Creyts, TT and Frearson, N (2013) Identification and control of subglacial water networks under Dome A, Antarctica. Journal of Geophysical Research: Earth Surface 118(1), 140154. doi: 10.1029/2012JF002555.Google Scholar
Wright, A and Siegert, M (2012) A fourth inventory of Antarctic subglacial lakes. Antarctic Science 24(6), 659664. doi: 10.1017/S095410201200048X.CrossRefGoogle Scholar
Wrona, Tand 5 others (2017) Position and variability of complex structures in the central East Antarctic ice sheet. In M, Siegert, S, Jamieson and D, White (eds), Exploration of Subsurface Antarctica: Uncovering Past Changes and Modern Processes, number 461 in Special Publication. Geological Society of London, pp. 113129. doi: 10.1144/SP461.12.Google Scholar
Wu, Xand 5 others (2011) Ice sheet bed mapping with airborne SAR tomography. IEEE Transactions on Geoscience and Remote Sensing 49(10), 37913802. doi: 10.1109/TGRS.2011.2132802.CrossRefGoogle Scholar
Xiong, S and Muller, JP (2019) Automated reconstruction of subsurface interfaces in Promethei Lingula near the Martian south pole by using SHARAD data. Planetary and Space Science 166, 5969. doi: 10.1016/j.pss.2018.08.001.CrossRefGoogle Scholar
Xiong, S, Muller, JP and Carretero, RC (2018) A new method for automatically tracing englacial layers from MCoRDS data in NW Greenland. Remote Sensing 10(1). doi: 10.3390/rs10010043.Google Scholar
Yan, JB, Gogineni, P and O'Neill, C (2018) L-band radar sounder for measuring ice basal conditions and ice-shelf melt rate. In IGARSS 2018-2018 IEEE International Geoscience and Remote Sensing Symposium. IEEE, Valencia, Spain, pp. 41354137. doi: 10.1109/IGARSS.2018.8518210.CrossRefGoogle Scholar
Young, DAand 9 others (2011) A dynamic early East Antarctic ice sheet suggested by ice-covered fjord landscapes. Nature 474(7349), 7275. doi: 10.1038/nature10114.CrossRefGoogle ScholarPubMed
Young, TJand 8 others (2018) Resolving the internal and basal geometry of ice masses using imaging phase-sensitive radar. Journal of Glaciology 64(246), 649660. doi: 10.1017/jog.2018.54.CrossRefGoogle Scholar
Young, DA, Schroeder, D, Blankenship, DD, Kempf, SD and Quartini, E (2016) The distribution of basal water between Antarctic subglacial lakes from radar sounding. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 374(2059), 20140297. doi: 10.1098/rsta.2014.0297.CrossRefGoogle ScholarPubMed