Articles | Volume 18, issue 5
https://doi.org/10.5194/tc-18-2455-2024
https://doi.org/10.5194/tc-18-2455-2024
Research article
 | 
17 May 2024
Research article |  | 17 May 2024

Mapping the vertical heterogeneity of Greenland's firn from 2011–2019 using airborne radar and laser altimetry

Anja Rutishauser, Kirk M. Scanlan, Baptiste Vandecrux, Nanna B. Karlsson, Nicolas Jullien, Andreas P. Ahlstrøm, Robert S. Fausto, and Penelope How
Abstract

The firn layer on the Greenland Ice Sheet (GrIS) plays a crucial role in buffering surface meltwater runoff, which is constrained by the available firn pore space and impermeable ice layers that limit deeper meltwater percolation. Understanding these firn properties is essential for predicting current and future meltwater runoff and its contribution to global sea-level rise. While very-high-frequency (VHF) radars have been extensively used for surveying the GrIS, their lower bandwidth restricts direct firn stratigraphy extraction. In this study, we use concurrent VHF airborne radar and laser altimetry data collected as part of Operation IceBridge over the 2011–2019 period to investigate our hypothesis that vertical heterogeneities in firn (i.e. ice layers) cause vertical offsets in the radar surface reflection (dz). Our results, corroborated by modelling and firn core analyses, show that a dz larger than 1 m is strongly related to the vertical heterogeneity of a firn profile and effectively delineates between vertically homogeneous and vertically heterogeneous firn profiles over a depth range of  4 m. Temporal variations in dz align with climatic events and reveal an expansion of heterogeneous firn between 2011–2013 covering an area of  350 815 km2, followed by firn replenishment over the years 2014–2019 spanning an area of  667 725 km2. Our approach reveals the firn evolution of key regions on the Greenland Ice Sheet, providing valuable insights for detecting potential alterations in meltwater runoff patterns.

1 Introduction

The Greenland Ice Sheet (GrIS) is a major contributor to global sea-level rise, where surface runoff currently accounts for about 50 % of the total GrIS mass loss (The IMBIE Team et al., 2020). Firn (compacted snow that has endured at least one melting season) is a key component in the GrIS surface mass balance and currently retains between 41 %–46 % of the surface meltwater (Pfeffer et al., 1991; Harper et al., 2012; Steger et al., 2017). This retention acts as a significant buffer against meltwater runoff and, consequently, sea-level rise. However, impermeable ice slabs and densification of firn restrict meltwater percolation and storage, thereby increasing surface runoff (Machguth et al., 2016; MacFerrin et al., 2019). These ice slabs typically form through the accretion of ice between, at the top, or at the bottom of pre-existing thin ice layers in the firn (Jullien et al., 2023; MacFerrin et al., 2019). Understanding firn properties and the extent of thin ice layers in firn is crucial for assessing current and future GrIS surface mass balance and important for interpreting satellite radar altimetry measurements, which can have ambiguities due to complex near-surface firn stratigraphy (Nilsson et al., 2015). Finally, knowledge of firn density is also required when converting surface height changes from repeat altimetry to mass balance estimates (Sørensen et al., 2011).

Research on firn properties and their impact on meltwater retention and surface mass balance can generally be grouped into three categories: first, detailed in situ measurements at selected sites, focusing on aspects like firn air content and melt features in firn cores (Rennermalm et al., 2021; Machguth et al., 2016; Kameda et al., 1995; Braithwaite et al., 1994), meltwater percolation monitoring via subsurface temperature measurements (Humphrey et al., 2012), time domain reflectometry (Samimi et al., 2021), upward-looking ground-penetrating radar (GPR) measurements (Heilig et al., 2018), or firn stratigraphy from local-scale GPR surveys (Brown et al., 2011). Second, airborne or spaceborne remote sensing techniques have been used to map spatially extensive firn properties. For instance, the ultra-high-frequency (UHF) accumulation radar (AR) deployed during Operation IceBridge (OIB) airborne surveys over the GrIS has been used to directly map ice layers (Culberg et al., 2021), ice slabs (Jullien et al., 2023; MacFerrin et al., 2019), and firn aquifers (Forster et al., 2014; Miège et al., 2016; Horlings et al., 2022) and retrieve past accumulation rates (Karlsson et al., 2016; Lewis et al., 2017). Extensive radar surveys have also been collected with very-high-frequency (VHF) radars over the GrIS; however, these are typically deployed to retrieve ice thickness and englacial layers, whereas their lower bandwidth prevents them from directly resolving the firn stratigraphy. Nonetheless, VHF datasets over Antarctica and the Canadian Arctic have been used to characterise firn properties through statistical analysis of the surface echo strengths (Grima et al., 2014; Rutishauser et al., 2016; Chan et al., 2023). Recent work demonstrated the use of passive microwave satellite measurements to map firn facies and liquid water in firn (Miller et al., 2022a, b; Colliander et al., 2023), while multi-frequency satellite radar can also be used to estimate the firn density and structure (Scanlan et al., 2023). Satellite optical observations have been used to track runoff from the firn area (Tedstone and Machguth, 2022). Lastly, physical or statistical firn models (Langen et al., 2017; Brils et al., 2022; Medley et al., 2022; Thompson-Munson et al., 2023; Vandecrux et al., 2019) are used to bridge the spatial gap between localised measurements. However, statistical models are limited by data availability, and physical models are still unable to capture the complex processes taking place in the firn of ice sheets (Lundin et al., 2017; Vandecrux et al., 2020).

Here, we leverage OIB measurements, specifically the VHF airborne radar (Multichannel Coherent Radar Depth Sounder, MCoRDS) and laser altimetry datasets, to investigate firn properties across the GrIS in the 2011–2019 period (Fig. 1) and to provide insights into how GrIS-wide firn conditions evolve over time. Our approach is based on the hypothesis that a vertically heterogeneous firn structure, resulting from repeated melting and refreezing events, affects the airborne VHF radar signal such that the returned surface reflection appears below the actual ice sheet surface. We first test this hypothesis through numerical modelling of the radar signal, examining how near-surface density heterogeneities impact the MCoRDS surface reflections. We then derive GrIS-wide radar surface reflection offsets via a comparison of the MCoRDS-derived ice sheet surface height to that measured with the Airborne Topographic Mapper (ATM) laser altimeter, assuming that the laser altimeter reveals the true ice sheet surface elevation. Finally, we compare the observed surface return offsets to in situ firn measurements and, together with the modelling results, classify zones of homogeneous and heterogeneous firn and track their evolution over the 2011–2019 period. While extensive UHF radar data are available over Greenland and could technically be used to resolve the firn stratigraphy, our proposed VHF radar-based method is relatively simple and allows for a classification of firn zones without tracing internal layers or structures from the radargram and is transferrable to areas such as Antarctica or other icy planetary surfaces, where no or little UHF radar data are available.

2 Data and methods

2.1 Radar sounding and laser altimetry data

We use data from the Multichannel Coherent Radar Depth Sounder (MCoRDS, versions 2 and 3) and Airborne Topographic Mapper (ATM) laser altimeter, both collected during NASA's OIB surveys over the GrIS between 2011–2019 (Fig. 1). The MCoRDS radar operates at a centre frequency of 195 MHz and has a 30 MHz bandwidth. The radar's surface return signal is affected by dielectric contrasts in the snow, firn, and ice within the depth volume determined by the theoretical vertical range resolution δr, given as

(1) δ r = c k t 2 B ε r ,

where c is the speed of light in vacuum, kt is a windowing factor, B is the bandwidth, and εr is the dielectric permittivity of the probed subsurface. For MCoRDS (kt=1.53, https://data.cresis.ku.edu/data/rds/rds_readme.pdf, last access: 25 March 2024), the theoretical vertical range resolution is 4.3 m in ice (εr=3.15) and 5.7 m in firn (εr=1.8, corresponding to a firn density of 384.4 kg m−3). The MCoRDS radar data have a nominal pulse-limited cross-track resolution (footprint) at the ice sheet surface (Dpl=2cktBR, where R is the aircraft height above the ice sheet) of  180 m, which becomes  25 m after focusing and incoherent summation, and a typical trace spacing of 14–30 m, depending on the survey year (see Table S1 in the Supplement). We extract the radar surface heights from the csarp_standard low-gain data product (Data_img_01), except for the 2011 survey year, where only the csarp_combined data were available.

The ATM laser altimetry data are collected simultaneously with the MCoRDS data. We utilise the ATM Level 2 product, which has been resampled along the tracks and contains three to five platelets spanning the 80 m cross-track swath of the ATM scan (Studinger, 2014).

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f01

Figure 1Map showing the MCoRDS radar profiles between 2011–2019 used to derive surface peak offsets, along with the Surface Mass Balance and Snow Depth on Sea Ice Working Group (SUMup) firn cores from overlapping survey years and the location of selected example traces shown in Fig. 2. Mean annual Modèle Atmosphérique Régional (MAR) surface mass balance (SMB) and melt estimates between 2009–2019 (Fettweis et al., 2017) are used to outline the ablation zone (SMB < 0), the percolation zone (SMB  0 and melt  50 mm w.e. yr−1), and a minimal-melt zone (melt < 50 mm w.e. yr−1), which we hereby refer to as the dry-snow zone. Catchment basins are derived from Mouginot and Rignot (2019).

2.2 Picking of the MCoRDS ice sheet surface reflection

Figure 2 presents examples of MCoRDS surface returns (hereby referred to as the broader surface signal of elevated amplitudes, typically  0.1–0.3 µs wide, and can encompass multiple peaks) of different glacier facies in central west Greenland (locations shown in Fig. 1). Generally, the surface returns from the ablation and dry-snow zones display distinct, narrow single peaks. In contrast, those from the percolation zone feature broader and often multi-peaked returns, complicating the identification of the ice sheet surface. While the Center for Remote Sensing and Integrated Systems (CReSIS) provides ice sheet surface picks along with the radar data, we find that these often correspond to the maximum peak amplitude. However, in instances where a smaller peak or a “bulge” precedes the maximum peak, that first peak likely represents the air–ice sheet interface, with subsequent peaks arising from internal firn density contrasts, such as ice layers (see Sect. 3).

To consistently identify the reflection most representative of the air–ice sheet interface, we apply a custom re-picking algorithm. This algorithm operates on a trace-by-trace basis and involves the following steps: (1) extracting a 40-sample (1.3 µs) vertical window centred around the CReSIS-provided surface pick, (2) upsampling the signal by a factor of 10 via padding of the signal in the frequency domain to enable sub-sample peak location estimates and the identification of bulges that often precede the maximum peak (e.g. Fig. 2c), (3) normalising the trace signal amplitudes between 0 and 1, and (4) selecting the first peak in the upsampled signal that exceeds a 0.1 amplitude threshold (10 % above the noise floor) in both the upsampled and original signals. Applying the 10 % threshold on the original signal minimises the risk of picking numerical artefacts introduced by upsampling, while allowing us to identify bulges that do not appear as peaks in the original MCoRDS signal (e.g. Fig. 2c).

Visual inspections of selected radar profiles confirm that the re-picked surface reflections are reasonable. A vertical picking error of 3.3 ns (approximately one fast-time sample on the upsampled signal) would translate to a surface height offset between 0.28 m (ice) and 0.37 m (firn). To mitigate the impact of outliers arising from potentially misidentified surface peaks, we apply a moving average window over the mean pulse-limited radar footprint diameter ( 180 m before focusing, calculated for each profile) to the re-picked ice sheet surface height.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f02

Figure 2Example of MCoRDS radar traces along profile A (Fig. 7) collected in 2017 (see Fig. 1 for trace location) showing typical surface returns over the ablation zone (a), percolation zone (bd, note the three traces are neighbouring traces and represented with one location marker in Fig. 1), and dry-snow zone (e). The black line is the measured MCoRDS signal, and the red line is the upsampled signal. The dashed black line shows the peak provided by CReSIS, whereas the red dot is the re-picked reflection used as the representation of the ice sheet surface and to calculate the radar peak offsets. The trace energy is normalised, and fast times (travel time) are presented relative to the re-picked surface.

Download

2.3 Derivation and calibration of radar surface peak offsets

To calculate the vertical offsets (dz) between the radar-derived ice sheet surface (hradar) and the actual ice sheet surface as measured by laser altimetry (hlaser), we use

(2) d z = h laser - h radar .

To ensure compatibility, hlaser is calculated as the mean of all laser observations within the radar's pulse-limited footprint.

Once the surface peak offset (dz) is calculated for each radar observation, we apply several exclusion criteria to ensure the best possible quality of the dz dataset: the MEaSUREs Greenland Ice Mapping Project (GIMP; 90 m resolution) ice mask (Howat et al., 2014) is used to remove data points outside ice-covered areas. Data points where the aircraft roll angle exceeded 0.1° are omitted to ensure nadir pointing of both instruments. For the 2011 and 2012 datasets lacking roll angle information, we use the deviation in the heading of 10 km long flight line segments as a roll angle proxy and exclude sections where the change in heading exceeds 0.02°. Locations with fewer than 10 laser points within the radar footprint are also excluded, as are data points where the aircraft's range to the ice sheet surface is below 400 m (due to interference/cut-off of the ice sheet surface reflection). Finally, a few poor-quality profiles were identified through visual inspection of the radargrams and the derived peak offsets (see Table S1). After all filtering, approximately 15.5 million dz observations for the 2011–2019 period remain.

The modelling results (see Sect. 3) suggest that radar- and laser-derived surfaces should align (dz=0) in the ablation zone, where the primary dielectric contrast is the air–ice transition. However, we observed systematic offsets between hlaser and hradar even over the ablation zone in some years and on some survey days. These could be caused by uncorrected timing issues (i.e. cable delays) in the radar measurements. To ensure consistency between measurements from different seasons, we defined a calibration area in the ablation zone with ample data from all seasons (Fig. S1 in the Supplement). The calibration zone started 20 km east from the western ice margin to avoid the steepest and most crevassed part of the ice and extended to the maximum end-of-summer snowline elevation between 2009–2018 (Fausto et al., 2018). For each season, and specific survey days when necessary, we adjusted all data by subtracting the median dz calculated over the calibration area (Table S1, Fig. S2).

After calibration, we remove physically unreasonable dz values that fall outside the 2–98 percentile range for each survey year, as these likely result from incorrect surface reflection picks. Finally, to analyse large-scale trends, dz data points are averaged onto a 10×10 km grid over the GrIS, where grid gaps smaller than 30 km are filled using linear interpolation.

2.4 Numerical modelling of the radar surface reflection

The radar return from a simple two-layer medium, such as the atmosphere over bare ice, typically consists of a single peak, with the peak centred at the interface between the two layers. However, over a medium with multiple interfaces within the radar range resolution depth – like firn containing several ice layers – the radar surface reflection is more complex. To investigate how firn properties affect the MCoRDS radar surface return, we employ RadSPy (https://github.com/scourvil/RadSPy, last access: 25 March 2024), a radar sounding simulator in Python (Courville and Perry, 2021; Courville et al., 2021). RadSPy is an open-source, one-dimensional electromagnetic wave forward-modelling software originally developed to simulate radar reflections observed with the Mars Reconnaissance Orbiter's Shallow Radar (SHARAD) (Courville and Perry, 2021). RadSPy assumes that the N-layered input model consists of flat layers without any roughness, a normally incident plane wave, and no dispersion of the radar signal through the media. To simulate the MCoRDS waveform, we use a 1 µs long chirp swept over the 180–210 MHz MCoRDS frequency range following Eq. 15 in Courville et al. (2021). We note that the vertical range resolution of the modelled signal, expressed as the pulse width 3 dB down from the peak (similar to the calculation of the MCoRDS windowing factor, https://data.cresis.ku.edu/data/rds/rds_readme.pdf, last access: 25 March 2024), is  8.4 m in ice and  11 m in firn. The larger range resolution is likely due to different windowing factors applied in the pulse compression (here we use the standard RadSPy Hanning window) and the input pulse not being an exact replica of the MCoRDS pulse. However, we expect the general behaviour of the modelled signal to be representative of the MCoRDS signal, where depths of layer interfaces should be considered in a relative sense to the different theoretical range resolutions. We employed various input models to represent different glacier profiles as summarised in Table 1. For the ablation zone, the model domain comprises two half-spaces – atmosphere and ice. Additionally, we added a snow layer of varying thickness (up to 2 m) to the ablation zone model (three-layer model), using a density of 341 kg m−3, corresponding to the mean density measured in the top 0.5 m across the GrIS (Fausto et al., 2018). In the dry-snow zone, we use a multi-layer model consisting of the atmosphere above an idealised ice-free firn density–depth profile (ρfirn=320.6 kg m-3+15.4 d, Fig. S3), which is constructed from in situ firn cores (see Sect. 2.5). The input profile representing the firn layer is segmented into 5 cm thick layers, where each layer contains the bulk density for the given depth range. For the percolation zone, we introduce single or multiple (evenly spaced) ice layers (density of 862 kg m−3 (Rennermalm et al., 2021)) at varying depths and thicknesses into the ice-free density profile.

The input density profiles (ρ in kg m−3) are converted to dielectric permittivity (ε) profiles using the empirical model by Kovacs et al. (1995):

(3) ε = 1 + 0.845 × 10 - 3 ρ 2 .

The RadSPy-modelled signals are normalised to their maximum amplitude, and the ice sheet surface reflection (representing the atmosphere–ice/snow/firn interface) is identified as the first peak exceeding a threshold of 10 % of the maximum amplitude, similar to our approach with MCoRDS data (see Sect. 2.2). We note that while we expect changes in reflection amplitude for different firn profiles, here we only focus on the vertical (fast-time) position of the surface peak and not the power itself.

Finally, the RadSPy-modelled signal has a higher sampling rate (1.2 ns) than the MCoRDS signal (33.3 ns). To assess the impacts of waveform sampling on the resultant peak offsets, we also examine how downsampling the modelled waveform influences the vertical position of the surface reflection peak.

Table 1Summary of the different modelling experiments representative of different glacier profiles.

Dielectric permittivities and densities used are a εr_air=1, b εr_ice=3.15, c εr_snow=1.7 (corresponding to 341 kg m−3; Fausto et al., 2018), d firn is an idealised ice-free density–depth profile following ρfirn=320.6 kg m-3+15.4 d (Fig. S3), and e εr_ice_layer=2.99 (corresponding to 862 kg m−3; Rennermalm et al., 2021).

Download Print Version | Download XLSX

2.5 SUMup firn cores

We use firn core density and stratigraphy from the Surface Mass Balance and Snow Depth on Sea Ice Working Group (SUMup) snow density subdataset, Greenland and Antarctica, 1952–2019 (Montgomery et al., 2018; Thompson-Munson et al., 2022).

We first use 401 SUMup firn cores collected between 2011–2019 to create a realistic ice-free firn density profile (Fig. S3). This profile also serves as a baseline to which ice layers can be added to simulate different firn scenarios in our waveform modelling. The ice-free firn density profile is derived for the top 10 m by fitting a linear regression to density values below 750 kg m−3, a threshold well below the typical ice layer density of 862 kg m−3 (Rennermalm et al., 2021).

Secondly, we evaluate the MCoRDS-derived surface peak offsets dz with 29 SUMup cores (Fig. 1) that are (i) collected within 5 km and the same year as the radar measurements and are (ii) at least 5.7 m deep, which is the theoretical depth sensitivity of the radar surface reflection.

3 Simulated imprints of ice layers on the radar surface reflection

Figure 3 presents RadSPy-simulated waveforms alongside observed MCoRDS radar traces (the same as in Fig. 2a, b, e) for density profiles typical of the ablation, percolation, and dry-snow zones. For both the ablation (Fig. 3b) and the dry-snow zones (Fig. 3d), the simulated surface signals exhibit a narrow surface return peak (0.15 µs peak width) at 0 µs. This aligns with the expected timing of the surface echo and confirms that, in the absence of subsurface density variations, the radar surface reflection accurately represents the air–firn or air–ice interface (dz=0). These findings are also in agreement with earlier studies that focused on the returned power of surface reflections (Grima et al., 2014).

Introducing a snow layer into the ablation zone model (Fig. 3e) generally has a minimal effect on the peak offset dz for most snow depths below 1 m but shows a trend of increasing dz with increasing snow depth (Fig. 3f). However, specific snow layer thicknesses – approximately 0.3, 0.9, and 1.5 m – introduce peak offsets of the order of ±5 m. These thicknesses correspond to odd multiples of a quarter MCoRDS wavelength in snow (wavelength of  1.2 m), suggesting that the observed peak offsets result from constructive interference between the air–snow and snow–ice interfaces. Indeed, the modelled waveforms reveal a double peak in the surface return or a bulge preceding the main peak around these particular snow layer thicknesses (Fig. 3g). Here, the primary peak (or bulge) likely corresponds to the air–snow interface, and the secondary peak is linked to the snow–ice interface. This waveform pattern affects the vertical positioning of the peak identified as the surface reflection such that the picked reflection is shifted upwards (negative fast time/dz) when the returned waveform displays two distinct peaks and downwards (positive time/dz) when there is a singular but broadened peak.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f03

Figure 3RadSPy model input depth–density profiles, representing the ablation (blue), percolation (teal), and dry-snow zones (brown). (b–c) RadSPy-modelled signals for the different input stratigraphies. The coloured curves represent the modelled signal, while the black curves represent typical MCoRDS traces observed over the different firn facies (Fig. 2). (e) RadSPy model input depth–density profile representing a snow layer on top of the ablation zone. (f) Modelled surface peak offset over the ablation zone as a function of the snow layer thickness. The x axis on the top shows the snow layer thickness scaled by the range resolution δr of the modelled signal in snow ( 11 m). (g) Examples of modelled waveforms for selected snow layer thicknesses (marked with crosses in f).

Download

For density profiles representative of the percolation zone, the vertical position (i.e. fast time) of the surface reflection peak varies significantly, with dz values ranging from 6.5 to 13.5 m (Figs. 4 and 5). Figure 4 shows how dz is influenced by the changing depth and thickness of a single ice layer. While dz generally increases with increasing ice layer depth and thickness (Fig. 4a, d, and g), it also oscillates at regular intervals, indicating a complex and non-linear relationship between these parameters and dz. Inspecting individual waveforms (Fig. 4e, f, i, and h) shows that the distinct negative and positive dz values are associated with the presence of a double peak (negative fast time/dz) and bulge (positive fast time/dz) preceding the main peak, respectively. Similar to the case for a snow layer in the ablation zone, the primary bulge/peak is likely associated with the air–firn transition and the secondary peak with the firn–ice layer interface. Here, the ice layer itself presents two interfaces (top and bottom), which both contribute to the surface return signal.

When an ice layer lies at depths greater than  0.7δr, thus  4 m for the MCoRDS signal in firn, the radar's surface return displays two distinct peaks. In this scenario, the secondary peak – attributed to the interface between the firn and the ice layer – no longer influences the primary peak, which represents the air–firn transition (light teal waveform in Fig. 4f). Thus, ice layers located deeper than  0.7δr have a negligible impact on the surface reflection peaks, leading to dz 0. Similarly, if an ice layer is sufficiently thick and positioned such that the lower boundary is deeper than  0.6δr, the bottom interface of the ice layer does not affect the surface signal (light teal waveform in Fig. 4i). However, in such cases, the surface reflection peak is still influenced by the upper firn–ice layer interface.

Simulations with a layer with densities of 500 and 600 kg m−3 (Figs. 4d and g, S4) reveal much smaller peak offset (e.g. 2.4 to 1.7 m for a 0.1 m thick 600 kg m−3 layer) compared to those with a typical ice layer density of 862 kg m−3 (6.5 to 8.3 m). This suggests that the highest dz values only occur when ice layers are present in the firn. Figure 5 shows the modelling results over a firn structure with multiple, evenly spaced ice layers of varying thicknesses. Here, dz tends to be the highest when firn contains 3–5 ice layers (i.e. 6–10 strong density contrasts); however, dz is also dependent on the thickness of the ice layers (Fig. 5a).

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f04

Figure 4(a) Surface peak offset (dz) for RadSPy-simulated MCoRDS surface returns over firn stratigraphies consisting of a single ice layer placed at various depths and with different thicknesses. Panels (b) and (c) show example model input profiles. Panel (d) shows the dz for a 0.1 m thick ice layer, as well as layers with densities of 500 and 600 kg m−3 at different depths, with waveforms for selected ice layer depths (marked with crosses) shown in (e) and (f). Panel (g) shows the dz for an ice layer, as well as layers with densities of 500 and 600 kg m−3, starting at 1 m depth and with increasing layer thickness, with waveforms for selected ice layer thicknesses (marked with crosses) shown in (h) and (i). The dotted lines in the waveform plots show the picked peak identified used to calculate dz. The x axes on the top of panels (d) and (g) show the ice layer depth and thickness scaled by the range resolution δr of the modelled signal in firn ( 11 m).

Download

Our modelling experiments show that over vertically homogeneous stratigraphies, such as in the ablation and dry-snow zones, the surface reflection peak accurately returns the true ice sheet surface (dz≈0). Conversely, over vertically heterogeneous stratigraphies, the surface reflection peak often deviates from the true ice sheet surface. While positive dz values are more common, specific combinations of ice layer depths and thicknesses can result in negative dz values. Overall, these results confirm the use of non-zero peak-offsets (dz≠0) in identifying areas with a vertically heterogeneous firn stratigraphy over approximately the theoretical range resolution depth. However, it is important to note that the same dz value can be produced by multiple heterogeneous firn stratigraphies, limiting the ability to infer specific firn characteristics based solely on dz.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f05

Figure 5(a) Surface peak offset (dz) for RadSPy-simulated MCoRDS surface returns over firn stratigraphies with multiple ice layers of varying thickness. (b) Two example input models and (c) the resulting simulated waveforms.

Download

3.1 Effects of waveform sampling

The shape of the waveform measured by MCoRDS is influenced by the radar's sampling rate. To assess its impact on the identification of the surface reflections and the resulting peak offsets (dz), we conducted tests using downsampled RadSPy-simulated waveforms in two scenarios: (i) a waveform with a single peak originating from a homogeneous subsurface (Fig. 6a) and (ii) a waveform with a double peak, generated by a firn profile containing a single ice layer (Fig. 6d). These RadSPy waveforms are downsampled to match the MCoRDS sampling rate (33.3 ns) but with varying time-zero offsets (0–33 ns) to simulate different distances between the radar receiver and the ice sheet surface.

For waveforms with a single peak, downsampling produces only minor variations in the waveform shape. When employing our upsampling approach to identify the surface reflection peak, the location of the identified peak shows minimal fluctuation, with dz offsets ranging between 0–0.5 m (Fig. 6a–c). In contrast, waveforms with double peaks may not be fully resolved at the MCoRDS sampling rate, making them irrecoverable even with the upsampling approach (Fig. 6f). In such instances, the peak offset dz fluctuates around ±6 m, where negative values occur when both peaks are captured and positive values arise when the first peak is missed during sampling. These examples suggest that any bulges observed at the leading edge of a peak reflection in the MCoRDS signal (e.g. Fig. 2c) likely represent the air–ice sheet interface. Therefore, using an upsampling approach based on a Fourier method that introduces oscillations that recovers closely spaced waveforms is a reasonable strategy for picking the first return. However, we note that depending on the relative position of the signal peak, this phenomenon can cause some uncertainties in dz.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f06

Figure 6Effects of downsampled signals on the MCoRDS sampling rate (sra) with varying time-zero offsets (0 ns a, d, 5 ns b, e, 15 ns c, f). Panels (a)(c) show the RadSPy-modelled waveform over the ablation zone (air–ice), and (d)(f) show the modelled signal over the percolation zone including a 10 cm thick ice layer at 2.1 m depth. The horizontal lines show which peak was identified from the modelled signal at the original sampling rate (black line) and downsampled signals (dashed red line).

Download

4 Observed radar surface peak offsets

4.1 Spatial pattern of surface peak offsets and surface reflection characteristics

Figure 7 presents yearly gridded radar surface peak offsets (dz) over the GrIS between 2011–2019 (see Fig. S5 for non-gridded dz along profile lines). While there are some annual variations (discussed in Sect. 4.3), the general spatial pattern of dz is consistent over all survey years. Specifically, low absolute dz values (< 1 m) are observed near the ice sheet margin and over the high-elevation interior, while a band of increased dz (> 1 m) is evident at mid-range elevations. The transition from low to high dz is relatively abrupt at lower elevations, aligning closely with the onset of the percolation facies, which here is derived where the mean 2009–2019 Modèle Atmosphérique Régional (MAR; Fettweis et al., 2017) surface mass balance (SMB) equals zero (dashed lines in Fig. 7). In the northern and northeastern regions, this transition lies at elevations slightly below the percolation facies. The transition from high to low dz at the upper percolation zone is more gradual and spreads across the boundary to the dry-snow facies. The dry-snow facies is derived where the mean 2009–2019 MAR melt equals 50 mm w.e. yr−1, outlining areas that receive little melting. Using all data from 2011–2019, we find that the median dz values (± its standard deviation) in the ablation, percolation, and dry-snow zones are 0.1±1.9, 2.2±2.7, and 0.3±1.2 m, respectively (Fig. S6).

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f07

Figure 7Maps showing the radar surface peak offsets (dz) over the GrIS between 2011–2019. For improved visualisation, the data have been gridded onto a 10×10 km grid, with data interpolated over empty grid cells less than 30 km from the nearest grid point with data (see Fig. S5 for non-gridded dz values along the radar profiles). The dark orange contour line represents dz of 2 m. The dashed black line indicates the boundary between the ablation and accumulation zones (derived from MAR SMB), and the dotted black line depicts the dry-snow facies (derived from MAR melt). The thin black lines are elevation contour lines at a 500 m interval.

Figure 8 shows dz along an MCoRDS profile (profile A) collected in western Greenland in 2017 (see Fig. 7e for profile location), in conjunction with accumulation radar (AR) data. Lower dz values over the ablation and dry-snow zones correspond to clear, narrow surface return peaks in the MCoRDS signal (Fig. 8d). Notably, the increase in dz at the transition between the ablation and percolation zones coincides with the upper elevation range of previously mapped ice slabs (Jullien et al., 2023; MacFerrin et al., 2019) (Figs. 8a, S5). A subsequent comparison reveals that dz decreases as ice slab thickness increases, stabilising at zero for ice slabs thicker than  8 m (Fig. S7).

The AR radargram in the upper dry-snow zone displays numerous bright, continuous reflectors, likely representing isochronous layers of past summer surfaces (Lewis et al., 2017). In contrast, the lower percolation facies features more diffuse reflectors (Fig. 8b), likely a result of annual surface melting causing the coalescence of individual ice layers, ice pipes between ice layers (Humphrey et al., 2012), and an overall higher firn ice content hampering the radar detection of annual layers (Culberg et al., 2021). These characteristics are mirrored in the MCoRDS surface signal (Fig. 8c), where the surface reflection broadens (in fast time) and often splits into multiple peaks (Fig. 8d), resulting in high and spatially variable dz. Figures S8 and S9 present additional profiles (B, C) comparing dz to MCoRDS and AR radargrams. The increased variability of dz can be observed over the entire percolation zone of the GrIS (Fig. S10). Combined with the modelling results suggesting that small changes in ice layer thickness or depth can cause large changes in dz, this indicates that the vertical structure of firn over the percolation zone has a high spatial variability.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f08

Figure 8Radar surface peak offset (dz) over the OIB CReSIS MCoRDS and accumulation radar (AR) profile A in central west Greenland (see Fig. 7e for profile location). (a) Ice sheet surface elevation (black) and dz (raw in light blue and smoothed over a 1 km window in dark blue) along the profile. Vertical dotted black–white lines indicate the firn zone transitions (ablation zone, percolation zone, dry-snow zone) derived from MAR, and the blue-shaded area marks the location of previously mapped ice slabs (Jullien et al., 2023). (b) AR data showing the firn stratigraphy, including ice slabs and isochronous layers of the uppermost 20 m. (c) MCoRDS data of the uppermost 50 m in firn/ice. The radargrams have been flattened with respect to the picked surface reflection. (d) MCoRDS traces along the profile showing the shape of the surface reflection along the profile. The traces are aligned with respect to the picked surface reflection and normalised to the maximum energy.

Download

In the dry-snow zone, a few localised areas exhibit elevated dz values that persist over multiple survey years. For instance, increased dz values near the summit region are consistent from 2012–2019 (Fig. 7). However, AR radargrams over these locations do not indicate abrupt changes in the near-surface stratigraphy (Fig. 9, also profile E in Fig. S11). Rather, the localised dz increase seems to correspond to a region where the MCoRDS surface signal consists of a broadened single peak, laterally bounded by areas with separated double peaks.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f09

Figure 9Radar surface peak offset (dz) over the OIB CReSIS MCoRDS and accumulation radar (AR) profile D over the dz anomaly region near the summit (see Fig. 7b for profile location). (a) Ice sheet surface elevation (black) and dz (blue) along the profile. (b) AR data showing the firn stratigraphy of the uppermost 20 m. (c) MCoRDS data of the uppermost 50 m in firn/ice. The radargrams have been flattened with respect to the picked surface reflection. (d) MCoRDS traces along the profile showing the shape of the surface reflection along the profile. The traces are aligned with respect to the picked surface reflection and normalised to the maximum energy.

Download

4.2 Correlation with in situ firn properties

To investigate the relationship between radar surface peak offsets (dz) and in situ firn conditions, we use four variables derived from the 29 spatial and contemporaneous firn cores (Sect. 2.5). These include the mean density; the standard deviation of the density; the number of density contrasts greater than 50 kg m−3; and the total amount of ice content, defined as the length percentage of core sections with density exceeding 862 kg m−3 (Rennermalm et al., 2021). The core stratigraphies used for the comparison are shown in Fig. S12.

The correlation between dz and the mean core density exhibits two opposing trends (Fig. 10a), complicating any direct inversion of dz to firn density. However, dz shows a linear increase with both the standard deviation of the density profile (Fig. 10b) and the number of density contrasts (Fig. 10c). We tested different density contrast thresholds ranging from 5–200 kg m−3 and found that the linear relationship starts to emerge for contrasts above  50 kg m−3 (Fig. S13). These parameters can be viewed as proxies for the vertical heterogeneity of the firn and thus support the hypothesis that high dz values are indicative of vertically heterogeneous firn profiles.

No clear correlation is found between dz and the firn ice content (Fig. 10d). While firn cores lacking any ice (all recorded densities < 862 kg m−3) generally show lower dz values, high ice content levels (> 10 %) typically result in dz values ranging from  1.3–4 m. However, a few exceptions occur where dz is high even for a 0 % ice content (Fig. S12). These firn core stratigraphies typically show a few strong density contrasts, which may cause the elevated dz.

Overall, low dz values appear to correspond with relatively uniform vertical firn (or ice) stratigraphy, while high dz values are associated with vertically heterogeneous firn profiles. However, the non-uniqueness of dz in relation to firn density and ice content makes it challenging to quantitatively relate dz to specific firn properties.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f10

Figure 10Radar surface peak offset (dz) versus firn core properties (extending from surface to 5.7 m depth): (a) mean density, (b) the standard deviation of the firn density, (c) the number of density contrasts above 50 kg m−3, and (d) the firn ice content (calculated as the length percentage of firn density exceeding 862 kg m−3). The blue markers indicate firn cores that contain ice layers.

Download

4.3 Response of radar peak offset to inter-annual climate variability

To analyse temporal variations in the near-surface facies, we established an empirical dz boundary of 1 m based on our modelling tests and firn core comparisons. Specifically, dz<1 m indicates a vertically homogeneous structure, such as massive ice slabs or firn in the dry-snow zone, while dz>1 m suggests a vertically heterogeneous structure, typically firn with single or multiple ice layers in the percolation zone. The outcomes of this classification and the most significant firn changes are presented in Fig. 11.

Between 2011–2013, areas of heterogeneous firn (dz>1 m) expanded upwards in elevations across all sectors of the GrIS, reaching their maximum extent in 2013 (Figs. 11 and 12). Particularly over the period from 2011–2012, dz values increased in the percolation zone in the CW region (Figs. 12 and S12). We note that in the southern region, the year of maximum heterogeneous extent is less certain due to limited data coverage in 2012/13 but was no later than 2014. During the same period (2011–2013), the lower boundary separating homogeneous and heterogeneous facies near the ablation–accumulation zone transition also shifted higher in elevation, especially in the northwestern regions ( 150 m in CW and  300 m in NW and NO, Fig. 12). Furthermore, we note that in the year 2013, the spatial extent of the dz-derived heterogeneous firn area matches well with the locations where the 2012 melt layer has been mapped in the firn (using a layer prominence > 0.5 (Culberg et al., 2021)) (Fig. 11c).

Comparing the minimal extent of dz>1 m in 2011 to its maximum extent over the 2012–2014 period, a total area of 350 815 km2 located in the upper percolation and lower dry-snow zones switched from homogeneous firn to heterogeneous firn after the spring of 2011 (Fig. 11g). This expansion of heterogeneous firn follows years where the ice sheet experienced relatively high surface melting, with the extreme melt event in the summer of 2012 (Fig. 12).

Between 2013 and 2014, areas of heterogeneous firn in the dry-snow zone reverted back to homogeneous firn (Fig. 12d), aligning with a year of reduced melting in 2013. This trend continued through 2019, with additional areas in the upper percolation zone reverting to homogeneous firn. Absolute dz values generally decreased after 2013 but saw an intermittent increase in the southwestern regions from 2014 to 2017, corresponding to relatively high melt years in 2015 and 2016, followed by another decrease from 2017–2019 (Fig. S14).

Comparing the maximum extent of dz>1 m during 2012–2014 to its minimum extent during 2017–2019, a total area of 667 725 km2 located in the upper percolation and dry-snow zones transitioned from heterogeneous to homogeneous (Fig. 11h). This change was most pronounced in the southern region and followed years of generally reduced surface melting (Fig. 12).

Repeat firn cores collected from 2015–2017 at NASA Southeast (NSE) and Saddle (SDL; locations in Fig. 11e) offer insights into in situ firn changes where dz-classified firn transitioned from heterogeneous to homogeneous between 2014–2019. Both locations exhibited a decline in the standard deviation of firn density variability between 2016–2017, indicating a shift towards more homogeneous vertical firn profiles (Figs. S15 and S16).

Lastly, while heterogeneous firn continued to retreat in the southwestern regions between 2017–2019, it expanded in the northeastern (NE) sector (Fig. 11e and f), coinciding with a slight positive melt anomaly in 2017 (Fig. 12). This expansion of heterogeneous firn in NE Greenland aligns with a reported increase in near-surface densities in 2018 (Scanlan et al., 2023).

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f11

Figure 11(a–f) Qualitative classification of radar surface peak offset (dz) into homogeneous firn/ice (dz<1 m, blue) and heterogeneous firn (dz>1 m, orange) areas. The dark-blue- and red-shaded areas represent the change in classification from heterogeneous to homogeneous and vice versa when compared to the previous survey year. The yellow dots in (c) show locations where the ice layer prominence is above 0.5, indicating the ice layer formed by the 2012 extreme melt likely occurs in the firn (Culberg et al., 2021). The triangles in (e) show the firn core locations at NASA Southeast (NSE) and Saddle (SDL). (g) Change in dz classification between 2011–2012/13/14, highlighting the switch from homogeneous to heterogeneous firn in the upper percolation zone/dry-snow zone (red). The years 2012–2014 are combined to represent the maximum extent of heterogeneous firn. (h) Change in dz classification between 2012/13/14 and 2017/19, highlighting the switch from heterogeneous to homogeneous firn in the upper percolation zone/dry-snow zone (blue). The 2017/19 data are combined to represent the minimum extent of heterogeneous firn.

https://tc.copernicus.org/articles/18/2455/2024/tc-18-2455-2024-f12

Figure 12(a) Spatio-temporal change in radar surface peak offset (dz) between 2011–2019. The dz values are binned into 50 m elevation bands separated by region, where the colour represents the mean dz value in the elevation bin measured in a specific year. (b) MAR-derived yearly melt anomalies from the mean annual melt rate between 2009–2018, binned into elevations that are the same as on the left.

Download

5 Discussion

Through the combination of modelling and observations, we have shown that it is possible to utilise the spatial distribution of surface reflection peak offsets (dz) to effectively delineate between vertically homogeneous and heterogeneous firn profiles across the GrIS. Our approach provides a rapid and consistent method for identifying firn influenced by surface melting and refreezing processes.

The results reveal an upward expansion of heterogeneous firn into the percolation and dry-snow zones between 2011–2013, followed by a reduction in heterogeneous firn between 2013–2019. The expansion of heterogeneous firn in the percolation and dry-snow zones between 2011–2013 is attributed to the formation of new ice layers within formerly homogeneous firn, following surface melt and refreeze events during the summers of 2011 and 2012. The good spatial alignment between the 2012 melt layer, which was typically found at 1 m depth in 2013 (Culberg et al., 2021), and the heterogeneous firn region identified in 2013 further substantiate that dz is sensitive to the presence of ice layers.

The switch of heterogeneous firn back to homogeneous firn over the 2014–2019 period aligns with localised firn pore space replenishment observations from firn core measurements (Rennermalm et al., 2021). These firn cores, collected in the percolation zone of CW Greenland, revealed a decline in the number of ice layers in newly formed firn in the top 3.75 m post-2013 (Rennermalm et al., 2021). Similarly, our analysis of the SDL and NSE firn cores indicated a decline in firn profile heterogeneity from 2015–2017. Thus, our findings indicate a widespread manifestation of these replenishment processes, covering at least  667 725 km2. As outlined in prior studies (Rennermalm et al., 2021), this likely stems from low surface melting and high accumulation rates between 2017–2018, culminating in the most positive mass balance anomaly compared to the 2003–2013 period (Sasgen et al., 2020). Furthermore, under such high mass balance conditions, the 2012 melt layer became gradually buried, reaching depths between 3–12 m by 2017 (Culberg et al., 2021), and surpassing 7.5 m in CW Greenland by 2018 (Rennermalm et al., 2021). Collectively, the observed reduction in dz and inferred switch from heterogeneous to homogeneous firn can most likely be attributed to the burial of the 2012 melt layer beyond the MCoRDS-sensitive depth and the overall decrease in melt layers and density contrasts within the firn.

The lower boundary between homogeneous and heterogeneous firn generally aligns with ice slabs around 6–7 m thick. The upward shift of this boundary from 2011 to 2019 implies that thin ice layers in the firn have grown into ice slabs as well as overall thickening of ice slabs as documented by Jullien et al. (2023). As supported by the modelling results, when ice slabs grow to occupy the top  4 m in firn, the radar signal perceives the surface as homogeneous, thereby shifting this boundary towards higher elevations as ice slabs grow with subsequent melt years.

Our findings also have implications for estimates of future ice slab formation and meltwater runoff. The presence of thin ice layers in firn can facilitate the formation of thick, impermeable ice slabs in subsequent melt years (Jullien et al., 2023; MacFerrin et al., 2019). Our results indicate that the 6-year period (2013–2019) with lower surface melting allowed firn in the high percolation zone to recover after the 2012 extreme melt event. If the 2012 extreme melt year was followed by another extreme melt event before the heterogeneous firn could be replenished, the consequent formation of ice slabs would likely have modified meltwater volumes. Thus, outlining regions with pronounced peak offsets (dz), indicating the presence of ice layers in firn, can help identify areas predisposed to future ice slab formation (e.g. dark orange contours for dz>2 m in Fig. 7). Additionally, identifying firn with ice layers is essential for interpreting satellite radar measurements, as internal reflectors can generate ambiguous signals (Nilsson et al., 2015).

Our results also reveal areas with complex surface processes, for example, localised high-dz areas in the dry-snow zone, particularly near the summit of the ice sheet. It is possible that these areas consist of a high small-scale density variability in the near surface (e.g. from wind scour or hoar formation) (Hörhold et al., 2011), causing strong enough density contrasts to affect the radar surface peak. Additionally, increased dz may be caused by increased surface roughness (e.g. from sastrugi), which introduces waveform interferences when the radar signal from facets with different heights is coherently integrated during processing. We note that this observation aligns with unexpectedly high surface densities in the summit region reported by Scanlan et al. (2023). The persistence of this anomalous feature over several years is unclear but may be attributed to the relatively low snowfall in the area, implying that the near surface varies only over long timescales.

While our modelling and observations generally align, some discrepancies exist. For instance, modelling predicts dz=0 over the ablation and dry-snow zones, yet our observations also include non-zero (and negative) values in these areas. Such deviations could stem from factors not accounted for in the model, such as the surface slope (given that the laser is nadir looking versus the fact that the radar records the nearest return, which may result in negative dz values over sloping surfaces), surface roughness, and surface anomalies like crevasses. Additionally, uncertainties in dz may arise from how the MCoRDS signal is sampled and timing issues (i.e. cable delays) in the radar measurements that were not fully identified and corrected in our calibration. We also note that dz represents a bulk near-surface heterogeneity over the MCoRDS surface-return-sensitive depth and cannot be used to identify heterogeneity/ice layers at a specific depth. Further, the non-unique nature of dz across different firn stratigraphies, especially in the percolation zone, limits the extraction of quantitative firn properties such as the density or ice content from dz.

Altogether, our results demonstrate that airborne VHF radar measurements, when combined with laser altimetry, can effectively discern different firn facies. This provides opportunities to assess firn properties in situations where only VHF radar is accessible (i.e. in the absence of UHF radar or in situ observations), particularly in contexts predating Operation IceBridge's accumulation radar. Furthermore, analogous techniques may prove invaluable for upcoming Earth-orbiting radar sounders when combined with existing laser altimetry missions (e.g. ICESat-2), as well as interplanetary missions such as ESA's JUpiter ICy moons Explorer (JUICE) spacecraft (Grasset et al., 2013), which carries both a radar sounder (RIME) and a laser altimeter (GALA). Future work could extend this methodology to higher-frequency ranges, as well as leverage satellite radar and laser altimetry for snowpack property delineation. Further, the OIB MCoRDS data could be reprocessed with a higher vertical sampling rate, which could eliminate the need for our upsampling approach to identify the surface reflection. Finally, we suggest that similar and potentially further firn properties could be derived by examining other surface reflection characteristics, such as the width and peakiness, similar to applications for characterising subglacial environments (e.g. Oswald and Gogineni, 2008; Jordan et al., 2017), which could also eliminate the need for concurrent laser observations.

6 Conclusions

In this study, we combined airborne radar sounding and laser altimetry measurements to derive radar surface reflection peak offsets (dz) over the GrIS. Our results, supported by modelling and in situ firn core analyses, demonstrate that dz serves as an effective tool to delineate between vertically homogeneous and heterogeneous firn profiles of  0–4 m depth, where temporal changes in dz over 2–5 years align well with known climatic events. This allows for a nuanced understanding of spatial and temporal variations in firn stratigraphy from 2011–2019, highlighting firn regions with important implications for meltwater runoff and ice sheet mass balance. For the first time, our results map and quantify areas (667 725 km2) where firn replenishment took place between 2014–2019. This phenomenon was previously observed only in localised in situ measurements, which emphasises the importance of spatially comprehensive observations. Importantly, dz can be used to outline firn areas that are predisposed to future ice slab formation, which has a considerable impact on future firn meltwater storage.

Code and data availability

The MCoRDS and accumulation radar data used in this study are available on the CReSIS public web page (https://data.cresis.ku.edu/, CReSIS, 2024). The IceBridge ATM data are available through NSIDC at https://doi.org/10.5067/CPRXXK3F39RV (Studinger, 2014). SUMup firn density measurements (https://doi.org/10.18739/A2NP1WK6M, Thompson-Munson et al., 2022) were accessed from the Arctic Data Center (https://arcticdata.io/, last access: 25 March 2024). Ice slab data (Jullien et al., 2023) were derived from https://doi.org/10.5281/zenodo.7505426 (Jullien, 2023), firn aquifer data from the Arctic Data Center (https://doi.org/10.18739/A2TM72225, Miège, 2018), and the snowline dataset (Fausto et al., 2018) from https://thredds.geus.dk/thredds/catalog/MODIS_snowline/catalog.html (last access: 25 March 2024). MAR data (MARv3.12_monthly_ERA5) (Fettweis et al., 2017) were accessed from ftp://ftp.climato.be/fettweis (last access: 21 February 2023). GrIS basins (Mouginot and Rignot, 2019) were derived from https://doi.org/10.7280/D1WT11. All data products resulting from this work are available on Dataverse (https://doi.org/10.22008/FK2/OLVPFG, Rutishauser, 2024).

Supplement

The supplement related to this article is available online at: https://doi.org/10.5194/tc-18-2455-2024-supplement.

Author contributions

AR conceptualised the study with inputs from all co-authors and led the methodology development and research investigations. KMS set up the modelling experiments, BV assisted with firn core analyses, and PH assisted with software implementations and data curation. NJ provided ice slab data resources and insights for interpreting the data. AR drafted the original manuscript and visualisations. All co-authors provided critical reviews and contributed to the final paper. APA acquired the funding supporting the study.

Competing interests

At least one of the (co-)authors is a member of the editorial board of The Cryosphere. The peer-review process was guided by an independent editor, and the authors also have no other competing interests to declare.

Disclaimer

Publisher’s note: Copernicus Publications remains neutral with regard to jurisdictional claims made in the text, published maps, institutional affiliations, or any other geographical representation in this paper. While Copernicus Publications makes every effort to include appropriate place names, the final responsibility lies with the authors.

Acknowledgements

We acknowledge the use of data from CReSIS generated with support from the University of Kansas; the NASA Operation IceBridge grant NNX16AH54G; NSF grants ACI-1443054, OPP-1739003, and IIS-1838230; Lilly Endowment Incorporated; and the Indiana METACyt Initiative. We employed OpenAI's GPT-4 language model (ChatGPT-4) for text editing.

Review statement

This paper was edited by Joseph MacGregor and reviewed by Tate Meehan and one anonymous referee.

Financial support

This study and all personnel at the Geological Survey of Denmark and Greenland were supported by the Greenland Climate Network (GC-Net) monitoring programme. Kirk M. Scanlan was supported by Villum Fonden (Villum Experiment Programme; project no. 50225), and Nicolas Jullien was supported by the European Research Council award 818994 (CASSANDRA).

References

Braithwaite, R. J., Laternser, M., and Pfeffer, W. T.: Variations of near-surface firn density in the lower accumulation area of the Greenland ice sheet, Pâkitsoq, West Greenland, J. Glaciol., 40, 477–485, https://doi.org/10.3189/s002214300001234x, 1994. 

Brils, M., Kuipers Munneke, P., van de Berg, W. J., and van den Broeke, M.: Improved representation of the contemporary Greenland ice sheet firn layer by IMAU-FDM v1.2G, Geosci. Model Dev., 15, 7121–7138, https://doi.org/10.5194/gmd-15-7121-2022, 2022. 

Brown, J., Harper, J., Pfeffer, W. T., Humphrey, N., and Bradford, J.: High-resolution study of layering within the percolation and soaked facies of the Greenland ice sheet, Ann. Glaciol., 52, 35–42, https://doi.org/10.3189/172756411799096286, 2011. 

Chan, K., Grima, C., Rutishauser, A., Young, D. A., Culberg, R., and Blankenship, D. D.: Spatial characterization of near-surface structure and meltwater runoff conditions across the Devon Ice Cap from dual-frequency radar reflectivity, The Cryosphere, 17, 1839–1852, https://doi.org/10.5194/tc-17-1839-2023, 2023. 

Colliander, A., Mousavi, M., Kimball, J. S., Miller, J. Z., and Burgin, M.: Spatial and temporal differences in surface and subsurface meltwater distribution over Greenland ice sheet using multi-frequency passive microwave observations, Remote Sens. Environ., 295, 113705, https://doi.org/10.1016/j.rse.2023.113705, 2023. 

Courville, S. W. and Perry, M.: scourvil/RadSPy: v1.0, Zenodo [code], https://doi.org/10.5281/zenodo.4429356, 2021. 

Courville, S. W., Perry, M. R., and Putzig, N. E.: Lower Bounds on the Thickness and Dust Content of Layers within the North Polar Layered Deposits of Mars from Radar Forward Modeling, Planet. Sci. J., 2, 28, https://doi.org/10.3847/psj/abda50, 2021. 

CReSIS: Radar Depth Sounder Data, Lawrence, Kansas, USA, Digital Media, http://data.cresis.ku.edu/ (last access: 25 March 2024), 2024. 

Culberg, R., Schroeder, D. M., and Chu, W.: Extreme melt season ice layers reduce firn permeability across Greenland, Nat. Commun., 12, 2336, https://doi.org/10.1038/s41467-021-22656-5, 2021. 

Fausto, R. S., Box, J. E., Vandecrux, B., van As, D., Steffen, K., Macferrin, M. J., Machguth, H., Colgan, W., Koenig, L. S., McGrath, D., Charalampidis, C., and Braithwaite, R. J.: A snow density dataset for improving surface boundary conditions in Greenland ice sheet firn modeling, Front. Earth Sci., 6, 51, https://doi.org/10.3389/feart.2018.00051, 2018. 

Fettweis, X., Box, J. E., Agosta, C., Amory, C., Kittel, C., Lang, C., van As, D., Machguth, H., and Gallée, H.: Reconstructions of the 1900–2015 Greenland ice sheet surface mass balance using the regional climate MAR model, The Cryosphere, 11, 1015–1033, https://doi.org/10.5194/tc-11-1015-2017, 2017. 

Forster, R. R., Box, J. E., van den Broeke, M. R., Miège, C., Burgess, E. W., van Angelen, J. H., Lenaerts, J. T. M., Koenig, L. S., Paden, J., Lewis, C., Gogineni, S. P., Leuschen, C., and McConnell, J. R.: Extensive liquid meltwater storage in firn within the Greenland ice sheet, Nat. Geosci., 7, 95–98, https://doi.org/10.1038/ngeo2043, 2014. 

Grasset, O., Dougherty, M. K., Coustenis, A., Bunce, E. J., Erd, C., Titov, D., Blanc, M., Coates, A., Drossart, P., Fletcher, L. N., Hussmann, H., Jaumann, R., Krupp, N., Lebreton, J.-P., Prieto-Ballesteros, O., Tortora, P., Tosi, F., and Van Hoolst, T.: JUpiter ICy moons Explorer (JUICE): An ESA mission to orbit Ganymede and to characterise the Jupiter system, Planet. Space Sci., 78, 1–21, 2013. 

Grima, C., Blankenship, D. D., Young, D. A., and Schroeder, D. M.: Surface slope control on firn density at Thwaites Glacier, West Antarctica: Results from airborne radar sounding, Geophys. Res. Lett., 41, 6787–6794, https://doi.org/10.1002/2014GL061635, 2014. 

Harper, J., Humphrey, N., Pfeffer, W. T., Brown, J., and Fettweis, X.: Greenland ice-sheet contribution to sea-level rise buffered by meltwater storage in firn, Nature, 491, 240–243, https://doi.org/10.1038/nature11566, 2012. 

Heilig, A., Eisen, O., MacFerrin, M., Tedesco, M., and Fettweis, X.: Seasonal monitoring of melt and accumulation within the deep percolation zone of the Greenland Ice Sheet and comparison with simulations of regional climate modeling, The Cryosphere, 12, 1851–1866, https://doi.org/10.5194/tc-12-1851-2018, 2018. 

Hörhold, M. W., Kipfstuhl, S., Wilhelms, F., Freitag, J., and Frenzel, A.: The densification of layered polar firn, J. Geophys. Res.-Earth, 116, F01001, https://doi.org/10.1029/2009JF001630, 2011. 

Horlings, A. N., Christianson, K., and Miège, C.: Expansion of Firn Aquifers in Southeast Greenland, J. Geophys. Res.-Earth, 127, e2022JF006753, https://doi.org/10.1029/2022JF006753, 2022. 

Howat, I. M., Negrete, A., and Smith, B. E.: The Greenland Ice Mapping Project (GIMP) land classification and surface elevation data sets, The Cryosphere, 8, 1509–1518, https://doi.org/10.5194/tc-8-1509-2014, 2014. 

Humphrey, N. F., Harper, J. T., and Pfeffer, W. T.: Thermal tracking of meltwater retention in Greenland's accumulation area, J. Geophys. Res.-Earth, 117, 1010, https://doi.org/10.1029/2011JF002083, 2012. 

Jordan, T. M., Cooper, M. A., Schroeder, D. M., Williams, C. N., Paden, J. D., Siegert, M. J., and Bamber, J. L.: Self-affine subglacial roughness: consequences for radar scattering and basal water discrimination in northern Greenland, The Cryosphere, 11, 1247–1264, https://doi.org/10.5194/tc-11-1247-2017, 2017. 

Jullien, N.: Data: Greenland Ice Sheet ice slab expansion and thickening (Version v4), Zenodo [data set], https://doi.org/10.5281/zenodo.7505426, 2023. 

Jullien, N., Tedstone, A. J., Machguth, H., Karlsson, N. B., and Helm, V.: Greenland Ice Sheet Ice Slab Expansion and Thickening, Geophys. Res. Lett., 50, e2022GL100911, https://doi.org/10.1029/2022GL100911, 2023. 

Kameda, T., Narita, H., Shoji, H., Nishio, F., Fujii, Y., and Watanabe, O.: Melt features in ice cores from Site J, southern Greenland: some implications for summer climate since AD 1550, Ann. Glaciol., 21, 51–58, https://doi.org/10.3189/S0260305500015597, 1995. 

Karlsson, N. B., Eisen, O., Dahl-Jensen, D., Freitag, J., Kipfstuhl, S., Lewis, C., Nielsen, L. T., Paden, J. D., Winter, A., and Wilhelms, F.: Accumulation Rates during 1311–2011 CE in North-Central Greenland Derived from Air-Borne Radar Data, Front. Earth Sci., 4, 97, https://doi.org/10.3389/feart.2016.00097, 2016. 

Kovacs, A., Gow, A. J., and Morey, R. M.: The in-situ dielectric constant of polar firn revisited, Cold Reg. Sci. Technol., 23, 245–256, https://doi.org/10.1016/0165-232X(94)00016-Q, 1995. 

Langen, P. L., Fausto, R. S., Vandecrux, B., Mottram, R. H., and Box, J. E.: Liquid water flow and retention on the Greenland ice sheet in the regional climate model HIRHAM5: Local and large-scale impacts, Front. Earth Sci., 4, 110, https://doi.org/10.3389/feart.2016.00110, 2017. 

Lewis, G., Osterberg, E., Hawley, R., Whitmore, B., Marshall, H. P., and Box, J.: Regional Greenland accumulation variability from Operation IceBridge airborne accumulation radar, The Cryosphere, 11, 773–788, https://doi.org/10.5194/tc-11-773-2017, 2017. 

Lundin, J. M. D., Stevens, C. M., Arthern, R., Buizert, C., Orsi, A., Ligtenberg, S. R. M., Simonsen, S. B., Cummings, E., Essery, R., Leahy, W., Harris, P., Helsen, M. M., and Waddington, E. D.: Firn Model Intercomparison Experiment (FirnMICE), J. Glaciol., 63, 401–422, https://doi.org/10.1017/jog.2016.114, 2017. 

MacFerrin, M., Machguth, H., As, D. van, Charalampidis, C., Stevens, C. M., Heilig, A., Vandecrux, B., Langen, P. L., Mottram, R., Fettweis, X., Broeke, M. R. van den, Pfeffer, W. T., Moussavi, M. S., and Abdalati, W.: Rapid expansion of Greenland's low-permeability ice slabs, Nature, 573, 403–407, https://doi.org/10.1038/s41586-019-1550-3, 2019. 

Machguth, H., MacFerrin, M., van As, D., Box, J. E., Charalampidis, C., Colgan, W., Fausto, R. S., Meijer, H. A. J., Mosley-Thompson, E., and van de Wal, R. S. W.: Greenland meltwater storage in firn limited by near-surface ice formation, Nat. Clim. Change, 6, 390–393, https://doi.org/10.1038/nclimate2899, 2016. 

Medley, B., Neumann, T. A., Zwally, H. J., Smith, B. E., and Stevens, C. M.: Simulations of firn processes over the Greenland and Antarctic ice sheets: 1980–2021, The Cryosphere, 16, 3971–4011, https://doi.org/10.5194/tc-16-3971-2022, 2022. 

Miège, C.: Spatial extent of Greenland firn aquifer detected by airborne radars, 2010–2017, Arctic Data Center [data set], https://doi.org/10.18739/A2TM72225, 2018. 

Miège, C., Forster, R. R., Brucker, L., Koenig, L. S., Solomon, D. K., Paden, J. D., Box, J. E., Burgess, E. W., Miller, J. Z., McNerney, L., Brautigam, N., Fausto, R. S., and Gogineni, S.: Spatial extent and temporal variability of Greenland firn aquifers detected by ground and airborne radars, J. Geophys. Res.-Earth, 121, 2381–2398, https://doi.org/10.1002/2016JF003869, 2016. 

Miller, J. Z., Culberg, R., Long, D. G., Shuman, C. A., Schroeder, D. M., and Brodzik, M. J.: An empirical algorithm to map perennial firn aquifers and ice slabs within the Greenland Ice Sheet using satellite L-band microwave radiometry, The Cryosphere, 16, 103–125, https://doi.org/10.5194/tc-16-103-2022, 2022a. 

Miller, J. Z., Long, D. G., Shuman, C. A., Culberg, R., Hardman, M. A., and Brodzik, M. J.: Mapping Firn Saturation Over Greenland Using NASA's Soil Moisture Active Passive Satellite, IEEE J. Select. Top. Appl., 15, 3714–3729, https://doi.org/10.1109/JSTARS.2022.3154968, 2022b. 

Montgomery, L., Koenig, L., and Alexander, P.: The SUMup dataset: compiled measurements of surface mass balance components over ice sheets and sea ice with analysis over Greenland, Earth Syst. Sci. Data, 10, 1959–1985, https://doi.org/10.5194/essd-10-1959-2018, 2018. 

Mouginot, J. and Rignot, E.: Glacier catchments/basins for the Greenland Ice Sheet, Dryad [data set], https://doi.org/10.7280/D1WT11, 2019. 

Nilsson, J., Vallelonga, P., Simonsen, S. B., Sørensen, L. S., Forsberg, R., Dahl-Jensen, D., Hirabayashi, M., Goto-Azuma, K., Hvidberg, C. S., Kjær, H. A., and Satow, K.: Greenland 2012 melt event effects on CryoSat-2 radar altimetry, Geophys. Res. Lett., 42, 3919–3926, https://doi.org/10.1002/2015GL063296, 2015. 

Oswald, G. K. A. and Gogineni, S. P.: Recovery of subglacial water extent from Greenland radar survey data, J. Glaciol., 54, 94–106, https://doi.org/10.3189/002214308784409107, 2008. 

Pfeffer, W. T., Meier, M. F., and Illangasekare, T. H.: Retention of Greenland runoff by refreezing: implications for projected future sea level change, J. Geophys. Res., 96, 22117–22124, https://doi.org/10.1029/91jc02502, 1991. 

Rennermalm, Å. K., Hock, R., Covi, F., Xiao, J., Corti, G., Kingslake, J., Leidman, S. Z., Miège, C., Macferrin, M., Machguth, H., Osterberg, E., Kameda, T., and McConnell, J. R.: Shallow firn cores 1989–2019 in southwest Greenland's percolation zone reveal decreasing density and ice layer thickness after 2012, J. Glaciol., 68, 431–442, https://doi.org/10.1017/jog.2021.102, 2021. 

Rutishauser, A.: Data for “Mapping the vertical heterogeneity of Greenland's firn from 2011–2019 using airborne radar and laser altimetry”, GEUS Dataverse V1 [data set], https://doi.org/10.22008/FK2/OLVPFG, 2024. 

Rutishauser, A., Grima, C., Sharp, M., Blankenship, D. D., Young, D. A., Cawkwell, F., and Dowdeswell, J. A.: Characterizing near-surface firn using the scattered signal component of the glacier surface return from airborne radio-echo sounding, Geophys. Res. Lett., 43, 12502–12510, https://doi.org/10.1002/2016GL071230, 2016. 

Samimi, S., Marshall, S. J., Vandecrux, B., and MacFerrin, M.: Time-Domain Reflectometry Measurements and Modeling of Firn Meltwater Infiltration at DYE-2, Greenland, J. Geophys. Res.-Earth, 126, e2021JF006295, https://doi.org/10.1029/2021JF006295, 2021. 

Sasgen, I., Wouters, B., Gardner, A. S., King, M. D., Tedesco, M., Landerer, F. W., Dahle, C., Save, H., and Fettweis, X.: Return to rapid ice loss in Greenland and record loss in 2019 detected by the GRACE-FO satellites, Commun. Earth Environ., 1, 1–8, https://doi.org/10.1038/s43247-020-0010-1, 2020. 

Scanlan, K. M., Rutishauser, A., and Simonsen, S. B.: Observing the Near-Surface Properties of the Greenland Ice Sheet, Geophys. Res. Lett., 50, e2022GL101702, https://doi.org/10.1029/2022GL101702, 2023. 

Sørensen, L. S., Simonsen, S. B., Nielsen, K., Lucas-Picher, P., Spada, G., Adalgeirsdottir, G., Forsberg, R., and Hvidberg, C. S.: Mass balance of the Greenland ice sheet (2003–2008) from ICESat data – the impact of interpolation, sampling and firn density, The Cryosphere, 5, 173–186, https://doi.org/10.5194/tc-5-173-2011, 2011. 

Steger, C. R., Reijmer, C. H., van den Broeke, M. R., Wever, N., Forster, R. R., Koenig, L. S., Munneke, P. K., Lehning, M., Lhermitte, S., Ligtenberg, S. R. M., Miège, C., and Noël, B. P. Y.: Firn meltwater retention on the greenland ice sheet: A model comparison, Front. Earth Sci., 5, 3, https://doi.org/10.3389/feart.2017.00003, 2017. 

Studinger, M.: IceBridge ATM L2 Icessn Elevation, Slope, and Roughness, Version 2, Boulder, Colorado USA, NASA National Snow and Ice [data set], https://doi.org/10.5067/CPRXXK3F39RV, 2014. 

Tedstone, A. J. and Machguth, H.: Increasing surface runoff from Greenland's firn areas, Nat. Clim. Change, 12, 672–676, https://doi.org/10.1038/S41558-022-01371-Z, 2022. 

The IMBIE Team, Shepherd, A., Ivins, E., Rignot, E., Smith, B., van den Broeke, M., Velicogna, I., Whitehouse, P., Briggs, K., Joughin, I., Krinner, G., Nowicki, S., Payne, T., Scambos, T., Schlegel, N., A, G., Agosta, C., Ahlstrøm, A., Babonis, G., Barletta, V. R., Bjørk, A. A., Blazquez, A., Bonin, J., Colgan, W., Csatho, B., Cullather, R., Engdahl, M. E., Felikson, D., Fettweis, X., Forsberg, R., Hogg, A. E., Gallee, H., Gardner, A., Gilbert, L., Gourmelen, N., Groh, A., Gunter, B., Hanna, E., Harig, C., Helm, V., Horvath, A., Horwath, M., Khan, S., Kjeldsen, K. K., Konrad, H., Langen, P. L., Lecavalier, B., Loomis, B., Luthcke, S., McMillan, M., Melini, D., Mernild, S., Mohajerani, Y., Moore, P., Mottram, R., Mouginot, J., Moyano, G., Muir, A., Nagler, T., Nield, G., Nilsson, J., Noël, B., Otosaka, I., Pattle, M. E., Peltier, W. R., Pie, N., Rietbroek, R., Rott, H., Sandberg Sørensen, L., Sasgen, I., Save, H., Scheuchl, B., Schrama, E., Schröder, L., Seo, K. W., Simonsen, S. B., Slater, T., Spada, G., Sutterley, T., Talpe, M., Tarasov, L., van de Berg, W. J., van der Wal, W., van Wessem, M., Vishwakarma, B. D., Wiese, D., Wilton, D., Wagner, T., Wouters, B., and Wuite, J.: Mass balance of the Greenland Ice Sheet from 1992 to 2018, Nature, 579, 233–239, https://doi.org/10.1038/s41586-019-1855-2, 2020. 

Thompson-Munson, M., Montgomery, L., Lenaerts, J., and Koenig, L.: Surface Mass Balance and Snow Depth on Sea Ice Working Group (SUMup) snow density subdataset, Greenland and Antartica, 1952–2019, Arctic Data Center [data set], https://doi.org/10.18739/A2NP1WK6M, 2022. 

Thompson-Munson, M., Wever, N., Stevens, C. M., Lenaerts, J. T. M., and Medley, B.: An evaluation of a physics-based firn model and a semi-empirical firn model across the Greenland Ice Sheet (1980–2020), The Cryosphere, 17, 2185–2209, https://doi.org/10.5194/tc-17-2185-2023, 2023.  

Vandecrux, B., MacFerrin, M., Machguth, H., Colgan, W. T., van As, D., Heilig, A., Stevens, C. M., Charalampidis, C., Fausto, R. S., Morris, E. M., Mosley-Thompson, E., Koenig, L., Montgomery, L. N., Miège, C., Simonsen, S. B., Ingeman-Nielsen, T., and Box, J. E.: Firn data compilation reveals widespread decrease of firn air content in western Greenland, The Cryosphere, 13, 845–859, https://doi.org/10.5194/tc-13-845-2019, 2019. 

Vandecrux, B., Mottram, R., Langen, P. L., Fausto, R. S., Olesen, M., Stevens, C. M., Verjans, V., Leeson, A., Ligtenberg, S., Kuipers Munneke, P., Marchenko, S., van Pelt, W., Meyer, C. R., Simonsen, S. B., Heilig, A., Samimi, S., Marshall, S., Machguth, H., MacFerrin, M., Niwano, M., Miller, O., Voss, C. I., and Box, J. E.: The firn meltwater Retention Model Intercomparison Project (RetMIP): evaluation of nine firn models at four weather station sites on the Greenland ice sheet, The Cryosphere, 14, 3785–3810, https://doi.org/10.5194/tc-14-3785-2020, 2020. 

Download
Short summary
The Greenland Ice Sheet interior is covered by a layer of firn, which is important for surface meltwater runoff and contributions to global sea-level rise. Here, we combine airborne radar sounding and laser altimetry measurements to delineate vertically homogeneous and heterogeneous firn. Our results reveal changes in firn between 2011–2019, aligning well with known climatic events. This approach can be used to outline firn areas primed for significantly changing future meltwater runoff.