Hostname: page-component-7bb8b95d7b-lvwk9 Total loading time: 0 Render date: 2024-10-04T19:26:07.417Z Has data issue: false hasContentIssue false

Geomorphological and historical records of the surge-type behaviour of Hansbreen (Svalbard)

Published online by Cambridge University Press:  02 October 2024

Aleksandra Osika*
Affiliation:
Faculty of Natural Sciences, University of Silesia, Institute of Earth Sciences, Bedzinska 60, 41-200 Sosnowiec, Poland
Jacek Jania
Affiliation:
Faculty of Natural Sciences, University of Silesia, Institute of Earth Sciences, Bedzinska 60, 41-200 Sosnowiec, Poland
*
Corresponding author: Aleksandra Osika; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

This paper presents geomorphological and historical records of the surge-type behaviour of Hansbreen, one of the most studied tidewater glaciers in Svalbard. The surge-type behaviour of the glacier has not been considered before due to the lack of evidence of this phenomenon. We integrate geomorphological mapping of the terrestrial and submarine forefields with historical data from the 19th and 20th centuries to reconstruct the glacier dynamics and identify the possible timing of surging. Landform assemblages are representative of the surging glacier landsystem, including crevasse-squeeze ridges (CSRs) and submarine streamlined glacial lineations. Abundant CSRs in the outer part of the terrestrial forefield were also documented in the 1980s, but most have been obliterated since then. We suggest the identified surge landsystem was produced during a surge of Hansbreen detected from photographs taken during the Austro-Hungarian expedition in 1872. Historical photogrammetric photos from the Norwegian expedition in 1918 revealed surge-diagnostic features in the glacier surface, including a folded medial moraine and a dense, complex network of crevasses. A potential next surge remains questionable in the following decades due to the low-lying accumulation area of the main stream hindering the mass build-up, but potential surges of the tributary glaciers should not be excluded.

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, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press on behalf of International Glaciological Society

Introduction

Surge-type glaciers exhibit multiyear, quasi-cyclical velocity fluctuations between a prolonged quiescent phase of slow flow and a shorter active phase of fast flow, during which ice velocities can increase by 10–1000 times (Meier and Post, Reference Meier and Post1969; Raymond, Reference Raymond1987; Sund and others, Reference Sund, Eiken, Hagen and Kääb2009). Most surge-type glaciers concentrate in spatial clusters within an optimal climatic envelope, such as Svalbard (74–81°N) (Sevestre and Benn, Reference Sevestre and Benn2015). Glaciers cover 57% of the land of this archipelago and more than 60% of the glaciated area is represented by tidewater glaciers (Błaszczyk and others, Reference Błaszczyk, Jania and Hagen2009; Nuth and others, Reference Nuth2013). There is general agreement that surging is a common phenomenon in Svalbard (e.g. Liestøl, Reference Liestøl1969; Jania, Reference Jania1988; Lefauconnier and Hagen, Reference Lefauconnier and Hagen1991; Hagen and others, Reference Hagen, Liestøl, Roland and Jørgensen1993; Sevestre and Benn, Reference Sevestre and Benn2015; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016; Benn and others, Reference Benn, Hewitt and Luckman2023), although the estimated percentage of surge-type glaciers varies from 13% (Jiskoot and others, Reference Jiskoot, Boyle and Murray1998) to 90% (Hagen and Liestøl, Reference Hagen and Liestøl1990). The duration of the quiescent phase typically ranges from 30 to 150 years, but can be as long as 500 years (Dowdeswell and others, Reference Dowdeswell, Hamilton and Hagen1991; Hagen and others, Reference Hagen, Liestøl, Roland and Jørgensen1993; Murray and others, Reference Murray, Strozzi, Luckman, Jiskoot and Christakos2003). Therefore, the length of surge cycles often exceeds the observation period, which poses a risk of misclassification of glaciers as non-surging if the active phase has not been observed and associated geomorphological and glaciological records of surging have been obliterated since then (Benn and Evans, Reference Benn and Evans2010).

The frequency of surges is influenced by climatic conditions through changes in the mass balance and modification of the thermal structure of glaciers (Dowdeswell and others, Reference Dowdeswell, Hodgkins, Nuttall, Hagen and Hamilton1995; Sevestre and others, Reference Sevestre, Benn, Hulton and Bælum2015). In recent centuries, Svalbard glaciers reached their maximum extent during the Little Ice Age (LIA) and many of these advances were surge-related (e.g. Liestøl, Reference Liestøl1969; Lefauconnier and Hagen, Reference Lefauconnier and Hagen1991; Christoffersen and others, Reference Christoffersen, Piotrowski and Larsen2005; Sevestre and others, Reference Sevestre, Benn, Hulton and Bælum2015; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016; Zagórski and others, Reference Zagórski2023). The main sources of data about the timing of the late LIA surges are historical observations of glaciers that have recently advanced (e.g. the surge of Kongsvegen dated to c. 1800; Liestøl, Reference Liestøl1988) or evidence of glaciers experiencing the active phase of a surge as recorded in historical drawings and maps (e.g. Recherchebreen in 1838; Liestøl, Reference Liestøl, Williams and Ferrigno1993, Kongsvegen in 1869; Liestøl, Reference Liestøl1988, Nathorstbreen and Paulabreen mapped in 1898; Ottesen and others, Reference Ottesen2008). Although the phenomenon of glacier surges was unknown at that time, this unique photographic and cartographic documentation can be reinterpreted to identify the evidence for past surge events (e.g. Zagórski and others, Reference Zagórski2023). However, in regions with less archival data on glacier fluctuations during the LIA, such as Hornsund, most surges have only been recorded in the second half of the 20th and in the 21st century (Błaszczyk and others, Reference Błaszczyk, Jania and Kolondra2013). The first maps presenting glacier termini in this area were prepared during the Austro-Hungarian North Pole Expedition 1872–1874 (Peterman, Reference Peterman1874) and the Swedish-Russian Arc-of-Meridian Expedition to Spitsbergen 1899–1902 (Wassiliew, Reference Wassiliew1925) when the glaciers were already extensive or even at their LIA maxima. The scales of these maps and the simplified cartographic representation of glaciers on the older map do not allow the identification of evidence of ongoing surges, such as extensive crevassing or folded medial moraines (Copland and others, Reference Copland, Sharp and Dowdeswell2003). Therefore, some surges may have occurred in this area during the LIA that are not documented in existing inventories.

In the case of sparse archival data, the application of surging glacier landsystem models based on characteristic landform assemblages in the terrestrial (Evans and Rea, Reference Evans and Rea1999, Reference Evans, Rea and Evans2003) and submarine forefields (Ottesen and Dowdeswell, Reference Ottesen and Dowdeswell2006; Ottesen and others, Reference Ottesen2008, Reference Ottesen, Dowdeswell, Bellec and Bjarnadóttir2017) allows the detection of surge-like behaviour of glaciers with no documented surging history (Christoffersen and others, Reference Christoffersen, Piotrowski and Larsen2005; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016, Reference Farnsworth2017; Flink and Noormets, Reference Flink and Noormets2018). The landforms are a record of a fully active surge with glacier advance (stage 3 according to Sund and others, Reference Sund, Eiken, Hagen and Kääb2009) marked with pro- and supraglacial moraines and characteristic meltwater drainage patterns (Lønne, Reference Lønne2016). One of the key features of the surge landsystems are crevasse-squeeze ridges (CSRs), which indicate basal crevassing and saturated basal sediments during the active surge phase, while the preservation of the ridges implies abrupt surge termination and ice stagnation during the quiescent phase (Evans and Rea, Reference Evans and Rea1999; Ben-Yehoshua and others, Reference Ben-Yehoshua, Aradóttir, Farnsworth, Benediktsson and Ingólfsson2023). Fast ice flow during the surge can be inferred from streamlined glacial lineations (Ottesen and Dowdeswell, Reference Ottesen and Dowdeswell2006). Glaciotectonic moraines, often associated with debris flow lobes in the marine environment, reflect an advance onto and deformation of proglacial sediments (Lovell and others, Reference Lovell2018). Ice-cored hummocky moraines record entrainment and transport of high volumes of sediments during the advance, and subsequent melt-out on stagnant ice (Evans and Rea, Reference Evans and Rea1999). Small submarine retreat moraines reflect minor winter advances of quiescent tidewater glaciers (Flink and others, Reference Flink2015).

Hansbreen (Hornsund, south Spitsbergen) is one of the most studied glaciers in Svalbard and its possible surge behaviour has long been debated (e.g. Jania, Reference Jania1988; Jania and Głowacki, Reference Jania and Głowacki1996; Rachlewicz and Szczuciński, Reference Rachlewicz and Szczuciński2000; Ćwiąkała and others, Reference Ćwiąkała2018). The glacier and surrounding area have been the subject of numerous investigations (e.g. Jania, Reference Jania1988; Glazovskiy and others, Reference Glazovskiy, Kolondra, Moskalevskiy and Jania1992; Vieli and others, Reference Vieli, Jania and Kolondra2002; Pälli and others, Reference Pälli, Moore, Jania, Kolondra and Glowacki2003; Grabiec and others, Reference Grabiec, Jania, Puczko, Kolondra and Budzik2012; De Andrés and other, Reference De Andrés2018; Laska and others, Reference Laska2022) and its contemporary dynamics are well understood (Błaszczyk and others, Reference Błaszczyk2021, Reference Błaszczyk2024). According to the Randolph Glacier Inventory (RGI) V7.0, there is ‘no evidence’ that Hansbreen is a surge-type glacier (Maussion and others, Reference Maussion2023; RGI 7.0 Consortium, 2023). In contrast to most medium-sized or large tidewater glaciers in Hornsund, Hansbreen has not surged since the regular monitoring of glacier dynamics initiated in 1982 and there is no unequivocal evidence for surging earlier in the 20th century (Błaszczyk and others, Reference Błaszczyk, Jania and Kolondra2013; Szafraniec, Reference Szafraniec2020). Over the last 40 years, three tidewater glaciers in Hornsund underwent a surge: Paierlbreen (advance of 380 m in 1992–1995), Mendeleevbreen (advance of 1.5 km in 1997–2002) and Svalisbreen (advance of c. 400 m in 2016–2020; Błaszczyk and others, Reference Błaszczyk2023). Evidence of an ongoing surge was also reported in the case of Mühlbacherbreen (in 1961; Sund and others, Reference Sund, Eiken, Hagen and Kääb2009) and Körberbreen (in 1938; Liestøl, Reference Liestøl1969). Marine-terminating Storbreen, Hornbreen and Samarinbreen are also surge-type glaciers (Fig. 1b; Jiskoot and others, Reference Jiskoot, Murray and Boyle2000). Minor advances (c. 100–300 m) of Hansbreen were observed in 1957–1959 (Kosiba, Reference Kosiba1960), 1973–1977 and 1993–1995, and an episode of increased surface velocity occurred in the early 1990s, but these events were interpreted as ‘mini-surges’ rather than fully developed surges (Jania, Reference Jania1988, Reference Jania and Kostrzewski1998). Landforms diagnostic of surging were not detected in the submarine forefield of Hansbreen when it was mapped with multibeam echosounder in 2008 (Tegowski and others, Reference Tegowski, Trzcinska, Kasprzak and Nowak2016; Ćwiąkała and others, Reference Ćwiąkała2018). Farnsworth and others (Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016) classified Hansbreen as a surge-type glacier based on the presence of CSRs in the terrestrial forefield. The geomorphology of the recently deglaciated submarine and terrestrial marginal zones presents an opportunity to investigate newly exposed geomorphological evidence for possible surging. Several studies have also demonstrated the importance of integrating terrestrial and submarine geomorphological records with historical datasets and observations in order to evaluate evidence for past surging of tidewater glaciers (e.g. Ottesen and others, Reference Ottesen2008; Lovell and others, Reference Lovell2018; Aradóttir and others, Reference Aradóttir2019; Zagórski and others, Reference Zagórski2023).

Figure 1. (a) Location of Hansbreen in the Svalbard archipelago and (b, c) Hornsund. (b) Glaciers: Ho, Hornbreen; K, Körberbreen; Me, Mendeleevbreen; Mu, Mühlbacherbreen; P, Paierlbreen; Sa, Samarinbreen; St, Storbreen; Sv, Svalisbreen. (c) Glaciers (black letters): D, Deileggbreen; F, Fuglebreen; K, Kvitungisen; S, Staszelisen; T, Tuvbreen; V, Vrangpeisbreen. Mountain ridges; peaks and saddles (yellow letters): Fa, Fannytoppen; Fb, Fugleberget; Fp, Flatpasset; Fr, Flatryggen; VT, Vesletuva. Peninsulas and islands (orange letters): B, Baranowski Peninsula (Baranowskiodden); Hh, Hansholmane; O, Oseanograftangen. Bays: Hb, Hansbukta; I, Isbjørnhamna. PPS, Polish Polar Station Hornsund. Figure map: © Norwegian Polar Institute (Norwegian Polar Institute, 2014). Sentinel-2B from 27 July 2023.

Here, we present the geomorphology of Hansbreen's terrestrial and submarine forefields and combine this with historical data to evaluate its surge-type behaviour.

Study site

Hansbreen is a polythermal tidewater glacier that terminates in Hansbukta on the northern side of Hornsund in south-west Spitsbergen (Fig. 1; Jania and others, Reference Jania, Mochnacki and Gądek1996; Błaszczyk and others, Reference Błaszczyk2019). The system consists of the main trunk flowing from north to south, four main tributary glaciers on the west (Fuglebreen, Tuvbreen, Deileggbreen and Staszelisen) and several minor tributaries on the east. The glacier is approximately 14 km long with a c. 2 km-long grounded calving front, 51.3 km2 in area (as of 2015), a total ice volume of 9.6 km3, a mean surface slope of 1.8° along the centreline and up to 5.6° along the centrelines of the tributaries (Grabiec and others, Reference Grabiec, Jania, Puczko, Kolondra and Budzik2012; Błaszczyk and others, Reference Błaszczyk2019, Reference Błaszczyk2024). There is a distinct ice divide between Hansbreen and Vrangpeisbreen at the highest part of the main trunk at c. 490 m a.s.l., whereas the tributary glaciers reach up to 565 m a.s.l. (Błaszczyk and others, Reference Błaszczyk2019). The delineation of the eastern border is not straightforward due to the transfluence of ice from the accumulation field of Hansbreen to Kvitungisen, a tributary of Paierlbreen. Around 2010, the mean and the maximum thickness of the glacier were 171 and 386 m, respectively, and the mean surface elevation was 306 m a.s.l. (Grabiec and others, Reference Grabiec, Jania, Puczko, Kolondra and Budzik2012). The surface mass balance has been surveyed since 1989 and the mean annual value in 2000–2019 was −0.26 m w.e. (Schuler and others, Reference Schuler2020). In 2007–2015, the mean annual velocity measured c. 3.5 km from the calving front varied from 49.3 to 88.2 m a−1 and the central part of the front was 4–5 times faster. The average retreat rate in 1992–2015 was 38 m a−1 with a maximum of 311 m a−1 (Błaszczyk and others, Reference Błaszczyk2021). The bedrock of the main trunk is overdeepened below sea level for 10 km from the calving front, and this overdeepening corresponds to a quarter of the total bedrock area beneath the glacier system (Grabiec and others, Reference Grabiec, Jania, Puczko, Kolondra and Budzik2012; Otero and others, Reference Otero2017). The geological structure is complex and diverse with several stratigraphic units of N-S orientation. The main valley has developed within Middle and Late Proterozoic phyllites, dolomites and conglomerates as well as Cambrian dolomites and quartzites. The west tributary glaciers have eroded Middle Proterozoic quartzites, marbles, phyllites and schists (Birkenmajer, Reference Birkenmajer1990; Ohta and Dallmann, Reference Ohta and Dallmann1999). In 1979–2018, mean annual air temperature and precipitation measured at the nearby Polish Polar Station were −3.7°C and 235 mm with increase trends of +1.14°C and +61.6 mm per decade (Wawrzyniak and Osuch, Reference Wawrzyniak and Osuch2019, Reference Wawrzyniak and Osuch2020).

Similar to other glaciers in Hornsund, the first maps of the Hansbreen frontal position were based on surveys by the Austro-Hungarian (1872–1874) and Swedish-Russian expeditions (1899–1902) (Peterman, Reference Peterman1874; Wassiliew, Reference Wassiliew1925). At this time, Hansbreen occupied Hansbukta, the Baranowski Peninsula (Baranowskiodden), part of Isbjørnhamna, Oseanograftangen and raised marine terraces at the foot of Fugleberget and Fannytoppen. The geomorphology of the forefield was investigated by Birkenmajer (Reference Birkenmajer1960), Karczewski and others (Reference Karczewski1984), Pękala (Reference Pękala1989), Lindner and others (Reference Lindner, Marks and Szczęsny1992), Rachlewicz and Szczuciński (Reference Rachlewicz and Szczuciński2000). Since that time, there have been no new studies of the entire terrestrial forefield. The submarine geomorphology of Hansbukta was outlined by Jania (Reference Jania1988, Reference Jania and Kostrzewski1998) and Tegowski and others (Reference Tegowski, Trzcinska, Kasprzak and Nowak2016), and described in detail by Ćwiąkała and others (Reference Ćwiąkała2018).

Materials and methods

We mapped the terrestrial forefield by combining field investigations with remote sensing data, as recommended by Chandler and others (Reference Chandler2018). Field observations were carried out in 2021–2023 with detailed mapping in August 2022. Our map is based on a very high-resolution orthophotomosaic and DEM generated and published by Błaszczyk and others (Reference Błaszczyk, Laska, Sivertsen and Jawak2022) from aerial photos acquired on 22 June 2020 using a Dornier DO228 aircraft. The resolution of the orthophotomosaic and the DEM are 0.0843 and 0.169 m, with a maximal vertical DEM error of 0.54 m. Mapping of the submarine forefield was based on two datasets of bathymetric data. We mapped most of the forefield using a high-resolution DEM (1 m) based on multibeam bathymetric data from the Norwegian Hydrographic Service collected in 2008 (The Norwegian Mapping Authority, Kartverket). This dataset was used in the previous geomorphological studies of Hansbukta (Tegowski and others, Reference Tegowski, Trzcinska, Kasprzak and Nowak2016; Ćwiąkała and others, Reference Ćwiąkała2018). In addition, we used a DEM generated by Błaszczyk and others (Reference Błaszczyk2021) from multiple data sources to map the areas not surveyed in 2008. The DEM was based on measurements from a small boat using a Norbit Wideband Multibeam Sonar with a positioning system (2014, 2016, 2017) and data collected from a small boat with a single beam echosounder Lowrance HDS5 with positioning (2015). The assessed accuracy of these bathymetric data was ±5 m (Błaszczyk and others, Reference Błaszczyk2021). We mapped the landforms in QGIS using the WGS84/UTM33N spatial reference system. Some small areas of the forefields were not covered by the bathymetry data or the orthophotomap and a small part of the seafloor with a surveyed topography remained unclassified in the geomorphological map (Figs 2 and 3). Changes in the frontal position of Hansbreen were taken from supplementary data to Błaszczyk and others (Reference Błaszczyk, Jania and Kolondra2013, Reference Błaszczyk2021) and vectorized from Sentinel-2 images in natural colour compositions for the period 2016–2018.

Figure 2. Hansbreen and its forefield. Orthophotomap from June 2020 after Błaszczyk and others (Reference Błaszczyk, Laska, Sivertsen and Jawak2022) and bathymetric map based on the Norwegian Hydrographic Service multibeam data (2008) combined with bathymetric data (2014–2017) after Błaszczyk and others (Reference Błaszczyk2021) were used to produce the geomorphological map (Fig. 3). Changes in the frontal position of Hansbreen after Błaszczyk and others (Reference Błaszczyk, Jania and Kolondra2013, Reference Błaszczyk2021), Sentinel-2A from 2 August 2016 and 18 September 2018, and Jania (Reference Jania1988). The dashed line marks the glacier extent in 1938 based on terraphotogrammetric measurements conducted by Pillewizer (Reference Pillewizer1939).

Figure 3. Geomorphological map of the forefield of Hansbreen. Glacier extent on 22 June 2020.

To characterize the changing glacier and its forefield over time, we used historical maps and photographs from 1872 (Austro-Hungarian North Pole expedition, Peterman, Reference Peterman1874), a map from 1899 (Swedish-Russian Arc-of-Meridian Expedition, Wassiliew, Reference Wassiliew1925), a map and archival photographs from the Norwegian expedition to Spitsbergen in 1918 (Hoel, Reference Hoel1929), oblique aerial photographs from 1936 by the Norwegian Polar Institute and the orthophotomap from these data (Geyman and others, Reference Geyman, van Pelt, Maloof, Aas and Kohler2022), terrophotogrammetric documentation of the Hansbreen forefield in 1982–1996 from the archives of the University of Silesia (Jania and Kolondra, Reference Jania and Kolondra1982) and the orthophotomap based on vertical aerial photographs from 2011 by the Norwegian Polar Institute, available in the TopoSvalbard online archive (https://toposvalbard.npolar.no/). The historical photographs from 1872 were acquired from the Austrian National Library Digital Image Archive (https://onb.digital/) and from 1918 from the digital image archive of the Norwegian Polar Institute (https://bildearkiv.npolar.no/fotoweb/). The purpose of the historical data was to help interpret the formation of landforms and to assess past glacier frontal positions and any geomorphological or glaciological evidence for surging. As the surge criteria, we considered (1) looped moraines, (2) deformed ice structures, (3) heavy surface crevassing, (4) a surge bulge and (5) shear margins on the glacier surface (Lefauconnier and Hagen, Reference Lefauconnier and Hagen1991; Copland and others, Reference Copland, Sharp and Dowdeswell2003; Grant and others, Reference Grant, Stokes and Evans2009).

Hansbreen surging glacier landsystem

Glacial and glacifluvial landforms are classified according to their origin and morphological features and presented in the context of their associations (subglacial, supraglacial, ice-marginal, proglacial, glacimarine). We describe the terrestrial and submarine marginal zones separately to highlight the similarities and contrasts between landform assemblages in different environments.

Glacier ice and debris covered glacier ice

The main tongue of Hansbreen is densely crevassed at the active calving front, and similarly crevassed at the confluence zone between Tuvbreen and the main trunk (Figs 1 and 2). The dominant crevasse pattern is transverse with additional sigmoidal crevasses near the confluence zone. A dense network of crevasses with different orientations and crevasse traces (healed crevasses) can be observed in the eastern lateral zone of Hansbreen, including the land-terminating part, which transitions into the ice-cored hummocky moraine. The extensively crevassed central and eastern parts of Hansbreen contrast with the almost uncrevassed surfaces of Tuvbreen and Fuglebreen, which are dominated by crevasse traces and exhibit a few supraglacial streams and moulins.

Two medial moraines can be traced on the western part of the glacier (Figs 1 and 2). The medial moraine between the main trunk of Hansbreen and Tuvbreen takes the form of a wide belt of highly dispersed debris a few centimetres thick, with fine clasts of metamorphic rocks from the nunatak Tuva. Snow patches partially mask a scattered medial moraine between Tuvbreen and Fuglebreen composed of fine-grained, homogeneous debris sourced from the metamorphic rocks of the Vesletuva mountain. We interpret the former as an ice-stream interaction medial moraine and the latter as an ablation-dominant moraine.

Terrestrial forefield

Ice-moulded bedrock: Ice-moulded bedrock is exposed on abundant glacially eroded skerries and capes in Hansbukta and Isbjørnhamna, which often take the form of roches moutonnées (Fig. 3). The most prominent of these features is the Baranowski Peninsula, which is covered by a thin, discontinuous layer of fluted till with numerous bedrock outcrops of amphibolites. The surface of the peninsula is streamlined with minor roches moutonnées and rock drumlins (Fig. 4a). Similarly, streamlined bedrock with a fluted till surface occurs at the foot of Fannytoppen, across Hansbukta from the Baranowski Peninsula (Fig. 4b). A large, 9.5 m high roche moutonné, partially draped with a thin layer of fluted till is located in the north-western forefield near the glacier terminus (Figs 5a and c). In the south-eastern forefield, ice-moulded bedrock is also represented by glacio-isostatically uplifted marine abrasive terraces and sea stacks covered with till.

Figure 4. Streamlined topography of (a) the Baranowski Peninsula and (b) at the foot of Fannytoppen with parallel-sided flutings on a thin layer of till. Note CSRs transverse and oblique to flutes on the Baranowski Peninsula. Shaded relief maps after Błaszczyk and others (Reference Błaszczyk, Laska, Sivertsen and Jawak2022).

Figure 5. Fluted till surface with CSRs at the foot of Fugleberget in the north-western part of the forefield. (a) Overview of the flutings and CSRs networks on the orthophotomap from 2020 (Błaszczyk and others, Reference Błaszczyk, Laska, Sivertsen and Jawak2022). (b) Flutings and CSRs emerging from downwasting glacier ice in August 2020. Hummocky moraine and trimline on the Fugleberget slopes in the background. (c) The same area in September 2023. Note two drumlins (dr) and a roche moutonné (rm) overprinted with flutings and well-developed CSRs. (d) A prominent CSR c. 0.8 m high with the extent marked with the dashed line. (b–d) Photos A. Osika.

Fluted till surface: Ice-moulded bedrock is strewn with a thin layer of coarse-grained, matrix-supported diamicton interpreted as subglacial traction till and a discontinuous layer of supraglacial diamicton of subglacial origin with similar properties (Skolasińska and others, Reference Skolasińska, Rachlewicz and Szczuciński2016). The clasts are sub-angular to sub-rounded, ranging from gravels to boulders up to 2 m in diameter. The total thickness of diamictons is usually less than 0.5–1 m (Rachlewicz and Szczuciński, Reference Rachlewicz and Szczuciński2000).

The till surface is densely fluted, except in areas reworked by meltwater streams, where only the most prominent flutes can be recognized. Flutes are low, flow-parallel ridges composed of till similar to the till plain, but with significantly smaller clasts. Their formation is connected with infilling of till into a subglacial cavity in the lee-side of a boulder (e.g. Boulton, Reference Boulton1976; Benn, Reference Benn1994; Ives and Iverson, Reference Ives and Iverson2019). Flutings concentrate in four spatial clusters: (1) the streamlined topography of the Baranowski Peninsula (Fig. 4), (2) the recently exposed north-western part of the forefield, at the foot of Fugleberget (Fig. 5), (3) raised marine terraces at the foot of Fannytoppen (Fig. 4) and (4) remnants of till plain on Oseanograftangen (Fig. 6). Most of the flutings are parallel-sided and up to 84 m long with nearly constant width between 0.3 and 2.4 m and height below 0.3 m, commonly with a boulder lodged on the stoss-side of the flute. There is a second group of short tapering flutes up to 14 m long, with boulders embedded at their heads and decreasing width and height in a downstream direction. The assemblages of long flutings associated with CSRs, which can be observed in the Hansbreen forefield, are vital components of the surging glacier landsystems (e.g. Evans and Rea, Reference Evans and Rea1999, Reference Evans, Rea and Evans2003; Christoffersen and others, Reference Christoffersen, Piotrowski and Larsen2005; Benn and Evans, Reference Benn and Evans2010; Aradóttir and others, Reference Aradóttir2019; Ives and Iverson, Reference Ives and Iverson2019).

Figure 6. (a, b) Roches moutonnées on Oseanograftangen overprinted with fluted till and CSRs. Note the extensive beaches and storm ridges around the hills and remaining of the plain south of the lake, indicating the flutes and CSRs reworking by coastal processes. (c) Levelled flutes (parallel; blue lines) and CSRs (oblique and transverse; red lines). The latter take the form of low ridges (0.3–0.5 m) or diamicton trails a few centimetres high.

Crevasse-squeeze ridges (CSRs): Low, diamicton ridges oriented transverse or oblique to ice-flow direction and superimposed on flutes are interpreted as CSRs. CSRs are composed of the same till as flutings and their formation involves till infilling of basal crevasses that have opened during a surge and subsequent emergence of ridges from stagnating ice during the quiescent phase (e.g. Boulton and others, Reference Boulton1996; Rea and Evans, Reference Rea and Evans2011; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016; Ben-Yehoshua and others, Reference Ben-Yehoshua, Aradóttir, Farnsworth, Benediktsson and Ingólfsson2023). The landforms are a key characteristic of the geomorphological record produced by surging glaciers and have been reported in the forefields of numerous glaciers in Svalbard with both confirmed and undocumented surge histories (e.g. Heintz, Reference Heintz1953; Boulton and others, Reference Boulton1996; Glasser and others, Reference Glasser, Hambrey, Crawford, Bennett and Huddart1998; Ottesen and Dowdeswell, Reference Ottesen and Dowdeswell2006; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016; Ottesen and others, Reference Ottesen, Dowdeswell, Bellec and Bjarnadóttir2017; Flink and Noormets, Reference Flink and Noormets2018; Aradóttir and others, Reference Aradóttir2019). The morphology of CSRs in the forefield of Hansbreen ranges from distinct, intersecting ridges c. 0.5 m high to remnants of ridges or mounds and sediment trails reflecting the characteristic criss-cross pattern of CSRs. A network of well-defined cross-cutting ridges c. 0.5 m high, up to 3.5 m wide and up to 50 m long that overlay dense flutes can be seen emerging from the downwasting glacier snout in the northwestern part of the forefield (Fig. 5). The most prominent ridges approach the height of 0.8 m (Fig. 5d). In contrast, CSRs on the Baranowski Peninsula appear as indistinct ridges in the field and are best identified on the orthophotomaps. However, in the 1980s they were visible as abundant, distinct diamicton ridges transverse or oblique to underlying flutes (Fig. 7). In the eastern forefield, CSRs occur less frequently and are mostly individual diamicton ridges and mounds 6–40 m long and 0.3–0.5 m high overprinted on highly degraded flutes (Figs 6 and 8). The ridges are located among numerous cross-cutting diamicton trails a few centimetres high, which mimic the networks of CSRs and flutings. Similar features interpreted as the remains of CSRs were described in western Svalbard by Ben-Yehoshua and others (Reference Ben-Yehoshua, Aradóttir, Farnsworth, Benediktsson and Ingólfsson2023). In the southernmost forefield, transverse and oblique CSRs are superimposed on fluted till draped on two roches moutonnées (Fig. 6). Their morphology ranges from distinct, short ridges and mounds approaching 0.3–0.5 m in height, to degraded, intersecting ridges a few centimetres high taking the form of diamicton trails.

Figure 7. (a, b) Terrestrial photogrammetric photograph of the Baranowski Peninsula from 1982. Note long, parallel-sided flutings and low CSRs (white arrows) oblique and transverse to the flutes. Today, CSRs in this site are sparse and have an indistinct appearance in the field.

Figure 8. (a) Low, obliterated CSRs c. 0.3–0.5 m high among sediment trails reflecting the cross-cutting pattern of CSRs and flutings at the foot of Fannytoppen. (b) Ice-cored hummocky moraine in this area. In the background, hummocky moraine and trimline on the slopes of Fugleberget. (c) Hummocky moraine (h) and trimline (t) on the Fannytoppen slopes. The arrow marks the lateral terminal moraine and the rectangle illustrates the location of the roches moutonnées with CSRs on Oseanograftangen. (a–c) Photos A. Osika.

Eskers: Low, locally eroded sinuous eskers are visible in the eastern forefield on the till plain with fluvial overprinting. The ridges are 70–90 m long, 1–3 m wide, up to 0.5 m high and dissected by meltwater channels. They can be recognized by their surface of compact, sub-angular to rounded clasts ranging from fine gravels to 30 cm boulders with absent finer fractions. Based on their sinuous morphology, we interpret them as landforms developed in stagnant ice during the quiescent phase.

Hummocky moraine: Hummocky moraine occurs on the ice-proximal side of the lateral terminal moraine (Fig. 3). In the western part of the forefield, it takes the form of low, ice-cored, aligned hummocks, depressions and ponds, transitioning to the north into regular, ice-cored, low-amplitude linear ridges, interpreted as controlled moraine (Evans, Reference Evans2009). The formation of hummocky and controlled moraines involves entrainment, transport and melt-out of debris on a stagnant glacier snout, leading to the accumulation of supraglacial debris cover (Boulton, Reference Boulton1968; Rachlewicz and Szczuciński, Reference Rachlewicz and Szczuciński2000; Schomacker and Kjær, Reference Schomacker and Kjær2008; Evans, Reference Evans2009; Ewertowski and others, Reference Ewertowski, Kasprzak, Szuman and Tomczyk2011). In controlled moraine, the supraglacial debris retains distinct linearity reflecting the inherited pattern of englacial debris concentrations (Evans, Reference Evans2009). In the eastern forefield, hummocky moraine occurs as a complex of irregular, partly ice-cored, low-amplitude mounds and depressions at the proximal zone of the lateral terminal moraine (Fig. 8). The second assemblage is a complex of low, recently exposed ice-cored hummocky moraines adjacent to the slopes of Fannytoppen and Flatryggen. The surficial sediments are clast-rich, matrix-supported diamictons with large clasts of metamorphic and sedimentary rocks (Skolasińska and others, Reference Skolasińska, Rachlewicz and Szczuciński2016). Abundant fragments of marine mollusc shells can be found up to the Flatpasset saddle between Fannytoppen and Flatryggen at c. 150 m a.s.l. (Fig. 2). The development of hummocky moraines requires entrainment and transport of large volumes of debris, which is particularly effective and widespread during the active phase of surging (Evans and Rea, Reference Evans and Rea1999; Lovell and others, Reference Lovell2015a). Therefore, they are an important component of the surging glacier landsystems, although they are not independently diagnostic of surge-type behaviour (Evans and Rea, Reference Evans, Rea and Evans2003).

Lateral terminal moraine: The outer, most prominent moraine belts are interpreted as lateral terminal moraines that extend underwater as a frontal terminal moraine (Fig. 3). In the western forefield, the lateral terminal moraine occurs as a conspicuous controlled moraine belt with sub-parallel units reflecting the former englacial structures in the glacier snout (cf. Evans, Reference Evans2009). The outer moraine belt is 1130 m long, up to 200 m wide, 30 m high and ice-cored as evidenced by electrical resistivity tomography (Glazer and others, Reference Glazer, Dobiński, Marciniak, Majdański and Błaszczyk2020), but the internal structure is unknown. The transverse profile is asymmetric with a steeper distal slope. In the south-eastern forefield, the lateral terminal moraine takes the form of two arcuate ridges 570 and 580 m long, 30–50 m wide and up to 8 m high, dissected by proglacial channels. The ridges are probably ice-cored based on the evidence of ongoing de-icing processes. The relief of their southern margins is subdued and resembles the adjacent hummocky moraine. In turn, the lateral terminal moraine on Flatpasset (Fig. 8c) occurs as a distinct, arcuate ridge 420 m long, up to 30 m wide, 6 m high and probably ice-cored. The northern part merges with ice-cored hummocky moraine, whereas the south-western part is associated with a trimline on the adjacent Fannytoppen slope. The lateral terminal moraines are composed of clast-rich diamicton with sub-angular to sub-rounded, and occasionally angular large clasts of metamorphic and sedimentary rocks. We interpret them as the outer and oldest ridges of controlled moraines, formed by debris entrainment during glacier advance and supraglacial melt-out of sediments from stagnating ice, preserving the ice core from degradation (cf. Evans, Reference Evans2009).

Trimline: Trimlines can be traced on the eastern and south-eastern slopes of Fugleberget at 174–70 m a.s.l. and the western slopes of Fannytoppen at 173–32 m a.s.l. (Figs 3 and 8). We interpret them as the maximum thickness of Hansbreen during the surge (cf. Yde and others, Reference Yde2019). The imprint is discontinuous due to paraglacial reworking and formation of talus cones. A gradual elevation decline corresponds to widening and lateral spread-out of the glacier tongue. Well-developed trimlines from the LIA are typical for Svalbard glaciers and their formation and distribution are primarily controlled by bedrock erodibility (Rootes and Clark, Reference Rootes and Clark2022). According to their study, erosional trimlines are slightly more common among surging and tidewater glaciers in Svalbard compared to land-terminating glaciers or glaciers with no known surge history.

Submarine forefield

Drumlinized seafloor and streamlined glacial lineations: The drumlinized seafloor comprises c. 750 m wide zone in the proximal part of the submarine forefield, taking the form of streamlined topography with drumlinoids, low-amplitude streamlined ridges and grooves (Figs 2 and 3). In the central zone, an oval, symmetrical ridge 260 m long and 90 m wide, interpreted as a drumlin, can be observed between a similar ridge with an indistinct appearance and a prominent drumlinoid 190 m wide. In the western part, an assemblage of distinct, flow-parallel, low-amplitude ridges and grooves up to 430 m long can be observed on the slopes of an overdeepening that extends to a depth of 94 m. The ridges are up to 5 m high, 30 m wide and the distance between their crests is usually 40–60 m. We interpret these as streamlined glacial lineations, which indicate high velocities during glacier advance and are a key characteristic of the submarine forefields of tidewater surge-type glaciers in Svalbard (e.g. Flink and others, Reference Flink2015, Reference Flink, Hill, Noormets and Kirchner2018; Dowdeswell and Ottesen, Reference Dowdeswell and Ottesen2016; Ottesen and others, Reference Ottesen, Dowdeswell, Bellec and Bjarnadóttir2017; Streuff and others, Reference Streuff, Cofaigh, Noormets and Lloyd2018). Based on their morphology and elongation ratios, the ridges represent a continuum between elongated drumlins and mega-scale glacial lineation, indicating subglacial streamlining (MSGL; Stokes and Clark, Reference Stokes and Clark2002; Spagnolo and others, Reference Spagnolo2014).

Esker: A single, sinusoidal ridge in the inner part of the forefield, which can be traced to an R channel in the nearby land-terminating part of the glacier is interpreted as an esker, resulting from the infilling of the subglacial conduit with sediments during decreasing meltwater discharge (Ottesen and others, Reference Ottesen2008). The ridge is 390 m long, up to 10 m wide, 2 m high and faces east for the first 185 m before turning south. The distinct morphology of the first segments contrasts with blurred distal part c. 90 m long ending at the recessional moraine. Eskers have been documented in the forefields of many tidewater glaciers in Svalbard and, in the case of surge-type glaciers, inferred to have developed after surge termination (Ottesen and others, Reference Ottesen2008; Dowdeswell and Ottesen, Reference Dowdeswell and Ottesen2016; Flink and others, Reference Flink, Hill, Noormets and Kirchner2018; Noormets and others, Reference Noormets, Flink and Kirchner2021).

Small moraine ridges – CSRs and / or De Geer moraines: Small ridges oriented subparallel to the calving front and overprinting the flow-parallel ridges of streamlined bedforms can be traced across the inner part of Hansbukta. The transverse, locally branching and cross-cutting ridges approach 10–20 m in width and 3 m in height with spacing of 15–50 m. Their spatial pattern and abundance compared with the number of winter glacier advances (Błaszczyk and others, Reference Błaszczyk2021) and the proximity of CSRs in the terrestrial forefield (Fig. 3) imply that the ridges are probably submarine CSRs formed during a surge (e.g. Ottesen and others, Reference Ottesen2008; Dowdeswell and Ottesen, Reference Dowdeswell and Ottesen2016). Based on a similar appearance, some ridges could be De Geer moraines, also known as small annual retreat moraines or push moraines (e.g. Streuff and others, Reference Streuff, Forwick, Szczuciński, Andreassen and Cofaigh2015; Flink and Noormets, Reference Flink and Noormets2018). The formation of De Geer moraines involves the pushing of sediments during seasonal winter advances (Boulton, Reference Boulton1986; Ottesen and Dowdeswell, Reference Ottesen and Dowdeswell2006). The landforms are widespread in the proximal but not visible in the central or outer parts of the forefield, possibly because they have been buried by sediments from meltwater plumes.

Recessional moraines: Several distinct transverse ridges, larger than those described above, can be observed primarily in the central and outer parts of the forefield. The ridges are oriented sub-parallel to the calving front, up to 10–15 m high and partly buried in glacimarine sediments (Ćwiąkała and others, Reference Ćwiąkała2018). In the proximal forefield, a chain of crescent ridges c. 5 m high corresponds to several winter advances to a similar position in 2013–2015. The most conspicuous ridges correspond with glacier advances in 1957–1959, 1973–1977 and during the 1990s (Jania, Reference Jania and Kostrzewski1998; Ćwiąkała and others, Reference Ćwiąkała2018; Błaszczyk and others, Reference Błaszczyk2021). We interpret the ridges as recessional moraines associated with minor readvances interrupting the overall glacier retreat (Robinson and Dowdeswell, Reference Robinson and Dowdeswell2011; Flink and others, Reference Flink2015; Streuff and others, Reference Streuff, Cofaigh, Noormets and Lloyd2018).

Terminal moraines: Two terminal moraines can be observed in the submarine forefield of Hansbreen, each of them associated with a debris flow lobe. The outermost terminal moraine takes the form of a large, crescent, transverse ridge beyond Hansbukta, emerging onshore as the ice-cored lateral terminal moraines and marking the LIA maximum of Hansbreen (Figs 2 and 3). The total length of the LIA terminal moraine approaches 5100 m with a submarine part of c. 3300 m. The arcuate form indicates the radial splaying of the glacier beyond the constraining mountain ridges. The indentation in the distal flank is related to the abrupt break in the slope into a deeper fjord basin, which has prevented the full development of the moraine (cf. Dowdeswell and others, Reference Dowdeswell, Ottesen and Bellec2020). The outermost terminal moraine of Hansbreen was described in detail by Ćwiąkała and others (Reference Ćwiąkała2018). The submarine ridge is plain-topped, c. 13 m high, with clear margins in the central, but extensively blurred in the lateral parts and located in a water depth of 10–25 m. Rock outcrops, fields of megaripples, iceberg pits and ploughmarks can be identified on the moraine surface (Ćwiąkała and others, Reference Ćwiąkała2018). On the proximal flank, four oval, flat-floored depressions 28–55 m in diameter and 2 m deep are interpreted as kettle holes. Near Oseanograftangen, the moraine surface is eroded by a channel-like feature interpreted as a glacifluvial incision, 2–7 m deep, 160–250 m wide, and with blurred margins reworked by sediment flows (Figs 2 and 3). In Isbjørnhamna, the outermost terminal moraine is subdued and overprinted with a debris flow lobe associated with an inner terminal moraine.

The inner terminal moraine occurs as transverse ridge segments up to 3 m high, semi-parallel to the outer terminal moraine and corresponding to the glacier extent in 1938 (Figs 2 and 3). We imply that terminal moraines have been formed by surges based on their resemblance to surge terminal moraines of Svalbard tidewater glaciers, in particular the association with well-developed debris flow lobes (e.g. Flink and others, Reference Flink2015; Burton and others, Reference Burton, Dowdeswell, Hogan and Noormets2016; Noormets and others, Reference Noormets, Flink and Kirchner2021). The morphology of the inner terminal moraine of Hansbreen is similar to the moraines of Tunabreen or Blomstrandbreen, corresponding to surges in the 20th century (Flink and others, Reference Flink2015; Burton and others, Reference Burton, Dowdeswell, Hogan and Noormets2016). In each site, the terminal moraines of subsequent surges are significantly smaller than the LIA terminal moraine and often discontinuous across the fjord. The formation of submarine terminal moraines involves thrusting and pushing of sediments during glacier advances as well as depositional processes during glacier stillstand, which may be long-lasting on a sill (e.g. Benn and Evans, Reference Benn and Evans2010; Ottesen and others, Reference Ottesen, Dowdeswell, Bellec and Bjarnadóttir2017; Lovell and others, Reference Lovell2018).

Debris flow lobes: We identified two glacigenic debris flow lobes in the forefield of Hansbreen. The outermost lobe overprints the distal flank of the large terminal moraine, the adjacent flat-floored area of Isbjørnhamna and a steep northern side of the Hornsund Fjord beyond the outermost terminal moraine (Figs 2 and 3). In Isbjørnhamna, the lobe is flat-surfaced with clear outer margins, several rock outcrops and small, circular depressions, which are probably iceberg pits. Towards the break in slope into a deeper part of Hornsund, the lobe surface becomes hummocky with abundant iceberg pits, ploughmarks and minor sediment slides and flows. Similar topography can be observed beyond the central and eastern parts of the outermost terminal moraine with several prominent downslope-oriented flow structures. The ice-proximal margins of the moraine are clear, apart from the area adjacent to the moraine incision affected by mass-wasting processes. The second debris flow lobe is associated with the inner terminal moraine and superimposed on the top of the outermost, large terminal moraine in Isbjørnhamna. The landform has distinct margins and flow direction to the south-west (Figs 2 and 3). According to published backscattering data, the sediment flow lobes are composed of fine-grained or mixed material contrasting with coarse-grained surficial sediments of the terminal moraines (Ćwiąkała and others, Reference Ćwiąkała2018). Glacigenic debris flow lobes result from the bulldozing or pushing of seafloor sediments in front of the advancing terminus during surging and subsequent slope failure on the distal side of the glaciotectonic moraine (Dowdeswell and Ottesen, Reference Dowdeswell and Ottesen2016; Lovell and others, Reference Lovell2018). The steep topography beyond the outermost terminal moraine has presumably controlled the development of the associated debris flow lobe, as they usually take the form of extensive gently sloped lobes in flat-floored areas (e.g. Dowdeswell and others, Reference Dowdeswell, Ottesen and Plassen2016; Flink and others, Reference Flink, Hill, Noormets and Kirchner2018; Streuff and others, Reference Streuff, Cofaigh, Noormets and Lloyd2018).

Smooth glacigenic seafloor: Smooth seafloor with flat-floored depressions separated by transverse moraine ridges comprises a major part of the submarine forefield (Figs 2 and 3). The maximum depth of depressions increases towards the current glacier front. We interpret these as glacially overdeepened basins filled with sediments from meltwater plumes, previously described by Tegowski and others (Reference Tegowski, Trzcinska, Kasprzak and Nowak2016) and Ćwiąkała and others (Reference Ćwiąkała2018). The basin-fill sediments partly bury the moraine ridges from minor readvances of Hansbreen. The sediment thickness ranges from 10 m in the central to 25 m in the southernmost part of the submarine forefield and the estimated mean sediment accumulation rate in particular basins ranged between 0.25 and 0.4 m a−1 (Ćwiąkała and others, Reference Ćwiąkała2018). In the topography of the smooth seafloor, Ćwiąkała and others (Reference Ćwiąkała2018) documented fields of megaripples, pockmarks, iceberg pits and ploughmarks.

Pockmarks: Pockmarks occur in the southernmost part of the forefield as small, circular depressions with diameters and depths approaching 15 and 1 m, respectively. The distribution of pockmarks in the basin adjacent to the frontal terminal moraine was documented by Ćwiąkała and others (Reference Ćwiąkała2018), who suggested their possible origin as gas migration along faults or seepage of pore water due to very high sedimentation rate. Widespread craters and mounds related to methane release develop in the Arctic seabed after deglaciation (Andreassen and others, Reference Andreassen2017). Apart from geological factors (e.g. Weniger and others, Reference Weniger2019), pockmarks in Svalbard fjords may form as a result of water escape from rapidly deposited glacigenic debris flow lobes related to surging (Ottesen and others, Reference Ottesen2008; Forwick and others, Reference Forwick, Baeten and Vorren2009). Pockmarks in Hansbukta may be of similar origin considering the high sediment accumulation rate in the basins (cf. smooth glacigenic seafloor; Ćwiąkała and others, Reference Ćwiąkała2018). In addition, we infer that their formation may be associated with the burying of organic-rich sediments during glacier advance.

Hansbreen on archival maps and photographs (1872–1936)

Historical maps

The first published sketch map of the Hansbreen snout (scale 1: 200 000) was prepared in 1872 during the Austro-Hungarian North Pole Expedition (Peterman, Reference Peterman1874). The glacier was also presented on a 1: 200 000 map from the Russian–Swedish expedition in 1899–1901 (Wassiliew, Reference Wassiliew1925). The frontal position can be traced on a 1: 100 000 map based on reconnaissance photogrammetric surveys of the Hornsund area in 1918 during the Norwegian expeditions in 1917–1921 (Fig. 9a; Hoel, Reference Hoel1929). In these maps, Hansbreen occupies Hansbukta and the north-eastern part of Isbjørnhamna, covering the Baranowski Peninsula and Oseanograftangen. However, the maps are not detailed enough to accurately assess the extent and possible fluctuations of the glacier. Hence, our study is primarily based on archival photographs documenting the extent and surface morphology of Hansbreen.

Figure 9. (a) Hansbreen on a 1: 100 000 map ‘Spitsbergens Kyst, Hornsund til Dunder Bay’ based on the Norwegian Spitsbergen expeditions in 1917–1921 with photogrammetric measurements in Hornsund in 1918 (public domain, https://data.npolar.no). (b–g) Terrestrial photogrammetric documentation of Hansbreen from 1918. (b) The radially expanded glacier snout after surge advance. Note a folded medial moraine in the eastern part of the glacier. (c–f) A complex network of narrow, cross-cutting crevasses and crevasse traces documented from Fannytoppen. (g) The glacier front in Isbjørnhamna. Note a waterfall from an active supraglacial channel in the background. (b) © Olaf Holtedahl, Norsk Polarinstitutt (modified), (c–f) © Jørgen Gløersen, Norsk Polarinstitutt (modified), (g) © Adolf Hoel, Norsk Polarinstitutt (cropped and modified).

Archival photographs

Unique photographic documentation of Hansbreen from the late 19th and early 20th centuries was collected by the Austro-Hungarian expedition in 1872 and the Norwegian expedition in 1918. In 1936, Hansbreen was documented on oblique aerial photographs during the Norwegian mapping campaign of Svalbard (Luncke, Reference Luncke1936). The photographs provide data on the glacier extent, surface morphology, ice structure in the ice cliff and geomorphological features, which shed light on the glacier dynamics. In addition, they are a vital source of data on geomorphological processes.

1872: Photographs from the Austro-Hungarian expedition in 1872 documented the terrestrial margins of Hansbreen and part of the calving front. In the western terrestrial zone, the glacial terminus had already reached its LIA maximum position (Fig. 10a). The marine-terminating zone was densely and chaotically crevassed with a conspicuous surface bulge (Fig. 10b). At the foot of Fannytoppen, the terrestrial margin of Hansbreen terminated as a steep ice cliff, exposing inclined and folded debris layers. The photograph also documented top-down and bottom-up crevasses and a convex surface profile (Figs 10c and d). In this part of the glacier, the terminus had not yet reached the LIA maximum, today marked by the lateral terminal moraine adjacent to rocky raised marine terraces. The proglacial outwash plains mapped outside the terminal moraine were also absent in this photograph (Figs 3 and 10c).

Figure 10. Photographs of Hansbreen from the Austro-Hungarian expedition in 1872. (a) The western terrestrial margin of the glacier. (b) Crevassed marine-terminating zone with a conspicuous surface bulge. (c, d) The eastern terrestrial margin, where the glacier had not yet reached the maximum extent. Note folded and inclined debris layers, top-down and bottom-up crevasses exposed on the ice cliff and a convex surface profile of the glacier. © ÖNB Vienna: Image ID 00299043, 00480587 and 00444839 (cropped and modified).

1918: The Norwegian Spitsbergen expedition in 1918, led by Adolf Hoel and Sverre Røvig, conducted reconnaissance photogrammetric measurements in Hornsund. The comprehensive documentation, including photogrammetric photos, overlays the entire glacier. At that time, the glacier front was anchored on the small Hansholmane islands and Oseanograftangen (Figs 9a and b). The extent of the terrestrial margins was similar to that of 1872, but the calving front positions are difficult to compare due to insufficient data from the previous photographs. Glacier-free unnamed islands within the terminal moraine belt suggest a slight retreat from the maximum LIA position. A photograph from the southern side of Hornsund revealed a folded median moraine between Hansbreen and one of the eastern tributaries (Fig. 9b). The two black features on the calving front can be observed in other photographs and could be interpreted as the mouths of subglacial channels. The photos from Fannytoppen documented a dense, complex network of intersecting crevasses and crevasse traces extending high up on the main trunk of the glacier (Figs 9c–f). Most crevasses were narrow, which allowed the expedition to cross the glacier (Hoel, Reference Hoel1929). The photographs also documented a steep ice cliff in Isbjørnhamna with supra- and englacial sediments and a waterfall at the ice front (Fig. 9g). Compared to 1872, this part of the glacier had a smooth topography in 1918.

1936: By 1936, the glacier had retreated beyond Hansholmane, but the terminus was still anchored on Oseanograftangen and the Baranowski Peninsula (Fig. 11). The glacier front was relatively straight and slightly expanded in the central zone. The downwasting of the stagnant terrestrial margins had led to the development of the ice-cored moraine complexes. The lateral moraines of the eastern tributary glaciers had merged into a wide belt along the glacier margin. As a result, the folded moraine of the 1918 photo was no longer visible as it was probably squeezed to the eastern lateral moraine. The surface morphology was dominated by crevasse traces and wide transverse crevasses in the frontal zone. Some crevasse traces reflected the intersecting pattern of crevasses documented in 1918.

Figure 11. Hansbreen in the oblique aerial photograph from 1936. © Norwegian Polar Institute.

Discussion

Geomorphological records of surging

The forefield of Hansbreen with CSRs, fluted till plains, hummocky moraines, high trimlines, drumlinized seafloor, glacial lineations, De Geer moraines, terminal moraines with debris flow lobes and other components (Fig. 3) indicates at least one surge event. Such landform assemblages are consistent with surging glacier landsystem models in terrestrial and submarine environments (Evans and Rea, Reference Evans and Rea1999, Reference Evans, Rea and Evans2003; Ottesen and Dowdeswell, Reference Ottesen and Dowdeswell2006; Ottesen and others, Reference Ottesen2008, Reference Ottesen, Dowdeswell, Bellec and Bjarnadóttir2017; Flink and others, Reference Flink2015; Aradóttir and others, Reference Aradóttir2019).

The key features diagnostic for the surges of Hansbreen are CSRs, glacial lineations and two glacitectonic terminal moraines with debris flow lobes. Terrestrial CSRs and flutings occur up to the southernmost part of field (Figs 2–8). In Svalbard, CSRs networks emerging from stagnating ice may be related to surge episodes from years to over a century ago (Lovell and others, Reference Lovell2015a; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016). CSRs and flutings in the terrestrial forefield of Hansbreen correspond with the submarine features. The rapid ice flow during the surge is evidenced by the streamlined glacial lineations (Figs 2 and 3; Ottesen and others, Reference Ottesen2008, Reference Ottesen, Dowdeswell, Bellec and Bjarnadóttir2017; Streuff and others, Reference Streuff, Cofaigh, Noormets and Lloyd2018), although their absence would not preclude a surge event (Aradóttir and others, Reference Aradóttir2019). The submarine CSRs and glacial lineations appear only in the inner part of Hansbukta, but they might have been more widespread in the past and buried later with glacimarine sediments. A restricted extent or absence of CSRs on the fjord floor, but occurrence on land, were evidenced in the forefield of surge-type Blomstrandbreen (Burton and others, Reference Burton, Dowdeswell, Hogan and Noormets2016; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016) and the glacial system of Trygghamna (Aradóttir and others, Reference Aradóttir2019). Hansbreen has not developed a terrestrial thrust-block moraine diagnostic of surging, but such landforms are not widespread in the lateral forefields of surge-type tidewater glaciers in Svalbard (Lønne, Reference Lønne2016; Lovell and Boston, Reference Lovell and Boston2017). The slightly curvilinear ridges of the lateral terminal moraines and the submarine large terminal moraine indicate the radial expansion of the glacier snout when entering the Hornsund Fjord (Figs 2, 3, 9). The occurrence of terrestrial CSRs up to the southernmost part of the forefield (Figs 3 and 6), the dimensions of the submarine terminal moraine and the occurrence of a debris flow lobe on its distal flank imply that the maximum LIA extent of Hansbreen was related to a surge. The lateral terminal moraine on the Flatpasset saddle (Figs 2 and 3) marks the glacier's attempt to pass the mountain ridge during this surge advance when the glacier entrained marine sediments with mollusc shells and deposited with diamicton up to 150 m a.s.l.

The inner terminal moraine with a debris flow lobe is the main evidence of a possible second surge. The lobe has overprinted the large terminal moraine and the flow direction of sediment was to the south-west, perpendicular to the inner moraine (Figs 2 and 3). The inner moraine corresponds to the frontal position of Hansbreen in 1938, which was similar or, in some parts, slightly more extensive than in 1936 (Fig. 2). The dimensions of the moraine and the sediment lobe are notably smaller than those from the LIA maximum. However, in Svalbard, surge terminal moraines from the 20th century tend to be less prominent than surge moraines from the LIA and may have similar morphology to recessional moraines (Burton and others, Reference Burton, Dowdeswell, Hogan and Noormets2016). Surge end moraines of Tunabreen from 1930, 1971 and 2004 are much lower than the outermost moraine belt from the 1870s surge and only two surge moraines have adjacent debris flow lobes (Flink and others, Reference Flink2015). The younger debris flow lobe of Hansbreen has similar dimensions to debris flow lobes on the outermost surge moraines of Blomstrandbreen (Streuff and others, Reference Streuff, Forwick, Szczuciński, Andreassen and Cofaigh2015, Reference Streuff, Cofaigh and Wintersteller2022). This glacier has not developed debris flow lobes at all during the subsequent surges in the 20th century (Burton and others, Reference Burton, Dowdeswell, Hogan and Noormets2016). The formation of the inner surge moraine of Hansbreen may have been limited by the scarce supply of sediment from resistant metamorphic rocks in the glacier basin that had been redeposited during the older surge. Low surge end moraines with small or without debris flow lobes may also reflect a short-active surge phase, which limits the sediment supply to the glacier terminus (Flink and others, Reference Flink2015). In addition, most of the moraine ridges in the central and southern Hansbukta have been partly buried with glacimarine sediments, which mask their primary size. If the inner terminal moraine had a continuation across Hansbukta, it also might have been buried and subdued.

Historical records of surging

1872 – ongoing surge: The photographs of Hansbreen from the Austro-Hungarian expedition in 1872 reveal characteristic features of an active surge phase (Fig. 10). The glacier surface was densely crevassed with a conspicuous bulge, which we interpret as a surge front propagating downglacier (e.g. Sund and others, Reference Sund, Eiken, Hagen and Kääb2009; Lovell and Fleming, Reference Lovell and Fleming2023). A surge bulge reflects the mass displacement from the reservoir area and builds up in a high friction zone near the glacier front (Haga and others, Reference Haga2020). The ice thickness at the south-western margin corresponded to the trimline on edge between the eastern and southern slopes of Fugleberget (Fig. 2). Most of the southern slope is overprinted with talus cones and it was not possible to identify a clear trimline over there (Figs 2, 3, 10a). The vertical ice extent in 1872 was not documented on the other mountain slopes. The south-eastern margin of Hansbreen had a convex surface profile and a steep ice cliff with debris-rich structures (Figs 10c, d). Similar features are typical for the termini of surging glaciers (e.g. Lovell and others, Reference Lovell2015a; Ingólfsson and others, Reference Ingólfsson2016; Sobota and others, Reference Sobota, Weckwerth and Nowak2016; Lovell and Fleming, Reference Lovell and Fleming2023) and we suggest they fed the subsequent controlled moraines in this area (Fig. 3). The glacier extent in this area was smaller than marked by the LIA lateral terminal moraine. Hence, we imply that the surge was underway and the glacier was advancing to the maximum LIA position.

1918 – quiescent phase: Photographic documentation from the Norwegian Spitsbergen expedition in 1918 provides evidence for a glacier in the quiescent phase. The radially expanded snout was anchored on capes and small islands, and the glacier extent corresponded to the outermost submarine terminal moraine and the lateral terminal moraines. The ice thickness was similar to the trimlines mapped on the Fugleberget slopes and the southern slope of Fannytoppen (Figs 3 and 9). The frontal position did not change significantly between 1872 and 1918 (Koryakin, Reference Koryakin1974; Jania, Reference Jania1988). We suggest it was linked to anchoring the terminus on the terminal moraine and skerries with the prevailing thinning of the glacier tongue by the surface ablation rather than limited frontal ablation in shallow water. The water depth on the terminal moraine ranges from 2 to 15 m and does not exceed 25 m on the ice-proximal side of the belt (Fig. 2). The southernmost extent of the calving front was on the Hansholmane islands, which were probably one of the pinning points within the moraine belt (Figs 1, 2, 3, 9a and b). The mouths of subglacial channels could be observed on the calving front (Figs 9a and b) and the location of one of them seemed to correspond with an incision in the submarine terminal moraine (Figs 2 and 3), which suggests possible glacifluvial origin of this feature. The surface morphology of Hansbreen revealed compelling evidence for a surge before 1918, such as a folded medial moraine and a dense, complex network of intersecting crevasses and crevasse traces extending high up on the main trunk of the glacier (Fig. 9). We infer that abundant healed crevasses and their cross-cutting pattern were associated with rapid ice flow while surging in 1872. In 1918, the glacier terminus in Isbjørnhamna had a smooth topography and an active supraglacial channel with a waterfall at the ice cliff, which suggests closed crevasses in this part of the glacier during the quiescent phase (Fig. 9g). Similar spectacular features can be observed on the front of Bråsvellbreen (https://toposvalbard.npolar.no/) after decades of the quiescent phase in the aftermath of its surge in 1938 when the glacier tongue was badly crevassed (Schytt, Reference Schytt1969).

Late 1930s – limited evidence for a subsequent surge: In the western part of the forefield, the inner terminal moraine with a debris flow lobe suggests a younger surge of Hansbreen, which might have occurred in the first half of the 20th century (Fig. 2). Oblique aerial photographs from 1936 do not indicate a fully active surge (stage 3 according to Sund and others, Reference Sund, Eiken, Hagen and Kääb2009) as the surface morphology was dominated by crevasse traces, wide transverse crevasses were confined to the frontal zone, the terrestrial margins were stagnant, and the calving front was relatively straight and slightly expanded only in the central zone (Fig. 11). These data alone do not allow a qualitative assessment of the upper part of the glacier for evidence for the first two stages of the active surge phase expressed by changes in the glacier thickness (Sund and others, Reference Sund, Eiken, Hagen and Kääb2009). In August 1938, Pillewizer (Reference Pillewizer1939) measured the unusually high surface velocity of Hansbreen using the terrestrial photogrammetric method of Finsterwalder (Reference Finsterwalder1931), which could suggest an active surge phase, and his map indicates an advance in some parts of the glacier front compared to 1936 (Fig. 2). If such event occurred, it may have been related to the inner terminal moraine with a debris flow lobe. However, additional evidence is necessary to support this interpretation in its current state.

Reconciling glacial geomorphology and past dynamics of Hansbreen

The possible surge-type behaviour of Hansbreen has long been debated for the lack of evidence of an ongoing surge in the observation period (e.g. Jania, Reference Jania1988; Jania and Głowacki, Reference Jania and Głowacki1996; Rachlewicz and Szczuciński, Reference Rachlewicz and Szczuciński2000; Ćwiąkała and others, Reference Ćwiąkała2018). We present the records of at least one surge event. Photographic documentation of the glacier extent, surface morphology and ice structure in the ice cliff revealed characteristic features for the active surge phase in 1872 (Fig. 10) and quiescence in 1918 (Fig. 9). This advance was marked by the outermost terminal moraine, trimlines and the lateral terminal moraines (Figs 2 and 3). Sparse historical data may suggest a minor surge episode in the 1930s, but further investigation is necessary to confirm this event. The imprints of the possible younger surge are much weaker than the LIA surge and the only compelling geomorphological evidence is the inner terminal moraine with a debris flow lobe (Fig. 3).

Our interpretation differs from previous investigations, which did not conclude that surges had occurred. Ćwiąkała and others (Reference Ćwiąkała2018) analysed the submarine forefield and excluded the possibility of surging in the last c. 120 years based on the absence of streamlined glacial lineations, overridden recessional moraines, CSRs or a surge moraine with a debris flow lobe. However, drumlinized seafloor, streamlined glacial lineations and CSRs are not visible in the multibeam data collected in 2008, which they used for geomorphological mapping. Such landforms occur in the inner zone of Hansbukta surveyed in 2014–2017 (Fig. 3). CSRs were also not identified in the terrestrial forefield of Hansbreen (Skolasińska and others, Reference Skolasińska, Rachlewicz and Szczuciński2016 and references therein), but part of the debris-rich structures in the passive ice cliff described by Rachlewicz and Szczuciński (Reference Rachlewicz and Szczuciński2000) resembled CSRs at the base of surge-type glaciers (e.g. Evans and Rea, Reference Evans, Rea and Evans2003; Lovell and others, Reference Lovell2015a). Apart from the north-western forefield, most CSRs are subdued and take the form of low ridges or sediment trails (Figs 6 and 8), similar to those described in Trygghamna (Aradóttir and others, Reference Aradóttir2019; Ben-Yehoshua and others, Reference Ben-Yehoshua, Aradóttir, Farnsworth, Benediktsson and Ingólfsson2023). CSRs in the Hansbreen forefield are less abundant, lower and worse preserved than in many surge-type glaciers in Svalbard, e.g. Nathorstbreen (Lovell and others, Reference Lovell2018), Sefströmbreen (Boulton and others, Reference Boulton1996), Tunabreen, Von Postbreen (Lovell and others, Reference Lovell2015a) or the Hornbreen-Hambergbreen glacier system (Osika and others, Reference Osika, Jania and Szafraniec2022). Therefore, previous investigations could have omitted them without high-resolution remote sensing data.

Our study shows the importance of compiling terrestrial and submarine geomorphology with historical records for reconstructing past glacier dynamics. However, if the archival data are sparse, the identification of surge-type glaciers is impeded by the low preservation potential of surge-diagnostic landforms or topographic or geological conditions meaning that they do not form in the first place (e.g. Hansen, Reference Hansen2003; Farnsworth and others, Reference Farnsworth, Ingólfsson, Retelle and Schomacker2016; Ingólfsson and others, Reference Ingólfsson2016; Gądek and others, Reference Gądek, Rojan and Suska-Malawska2022). Therefore, we infer that the number of Svalbard surge-type glaciers or decaying glaciers that surged in the past can be underestimated if the geomorphological fingerprints have been obliterated and historical records are lacking or unrecognized. One example may be a cold-based valley Tellbreen, which was thought not to have ever surged (Bælum and Benn, Reference Bælum and Benn2011). However, the basal sequence and glaciological structures revealed that it had experienced warm-based conditions and dynamic ice flow, most probably during the LIA (Lovell and others, Reference Lovell2015b).

Interestingly, the 1870s surge of Hansbreen occurred in a period of numerous surge events in Svalbard between 1860s–1900s (e.g. Fridtjovbreen in 1861; Liestøl, Reference Liestøl, Williams and Ferrigno1993, Kronebreen, Kongsvegen and Kongsbreen in 1869; Liestøl, Reference Liestøl1988, Von Postbreen in 1870; Liestøl, Reference Liestøl, Williams and Ferrigno1993, Nathorstbreen in the 1870s or 1880s; Liestøl, Reference Liestøl1977; Ottesen and others, Reference Ottesen2008, Sefströmbreen in 1882–1886; De Geer, Reference De Geer1910, Recherchebreen and Renardbreen at the turn of the 1870s and 1880s; Zagórski and others, Reference Zagórski2023, Scottbreen in 1880s; Zagórski and others, Reference Zagórski2023, Hambergbreen c. 1900; Wassiliew, Reference Wassiliew1925 or Paulabreen before 1898; Ottesen and others, Reference Ottesen2008). The 1930s was another decade of apparent surge clustering (e.g. Bråsvellbreen in 1936–1938; Schytt, Reference Schytt1969, Negribreen in 1935–1936; Liestøl, Reference Liestøl1969, Körberbreen in 1938; Liestøl, Reference Liestøl1969, Markhambreen in 1930–1936; Hagen and others, Reference Hagen, Liestøl, Roland and Jørgensen1993, Arnesenbreen in 1925–1935; Hagen and others, Reference Hagen, Liestøl, Roland and Jørgensen1993 or Etonbreen, Rijpbreen and Clasebreen in 1938; Hagen and others, Reference Hagen, Liestøl, Roland and Jørgensen1993), although, despite some evidence, a surge of Hansbreen in this period remains questionable. Such synchronicity of surge advances might suggest regional factors affecting glacier dynamics. The reconstructed winter surface air temperatures in Longyearbyen were higher in the 1860s and early 1870s than in the 1850s or late 1870s (Divine and others, Reference Divine2011). The 1920s and 1930s corresponded to the early 20th-century climate warming in Svalbard (Nordli and others, Reference Nordli2020) and we speculate that these surges may have been related to increased water content at the glacier beds from enhanced summer melt and produced by frictional heating (cf. enthalpy balance theory by Benn and others, Reference Benn, Fowler, Hewitt and Sevestre2019). However, this observation requires further studies on the LIA climate conditions and investigation into triggering factors of glacier surges (e.g. Benn and others, Reference Benn, Hewitt and Luckman2023).

Repeated surge events of many Svalbard glaciers with different durations of the quiescent phase (e.g. Tunabreen; Flink and others, Reference Flink2015, Blomstrandbreen; Burton and others, Reference Burton, Dowdeswell, Hogan and Noormets2016, Negribreen; Haga and others, Reference Haga2020) raise the question about a potential future surge of Hansbreen. In our opinion, the surface topography of Hansbreen, i.e. the low-lying accumulation area of the main trunk and significant ice supply by the western tributary glaciers, as well as the subglacial topography with overdeepenings along the centreline, will hinder the build-up of mass and triggering the next surge. A dramatic increase in solid precipitation will be critical to fill up the reservoir area(s) and allow the rapid transfer of large amounts of ice to the receiving zone. We can speculate that such conditions will not be fulfilled in the following decades under projected climatic conditions in Svalbard (Hanssen-Bauer and others, Reference Hanssen-Bauer2019). However, one should not exclude possible dynamic advances of the western tributary glaciers after retreat of the main trunk. The loose of the back-stress from the Hansbreen main tongue may lead to re-organization of the surface profiles of the tributary glaciers through re-advances (Farnsworth and others, Reference Farnsworth2017).

Conclusions

We present a reconstruction of the past dynamics of Hansbreen by combining geomorphological mapping of the terrestrial and submarine forefields with historical data from Spitsbergen expeditions in the 19th and 20th centuries. Landform assemblages, including CSRs, flutings, glacial lineations and terminal moraines with debris flow lobes are consistent with the published surging glacier landsystem models and indicate at least one surge of Hansbreen.

Photographic documentation from the Austro-Hungarian expedition in 1872 revealed evidence of an ongoing active surge phase, such as a surge bulge, extensive surface crevassing, top-down and bottom-up crevasses and a convex surface profile of the terrestrial margin. The terminus had not yet reached the surge maximum. This surge corresponded to the maximum LIA extent of Hansbreen, marked by the outermost submarine terminal moraine with a debris flow lobe, terrestrial lateral terminal moraines and trimlines.

In 1918, the glacier was in the quiescent phase and the calving front was anchored on small islands and capes. The photogrammetric photos from the Norwegian expedition revealed further evidence of the surge before 1918, including a folded medial moraine and a dense, complex network of intersecting healed crevasses extending high up on the main trunk of Hansbreen. There may have been a subsequent surge in the 1930s based on a minor submarine terminal moraine with a debris flow lobe but this potential event requires further investigation.

Under projected climatic conditions, the low-lying accumulation area of the main trunk will hinder the mass build-up, and a potential next surge will probably remain questionable in the following decades. However, surges of the western tributary glaciers should not be excluded, in particular as a dynamic response when they lose the back-stress from the receding main tongue.

Acknowledgements

This study was funded by the National Science Centre of Poland (grant no. 2021/41/N/ST10/02070). The publication was co-financed by the Institute of Earth Sciences at the University of Silesia in Katowice. We sincerely thank Joanna Szafraniec for her comments, and Małgorzata Błaszczyk, Michał Ciepły, Dawid Saferna and Kamil Kachniarz (University of Silesia in Katowice) for their support during the fieldwork. We thank the crew of the Polish Polar Station Hornsund for logistic support and hospitality. The research and logistic equipment of the Polar Laboratory of the University of Silesia in Katowice was used during the fieldwork. We thank the Norwegian Hydrographic Service for the bathymetric data from Hansbukta and Isbjørnhamna. We are especially grateful to Österreichische Nationalbibliothek and Norsk Polarinstitutt for providing historical photographs from the polar expeditions to Hornsund in 1872 and 1918. We thank the Editor Andrew Fowler and in particular Harold Lovell for constructive comments and the insightful review, which significantly improved this manuscript.

Author contributions

A.O.: conceptualization, field investigation, data analysis. J.J.: conceptualization, field investigation, data analysis, supervision. A.O. structured the paper and wrote the manuscript with support from J.J.

Conflict of interest

The authors declare none.

Data availability

We re-used several published datasets available at individual databases. A very high-resolution orthophotomap of the Hansbreen forefield is available at https://ppdb.us.edu.pl/geonetwork/srv/eng/catalog.search#/metadata/9b980572-2bcb-401f-a2e6-2a7a16428dd7 (supplementary data to Błaszczyk and others, Reference Błaszczyk, Laska, Sivertsen and Jawak2022). A DEM of these areas is available at https://ppdb.us.edu.pl/geonetwork/srv/eng/catalog.search#/metadata/69987fe3-f0ed-4fdc-b597-41dd5ae7d21a (supplementary data to Błaszczyk and others, Reference Błaszczyk, Laska, Sivertsen and Jawak2022). The bathymetric data of Hansbukta are available at https://ppdb.us.edu.pl/geonetwork/srv/eng/catalog.search#/metadata/27151f9a-b3c4-4955-8db7-0365861f3dd6 (supplementary data to Błaszczyk and others, Reference Błaszczyk2021). Front positions of Hansbreen are available at https://ppdb.us.edu.pl/geonetwork/srv/eng/catalog.search#/metadata/05e2c69a-5645-4488-bc88-630beb03a462 and https://ppdb.us.edu.pl/geonetwork/srv/eng/catalog.search#/metadata/77c6fbd4-1a07-47b1-a663-559970517841 (supplementary data to Błaszczyk and others, Reference Błaszczyk, Jania and Kolondra2013, Reference Błaszczyk2021). Archival photographs from 1872 are available at https://onb.digital/search/766393 and from 1918 at https://bildearkiv.npolar.no/fotoweb/. Upon publication, the vector layers (shapefiles) of the geomorphological map will be available in the open database ZENODO: https://doi.org/10.5281/zenodo.13646525.

References

Andreassen, K and 11 others (2017) Massive blow-out craters formed by hydrate-controlled methane expulsion from the Arctic seafloor. Science (New York, N.Y.) 356(6341), 948953. doi: 10.1126/science.aal4500CrossRefGoogle ScholarPubMed
Aradóttir, N and 6 others (2019) Glacial geomorphology of Trygghamna, western Svalbard – Integrating terrestrial and submarine archives for a better understanding of past glacial dynamics. Geomorphology 344, 7589. doi: 10.1016/j.geomorph.2019.07.007CrossRefGoogle Scholar
Bælum, K and Benn, DI (2011) Thermal structure and drainage system of a small valley glacier (Tellbreen, Svalbard), investigated by ground penetrating radar. Cryosphere 5, 139149. doi: 10.5194/tc-5-139-2011CrossRefGoogle Scholar
Benn, DI (1994) Fluted moraine formation and till genesis below a temperate valley glacier: Slettmarkbreen, Jotunheimen, southern Norway. Sedimentology 41(2), 279292. doi: 10.1111/j.1365-3091.1994.tb01406.xCrossRefGoogle Scholar
Benn, DI and Evans, DJA (2010) Glaciers & Glaciation, 2nd Edn. London: Routledge. doi:10.5860/choice.35-6240Google Scholar
Benn, DI, Fowler, AC, Hewitt, I and Sevestre, H (2019) A general theory of glacier surges. Journal of Glaciology 65(253), 701716. doi: 10.1017/jog.2019.62CrossRefGoogle Scholar
Benn, DI, Hewitt, I and Luckman, A (2023) Enthalpy balance theory unifies diverse glacier surge behaviour. Annals of Glaciology 63, 8894. doi: 10.1017/aog.2023.23CrossRefGoogle Scholar
Ben-Yehoshua, D, Aradóttir, N, Farnsworth, WR, Benediktsson, ÍÖ and Ingólfsson, Ó (2023) Formation of crevasse-squeeze ridges at Trygghamna, Svalbard. Earth Surface Processes and Landforms 48(12), 23342348. doi: 10.1002/esp.5631CrossRefGoogle Scholar
Birkenmajer, K (1960) Raised marine features of the Hornsund area, Vestspitsbergen. Studia Geologica Polonica 5, 395.Google Scholar
Birkenmajer, K (1990) Geological map of the Hornsund Area 1:75 000. Katowice: Uniwersytet Śląski.Google Scholar
Błaszczyk, M, Jania, JA and Hagen, JO (2009) Tidewater glaciers of Svalbard: recent changes and estimates of calving fluxes. Polish Polar Research 30(2), 85142.Google Scholar
Błaszczyk, M, Jania, JA and Kolondra, L (2013) Fluctuations of tidewater glaciers in Hornsund Fjord (Southern Svalbard) since the beginning of the 20th century. Polish Polar Research 34(4), 327352. doi: 10.2478/popore-2013-0024CrossRefGoogle Scholar
Błaszczyk, M and 10 others (2019) Quality assessment and glaciological applications of digital elevation models derived from space-borne and aerial images over two tidewater glaciers of southern Spitsbergen. Remote Sensing 11(9), 1121. doi: 10.3390/rs11091121CrossRefGoogle Scholar
Błaszczyk, M and 12 others (2021) Factors controlling terminus position of Hansbreen, a tidewater glacier in Svalbard. Journal of Geophysical Research: Earth Surface 126(2), e2020JF005763. doi: 10.1029/2020JF005763CrossRefGoogle Scholar
Błaszczyk, M, Laska, M, Sivertsen, A and Jawak, SD (2022) Combined use of aerial photogrammetry and terrestrial laser scanning for detecting geomorphological changes in Hornsund, Svalbard. Remote Sensing 14(3), 601. doi: 10.3390/rs14030601CrossRefGoogle Scholar
Błaszczyk, M and 9 others (2023) The response of tidewater glacier termini positions in Hornsund (Svalbard) to climate forcing, 1992–2020. Journal of Geophysical Research: Earth Surface 128(5), e2022JF006911. doi: 10.1029/2022JF006911CrossRefGoogle Scholar
Błaszczyk, M and 7 others (2024) High temporal resolution records of the velocity of Hansbreen, a tidewater glacier in Svalbard. Earth Syst. Sci. Data 16(4), 18471860. doi: 10.5194/essd-16-1847-2024CrossRefGoogle Scholar
Boulton, GS (1968) Flow tills and related deposits on some Vestspitsbergen glaciers. Journal of Glaciology 7(51), 391412. doi: 10.3189/S0022143000020608CrossRefGoogle Scholar
Boulton, GS (1976) The origin of glacially fluted surfaces – Observations and theory. Journal of Glaciology 17(76), 287309. doi: 10.3189/S0022143000013605CrossRefGoogle Scholar
Boulton, GS (1986) Push-moraines and glacier-contact fans in marine and terrestrial environments. Sedimentology 33(5), 677698. doi: 10.1111/j.1365-3091.1986.tb01969.xCrossRefGoogle Scholar
Boulton, GS and 6 others (1996) Till and moraine emplacement in a deforming bed surge – an example from a marine environment. Quaternary Science Reviews 15(10), 961987. doi: 10.1016/0277-3791(95)00091-7CrossRefGoogle Scholar
Burton, DJ, Dowdeswell, JA, Hogan, KA and Noormets, R (2016) Marginal fluctuations of a Svalbard surge-type tidewater glacier, Blomstrandbreen, since the little ice age: a record of three surges. Arctic, Antarctic, and Alpine Research 48(2), 411426. doi: 10.1657/AAAR0014-094CrossRefGoogle Scholar
Chandler, BMP and 17 others (2018) Glacial geomorphological mapping: a review of approaches and frameworks for best practice. Earth-Science Reviews 185, 806846. doi: 10.1016/j.earscirev.2018.07.015CrossRefGoogle Scholar
Christoffersen, P, Piotrowski, JA and Larsen, NK (2005) Basal processes beneath an Arctic glacier and their geomorphic imprint after a surge, Elisebreen, Svalbard. Quaternary Research 64(2), 125137. doi: 10.1016/j.yqres.2005.05.009CrossRefGoogle Scholar
Copland, L, Sharp, MJ and Dowdeswell, JA (2003) The distribution and flow characteristics of surge-type glaciers in the Canadian high Arctic. Annals of Glaciology 36, 7381. doi: 10.3189/172756403781816301CrossRefGoogle Scholar
Ćwiąkała, J and 5 others (2018) Submarine geomorphology at the front of the retreating Hansbreen tidewater glacier, Hornsund fjord, southwest Spitsbergen. Journal of Maps 14(2), 123134. doi: 10.1080/17445647.2018.1441757CrossRefGoogle Scholar
De Andrés, E and 5 others (2018) A two-dimensional glacier–fjord coupled model applied to estimate submarine melt rates and front position changes of Hansbreen, Svalbard. Journal of Glaciology 64(247), 745758. doi: 10.1017/jog.2018.61CrossRefGoogle Scholar
De Geer, G (1910) Guide de l'excursion au Spitzberg. XIe Congrés Géologique Internationale, Stockholm.Google Scholar
Divine, D and 7 others (2011) Thousand years of winter surface air temperature variations in Svalbard and northern Norway reconstructed from ice core data. Polar Research 30, 7379. doi: 10.3402/polar.v30i0.7379CrossRefGoogle Scholar
Dowdeswell, JA and Ottesen, D (2016) Submarine landform assemblage for Svalbard surge-type tidewater glaciers. Geological Society, London, Memoirs 46, 151154. doi: 10.1144/M46.160CrossRefGoogle Scholar
Dowdeswell, JA, Hamilton, GS and Hagen, JO (1991) The duration of the active phase on surge-type glaciers: contrasts between Svalbard and other regions. Journal of Glaciology 37(127), 388400. doi: 10.3189/S0022143000005827CrossRefGoogle Scholar
Dowdeswell, JA, Hodgkins, R, Nuttall, M, Hagen, JO and Hamilton, GS (1995) Mass balance change as a control on the frequency and occurrence of glacier surges in Svalbard, Norwegian High Arctic. Geophysical Research Letters 22(21), 29092912. doi: 10.1029/95GL02821CrossRefGoogle Scholar
Dowdeswell, JA, Ottesen, D and Plassen, L (2016) Debris-flow lobes on the distal flanks of terminal moraines in Spitsbergen fjords. Geological Society, London, Memoirs 46(1), 7778. doi: 10.1144/M46.97CrossRefGoogle Scholar
Dowdeswell, JA, Ottesen, D and Bellec, VK (2020) The changing extent of marine-terminating glaciers and ice caps in northeastern Svalbard since the ‘Little Ice Age’ from marine-geophysical records. The Holocene 30(3), 389401. doi: 10.1177/0959683619887429CrossRefGoogle Scholar
Evans, DJA (2009) Controlled moraines: origins, characteristics and palaeoglaciological implications. Quaternary Science Reviews 28(3–4), 183208. doi: 10.1016/j.quascirev.2008.10.024CrossRefGoogle Scholar
Evans, DJA and Rea, BR (1999) Geomorphology and sedimentology of surging glaciers: a land-systems approach. Annals of Glaciology 28, 7582. doi: 10.3189/172756499781821823CrossRefGoogle Scholar
Evans, DJA and Rea, BR (2003) Surging glacier landsystem. In Evans, DJA (ed.), Glacial Landsystems. London: Arnold, pp. 259288.Google Scholar
Ewertowski, M, Kasprzak, L, Szuman, I and Tomczyk, AM (2011) Controlled, ice-cored moraines: sediments and geomorphology. An example from Ragnarbreen, Svalbard. Zeitschrift für Geomorphologie 56(1), 5374. doi: 10.1127/0372-8854/2011/0049CrossRefGoogle Scholar
Farnsworth, WR, Ingólfsson, Ó, Retelle, M and Schomacker, A (2016) Over 400 previously undocumented Svalbard surge-type glaciers identified. Geomorphology 264, 5260. doi: 10.1016/j.geomorph.2016.03.025CrossRefGoogle Scholar
Farnsworth, WR and 6 others (2017) Dynamic Holocene glacial history of St. Jonsfjorden, Svalbard. Boreas 46(3), 585603. doi: 10.1111/bor.12269CrossRefGoogle Scholar
Finsterwalder, R (1931) Geschwindigkeitsmessungen an Gletschern mittels Photogrammetrie. Z. Gletscherk 19(4–5), 251262.Google Scholar
Flink, AE and Noormets, R (2018) Submarine glacial landforms and sedimentary environments in Vaigattbogen, northeastern Spitsbergen. Marine Geology 402, 244263. doi: 10.1016/j.margeo.2017.07.019CrossRefGoogle Scholar
Flink, AE and 5 others (2015) The evolution of a submarine landform record following recent and multiple surges of Tunabreen glacier, Svalbard. Quaternary Science Reviews 108, 3750. doi: 10.1016/j.quascirev.2014.11.006CrossRefGoogle Scholar
Flink, AE, Hill, P, Noormets, R and Kirchner, N (2018) Holocene glacial evolution of Mohnbukta in eastern Spitsbergen. Boreas 47, 390409. doi: 10.1111/bor.12277CrossRefGoogle Scholar
Forwick, M, Baeten, NJ and Vorren, TO (2009) Pockmarks in Spitsbergen fjords. Norwegian Journal of Geology 89, 6577.Google Scholar
Gądek, B, Rojan, E and Suska-Malawska, M (2022) Surge-type Uisu glacier and its undisturbed forefield relief, Eastern Pamir, Tajikistan. Misc. Geogr 26(4), 227236. doi: 10.2478/mgrsd-2022-0009Google Scholar
Geyman, EC, van Pelt, WJJ, Maloof, AC, Aas, HF and Kohler, J (2022) Historical glacier change on Svalbard predicts doubling of mass loss by 2100. Nature 601(7893), 374379. doi: 10.1038/s41586-021-04314-4CrossRefGoogle ScholarPubMed
Glasser, NF, Hambrey, MJ, Crawford, KR, Bennett, MR and Huddart, D (1998) The structural glaciology of Kongsvegen, Svalbard, and its role in landform genesis. Journal of Glaciology 44(146), 136148. doi: 10.3189/S0022143000002422CrossRefGoogle Scholar
Glazer, M, Dobiński, W, Marciniak, A, Majdański, M and Błaszczyk, M (2020) Spatial distribution and controls of permafrost development in non-glacial Arctic catchment over the Holocene, Fuglebekken, SW Spitsbergen. Geomorphology 358, 107128. doi: 10.1016/j.geomorph.2020.107128CrossRefGoogle Scholar
Glazovskiy, AF, Kolondra, L, Moskalevskiy, MY and Jania, J (1992) Studies on the tidewater glacier Hansbreen on Spitsbergen. Polar Geography and Geology 16, 243252. doi: 10.1080/10889379209377492CrossRefGoogle Scholar
Grabiec, M, Jania, JA, Puczko, D, Kolondra, L and Budzik, T (2012) Surface and bed morphology of Hansbreen, a tidewater glacier in Spitsbergen. Polish Polar Research 33(2), 111138. doi: 10.2478/v10183-012-0010-7CrossRefGoogle Scholar
Grant, KL, Stokes, CR and Evans, IS (2009) Identification and characteristics of surge-type glaciers on Novaya Zemlya, Russian Arctic. Journal of Glaciology 55(194), 960972. doi: 10.3189/002214309790794940CrossRefGoogle Scholar
Haga, ON and 5 others (2020) From high friction zone to frontal collapse: dynamics of an ongoing tidewater glacier surge, Negribreen, Svalbard. Journal of Glaciology 66(259), 742754. doi: 10.1017/jog.2020.43CrossRefGoogle Scholar
Hagen, JO and Liestøl, O (1990) Long-term glacier mass-balance investigations in Svalbard, 1950–88. Annals of Glaciology 14, 102106. doi: 10.3189/S0260305500008351CrossRefGoogle Scholar
Hagen, JO, Liestøl, O, Roland, E and Jørgensen, T (1993) Glacier atlas of Svalbard and Jan Mayen. Nor. Polarinst. Medd. 129.Google Scholar
Hansen, S (2003) From surge-type to non-surge-type glacier behaviour: midre Lovénbreen, Svalbard. Annals of Glaciology 36, 97102. doi: 10.3189/172756403781816383CrossRefGoogle Scholar
Hanssen-Bauer, I and 5 others (eds) (2019) Climate in Svalbard 2100 – a knowledge base for climate adaptation. Norwegian Centre of Climate Services (NCCS) for Norwegian Environment Agency (Miljødirektoratet). doi: 10.25607/OBP-888CrossRefGoogle Scholar
Heintz, A (1953) Non iakktagelser over isbreenes tilbakegang i Hornsund. Norsk Geografisk Tidsskrift 31, 736.Google Scholar
Hoel, A (1929) The Norwegian Svalbard Expeditions 1906-1926, vol. 1. Oslo: Skrifter om Svalbard og Ishavet.Google Scholar
Ingólfsson, Ó and 7 others (2016) Glacial geological studies of surge-type glaciers in Iceland – research status and future challenges. Earth-Science Reviews 152, 3769. doi: 10.1016/j.earscirev.2015.11.008CrossRefGoogle Scholar
Ives, LRW and Iverson, NR (2019) Genesis of glacial flutes inferred from observations at Múlajökull, Iceland. Geology 47(5), 387390. doi: 10.1130/G45714.1CrossRefGoogle Scholar
Jania, J (1988) Dynamiczne procesy glacjalne na południowym Spitsbergenie (w świetle badań fotointerpretacyjnych i fotogrametrycznych) [Dynamic glacial processes in south Spitsbergen (in the light of photointerpretation and photogrammetric research)]. Katowice: Wydawnictwo Uniwersytetu Śląskiego (in Polish with English summary).Google Scholar
Jania, J (1998) Dynamika Lodowca Hansa (Spitsbergen, Svalbard), a wybrane elementy rzeźby na jego przedpolu [Dynamics of the Hans Glacier (Spitsbergen, Svalbard) and selected elements of relief in its forefield. In Kostrzewski, A (ed.), Rzeźba i osady czwartorzędowe obszarów współczesnego i plejstoceńskiego zlodowacenia półkuli północnej. Poznań: Wydawnictwo Naukowe UAM, pp. 8196 (In Polish with English summary).Google Scholar
Jania, J and Głowacki, P (1996) Is the Hansbreen in South Spitsbergen (Svalbard) a surge-type glacier? In Krawczyk WE ed. 23rd Polar Symposium, Uniwersytet Śląski, Sosnowiec, 2743.Google Scholar
Jania, J and Kolondra, L (1982) Field investigations performed during the Glaciological Spitsbergen Expedition in the Summer of 1982. Interim Report, Uniwersytet Śląski, Katowice.Google Scholar
Jania, J, Mochnacki, D and Gądek, B (1996) The thermal structure of Hansbreen, a tidewater glacier in southern Spitsbergen, Svalbard. Polar Research 15, 5366. doi: 10.1111/j.1751-8369.1996.tb00458.xCrossRefGoogle Scholar
Jiskoot, H, Boyle, P and Murray, T (1998) The incidence of glacier surging in Svalbard: evidence from multivariate statistics. Computers & Geosciences 24(4), 387399. doi: 10.1016/S0098-3004(98)00033-8CrossRefGoogle Scholar
Jiskoot, H, Murray, T and Boyle, P (2000) Controls on the distribution of surge-type glaciers in Svalbard. Journal of Glaciology 46(154), 412422. doi: 10.3189/172756500781833115CrossRefGoogle Scholar
Karczewski, A and 13 others (1984) Hornsund, Spitsbergen – Geomorphology, 1:75000. Katowice: University of Silesia.Google Scholar
Koryakin, VM (1974) Izmjenienje razmerov lednikov Szpicbergena (Svalnarda). Materials of the Investigations in the Glaciation Region of the Spitsbergen (Svalbard). Moscow: AN SSSR [Acad. of Sci. of the USSR], pp. 2944.Google Scholar
Kosiba, A (1960) Some Results of Glaciological Investigations in SW-Spitsbergen Carried Out During the Polish I. G. Y. Spitsbergen Expeditions in 1957, 1958 and 1959. Wrocław: Uniwersytet Wrocławski.Google Scholar
Laska, M and 8 others (2022) Hansbreen snowpit dataset – over 30-year of detailed snow research on an Arctic glacier. Scientific Data 9, 656. doi: 10.1038/s41597-022-01767-8CrossRefGoogle Scholar
Lefauconnier, B and Hagen, JO (1991) Surging and calving glaciers in eastern Svalbard. Nor. Polarinst. Medd 116, 1130.Google Scholar
Liestøl, O (1969) Glacier surges in West Spitsbergen. Canadian Journal of Earth Sciences 6(4), 895897. doi: 10.1139/e69-092CrossRefGoogle Scholar
Liestøl, O (1977) Årsmorener foran Nathorstbreen? Nor. Polarinst. Årb 1976, 361363.Google Scholar
Liestøl, O (1988) The glaciers in the Kongsfjorden area, Spitsbergen. Nor. Geogr. Tidsskr 42(4), 231238. doi: 10.1080/00291958808552205CrossRefGoogle Scholar
Liestøl, O (1993) Glaciers of Svalbard, Norway. In Williams, RS Jr and Ferrigno, JG (eds), Satellite Image Atlas of Glaciers of the World. Washington, DC: US Geological Survey, pp. 127151.Google Scholar
Lindner, L, Marks, L and Szczęsny, R (1992) Quaternary landforms, sediments and morphogenetic evolution of Hansbreen-Sofiekammen region, Wedel Jarlsberg Land, Spitsbergen. Polish Polar Research 13(2), 91101.Google Scholar
Lønne, I (2016) A new concept for glacial geological investigations of surges, based on high-Arctic examples (Svalbard). Quaternary Science Reviews 132, 74100. doi: 10.1016/j.quascirev.2015.11.009CrossRefGoogle Scholar
Lovell, H and Boston, CM (2017) Glacitectonic composite ridge systems and surge-type glaciers: an updated correlation based on Svalbard, Norway. Arktos 3, 2. doi: 10.1007/s41063-017-0028-5CrossRefGoogle Scholar
Lovell, H and Fleming, E (2023) Structural evolution during a surge in the Paulabreen glacier system, Svalbard. Journal of Glaciology 69(273), 141152. doi: 10.1017/jog.2022.53CrossRefGoogle Scholar
Lovell, H and 7 others (2015a) Debris entrainment and landform genesis during tidewater glacier surges. Journal of Geophysical Research: Earth Surface 120(8), 15741595. doi: 10.1002/2015JF003509CrossRefGoogle Scholar
Lovell, H and 5 others (2015b) Former dynamic behaviour of a cold-based valley glacier on Svalbard revealed by basal ice and structural glaciology investigations. Journal of Glaciology 61(226), 309328. doi: 10.3189/2015JoG14J120CrossRefGoogle Scholar
Lovell, H and 8 others (2018) Multiple Late Holocene surges of a high-Arctic tidewater glacier system in Svalbard. Quaternary Science Reviews 201, 162185. doi: 10.1016/j.quascirev.2018.10.024CrossRefGoogle Scholar
Luncke, B (1936) Luftkartlegningen på Svalbard 1936. Norsk Geogr. Tidsskr 6, 145154.CrossRefGoogle Scholar
Maussion, F and 11 others (2023) The Randolph Glacier Inventory version 7.0, User guide v1.0. doi: 10.5281/zenodo.8362857CrossRefGoogle Scholar
Meier, MF and Post, A (1969) What are glacier surges? Canadian Journal of Earth Sciences 6(4), 807817. doi: 10.1139/e69-081CrossRefGoogle Scholar
Murray, T, Strozzi, T, Luckman, A, Jiskoot, H and Christakos, P (2003) Is there a single surge mechanism? Contrasts in dynamics between glacier surges in Svalbard and other regions. Journal of Geophysical Research: Solid Earth 108(B5), 2237. doi: 10.1029/2002JB001906CrossRefGoogle Scholar
Noormets, R, Flink, A and Kirchner, N (2021) Glacial dynamics and deglaciation history of Hambergbukta reconstructed from submarine landforms and sediment cores, SE Spitsbergen, Svalbard. Boreas 50, 2950. doi: 10.1111/bor.12488CrossRefGoogle Scholar
Nordli, Ø and 6 others (2020) Revisiting the extended Svalbard Airport monthly temperature series, and the compiled corresponding daily series 1898–2018. Polar Research 39, 3614. doi: 10.33265/polar.v39.3614CrossRefGoogle Scholar
Norwegian Polar Institute (2014) Kartdata Svalbard 1:1 000 000 (S1000 Kartdata) [Data set]. Tromsø: Norwegian Polar Institute. doi: 10.21334/npolar.2014.63730e2eGoogle Scholar
Nuth, C and 7 others (2013) Decadal changes from a multi-temporal glacier inventory of Svalbard. Cryosphere 7, 16031621. doi: 10.5194/tc-7-1603-2013CrossRefGoogle Scholar
Ohta, Y and Dallmann, WK (eds) (1999) Geological map of Svalbard 1:100 000. Sheet B12 G Torellbreen. Norsk Polarinstitutt Temakart Nr. 29.Google Scholar
Osika, A, Jania, J and Szafraniec, JE (2022) Holocene ice-free strait followed by dynamic Neoglacial fluctuations: Hornsund, Svalbard. The Holocene 32(7), 664679. doi: 10.1177/09596836221088232CrossRefGoogle Scholar
Otero, J and 5 others (2017) Modeling the controls on the front position of a tidewater glacier in Svalbard. Frontiers in Earth Science 5, 243157. doi: 10.3389/feart.2017.00029CrossRefGoogle Scholar
Ottesen, D and Dowdeswell, JA (2006) Assemblages of submarine landforms produced by tidewater glaciers in Svalbard. Journal of Geophysical Research: Earth Surface 111, F01016. doi: doi:10.1029/2005JF000330CrossRefGoogle Scholar
Ottesen, D and 9 others (2008) Submarine landforms characteristic of glacier surges in two Spitsbergen fjords. Quaternary Science Reviews 27(15–16), 15831599. doi: 10.1016/j.quascirev.2008.05.007CrossRefGoogle Scholar
Ottesen, D, Dowdeswell, J, Bellec, V and Bjarnadóttir, L (2017) The geomorphic imprint of glacier surges into open-marine waters: examples from eastern Svalbard. Marine Geology 392, 129. doi: 10.1016/j.margeo.2017.08.007CrossRefGoogle Scholar
Pälli, A, Moore, JC, Jania, J, Kolondra, L and Glowacki, P (2003) The drainage pattern of Hansbreen and Werenskioldbreen, two polythermal glaciers in Svalbard. Polar Research 22(2), 355371. doi: 10.3402/polar.v22i2.6465CrossRefGoogle Scholar
Pękala, K (1989) Quaternary deposits of the Hans Glacier forefield (Hornsund, Spitsbergen). Polar Session. Natural Environment Research of West Spitsbergen. Lublin: Wydawnictwo UMCS, pp. 191204.Google Scholar
Peterman, A (1874) Graf Wiltschek's Nordpolarfahrt im Jahre 1872 / nach den Aufzeichnungen des Contre-Admirals Max Freiherrn Daublebsky v. Sterneck und Ehrenstein. In Peterman A ed. Mittheilungen aus Justus Perthes' Geographischer Anstalt über wichtige neue Erforschungen auf dem Gesammtgebiete der Geographie, 20 Band, Justus Perthes' Geographische Anstalt Gotha, 6572.Google Scholar
Pillewizer, W (1939) Die kartographischen und gletscherkundlichen Ergebnisse der deutschen Spitzbergen-Expedition 1938, Justus Perthes Verlag Gotha.Google Scholar
Rachlewicz, G and Szczuciński, W (2000) Ice tectonics and bedrock relief control on glacial sedimentation – an example from Hansbreen, Spitsbergen. In Grześ M, Lankauf KR and Sobota I eds. Polish Polar Studies. 27th Polar Symposium, 1–3 December 2000, Toruń, Pracownia Sztuk Plastycznych, 259275.Google Scholar
Raymond, CF (1987) How do glaciers surge? A review. Journal of Geophysical Research: Solid Earth 92(B9), 91219134. doi: 10.1029/JB092iB09p09121CrossRefGoogle Scholar
Rea, BR and Evans, DJA (2011) An assessment of surge-induced crevassing and the formation of crevasse squeeze ridges. Journal of Geophysical Research: Earth Surface 116, F04005. doi: 10.1029/2011JF001970CrossRefGoogle Scholar
RGI 7.0 Consortium (2023) Randolph Glacier Inventory – A Dataset of Global Glacier Outlines, Version 7.0. Boulder, Colorado, USA: NSIDC: National Snow and Ice Data Center. doi: 10.5067/f6jmovy5navzGoogle Scholar
Robinson, P and Dowdeswell, JA (2011) Submarine landforms and the behavior of a surging ice cap since the last glacial maximum: the open-marine setting of eastern Austfonna, Svalbard. Marine Geology 286(1–4), 8294. doi: 10.1016/j.margeo.2011.06.004CrossRefGoogle Scholar
Rootes, CM and Clark, CD (2022) On the expression and distribution of glacial trimlines: a case study of Little Ice Age trimlines on Svalbard. E&G Quaternary Science Journal 71, 111122. doi: 10.5194/egqsj-71-111-2022Google Scholar
Schomacker, A and Kjær, KH (2008) Quantification of dead-ice melting in ice-cored moraines at Holmströmbreen. Svalbard. Boreas 37(2), 211225. doi: 10.1111/j.1502-3885.2007.00014.xCrossRefGoogle Scholar
Schuler, TV and 12 others (2020) Reconciling Svalbard glacier mass balance. Frontiers in Earth Science 8, 523648. doi: 10.3389/feart.2020.00156CrossRefGoogle Scholar
Schytt, V (1969) Some comments on glacier surges in eastern Svalbard. Canadian Journal of Earth Sciences 6(4), 867873. doi: 10.1139/e69-088CrossRefGoogle Scholar
Sevestre, H and Benn, D (2015) Climatic and geometric controls on the global distribution of surge-type glaciers: implications for a unifying model of surging. Journal of Glaciology 61(228), 646662. doi: 10.3189/2015JoG14J136CrossRefGoogle Scholar
Sevestre, H, Benn, DI, Hulton, NJR and Bælum, K (2015) Thermal structure of Svalbard glaciers and implications for thermal switch models of glacier surging. Journal of Geophysical Research: Earth Surface 120(10), 22202236. doi: 10.1002/2015JF003517CrossRefGoogle Scholar
Skolasińska, K, Rachlewicz, G and Szczuciński, W (2016) Micromorphology of modern tills in southwestern Spitsbergen – insights into depositional and post-depositional processes. Polish Polar Research 37(4), 435456. doi: 10.1515/popore-2016-0023CrossRefGoogle Scholar
Sobota, I, Weckwerth, P and Nowak, M (2016) Surge dynamics of Aavatsmarkbreen, Svalbard, inferred from the geomorphological record. Boreas 45(2), 360376. doi: 10.1111/bor.12160CrossRefGoogle Scholar
Spagnolo, M and 7 others (2014) Size, shape and spatial arrangement of mega-scale glacial lineations from a large and diverse dataset. Earth Surface Processes and Landforms 39(11), 14321448. doi: 10.1002/esp.3532CrossRefGoogle Scholar
Stokes, CR and Clark, CD (2002) Are long subglacial bedforms indicative of fast ice flow? Boreas 31(3), 239249. doi: 10.1111/j.1502-3885.2002.tb01070.xCrossRefGoogle Scholar
Streuff, K, Forwick, M, Szczuciński, W, Andreassen, K and Cofaigh, (2015) Submarine landform assemblages and sedimentary processes related to glacier surging in Kongsfjorden, Svalbard. Arktos 1, 14. doi: 10.1007/s41063-015-0003-yCrossRefGoogle Scholar
Streuff, K, Cofaigh, , Noormets, R and Lloyd, J (2018) Submarine landform assemblages and sedimentary processes in front of Spitsbergen tidewater glaciers. Marine geology 402, 209227. doi: 10.1016/j.margeo.2017.09.006CrossRefGoogle Scholar
Streuff, KT, Cofaigh, and Wintersteller, P (2022) Glacidat – a GIS database of submarine glacial landforms and sediments in the Arctic. Boreas 51(3), 517531. doi: 10.1111/bor.12577CrossRefGoogle Scholar
Sund, M, Eiken, T, Hagen, JO and Kääb, A (2009) Svalbard surge dynamics derived from geometric changes. Annals of Glaciology 50(52), 5060. doi: 10.3189/172756409789624265CrossRefGoogle Scholar
Szafraniec, JE (2020) Ice-cliff morphometry in identifying the surge phenomenon of tidewater glaciers (Spitsbergen, Svalbard). Geosciences 10, 328. doi: 10.3390/geosciences10090328CrossRefGoogle Scholar
Tegowski, J, Trzcinska, K, Kasprzak, M and Nowak, J (2016) Statistical and spectral features of corrugated seafloor shaped by the Hans glacier in Svalbard. Remote Sensing 8(9), 744. doi: 10.3390/rs8090744CrossRefGoogle Scholar
Vieli, A, Jania, J and Kolondra, L (2002) The retreat of a tidewater glacier: observations and model calculations on Hansbreen, Spitsbergen. Journal of Glaciology 48(163), 592600. doi: 10.3189/172756502781831089CrossRefGoogle Scholar
Wassiliew, AS (1925) Océan Glacial arctique. Spitzberg. Région des travaux de l'expédition de l'Académies de sciences de Russie pour la mesure d'un arc de méridien en 1899–1901. Carte dresie d'après les matériaux de l'expédition sous la rédactin de O.E. Stubendorff par A.S. Wassiliew. Échelle de 1:200 000.Google Scholar
Wawrzyniak, T and Osuch, M (2019) A consistent high Arctic climatological dataset (1979-2018) of the Polish polar station Hornsund (SW Spitsbergen, Svalbard). PANGAEA. doi: 10.1594/PANGAEA.909042Google Scholar
Wawrzyniak, T and Osuch, M (2020) A 40-year high Arctic climatological dataset of the Polish polar station Hornsund (SW Spitsbergen, Svalbard). Earth Syst. Sci. Data 12, 805815. doi: 10.5194/essd-12-805-2020CrossRefGoogle Scholar
Weniger, P and 6 others (2019) Origin of near-surface hydrocarbon gases bound in northern Barents Sea sediments. Marine and Petroleum Geology 102, 455476. doi: 10.1016/j.marpetgeo.2018.12.036CrossRefGoogle Scholar
Yde, JC and 5 others (2019) Kuannersuit Glacier revisited: constraining ice dynamics, landform formations and glaciomorphological changes in the early quiescent phase following the 1995–98 surge event. Geomorphology 330, 8999. doi: 10.1016/j.geomorph.2019.01.012CrossRefGoogle Scholar
Zagórski, P and 5 others (2023) Surges in three Svalbard glaciers derived from historic sources and geomorphic features. Annals of the American Association of Geographers 113(8), 18351855. doi: 10.1080/24694452.2023.2200487CrossRefGoogle Scholar
Figure 0

Figure 1. (a) Location of Hansbreen in the Svalbard archipelago and (b, c) Hornsund. (b) Glaciers: Ho, Hornbreen; K, Körberbreen; Me, Mendeleevbreen; Mu, Mühlbacherbreen; P, Paierlbreen; Sa, Samarinbreen; St, Storbreen; Sv, Svalisbreen. (c) Glaciers (black letters): D, Deileggbreen; F, Fuglebreen; K, Kvitungisen; S, Staszelisen; T, Tuvbreen; V, Vrangpeisbreen. Mountain ridges; peaks and saddles (yellow letters): Fa, Fannytoppen; Fb, Fugleberget; Fp, Flatpasset; Fr, Flatryggen; VT, Vesletuva. Peninsulas and islands (orange letters): B, Baranowski Peninsula (Baranowskiodden); Hh, Hansholmane; O, Oseanograftangen. Bays: Hb, Hansbukta; I, Isbjørnhamna. PPS, Polish Polar Station Hornsund. Figure map: © Norwegian Polar Institute (Norwegian Polar Institute, 2014). Sentinel-2B from 27 July 2023.

Figure 1

Figure 2. Hansbreen and its forefield. Orthophotomap from June 2020 after Błaszczyk and others (2022) and bathymetric map based on the Norwegian Hydrographic Service multibeam data (2008) combined with bathymetric data (2014–2017) after Błaszczyk and others (2021) were used to produce the geomorphological map (Fig. 3). Changes in the frontal position of Hansbreen after Błaszczyk and others (2013, 2021), Sentinel-2A from 2 August 2016 and 18 September 2018, and Jania (1988). The dashed line marks the glacier extent in 1938 based on terraphotogrammetric measurements conducted by Pillewizer (1939).

Figure 2

Figure 3. Geomorphological map of the forefield of Hansbreen. Glacier extent on 22 June 2020.

Figure 3

Figure 4. Streamlined topography of (a) the Baranowski Peninsula and (b) at the foot of Fannytoppen with parallel-sided flutings on a thin layer of till. Note CSRs transverse and oblique to flutes on the Baranowski Peninsula. Shaded relief maps after Błaszczyk and others (2022).

Figure 4

Figure 5. Fluted till surface with CSRs at the foot of Fugleberget in the north-western part of the forefield. (a) Overview of the flutings and CSRs networks on the orthophotomap from 2020 (Błaszczyk and others, 2022). (b) Flutings and CSRs emerging from downwasting glacier ice in August 2020. Hummocky moraine and trimline on the Fugleberget slopes in the background. (c) The same area in September 2023. Note two drumlins (dr) and a roche moutonné (rm) overprinted with flutings and well-developed CSRs. (d) A prominent CSR c. 0.8 m high with the extent marked with the dashed line. (b–d) Photos A. Osika.

Figure 5

Figure 6. (a, b) Roches moutonnées on Oseanograftangen overprinted with fluted till and CSRs. Note the extensive beaches and storm ridges around the hills and remaining of the plain south of the lake, indicating the flutes and CSRs reworking by coastal processes. (c) Levelled flutes (parallel; blue lines) and CSRs (oblique and transverse; red lines). The latter take the form of low ridges (0.3–0.5 m) or diamicton trails a few centimetres high.

Figure 6

Figure 7. (a, b) Terrestrial photogrammetric photograph of the Baranowski Peninsula from 1982. Note long, parallel-sided flutings and low CSRs (white arrows) oblique and transverse to the flutes. Today, CSRs in this site are sparse and have an indistinct appearance in the field.

Figure 7

Figure 8. (a) Low, obliterated CSRs c. 0.3–0.5 m high among sediment trails reflecting the cross-cutting pattern of CSRs and flutings at the foot of Fannytoppen. (b) Ice-cored hummocky moraine in this area. In the background, hummocky moraine and trimline on the slopes of Fugleberget. (c) Hummocky moraine (h) and trimline (t) on the Fannytoppen slopes. The arrow marks the lateral terminal moraine and the rectangle illustrates the location of the roches moutonnées with CSRs on Oseanograftangen. (a–c) Photos A. Osika.

Figure 8

Figure 9. (a) Hansbreen on a 1: 100 000 map ‘Spitsbergens Kyst, Hornsund til Dunder Bay’ based on the Norwegian Spitsbergen expeditions in 1917–1921 with photogrammetric measurements in Hornsund in 1918 (public domain, https://data.npolar.no). (b–g) Terrestrial photogrammetric documentation of Hansbreen from 1918. (b) The radially expanded glacier snout after surge advance. Note a folded medial moraine in the eastern part of the glacier. (c–f) A complex network of narrow, cross-cutting crevasses and crevasse traces documented from Fannytoppen. (g) The glacier front in Isbjørnhamna. Note a waterfall from an active supraglacial channel in the background. (b) © Olaf Holtedahl, Norsk Polarinstitutt (modified), (c–f) © Jørgen Gløersen, Norsk Polarinstitutt (modified), (g) © Adolf Hoel, Norsk Polarinstitutt (cropped and modified).

Figure 9

Figure 10. Photographs of Hansbreen from the Austro-Hungarian expedition in 1872. (a) The western terrestrial margin of the glacier. (b) Crevassed marine-terminating zone with a conspicuous surface bulge. (c, d) The eastern terrestrial margin, where the glacier had not yet reached the maximum extent. Note folded and inclined debris layers, top-down and bottom-up crevasses exposed on the ice cliff and a convex surface profile of the glacier. © ÖNB Vienna: Image ID 00299043, 00480587 and 00444839 (cropped and modified).

Figure 10

Figure 11. Hansbreen in the oblique aerial photograph from 1936. © Norwegian Polar Institute.