Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Historical glacier change on Svalbard predicts doubling of mass loss by 2100

Abstract

The melting of glaciers and ice caps accounts for about one-third of current sea-level rise1,2,3, exceeding the mass loss from the more voluminous Greenland or Antarctic Ice Sheets3,4. The Arctic archipelago of Svalbard, which hosts spatial climate gradients that are larger than the expected temporal climate shifts over the next century5,6, is a natural laboratory to constrain the climate sensitivity of glaciers and predict their response to future warming. Here we link historical and modern glacier observations to predict that twenty-first century glacier thinning rates will more than double those from 1936 to 2010. Making use of an archive of historical aerial imagery7 from 1936 and 1938, we use structure-from-motion photogrammetry to reconstruct the three-dimensional geometry of 1,594 glaciers across Svalbard. We compare these reconstructions to modern ice elevation data to derive the spatial pattern of mass balance over a more than 70-year timespan, enabling us to see through the noise of annual and decadal variability to quantify how variables such as temperature and precipitation control ice loss. We find a robust temperature dependence of melt rates, whereby a 1 °C rise in mean summer temperature corresponds to a decrease in area-normalized mass balance of −0.28 m yr−1 of water equivalent. Finally, we design a space-for-time substitution8 to combine our historical glacier observations with climate projections and make first-order predictions of twenty-first century glacier change across Svalbard.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Three-dimensional models of Svalbard glaciers.
Fig. 2: Elevation change from 1936 to approximately 2010.
Fig. 3: Temperature control on glacier mass balance.
Fig. 4: Space-for-time substitution.

Similar content being viewed by others

Data availability

The 1936/1938 Svalbard glacier inventory presented here consists of raster DEMs and orthophotos (5 m resolution), and vector outlines of glacier extents (Extended Data Fig. 3). All data are publicly available on the NPI website (https://www.doi.org/10.21334/npolar.2021.f6afca5c) and on Zenodo (https://doi.org/10.5281/zenodo.5806388). In these repositories, we also provide the raw (unprocessed) 3D point clouds as .laz files and a spreadsheet (.xlsx file) containing glacier-by-glacier estimates of area, volume, hypsometry, ∆h/∆t, bed slope, DEM uncertainty and climate fields (mean annual temperature, mean summer temperature, PDDs, precipitation as snow and total precipitation). The original 1936/1938 aerial images and their locations can be viewed at https://toposvalbard.npolar.no/. The 5 m regional DEMs from the 2008–2012 survey33 are available as .tif files from https://publicdatasets.data.npolar.no/kartdata/S0_Terrengmodell/Delmodell/ and the associated 50 cm orthophotomosaic is available as a WMTS layer from https://geodata.npolar.no/#basemap-data.

Code availability

The code developed to analyse the 1936–2010 mass balance data and implement the space-for-time substitution is available on Zenodo (https://doi.org/10.5281/zenodo.5643856).

References

  1. Meier, M. F. et al. Glaciers dominate eustatic sea-level rise in the 21st century. Science 317, 1064–1067 (2007).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Gardner, A. S. et al. A reconciled estimate of glacier contributions to sea level rise: 2003 to 2009. Science 340, 852–857 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  3. Zemp, M. et al. Global glacier mass changes and their contributions to sea-level rise from 1961 to 2016. Nature 568, 382–386 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Hugonnet, R. et al. Accelerated global glacier mass loss in the early twenty-first century. Nature 592, 726–731 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  5. van Pelt, W. et al. A long-term dataset of climatic mass balance, snow conditions, and runoff in Svalbard (1957–2018). Cryosphere 13, 2259–2280 (2019).

    Article  ADS  Google Scholar 

  6. Hanssen-Bauer, I. et al. Climate in Svalbard 2100 – A Knowledge Base for Climate Adaptation NCCS Report no 1/2019 commissioned by the Norwegian Environment Agency (Miljødirektoratet) (Norwegian Centre for Climate Services, 2019).

  7. Luncke, B. Luftkartlegningen på Svalbard 1936. Norsk Geogr. Tidsskr. 6, 145–154 (1936).

    Article  Google Scholar 

  8. Blois, J. L., Williams, J. W., Fitzpatrick, M. C., Jackson, S. T. & Ferrier, S. Space can substitute for time in predicting climate-change effects on biodiversity. Proc. Natl Acad. Sci. USA 110, 9374–9379 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Serreze, M. C. & Barry, R. G. Processes and impacts of Arctic amplification: a research synthesis. Global Planet. Change 77, 85–96 (2011).

    Article  ADS  Google Scholar 

  10. Isaksen, K. et al. Recent warming on Spitsbergen – influence of atmospheric circulation and sea ice cover. J. Geophys. Res. Atmos. 121, 11–913 (2016).

    Article  Google Scholar 

  11. Nordli, Ø. et al. Revisiting the extended Svalbard Airport monthly temperature series, and the compiled corresponding daily series 1898–2018. Polar Res. (2020).

  12. Nuth, C., Moholdt, G., Kohler, J., Hagen, J. O. & Kääb, A. Svalbard glacier elevation changes and contribution to sea level rise. J. Geophys. Res. Earth Surf. 115 (2010).

  13. Dowdeswell, J. A. et al. The mass balance of circum-Arctic glaciers and recent climate change. Quat. Res. 48, 1–14 (1997).

    Article  Google Scholar 

  14. Hagen, J. O., Kohler, J., Melvold, K. & Winther, J.-G. Glaciers in Svalbard: mass balance, runoff and freshwater flux. Polar Res. 22, 145–159 (2003).

    Article  Google Scholar 

  15. Hagen, J. O., Melvold, K., Pinglot, F. & Dowdeswell, J. A. On the net mass balance of the glaciers and ice caps in Svalbard, Norwegian Arctic. Arct. Antarct. Alp. Res. 35, 264–270 (2003).

    Article  Google Scholar 

  16. Möller, M. & Kohler, J. Differing climatic mass balance evolution across Svalbard glacier regions over 1900–2010. Front. Earth Sci. 6, 128 (2018).

    Article  ADS  Google Scholar 

  17. Østby, T. I. et al. Diagnosing the decline in climatic mass balance of glaciers in Svalbard over 1957-2014. Cryosphere 11, 191–215 (2017).

    Article  ADS  Google Scholar 

  18. Aas, K. S. et al. The climatic mass balance of Svalbard glaciers: a 10-year simulation with a coupled atmosphere-glacier mass balance model. Cryosphere 10, 1089–1104 (2016).

    Article  ADS  Google Scholar 

  19. Schuler, T. V. et al. Reconciling Svalbard glacier mass balance. Front. Earth Sci. 8, 156 (2020).

    Article  ADS  Google Scholar 

  20. Bjørk, A. A. et al. An aerial view of 80 years of climate-related glacier fluctuations in southeast Greenland. Nat. Geosci. 5, 427–432 (2012).

    Article  ADS  Google Scholar 

  21. Kjeldsen, K. K. et al. Spatial and temporal distribution of mass loss from the Greenland Ice Sheet since AD 1900. Nature 528, 396–400 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Child, S. F., Stearns, L. A., Girod, L. & Brecher, H. H. Structure-from-motion photogrammetry of Antarctic historical aerial photographs in conjunction with ground control derived from satellite data. Remote Sens. 13, 21 (2021).

    Article  ADS  Google Scholar 

  23. Dyurgerov, M. B. & Meier, M. F. Twentieth century climate change: evidence from small glaciers. Proc. Natl Acad. Sci. USA 97, 1406–1411 (2000).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. WGMS. Fluctuations of Glaciers Database (World Glacier Monitoring Service, accessed 15 February 2020).

  25. Beaulieu, C. & Killick, R. Distinguishing trends and shifts from memory in climate data. J. Clim. 31, 9519–9543 (2018).

    Article  ADS  Google Scholar 

  26. Sasgen, I. et al. Return to rapid ice loss in Greenland and record loss in 2019 detected by the GRACE-FO satellites. Commun. Earth Environ. 1, 8 (2020).

    Article  ADS  Google Scholar 

  27. Velicogna, I., Sutterley, T. & Van Den Broeke, M. Regional acceleration in ice mass loss from Greenland and Antarctica using GRACE time-variable gravity data. Geophys. Res. Lett. 41, 8130–8137 (2014).

    Article  ADS  Google Scholar 

  28. Geyman, E. C., Maloof, A. C. & Dyer, B. How is sea level change encoded in carbonate stratigraphy? Earth Planet. Sci. Lett. 560, 116790 (2021).

    Article  CAS  Google Scholar 

  29. Sullivan, M. J. et al. Long-term thermal sensitivity of Earth’s tropical forests. Science 368, 869–874 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Liu, J., Wennberg, P. O., Parazoo, N. C., Yin, Y. & Frankenberg, C. Observational constraints on the response of high-latitude northern forests to warming. AGU Adv. e2020AV000228 (2020).

  31. Luckman, A. et al. Calving rates at tidewater glaciers vary strongly with ocean temperature. Nat. Commun. 6, 8566 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Koenderink, J. J. & Van Doorn, A. J. Affine structure from motion. J. Opt. Soc. Am. A 8, 377–385 (1991).

    Article  ADS  CAS  PubMed  Google Scholar 

  33. Terrengmodell Svalbard (S0 Terrengmodell) (Norwegian Polar Institute, 2014).

  34. Hay, C. C., Morrow, E., Kopp, R. E. & Mitrovica, J. X. Probabilistic reanalysis of twentieth-century sea-level rise. Nature 517, 481–484 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  35. Oerlemans, J. Glaciers and Climate Change (CRC Press, 2001).

  36. Oerlemans, J. & Fortuin, J. Sensitivity of glaciers and small ice caps to greenhouse warming. Science 258, 115–117 (1992).

    Article  ADS  CAS  PubMed  Google Scholar 

  37. Raper, S. C. & Braithwaite, R. J. Low sea level rise projections from mountain glaciers and ice caps under global warming. Nature 439, 311–313 (2006).

    Article  ADS  CAS  PubMed  Google Scholar 

  38. Nöel, B. et al. Low elevation of Svalbard glaciers drives high mass loss variability. Nat. Commun. 11, 4597 (2020).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  39. Porter, C. et al. ArcticDEM (2018).

  40. Girod, L., Nielsen, N. I., Couderette, F., Nuth, C. & Kääb, A. Precise DEM extraction from Svalbard using 1936 high oblique imagery. Geosci. Instrum. Methods Data Syst. 7, 277–288 (2018).

    Article  ADS  Google Scholar 

  41. Midgley, N. & Tonkin, T. Reconstruction of former glacier surface topography from archive oblique aerial images. Geomorphology 282, 18–26 (2017).

    Article  ADS  Google Scholar 

  42. Mertes, J. R., Gulley, J. D., Benn, D. I., Thompson, S. S. & Nicholson, L. I. Using structure-from-motion to create glacier DEMs and orthoimagery from historical terrestrial and oblique aerial imagery. Earth Surf. Processes Landforms 42, 2350–2364 (2017).

    Article  ADS  Google Scholar 

  43. Kavan, J. Early twentieth century evolution of Ferdinand glacier, Svalbard, based on historic photographs and structure-from-motion technique. Geogr. Ann. A 102, 57–67 (2020).

    Article  Google Scholar 

  44. Holmlund, E. S. Aldegondabreen glacier change since 1910 from structure-from-motion photogrammetry of archived terrestrial and aerial photographs: utility of a historic archive to obtain century-scale Svalbard glacier mass losses. J. Glaciol. 67, 107–116 (2021).

    Article  ADS  Google Scholar 

  45. Loop, C. & Zhang, Z. Computing rectifying homographies for stereo vision. In Proc. 1999 IEEE Computer Society Conference on Computer Vision and Pattern Recognition (Cat. No PR00149) Vol. 1, 125–131 (IEEE, 1999).

  46. Bretscher, O. Linear Algebra with Applications 5th edn (Pearson, 2013).

  47. Westoby, M. J., Brasington, J., Glasser, N. F., Hambrey, M. J. & Reynolds, J. M. ‘Structure-from-Motion’ photogrammetry: a low-cost, effective tool for geoscience applications. Geomorphology 179, 300–314 (2012).

    Article  ADS  Google Scholar 

  48. McNabb, R., Nuth, C., Kääb, A. & Girod, L. Sensitivity of glacier volume change estimation to DEM void interpolation. Cryosphere 13, 895–910 (2019).

    Article  ADS  Google Scholar 

  49. Williams, C. & Rasmussen, C. E. Gaussian Processes for Machine Learning Vol. 2, 302, 303 (MIT Press, 2006).

  50. Nuth, C. & Kääb, A. Co-registration and bias corrections of satellite elevation data sets for quantifying glacier thickness change. Cryosphere 5, 271–290 (2011).

    Article  ADS  Google Scholar 

  51. Kääb, A. Remote Sensing of Mountain Glaciers and Permafrost Creep 48 (Geographisches Institutder Univ., 2005).

  52. James, M. R., Robson, S., d’Oleire Oltmanns, S. & Niethammer, U. Optimising UAV topographic surveys processed with structure-from-motion: ground control quality, quantity and bundle adjustment. Geomorphology 280, 51–66 (2017).

    Article  ADS  Google Scholar 

  53. Rolstad, C., Haug, T. & Denby, B. Spatially integrated geodetic glacier mass balance and its uncertainty based on geostatistical analysis: application to the western Svartisen ice cap, Norway. J. Glaciol. 55, 666–680 (2009).

    Article  ADS  Google Scholar 

  54. Granshaw, F. D. & Fountain, A. G. Glacier change (1958–1998) in the north Cascades national park complex, Washington, USA. J. Glaciol. 52, 251–256 (2006).

    Article  ADS  CAS  Google Scholar 

  55. König, M., Nuth, C., Kohler, J., Moholdt, G. & Pettersen, R. A Digital Glacier Database for Svalbard 229–239 (Springer, 2014).

  56. Nuth, C. et al. Decadal changes from a multi-temporal glacier inventory of Svalbard. Cryosphere 7, 1603–1621 (2013).

    Article  ADS  Google Scholar 

  57. Huss, M. Density assumptions for converting geodetic glacier volume change to mass change. Cryosphere 7, 877–887 (2013).

    Article  ADS  Google Scholar 

  58. van Pelt, W. J. J. et al. Multidecadal climate and seasonal snow conditions in Svalbard. J. Geophys. Res. Earth Surf. 121, 2100–2117 (2016).

    Article  ADS  Google Scholar 

  59. Fürst, J. J. et al. The ice-free topography of Svalbard. Geophys. Res. Lett. 45, 760–769 (2018).

    Article  Google Scholar 

  60. Fürst, J. J. et al. Application of a two-step approach for mapping ice thickness to various glacier types on Svalbard. Cryosphere 11, 2003–2032 (2017).

    Article  ADS  Google Scholar 

  61. Morlighem, M. et al. A mass conservation approach for mapping glacier ice thickness. Geophys. Res. Lett. 38 (2011).

  62. Hagedoorn, J. & Wolf, D. Pleistocene and recent deglaciation in Svalbard: implications for tide-gauge, GPS and VLBI measurements. J. Geodyn. 35, 415–423 (2003).

    Article  Google Scholar 

  63. Sato, T. et al. A geophysical interpretation of the secular displacement and gravity rates observed at Ny-Ålesund, Svalbard in the Arctic–effects of post-glacial rebound and present-day ice melting. Geophys. J. Int. 165, 729–743 (2006).

    Article  ADS  Google Scholar 

  64. Kierulf, H. P., Plag, H. P. & Kohler, J. Surface deformation induced by present-day ice melting in Svalbard. Geophys. J. Int. 179, 1–13 (2009).

    Article  ADS  Google Scholar 

  65. Davis, J. L. & Mitrovica, J. X. Glacial isostatic adjustment and the anomalous tide gauge record of eastern North America. Nature 379, 331–333 (1996).

    Article  ADS  CAS  Google Scholar 

  66. Uppala, S. M. et al. The ERA-40 re-analysis. Q. J. R. Meteorolog. Soc. 131, 2961–3012 (2005).

    Article  ADS  Google Scholar 

  67. Reistad, M. et al. A high-resolution hindcast of wind and waves for the North Sea, the Norwegian Sea, and the Barents Sea J. Geophys. Res. Oceans 116 (2011).

  68. Førland, E. J. et al. Measured and modeled historical precipitation trends for Svalbard. J. Hydrometeorol. 21, 1279–1296 (2020).

    Article  ADS  Google Scholar 

  69. Huybers, P. & Curry, W. Links between annual, Milankovitch and continuum temperature variability. Nature 441, 329–332 (2006).

    Article  ADS  CAS  PubMed  Google Scholar 

  70. Amdur, T., Stine, A. & Huybers, P. Global surface temperature response to 11-yr solar cycle forcing consistent with general circulation model results. J. Clim. 34, 2893–2903 (2021).

    Article  ADS  Google Scholar 

  71. Hays, J. D., Imbrie, J. & Shackleton, N. J. Variations in the Earth’s orbit: pacemaker of the ice ages. Science 194, 1121–1132 (1976).

    Article  ADS  CAS  PubMed  Google Scholar 

  72. Huybers, P. Early Pleistocene glacial cycles and the integrated summer insolation forcing. Science 313, 508–511 (2006).

    Article  ADS  CAS  PubMed  Google Scholar 

  73. Benn, D. I. et al. Melt-under-cutting and buoyancy-driven calving from tidewater glaciers: new insights from discrete element and continuum model simulations. J. Glaciol. 63, 691–702 (2017).

    Article  ADS  Google Scholar 

  74. Cuffey, K. M. & Paterson, W. S. B. The Physics of Glaciers (Academic Press, 2010).

  75. Gabbi, J., Carenzo, M., Pellicciotti, F., Bauder, A. & Funk, M. A comparison of empirical and physically based glacier surface melt models for long-term simulations of glacier response. J. Glaciol. 60, 1140–1154 (2014).

    Article  ADS  Google Scholar 

  76. Fujita, K. & Ageta, Y. Effect of summer accumulation on glacier mass balance on the Tibetan Plateau revealed by mass-balance model. J. Glaciol. 46, 244–252 (2000).

    Article  ADS  Google Scholar 

  77. Naegeli, K. & Huss, M. Sensitivity of mountain glacier mass balance to changes in bare-ice albedo. Ann. Glaciol. 58, 119–129 (2017).

    Article  ADS  Google Scholar 

  78. MacAyeal, D. Binge/purge oscillations of the Laurentide ice sheet as a cause of the North Atlantic’s Heinrich events. Paleoceanography 8, 775–784 (1993).

    Article  ADS  Google Scholar 

  79. Oerlemans, J. Minimal Glacier Models (Utrecht Publishing & Archiving Services, 2008).

  80. Oerlemans, J., Jania, J. & Kolondra, L. Application of a minimal glacier model to Hansbreen, Svalbard. Cryosphere 5, 1–11 (2011).

    Article  ADS  Google Scholar 

  81. Florentine, C., Harper, J., Fagre, D., Moore, J. & Peitzsch, E. Local topography increasingly influences the mass balance of a retreating cirque glacier. Cryosphere 12, 2109–2122 (2018).

    Article  ADS  Google Scholar 

  82. Mott, R. et al. Avalanches and micrometeorology driving mass and energy balance of the lowest perennial ice field of the Alps: a case study. Cryosphere 13, 1247–1265 (2019).

    Article  ADS  Google Scholar 

  83. Kingslake, J. et al. Extensive retreat and re-advance of the West Antarctic Ice Sheet during the Holocene. Nature 558, 430–434 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  84. Etzelmüller, B. et al. Modeling the temperature evolution of Svalbard permafrost during the 20th and 21st century. Cryosphere 5, 67–79 (2011).

    Article  ADS  Google Scholar 

  85. Schuler, T. V. et al. Reconciling Svalbard glacier mass balance. Front. Earth Sci. 8, 156 (2020).

    Article  ADS  Google Scholar 

  86. Błaszczyk, M., Jania, J. & Hagen, J. O. Tidewater glaciers of Svalbard: recent changes and estimates of calving fluxes. Polish Polar Res. 30 (2009).

  87. Huss, M. & Hock, R. A new model for global glacier change and sea-level rise. Front. Earth Sci. 3, 54 (2015).

    Article  ADS  Google Scholar 

  88. Radíc, V. et al. Regional and global projections of twenty-first century glacier mass changes in response to climate scenarios from global climate models. Clim. Dyn. 42, 37–58 (2014).

    Article  Google Scholar 

  89. Marzeion, B., Jarosch, A. & Hofer, M. Past and future sea-level change from the surface mass balance of glaciers. Cryosphere 6, 1295–1322 (2012).

    Article  ADS  Google Scholar 

  90. Jacob, T., Wahr, J., Pfeffer, W. T. & Swenson, S. Recent contributions of glaciers and ice caps to sea level rise. Nature 482, 514–518 (2012).

    Article  ADS  CAS  PubMed  Google Scholar 

  91. Matsuo, K. & Heki, K. Current ice loss in small glacier systems of the Arctic Islands (Iceland, Svalbard, and the Russian High Arctic) from satellite gravimetry. Terr. Atmospheric Ocean. Sci. 24, 657–670 (2013).

    Article  Google Scholar 

  92. Moholdt, G., Nuth, C., Hagen, J. O. & Kohler, J. Recent elevation changes of Svalbard glaciers derived from ICESat laser altimetry. Remote Sens. Environ. 114, 2756–2767 (2010).

    Article  ADS  Google Scholar 

  93. Lang, C., Fettweis, X. & Erpicum, M. Stable climate and surface mass balance in Svalbard over 1979–2013 despite the Arctic warming. Cryosphere 9, 83–101 (2015).

    Article  ADS  Google Scholar 

  94. Oerlemans, J. et al. Estimating the contribution of Arctic glaciers to sea-level change in the next 100 years. Ann. Glaciol. 42, 230–236 (2005).

    Article  ADS  Google Scholar 

  95. Wouters, B., Gardner, A. S. & Moholdt, G. Global glacier mass loss during the GRACE satellite mission (2002-2016). Front. Earth Sci. 7, 96 (2019).

    Article  ADS  Google Scholar 

  96. Box, J. E. et al. Global sea-level contribution from Arctic land ice: 1971–2017. Environ. Res. Lett. 13, 125012 (2018).

    Article  Google Scholar 

  97. Ciracì, E., Velicogna, I. & Swenson, S. Continuity of the mass loss of the world’s glaciers and ice caps from the GRACE and GRACE Follow-On missions. Geophys. Res. Lett. 47, e2019GL086926 (2020).

    Article  ADS  Google Scholar 

  98. Radíc, V. & Hock, R. Regionally differentiated contribution of mountain glaciers and ice caps to future sea-level rise. Nat. Geosci. 4, 91–94 (2011).

    Article  ADS  Google Scholar 

  99. Straneo,F. et al. Impact of fjord dynamics and glacial runoff on the circulation near Helheim Glacier. Nat. Geosci. 4, 322–327 (2011).

    Article  ADS  CAS  Google Scholar 

  100. Bonan, D. B., Christian, J. E. & Christianson, K. Influence of North Atlantic climate variability on glacier mass balance in Norway, Sweden and Svalbard. J. Glaciol. 65, 580–594 (2019).

    Article  ADS  Google Scholar 

  101. Benn, D., Fowler, A. C., Hewitt, I. & Sevestre, H. A general theory of glacier surges. J. Glaciol. 65, 701–716 (2019).

    Article  ADS  Google Scholar 

  102. Dunse, T. et al. Glacier-surge mechanisms promoted by a hydro-thermodynamic feedback to summer melt. Cryosphere 9, 197–215 (2015).

    Article  ADS  Google Scholar 

  103. Morris, A., Moholdt, G. & Gray, L. Spread of Svalbard glacier mass loss to Barents Sea margins revealed by CryoSat-2. J. Geophys. Res. Earth Surf. 125, e2019JF005357 (2020).

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank F. Simons, C.-Y. Lai, P. Wennberg, P. Moore, B. Dyer, G. Moholdt, R.A. Morris, E. Isaksson, A. Schomacker, C. Nuth, E. Schytt Holmlund and B. Geyman for conversations that improved the manuscript. W.J.J.v.P. acknowledges funding from the Swedish National Space Agency (project 189/18). E.C.G. was supported by a Daniel M. Sachs Class of 1960 Global Scholarship at Princeton University, a Svalbard Science Forum Arctic Field Grant, and the Fannie and John Hertz Foundation.

Author information

Authors and Affiliations

Authors

Contributions

E.C.G. and J.K. designed the study. E.C.G. performed the SfM reconstructions and the analysis of the data. W.J.J.v.P. assembled and downscaled the regional climate model results. A.C.M. contributed to the formulation of the space-for-time substitution. H.F.A. oversaw the digitization of the 1936/1938 image archive. E.C.G. wrote the manuscript, with edits from all authors.

Corresponding author

Correspondence to Emily C. Geyman.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review information

Nature thanks Jaime Otero and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Svalbard’s estimated contribution to sea level rise: past and future.

The x-axis represents the total ice loss, expressed in terms of mm of global sea level equivalent (SLE). The SLE is normalized to a 100-yr period to facilitate comparison between historical observations of varying length and projections for the 21st century. For the historical models that only simulate climatic mass balance19, we added a calving flux86 of −6.75 ± 1.7 Gt yr−1 to estimate the total mass balance and facilitate comparison with the geodetic and gravimetric observations. Since the data are compiled from many sources2,3,4,5, 12,13,14,15,16,17, 85, 87,88,89,90,91,92,93,94,95,96,97,98, 103, the error bars represent different quantities. In most cases, the shaded bars represent the reported ± 1σ uncertainty. For Marzeion et al. (2012), the bars represent the range of results from different ensemble runs, and for ‘This study’, the bars represent the range from using the 5th to 95th percentile of future precipitation estimates (Fig. 4). The 21st century projections for this study are broken into 3 groups: (1) those based on the 1936-2010 mass balance observations using mean summer temperature as the explanatory variable for ice loss (†; Figs. 34), (2) those based on the 2000-2019 satellite era observations4 using mean summer temperature (; Extended Data Fig. 10), and (3) those based on the 1936-2010 observations using positive degree days (‡; Extended Data Fig. 9). All three methods produce similar estimates for 21st century ice loss. The predictions begin to diverge under the most extreme warming scenario (RCP8.5)6, with the space-for-time substitution based on the positive degree day (PDD) approach (Extended Data Fig. 9) producing the highest estimates of sea level rise. Note that the total ice volume of Svalbard is estimated59 to be 6,199 km3 which is equivalent to 15 mm of sea level rise. Thus, the 21st century predictions exceeding 15 mm of SLE87,88,89 are not feasible.

Extended Data Fig. 2 Past and future estimates of temperature and precipitation across Svalbard.

(a-c) The 1,888 glaciers in Svalbard span an elevation range38 of >1,200 m, a mean annual temperature range of >10 °C, and a >4-fold change in precipitation (<0.5 to 2.0 m.w.e. yr−1). The elevation map in (a) uses the Norwegian Polar Institute S0 terrain model33, and the mean annual air temperature (b) and mean annual precipitation (c) estimates use the downscaled NORA10 dataset5 (1 km resolution, 1957-2018). (d-i) Glacier-averaged summer temperature (d-f) and solid precipitation estimates (g-i) for 2010-2100 from Arctic CORDEX under the RCP2.6, RCP4.5, and RCP8.5 scenarios6. The color bar for the mean summer temperature (Ts) maps in (d-f) is centered on the upper bound of the 1936-2010 Ts (95th percentile = 2.2 °C; Fig. 3). Thus, brown colors indicate that, in the space-for-time substitution in Fig. 4, we are predicting glacier behavior in response to temperatures that rise higher than those in the observational data that calibrate the model.

Extended Data Fig. 3 An overview of the 1936/1938 aerial survey7 and the new datasets available from this study.

(a) Locations of the 5,507 high-oblique aerial photographs acquired during the mapping campaigns7 of 1936 and 1938. The coordinates are displayed in WGS84 / UTM zone 33N. (b) We divided the images into 17 groups with convergent camera geometries for SfM reconstruction in Agisoft Metashape (Extended Data Fig. 4). We provide the Svalbard-wide (c) 1936 orthophotomosaic, (d) 1936 digital elevation model (DEM), and (e) 1936-2010 elevation change (∆h) at 20 m and 50 m resolution. We also provide the 1936-2010 ∆h datasets at 5 m resolution, although, because of file size constraints, these data are divided into 8 files for each of the Svalbard zones depicted in Fig. 2d. The unprocessed point clouds from the 17 regional SfM reconstructions in (c) are available as .laz files. In addition to the raster datasets in (c-e), we provide a .shp file inventorying the 1936 extents of Svalbard glaciers and a spreadsheet recording statistics such as ∆h/∆t, ∆M/∆t, DEM uncertainty, and NORA10 climate fields5 (mean summer temperature, precipitation as snow, etc.). See Data availability.

Extended Data Fig. 4 An overview of the image pre-processing and structure from motion (SfM) pipelines.

To improve feature selection during the SfM reconstruction, we enhance the digitized images by increasing contrast through histogram stretching and sharpening features using the Dehaze Tool in Adobe Lightroom. This radiometric enhancement step improves photogrammetric reconstructions over ice and snow, which tend to be lower contrast than the surrounding land. Finally, since scanning does not preserve the internal geometry (images can be rotated, translated, and warped), we locate the four fiducial marks on the edges of each image and apply a projective transformation that maps the images to a standardized internal geometry. Owing to the large number of images in the dataset, we use an automated pipeline for fiducial mark identification. We convolve the edges of the image with an idealized fiducial template to identify target regions. Next, inside the target regions, we convolve the image with a Laplacian of Gaussian filter to locate the fiducial spot. We process the aerial photographs in a standard photogrammetric workflow in Agisoft Metashape 1.6.0. In brief, we first extract up to 40,000 keypoints from each image. Keypoint matching across all the images provides the constraints to solve for the unknown parameters, including the relative camera locations/orientations and the camera distortion parameters. Adding ground control points (GCPs), with specified (x,y,z) positions, enables the absolute georeferencing of the model. Finally, a multi- view stereo (MVS) reconstruction converts the sparse 3D model to a dense 3D point cloud. We perform the MVS reconstruction with a Dense Quality of Medium (meaning that depth maps are generated at 1/4 the image resolution) and Dense Filtering at Moderate to Aggressive.

Extended Data Fig. 5 Example 3D comparisons of Svalbard glaciers in 1936 and 2008-2011.

(a) Austre/Vestre Brøggerbreen, (b) Midtre/Austre Lovénbreen, (c) Grønfjordbreen, (d) Tungebreen, (e) Gløttfjellbreen, (f) Pedâsenkobreen. Glacier volume decreased by 11% across Svalbard during the interval 1936-2010 (Fig. 2). The 2008-2011 models use the NPI 5 m regional DEMs33 and associated 50 cm orthophotos (https://geodata.npolar.no/).

Extended Data Fig. 6 Data coverage and void-filling with Gaussian process (GP) regression.

(a) Black areas denote regions with 3D photogrammetric (SfM) constraints from the 1936/1938 aerial images (Fig. 1) and white regions denote void areas infilled with the GP regression (Methods). The SfM-generated point clouds have void areas because of occlusion and poor feature matching in low-contrast areas. There is no reconstruction for the eastern portion of Austfonna (Fig. 2), since no photographs of that region were acquired during the 1936/1938 expeditions7 (Extended Data Fig. 3a). (b-e) An illustration of the void filling procedure, applied to Oscar II Land in western Svalbard. To fill the holes in the 1936 DEM, we first compute the ∆h map, differencing the 2010 reference DEM to the 1936 reconstruction (b). Next, we train a GP regression to estimate the ∆h values in the void areas. The GP regression is trained using x, y, and z (the 2010 elevation) as predictor variables to infer ∆h as the response variable, and thereby incorporates both the spatial information in (a) and the elevation-dependence of ∆h illustrated in (c). (d) The error of the GP-regression-infilled values is estimated on random subsets of data points (60%) held-out from model training. Finally, subtracting the infilled ∆h map in (e) from the 2010 reference DEM yields the 1936 surface (Extended Data Fig. 3d).

Extended Data Fig. 7 Climatic and geometric controls on glacier mass balance.

(a-b) Temperature control on ice loss. The scatter plots in (a-b) show a similar relationship as Main text, Fig. 3f-g (Ts vs. ∆h/∆t), except the y-axis in (a-b) also includes the solid precipitation component. Specifically, Fout = Psolid − b, where Psolid is extracted from the downscaled NORA10 dataset5, and b is ∆h/∆t converted to m.w.e. yr−1 using a density57 of 850 kg m−3. Since the glacier-specific precipitation estimates are noisy, the plots in (a-b) show considerably more scatter than those in Main text, Fig. 3f-g. The advantage of examining the data in terms of Fout is that it enables us to extract the physical quantity k1, which describes the expected increase in ice loss (m.w.e. yr−1) for each 1 °C rise in mean summer temperature. The gray bands in (a-b) depict the 25th–75th percentile uncertainty envelopes of the Ts vs. Fout regressions (all glaciers). Note that the ice loss in marine-terminating glaciers is regulated not only by air temperature driving Fmelt, but also by fjord temperature, bathymetry, and circulation controlling Fcalving31,73,99. We take a first-order approach and fit different k constants to land- vs. marine-terminating glaciers. In both (a-b), the k1 coefficients are larger in land-terminating glaciers than marine-terminating glaciers. This observation of a weaker sensitivity of Fout to air temperature is consistent with the Fout in marine-terminating glaciers being driven partly by fjord processes that are decoupled from air temperature. Since the satellite-era observations (b) represent a shorter interval and therefore the ∆h/∆t data are more influenced by annual variability100 and surge cycles101, we only fit glaciers with ∆h/∆t within the 10th–90th percentile range. Glaciers outside this range are depicted with gray dots. Note that the estimated k1 coefficients from the 1936-2010 observations (a) and satellite-era datasets4 (b) are within uncertainty of each other. (c-d) Glacier slope modulates sensitivity to warming. (a) Simple theoretical glacier models79 predict that glaciers with steeper slopes should be less sensitive to temperature rise. The rationale is that, for a lower-slope glacier, a given ELA rise of x meters will transfer a larger fraction of the glacier’s area from the accumulation zone to the ablation zone38, causing a more substantial decrease in glacier-averaged ∆h/∆t. (b) We test whether there is evidence for a bed-slope control on glacier sensitivity to temperature rise79 in our 1936-2010 dataset (Fig. 2) by estimating the relationship between Ts and ∆h/∆t for low, medium, and high slope glaciers. The bed slope is calculated as ∆z/∆x of the bed topography59 along each glacier’s centerline56. The distributions in (b) represent the regression slopes derived from weighted total least squares regressions on repeated random 50% subsamples of the dataset.

Extended Data Fig. 8 Visual and quantitative comparison of the regional Ts vs. ∆h/∆t behavior of Svalbard glaciers.

(a-b) A visual comparison between glaciers in (a) NE Spitsbergen and (b) Edgeøya near the end of the 2020 melt season. Note that, in contrast to the glaciers in (a), the glaciers in (b) have no remaining winter snow, as evinced by the darker, debris-rich ice exposed even at the highest reaches of the glaciers. In other words, 100% of the glacier surface lies within the ablation zone. The Sentinel-2 imagery is from August 01, 2020 and the coordinates are in UTM zone 33N. (c-d) Glacier extents in 1936 and 2010. (e-i) Searching for evidence of threshold behavior in the Svalbard-wide 1936-2010 dataset. (e-f) If a strong tipping point already had been reached, such that, at high Ts, glaciers diverged from the linear behavior in Fig. 3, one might expect a Ts vs. ∆h/∆t relationship like that depicted in (f). However, (e) does not show strong support for the model in (f). Next, we look for a regional signal in the Ts vs. ∆h/∆t relationship. We divide the Svalbard-wide dataset into the 8 regions shown in (h), each of which has a different average Ts (g) and ∆h/∆t. (i) For each region, we study the residual between the Svalbard-wide linear Ts vs. ∆h/∆t (e) and the regional observations. The residuals are computed as predicted minus observed, so negative values indicate that the observed mass balance is more negative than the predictions. The regions in (i) are ordered according to the region-averaged glacier bed slope (Extended Data Fig. 7c), from smallest to largest. The glaciers like those in (b) that are committed to a path of pure melting (no accumulation) appear to follow similar Ts vs. ∆h/∆t relationships as the healthier (close to balance) glaciers in (a).

Extended Data Fig. 9 Glacier sensitivity to warming  and 21st century predictions based on positive degree day (PDD) estimates.

(a-g) An analogous figure to Fig. 3, except using positive degree days (PDDs) to model the ice loss flux (eqns. 1213) rather than mean summer temperature. In (g), only bins representing n ≥ 20 glaciers are shown. (h-k) PDDs are a non-linear function of mean summer temperature. For each glacier, we use the 1957-present daily temperature time series5 (h) to understand the relationship between mean summer temperature and PDDs (k). Specifically, we iteratively shift the time series in (h) up at 0.5 °C steps and compute the new number of PDDs. The red dashed line in (k) depicts what the PDD estimate would be if the melt season didn’t get any longer, which is analogous to what the Ts model does (Figs. 34). As illustrated in (j), the number of PDDs increases not only because the mean summer temperature rises, but also because the duration of the melt season increases. The plots in (h-k) are for the glacier Bungebreen (76.814N, 16.097E). We repeat the analysis for all glaciers on Svalbard to produce glacier-specific relationships between mean summer temperature and PDDs (k). (l) We test the PDD-based space-for-time substitution using 1936-1990 data to calibrate the model, and then compare predictions for the 1990-2010 interval to independent DEM-derived mass balance estimates. We find excellent agreement between the model and observations. Next, we use the model to predict 21st century (2010-2100) mass balance. (m-n) The predictions using the PDD method are nearly identical to the predictions using mean summer temperature (Fig. 4) under the RCP2.6 (m) and RCP4.5 (n) scenarios. However, the PDD method produces more negative mass balance estimates under RCP8.5, because of the divergence of the PDD curve from a simple linear function at higher summer temperatures (k). The brackets in (m-o) show results from simulations using the 5th and 95th percentiles of predicted winter precipitation6.

Extended Data Fig. 10 Temperature control on glacier mass balance in the satellite era4 (2000-2019).

An analogous figure to Figs. 34, but using 2000-2019 ∆h/∆t derived from ASTER DEMs4. The mean summer temperature values are extracted from the downscaled NORA10 dataset5. (c) The very negative ∆h/∆t at Basin-3 on Austfonna represents an ongoing surge there with an estimated calving flux102 of 4.2 ± 1.6 Gt yr−1, which represents about a quarter of the Svalbard-wide ice loss103. (e-g) To reduce the influence of outliers in the regression analysis, we only include the glaciers that have mean summer temperature and ∆h/∆t values that fall within the 10th to 90th percentiles of the Svalbard-wide dataset. (e) As in Fig. 3e, the distribution of regression slopes from the bootstrap resampling scheme does not overlap 0 m yr−1 °C−1, indicating a significant temperature dependence of ∆h/∆t. The estimated regression slope is −0.28 [−0.36, −0.22] m yr−1 °C−1 (median with 25-75th percentile range), which is slightly less negative, but still within the uncertainty envelope of the value estimated in Fig. 3 (−0.37 [−0.43, −0.29] m yr−1 °C−1). (h) To test whether the 2000-2019 observations4 are sufficiently long to characterize the temperature and precipitation control on ice loss (Extended Data Fig. 7), we train a space-for-time substitution using the 2000-2019 observations and the NORA10 temperature and precipitation estimates5 for 1936-1990 to estimate mass balance during the period 1936-1990. Comparison with the independent, DEM-derived observations of ∆h/∆t indicates that the space-for-time estimates have an error of 0.05 m yr−1. The simulated 1936-1990 mass balance in (h) may have a larger spread than the observed mass balance due to the somewhat more noisy (affected by surges, interannual variability100, etc.) satellite-era data used to calibrate the model. (i-k) We use a space-for-time substitution, trained using the 2000-2019 observations4 to predict 21st century Svalbard-wide ∆h/∆t under the same three climate scenarios6 as in Fig. 4b-d. The brackets show the model runs with the 5th and 95th percentiles of modeled winter precipitation6.

Extended Data Table 1 Svalbard glacier changes, 1936-2010

Supplementary information

Supplementary Information

Additional clarification and details about the methodology. The code and the associated datasets required to reproduce the space-for-time analysis can be downloaded from https://doi.org/10.5281/zenodo.5643856. The full dataset of 1936/1938 3D glacier reconstructions can be downloaded from https://doi.org/10.5281/zenodo.5644415.

Peer Review Information

Supplementary Data

Glacier-by-glacier statistics such as ice loss, climate parameters and twenty-first century predictions.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Geyman, E.C., J. J. van Pelt, W., Maloof, A.C. et al. Historical glacier change on Svalbard predicts doubling of mass loss by 2100. Nature 601, 374–379 (2022). https://doi.org/10.1038/s41586-021-04314-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-04314-4

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing