Elsevier

Earth and Planetary Science Letters

Volume 452, 15 October 2016, Pages 206-215
Earth and Planetary Science Letters

Crustal deformation across the Southern Patagonian Icefield observed by GNSS

https://doi.org/10.1016/j.epsl.2016.07.042Get rights and content

Highlights

  • Crustal deformation at the Southern Patagonian Icefield observed at 43 GNSS sites.

  • Glacial-isostatic adjustment generates crustal uplift of up to 4 cm per year.

  • Plate collision and the Patagonian slab window contribute also to horizontal strain.

Abstract

Geodetic GNSS observations at 43 sites well distributed over the Southern Patagonian Icefield region yield site velocities with a mean accuracy of 1 mm/a and 6 mm/a for the horizontal and vertical components, respectively. These velocities are analyzed to reveal the magnitudes and patterns of vertical and horizontal present-day crustal deformation as well as their primary driving processes. The observed vertical velocities confirm a rapid uplift, with rates peaking at 41 mm/a, causally related to glacial-isostatic adjustment (GIA). They yield now an unambiguous preference between two competing GIA models. Remaining discrepancies between the preferred model and our observations point toward an effective upper mantle viscosity even lower than 1.61018 Pas and effects of lateral rheological heterogeneities. An analysis of the horizontal strain and strain-rate fields reveals some complex superposition, with compression dominating in the west and extension in the east. This deformation field suggests significant contributions from three processes: GIA, a western interseismic tectonic deformation field related to plate subduction, and an extensional strain-rate field related to active Patagonian slab window tectonics.

Introduction

The Southern Patagonian Icefield (SPI) has been identified as a locus of exceptionally rapid crustal uplift (Dietrich et al., 2010, Lange et al., 2014). Models show that observed vertical deformation in this region can be explained by glacial-isostatic adjustment (GIA) (Ivins and James, 1999, Ivins and James, 2004; Klemann et al., 2007). Vertical site velocities observed in Patagonia exceed those reported for other regions affected by glacier retreat such as Alaska (Larsen et al., 2005), Greenland (Khan et al., 2007, Dietrich et al., 2005) and West Antarctica (Groh et al., 2012).

The large-amplitude GIA response in Patagonia results from the coincidence of a large, rapidly changing temperate ice mass and a unique tectonic setting. The Patagonian icefields, representing the largest extra-polar ice mass in the southern hemisphere, have experienced substantial fluctuations in glacier extent and mass throughout the Pleistocene (e.g. Strelin et al., 2011, Strelin et al., 1999; Mercer, 1976) and Holocene (e.g. Strelin et al., 2014, Aniya, 2013, Masiokas et al., 2009) until present (e.g. Floricioiu et al., 2012, Willis et al., 2012, Glasser et al., 2011, Casassa et al., 2002). GIA models suggest, however, that the uplift observed today in Patagonia is caused by ice-mass changes since the Little Ice Age (LIA) with its maximum extent between AD 1630 and 1870 (Ivins and James, 2004, Lange et al., 2014). The uplift predicted from the elastic crustal response to ongoing ice loss shows a more localized pattern than the complete gravitational visco-elastic rebound, but may reach about half the total uplift rate close to fast-retreating glaciers like Upsala (Lange et al., 2014, Fig. 1c).

The short-lived memory of the solid earth for load changes in southern Patagonia, compared to other regions affected by GIA, is due to the peculiar regional rheology characterized by a thin lithosphere and very low viscosities in the asthenosphere and upper mantle (Lange et al., 2014). These rheological conditions, in turn, are imposed by the geologically young thermomechanics of the region. At the Chile Triple Junction, some 200 km NNW from the SPI, material of the relatively hot Chile Ridge is subducted beneath the South American plate (Fig. 1a). The subduction of this active spreading-ridge system between Nazca and Antarctic plates is accompanied by the opening of the Patagonian slab window. The existence and extent of this slab window has been established by geochemical analysis of Patagonian lavas and basalts (Gorring and Kay, 2001, Boutonnet et al., 2010), seismic imaging (Russo et al., 2010a, Gallego et al., 2010), shear wave splitting measurements (Murdie and Russo, 1999, Russo et al., 2010b) and plate kinematic modeling (Breitsprecher and Thorkelson, 2009). The slab window environment beneath the southernmost South American continent is characterized by upwelling of hot mantle material and enhanced mantle flow through the window with a strong ridge-parallel component (Russo et al., 2010b).

From structural-geological point of view, most of our sites pertain to the Southern Patagonian Andes fold-thrust belt. Most authors agree that fold deformation and activity of minor faults in our study area are of late-Miocene age or older (e.g. Giacosa et al., 2012; Fosdick et al., 2011, Ghiglione et al., 2009, Kraemer, 1998). Coutand et al. (1999) suspect that basement faults along the cordillera in the Lago Viedma area “appear to be active in strike-slip reverse mode”. Currently, evidence for active faulting in the SPI region is difficult to find. International seismological catalogues (ISC, 2015, USGS, 2015, Villaseñor et al., 1997, Sabbione et al., 2007) report also only a small number and low magnitude of seismic events for this area. There are active volcanoes along the crest of the Southern Patagonian Andes. However, the deformation associated with volcanic activity is usually restricted to the vicinity of the volcano edifice, and none of our sites are located close enough to be affected by volcanic deformation. We assume, therefore, that our observed site velocities essentially represent the long-wavelength deformation field.

Both plate collision and slab window opening are expected in addition to GIA to produce large-scale surface deformations affecting the SPI region. Tectonic deformations are reflected primarily in the horizontal components (e.g., Elliott et al., 2010). However, previous works on crustal deformation in southern Patagonia focused on the vertical component (Dietrich et al., 2010, Lange et al., 2014). Global compilations of the horizontal deformation field, in turn, expose Patagonia as a region lacking geodetic observations (Kreemer et al., 2014).

This paper presents results of GNSS observations in a regional network of 43 sites. Compared to previous work, the network extension and geometry are improved and allow, for the first time, a detailed analysis of the horizontal velocity components. The observed vertical and horizontal site velocities are interpreted with regard to the magnitude, patterns and driving processes of present-day crustal deformation across the SPI.

Section snippets

Observations

The first geodetic GNSS observations in the area of the SPI were carried out as early as 1996 (Lange et al., 2014). Until 2001, seven sites were observed, mainly on top of the hard-to-reach icefield on Chilean territory. In 2003, a network of eight sites was set up in the area of Lago O'Higgins (Chile; “S” in Fig. 1b) in the northern part of the SPI for a systematic observation of GIA (Dietrich et al., 2010). This was complemented by a network of eleven sites set up in 2009 across the southern

GNSS data processing

The entire set of GNSS data was processed using the Bernese GNSS Software 5.1 (Dach et al., 2007). Products derived from a joint reprocessing of global GPS, GLONASS and SLR data (Fritsche et al., 2014), including IGS station coordinates, satellite orbits and earth orientation parameters, are introduced and provide state-of-the-art consistency and stability. Absolute phase-center corrections are applied for both satellite and receiver antennas. The Vienna mapping function (Kouba, 2008) and

Vertical velocities

The vertical velocities observed at our GNSS sites, together with their estimated uncertainties, are included in Table 1 and displayed in Fig. 3a. All the observed vertical velocities are positive, indicating uplift with respect to the global reference frame. They range from 2.3 mm/a at our southernmost site 31 to 41.0 mm/a at site 10. The site-specific uncertainties of the vertical velocities range from 3.9 to 11.4 mm/a. The consistency found in general between neighboring sites suggests that

Vertical velocities

Fig. 3a shows the observed vertical velocities together with the vertical deformation predicted by the GIA model A in Lange et al. (2014). The general agreement between model and observations, especially regarding the concentrical, dome-shaped uplift pattern, demonstrates that the visco-elastic response to regional ice-mass changes since the LIA is the primary process reflected in the vertical velocities. This model results from a fit, via the adjustment of the earth model parameters, to a

Conclusions

Our GNSS network extents for the first time into Argentina and completes now a circumferential coverage around the SPI. The observed vertical velocities confirm a rapid uplift with a maximum rate slightly exceeding 4 cm/a and GIA as the primary cause (Lange et al., 2014, Dietrich et al., 2010). But our updated and extended data set yields now an unambiguous preference for one of the two competing GIA models presented by Lange et al. (2014). And the improved network geometry allows now also the

Acknowledgments

The German part of the project was funded by the German Research Foundation DFG (grants RI 2340/1-1, DI 473/40-1). E. Ivins was funded at the Jet Propulsion Laboratory, California Institute of Technology, by the Cryosphere Program, the Earth Surface and Interior Focus Area and as part of both the GRACE Science and NASA Sea-level Change Teams. We thank Gerardo Connon, Luis Barbero, Anja Wendt, Andrés Rivera, Rodrigo Traub, Marcelo Arévalo and Hans Silva for their valuable help in the field. We

References (52)

  • J.A. Strelin et al.

    New evidences concerning the Plio-Pleistocene landscape evolution of southern Santa Cruz region

    J. South Am. Earth Sci.

    (1999)
  • J.A. Strelin et al.

    Radiocarbon chronology of the late-glacial Puerto Bandera moraines, Southern Patagonian Icefield, Argentina

    Quat. Sci. Rev.

    (2011)
  • J.A. Strelin et al.

    Holocene glacier history of the Lago Argentino basin, Southern Patagonian Icefield

    Quat. Sci. Rev.

    (2014)
  • M. Aniya

    Holocene glaciations of Hielo Patagonico (Patagonia Icefield), South America. A brief review

    Geochem. J.

    (2013)
  • M. Bevis et al.

    Trajectory models and reference frames for crustal motion geodesy

    J. Geod.

    (2014)
  • P. Bird

    An updated digital model of plate boundaries

    Geochem. Geophys. Geosyst.

    (2003)
  • G. Blewitt et al.

    MIDAS robust trend estimator for accurate GPS station velocities without step detection

    J. Geophys. Res., Solid Earth

    (2016)
  • G. Casassa et al.

    Current knowledge of the Southern Patagonia Icefield

  • R. Dietrich et al.

    Present-day vertical crustal deformations in West Greenland from repeated GPS observations

    Geophys. J. Int.

    (2005)
  • J.L. Elliott et al.

    Tectonic block motion and glacial isostatic adjustment in southeast Alaska and adjacent Canada constrained by GPS measurements

    J. Geophys. Res.

    (2010)
  • Floricioiu, D., Abdel, J.W., Rott, H., 2012. Surface elevation changes and velocities on the Southern Patagonia...
  • J.C. Fosdick et al.

    Kinematic evolution of the Patagonian retroarc fold-and-thrust belt and Magallanes foreland basin, Chile and Argentina, 51°30′S

    Geol. Soc. Am. Bull.

    (2011)
  • M. Fritsche et al.

    Impact of higher-order ionospheric terms on GPS estimates

    Geophys. Res. Lett.

    (2005)
  • M. Fritsche et al.

    Homogeneous reprocessing of GPS, GLONASS and SLR observations

    J. Geod.

    (2014)
  • A. Gallego et al.

    Seismic noise tomography in the Chile ridge subduction region

    Geophys. J. Int.

    (2010)
  • Cited by (45)

    • Glacial isostatic adjustment near the center of the former Patagonian Ice Sheet (48°S) during the last 16.5 kyr

      2022, Quaternary Science Reviews
      Citation Excerpt :

      Observations of post-Last Glacial Maximum (LGM) glacial isostatic adjustment across Patagonia are scarce. The available reconstructions (Bourgois et al., 2016; Kilian et al., 2007; Ríos et al., 2020) and GPS measurements (Richter et al., 2016) suggest that postglacial isostatic rebound might have fluctuated considerably in both space and time. East of the Northern Patagonian Icefield, within the Lake General Carrera basin (Fig. 1), Bourgois et al. (2016) reconstructed glacial isostatic adjustment rates of 1.5 – 3.4 cm/yr for the past 7.9 kyr.

    • Perito Moreno Glacier dam rupture - A recurrent natural experiment to probe solid-earth elasticity

      2020, Journal of South American Earth Sciences
      Citation Excerpt :

      In this case, observations of loading effects would provide a means to derive parameter values for such a model, valid for the rheologically complex SPI region. This would help to disentangle the elastic and visco-elastic contributions to observed, ice-mass induced uplift rates (Richter et al., 2016a) and to explain the 1 h lag of the ocean-tidal loading contribution to the lake tides observed in Lago Argentino and Lago Viedma (Richter et al., 2016b). This proposed further observation and analysis calls for a sub-daily resolution of both the BR + BS water level and the vertical deformation derived from GNSS observations during and following the outburst.

    • A recent increase in megathrust locking in the southernmost rupture area of the giant 1960 Chile earthquake

      2020, Earth and Planetary Science Letters
      Citation Excerpt :

      If hydrological loading in this area (e.g., Rodell et al., 2018) is responsible for the faster vertical motion, it must be operating at a timescale long enough to affect both measurement periods similarly and cannot be the main cause for the rapid velocity change between the two periods which is predominantly horizontal. The effects of climatically induced changes in ice loading are more important south of our study region (Dietrich et al., 2010; Richter et al., 2016) and deserve more detailed research. We thus must conclude that the sudden increase in eastward velocity in the early 21st century, both at COYQ and at a regional scale, is a true tectonic signal.

    View all citing articles on Scopus
    View full text