Skip to main content

Main menu

  • Home
    • Journal home
    • Lyell Collection home
    • Geological Society home
  • Content
    • Online First
    • Issue in progress
    • All issues
    • Thematic Collections
    • Supplementary publications
    • Open Access
  • Subscribe
    • GSL fellows
    • Institutions
    • Corporate
    • Other member types
  • Info
    • Authors
    • Librarians
    • Readers
    • GSL Fellows access
    • Other member type access
    • Press office
    • Accessibility
    • Help
    • Metrics
  • Alert sign up
    • eTOC alerts
    • Online First alerts
    • RSS feeds
    • Newsletters
    • GSL blog
  • Submit
  • Geological Society of London Publications
    • Engineering Geology Special Publications
    • Geochemistry: Exploration, Environment, Analysis
    • Journal of Micropalaeontology
    • Journal of the Geological Society
    • Lyell Collection home
    • Memoirs
    • Petroleum Geology Conference Series
    • Petroleum Geoscience
    • Proceedings of the Yorkshire Geological Society
    • Quarterly Journal of Engineering Geology and Hydrogeology
    • Quarterly Journal of the Geological Society
    • Scottish Journal of Geology
    • Special Publications
    • Transactions of the Edinburgh Geological Society
    • Transactions of the Geological Society of Glasgow
    • Transactions of the Geological Society of London

User menu

  • My alerts
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
Journal of the Geological Society
  • Geological Society of London Publications
    • Engineering Geology Special Publications
    • Geochemistry: Exploration, Environment, Analysis
    • Journal of Micropalaeontology
    • Journal of the Geological Society
    • Lyell Collection home
    • Memoirs
    • Petroleum Geology Conference Series
    • Petroleum Geoscience
    • Proceedings of the Yorkshire Geological Society
    • Quarterly Journal of Engineering Geology and Hydrogeology
    • Quarterly Journal of the Geological Society
    • Scottish Journal of Geology
    • Special Publications
    • Transactions of the Edinburgh Geological Society
    • Transactions of the Geological Society of Glasgow
    • Transactions of the Geological Society of London
  • My alerts
  • Log in
  • Log out
  • My Cart
  • Follow gsl on Twitter
  • Visit gsl on Facebook
  • Visit gsl on Youtube
  • Visit gsl on Linkedin
Journal of the Geological Society

Advanced search

  • Home
    • Journal home
    • Lyell Collection home
    • Geological Society home
  • Content
    • Online First
    • Issue in progress
    • All issues
    • Thematic Collections
    • Supplementary publications
    • Open Access
  • Subscribe
    • GSL fellows
    • Institutions
    • Corporate
    • Other member types
  • Info
    • Authors
    • Librarians
    • Readers
    • GSL Fellows access
    • Other member type access
    • Press office
    • Accessibility
    • Help
    • Metrics
  • Alert sign up
    • eTOC alerts
    • Online First alerts
    • RSS feeds
    • Newsletters
    • GSL blog
  • Submit

Mountains of southernmost Norway: uplifted Miocene peneplains and re-exposed Mesozoic surfaces

View ORCID ProfilePeter Japsen, Paul F. Green, View ORCID ProfileJames A. Chalmers and View ORCID ProfileJohan M. Bonow
Journal of the Geological Society, 175, 721-741, 12 July 2018, https://doi.org/10.1144/jgs2017-157
Peter Japsen
1Geological Survey of Denmark and Greenland (GEUS), Øster Voldgade 10, DK-1350 Copenhagen, Denmark
  • Find this author on Google Scholar
  • Search for this author on this site
  • ORCID record for Peter Japsen
  • For correspondence: pj@geus.dk
Paul F. Green
2Geotrack International, 37 Melville Road, Brunswick West, VIC 3055, Australia
  • Find this author on Google Scholar
  • Search for this author on this site
James A. Chalmers
1Geological Survey of Denmark and Greenland (GEUS), Øster Voldgade 10, DK-1350 Copenhagen, Denmark
  • Find this author on Google Scholar
  • Search for this author on this site
  • ORCID record for James A. Chalmers
Johan M. Bonow
3Geovisiona AB, Högbyvägen 168, SE-17554 Järfälla, Sweden
4Uppsala University, Box 513, SE-75120 Uppsala, Sweden
  • Find this author on Google Scholar
  • Search for this author on this site
  • ORCID record for Johan M. Bonow
PreviousNext
  • Article
  • Figures & Data
  • Info & Metrics
  • PDF
Loading

Abstract

The origin of the Norwegian mountains (the Scandes) is a key controversy in modern geoscience. Are they remnants from the Caledonian Orogeny, modified shoulders of late Mesozoic rifts, or are they evidence of Neogene uplifts? Our synthesis of geological data, landscape analysis and new thermochronological data from Norway south of c. 60°N, combined with previously published data from southern Sweden, reveals a four-stage history: (1) Middle Triassic and Middle Jurassic exhumation produced a weathered basement surface with a hilly relief; (2) after late Mesozoic rifting, Upper Jurassic–Oligocene sediments accumulated across most of the area; (3) early Miocene uplift and erosion to the base level of the adjacent ocean led to formation of a peneplain that extended across sedimentary basins and Caledonian rocks; the subhorizontal Hardangervidda plateau represents this peneplain; (4) early Pliocene uplift raised Hardangervidda to its present elevation of c. 1200 m above sea-level and led to re-exposure of the tilted, Mesozoic surface at lower elevations. The Southern Scandes are thus, like other elevated passive continental margins around the world, the product of post-breakup uplift. Identification of the mechanisms driving these uplifts awaits geodynamic modelling constrained by observations such as those presented in this study.

Supplementary material: Table S1: AFTA data, sample details and associated thermal history interpretations; southernmost Norway. S2: analytical details and thermal history reconstructions for all apatite fission-track analysis (AFTA) samples as well as analytical data for one vitrinite reflectance sample, together with background information on the AFTA technique, are available at https://doi.org/10.6084/m9.figshare.c.4085996

Why are there mountains in Norway? And why are there mountains along many other passive continental margins around the world; for example, along the Atlantic margin of Brazil, in East and West Greenland and in SE Australia? These questions have been debated intensely in recent years, but answers cannot be given without knowing the time at which the elevated topography formed. In the Norwegian case, the rocks that form the mountains are more than 400 myr old and thus do not provide any direct evidence of the geological development over that time span.

This huge gap in the geological record of the Norwegian mountains (also known as the Northern and Southern Scandes; Fig. 1a), together with similar time gaps for other elevated passive continental margins (EPCMs) around the world, has led to a variety of estimates of the age of the mountains revealing disagreements of a spectacular magnitude (Japsen et al. 2012a; Green et al. 2013). Are the Scandes remnants from the mid-Silurian–early Devonian Caledonian Orogeny (Nielsen et al. 2009; Pedersen et al. 2016, 2018)? Are they modifications of rift shoulders formed during late Mesozoic rifting or Paleogene break-up in the NE Atlantic (Redfield & Osmundsen 2013)? Or are they the result of Neogene uplifts (Reusch 1901; Riis 1996; Lidmar-Bergström et al. 2000)? These conflicting views illustrate fundamental disagreements, not only about first-order geological observations of the margin of NW Europe but also about the geodynamic processes that shape the surface of the Earth.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

(a) Regional geology. Location of AFTA samples in this study and in southern Sweden (Japsen et al. 2016). (Based on Gee and Sturt, 1985; Sigmond 1993; Lidmar-Bergström et al. 1999; Knudsen & Fossen 2001; Japsen et al. 2007, 2016; Bøe et al. 2010; Fredin et al. 2017.) (b) Elevation of southernmost Norway. (c) Location of samples and landscape elements of the Southern Scandes; in particular the subhorizontal Hardangervidda. The flanks of the Southern Scandes are characterized by undulating hilly relief and fracture valleys (Fig. 4b; Green et al. 2013; Lidmar-Bergström et al. 2013). Both types of relief are generated by formation of kaolinitic saprolites by deep weathering in bedrock fractures in subtropical climates during the Mesozoic (Lidmar-Bergström 1995). Deep weathering of assumed Mesozoic age in the Oslo region was reported by Baranwal et al. (2016). DB, Danish Basin; FF, Folgefonni Glacier; OR, Oslo Rift; NSB, North Sea Basin; NS and SS, Northern and Southern Scandes; SD, Setesdalen; SG, Skagerrak Graben; SSD, South Swedish Dome; STZ, Sorgenfrei–Tornquist Zone.

Here we present new apatite fission-track analysis (AFTA) data from Norway south of c. 60°N that shed light on the history of burial, uplift and erosion episodes that affected the region (Fig. 1). We do so by combining AFTA data from 28 samples with observations from the geological record onshore and offshore Norway and with results from stratigraphic landscape analysis (Green et al. 2013; Lidmar-Bergström et al. 2013), similar to our approach in a study of southern Sweden (Japsen et al. 2016). These results allow us to present a new model for the formation of the Southern Scandes.

There has been considerable discussion in the literature concerning the use of the word ‘uplift’. We use the term to describe vertical movement of rock against gravity. If erosion has resulted from that movement, we refer to ‘uplift and erosion’ or to ‘exhumation’. If a landmass has moved upwards, we may refer to ‘surface uplift’, but in most circumstances we feel that the meaning should be clear.

Geological outline

We present here a very simplified geological history of southern Norway; more detail has been provided by Ramberg et al. (2008) and references therein. References are given to sources of information not covered by Ramberg et al. (2008).

The Southern Scandes (south of Trondheim; 63°N) consists of four geological provinces. (1) A SW–NE-trending belt of Caledonian metamorphic rocks thrust over the basement as a succession of nappes. NW of this belt are (2) Proterozoic high-grade metamorphic rocks that were folded and metamorphosed in the Caledonian event. SE of the belt is (3) Proterozoic high-grade metamorphic basement unaffected by the Caledonian Orogeny, which is separated from its continuation eastward into Sweden by (4) the Permo-Carboniferous Oslo Rift. The Oslo Rift contains Cambrian–Silurian sediments and a complex succession of Permo-Carboniferous as well as Lower Triassic sediments, and volcanic and intrusive rocks.

Paleozoic

The metamorphic history of the Precambrian basement need not concern us here. Precambrian basement rocks were, however, at the surface in the Cambrian when they were eroded to form a very flat surface, the Sub-Cambrian Peneplain (Lidmar-Bergström et al. 2013; Gabrielsen et al. 2015). This surface was transgressed during the Cambrian–Ordovician and buried below a sequence of marine mudstones and sandstones. The Caledonian Orogeny had a long, complex history between 500 and 410 Ma, culminating in the collision between Laurentia and Baltica between 420 and 410 Ma when a succession of nappes were thrust eastwards over the basement and some of the Cambrian–Ordovician sediments. The lowermost thrusts are found at the base of and within the sediments that acted as a décollement, and that were metamorphosed to greenschist facies (the so-called phyllites).

The Caledonian mountains collapsed in the early Devonian on extensional faults both within and to the NW of the preserved Caledonian metamorphic rocks. The amount of extension was large, shown by the presence of eclogite-facies rocks, brought up from depths of 60–100 km, in contact with overlying Devonian sediments in west Norway. The Cambrian–Silurian sediments within the Oslo Graben appear to be unmetamorphosed equivalents of the phyllites, and remnants of similar sediments are found scattered over much of southern Sweden, showing that the transgression over the Sub-Cambrian Peneplain was very widespread.

Further clues to the late Paleozoic history of vertical motion of the study area may be derived from the sedimentary basins offshore Norway and in Greenland. By mid-Permian times, the Caledonian mountains in Greenland had collapsed, and the sea transgressed a flat peneplain (Haller 1971). To the west and south of Norway, a similar development took place at that time. Marine Zechstein (upper Permian) sediments lie above the eroded remnants of the Caledonian mountains in the northern North Sea (between southern Norway and Scotland), above subsided Caledonian rocks in the southern North Sea and above remnants of the Variscan mountains in northern Germany (Ziegler 1985; Glennie et al. 2003; Peryt et al. 2010). It is therefore possible that the low-lying, late Permian landscape also included what is now the Southern Scandes.

Mesozoic

Three phases of rifting affected the area to the west of southern Norway, in the Permo-Triassic (c. 250 Ma), Late Jurassic (c. 165 Ma) and Early Cretaceous (c. 110 Ma). Each was followed by a period of continued sedimentation as the basins subsided thermally.

Along much of the coast of southwestern Norway, Jurassic sediments rest on basement, and further offshore they are progressively more deeply buried below Cretaceous and younger sediments (Figs 1a and 2). Three outliers of Jurassic sediments occur along the west coast of Norway (Bøe et al. 2010). Of particular interest for this study is the Upper Jurassic (early to middle Oxfordian), marine Bjorøy Formation encountered near Bergen (Fossen et al. 1997), similar to much of the offshore region, just west of the study area, where Jurassic sediments rest on basement (Fig. 2). Andersen et al. (1999) suggested that several of the Devonian extensional faults within the Caledonian nappes were reactivated as extensional faults during the Permian and Mesozoic, and this view was recently supported by Ksienzyk et al. (2016), who showed that faults in the Bergen area were active at 300–280, 210–190 and 120 Ma. Such faulting is consistent with the former presence of late Paleozoic and/or Mesozoic basins above the rocks that today form the Southern Scandes.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Seismic profile off SW Norway. Noteworthy features are the irregular basement topography (hilly relief, see enlargment) covered by Upper Jurassic sediments (see Fig. 4b) and the pronounced tilting and truncation of Neogene and older strata towards the present-day landmass below Quaternary strata, a configuration that is indicative of late uplift, as seen offshore many elevated passive continental margins around the world (Japsen et al. 2012a). The base of the Neogene strata also represents a pronounced low-angular unconformity representing removal of c. 750 m of section (Japsen et al. 2010). Data courtesy of Statoil; seismic interpretation by L. N. Jensen, Statoil and C. Andersen, GEUS.

In Late Cretaceous–Danian time, pelagic ooze accumulated at the bottom of the epicontinental sea that covered much of NW Europe. Today these deposits, which contain very limited amounts of siliciclastic material, form the main constituent of the Chalk Group (or Shetland Group), which reaches thicknesses greater than 1 km along the coast of Norway south of 59°N but is totally eroded below the base-Quaternary unconformity towards the landmass (Scholle 1977; Ziegler 1990; Japsen 1998). Sømme et al. (2013b) found that the Chalk Group in this region contains no mappable siliciclastic units and that there is no evidence of progradational units originating from a hinterland to the east. Sømme et al. (2013b) combined these observations with the volume of sediment stored in point-sourced depocentres along the margin to conclude that the relief of what is now onshore southernmost Norway was less than 500 m in the Late Cretaceous.

Cenozoic

Prior to the onset of seafloor spreading in the NE Atlantic at 55 Ma, there appears to have been minor uplift and erosion of a generally low-lying Norwegian area, shown by the appearance of small sand fans off western Norway during the Paleocene (Ahmadi et al. 2003). Uplift of Norway at this time must, however, have been much less than that of Scotland, as most of the sand forming the Paleocene fans in the Viking and Central Graben areas came from the Moray Firth area and the Orkney–Shetland ridge (Ahmadi et al. 2003). Sømme et al. (2013b) estimated the Paleocene palaeotopography in southernmost Norway as less than 500 m.

Hemipelagic, deep-marine sediments of Eocene age are present offshore along the coast of southern Norway (Heilmann-Clausen et al. 1985; Faleide et al. 2002). In fact, a compilation of stratigraphic and sedimentological evidence from NW Europe led Knox et al. (2010) to suggest the presence of an Eocene Baltic Seaway from western Siberia to the North Sea, across southern Scandinavia. In that case, only the present-day highest summits in southern Norway would have emerged from the Eocene seas. In agreement with this conclusion, Sømme et al. (2013a) estimated the maximum early Eocene palaeotopography of southwestern Norway to be only c. 200 m.

During the early Oligocene, major clastic wedges prograded southwards from present-day Norway and reached their maximum extent at 29 Ma (Schiøler et al. 2007). This change in depositional environment from Eocene times indicates a phase of uplift and erosion of southern Norway that started at about 33 Ma, a few million years before the maximum extent of the Oligocene prograding wedges (Gregersen et al. 1998; Clausen et al. 1999; Faleide et al. 2002; Jarsve et al. 2015). Jarsve et al. (2015) showed that the creation of accommodation space during the Oligocene subsidence was out of phase relative to eustatic sea-level changes, and was controlled mainly by regional-scale differential vertical movements; that is, uplift and erosion of the hinterland landmasses (southern Norway) occurred concurrently with basin subsidence. Sømme et al. (2013a) estimated the maximum early Oligocene palaeotopography of southwestern Norway to reach elevations above 1.6 km north of Sognefjord (north of 61°N).

The Oligocene–Miocene transition was a period of distinct change in the depositional environment in the eastern North Sea Basin, south and SW of Norway (Rasmussen 2004, 2017). In the late Oligocene, the Danish Basin south of the Oligocene deltas was sediment-starved and hemipelagic sediments accumulated, but in the earliest Miocene, delta systems prograded into the basin from the NNE from present-day Norway and southern Sweden. Three periods of delta progradation took place, but whereas braided river systems dominated the fluvial system in earliest Miocene, meandering rivers became increasingly dominant over time. According to Rasmussen (2014, 2017), early Miocene uplift and erosion in the Scandinavian hinterland resulted in progradation of deltas up to 300 m thick and to formation of the braided fluvial systems and deposition of thick, coarse-grained sequences across much of the present-day Danish onshore area (see also Rasmussen et al. 2010). The occurrence of both immature and mature sediments in these sequences reflects erosion of weathered as well as newly exposed basement and shows that these sequences contain both reworked sediments and erosional products from strongly weathered basement. Early Miocene exhumation also affected the present-day offshore areas west and south of southern Norway (Japsen et al. 2007, 2010) as well as southern Sweden (Japsen et al. 2016).

In the late Miocene, marine sediments reaching a thickness of 200–300 m were deposited in a gently subsiding basin west of present-day southern Norway (Fyfe et al. 2003) in water no more than a few hundred metres deep. However, uplift of mid-Norway and the Northern Scandes in the late Miocene–early Pliocene resulted in deposition of the prograding Molo Formation (Eidvin et al. 2007; Løseth et al. 2017). Major progradation into a marine basin at least a kilometre deep also occurred off southern Norway during the Pliocene (Fyfe et al. 2003), indicating both significant uplift of the source area onshore and subsidence of the basin to provide the accommodation space. Japsen et al. (2007) used palaeothermal and palaeoburial data to show that a final phase of exhumation of the Danish Basin began in the early Pliocene.

In summary, geological evidence defines an episodic history of burial and exhumation during the Phanerozoic development of southernmost Norway and surrounding regions. Thick covers of Paleozoic sediment and Caledonian nappes accumulated over the Precambrian basement, but across most of the area, the Precambrian basement was subsequently exhumed; where Jurassic sediments rest on Precambrian basement, the exhumation occurred prior to Jurassic reburial. Subsequently, three main phases of uplift affected southern Norway after the onset of seafloor spreading in the NE Atlantic: at the Eocene–Oligocene transition at c. 33 Ma, in the early Miocene at c. 23 Ma and in the early Pliocene at c. 4 Ma (Stoker et al. 2005; Japsen et al. 2007, 2010).

The landscape of the Southern Scandes

Elevated plateaux as evidence of Cenozoic uplift

Early geomorphologists regarded the combination of elevated plateaux and deeply incised valleys of the Southern Scandes to be the result of Tertiary uplift. Reusch (1901) therefore made a clear distinction between the young, pre-glacial (fluvial) incised valleys and the near-flat, elevated plateaux, for which he coined the term palaeic (old) relief (e.g. Hardangervidda at 1.2 km above sea-level (a.s.l.); Figs 3 and 4). Wråk (1908) identified four stepped surfaces in southern Norway, which could be followed eastwards along the flanks of the major valleys at high positions in the landscape. He argued that these steps had formed as a result of pronounced uplift events along the western margin of Norway, and that some of these valleys had later been captured by west-flowing rivers. Ahlmann (1919) had a similar view to Wråk (1908) and regarded the ‘viddas’ (low-relief surfaces at high elevation) of southern Norway as representing uplifted peneplains. He also pointed out that the earlier water divide had a more westerly position than at present. Gjessing (1967) was inspired by the idea of Büdel (1963), who argued that deep weathering was important for the evolution of landscapes. Gjessing (1967) thus described the Palaeic Relief as ‘hills and basins interconnected with passes’ and showed that deep weathering and subaerial slope processes were important for the shape of the palaeic forms. He thus regarded the landscape as ‘the paleic surface’ (i.e. one surface and unrelated to base level), and thought that the landforms were formed during warm climates and not by tectonic events.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Topographic profile across the western flank of the Southern Scandes, from the coast to Hardangervidda (location is shown in Fig. 1a and c). Three lines are shown on the topographic profile; the actual profile is represented by the middle line, and the maximum and minimum topographies within a swath 10 km on each side of the profile are shown as the upper and lower lines, respectively. The flank dips steadily from elevations around 1000 m a.s.l. to the coast, but between the deep valleys that cut into the flank, the landscape is dominated by irregular hills cut by linear valleys; the ‘hilly relief’ and ‘fracture valleys’ (Rudberg 1960; Lidmar-Bergström 1995), characteristic features of Mesozoic weathering during a warm and humid climate. The presence of an Upper Jurassic outlier near Bergen confirms the Mesozoic age of the sloping basement surface that is characterized by a hilly relief to high elevations (Fig. 1c), and its thermal maturity (VR) shows that it has been buried by at least 1 km of Jurassic and younger sediments. This Mesozoic peneplain or denudation surface continues offshore, west of Bergen, where it is preserved below a cover of Jurassic and younger rocks (Fig. 2). At elevations of c. 1200 m a.s.l., the profile is dominated by the subhorizontal Hardangervidda, which is a low-relief plateau formed by post-Caledonian erosion (Fig. 4a) and the third-lowest of the four steps (Level III) in the Palaeic Relief as defined by Lidmar-Bergström et al. (2000). The profile thus reveals the two main components of the landscape of the southernmost Scandes: a tilted Mesozoic peneplain (green dashed line) with a hilly relief along the SW coast, cut by deep valleys and truncated by the subhorizontal, Hardangervidda surface (blue dashed line). Comparison with the similar configuration of surfaces defining the South Swedish Dome suggests that Hardangervidda is a Cenozoic erosion surface (Lidmar-Bergström et al. 2013) and our AFTA data show that it formed during the Miocene.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Contrasting landscape elements of the Southern Scandes. (a) Hardangervidda. (a1) View towards the west from the northwestern part of Hardangervidda at c. 1150 m a.s.l. (photo location is shown in (a2)). The rolling topography of the low-relief surface at 1100–1200 m a.s.l. is cut by deep glacially eroded valleys of the Hardangerfjord. (a2) Elevation illustrating the low relative relief of Hardangervidda; detail from Figure 1. (a3) Geological profile across Hardangervidda (location is shown in (a2)). The metamorphosed Cambrian–Ordovician sediments (phyllites) rest in place on the Sub-Cambrian Peneplain where the basal Caledonian thrusts are within the phyllites. Elsewhere, the basal Caledonian thrust forms the contact between the phyllites and basement (Schiphull 1974). The phyllites (and thus the Sub-Cambrian Peneplain) are preserved below Hardangervidda in several synclinal folds, and this demonstrates that Hardangervidda is a post-Caledonian erosion surface. (b) Remnants of sub-Mesozoic hilly relief and fracture valleys (‘sprekkedale’). The photographs show areas between 800 and 1100 m a.s.l. Photo locations are shown in (b4). (b1), (b2) Hilly relief at elevations around 1000 m (see also Green et al. 2013, fig. 4). The hilly relief has been modified by overriding ice, resulting in enhanced relief (see Olvmo et al. 1999); some glacial features have formed, such as a small cirque on the flank of the left-hand hill in (b1) (Kvernafjellet, 965 m a.s.l., rises 190 m above the lake in the foreground), and there are ubiquitous minor forms such as striae and shatter marks. (b3) Detail of a typical, linear fracture valley at an elevation of c. 950 m a.s.l. It should be noted how the fracture pattern controls the side of the valley as well as its direction. (b4) Digital elevation model of the area east of Stavanger shown in Figure 1c. The area is criss-crossed by fractures that have weathered to form fracture valleys and hilly relief structures, characteristic features of sub-Mesozoic denudation surfaces elsewhere in Scandinavia. The deepest of the valleys (e.g. Lysefjord) have been incised to their present depths by large valley glaciers, but the entire area has been modified by glacial erosion to some extent. Nonetheless, pre-glacial landscape elements occur throughout the landscape, in agreement with the conclusion of Lidmar-Bergström et al. (2013) that the flanks of the Southern Scandes are remnants of a sub-Mesozoic peneplain.

Riis (1996) correlated offshore geology and onshore morphological elements in Scandinavia and interpreted deep weathering profiles encountered in the mountainous areas to be relics of Mesozoic denudation. From this, Riis (1996) inferred that the enveloping summit level originated as a peneplain in the Jurassic. He correlated the main Palaeic Relief of the Southern Scandes with the Tertiary plains with residual hills in northern Sweden and based on offshore sediment volumes he concluded that Scandinavia had been uplifted in two phases, Paleogene and Neogene.

Lidmar-Bergström et al. (2000) used the summit surfaces to identify four main steps with a vertical separation of 1500 m within the Palaeic Relief of the Southern Scandes, and interpreted each of them as the result of different episodes of fluvial incision to a base level following uplift. Lidmar-Bergström et al. (2013) argued that the Palaeic Relief reached its present-day elevation after major late Cenozoic uplift because the incision of the deep and narrow valleys of the Southern Scandes begins at the lowest level of the Palaeic Relief at about 800–1200 m a.s.l. No modern mapping of the surfaces in Norway has been done, except for an area around northern Gudbrandsdalen valley at c. 62°N (Bonow et al. 2003). However, this mapping documented the presence of stepped surfaces at high elevation along the main valleys as well as the fluvial origin of these valleys. The study thus confirmed the results of earlier studies of the Palaeic Relief (e.g. Reusch 1901; Lidmar-Bergström et al. 2000).

Some researchers have suggested that the Palaeic Relief is the result of highly efficient and extensive glacial and periglacial erosion (Steer et al. 2012; Egholm et al. 2017). However, Kleman (2008) in Baffin Island and Hall et al. (2013) in western Norway demonstrated that the plateaux were mainly unaffected during the glaciations owing to cold-based, non-erosive, conditions. Significant erosion took place predominantly in valleys where the ice was warm-based. Recently, Andersen et al. (2018) used cosmogenic isotope data to show that plateau erosion accounts for only 10% of the Quaternary erosion output from the Sognefjorden drainage basin (just north of our study area). Further, Steer et al. (2012) and Egholm et al. (2017) ignored the observation that the glacial valleys exploited pre-existing fluvial valleys (Bonow et al. 2003). However, the preservation of fluvial valley shoulders on the plateaux shows that glacial erosion cannot have removed a 300–400 m thick rock column as suggested by Steer et al. (2012). They also failed to take into account that low-relief plateaux similar to those in the Norwegian landscape occur in both glaciated and non-glaciated areas (Lidmar-Bergström et al. 2000; Bonow et al. 2009; Green et al. 2013); for example, Lidmar-Bergström et al. (2000) showed that major landscape features such as high plateaux, a great escarpment and a coastal plain are similar in southern Norway and eastern Australia. These observations show that the formation of elevated plateaux along passive continental margins is not related to glacial action.

Clayey saprolite, fracture valleys and hilly relief as evidence for Mesozoic weathering

Lidmar-Bergström (1995) showed that deep weathering of the fractures in basement rocks containing feldspars by meteoric water in a warm and humid climate can lead to the formation of distinctive landscape types characterized by undulating hilly relief and by fracture valleys (‘sprekkedale’). The meteoric water penetrated fractures and altered feldspar to kaolinite that was subsequently preferentially eroded, forming valleys typically up to 200 m deep between areas of unweathered rock. The warm and humid conditions necessary for the formation of kaolinitic saprolites prevailed in Scandinavia during the Rhaetian–Campanian (Lidmar-Bergström 1982). The resulting hilly relief with associated saprolites was subsequently preserved below a cover of Mesozoic and younger rocks, as at Ivö in southern Sweden where the hilly relief is capped by Upper Cretaceous limestone and chalk (Lidmar-Bergström 1989). Where these protective sedimentary covers have been removed in the recent past, the hilly relief with remnants of kaolinitic saprolite between the hills has been re-exposed and the surface of the basement hills slightly altered, in particular by glacial action (see Lidmar-Bergström et al. 2017, fig. 3).

There is evidence of late Mesozoic weathering of basement both onshore and offshore southern Norway. Seismic lines around Norway show a pattern corresponding to hilly relief at top basement below Jurassic sediments (Fig. 2), and Riber et al. (2015) showed evidence of subaerially weathered basement rocks on the Utsira High (off SW Norway) prior to burial at the Jurassic–Cretaceous transition. Migón & Lidmar-Bergström (2001) mentioned reports of kaolin at several localities around southern Norway. Lidmar-Bergström & Näslund (2002, fig. 4) showed an extensive area of hilly relief to the east of the Oslo Rift, where Olesen et al. (2007) found that structural weakness zones containing clay minerals, including smectite and kaolinite that, to a large extent, are the result of chemical weathering of pre-existing fracture zones during subtropical conditions in the late Mesozoic. Olesen et al. (2007) argued that the weathering penetrated deep into the fracture zones and was preserved below a sedimentary cover until exhumation removed the cover and most of the weathering products while preserving clay zones at depths of 200–300 m.

The Norwegian strandflat is an uneven and partly submerged rock platform along the coast (Reusch 1894). Holtedahl (1998) thought that the strandflat probably formed in late Pliocene or Pleistocene times as a product of glacial erosion, marine erosion and subaerial weathering, and that it coincides with an exhumed pre-Jurassic surface in some coastal areas. Fredin et al. (2017) obtained ages of 221–206 Ma (Norian, Late Triassic) from K–Ar dating of authigenic, syn-weathering illite from saprolitic remnants at Bømlo on the strandflat of southwestern Norway (Fig. 1a). Assuming the K–Ar dates to be accurate, they imply that a surface close to at least part of the strandflat must have been near the Mesozoic surface at that time. Its preservation is unlikely unless it had been covered by sediments in Mesozoic and Cenozoic times that were stripped off again more recently. The known kaolinitic saprolites on the strandflat of west and south Norway (Migón & Lidmar-Bergström 2001) cannot, however, have formed during the Norian because formation of kaolin requires warm wet conditions, and northern Europe was a desert during the Norian. Warm and wet conditions prevailed from the Rhaetian to Campanian, so kaolin is more likely to have formed during this period (Lidmar-Bergström 1982).

Peneplains and topographic highs of Scandinavia

Stratigraphic landscape analysis provides evidence of uplift and subsidence in shield areas, using cross-cutting relationships between peneplains and stratigraphic constraints from the sedimentary cover that preserved them to construct a relative chronology for surface formation and tectonic events (Lidmar-Bergström et al. 2013). Here we adopt the definition that all surfaces graded to a base level are peneplains, irrespective of their detailed form (hilly or flat) (Green et al. 2013; Lidmar-Bergström et al. 2017). In any case, the formation of a peneplain involves kilometre-scale erosion of rock during valley incision and valley widening by running water and slope processes, whereas the final shape of the peneplain depends on the climatic conditions during its formation.

Lidmar-Bergström et al. (2013) described three topographic highs that characterize Scandinavia: the Northern Scandes, the Southern Scandes and the low South Swedish Dome (2.1, 2.5 and 0.4 km a.s.l., respectively; Fig. 1a). The South Swedish Dome emerges from below Cambrian cover rocks in the north and east and from below Cretaceous cover rocks in the south and west, and is delimited to the SW by the Sorgenfrei–Tornquist Zone. Lidmar-Bergström et al. (2013) emphasized that these geological constraints permitted a detailed stratigraphic landscape analysis of the South Swedish Dome, which provides the key to understanding the Phanerozoic history of vertical movements of Scandinavia.

Three major peneplains define the South Swedish Dome of which two are re-exposed peneplains (Lidmar-Bergström et al. 2017). The sub-Cretaceous peneplain, characterized by undulating hilly relief and showing up to 200 m difference between the summit of basement hills and the fresh basement below thick kaolinized basement rocks (saprolite), stands in stark contrast to the extremely flat Sub-Cambrian Peneplain. Both re-exposed surfaces developed close to former sea-levels, were subsequently transgressed, and were buried below sedimentary covers. The preservation of these peneplains below Paleozoic and Upper Cretaceous cover rocks, respectively, documents that uplift of the land surface had been followed by subsidence and burial, and that uplift and erosion in the recent past has brought the peneplains back to the surface. Cross-cutting relationships between these re-exposed and tilted peneplains and a third, horizontal plain with residual hills (South Småland Peneplain) demonstrate that the latter is younger than the sub-Cretaceous peneplain and is thus of post-Cretaceous age.

Lidmar-Bergström et al. (2013) showed that the three relief types of the South Swedish Dome can be identified all around Scandinavia, and that their presence demonstrates phases of uplift/denudation and subsidence/burial of Scandinavia during the Phanerozoic. They argued that the subhorizontal peneplains of the Northern Scandes, the Southern Scandes and the South Swedish Dome are Cenozoic erosion surfaces because these subhorizontal plains truncate tilted, sub-Mesozoic peneplains (Fig. 3). The flanks were thus at the surface and weathered during the Mesozoic but were subsequently covered and tilted prior to their re-exposure during relatively recent uplift.

Key landscape elements of the southernmost Scandes

Overall, our study area, the southern half of the Southern Scandes, represents half of an elongated dome with slopes dipping to the west, south and SE from high ground in the centre (Figs 1b and 3). A subhorizontal plateau, Hardangervidda, with a very low relative relief, occupies much of the higher ground, at elevations of c. 1200 m. Hardangervidda corresponds mainly to the third-lowest of the four steps (Level III) in the Palaeic Relief as defined by Lidmar-Bergström et al. (2000). Level IV occurs some hundreds of metres below Hardangervidda and corresponds to the bottom level of the Palaeic Relief.

The slopes are incised by deep and narrow valleys below Level IV (Lidmar-Bergström et al. 2000), commonly showing distinct fluvial character, whose final development involved enlargement by valley glaciers (Bonow et al. 2003). Many irregular hills cut by linear valleys also occur between the deep valleys. At low levels around much of the region, we find irregular bedrock surfaces, which Lidmar-Bergström (e.g. Lidmar-Bergström 1995) described as undulating hilly relief and fracture valleys. In coastal regions, an uneven and partly submerged rock platform, the strandflat, occurs (Holtedahl 1998). In places, the amount of glacial erosion has been so great that it has obliterated much of any older landscape elements, but as we discuss below, there appear to be areas both on the top and on the flanks of the dome where substantial pre-Pleistocene landscape elements have been preserved.

Hardangervidda

The central and eastern parts of Hardangervidda constitute a very low-relief plateau at an elevation of c. 1200 m a.s.l. (Figs 1c and 2). The plateau is bounded to the north by the 600 m high escarpment forming the southern flank of Hallingskarvet mountain, which consists of Caledonian-age rocks of the Jotun Nappe lying on Cambro-Ordovician metasediments (the so-called phyllites). The lowest thrusts of the Caledonian Orogeny are either within or at the base of the phyllites, which appear to have acted as a décollement to the nappes. The phyllites lie in turn on the Sub-Cambrian Peneplain.

Southern and western Hardangervidda has much more relief than other parts of the plateau. This relief consists of Caledonian rocks; these are mostly phyllites, but also Jotun Nappe rocks in a few places.

Eastern and central Hardangervidda has a very low relative relief. Gabrielsen et al. (2015) argued that this is because this area coincides with the Sub-Cambrian Peneplain, and the surface is certainly very close to the Sub-Cambrian Peneplain. But this was folded slightly during the Caledonian Orogeny, and in some places the Hardangervidda surface has definitely been eroded below the Sub-Cambrian Peneplain, so the two do not coincide (Fig. 4a3). Preservation of phyllites, and thus the Sub-Cambrian Peneplain, below the Hardangervidda surface in several synclinal folds shows that Hardangervidda is a post-Caledonian feature. This interpretation agrees with that of Lidmar-Bergström et al. (2013), who argued that the near-horizontal plain of Hardangervidda formed by erosion to the base level of the adjacent ocean in the Cenozoic, in contrast to the inclined sub-Mesozoic peneplains that define the flanks of the Southern Scandes.

The southwestern flank of the Southern Scandes

The southwestern flank of the Southern Scandes dips steadily from elevations around 1000 m a.s.l. to the coast (Figs 1b and 3). It is incised by deep valleys that have been modified by glacial action, but the slope is not uniform between the deep valleys and consists of many irregular hills cut by linear valleys even at elevations of 1000 m a.s.l. (Fig. 4b). The valleys appear to have been eroded along pre-existing fractures and are very reminiscent of the joint-valleys and hilly relief described by Lidmar-Bergström (1995). Both types of relief may be generated by formation of kaolinitic saprolites by weathering of feldspars in fractures in the bedrock by meteoric water in subtropical climates.

Hilly relief and fracture valleys thus dominate the landscape patterns at elevations reaching 1 km a.s.l. around the flanks of the Southern Scandes. This agrees with the conclusion of Lidmar-Bergström et al. (2013) that the flanks formed as a sub-Mesozoic peneplain by denudation during the warm and humid late Mesozoic, and that the landscapes have been buried and protected from erosion since then by sediment that has been removed only in recent geological time. The hilly relief is everywhere located at lower locations compared with the near-horizontal, Cenozoic plain of Hardangervidda (Lidmar-Bergström et al. 2000, 2013; Green et al. 2013, fig. 4). The sub-Mesozoic peneplain continues offshore where it is preserved below a thick cover of Jurassic and younger rocks (Fig. 2), and the presence of an Upper Jurassic outlier near Bergen shows that Jurassic sedimentation onto basement extended at least into the coastal areas of present-day Norway (Fossen et al. 1997).

Apatite fission-track data from southern Norway

Previous thermal history studies

A number of studies have applied apatite fission-track thermochronology in different parts of southern Norway to investigate the exhumation history and the development of the present-day topography of the region. Earliest studies revealed that Precambrian basement rocks at outcrop are characterized by apatite fission-track ages of c. 250 Ma or younger (Andriessen 1990; Grønlie et al. 1994; Rohrman et al. 1995, 1996). These results, together with track-length data, were interpreted as showing that the samples cooled from temperatures of 100°C or above in Triassic–Jurassic times, reflecting Mesozoic exhumation from depths of several kilometres. Olaussen et al. (1994) used onshore apatite fission-track data and offshore seismic data to infer that Early Triassic and ‘mid’ Jurassic episodes of exhumation removed any post-rift basin fill from the Upper Carboniferous–Permian Oslo Rift and the Skagerrak Graben. Rohrman et al. (1995) further interpreted their data as defining a later, Neogene phase of exhumation that formed a domal uplift of southern Norway, responsible for the modern-day topography of the region. Early apatite fission-track studies in southern Sweden reached broadly similar conclusions (Zeck et al. 1988; Larson et al. 1999; Cederbom et al. 2000; Cederbom 2001; Huigen & Andriessen 2004).

Redfield et al. (2004, 2005a,b) confirmed the importance of Mesozoic cooling and revealed major offsets across fault zones along the west coast of central and northern Norway. Data from a number of studies were compiled by Hendriks et al. (2007), demonstrating a consistent regional pattern of variation in apatite fission-track age across the study area (Fig. 5a) and further afield into Sweden and Finland. Youngest ages, typically less than 150 Ma, are found in the centre of the Southern Scandes and further north, somewhat older ages up to c. 250 Ma are found around the coast of southern Norway and into Sweden, and much older ages in central and northern Sweden and Finland. The compiled data show a complex relationship between apatite fission-track age and mean confined track length (Fig. 5b), suggesting significant variation in thermal history across the region.

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Apatite fission-track data from this and previous studies (Hendriks et al. 2007; Japsen et al. 2016). (a) Map of apatite fission-track ages in southernmost Norway. (b) Mean track length v. apatite fission-track age for Scandinavian outcrop samples. Data from this study are consistent not only with compiled data from Norway (Hendriks et al. 2007) but also with data from southern Sweden (Japsen et al. 2016), suggesting a coherent thermal history framework across the region. It should be noted that measured fission-track ages from outcrops across southern Norway are dominated by values less than 250 Ma, emphasizing the importance of post-Paleozoic events in controlling the development of the Southern Scandes.

Japsen et al. (2007, 2010) applied AFTA, vitrinite reflectance (VR) and sonic data in wells to study the exhumation history of the offshore sedimentary basins south and west of southern Norway (Fig. 1a). Japsen et al. (2010) integrated their results with evidence from offshore stratigraphy and suggested that southern Norway had undergone three phases of Cenozoic uplift and erosion, starting at c. 33, c. 24 and at c. 4 Ma. Thus, a number of AFT studies have supported earlier inferences from geomorphological investigations that the mountains of southern Norway were uplifted within the relatively recent geological past.

Based on a reinterpretation of a number of apatite fission-track datasets, including Rohrman et al. (1995) and other published and unpublished sources, Nielsen et al. (2009) presented an alternative view, involving more or less continuous cooling/denudation since the Caledonian Orogeny. We discuss this view further in the section on mechanisms.

Later apatite fission-track studies of Norway have been focused around fault movements in the north of the country (Redfield & Osmundsen 2009; Hendriks et al. 2010; Osmundsen & Redfield 2011; Davids et al. 2013). Ksienzyk et al. (2014) provided new apatite fission-track data from southern Norway in the vicinity of Bergen, and also reported the first apatite (U–Th)/He ages from Norway. Their results revealed significant Mesozoic fault offsets, and concluded that the modern-day topography is the result of repeated pre- and post-rift tectonic episodes.

Thus, despite a number of studies over the last 30 years, opinions remain divided in regard to the interpretation of low-temperature thermochronology data and the development of the modern-day topography of southern Norway.

New AFTA data from southernmost Norway

Observations

A total of 28 samples of Precambrian basement were collected for AFTA across southernmost Norway (Supplementary Table S1 and Supplementary Material S2). Most samples were from outcrop, but in addition to the regional spread of AFTA samples, we collected two near-vertical profiles, and we have access to three samples from a borehole near Stavanger (Fig. 1c). Apatite fission-track ages vary from 156 ± 15 Ma to 256 ± 16 Ma, and mean confined track lengths are between 10.3 ± 0.8 µm and 14.41 ± 0.1 µm. The data define a similar relationship between mean track length and fission-track age to that shown by the compiled data from Hendriks et al. (2007) (Fig. 5b), typified by a gradual decrease in mean track length with decreasing fission-track age. Data from southern Sweden reported by Japsen et al. (2016) define a similar trend (with a few older ages, representing samples that were not totally annealed during the Mesozoic palaeothermal maximum). As with the Hendriks et al. (2007) dataset, apatite fission-track ages across southern Norway show little variation (Fig. 5a). The AFTA data from southern Sweden and southernmost Norway are very similar, suggesting a uniform thermal history across the region. This is confirmed by the thermal history interpretation of individual samples, discussed in the next section.

Interpretation

The principles involved in extracting thermal history constraints from AFTA data in studies such as this have been discussed in detail elsewhere (Green & Duddy 2012; Green et al. 2013). The controlling principle is that in samples that are heated to a maximum palaeotemperature and then cooled, the data are dominated by the palaeothermal maximum, and the prior history is overprinted. For this reason, it is not possible to constrain the entire history, so instead we focus on defining the palaeothermal maximum and subsequent palaeothermal peaks, within an overall framework of episodic heating and cooling reflecting the episodic nature of the Phanerozoic development of the study area documented by the geological record and by the contrasting types of bedrock relief as summarized in the previous sections. As explained in detail elsewhere we believe that this type of history is more geologically realistic than those involving slow, monotonic cooling (Green & Duddy 2012; Green et al. 2013, 2018). This belief is based on AFTA data in situations where geological constraints provide key insights into the history, defining not only episodes of cooling/exhumation but also prior periods of heating/burial, which cannot be explained by slow continuous cooling/exhumation histories.

Thermal history solutions, represented by 95% confidence limits on the maximum or peak palaeotemperature and the time at which cooling from that palaeotemperature began, are obtained by comparing measured data with values predicted from candidate thermal histories involving up to three episodes of heating and cooling (three being the maximum number of cooling episodes that can be obtained from AFTA data in most cases). Where three episodes are defined, the earliest episode is defined primarily from the fission-track age data (and represents the episode in which samples cool below c. 110°C and began to retain tracks), whereas the two more recent episodes are defined principally from the distribution of track lengths (see Fig. C.21 of Appendix C in Supplementary material S2). By varying the magnitude of maximum paleotemperature and the onset of cooling in each episode (using assumed heating and cooling rates each of 1°C Ma−1), the range of values giving predictions that are consistent with the measured data within 95% confidence limits is defined.

Thermal history constraints extracted from the AFTA data in each sample, following the strategy outlined above, are summarized in Supplementary Table S1. Figure 6 shows a comparison of the constraints (95% confidence intervals) on the onset of cooling in up to three palaeothermal episodes in each sample; that is, times when samples were hotter than they are today. If we assume that data in individual samples represent synchronous cooling episodes resulting from regional processes such as exhumation or reduction in heat flow, the results in almost all samples from this study can be explained in terms of just three Phanerozoic cooling episodes, beginning in the following intervals (assignment of chronostratigraphic age based on Gradstein et al. 2012): 251–243 Ma; Early to Middle Triassic; 172–164 Ma; Middle Jurassic; 30–21 Ma; Oligocene–early Miocene.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

Constraints on the timing of onset of cooling in palaeothermal episodes recognized from AFTA data in this study (Supplementary table S1). Synthesis of all data suggests three dominant episodes of cooling, as indicated by the vertical bands with different colours. In addition, an earlier onset of cooling is seen in three samples, broadly defining a Paleozoic cooling event. The stratigraphic age of Precambrian basement samples plots outside the diagram. Constraints that are not consistent with the preferred timing for the major regional events are highlighted with a yellow background.

It should be noted that these intervals define the uncertainty on the time at which cooling began in each episode, and we do not imply that all cooling took place within these intervals. Three timing constraints that are not consistent with the above synthesis, highlighted in Figure 6, are regarded as statistical outliers. In addition, an earlier cooling episode is seen in three samples, consistent with cooling that began between 411 and 333 Ma, probably reflecting late-stage Caledonian cooling.

Regional synthesis

As we noted earlier, the AFTA data in samples from this study are highly consistent with those from our study of southern Sweden (Japsen et al. 2016), and the three regional cooling events defined above closely correlate with cooling events defined in Sweden; namely, 245–240, 171–142 and 23–15 Ma. It should be noted that the data from southern Sweden defined a late Carboniferous event of cooling and exhumation (beginning between 314 and 307 Ma) that could not be resolved in the data presented here, possibly because of the high palaeotemperatures involved in the Middle Triassic thermal event. Furthermore, the dataset from Sweden also defined a mid-Cretaceous cooling event beginning between 107 and 92 Ma (Albian–Turonian), which reflects the onset of inversion of the Sorgenfrei–Tornquist Zone.

Given the consistency of the two datasets we conclude that results from both areas reflect common events, and that the best estimates of the onset of cooling in these events are as follows:

  • 245–243 Ma, Middle Triassic (Anisian);

  • 171–164 Ma, Middle Jurassic (Bajocian–Bathonian);

  • 23–21 Ma, early Miocene (Aquitanian).

Given the wide timing constraints in individual samples (Fig. 6), somewhat wider uncertainty ranges should probably be applied to the Triassic and Miocene episodes. In the following discussion, we follow this interpretation in terms of three regionally synchronous cooling episodes across southern Scandinavia. On the basis of the regional extent of each of these episodes, we suggest that an explanation in terms of cooling owing to exhumation is the most appropriate explanation for each of these cooling episodes, as also concluded by Japsen et al. (2016) in regard to southern Sweden.

Regional events defined from AFTA and their relation to geology and landscapes

Middle Triassic exhumation leading to removal of Paleozoic cover on the flank of the earliest North Sea rift

Almost all samples cooled from palaeotemperatures above 100°C and began to retain tracks in the Middle Triassic episode at c. 245 Ma. All samples forming the near-vertical profile from Rjukan (located at the southeastern edge of Hardangervidda; Figs 1c and 7a) define Middle Triassic palaeotemperatures >100°C, corresponding to burial below a cover many kilometres thick for any reasonable palaeogeothermal gradient. This cover is likely to have been a thick succession of Caledonian nappes and probably upper Paleozoic–Lower Triassic sediments that covered what is now Hardangervidda prior to the onset of Middle Triassic exhumation. Only three samples, all located close to the west coast, cooled from above 100°C in an earlier episode and these cooled from palaeotemperatures around 90–100°C in the Middle Triassic episode (Fig. 6). Because the rock that was removed was likely to have lithologies similar to the rock that remains, issues connected with possible non-linear thermal profiles discussed below are not significant here. These palaeotemperatures correspond to burial below c. 3 km for a palaeogeothermal gradient of 25°C km−1 or c. 1.5 km for a palaeogeothermal gradient of 50°C km−1 (assuming a palaeo-surface temperature of 20°C).

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

Palaeotemperature constraints from AFTA v. elevation compared with assumed palaeogeothermal gradients of 25 and 35°C km−1. (a) Constraints from a vertical transect from Rjukan showing palaeotemperatures for the Middle Triassic, Middle Jurassic and early Miocene paleo-thermal episodes. Extrapolation of these constraints along the assumed palaeogeothermal gradient of 25°C km−1 yields palaeotemperatures at present-day sea-level of more than 100°C, about 90°C and about 70°C, respectively. If the early Miocene palaeotemperatures are extrapolated along that gradient they reach the likely palaeo-surface temperature of 20°C at an elevation of 2 km a.s.l. (the projected surface elevation, PSE; Japsen et al. 2014). This implies that prior to the onset of early Miocene cooling, the land surface was well above present-day terrain. (b) Early Miocene paleotemperature constraints for all samples colour-coded by location. The constraints from the Rjukan and Lysefjord profile match a palaeotemperature gradient of 25°C km−1. For this gradient, the projected surface elevation for inland locations is about 2 km a.s.l. (red circle a). For the coastal locations it is slightly lower, about 1.7 km a.s.l. (red circle b). For a palaeotemperature gradient of 35°C km−1, corresponding to measured values for the sedimentary cover offshore (Japsen et al. 2010), the projected surface elevation for coastal locations is about 1.3 km a.s.l. (red circle c). The estimated palaeotemperature for a sample of the Jurassic outlier in Bergen (Fossen et al. 1997) based on a new measurement of vitrinite reflectance (0.38%) is also shown. Coastal locations include Setesdalen.

Rohrman et al. (1995) also reported evidence for Triassic exhumation in southern Norway based on apatite fission-track data, and Johannessen et al. (2013) used apatite fission-track data combined with (U–Th)/He data from the inner Hardangerfjord to define accelerated Permo-Triassic cooling. Fauconnier et al. (2014) found that the Cambro-Silurian sediments below the Jotun Nappe Basal Thrust, just north of our study area, had been heated to temperatures ranging from 320 to 500°C during the main Caledonian burial stage and thus prior to the Middle Triassic exhumation event (Fig. 8). A kilometres-thick cover of Paleozoic–Lower Triassic rocks also accumulated on the Sub-Cambrian Peneplain in southern Sweden prior to Middle Triassic exhumation (Zeck et al. 1988; Japsen et al. 2016). Up to 4 km of Paleozoic strata are preserved in the Baltic Sea and in the Sorgenfrei–Tornquist Zone in the Kattegat (Sigmond 1993; Calner et al. 2013).

Fig. 8.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 8.

Heating and cooling (burial and exhumation) history of basement rocks now exposed from below Upper Jurassic sediments along the coast of SW Norway (samples 361-21, 22, 25). The Sub-Cambrian Peneplain and Hardangervidda are found at a short distance from the coast at elevations around 1 km a.s.l. (Figs 1 and 4). A, The Sub-Cambrian Peneplain that extended across Scandinavia (Lidmar-Bergström et al. 2013) formed after uplift and exhumation that began around 600 Ma in the latest Neoproterozoic (Japsen et al. 2016). B, When the peneplain was fully developed prior to Cambrian transgression, rocks at this surface had cooled to the palaeosurface temperature of c. 20°C. C, Heating to 300°C or more below thick Caledonian nappes (Fauconnier et al. 2014). D, Cooling below palaeotemperatures of 100°C after Middle Triassic exhumation. E, Cooling to near-surface conditions and deep weathering in the Late Triassic (Fredin et al. 2017) after removal of the Paleozoic cover and destruction of the Sub-Cambrian Peneplain. F, Heating to c. 80°C below Upper Triassic to Middle Jurassic sediments. G, Exhumation after Middle Jurassic uplift leading to exposure of the bedrock and formation of a deeply weathered, hilly bedrock surface (peneplain) (Figs 2 and 4b). H, Heating to c. 60°C below Upper Jurassic to Oligocene sediments and preservation of the Mesozoic peneplain. I, Exhumation after early Miocene uplift with samples cooling to a surface temperature of c. 10°C on the strandflat.

The Middle Triassic event is likely to have exhumed Precambrian basement to near-surface conditions in the Late Triassic over parts of our study area. This is supported by K–Ar ages of epigenetic mineral deposits in the Oslo Rift (Ineson et al. 1975) and K–Ar ages of illite from remnants of saprolite on the Utsira High and on the strandflat at Bømlo, SW Norway (Fig. 1c; Fredin et al. 2017).

Kilometre-thick successions of coarse clastic sediments of the Middle–Upper Triassic Skagerrak Formation (Hegre Group) are present west and south of southern Norway and in the Danish Basin (Bertelsen 1980; Goldsmith et al. 2003; www.npd.no). The results presented here and by Japsen et al. (2016) suggest that the Skagerrak Formation includes the erosional products derived from southern Scandinavia after Middle Triassic uplift. Jarsve et al. (2015) likewise concluded that southern Norway was the main source area for the Triassic deposits in the central and eastern North Sea. The Middle Triassic event coincided with the concluding stage of the Oslo Rift, which was characterized by the deep intrusion of minor granites at 249–241 Ma (Larsen et al. 2008).

Middle Jurassic exhumation leading to formation of a Mesozoic peneplain

The effects of the Middle Jurassic cooling and exhumation beginning between 171 and 164 Ma are recorded by most of the samples analysed for this study. Samples at elevations near sea-level cooled from palaeotemperatures around 80–90°C, corresponding to burial below 1.2–2.8 km for palaeogeothermal gradients from 50 to 25°C km−1 (palaeo-surface temperature 20°C). Fredin et al. (2017) showed that the basement in western coastal regions was at the surface in the Late Triassic, so these palaeotemperatures indicate that, along the west coast, a kilometres-thick cover of Upper Triassic–Lower Jurassic sediment must have accumulated onto the basement. As Upper Jurassic sediments rest on basement near Bergen (Fossen et al. 1997) and further west offshore (Fig. 2), the sedimentary cover removed during Middle Jurassic exhumation must thus have been of Late Triassic–Middle Jurassic age. This cover was thus removed prior to re-exposure of the basement and subsequent reburial in Oxfordian time (Fig. 8), thus giving a maximum time of about 14 myr between the onset of Middle Jurassic exhumation and Oxfordian reburial. Because the events we are discussing are of regional extent, it seems likely that this former sedimentary cover extended over a wide area of the present-day onshore prior to Middle Jurassic exhumation.

Jurassic palaeotemperatures in samples from the Rjukan vertical profile (Fig. 7a) are consistent with a palaeogeothermal gradient of 25°C km−1 (although the deepest value seems to be anomalously high probably because the Triassic event is not resolved in this sample, and the single constraint probably represents the unresolved effects of Jurassic and Triassic cooling). A palaeotemperature intercept of 90°C at sea-level implies the presence of a rock column in the Middle Jurassic that reached almost 3 km above present-day sea-level. This cover is likely to have been remnant Cambro-Ordovician metasediments and rocks of the Caledonian Jotun Nappe and any Upper Triassic to Lower Jurassic sedimentary cover that extended this far. Because the Jotun Nappe rocks today extend to an altitude of nearly 2 km a.s.l. just north of Hardangervidda, there may not have been more than a few hundred metres of Mesozoic sediment above the Caledonian rocks in this central part of the study area.

A deeply weathered, hilly bedrock surface (peneplain) formed after the Middle Jurassic denudation event. Today the surface is preserved below a cover of Jurassic and younger sediments in the offshore region (Fig. 2), and it is re-exposed along the western and southern flanks of the Southern Scandes from sea-level to about 1 km altitude, just below Hardangervidda (Figs 1c and 3). We will refer to this surface as a Mesozoic peneplain as it was formed after removal of kilometres-thick rock columns and finally graded to a base level, in this case the sea as documented by the marine, Upper Jurassic sediments in Bergen (Fossen et al. 1997). The peneplain formed by weathering during the warm and humid climate in the late Mesozoic and is thus characterized by hilly relief and fracture valleys. Jurassic and possibly Cretaceous sediments accumulated on the peneplain, which consequently was preserved until its recent exposure. Stratigraphic landscape analysis thus assigns ‘sub-Mesozoic’ as the relative age of the denudation surface, whereas our results document that it is a Mesozoic surface.

Mid-Jurassic uplift affected a wide region in the North Sea (Ziegler 1990), both west and south of the present study area. Underhill & Partington (1993) combined stratigraphic observations with evidence for igneous activity from the central North Sea to conclude that regional, domal uplift resulted from the impingement of a broad (>600 km radius), transient plume head at the base of the lithosphere (‘the central North Sea Dome’). Underhill & Partington (1993) did not, however, consider the continuation of this zone of uplift towards the coast of Norway, which they classified as ‘data poor’. Rohrman et al. (1995), however, presented evidence for major Jurassic exhumation in southern Norway based on apatite fission-track data.

The stratigraphic record of the Danish Basin and the Sorgenfrei–Tornquist Zone contains ample evidence of a pronounced, mid-Jurassic phase of uplift and erosion that began in the Aalenian at c. 175 Ma coeval with pronounced volcanic activity at the southwestern margin of the Baltic Shield (Norling & Bergström 1987; Nielsen 2003; Tappe 2004). The cooling and exhumation resulting from this event is also revealed by AFTA data from southern Sweden and the Danish Basin (Japsen et al. 2007, 2016).

Early Miocene exhumation leading to formation of Hardangervidda

The effects of the early Miocene cooling and exhumation beginning between 23 and 21 Ma are recorded by the AFTA data in all but one sample. Samples close to sea-level cooled below palaeotemperatures around 60–70°C (Fig. 7). This is consistent with a new VR value of 0.38% (see Supplementary material S2) for the Upper Jurassic outlier near Bergen (Fossen et al. 1997). This VR value suggests that the Jurassic sediments were heated to a maximum palaeotemperature of 63°C (Burnham & Sweeney 1989, assuming a heating rate of 1°C Ma−1, consistent with that employed in interpreting the AFTA data) prior to exhumation. The AFTA data also suggest that the most likely timing for the maximum heating was in the Miocene. Such paleotemperatures imply burial by a considerable thickness of younger section, subsequently removed during exhumation.

Results from elevation profiles at Rjukan and other locations in Figure 7 are all consistent with a palaeogeothermal gradient of c. 25°C km−1 in basement rocks for this episode, although coastal locations appear to define a profile that is offset to lower temperatures compared with the inland locations. Linear extrapolation of the palaeotemperature profile from inland locations to a palaeo-surface temperature of 20°C implies that a cover of rocks extended to c. 2 km above present sea-level prior to the onset of Miocene exhumation, whereas the cover above coastal locations reached c. 1.6 km above present sea-level. Whereas rocks from the highest elevations tend to define fairly wide temperature constraints in this event (owing to the low temperatures involved), as shown in Figure 7 extrapolation of the profiles suggest that samples at inland elevations close to that of present-day Hardangervidda cooled from around 40°C. Thus, in this scenario the rock column that was present in the early Miocene must have extended well above all of the present-day landscape within the study area.

Estimating the amount of section that was present above the present-day surface by linear extrapolation of the palaeotemperature profile is not straightforward because the analysis depends critically on the assumption that the palaeogeothermal gradient was linear throughout the entire section when cooling began. This is equivalent to assuming that the missing section and the preserved section were of similar lithologies, such that no significant contrasts in thermal conductivity existed that may have resulted in a non-linear palaeotemperature profile (see fig. 32 of Green et al. 2013). Thermal conductivities measured in rocks of different types show wide variation from around 1 W m−1 K−1 or less to over 5 W m−1 K−1, related to lithology, water content, porosity, temperature and other factors (e.g. Robertson 1988; Clauser & Huenges 1995; Eppelbaum et al. 2014). If the cover rocks that were removed during Miocene exhumation were relatively unconsolidated sediments with lower thermal conductivities compared with the underlying basement rocks, then the palaeogeothermal gradient in the overlying section would have been higher. In that case, the amount of cover required to explain the Miocene palaeotemperatures would be correspondingly reduced, perhaps by a factor of two, based on the range of typical thermal conductivities in the studies cited above.

Therefore we can define two likely end-member scenarios for the thickness of the cover that was present above the level of the present-day Hardangervidda immediately prior to the onset of Miocene exhumation. If the rocks that were eroded during Miocene exhumation were Jotun Nappes and phyllites, then linear extrapolation of the profile in Figure 7 from c. 40°C at 1.2 km a.s.l. suggests that around 0.8 km of rock was present above the level of Hardangervidda. The thickness required could possibly be higher if we apply present-day temperature gradients in the range 15–20°C km−1 as measured in boreholes through Proterozoic–Paleozoic rocks in southern Norway (Slagstad et al. 2009). Alternatively, if the rock that was removed was mostly a Mesozoic to Paleogene sedimentary cover, the amount of rock removed above Hardangervidda could have been as little as 0.4 km. We consider the former scenario is more realistic (Fig. 9), but in either case, a significant thickness of rock must have been removed following the onset of Miocene exhumation, resulting in creation of the planation surface that now forms Hardangervidda.

Fig. 9.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 9.

Schematic illustration of the burial and exhumation history of Southern Scandes along an onshore–offshore profile reaching from the eastern North Sea to Hardangervidda (location shown in Fig. 1). (a) Present-day: the interpretation of the seismic line in Figure 2 is shown depth-converted on the left and a schematic representation of the onshore landscapes (Fig. 4) and geology on the right. Red dashed line: the ‘early Miocene surface’ marks the top of the rock column that was present above the present-day surface prior to early Miocene exhumation estimated from the palaeotemperature constraints from AFTA data onshore (this study) and offshore (Japsen et al. 2010). Green dashed line: the tilted Mesozoic peneplain with a hilly relief below the Jurassic sediments on the seismic line and on the sloping surface SW of Hardangervidda. Blue dashed line: the subhorizontal Hardangervidda surface. (b) Mid- to late Miocene: formation of the Hardangervidda planation surface as part of an extensive peneplain (blue dashed line) after early Miocene uplift leading to erosion by river incision to a base level interpreted as adjacent sea-level. Hardangervidda was then uplifted to its present elevation of c. 1200 m a.s.l. as shown in (a). This profile is produced by removing 800 m from the top of the profile in (c) in agreement with estimates of removed covers owing to early Miocene exhumation for Hardangervidda and for wells offshore (Japsen et al. 2010). (c) Early Miocene: maximum Cenozoic burial following rifting in the Late Jurassic and subsequent post-rift subsidence and burial. This profile is produced by flattening the early Miocene surface in profile in (a) to early Miocene palaeo-sea-level. It should be noted that a rock column of about 800 m covered Hardangervidda, interpreted here as Caledonian rocks, and that a column of Jurassic and younger sediments covered the present-day land surface from the coast to the western margin of Hardangervidda. Jurassic sediments extended at least as far east as the present-day coast (Fossen et al. 1997), and we have chosen to show them as onlapping the basement SW of the Folgefonni Glacier, implying that the etch surface at present-day higher elevations was produced during the Cretaceous. FF, Folgefonni Glacier. AFTA data from Stavanger (ST), LF (Lysefjord) and R (Rjukan).

Where the Mesozoic peneplain is exposed onshore today (Fig. 1c), our results suggest that a protective cover of Upper Jurassic to Oligocene sediment was present prior to the onset of early Miocene exhumation, similar to what is preserved offshore today (Fig. 2). In this case, a significant thermal conductivity contrast is likely between basement and the removed cover, in which case linear extrapolation of the palaeotemperatures to higher elevations, as discussed above, would not be valid. Whereas linear extrapolation of the profile for coastal locations in Figure 7 implies a rock column extending to 1.7 km above present-day sea-level, the arguments above suggest that as little as c. 1 km of overburden may have been present. However, we have adopted a thickness of the sedimentary cover on the coast of 1.3 km, based on a palaeogeothermal gradient of 35°C km−1 through the removed section, which corresponds to the upper limit for present-day values of the geothermal gradient for Mesozoic–Cenozoic sediments in wells in quadrant 17 (Fig. 1a; Japsen et al. 2010).

The early Miocene tectonic phase had great impact on the region. It produced a pronounced regional unconformity in the sedimentary basins offshore the margin of NW Europe (Stoker et al. 2005) and caused the removal of a sedimentary cover of 750–1000 m SW and south of Norway (Fig. 10; Japsen et al. 2007, 2010). In the eastern North Sea Basin, coarse-grained, braided river systems from Norway reached western Denmark at c. 22 Ma, followed by deltas that prograded towards the SSW (Rasmussen 2014). The main source area for these lower Miocene sand deposits was within southern Norway and southwestern Sweden (Olivarius et al. 2014). Olivarius et al. (2014) also found that the high maturity of the heavy-mineral assemblages of the Miocene sand, a relatively large zircon content and paucity of feldspars, indicates weathering and kaolinization in the provenance area. It could also indicate that many of the minerals were from redeposited sediments, a suggestion supported by the presence of Archean-age zircons whose nearest source area is in Finland (Olivarius et al. 2014). We suggest that the sources for the lower Miocene sands that reached the Danish Basin from Norway consisted of both sedimentary covers as well as the deeply weathered bedrock of the Mesozoic peneplain that was re-exposed after early Miocene exhumation.

Fig. 10.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 10.

Outline of southern Scandinavia with evidence for two phases of Neogene uplift beginning between 23 and 21 Ma and at c. 4 Ma. Neogene uplift began in the early Miocene within and north of the Sorgenfrei–Tornquist Zone. However, Neogene uplift began in the Pliocene south of the zone, where Miocene sediments had accumulated in the Danish Basin. Sources: early Miocene uplift of southernmost Norway, this study; offshore SW Norway, Stoker et al. (2005) and Japsen et al. (2010); southern Sweden, Japsen et al. (2016); the Sorgenfrei–Tornquist Zone, Japsen et al. (2007, 2016); early Pliocene uplift of southernmost Norway, this study; offshore SW Norway, Stoker et al. (2005) and Japsen et al. (2010); southern Sweden, Japsen et al. (2016); Danish Basin, Japsen et al. (2007).

Neogene uplift began in the early Miocene within and north of the Sorgenfrei–Tornquist Zone (Fig. 10). However, Neogene uplift began in the Pliocene south of the zone, where Miocene sediments had accumulated in the Danish Basin (Japsen et al. 2007; Rasmussen et al. 2008). It thus seems likely that the Sorgenfrei–Tornquist Zone acted as a hinge zone during the early Miocene between the uplifting Baltic Shield and the subsiding Danish Basin.

Reconstruction of the late Cenozoic development of the southernmost Scandes

Figure 9c illustrates our interpretation of the situation prior to early Miocene exhumation and the subsequent formation of Hardangervidda. As discussed above, a significant thickness of rock extended above the present-day level of Hardangervidda, and more than 1 km of Jurassic–Oligocene sediments covered the present-day coast prior to early Miocene exhumation. The presence of a Mesozoic peneplain characterized by hilly relief and joint-aligned valleys from the coast to elevations of about 1000 m a.s.l. indicates the extent of the Jurassic–Cretaceous cover (Figs 1c and 3). Prior to the onset of exhumation, the early Miocene surface (red dashed line in Fig. 9c) thus extended across Paleogene, Mesozoic and Paleozoic rocks, and it may have been a depositional surface as shown in the figure. The result of the exhumation was thus a Miocene peneplain (dashed blue line in Fig. 9b) that extended, not only across Hardangervidda, but also across a sedimentary cover that preserved the sub-Mesozoic relief on basement rocks underlying the sedimentary section. The focus of this reconstruction is the formation of Hardangervidda (which mainly corresponds to Level III of Lidmar-Bergström et al. 2000) as the result of early Miocene exhumation; the lower Level IV (the bottom level of the Palaeic Relief) must have formed in response to a later phase of uplift and erosion.

The low-relief Hardangervidda plateau cuts across bedrock of different age and resistance (Fig. 4a3), and as documented here, it is a Miocene erosion surface, which must have been graded to a distinct base level. Following the arguments presented by Japsen et al. (2009), we suggest that the parsimonious (Occam's razor) explanation is, in the absence of other options, that base level was the level of the adjacent sea at the time. Where a study area is known to have been near the sea at the time in question, and there is no resistant level in or at the base of a sedimentary sequence, this is the obvious explanation. In the case of Hardangervidda, it is clear that the base level in the Miocene was the level of the adjacent Atlantic Ocean. We thus conclude that the low-relief Hardangervidda plateau represents a remnant of a peneplain graded to sea-level following the onset of early Miocene exhumation (Fig. 9b), and that the plateau consequently reached its present elevation of c. 1200 m a.s.l. after its formation during the Miocene (Fig. 9a). Removal of the Mesozoic–Paleogene cover as a result of this last phase of uplift led to re-exposure of both the Mesozoic peneplain on the western flank of the Southern Scandes and the Sub-Cambrian Surface over parts of Hardangervidda. The preservation of the remnants of these relief types to the present day indicates that the re-exposure of the Mesozoic peneplain must be geologically very recent.

Several lines of evidence indicate that the main phase of uplift of Hardangervidda (and thus of the Palaeic Relief as such) took place in the early Pliocene. Stoker et al. (2005) used stratigraphic arguments to show that there are two Neogene unconformities within the sedimentary succession along the NW European margin; a base Miocene and an intra-Pliocene unconformity (c. 23 and c. 4 Ma, respectively). Stoker et al. (2005) used the patterns of contourite sedimentation coeval with the onset of rapid seaward progradation along the NW European continental margin to show that the lower Pliocene–Holocene succession there records a regional change at about 4 Ma. Stoker et al. (2005) inferred that this progradation reflects a marked increase in sediment supply in response to uplift and tilting of the continental margin. Although glacial sediments form a major component of the prograding wedges, Stoker et al. (2005) used stratigraphic data to show that the onset of progradation predated significant high-latitude glaciation by c. 1 myr and extensive northern hemisphere glaciation by at least 3 myr. The incision of the deep valleys of the Southern Scandes below Level IV (800–1200 m a.s.l.) further supports that the dominant uplift of the Palaeic Relief, including Hardangervidda, took place in the late Cenozoic (Lidmar-Bergström et al. 2013). The formation and uplift of the lower levels of the Palaeic Relief of the Southern Scandes thus occurred in a similar fashion and synchronous with the formation and uplift of the South Småland Peneplain across the South Swedish Dome (Lidmar-Bergström et al. 2013; Japsen et al. 2016).

Several other researchers have used stratigraphic arguments to show that a phase of uplift and erosion affected southern Norway and the adjacent sedimentary basins in the late Cenozoic; for example, the eastern North Sea Basin, the Skagerrak and the Danish Basin (Jensen & Schmidt 1992; Hansen 1996; Riis 1996; Japsen 1998; Japsen & Chalmers 2000; Japsen et al. 2002). Japsen et al. (2002) argued that this phase took place in the Pliocene because this timing matches a regional erosional unconformity that separates upper Miocene strata from Plio-Pleistocene strata in southwestern Denmark.

Another line of evidence for a late Neogene tectonic phase in Scandinavia is based on sedimentological arguments; for example, Løseth et al. (2017) used sequence-stratigraphic techniques to show that there had been 500 m of uplift offshore mid-Norway at the Miocene–Pliocene transition. The central North Sea Basin experienced pronounced subsidence in the early Pliocene, renewed progradation from Scandinavia and a final, major westward tilting of the eastern North Sea Basin at the Pliocene–Quaternary transition (Rasmussen 2005). As a counterpart to this subsidence, the Scandinavian mainland was uplifted in the late Neogene as illustrated by the progressive truncation of Neogene and Jurassic strata towards Norway (Figs 1 and 2).

Finally, thermochronological data from the Danish Basin provide evidence for a phase of Pliocene uplift and erosion. Japsen et al. (2007) interpreted AFTA, VR and sonic data from four wells there to show that 450–850 m of mainly Cenozoic sediments were removed by erosion during a Pliocene event that was dated to c. 4 Ma (Japsen et al. 2007). However, Pliocene cooling cannot be resolved in the AFTA data from southern Norway presented here, because of the relatively low palaeotemperatures involved.

Discussion

Our investigations have defined a sequence of events that integrates the new results from AFTA with the stratigraphic landscape analysis presented by Lidmar-Bergström et al. (2000, 2013). The new data allow confirmation of the timing and magnitude of Middle Triassic and Middle Jurassic denudation defined by previous workers (e.g. Rohrman et al. 1995), which ultimately led to formation of the Mesozoic peneplain of which remnants are preserved at elevations up to over 1000 m a.s.l. across southwestern Norway. Our results also define an early Miocene uplift event that led to the formation of Hardangervidda by erosion to the base level of the adjacent ocean.

Lower Paleozoic sediments and Caledonian nappes are present across southern Scandinavia, but the results presented here and those of Japsen et al. (2016) show that significant additional thicknesses of sediment accumulated across the region in separate episodes in post-Caledonian times; during the Devonian–Early Triassic, Late Triassic–Early Jurassic and Late Jurassic–Oligocene. However, all of these covers (apart from the Jurassic outlier near Bergen) were removed from the study area during three episodes of exhumation prior to deposition of the later sequences. Devonian–Lower Triassic sediments were probably all removed prior to deposition of a Upper Triassic–Lower Jurassic sequence, which in turn was removed prior to deposition of Upper Jurassic–Oligocene sediments. The latter have been removed as a result of Neogene uplift of the Southern Scandes.

We interpret the Middle Triassic and Middle Jurassic episodes of exhumation of southernmost Norway and southern Sweden as the result of epeirogenic uplifts accompanying fragmentation of Pangaea, caused by accumulation of mantle heat beneath the supercontinent (Nance et al. 2014). A similar explanation was invoked by Japsen et al. (2016), who found that these events (and a late Carboniferous event that is not identified in southern Norway) affected southern Sweden and wide areas beyond the Baltic Shield (Veevers 2004).

The results of this study suggest that the major components of the landscape of the southernmost Scandes are (1) a tilted Mesozoic peneplain rising from the south-west coast to altitudes around 1000 m a.s.l., (2) the sub-horizontal, four-stepped Palaeic Relief (including the Miocene Hardangervidda) which truncates the Mesozoic peneplain and (3) incised valleys, dissecting the Mesozoic peneplain and the Palaeic Relief (Fig. 3). This result is in agreement with the studies of Riis (1996) and Lidmar-Bergström et al. (2000, 2013), but conflicts with the often-made claim (e.g. Steer et al. 2012; Ksienzyk et al. 2014; Pedersen et al. 2016, 2018) that the subhorizontal Palaeic Relief is Mesozoic in age.

Fredin et al. (2017) interpreted their documentation of a relict, Mesozoic landscape at Bømlo in terms of relative tectonic stability of the entire strandflat. They did not, however, take into account that their study concerned only that locality whereas the geomorphological analysis by Holtedahl (1998) showed that the strandflat is a Plio-Pleistocene feature that coincides with an exhumed pre-Jurassic surface in some coastal areas. Fredin et al. (2017) also argued that the landscape could not have been very deeply buried because diagenetic alteration products would have formed and potentially been preserved. They cited low-temperature thermochronology studies indicating that heating could not have exceeded 50°C, which they thought corresponded to burial depths less than 1 km. However, the AFTA data from coastal locations and the new VR data from near Bergen suggest that Jurassic sediments and underlying basement reached palaeotemperatures of 60°C or more, corresponding to burial of the order of 1 km. Significantly, Fredin et al. (2017) reported an absence of diagenetic alteration products in basement buried by almost 2 km of sedimentary cover in two wells on the Utsira High (Fig. 1), negating their argument regarding observations from Bømlo. The age of the weathering at Bømlo and the age of the Jurassic outlier in Bergen thus represent single points on the Mesozoic peneplain that extends from about 1 km a.s.l. to depths of several kilometres below sea-level, and thus do not indicate tectonic stability since the Mesozoic.

Mechanisms

Episodic burial and exhumation

The Norwegian margin is in many aspects a typical example of an elevated passive continental margin (EPCM) with a Mesozoic rift system parallel to the coast, a transition from continental to oceanic crust further offshore and elevated plateaux cut by deeply incised valleys, separated from a coastal plain (the strandflat) by one or more escarpments (Japsen et al., 2012a). In Norway, as at other EPCMs, these plateaux formed by erosion to the base level of the adjacent sea long after break-up and were then raised to their present elevations (Fig. 9). Furthermore, sediments deposited horizontally in the post-rift section now dip seaward and are truncated by post-rift erosional unconformities, and post-rift sedimentary systems prograde away from the margin (Figs 1 and 2). The many similarities between the Norwegian EPCM and EPCMs elsewhere in the world (e.g. in Brazil, southern Africa and SE Australia; Lidmar-Bergström et al. 2000; Japsen et al. 2012b; Green et al. 2013, 2018), lead us to suggest that the formation of the Norwegian margin is controlled by processes common to all EPCMs, and that these processes are probably plate-scale.

The processes responsible for the Neogene rise of the Southern Scandes are, however, not understood. That the controlling processes are tectonic rather than climatic is demonstrated by the separation between early and late Neogene onset of exhumation in southern Scandinavia by the Sorgenfrei–Tornquist Zone (Fig. 10), and by the fact that the major Neogene unconformities along the NW European margin predate the onset of widespread glaciations (Stoker et al. 2005). Rasmussen (2014) highlighted the tectonic origin of the changes that led to the formation of the early Miocene, braided fluvial system from the Southern Scandes into the Danish Basin. This system was capable of transporting clasts with a diameter of up to 4 cm several hundred kilometres into the basin, indicating a high-energy fluvial system and hence uplift of the source area.

Vertical motion forming EPCMs in general and the Southern Scandes in particular may be due either to sub-lithospheric processes (e.g. dynamic support from the asthenosphere; Pekeris 1935; Griffiths & Campbell 1990; Lovell 2010; Colli et al. 2014), or to compressive stresses that build up during changes in plate motion (Cobbold et al. 2007; Cloetingh & Burov 2010). Such motion may in turn be amplified by a flexural, isostatic response to erosional unloading of the landmass and depositional loading of adjacent basins (Molnar & England 1990), but the initial uplift of peneplains formed by erosion to the base level of the adjacent sea requires a tectonic trigger.

We have shown that the southernmost Scandes formed as a result of three Mesozoic–Cenozoic episodes of burial and exhumation. The rocks that are now at the surface have thus been at very different elevations above and below sea-level during that time span, which is in agreement with the quantitative estimates of the palaeotopography for Norway south of 62°N by Sømme et al. (2013a,b). This complex development is reflected in the overall configuration of the Southern Scandes where subhorizontal Palaeic surfaces (including Hardangervidda) cut off a tilted, Mesozoic peneplain (Figs 2 and 3). Here we have shown that the main Palaeic surfaces in the southernmost Scandes formed as the result of Miocene denudation, and that a cover of Jurassic–Oligocene sediments on the flanks caused the early Miocene palaeotemperatures defined by AFTA and VR. It is thus clear that the Scandes are not the erosional remnants of the Caledonian Mountains (Nielsen et al. 2009; Pedersen et al. 2016, 2018; see also Lidmar-Bergström & Bonow, 2009; Chalmers et al., 2010; Gabrielsen et al. 2010).

Erosion and isostatic adjustment of the Caledonian Mountains

Nielsen et al. (2009) hypothesized that the topography of the Southern Scandes represents the combined effects of continued erosion since the Caledonian Orogeny as a result of climatic deterioration, magnified by isostasy supported by a crustal root under the Scandes. This argument was continued by Pedersen et al. (2016), who argued that the spatial offset between the small root observed by Stratford & Thybo (2011) and the highest topography could be accounted for by flexural rigidity of the crust. Chalmers et al. (2010) showed, however, that a root 2–3 times the size of that observed by Stratford & Thybo (2011) would be necessary to support the southern Scandes. In addition, no root whatsoever was observed under the Northern Scandes by England & Ebbing (2012), although a transition to a high-velocity lower crust was observed east of the Scandes. It is true that the low-lying topography to the east, under Sweden and Finland, is underlain by thicker crust and the deep crust probably contains a high proportion of high-density eclogite (Kozlovskaya et al. 2004). Eclogite is also known from exhumed lower crust near Sognefjord in west Norway (Milnes et al. 1997), so there is likely to be eclogite in unknown proportions under all the Scandinavian crust, and, as the distribution of eclogite is unknown, arguments based on isostasy are inconclusive. Pedersen et al. (2016) did not consider that dynamic topography is reflected in the gravity signal in a complicated, wavelength-dependent manner (Colli et al. 2016), and they consequently arrived at models that are at odds with observations available in the literature and with those presented in this study.

The ideas presented by Nielsen et al. (2009) have also been criticized on a number of other grounds including geomorphological, geophysical, thermochronological, sedimentological and basic geological arguments (Lidmar-Bergström & Bonow 2009; Chalmers et al. 2010; Gabrielsen et al. 2010). We therefore do not think that arguments based on isostasy can be used to come to any definitive conclusion about the presence of the Norwegian mountains, nor any other mountains near so-called ‘passive’ margins. We need to look for other evidence, and we consider that this paper is a contribution to such evidence.

More recently, Pedersen et al. (2018) used inverse landscape evolution modelling in an attempt to constrain the topographic evolution of Scandinavia between 54 and 4 Ma, based on sediment volumes west of southern Norway and reconstructed pre-glacial topography. Their preferred model involved the presence of high (2 km) topography since 54 Ma. However, those researchers failed to consider several simple observations that contradict their conclusions and render their modelling exercise meaningless, including the following. (1) The maturity of Jurassic sediments reflects that they have been buried by a significant thickness of cover (Fossen et al. 1997), as supported by the erosional truncation of Jurassic and younger sequences towards the coast of Norway. (2) Major Paleogene and Miocene sediment accumulations occur in the Danish Basin are not included in their calculations, and notably uplift and erosion in the Scandinavian hinterland resulted in progradation of major delta systems into the basin in the earliest Miocene (Rasmussen 2014, 2017). (3) The thick Eocene sediments included in their volume calculations are primarily derived from the Shetland Platform, not from the Scandes (Ahmadi et al. 2003; Knox et al. 2010). (4) Significant volumes of sediment were removed by Neogene erosion west and south of Norway (Jensen & Schmidt 1992; Hansen 1996; Japsen et al. 2007, 2010), and unknown volumes of sediment may have been deposited east of Norway and subsequently removed. (5) The Palaeic Relief of the Southern Scandes consists of Cenozoic erosion surfaces in four steps with a vertical separation of 1500 m (Lidmar-Bergström et al. 2000, 2013), and thus the traditional interpretation of the topography of Norway as ‘remnants of a peneplain eroded to sea-level in the Mesozoic’ is incorrect (Pedersen et al. 2018).

Erosion and isostatic adjustment of Mesozoic rift shoulders

It was argued (Osmundsen & Redfield 2011; Redfield & Osmundsen 2013) that the post-rift evolution of the Scandes was driven by flexed isostatic adjustments to rift shoulders formed during Late Jurassic and Early Cretaceous rifting. They argued that accommodation of sediments above the rifted basins offshore was driven by thermal cooling, and that continued erosion of the rift-margin uplifts resulted in isostatic flexural arching of the footwall area. The uplift was sufficient that the rift-margin faults moved repeatedly, as they continue to do. Redfield & Osmundsen (2013) argued that this mechanism has acted several times in distinct episodes since the Early Cretaceous and that the stepped topography identified by Lidmar-Bergström et al. (2000) is the result of each distinct episode, but they did not, however, provide a mechanism for the episodic nature of the uplifts.

The interpretation by Redfield & Osmundsen (2013) may explain some of the effects affecting the area within and on the hanging wall of the Møre–Trøndelag Fault Complex. It is inadequate, however, to explain the uplift of the whole of the Southern Scandes. As Chalmers et al. (2010) pointed out, the width of footwall uplift of a crust with an elastic thickness of about 30 km is about 70 km, not the 300 km or more width of the Southern Scandes. Medvedev & Hartz (2014)  also showed that the flexed isostatic response of erosion of the deep valleys is of the order of 800 m, which is inadequate to explain all the modern topography.

Our AFTA data also show that post-rift subsidence and burial is not restricted to the present-day offshore domain but extended into the region that is onshore Norway today. This is demonstrated by high, early Miocene palaeotemperatures of rocks along the coastal zone where Upper Jurassic sediments rest on basement (Figs 1a and 2). The present-day, seaward facing escarpment along the west coast of the Southern Scandes is thus not a modified, Mesozoic rift shoulder, as suggested by Redfield & Osmundsen (2013). It is, instead, a Neogene feature and the remnant landscape on it (hilly relief) was formed prior to burial below late Mesozoic sediments and subsequently exhumed. This conclusion is supported by the quantitative estimates of Sømme et al. (2013b), who found that the land surface along the western margin of southern Norway did not exceed 200 m a.s.l. during the early Eocene. In fact, much of southern Scandinavia may have been below sea-level in the Eocene (Knox et al. 2010).

Dynamic topography

Rickers et al. (2013) used full waveform seismic tomography to map a low-velocity layer extending from the Iceland and Jan Mayen hotspots to beneath the continental lithosphere of the Southern Scandes, the Danish Basin, part of the British Isles and eastern Greenland, and they noted that these regions experienced uplift in Neogene times. Schoonman et al. (2017)  argued that this spatial correlation between the low-velocity layer and uplifted regions indicates dynamic support by low-density asthenosphere. However, they did not comment on the apparent absence of low-density asthenosphere under, for example, West Greenland and the Barents Sea where Neogene uplift is well documented (Bonow et al. 2006; Japsen et al. 2006; Green & Duddy 2010). Japsen et al. (2014) argued that the very high elevations in the parts of East Greenland nearest Iceland (in contrast to the margin further north and south as well as West Greenland) were the result of Pliocene uplift owing to dynamic support from the adjacent Iceland Plume superimposed on whatever process was causing the other EPCMs around Greenland. Thus, although we find it possible that dynamic support by low-density asthenosphere originating from the Iceland Plume may have contributed to the Pliocene uplift of the Southern Scandes, we also want to point out that the presence of similar EPCMs in, for example, eastern Australia and eastern Brazil cannot be explained by influence from nearby plumes.

The magnitude of the late Cenozoic (Pliocene) uplift phase in southern Scandinavia declines rapidly away from the Atlantic margin (Lidmar-Bergström et al. 2013): from 800–1200 m in the Southern Scandes to 150 m for the South Swedish Dome. This is in contrast to the similar amounts of section removed during early Miocene exhumation across southern Scandinavia, as illustrated by similar early Miocene palaeotemperatures from southernmost Norway to southern Sweden in samples at elevations close to sea-level (Japsen et al. 2016). In contrast to Scandinavia, Neogene uplift of the margins of East and West Greenland did not begin in the early Miocene, but in the late Miocene (c. 10 Ma; Japsen et al. 2006, 2014). Because the processes causing these vertical motions are not understood, it is not clear if the differences in magnitude and timing of major episodes in different areas are due to different processes or are controlled by varying responses of the lithosphere in different areas. Further insights into these questions must await improved insights of the underlying geodynamic processes.

Conclusions

Our analysis of new AFTA data from southernmost Norway has revealed three episodes of cooling caused by denudation, beginning in the Middle Triassic, Middle Jurassic and early Miocene.

The palaeotemperatures inferred from the AFTA data imply that a kilometres-thick succession of Paleozoic–Lower Triassic rocks covered the study area prior to Middle Triassic exhumation that resulted in basement exposure in the present-day coastal regions (Fredin et al. 2017). Furthermore, a thick cover of Upper Triassic–Lower Jurassic sediment had accumulated onto the basement prior to Middle Jurassic exhumation.

The palaeotemperature constraints for the early Miocene episode show that the rock column that was present at that time extended well above all of the present-day landscape within the study area. This leads us to the key conclusion that Hardangervidda formed during the Miocene by erosion to the base level of the adjacent ocean. The early Miocene uplift and erosion caused a distinct change in the depositional environment in the eastern North Sea Basin and to formation of coarse-grained, braided river systems flowing southwards from Norway (Rasmussen 2014). Hardangervidda reached its present elevation of c. 1200 m a.s.l. after its formation, and various lines of evidence indicate that the final phase of uplift of Hardangervidda and of the other steps in the Palaeic Relief of the Southern Scandes started in the early Pliocene.

An earliest Oligocene phase of uplift and erosion in the Scandinavian hinterland is documented by the presence of prograding clastic wedges south of Norway (e.g. Faleide et al. 2002), but the AFTA data do not resolve cooling related to this event. Nor do they resolve cooling events following the early Miocene phase because of the relatively low paleotemperatures of outcrop samples prior to the final phase of exhumation.

The contrast between the hilly relief on the flanks of the Scandes and the plains of the Palaeic Relief supports our conclusions. The hilly relief formed on a Mesozoic peneplain by chemical weathering in basement fractures; offshore it is preserved below a cover of Mesozoic and younger sediments, and onshore remnants of the Mesozoic peneplain extend from the coast to elevations of about 1000 m a.s.l. The presence of this denudation surface onshore implies that (1) there was a protective cover of Jurassic–Cretaceous sediments over that ground, (2) Hardangervidda must have formed as part of a Miocene peneplain that also extended across that sedimentary cover and (3) the re-exposure of the Mesozoic peneplain followed Pliocene uplift and erosion. Our AFTA data, supported by new VR data from the Upper Jurassic outlier near Bergen, also show that at least 1 km of Upper Jurassic to Oligocene sediments had accumulated above the basement rocks that are now exposed along the coast prior to early Miocene exhumation.

The Southern Scandes resemble elevated passive continental margins around the world with young landscapes dominated by high plateaux cut by deeply incised valleys adjacent to sedimentary basins where erosional unconformities truncate post-rift sequences tilted away from the landmass. We have shown here that, as on other margins, these characteristic features of the Southern Scandes formed through events of episodic burial and exhumation long after rifting and breakup in the adjacent basins. Identification of the mechanisms driving the uplifts of passive continental margins in general and of the Southern Scandes in particular, awaits geodynamic modelling constrained by observations such as those presented in this study.

Acknowledgements

We thank K. Lidmar-Bergström and reviewers A. Hall and C. Leighton as well as one anonymous reviewer for their very constructive comments that improved our paper significantly. Published with the permission of the Geological Survey of Denmark and Greenland (GEUS).

Funding

This research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

Scientific editing by Andrew Carter

  • © 2018 The Author(s). Published by The Geological Society of London. All rights reserved

References

  1. ↵
    1. Ahlmann, H.W.
    1919. Geomorphological studies in Norway. Geografiska Annaler, 1, 3–320.
    OpenUrl
  2. ↵
    1. Ahmadi, Z.,
    2. Sawyers, M.,
    3. Kenyon-Roberts, S.,
    4. Stanworth, B.,
    5. Kugler, K.,
    6. Kristensen, J. &
    7. Fugelli, E.
    2003. Paleocene. In: Evans, D., Graham, C., Armour, A. & Bathurst, P. (eds) The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea. Geological Society, London, 235–259.
  3. ↵
    1. Andersen, T.B.,
    2. Torsvik, T.H.,
    3. Eide, E.A.,
    4. Osmundsen, P.T. &
    5. Faleide, J.I.
    1999. Permian and Mesozoic extensional faulting within the Caledonides of central south Norway. Journal of the Geological Society, London, 156, 1073–1080, https://doi.org/10.1144/gsjgs.156.6.1073
    OpenUrl
  4. ↵
    1. Andersen, J.L.,
    2. Egholm, D.L. et al.
    2018. Widespread erosion on high plateaus during recent glaciations in Scandinavia. Nature Communications, 9, 830.
    OpenUrl
  5. ↵
    1. Andriessen, P.A.M.
    1990. Anomalous fission-track apatite ages of the Precambrian basement in the Hunnedalen region, south-western Norway. Nuclear Tracks and Radiation Measurements, 17, 285–291.
    OpenUrl
  6. ↵
    1. Baranwal, V.C.,
    2. Olesen, O. &
    3. Rønning, J.S.
    2016. Action map for tunnel planning, Oslofjord – Telemark region: Mapping of deeply weathered weakness zones. NGU Report, 2016.015.
  7. ↵
    1. Bertelsen, F.
    1980. Lithostratigraphy and depositional history of the Danish Triassic. Geological Survey of Denmark Series B, 4. Geological Survey of Denmark, Copenhagen, 1–59.
  8. ↵
    1. Bøe, R.,
    2. Fossen, H. &
    3. Smelror, M.
    2010. Mesozoic sediments and structures onshore Norway and in the coastal zone. Norwegian Journal of Geology, 450, 15–32.
    OpenUrl
  9. ↵
    1. Bonow, J.M.,
    2. Lidmar-Bergström, K. &
    3. Näslund, J.O.
    2003. Palaeosurfaces and major valleys in the area of the Kjølen Mountains, southern Norway – consequences of uplift and climate change. Norsk Geografisk Tidsskrift, 57, 83–101.
    OpenUrlCrossRef
  10. ↵
    1. Bonow, J.M.,
    2. Japsen, P.,
    3. Lidmar-Bergström, K.,
    4. Chalmers, J.A. &
    5. Pedersen, A.K.
    2006. Cenozoic uplift of Nuussuaq and Disko, West Greenland – elevated erosion surfaces as uplift markers of a passive margin. Geomorphology, 80, 325–337.
    OpenUrlCrossRefWeb of Science
  11. ↵
    1. Bonow, J.M.,
    2. Japsen, P.,
    3. Green, P.F.,
    4. Cobbold, P.R.,
    5. Pedreira, A.J.,
    6. Lilletveit, R. &
    7. Chiossi, D.
    2009. Post-rift landscape development of north-east Brazil. Geological Survey of Denmark and Greenland Bulletin, 17, 81–84.
    OpenUrl
  12. ↵
    1. Büdel, J.
    1963. Die pliozänen und quartären Pluvialzeiten der Sahara. E&G – Quaternary Science Journal, 14, 161–187.
    OpenUrl
  13. ↵
    1. Burnham, A.K. &
    2. Sweeney, J.J.
    1989. A chemical kinetic model of vitrinite reflectance maturation. Geochimica et Cosmochimica Acta, 53, 2649–2657.
    OpenUrlCrossRefWeb of Science
  14. ↵
    1. Calner, M.,
    2. Ahlberg, P.,
    3. Lehnert, O. &
    4. Erlström, M.
    2013. The Lower Palaeozoic of Southern Sweden and the Oslo Region, Norway. Field Guide for the 3rd Annual Meeting of the IGCP Project 591, Sveriges Geologiska Undersökning, Palaeozoic of southern Sweden and the Oslo Region, Norway. Sveriges Geologiska Undersökning Rapporter och Meddelanden, 133.
  15. ↵
    1. Cederbom, C.
    2001. Phanerozoic, pre-Cretaceous thermotectonic events in southern Sweden revealed by fission track thermochronology. Earth and Planetary Science Letters, 188, 199–209.
    OpenUrlCrossRefWeb of Science
  16. ↵
    1. Cederbom, C.,
    2. Larson, S.Å.,
    3. Tullborg, E.L. &
    4. Stiberg, J.P.
    2000. Fission track thermochronology applied to Phanerozoic thermotectonic events in central and southern Sweden. Tectonophysics, 316, 153–167.
    OpenUrlCrossRefWeb of Science
  17. ↵
    1. Chalmers, J.A.,
    2. Green, P.,
    3. Japsen, P. &
    4. Rasmussen, E.S.
    2010. The Scandinavian mountains have not persisted since the Caledonian orogeny. A comment on Nielsen et al. (2009a). Journal of Geodynamics, 50, 94–101.
    OpenUrlCrossRefWeb of Science
  18. ↵
    1. Clausen, O.R.,
    2. Gregersen, U.,
    3. Michelsen, O. &
    4. Sørensen, J.C.
    1999. Factors controlling the Cenozoic sequence development in the eastern parts of the North Sea. Journal of the Geological Society, London, 156, 809–816, https://doi.org/10.1144/gsjgs.156.4.0809
    OpenUrl
  19. ↵
    1. Clauser, C. &
    2. Huenges, E.
    1995. Thermal conductivity of rocks and minerals. In: Handbook of Physical Constants. AGU Reference Shelf, 3, 105–126.
    OpenUrl
  20. ↵
    1. Cloetingh, S. &
    2. Burov, E.
    2010. Lithospheric folding and sedimentary basin evolution: a review and analysis of formation mechanisms. Basin Research, 23, 1–34.
    OpenUrl
  21. ↵
    1. Cobbold, P.R.,
    2. Rossello, E.A.,
    3. Roperch, P.,
    4. Arriagada, C.,
    5. Gómez, L.A. &
    6. Lima, C.
    2007. Distribution, timing, and causes of Andean deformation across South America. In: Ries, A.C., Butler, A.W.H. & Graham, A.H. (eds) Deformation of the Continental Crust: The Legacy of Mike Coward. Geological Society, London, Special Publications, 272, 321–343, https://doi.org/10.1144/GSL.SP.2007.272.01.17
    OpenUrlAbstract/FREE Full Text
  22. ↵
    1. Colli, L.,
    2. Stotz, I. et al.
    2014. Rapid South Atlantic spreading changes and coeval vertical motion in surrounding continents: evidence for temporal changes of pressure-driven upper mantle flow. Tectonics, 33, 1304–1321.
    OpenUrlCrossRef
  23. ↵
    1. Colli, L.,
    2. Ghelichkhan, S. &
    3. Bunge, H.P.
    2016. On the ratio of dynamic topography and gravity anomalies in a dynamic Earth. Geophysical Research Letters, 43, 2510–2516, https://doi.org/10.1002/2016GL067929
    OpenUrl
  24. ↵
    1. Davids, C.,
    2. Wemmer, K.,
    3. Zwingmann, H.,
    4. Kohlmann, F.,
    5. Jacobs, J. &
    6. Bergh, S.G.
    2013. K–Ar illite and apatite fission track constraints on brittle faulting and the evolution of the northern Norwegian passive margin. Tectonophysics, 608, 196–211.
    OpenUrlCrossRefWeb of Science
  25. ↵
    1. Egholm, D.L.,
    2. Jansen, J.D.,
    3. Brædstrup, C.F.,
    4. Pedersen, V.K.,
    5. Andersen, J.L.,
    6. Ugelvig, S.V.,
    7. Larsen, N.K. &
    8. Knudsen, M.F.
    2017. Formation of plateau landscapes on glaciated continental margins. Nature Geoscience, 10, 592–597.
    OpenUrl
  26. ↵
    1. Eidvin, T.,
    2. Bugge, T. &
    3. Smelror, M.
    2007. The Molo Formation, deposited by coastal progradation on the inner Mid-Norwegian continental shelf, coeval with the Kai Formation to the west and the Utsira Formation in the North Sea. Norwegian Journal of Geology, 87, 75–142.
    OpenUrlWeb of Science
  27. ↵
    1. England, R. &
    2. Ebbing, J.
    2012. Crustal structure of central Norway and Sweden from integrated modelling of teleseismic receiver functions and the gravity anomaly. Geophysical Journal International, 191, 1–11.
    OpenUrlCrossRefWeb of Science
  28. ↵
    1. Eppelbaum, L.,
    2. Kutasov, I. &
    3. Pilchin, A.
    2014. Applied Geothermics. Springer, Berlin.
  29. ↵
    1. Faleide, J.I.,
    2. Kyrkjebø, R.,
    3. Kjennerud, T.,
    4. Gabrielsen, R.H.,
    5. Jordt, H.,
    6. Fanavoll, S. &
    7. Bjerke, M.
    2002. Tectonic impact on sedimentary processes during Cenozoic evolution of the northern North Sea and surrounding areas. In: Doré, A.G., Cartwright, J.A., Stoker, M.S., Turner, J. & White, N. (eds) Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration. Geological Society, London, Special Publications, 196, 235–269, https://doi.org/10.1144/GSL.SP.2002.196.01.14
    OpenUrlAbstract/FREE Full Text
  30. ↵
    1. Fauconnier, J.,
    2. Labrousse, L.,
    3. Andersen, T.B.,
    4. Beyssac, O.,
    5. Duprat-Oualid, S. &
    6. Yamato, P.
    2014. Thermal structure of a major crustal shear zone, the basal thrust in the Scandinavian Caledonides. Earth and Planetary Science Letters, 385, 162–171.
    OpenUrlCrossRefWeb of Science
  31. ↵
    1. Fossen, H.,
    2. Mangerud, G.,
    3. Hesthammer, J.,
    4. Bugge, T. &
    5. Gabrielsen, R.H.
    1997. The Bjorøy Formation: a newly discovered occurrence of Jurassic sediments in the Bergen Arc System. Norsk Geologisk Tidsskrift, 77, 269–287.
    OpenUrlWeb of Science
  32. ↵
    1. Fredin, O.,
    2. Viola, G. et al.
    2017. The inheritance of a Mesozoic landscape in western Scandinavia. Nature Communications, 8, 1–11, 14879, https://doi.org/10.1038/ncomms14879
    OpenUrl
  33. ↵
    1. Fyfe, A.,
    2. Gregersen, U. et al.
    2003. Oligocene to Holocene. In: Evans, D., Graham, C., Armour, A. & Bathurst, P. (eds) The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea. Geological Society, London, 279–287.
  34. ↵
    1. Gabrielsen, R.H.,
    2. Faleide, J.I.,
    3. Pascal, C.,
    4. Braathen, A.,
    5. Nystuen, J.P.,
    6. Etzelmüller, B. &
    7. O'Donnell, S.
    2010. Reply to discussion of Gabrielsen et al. (2010) by Nielsen et al. (this volume): Latest Caledonian to present tectonomorphological development of southern Norway. Marine and Petroleum Geology, 27, 1290–1295.
    OpenUrlCrossRefWeb of Science
  35. ↵
    1. Gabrielsen, R.H.,
    2. Nystuen, J.P.,
    3. Jarsve, E.M. &
    4. Lundmark, A.M.
    2015. The Sub-Cambrian Peneplain in southern Norway: its geological significance and its implications for post-Caledonian faulting, uplift and denudation. Journal of the Geological Society, London, 172, 777–791, https://doi.org/10.1144/jgs2014-154
    OpenUrlAbstract/FREE Full Text
  36. ↵
    1. Gee, D.G. &
    2. Sturt, B.
    (eds) 1985. The Caledonide orogen: Scandinavia and related areas, John Wiley and Sons, Chichester.
  37. ↵
    1. Gjessing, J.
    1967. Norway's paleic surface. Norsk Geografisk Tidsskrift, 21, 69–132.
    OpenUrlCrossRef
  38. ↵
    1. Glennie, K.W.,
    2. Higham, J. &
    3. Stemmerik, L.
    2003. Permian. In: Evans, D., Graham, C., Armour, A. & Bathurst, P. (eds) The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea. Geological Society, London, 91–103.
  39. ↵
    1. Goldsmith, P.,
    2. Hudson, G. &
    3. Van Veen, P.
    2003. Triassic. In: Evans, D., Graham, C., Armour, A. & Bathurst, P. (eds) The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea. Geological Society, London, 105–127.
  40. ↵
    1. Gradstein, F.M.,
    2. Ogg, J.G. &
    3. Hilgen, F.J.
    2012. On the Geologic Time Scale. Newsletters on Stratigraphy, 45, 171–188.
    OpenUrlCrossRefWeb of Science
  41. ↵
    1. Green, P.F. &
    2. Duddy, I.R.
    2010. Synchronous exhumation events around the Arctic including examples from Barents Sea and Alaska North Slope. In: Vining, B.A. & Pickering, S.C. (eds) Petroleum Geology: From Mature Basins to New Frontiers – Proceedings of the 7th Petroleum Geology Conference. Geological Society, London, 633–644, https://doi.org/10.1144/0070633
  42. ↵
    1. Green, P.F. &
    2. Duddy, I.R.
    2012. Thermal history reconstruction in sedimentary basins using apatite fission-track analysis and related techniques. In: Analyzing the Thermal History of Sedimentary Basins: Methods and Case Studies, SEPM Special Publication, 103, 65–104.
    OpenUrl
  43. ↵
    1. Green, P.F.,
    2. Lidmar-Bergström, K.,
    3. Japsen, P.,
    4. Bonow, J.M. &
    5. Chalmers, J.A.
    2013. Stratigraphic landscape analysis, thermochronology and the episodic development of elevated passive continental margins. Geological Survey of Denmark and Greenland Bulletin, 2013/30.
  44. ↵
    1. Green, P.F.,
    2. Japsen, P.,
    3. Chalmers, J.A.,
    4. Bonow, J.M. &
    5. Duddy, I.R.
    2018. Post-breakup burial and exhumation of passive continental margins: Seven propositions to inform geodynamic models. Gondwana Research, 53, 58–81, https://doi.org/10.1016/j.gr.2017.03.007
    OpenUrl
  45. ↵
    1. Gregersen, U.,
    2. Michelsen, O. &
    3. Sørensen, J.C.
    1998. Stratigraphy and facies distribution of the Utsira Formation and the Pliocene sequences in the northern North Sea. Marine and Petroleum Geology, 14, 893–914.
    OpenUrl
  46. ↵
    1. Griffiths, R.W. &
    2. Campbell, I.H.
    1990. Stirring and structure in mantle starting plumes. Earth and Planetary Science Letters, 99, 66–78.
    OpenUrlCrossRefWeb of Science
  47. ↵
    1. Grønlie, A.,
    2. Naeser, C.,
    3. Naeser, N.,
    4. Mitchell, J.,
    5. Sturt, B.A. &
    6. Ineson, P.
    1994. Fission-track and K–Ar dating of tectonic activity in a transect across the Møre–Trøndelag Fault Zone, central Norway. Norsk Geologisk Tidsskrift, 74, 24–34.
    OpenUrlWeb of Science
  48. ↵
    1. Hall, A.M.,
    2. Ebert, K.,
    3. Kleman, J.,
    4. Nesje, A. &
    5. Ottesen, D.
    2013. Selective glacial erosion on the Norwegian passive margin. Geology, 41, 1203–1206.
    OpenUrlAbstract/FREE Full Text
  49. ↵
    1. Haller, J.
    1971. Geology of the East Greenland Caledonides. Interscience, London.
  50. ↵
    1. Hansen, S.
    1996. Quantification of net uplift and erosion on the Norwegian Shelf south of 66°N from sonic transit times of shale. Norsk Geologisk Tidsskrift, 76, 245–252.
    OpenUrlWeb of Science
  51. ↵
    1. Heilmann-Clausen, C.,
    2. Nielsen, O.B. &
    3. Gersner, F.
    1985. Lithostratigraphy and depositional environments in the upper Paleocene and Eocene of Denmark. Bulletin of the Geological Society of Denmark, 33, 287–323.
    OpenUrl
  52. ↵
    1. Hendriks, B.,
    2. Andriessen, P.,
    3. Huigen, Y.,
    4. Leighton, C.,
    5. Redfield, T.,
    6. Murrell, G.,
    7. Gallagher, K. &
    8. Nielsen, S.B.
    2007. A fission track data compilation for Fennoscandia. Norwegian Journal of Geology, 87, 143–155.
    OpenUrlWeb of Science
  53. ↵
    1. Hendriks, B.,
    2. Osmundsen, P. &
    3. Redfield, T.
    2010. Normal faulting and block tilting in Lofoten and Vesterålen constrained by apatite fission track data. Tectonophysics, 485, 154–163.
    OpenUrlCrossRefWeb of Science
  54. ↵
    1. Holtedahl, H.
    1998. The Norwegian strandflat – a geomorphological puzzle. Norsk Geologisk Tidsskrift, 78, 47–66.
    OpenUrlWeb of Science
  55. ↵
    1. Huigen, Y. &
    2. Andriessen, P.
    2004. Thermal effects of Caledonian foreland basin formation, based on fission track analyses applied on basement rocks in central Sweden. Physics and Chemistry of the Earth, 29, 683–694.
    OpenUrlCrossRefWeb of Science
  56. ↵
    1. Ineson, P.,
    2. Mitchell, J. &
    3. Vokes, F.
    1975. K–Ar dating of epigenetic mineral deposits; an investigation of the Permian metallogenic province of the Oslo region, southern Norway. Economic Geology, 70, 1426–1436.
    OpenUrlAbstract/FREE Full Text
  57. ↵
    1. Japsen, P.
    1998. Regional velocity–depth anomalies, North Sea Chalk: a record of overpressure and Neogene uplift and erosion. AAPG Bulletin, 82, 2031–2074.
    OpenUrlAbstract
  58. ↵
    1. Japsen, P. &
    2. Chalmers, J.A.
    2000. Neogene uplift and tectonics around the North Atlantic: Overview. Global and Planetary Change, 24, 165–173.
    OpenUrlCrossRefWeb of Science
  59. ↵
    1. Japsen, P.,
    2. Bidstrup, T. &
    3. Lidmar-Bergström, K.
    2002. Neogene uplift and erosion of southern Scandinavia induced by the rise of the South Swedish Dome. In: Doré, A.G., Cartwright, J.A., Stoker, M.S., Turner, J.P. & White, N. (eds) Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration. Geological Society, London, Special Publications, 196, 183–207, https://doi.org/10.1144/GSL.SP.2002.196.01.12
    OpenUrlAbstract/FREE Full Text
  60. ↵
    1. Japsen, P.,
    2. Bonow, J.M.,
    3. Green, P.F.,
    4. Chalmers, J.A. &
    5. Lidmar-Bergström, K.
    2006. Elevated, passive continental margins: Long-term highs or Neogene uplifts? New evidence from West Greenland. Earth and Planetary Science Letters, 248, 315–324.
    OpenUrl
  61. ↵
    1. Japsen, P.,
    2. Green, P.F.,
    3. Nielsen, L.H.,
    4. Rasmussen, E.S. &
    5. Bidstrup, T.
    2007. Mesozoic–Cenozoic exhumation events in the eastern North Sea Basin: a multi-disciplinary study based on palaeothermal, palaeoburial, stratigraphic and seismic data. Basin Research, 19, 451–490.
    OpenUrlCrossRefWeb of Science
  62. ↵
    1. Japsen, P.,
    2. Bonow, J.M.,
    3. Green, P.F.,
    4. Chalmers, J.A. &
    5. Lidmar-Bergström, K.
    2009. Formation, uplift and dissection of planation surfaces at passive continental margins – a new approach. Earth Surface Processes and Landforms, 34, 683–699.
    OpenUrlCrossRefWeb of Science
  63. ↵
    1. Japsen, P.,
    2. Green, P.F.,
    3. Bonow, J.M.,
    4. Rasmussen, E.S.,
    5. Chalmers, J.A. &
    6. Kjennerud, T.
    2010. Episodic uplift and exhumation along North Atlantic passive margins: implications for hydrocarbon prospectivity. In: Vining, B.A. & Pickering, S.C. (eds) Petroleum Geology: From Mature Basins to New Frontiers – Proceedings of the 7th Petroleum Geology Conference. Geological Society, London, 979–1004, https://doi.org/10.1144/0070979
  64. ↵
    1. Japsen, P.,
    2. Chalmers, J.A.,
    3. Green, P.F. &
    4. Bonow, J.M.
    2012a. Elevated, passive continental margins: Not rift shoulders, but expressions of episodic, post-rift burial and exhumation. Global and Planetary Change, 90–91, 73–86.
    OpenUrl
  65. ↵
    1. Japsen, P.,
    2. Bonow, J.M.,
    3. Green, P.F.,
    4. Cobbold, P.R.,
    5. Chiossi, D.,
    6. Lilletveit, R.,
    7. Magnavita, L.P. &
    8. Pedreira, A.J.
    2012b. Episodic burial and exhumation history of NE Brazil after opening of the South Atlantic. Geological Society of America Bulletin, 124, 800–816.
    OpenUrlAbstract/FREE Full Text
  66. ↵
    1. Japsen, P.,
    2. Green, P.F.,
    3. Bonow, J.M.,
    4. Nielsen, T.F.D. &
    5. Chalmers, J.A.
    2014. From volcanic plains to glaciated peaks: Burial and exhumation history of southern East Greenland after opening of the NE Atlantic. Global and Planetary Change, 116, 91–114.
    OpenUrl
  67. ↵
    1. Japsen, P.,
    2. Green, P.F.,
    3. Bonow, J.M. &
    4. Erlström, M.
    2016. Episodic burial and exhumation of the southern Baltic Shield: Epeirogenic uplifts during and after break-up of Pangea. Gondwana Research, 35, 357–377.
    OpenUrl
  68. ↵
    1. Jarsve, E.,
    2. Eidvin, T.,
    3. Nystuen, J.,
    4. Faleide, J.,
    5. Gabrielsen, R. &
    6. Thyberg, B.
    2015. The Oligocene succession in the eastern North Sea: basin development and depositional systems. Geological Magazine, 152, 668–693.
    OpenUrlAbstract/FREE Full Text
  69. ↵
    1. Jensen, L.N. &
    2. Schmidt, B.J.
    1992. Late Tertiary uplift and erosion in the Skagerrak area; magnitude and consequences. Norsk Geologisk Tidsskrift, 72, 275–279.
    OpenUrlWeb of Science
  70. ↵
    1. Johannessen, K.C.,
    2. Kohlman, F.,
    3. Ksienzyk, A.K.,
    4. Dunkl, I. &
    5. Jacobs, J.
    2013. Tectonic evolution of the SW Norwegian passive margin based on low-temperature thermochronology from the innermost Hardangerfjord area. Norwegian Journal of Geology, 93, 243–260.
    OpenUrl
  71. ↵
    1. Kleman, J.
    2008. Where glaciers cut deep. Nature Geoscience, 1, 343–344.
    OpenUrl
  72. ↵
    1. Knox, R.,
    2. Bosch, A. et al.
    2010. Cenozoic. In: Doornenbal, H. & Stevenson, A. (eds) Petroleum Geological Atlas of the Southern Permian Basin Area. EAGE, Houten, 211–220.
  73. ↵
    1. Knudsen, T.L. &
    2. Fossen, H.
    2001. The Late Jurassic Bjorøy Formation: A provenance indicator for offshore sediments derived from SW Norway as based on single zircon (SIMS) data. Norsk Geologisk Tidsskrift, 81, 283–292.
    OpenUrl
  74. ↵
    1. Kozlovskaya, E.,
    2. Elo, S.,
    3. Hjelt, S.-E.,
    4. Yliniemi, J. &
    5. Pirttijaärvi, M. & SVEKALAPKO Seismic Tomography Working Group
    2004. 3-D density model of the crust of southern and central Finland obtained from joint interpretation of the SVEKALAPKO crustal P-wave velocity models and gravity data. Geophysical Journal International, 158, 827–848.
    OpenUrlCrossRefWeb of Science
  75. ↵
    1. Ksienzyk, A.K.,
    2. Dunkl, I.,
    3. Jacobs, J.,
    4. Fossen, H. &
    5. Kohlmann, F.
    2014. From orogen to passive margin: constraints from fission track and (U–Th)/He analyses on Mesozoic uplift and fault reactivation in SW Norway. In: Corfu, F., Gasser, D. & Chew, D.M. (eds) New Perspectives on the Caledonides of Scandinavia and Related Areas. Geological Society, London, Special Publications, 390, 679–702, https://doi.org/10.1144/SP390.27
    OpenUrlAbstract/FREE Full Text
  76. ↵
    1. Ksienzyk, A.K.,
    2. Wemmer, K.,
    3. Jacobs, J.,
    4. Fossen, H.,
    5. Schomberg, A.C.,
    6. Süssenberger, A.,
    7. Lünsdorf, N. &
    8. Bastesen, E.
    2016. Post-Caledonian brittle deformation in the Bergen area, West Norway: results from K–Ar illite fault gouge dating. Norwegian Journal of Geology, 96, 275–299.
    OpenUrl
  77. ↵
    1. Larsen, B.,
    2. Olaussen, S.,
    3. Sundvoll, B. &
    4. Heeremans, M.
    2008. Volcanoes and faulting in an arid climate. The Oslo Rift and North Sea in the Carboniferous and Permian, 359–251 million years ago. In: Ramberg, I.B., Bryhni, I., Nøttvedt, A. & Rangnes, K. (eds) The making of a land: geology of Norway. Norwegian Geological Association, Trondheim, 260–303.
  78. ↵
    1. Larson, S.Å.,
    2. Tullborg, E.L.,
    3. Cederbom, C. &
    4. Stiberg, J.P.
    1999. Sveconorwegian and Caledonian foreland basins in the Baltic Shield revealed by fission-track thermochronology. Terra Nova, 11, 215.
    OpenUrl
  79. ↵
    1. Lidmar-Bergström, K.
    1982. Pre-Quaternary geomorphological evolution in southern Fennoscandia. Sveriges Geologiska Undersökning C. Sveriges Geologiska Undersökning, Uppsala.
  80. ↵
    1. Lidmar-Bergström, K.
    1989. Exhumed Cretaceous landforms in south Sweden. Zeitschrift für Geomorphologie, Neue Folge, Supplementband, 72, 21–40.
    OpenUrl
  81. ↵
    1. Lidmar-Bergström, K.
    1995. Relief and saprolites through time on the Baltic Shield. Geomorphology, 12, 45–61.
    OpenUrlCrossRefWeb of Science
  82. ↵
    1. Lidmar-Bergström, K. &
    2. Bonow, J.M.
    2009. Hypotheses and observations on the origin of the landscape of southern Norway – A comment regarding the isostasy–climate–erosion hypothesis by Nielsen et al. 2008. Journal of Geodynamics, 48, 95–100.
    OpenUrlCrossRefWeb of Science
  83. ↵
    1. Lidmar-Bergström, K. &
    2. Näslund, J.O.
    2002. Landforms and uplift in Scandinavia. In: Doré, A.G., Cartwright, J.A., Stoker, M.S., Turner, J.P. & White, N. (eds) Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration. Geological Society, London, Special Publications, 196, 103–116, https://doi.org/10.1144/GSL.SP.2002.196.01.07
    OpenUrlAbstract/FREE Full Text
  84. ↵
    1. Lidmar-Bergström, K.,
    2. Olsson, S. &
    3. Roaldset, R.
    1999. Relief features and palaeoweathering remnants in formerly glaciated Scandinavian basement areas. In: Thiry, M. & Simon-Coinçon, R. (eds) Palaeoweathering, Palaeosurfaces and Related Continental Deposits. Special Publication of the International Association of Sedimentologists, 27, 275–301.
    OpenUrl
  85. ↵
    1. Lidmar-Bergström, K.,
    2. Ollier, C.D. &
    3. Sulebak, J.C.
    2000. Landforms and uplift history of southern Norway. Global and Planetary Change, 24, 211–231.
    OpenUrlCrossRefWeb of Science
  86. ↵
    1. Lidmar-Bergström, K.,
    2. Bonow, J.M. &
    3. Japsen, P.
    2013. Stratigraphic landscape analysis and geomorphological paradigms: Scandinavia as an example of Phanerozoic uplift and subsidence. Global and Planetary Change, 100, 153–171.
    OpenUrlCrossRef
  87. ↵
    1. Lidmar-Bergström, K.,
    2. Olvmo, M. &
    3. Bonow, J.M.
    2017. The South Swedish Dome: a key structure for identification of peneplains and conclusions on Phanerozoic tectonics of an ancient shield. GFF, 2017, 1–16.
    OpenUrl
  88. ↵
    1. Løseth, H.,
    2. Kyrkyebø, R.,
    3. Hilde, E.,
    4. Wild, R.J. &
    5. Bunkholt, H.
    2017. 500 m of rapid base level rise along an inner passive margin – Seismic observations from the Pliocene Molo Formation, mid-Norway. Marine and Petroleum Geology, 86, 268–287.
    OpenUrl
  89. ↵
    1. Lovell, B.
    2010. A pulse in the planet: regional control of high-frequency changes in relative sea level by mantle convection. Journal of the Geological Society, London, 167, 637–648, https://doi.org/10.1144/0016-76492009-127
    OpenUrlAbstract/FREE Full Text
  90. ↵
    1. Medvedev, S. &
    2. Hartz, E.H.
    2014. Evolution of topography of post-Devonian Scandinavia: Effects and rates of erosion. Geomorphology, 231, 229–245.
    OpenUrl
  91. ↵
    1. Migón, P. &
    2. Lidmar-Bergström, K.
    2001. Weathering mantles and their significance for geomorphological evolution of central and northern Europe since the Mesozoic. Earth-Science Reviews, 56, 285–324.
    OpenUrlCrossRef
  92. ↵
    1. Milnes, A.,
    2. Wennberg, O.,
    3. Skår, Ø. &
    4. Koestler, A.
    1997. Contraction, extension and timing in the South Norwegian Caledonides: the Sognefjord transect. In: Burg, J.-P. & Ford, M. (eds) Orogeny through Time. Geological Society, London, Special Publications, 121, 123–148, https://doi.org/10.1144/GSL.SP.1997.121.01.06
    OpenUrlAbstract/FREE Full Text
  93. ↵
    1. Molnar, P. &
    2. England, P.C.
    1990. Late Cenozoic uplift of mountain ranges and global climate change; chicken or egg? Nature, 346, 29–34.
    OpenUrlCrossRefWeb of Science
  94. ↵
    1. Nance, R.D.,
    2. Murphy, J.B. &
    3. Santosh, M.
    2014. The supercontinent cycle: A retrospective essay. Gondwana Research, 25, 4–29.
    OpenUrlCrossRefWeb of Science
  95. ↵
    1. Nielsen, L.H.
    2003. Late Triassic–Jurassic development of the Danish Basin and the Fennoscandian Border Zone, southern Scandinavia. In: Ineson, J. & Surlyk, F. (eds) The Jurassic of Denmark and Greenland. GEUS Bulletin, 2003/1, 459–526.
    OpenUrl
  96. ↵
    1. Nielsen, S.B.,
    2. Gallagher, K. et al.
    2009. The evolution of western Scandinavian topography: a review of Neogene uplift versus the ICE (isostasy–climate–erosion) hypothesis. Journal of Geodynamics, 47, 72–95.
    OpenUrlCrossRefWeb of Science
  97. ↵
    1. Norling, E. &
    2. Bergström, J.
    1987. Mesozoic and Cenozoic tectonic evolution of Scania, southern Sweden. Tectonophysics, 137, 7–19.
    OpenUrlCrossRefWeb of Science
  98. ↵
    1. Olaussen, S.,
    2. Larsen, B.T. &
    3. Steel, R.
    1994. The Upper Carboniferous–Permian Oslo Rift; basin fill in relation to tectonic development. In: Pangea: Global Environments and Resources. Canadian Society of Petroleum Geologists, Memoir, 17, 175–197.
    OpenUrl
  99. ↵
    1. Olesen, O.,
    2. Dehs, J.F.,
    3. Ebbing, J.,
    4. Henriksen, H.,
    5. Kihe, O. &
    6. Lundin, E.
    2007. Aeromagnetic mapping of deep-weathered fracture zones in the Oslo Region – a new tool for improved planning of tunnels. Norwegian Journal of Geology, 87, 253–267.
    OpenUrl
  100. ↵
    1. Olivarius, M.,
    2. Rasmussen, E.S.,
    3. Siersma, V.,
    4. Knudsen, C.,
    5. Kokfelt, T.F. &
    6. Keulen, N.
    2014. Provenance signal variations caused by facies and tectonics: Zircon age and heavy mineral evidence from Miocene sand in the north-eastern North Sea Basin. Marine and Petroleum Geology, 49, 1–14.
    OpenUrl
  101. ↵
    1. Olvmo, M.,
    2. Lidmar-Bergström, K. &
    3. Lindberg, G.
    1999. The glacial impact on an exhumed sub-Mesozoic etch surface in southwestern Sweden. Annals of Glaciology, 28, 153–160.
    OpenUrlCrossRefWeb of Science
  102. ↵
    1. Osmundsen, P.T. &
    2. Redfield, T.
    2011. Crustal taper and topography at passive continental margins. Terra Nova, 23, 1–13.
    OpenUrlCrossRefWeb of Science
  103. ↵
    1. Pedersen, V.K.,
    2. Huismans, R.S. &
    3. Moucha, R.
    2016. Isostatic and dynamic support of high topography on a North Atlantic passive margin. Earth and Planetary Science Letters, 446, 1–9.
    OpenUrl
  104. ↵
    1. Pedersen, V.K.,
    2. Braun, J. &
    3. Huismans, R.S.
    2018. Eocene to mid-Pliocene landscape evolution in Scandinavia inferred from offshore sediment volumes and pre-glacial topography using inverse modelling. Geomorphology, 303, 467–485.
    OpenUrl
  105. ↵
    1. Pekeris, C.L.
    1935. Thermal convection in the interior of the Earth. Geophysical Journal International, 3, 343–367.
    OpenUrlCrossRef
  106. ↵
    1. Peryt, T.,
    2. Geluk, M.,
    3. Mathiesen, A.,
    4. Paul, J. &
    5. Smith, K.
    2010. Zechstein. In: Doornenbal., H. & Stevenson, A. (eds) Petroleum Geological Atlas of the Southern Permian Basin Area. EAGE, Houten, 123–147.
  107. ↵
    1. Ramberg, I.B.,
    2. Bryhni, I.,
    3. Nøttvedt, A. &
    4. Rangnes, K.
    (eds) 2008. The making of a land: geology of Norway. Norwegian Geological Association, Trondheim.
  108. ↵
    1. Rasmussen, E.S.
    2004. The interplay between true eustatic sea-level changes, tectonics, and climatic changes: what is the dominating factor in sequence formation of the Upper Oligocene–Miocene succession in the eastern North Sea Basin, Denmark? Global and Planetary Change, 41, 15–30.
    OpenUrlCrossRefWeb of Science
  109. ↵
    1. Rasmussen, E.S.
    2005. The geology of the upper Middle–Upper Miocene Gram Formation in the Danish area. In: Roth, F. & Hoedemarkers, K. (eds) The Gram Book. Palaeontos, 7, 5–18.
    OpenUrl
  110. ↵
    1. Rasmussen, E.S.
    2014. Development of an incised-valley fill under the influence of tectonism and glacio-eustatic sea-level change: valley morphology, fluvial style, and lithology. Journal of Sedimentary Research, 84, 278–300.
    OpenUrlAbstract/FREE Full Text
  111. ↵
    1. Rasmussen, E.S.
    2017. Sedimentology and sequence stratigraphy of the uppermost upper Oligocene–Miocene fluvio-deltaic system in the eastern North Sea Basin: the influence of tectonism, eustacy and climate. Doctoral thesis, University of Copenhagen. GEUS, Copenhagen, http://www.geus.dk/publications/theses/erik_skovbjerg_rasmussen-doctoral_thesis_2017.pdf
  112. ↵
    1. Rasmussen, E.S.,
    2. Heilmann-Clausen, C.,
    3. Waagstein, R. &
    4. Eidvin, T.
    2008. The Tertiary of Norden. Episodes, 31, 66–72.
    OpenUrlWeb of Science
  113. ↵
    1. Rasmussen, E.S.,
    2. Dybkjær, K. &
    3. Piasecki, S.
    2010. Lithostratigraphy of the Upper Oligocene–Miocene succession of Denmark. Geological Survey of Denmark and Greenland Bulletin, 2010/22.
  114. ↵
    1. Redfield, T. &
    2. Osmundsen, P.
    2009. The Tjellefonna fault system of Western Norway: Linking late-Caledonian extension, post-Caledonian normal faulting, and Tertiary rock column uplift with the landslide-generated tsunami event of 1756. Tectonophysics, 474, 106–123.
    OpenUrlCrossRefWeb of Science
  115. ↵
    1. Redfield, T.F. &
    2. Osmundsen, P.T.
    2013. The long-term topographic response of a continent adjacent to a hyperextended margin: A case study from Scandinavia. Geological Society of America Bulletin, 125, 184–200.
    OpenUrlAbstract/FREE Full Text
  116. ↵
    1. Redfield, T.F.,
    2. Torsvik, T.H.,
    3. Andriessen, P.A.M. &
    4. Gabrielsen, R.H.
    2004. Mesozoic and Cenozoic tectonics of the More Trøndelag Fault Complex, central Norway: constraints from new apatite fission track data. Physics and Chemistry of the Earth, 29, 673–682.
    OpenUrlCrossRefWeb of Science
  117. ↵
    1. Redfield, T.F.,
    2. Osmundsen, P.T. &
    3. Hendriks, B.W.H.
    2005a. The role of fault reactivation and growth in the uplift of western Fennoscandia. Journal of the Geological Society, London, 162, 1013–1030, https://doi.org/10.1144/0016-764904-149
    OpenUrlAbstract/FREE Full Text
  118. ↵
    1. Redfield, T.F.,
    2. Braathen, A.,
    3. Gabrielsen, R.H.,
    4. Osmundsen, P.T.,
    5. Torsvik, T.H. &
    6. Andriessen, P.A.M.
    2005b. Late Mesozoic to early Cenozoic components of vertical separation across the More–Trøndelag Fault Complex, Norway. Tectonophysics, 395, 233–249.
    OpenUrlCrossRefWeb of Science
  119. ↵
    1. Reusch, H.
    1894. Strandfladen, et nyt træk i Norges geografi. Norges geologiske undersøkelse, 14, 1–12.
    OpenUrl
  120. ↵
    1. Reusch, H.
    1901. En ejendommelighet ved Skandinaviens hovedvannskille. Norges Geologiske Undersøkelse, 32, 124–263.
    OpenUrl
  121. ↵
    1. Riber, L.,
    2. Dypvik, H. &
    3. Sørlie, R.
    2015. Altered basement rocks on the Utsira High and its surroundings, Norwegian North Sea. Norwegian Journal of Geology, 95, 57–89.
    OpenUrl
  122. ↵
    1. Rickers, F.,
    2. Fichtner, A. &
    3. Trampert, J.
    2013. The Iceland–Jan Mayen plume system and its impact on mantle dynamics in the North Atlantic region: Evidence from full-waveform inversion. Earth and Planetary Science Letters, 367, 39–51.
    OpenUrlCrossRefWeb of Science
  123. ↵
    1. Riis, F.
    1996. Quantification of Cenozoic vertical movements of Scandinavia by correlation of morphological surfaces with offshore data. Global and Planetary Change, 12, 331–357.
    OpenUrlCrossRefWeb of Science
  124. ↵
    1. Robertson, E.
    1988. Thermal properties of rocks. US Geological Survey, Open-File Report, 88-441.
  125. ↵
    1. Rohrman, M.,
    2. van der Beek, P.,
    3. Andriessen, P. &
    4. Cloetingh, S.A.P.L.
    1995. Meso-Cenozoic morphotectonic evolution of southern Norway: Neogene domal uplift inferred from apatite fission track thermochronology. Tectonics, 14, 700–714.
    OpenUrl
  126. ↵
    1. Rohrman, M.,
    2. Andriessen, P. &
    3. van der Beek, P.
    1996. The relationship between basin and margin thermal evolution assessed by fission track thermochronology: an application to offshore southern Norway. Basin Research, 8, 45–63.
    OpenUrlCrossRefWeb of Science
  127. ↵
    1. Rudberg, S.
    1960. Geology and geomorphology. In: Sømme, S. (ed.) A Geography of Norden. Cappelens, Oslo, 27–40.
  128. ↵
    1. Schiøler, P.,
    2. Andsbjerg, J. et al.
    2007. Lithostratigraphy of the Palaeogene to Lower Neogene (Rogaland to Westray groups) sediments of the Danish sector of the North Sea. GEUS Bulletin, 2007/12.
  129. ↵
    1. Schiphull, K.
    1974. Geomorphologische Studien in zentralen Südnorwegen mit Beiträgen über Regelungs- und Steuerungssysteme in der Geomorphologie. Hamburger Geographische Studien, 31.
  130. ↵
    1. Scholle, P.A.
    1977. Chalk diagenesis and its relation to petroleum exploration; oil from chalks, a modern miracle? AAPG Bulletin, 61, 982–1009.
    OpenUrlAbstract
  131. ↵
    1. Schoonman, C.,
    2. White, N. &
    3. Pritchard, D.
    2017. Radial viscous fingering of hot asthenosphere within the Icelandic plume beneath the North Atlantic Ocean. Earth and Planetary Science Letters, 468, 51–61.
    OpenUrl
  132. ↵
    1. Sigmond, E.M.O.
    1993. Bedrock map of Norway and adjacent ocean areas. 1:3 000 000. Geological Survey of Norway, Oslo.
  133. ↵
    1. Slagstad, T.,
    2. Balling, N.,
    3. Elvebakk, H.,
    4. Midttømme, K.,
    5. Olesen, O.,
    6. Olsen, L. &
    7. Pascal, C.
    2009. Heat-flow measurements in Late Palaeoproterozoic to Permian geological provinces in south and central Norway and a new heat-flow map of Fennoscandia and the Norwegian–Greenland Sea. Tectonophysics, 473, 341–361.
    OpenUrlCrossRefWeb of Science
  134. ↵
    1. Sømme, T.O.,
    2. Helland-Hansen, W. &
    3. Martinsen, O.J.
    2013a. Quantitative aspects of stratigraphic onshore–offshore relationships along the western margin of southern Norway: implications for Late Mesozoic and Cenozoic topographic evolution. Norwegian Journal of Geology, 93, 261–276.
    OpenUrl
  135. ↵
    1. Sømme, T.O.,
    2. Martinsen, O.J. &
    3. Lunt, I.
    2013b. Linking offshore stratigraphy to onshore paleotopography: The Late Jurassic–Paleocene evolution of the south Norwegian margin. Geological Society of America Bulletin, 125, 23.
    OpenUrl
  136. ↵
    1. Steer, P.,
    2. Huismans, R.S.,
    3. Valla, P.G.,
    4. Gac, S. &
    5. Herman, F.
    2012. Bimodal Plio-Quaternary glacial erosion of fjords and low-relief surfaces in Scandinavia. Nature Geoscience, 5, 635–639.
    OpenUrlCrossRef
  137. ↵
    1. Stoker, M.S.,
    2. Praeg, D.,
    3. Hjelstuen, B.O.,
    4. Laberg, J.S.,
    5. Nielsen, T. &
    6. Shannon, P.M.
    2005. Neogene stratigraphy and the sedimentary and oceanographic development of the NW European Atlantic margin. Marine and Petroleum Geology, 22, 977–1005.
    OpenUrlCrossRefWeb of Science
  138. ↵
    1. Stratford, W. &
    2. Thybo, H.
    2011. Seismic structure and composition of the crust beneath the southern Scandes. Norway Tectonophysics, 502, 364–382.
    OpenUrl
  139. ↵
    1. Tappe, S.
    2004. Mesozoic mafic alkaline magmatism of southern Scandinavia. Contributions to Mineralogy and Petrology, 148, 312–334.
    OpenUrlCrossRefWeb of Science
  140. ↵
    1. Underhill, J.R. &
    2. Partington, M.A.
    1993. Jurassic thermal doming and deflation in the North Sea: implications of the sequence stratigraphic evidence. In: Parker, J.R. (ed.) Petroleum Geology of North West Europe: Proceedings of the 4th Conference. Geological Society, London, 337–345, https://doi.org/10.1144/0040337
  141. ↵
    1. Veevers, J.J.
    2004. Gondwanaland from 650–500 Ma assembly through 320 Ma merger in Pangea to 185–100 Ma breakup: supercontinental tectonics via stratigraphy and radiometric dating. Earth-Science Reviews, 68, 1–132.
    OpenUrlCrossRef
  142. ↵
    1. Wråk, W.
    1908. Bidrag til Skandinaviens reliefkronologi. Ymer, 254–300.
  143. ↵
    1. Zeck, H.,
    2. Andriessen, P.,
    3. Hansen, K.,
    4. Jensen, P. &
    5. Rasmussen, B.
    1988. Paleozoic paleo-cover of the southern part of the Fennoscandian Shield– fission track constraints. Tectonophysics, 149, 61–66.
    OpenUrlCrossRefWeb of Science
  144. ↵
    1. Ziegler, P.
    1985. Late Caledonian framework of western and central Europe. In: Gee, D.G. & Sturt, B.A. (eds) The Caledonide Orogen – Scandinavia and Related Areas. Wiley, Chichester, 3–18.
  145. ↵
    1. Ziegler, P.A.
    1990. Geological Atlas of Western and Central Europe. Shell International, The Hague.
View Abstract
PreviousNext
Back to top

In this issue

Journal of the Geological Society: 175 (5)
Journal of the Geological Society
Volume 175, Issue 5
September 2018
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Back Matter (PDF)
  • Front Matter (PDF)
Alerts
Sign In to Email Alerts with your Email Address
Citation tools

Mountains of southernmost Norway: uplifted Miocene peneplains and re-exposed Mesozoic surfaces

Peter Japsen, Paul F. Green, James A. Chalmers and Johan M. Bonow
Journal of the Geological Society, 175, 721-741, 12 July 2018, https://doi.org/10.1144/jgs2017-157
Peter Japsen
1Geological Survey of Denmark and Greenland (GEUS), Øster Voldgade 10, DK-1350 Copenhagen, Denmark
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Peter Japsen
  • For correspondence: pj@geus.dk
Paul F. Green
2Geotrack International, 37 Melville Road, Brunswick West, VIC 3055, Australia
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
James A. Chalmers
1Geological Survey of Denmark and Greenland (GEUS), Øster Voldgade 10, DK-1350 Copenhagen, Denmark
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for James A. Chalmers
Johan M. Bonow
3Geovisiona AB, Högbyvägen 168, SE-17554 Järfälla, Sweden
4Uppsala University, Box 513, SE-75120 Uppsala, Sweden
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Johan M. Bonow

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Permissions
View PDF
Share

Mountains of southernmost Norway: uplifted Miocene peneplains and re-exposed Mesozoic surfaces

Peter Japsen, Paul F. Green, James A. Chalmers and Johan M. Bonow
Journal of the Geological Society, 175, 721-741, 12 July 2018, https://doi.org/10.1144/jgs2017-157
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
Email to

Thank you for sharing this Journal of the Geological Society article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Mountains of southernmost Norway: uplifted Miocene peneplains and re-exposed Mesozoic surfaces
(Your Name) has forwarded a page to you from Journal of the Geological Society
(Your Name) thought you would be interested in this article in Journal of the Geological Society.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
Download PPT
  • Tweet Widget
  • Facebook Like
  • Google Plus One
  • Article
    • Abstract
    • Geological outline
    • The landscape of the Southern Scandes
    • Apatite fission-track data from southern Norway
    • Regional events defined from AFTA and their relation to geology and landscapes
    • Reconstruction of the late Cenozoic development of the southernmost Scandes
    • Discussion
    • Mechanisms
    • Conclusions
    • Acknowledgements
    • Funding
    • References
  • Figures & Data
  • Info & Metrics
  • PDF

Related Articles

Similar Articles

Cited By...

More in this TOC Section

  • New sequence stratigraphic framework on a ‘passive’ margin: implications for the post-break-up depositional environment and onset of glaciomarine conditions in NE Greenland
  • Tectonic significance of the Variscan suture between Brunovistulia and the Bohemian Massif
  • Sedimentology and provenance of the Lower Old Red Sandstone Grampian outliers: implications for Caledonian orogenic basin development and the northward extension of the Midland Valley Basin
Show more: Research article
  • Most read
  • Most cited
Loading
  • Geological Society of London Scientific Statement: what the geological record tells us about our present and future climate
  • The Nonesuch Formation Lagerstätte: a rare window into freshwater life one billion years ago
  • The Shibantan Lagerstätte: insights into the Proterozoic–Phanerozoic transition
  • Linking surface and subsurface volcanic stratigraphy in the Turkana Depression of the East African Rift system
  • Terrestrial stratigraphical division in the Quaternary and its correlation
More...

Journal of the Geological Society

  • About the journal
  • Editorial Board
  • Submit a manuscript
  • Author information
  • Supplementary Publications
  • Subscribe
  • Pay per view
  • Alerts & RSS
  • Copyright & Permissions
  • Activate Online Subscription
  • Feedback
  • Help

Lyell Collection

  • About the Lyell Collection
  • Lyell Collection homepage
  • Collections
  • Open Access Collection
  • Open Access Policy
  • Lyell Collection access help
  • Recommend to your Library
  • Lyell Collection Sponsors
  • MARC records
  • Digital preservation
  • Developing countries
  • Geofacets
  • Manage your account
  • Cookies

The Geological Society

  • About the Society
  • Join the Society
  • Benefits for Members
  • Online Bookshop
  • Publishing policies
  • Awards, Grants & Bursaries
  • Education & Careers
  • Events
  • Geoscientist Online
  • Library & Information Services
  • Policy & Media
  • Society blog
  • Contact the Society

Published by The Geological Society of London, registered charity number 210161

Print ISSN 
0016-7649
Online ISSN 
2041-479X

Copyright © 2021 Geological Society of London