Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-29T09:06:44.972Z Has data issue: false hasContentIssue false

Disruption and localization of sediment pathways by continental extension: Detrital-zircon provenance change from upper Triassic to lower Jurassic in the northern Sverdrup Basin, Nunavut

Published online by Cambridge University Press:  02 April 2024

Derrick Midwinter
Affiliation:
Geological Survey of Canada, 3303-33rd St NW, Calgary, AB T2L 2A7, Canada
Thomas Hadlari*
Affiliation:
Geological Survey of Canada, 3303-33rd St NW, Calgary, AB T2L 2A7, Canada
Keith Dewing
Affiliation:
Geological Survey of Canada, 3303-33rd St NW, Calgary, AB T2L 2A7, Canada
William A Matthews
Affiliation:
University of Calgary, Department of Geosciences, Calgary, AB T2N 1N4, Canada
*
Corresponding author: Thomas Hadlari; Email: thomas.hadlari@canada.ca
Rights & Permissions [Opens in a new window]

Abstract

Constraints on the tectonic setting of the upper Triassic to lower Jurassic in the Sverdrup Basin can be elucidated from detrital-zircon U-Pb ages. During the Triassic, there was a dual provenance system into sedimentary basins along the western and northern margins of Laurentia. One of the sediment sources was from an extra-basinal igneous source of Permian-Triassic zircon while the other source was recycled sediment eroded from older sedimentary basins. The Heiberg Formation/Group was deposited during a period of significant siliciclastic sedimentation into the basin from the upper Triassic to the lower Jurassic and comprises three members: Romulus, Fosheim and Remus. Previous work has interpreted that the Carboniferous-Permian-Triassic detrital zircon had stopped reaching the northern part of the Sverdrup Basin by deposition of the upper Heiberg Formation (lower Jurassic). New detrital-zircon age analyses from samples along the northern part of the basin spanning different horizons in the Heiberg Formation show that the typical extra-basinal signature, with abundant Carboniferous-Permian-Triassic ages, was no longer recorded during the initial deposition of the Fosheim Member during the latest Triassic. Previously published basin analysis from the Sverdrup Basin interprets syn-Jurassic extensional faults and so we relate the provenance change to the onset of extension. It is interpreted that the Sverdrup Basin transitioned from a basin that received sediment from a northern extra-basinal igneous source during deposition of the Romulus Member to an extensional basin by the deposition of the Fosheim Member in the latest Triassic, as the northern sediment source was interrupted by intervening extensional basins of the proto-Amerasia Basin.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

Detrital-zircon geochronology from siliciclastic sedimentary successions can provide important insight into tectonic setting and tectonic evolution of a sedimentary basin. Rift basins, in particular, are typically characterized by rift flank erosion and relatively local sedimentary provenance (e.g., Withjack et al. Reference Withjack, Schlische and Olsen2002; Cawood et al. Reference Cawood, Hawkesworth and Dhuime2012). The Sverdrup Basin formed initially as a Carboniferous rift basin and subsequent Jurassic-Cretaceous rifting culminated in opening of the Amerasia Basin (e.g., Embry & Beauchamp, Reference Embry, Beauchamp and Miall2008). The Triassic tectonic setting is less clear but, from previous studies, two predominant sediment sources into the Sverdrup Basin during the Triassic have been identified (e.g. Embry, Reference Embry2009; Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016); one from the southern and eastern margins of the basin, and another derived from the northern margin.

The nature of the tectonic setting of the Sverdrup Basin during the Triassic was long considered tectonically quiescent (e.g. Embry & Beauchamp, Reference Embry, Beauchamp and Miall2008). The presence of Triassic detrital zircon within Triassic strata (Miller et al. Reference Miller, Toro, Gehrels, Amato, Prokopiev, Tuchkova, Akinin, Dumitru, Moore and Cecile2006; Omma et al. Reference Omma, Pease, Scott, Spencer, Gautier, Stoupakova, Embry and Sørensen2011) led to interpretations that sedimentary provenance was from the Polar Urals (e.g., Miller et al. Reference Miller, Soloviev, Prokopiev, Toro, Harris, Kuzmichev and Gehrels2013; Anfinson et al. Reference Anfinson, Embry and Stockli2016) and that sedimentary systems that traversed the greater Barents Sea Basin (e.g., Gilmullina et al. Reference Gilmullina, Klausen, Dore, Sirevaag, Suslova and Eide2022; Klausen et al. Reference Klausen, Rismyhr, Muller and Olaussen2022) directed sediment into the Sverdrup Basin (Miller et al. Reference Miller, Soloviev, Prokopiev, Toro, Harris, Kuzmichev and Gehrels2013). Observations of volcanic ash beds and syn-depositional detrital-zircon ages through the late Permian to late Triassic indicate volcanism proximal to the northwestern margin of the basin (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016; Hadlari et al. Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018; Alonso-Torres et al. Reference Alonso-Torres, Beauchamp, Guest, Hadlari and Matthews2018). From the combination of Permian-Triassic volcanism and syn-depositional zircon, it has been proposed that there was a magmatic arc to the north of the basin during the late Permian and Triassic, and that the Sverdrup Basin was in a retro-arc position during that time (Hadlari et al. Reference Hadlari, Midwinter, Poulton and Matthews2017; Alonso-Torres et al. Reference Alonso-Torres, Beauchamp, Guest, Hadlari and Matthews2018; Hadlari et al. Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018). The detrital-zircon U-Pb age signature along the northern margin of the Sverdrup Basin is characterized by syn-depositional ages in upper Triassic strata that are absent from lower and middle Jurassic strata (Omma et al. Reference Omma, Pease, Scott, Spencer, Gautier, Stoupakova, Embry and Sørensen2011; Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016), indicating the diminution of the hypothetical magmatic arc and/or the Polar Urals as a sediment source, potentially as initial rifting of the Amerasia Basin isolated the Sverdrup Basin from the extra-basinal source(s) (Embry, Reference Embry2009). The Heiberg Formation/Group was deposited during this key interval with the Romulus Member in a late Triassic pre-rift setting, whereas the transition to syn-rift had taken place by deposition of the Remus member in the Early Jurassic (Pleinsbachian) (Embry, Reference Embry2009; Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016). To accommodate uncertainties regarding the sub-Fosheim basin setting, we simply refer to the Romulus Member and older strata as ‘pre-rift’ (Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016). Deformation, uplift and erosion during the Eurekan Orogeny in the Cenozoic have obscured older faults on Axel Heiberg and Ellesmere Islands (e.g., Piepjohn & von Gosen, Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018). The best evidence for syn-rift normal faults is seismic data from Prince Patrick Island that show half-graben basins, some with the upper Heiberg Formation/Group at the base, consistent with the onset of rifting within the Heiberg Formation/Group (see discussion, Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016). The change in detrital-zircon provenance is determined by detrital-zircon samples from the lower (Romulus) and upper (Remus) members of the Heiberg Formation, which are grab samples collected during regional mapping (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016), with no indication where this important change takes place within the intervening stratigraphy. This paper provides new detrital-zircon data from measured sections of this interval and four samples that span the boundary between the Romulus and Fosheim members (lower and middle Heiberg Formation), which serves to constrain the timing of the provenance change that is inferred to record onset of rifting in the proto-Amerasia Basin.

2. Geological setting

2.a. Sverdrup Basin

The Sverdrup Basin (Figure 1) is underlain by an up to 10 km-thick sedimentary succession of Devonian clastic wedge strata that were deformed during the Late Devonian-Early Carboniferous Ellesmerian orogeny (Embry, Reference Embry and Trettin1991). The Ellesmerian orogeny was succeeded by initial rifting of the Sverdrup Basin, beginning in the Carboniferous where up to 5 km of Upper Paleozoic strata accumulated in a shelf to deep-basin setting (Balkwill, Reference Balkwill1978), and subsequently underwent significant subsidence in the Triassic followed by Early Jurassic – Early Cretaceous rifting resulting in the opening of the Amerasia Basin (e.g. Grantz et al. Reference Grantz, Eittreim and Dinter1979, Reference Grantz, Hart and Childers2011; Houseknecht & Bird, Reference Houseknecht and Bird2011; Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016). Sediment deposited along the northern margin of the basin has been interpreted to be derived from a poorly constrained landmass that lay to the north of the basin based on sediment provenance direction (Embry, Reference Embry2009). While the basin was long considered tectonically quiescent during the Triassic (e.g. Embry & Beauchamp, Reference Embry, Beauchamp and Miall2008), new data support the idea that the basin had tectonic activity, such as the observation of volcanic ash beds (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016) and the undefined Triassic faults and Jurassic extensional faults at the eastern end of the Sverdrup Basin, which are parallel to the eastern flank of the Amerasia Basin (Lopez-Mir et al. Reference Lopez-Mir, Schneider and Hülse2018).

Figure 1. (a) Map of the Sverdrup Basin showing the location of detrital-zircon samples from the Heiberg Formation/Group. Surface and sea bottom bedrock geology is from Okulitch (Reference Okulitch1991). Detrital-zircon sample locations are from this study, Midwinter et al. (Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016), Hadlari et al. (Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018). See Figures 2 and 3 for stratigraphic locations of samples. (b) Map of the Arctic, depicting the basins and regions, mentioned in the text. Outline of AACM (black line) after Drachev (Reference Drachev2011). Rotation of AACM away from the Canadian margin illustrated by black arrows (after). Red dashed line denotes location of Figure 1a. Red circle indicates the approximate location of Mendeleev rise from the Shamshur Seamount sample collected by Tuchkova et al. (Reference Tuchkova, Shokalsky, Petrov, Sokolov, Sergeev and Moiseev2020). AA, Arctic Alaska; AACM, Arctic Alaska – Chukotka Microplate; Ch, Chukotka; M, Mendeleev rise; WI, Wrangel Island.

The Heiberg Formation in the central and eastern part of the basin is subdivided into three members, Romulus, Fosheim and Remus (Figure 2; Embry, Reference Embry1983a), and is approximately 1 km thick along the eastern margin of the basin (e.g. Midwinter et al. Reference Midwinter, Hadlari and Dewing2022). The lowermost strata in the Heiberg Formation, the Romulus Member, records a coarsening-upward succession of mudstone to fine-grained sandstone from a prodelta to delta plain environment; the overlying Fosheim Member is a coal-bearing and sandstone-rich interval deposited in a mixed alluvial-marine environment. The Romulus Member is more arkosic than the Fosheim Member, which is more quartzose. The Remus Member is a sandstone-rich unit representative of shallow marine deposits (Embry, Reference Embry1983a; Midwinter et al. Reference Midwinter, Hadlari and Dewing2022). The Heiberg Formation is stratigraphically equivalent to the Heiberg Group in the western part of the basin (Embry, Reference Embry1983b).

Figure 2. Triassic to Early Cretaceous stratigraphic chart of the Sverdrup Basin (Embry & Beauchamp, Reference Embry, Beauchamp and Miall2008; Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016) with location and reference of detrital-zircon samples with respect to stratigraphy.

The Heiberg Group is comprised of five formations: the sandstone-rich Skybattle Formation (equivalent to the Romulus Member), overlain by marine mudstone of the Grosvenor Island Formation, then coarsening upward into the deltaic Maclean Strait Formation, which is subsequently overlain by marine shale of the Lougheed Island Formation and ultimately capped by the sandstone-dominant King Christian Formation (Figure 2). As identified from palynological studies (Embry & Suneby, Reference Embry and Suneby1994), the Grosvenor Island Formation contains the Triassic-Jurassic boundary. This basin centre mudstone is correlative with thin marine mudstone in the lower portion of the Fosheim Member (Suneby & Hills, Reference Suneby and Hills1988; also see Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016). Additional evidence for this mudstone being correlative to the Grosvenor Island Formation is the presence of ironstone rubble (Midwinter et al. Reference Midwinter, Hadlari and Dewing2022), as ironstone was observed within the Grosvenor Island Formation along the southwestern margin of the Sverdrup Basin, indicative of a starved shelf (Embry & Johannessen, Reference Embry, Johannessen and Vorren1993).

The Heiberg Formation records the transition of the pre-rift stage into the syn-rift stage of the Sverdrup Basin as shallow marine sandstones prograded across the basin and multiple unconformities within the formation indicate a basin-filled state with the rift onset unconformity likely within the Fosheim Member, previously estimated to be ∼200-190 Ma (Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016) or ∼170 Ma (Embry & Beauchamp, Reference Embry, Beauchamp and Miall2008). This was supported by detrital-zircon geochronology suggesting the King Christian Formation (coeval to the upper Fosheim and Remus members) had a change in sediment source relative to the lower Heiberg Group/Formation due to the onset of rifting to the north of the Sverdrup Basin (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016). This period marked the initial extension of the proto-Amerasia Basin, which could have developed as a retroarc system driven by slab rollback, which in turn rifted the Arctic Alaska-Chukotka Microplate (AACM) from northern Laurentia, thus disrupting the pathway for the northern sediment source to the Sverdrup Basin. This key transition occurred within the sedimentary succession of the Heiberg Formation, which makes samples tied to stratigraphic sections measured in the Blue Mountains and at Depot Point (Midwinter et al. Reference Midwinter, Hadlari and Dewing2022) vital to better constrain the timing of the extension.

2.a. Summary of published U-Pb provenance: Sverdrup Basin

U-Pb detrital-zircon provenance data reported from Triassic to Jurassic-aged strata from the Sverdrup Basin have a sampling gap from the upper Triassic to lower Jurassic of the Heiberg Formation/Group. Previous detrital-zircon U-Pb age studies of Triassic-aged strata (e.g. Miller et al. Reference Miller, Toro, Gehrels, Amato, Prokopiev, Tuchkova, Akinin, Dumitru, Moore and Cecile2006; Omma et al. Reference Omma, Pease, Scott, Spencer, Gautier, Stoupakova, Embry and Sørensen2011; Anfinson et al. Reference Anfinson, Embry and Stockli2016; Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016; Hadlari et al. Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018) identified two different detrital-zircon signatures: a spectrum similar to the upper portion of the Devonian clastic wedge with a characteristic signature of 700–360 Ma ages and a broad spectrum of Meso- to Paleoproterozoic ages (Anfinson et al. Reference Anfinson, Leier, Embry and Dewing2012a, Reference Anfinson, Leier, Gaschnig, Embry and Dewing2012b), and Permian-Triassic grains that are near the age of deposition. Detrital-zircon samples from lower Jurassic-aged strata collected along the northern margin do not record the Carboniferous-Permian-Triassic age spectrum (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016).

Previous studies used samples collected from the Heiberg Formation/Group; however, samples that are not well constrained by measured sections and/or mapping are not included in this study (e.g., Romulus Member samples from Anfinson et al. (Reference Anfinson, Embry and Stockli2016) and the Remus Member sample from Hadlari et al. Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018). Additionally, the C-numbered samples from Omma et al. (Reference Omma, Pease, Scott, Spencer, Gautier, Stoupakova, Embry and Sørensen2011) are from the field collections of the Geological Survey of Canada (GSC), and the GSC scientist who donated them to that study asserts that the Heiberg sample is not from the upper Heiberg (Embry pers.comm., 2015), and so it is omitted. Three samples, two from the Romulus Member (measured sections) and one from the King Christian Formation (map description) (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016; Hadlari et al. Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018), are included. In summary, previous detrital-zircon studies of the Sverdrup Basin indicate that an influx of sediment from an igneous source to the northwest persisted through the Triassic until some point between the late Triassic and Pleinsbachian.

3. Materials and methodology

A total of four new field samples are used in this study and the Supplementary Table (DR2) lists the sources of published data. Two samples were collected in 2015 from a measured section at Depot Point that spans from the upper Triassic Barrow Formation to unsubdivided Jurassic strata over an interval of 1,140 m. The Depot Point section is 40 km west of the type section for the Heiberg Formation at Buchanan Lake (Souther, Reference Souther1963). One sample from the lower Romulus Member, near the contact with the Barrow Formation, was collected from a field visit to the same location at Depot Point in 2011 with detrital-zircon geochronology results previously published (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016). Two samples were collected from a measured section at the Blue Mountains on northern Ellesmere Island in 2015. The stratigraphic section started in the upper Triassic Barrow Formation and ended in the lower Jurassic Jameson Bay Formation over approximately 1,200 m. For more details on the sedimentology and stratigraphy of these measured sections, see Midwinter et al. (Reference Midwinter, Hadlari and Dewing2022).

Detrital zircon was separated using standard separation techniques and isotopic signal intensities were measured for 300 grains per sample by quadrupole LA-ICP-MS (Matthews & Guest, Reference Matthews and Guest2017). Measurements were filtered using the calculated probability of concordance. Measurements with <1% probability of concordance were filtered from the dataset (812 of 1510 total). Dates used in the figures are 206Pb/238U for dates <1500 Ma and 207Pb/206Pb for older dates. Devonian clastic wedge reference spectra are derived from Anfinson et al. (Reference Anfinson, Leier, Embry and Dewing2012a, Reference Anfinson, Leier, Gaschnig, Embry and Dewing2012b). Maximum depositional ages (MDA) were calculated using the YGC2σ method (Dickinson & Gehrels, Reference Dickinson and Gehrels2009), which is the weighted mean of the youngest cluster of three or more dates that overlap at 2σ uncertainty. This approach has been shown to produce a conservative age estimate (Coutts et al. Reference Coutts, Matthews and Hubbard2019).

4. Detrital-zircon geochronology results

Isotopic measurements yielded 812 dates that passed our filtering criteria for use in this study. Results are presented as normalized probability density and cumulative probability plots in Figure 3. Details of the isotopic results and the MDA calculations are presented in the online Supplementary Material at http://journals.cambridge.org/geo.

Figure 3. Measured sections (a) with locations of detrital-zircon data (b) presented as normalized probability density functions. Sedimentological data and exact locations of the Blue Mountains and Depot Point sections can be found in Midwinter et al. 2021. All U-Pb data from this study, except for the Romulus Member sample from Depot Point (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016).

Sample Lower Romulus – Blue Mtns (SHF3) was collected from the lowermost sandstone of the Romulus Member (97 m, Figure 3) in the Blue Mountains, 1–2 m above the contact with the mudstone-rich Barrow Formation. The sandstone bed is fine-grained and trough and planar cross-bedded. The proportions of age groupings from n = 184 measurements were Archean (5%), Paleoproterozoic (18%), Mesoproterozoic (13%), 700–360 Ma (29%), Carboniferous (14%), Permian (9%) and Triassic (6%) with a near-continuous distribution of dates between 610 Ma and 209 Ma. A geologically important age fraction from 350 Ma to 200 Ma consists of 51 of 184 measurements (28%). The calculated MDA incorporates the five youngest grains and yielded a weighted mean age of 215.2 Ma ± 9.4 (2σ).

Sample Lower Fosheim - Depot Pt. (15-DTA-32-580m) was collected from a fine-grained sandstone bed immediately above the contact with the Romulus Member (580 m, Figure 3). The quartzose sandstone bed with erosional base was above the bioturbated sandstone of the Romulus Member and below a thin coal bed (Figure 3; Midwinter et al. Reference Midwinter, Hadlari and Dewing2022). The majority of the n = 172 measurements were Mesoproterozoic (1600-1000 Ma; 41%) and Paleoproterozoic (2500-1600 Ma; 32%). Lesser fractions were Archean (>2500 Ma; 8%) and 700–360 Ma (13%). The vast majority of measurements filtered out by probability of concordance are Proterozoic, with no apparent 206Pb/238U ages younger than 350 Ma (discordant or concordant, no ages younger than 350 Ma have been filtered out). The youngest date was 353.7 Ma ± 16.5 (2σ); therefore, there are no detrital-zircon grains in this sample dated between 350 Ma and 200 Ma.

Sample Lower Fosheim – Depot Pt. (15-DTA-32-660m) was collected from a medium-grained, quartzose, trough cross-bedded sandstone approximately 80 m above the contact with the Romulus Member (660 m, Figure 3). This sample yielded n = 246 measurements that passed the filtering criteria, and those that were filtered out are mainly Proterozoic with none younger than 350 Ma. The age spectrum is very similar to the basal lower Fosheim Member sample with the majority of the measurements yielding Mesoproterozoic (43%) and Paleoproterozoic (32%) dates. Other age groupings were Archean (8%) and 700–360 Ma (13%). The youngest date was 406.2 Ma ± 9.7 (2σ); therefore, there are no detrital-zircon grains in this sample dated between 350 Ma and 200 Ma.

Sample Upper Fosheim – Blue Mtns (SHF9) was collected from thinly interbedded very fine-grained quartzose sandstone with carbonaceous mudstone in the Blue Mountains (876 m, Figure 3). Additional structures observed in this sedimentary bed were wavy, ripple cross-laminations. This interval is approximately 150 m below the contact to the Remus Member. From n = 210 measurements, 7% were Archean, 36% were Paleoproterozoic, 46% were Mesoproterozoic and 7% were between 700 Ma and 360 Ma. A single grain dated at 261.5 Ma ± 12.2 (2σ) was found between 350 Ma and 200 Ma, and none are Triassic.

5. Discussion

The collection of samples from the Heiberg Formation indicates a vertical stratigraphic transition between two provenance signatures (Figure 4). All detrital-zircon samples of the Romulus Member have a consistent and continuous age spectrum from the Carboniferous to Permian through the Triassic where the youngest-aged fraction is near the age of deposition. For example, the 350 Ma to 200 Ma age fraction in sample SHF3 composes 28% of the detrital zircon (51 of 184 measurements). Conversely, above the contact between the Romulus and Fosheim members, the detrital-zircon signature changes to a spectrum with almost no Carboniferous-Triassic age probability (1 of 628 measurements; no Triassic) that is similar to the Devonian clastic wedge reference spectrum from the Franklinian Basin that has a moderate amount of 700–360 Ma ages, a prominent peak at ca. 430–410 Ma and abundant Meso- and Paleoproterozoic ages (Figure 4). This change in detrital-zircon age signature corresponds to the lithological change from arkosic in the Romulus Member to quartzose sandstone in the Fosheim Member. Data presented here for the Fosheim Member are consistent with U-Pb detrital-zircon data also from the Fosheim Member presented by Pointon et al. (Reference Pointon, Smyth, Omma, Morton, Schneider, Hulse, Rippington, Lopez-Mir, Crowley, Millar, Whitehouse, Frei, Scott and Flowerdew2023). Given that the Fosheim Member along the northern portion of the basin has only the older recycled detrital-zircon age spectrum, it can be interpreted that a shift in provenance likely occurs at the boundary between the Romulus and Fosheim members. Based on palynology from Ellesmere Island and interpreted by correlation of the member boundary to our sections, the Romulus Member is assigned a Norian to late Rhaetian age and the basal Fosheim Member consists of latest Rhaetian strata succeeded by the Triassic-Jurassic boundary interval (Suneby & Hills, Reference Suneby and Hills1988). In the southern portion of the Sverdrup Basin, the correlative Triassic-Jurassic boundary interval lies within the Grosvenor Formation (Embry & Suneby, Reference Embry and Suneby1994), and we speculate from regional correlations that the interval of the lower Fosheim Member from 588 to 607 m in the Depot Point section contains the Triassic-Jurassic boundary (Midwinter et al. Reference Midwinter, Hadlari and Dewing2022). This would place the change in provenance and rift onset unconformity as younger than Rhaetian fossils within the Romulus Member, older than the deposition of the Fosheim Member starting in the latest Rhaetian, and therefore late Rhaetian. Our estimation of the timing is dependent upon stratigraphic correlation from the sections with biostratigraphy to wells logs and our outcrop sections, which is a potential source of error and would benefit from additional geological age data.

Figure 4. Compiled normalized probability (a) and cumulative probability (b) plots with corresponding number of analyses. Data sources: the Devonian clastic wedge reference spectrum (Anfinson et al. Reference Anfinson, Leier, Embry and Dewing2012a, 2012b); Triassic sediment of Svalbard (Bue & Andresen, Reference Bue and Andresen2014); Triassic sediment of Chukotka (Miller et al. Reference Miller, Toro, Gehrels, Amato, Prokopiev, Tuchkova, Akinin, Dumitru, Moore and Cecile2006; Tuchkova et al. Reference Tuchkova, Sokolov, Khudoley, Verzhbitsky, Hayasaka and Moiseev2011; Amato et al. Reference Amato, Toro, Akinin, Hampton, Salnikov and Tuchkova2015); Triassic sediment of Wrangel Island (Miller et al. Reference Miller, Gehrels, Pease and Sokolov2010); Romulus Mbr. (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016; Hadlari et al. Reference Hadlari, Dewing, Matthews, Alonso-Torres and Midwinter2018; this study), Fosheim Mbr./King Christian Fm. (Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016; this study). ROU: Rift onset unconformity.

These two distinct assemblages, a recycled source and a volcanic arc source, are exhibited in Triassic strata in sedimentary basins along the western and northern Laurentian margin, including the Sverdrup Basin, Arctic Alaska, Yukon-Tanana, Southern Canada (Quesnel and Western Interior Basin) and the Southwest U.S (Hadlari et al. Reference Hadlari, Midwinter, Poulton and Matthews2017 and references therein). Triassic sedimentary rocks in Chukotka and Wrangel Island, part of the Arctic Alaska – Chukotka microplate, are potentially representative of the pre-rift source area and were most likely adjacent to the northwestern Canadian Arctic during this period (e.g. Gottlieb et al. Reference Gottlieb, Meisling, Miller and Charles2014; Tuchkova et al. Reference Tuchkova, Shokalsky, Petrov, Sokolov, Sergeev and Moiseev2020). The detrital-zircon signature of these predominantly upper Triassic rocks of Chukotka (Miller et al. Reference Miller, Toro, Gehrels, Amato, Prokopiev, Tuchkova, Akinin, Dumitru, Moore and Cecile2006; Tuchkova et al. Reference Tuchkova, Sokolov, Khudoley, Verzhbitsky, Hayasaka and Moiseev2011; Amato et al. Reference Amato, Toro, Akinin, Hampton, Salnikov and Tuchkova2015) and Wrangel Island (Miller et al. Reference Miller, Gehrels, Pease and Sokolov2010) resemble the Romulus Member detrital-zircon age spectrum (Figure 4). A detrital-zircon sample from the Mendeleev Rise is similar to upper Triassic samples from Chukotka and Wrangel Island, but with a greater proportion of near-depositional detrital-zircon ages (Tuchkova et al. Reference Tuchkova, Shokalsky, Petrov, Sokolov, Sergeev and Moiseev2020). The Permo-Triassic detrital-zircon age peaks are also found in upper Triassic strata on the Barents Sea side of the Amerasia Basin (Klausen et al. Reference Klausen, Muller, Slama and Helland-Hansen2017), in Svalbard (Bue & Andresen, Reference Bue and Andresen2014) and Taimyr (Zhang et al. Reference Zhang, Pease, Skogseid and Wohlgemuth-Ueberwasser2016), where they persist through Jurassic strata. The presence of abundant Permo-Triassic detrital zircon in Jurassic strata on Svalbard and near absence in the Sverdrup Basin indicates that the Sverdrup Basin was no longer in sedimentary communication with Svalbard or the Barents shelf in the Early Jurassic (Figure 5). The comparison of detrital-zircon data does not eliminate the transport of sediment from the Sverdrup Basin to the Barents region, but it does preclude the transport of sediment from the Barents region to the Sverdrup Basin. The interpretation of detrital zircon in the Barents shelf region is of provenance from the northern Urals (e.g., Miller et al. Reference Miller, Soloviev, Prokopiev, Toro, Harris, Kuzmichev and Gehrels2013; Klausen et al. Reference Klausen, Muller, Slama and Helland-Hansen2017; Klausen et al. Reference Klausen, Rismyhr, Muller and Olaussen2022), and so if the Sverdrup Basin received such sediment in the Late Triassic, this flux had ceased by the Early Jurassic.

Figure 5. The early Mesozoic position of Svalbard relative to the Sverdrup Basin, based on the palaeogeographic reconstruction of the Amerasia Basin by Grantz (2011). Detrital-zircon U-Pb age spectra from Lower Jurassic strata of the Sverdrup Basin are compared to Svalbard, note the difference in the Permian-Triassic interval (Sverdrup Basin, Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016, and this study; Bue & Andresen, Reference Bue and Andresen2014).

It is possible that part of the provenance change could be due to a cessation of magmatism in the source area, in which case, the 350–215 Ma age fraction would still be present. The virtual absence of the 350–215 Ma fraction indicates that there was an extra-basinal change in sediment routing patterns and even precludes recycling of upper Triassic deposits from the Sverdrup, Barents, Taimyr and the Arctic-Alaska microplate.

Previous work discussed how the Heiberg Formation/Group marks a tectonic transition in the Sverdrup Basin in terms of basin-filling patterns, normal faults and detrital-zircon provenance (Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016; Midwinter et al. Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016). From new U-Pb detrital-zircon dating, we show that the transition occurs at the boundary between the Romulus and Fosheim members. We rely on previous biostratigraphy of the Romulus and Fosheim members to indicate that the timing of this transition is during the latest Triassic (latest Rhaetian). The detrital-zircon age signature of the Fosheim Member is distinct from upper Triassic strata of the Arctic Alaska – Chukotka microplate, and from Jurassic strata of Taimyr and Svalbard. We can therefore state that sediment from the extra-basinal source of 350–200 Ma detrital zircon had no longer reached the Sverdrup Basin in the lower Jurassic. Based on the similarity to the Devonian clastic wedge, we interpret that the detrital-zircon provenance of the Fosheim Member was sedimentary recycling from erosion of relatively local sedimentary basins, such as the Franklinian Basin, which underlies the Sverdrup Basin (cf. Hadlari et al. Reference Hadlari, Swindles, Galloway, Bell, Sulphur, Heaman, Beranek and Fallas2015). Furthermore, we infer that the provenance transition was a response to regional extension (see Hadlari et al. Reference Hadlari, Midwinter, Galloway, Dewing and Durbano2016), which disrupted sediment transport pathways by the formation of extensional basins that trapped sediment, acting as a sediment sink (cf. Withjack et al. Reference Withjack, Schlische and Olsen2002; Oliebrook et al. Reference Oliebrook, Barham, Fitzsimons, Timms, Jiang, Evans and Mcdonald2019). Between the newly-developed proto-Amerasia Basin and the Sverdrup Basin was a palaeo-high, the Sverdrup Rim (Meneley et al. Reference Meneley, Henao, Merritt and Yorath1975), which was a horst that likely prevented sediment from entering the northern part of the Sverdrup Basin and likely acted as a source of sediment for both basins (Figure 6). We interpret that by formation of localized sediment sinks and barriers, the tectonic localization of provenance thereby prevented sediment from the extra-basinal igneous source from reaching the Sverdrup Basin.

Figure 6. Tectonic environment of northern Laurentia adapted from palaeogeographical maps of Plafker & Berg (Reference Plafker, Berg, Plafker and Berg1994), Nokleberg et al. (Reference Nokleberg, Parfenov, Monger, Norton, Chanďuk, Stone, Scotese, Scholl and Fujita2000), and Midwinter et al. (Reference Midwinter, Hadlari, Davis, Dewing and Arnott2016).

6. Conclusion

The upper Triassic to lower Jurassic Heiberg Formation records significant change within the Sverdrup Basin. During deposition of the Romulus Member, the Sverdrup Basin received a large amount of sediment from an extra-basinal igneous source to the northwest characterized by Carboniferous-Permian-Triassic detrital zircon, and the basin was effectively filled before the end of the late Triassic. Detrital zircon from the Fosheim Member has an age signature that is consistent with an older recycled and more proximal sediment source, similar to the Franklinian Basin, as opposed to a hypothetical Permo-Triassic magmatic arc. We attribute the provenance change from the Romulus to Fosheim members to the onset of extension and rifting that resulted in the initiation of the proto-Amerasia Basin, which acted as a sink and thereby prevented sediment from the northern extra-basinal source area from reaching the Sverdrup Basin.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S0016756824000050

Acknowledgements

This work was supported by the Geological Survey of Canada and the University of Ottawa. Fieldwork was assisted by the Polar Continental Shelf Program. The manuscript was reviewed by Dr. T.G. Klausen, Dr. J.T. Gooley and five anonymous reviewers.

Competing interests

The authors have no declarations of interest.

References

Amato, JM, Toro, J, Akinin, VV, Hampton, BA, Salnikov, AS and Tuchkova, MI (2015) Tectonic evolution of the Mesozoic South Anyui suture zone, eastern Russia: a critical component of paleogeographic reconstructions of the Arctic region. Geosphere 11(5), 15301564.CrossRefGoogle Scholar
Anfinson, OA, Embry, AF and Stockli, DF (2016) Geochronologic constraints on the Permian–Triassic northern source region of the Sverdrup Basin, Canadian Arctic Islands. Tectonophysics 691, 206219.CrossRefGoogle Scholar
Anfinson, OA, Leier, AL, Embry, AF and Dewing, K (2012a) Detrital zircon geochronology and provenance of the Neoproterozoic to Late Devonian Franklinian basin, Canadian Arctic Islands. Geological Society of America Bulletin 124(3-4), 415430.CrossRefGoogle Scholar
Anfinson, OA, Leier, AL, Gaschnig, R, Embry, AF and Dewing, K (2012b) U–Pb and Hf isotopic data from Franklinian Basin strata: insights into the nature of Crockerland and the timing of accretion, Canadian Arctic Islands. Canadian Journal of Earth Sciences 49(11), 13161328.CrossRefGoogle Scholar
Alonso-Torres, D, Beauchamp, B, Guest, B, Hadlari, T and Matthews, WA (2018) Late Paleozoic to Triassic arc magmatism north of the Sverdrup Basin in the Canadian Arctic: Evidence from detrital zircon U-Pb geochronology. Lithosphere 10(3), 426445.CrossRefGoogle Scholar
Balkwill, HR (1978) Evolution of Sverdrup Basin. American Association of Petroleum Geologists Bulletin 62, 10041028.Google Scholar
Bue, EP and Andresen, A (2014) Constraining depositional models in the Barents Sea region using detrital zircon U-Pb data from Mesozoic sediments in Svalbard. Geological Society, London, Special Publications 386, 261279.Google Scholar
Cawood, PA, Hawkesworth, C and Dhuime, B (2012) Detrital zircon record and tectonic setting. Geology 40, 875878.CrossRefGoogle Scholar
Coutts, DS, Matthews, WA and Hubbard, SM (2019) Assessment of widely used methods to derive depositional ages from detrital zircon populations. Geoscience Frontiers 10, 14211435.CrossRefGoogle Scholar
Dickinson, WR and Gehrels, GE (2009) Use of U–Pb ages of detrital zircons to infer maximum depositional ages of strata: a test against a Colorado Plateau Mesozoic database. Earth and Planetary Science Letters 288(1-2), 115125.CrossRefGoogle Scholar
Drachev, SS (2011) Tectonic setting, structure and petroleum geology of the Siberian Arctic offshore sedimentary basins. Geological Society of London Memoirs 35, 369394.CrossRefGoogle Scholar
Embry, AF (1983a) Stratigraphic subdivision of the Heiberg Formation, eastern and central Sverdrup Basin, Arctic Islands. Geological Survey of Canada Current Research 83-1B, 205213.CrossRefGoogle Scholar
Embry, AF (1983b) The Heiberg Group, western Sverdrup Basin, Arctic Islands. Geological Survey of Canada Current Research 83-1B, 381389.CrossRefGoogle Scholar
Embry, AF (1991) Mesozoic history of the Arctic Islands. In Inuitian Orogen and Arctic Platform: Canada and Greenland Volume 3 (ed Trettin, HP), pp. 369433. Geological Survey of Canada, Geology of Canada (DNAG).CrossRefGoogle Scholar
Embry, AF (2009) Crockerland–the source area for the Triassic to Middle Jurassic Strata of Northern Axel Heiberg Island, Canadian Arctic Islands. Bulletin of Canadian Petroleum Geology 57(2), 129140.CrossRefGoogle Scholar
Embry, AF and Beauchamp, B (2008) Sverdrup Basin. In The Sedimentary Basins of the United States and Canada (ed Miall, AD), pp. 451471. Amsterdam, The Netherlands: Elsevier.CrossRefGoogle Scholar
Embry, AF and Johannessen, EP (1993) T-R sequence stratigraphy, facies analysis and reservoir distribution in the uppermost Triassic-Lower Jurassic succession, western Sverdrup Basin, Arctic Canada. In Arctic Geology and Petroleum Potential (eds Vorren, TO), pp. 121146. Norwegian Petroleum Society, Special Publication no. 2.CrossRefGoogle Scholar
Embry, AF and Suneby, LB (1994) The Triassic-Jurassic boundary in the Sverdrup Basin, Arctic Canada. Canadian Society of Petroleum Geologists, Memoir 17, 858869.Google Scholar
Gilmullina, A, Klausen, TG, Dore, AG, Sirevaag, H, Suslova, A and Eide, CH (2022) Arctic sediment routing during the Triassic: sinking the Arctic Atlantis. Journal of the Geological Society 180(1): jgs2022-018.CrossRefGoogle Scholar
Gottlieb, ES, Meisling, KE, Miller, EL and Charles, G (2014) Closing the Canada Basin: Detrital zircon geochronology relationships between the North Slope of Arctic Alaska and the Franklinian mobile belt of Arctic Canada. Geosphere 10(6), 13661384.CrossRefGoogle Scholar
Grantz, A, Eittreim, S and Dinter, DA (1979) Geology and tectonic development of the continental margin north of Alaska. Tectonophysics 59(1), 263291.CrossRefGoogle Scholar
Grantz, A, Hart, PE and Childers, VA (2011) Geology and tectonic development of the Amerasia and Canada Basins, Arctic Ocean. Geological Society of London Memoirs 35, 771799.CrossRefGoogle Scholar
Hadlari, T, Dewing, K, Matthews, WA, Alonso-Torres, D and Midwinter, D (2018) Early Triassic development of a foreland basin in the Canadian high Arctic: Implications for a Pangean Rim of Fire. Tectonophysics 736, 7584.CrossRefGoogle Scholar
Hadlari, T, Midwinter, D, Galloway, JM, Dewing, K and Durbano, AM (2016) Mesozoic rift to post-rift tectonostratigraphy of the Sverdrup Basin, Canadian Arctic. Marine and Petroleum Geology 76, 148158.CrossRefGoogle Scholar
Hadlari, T, Midwinter, D, Poulton, TP and Matthews, WA (2017) A Pangean rim of fire: Reviewing the Triassic of western Laurentia. Lithosphere 9(4), 579582.Google Scholar
Hadlari, T, Swindles, GT, Galloway, JM, Bell, KM, Sulphur, KC, Heaman, LM, Beranek, LP and Fallas, KM (2015) 1.8 billion years of sedimentary recycling calibrates a refractory part of Earth’s sedimentary cycle. PLoS ONE 10, e0144727.Google Scholar
Houseknecht, DW and Bird, KJ (2011) Geology and petroleum potential of the rifted margins of the Canada Basin. Geological Society of London Memoirs 35, 509526.CrossRefGoogle Scholar
Klausen, TG, Muller, R, Slama, J and Helland-Hansen, W (2017) Evidence for Late Triassic provenance areas and Early Jurassic sediment supply turnover in the Berents Sea Basin of northerns Pangea. Lithosphere 9, 1428.CrossRefGoogle Scholar
Klausen, TG, Rismyhr, B, Muller, R and Olaussen, S (2022) Changing provenance and stratigraphic signatures across the Triassic–Jurassic boundary in eastern Spitsbergen and the subsurface Barents Sea. Norwegian Journal of Geology 102, 202205 Google Scholar
Lopez-Mir, B, Schneider, S and Hülse, P (2018) Fault activity and diapirism in the Mississippian to Late Cretaceous Sverdrup Basin: New insights into the tectonic evolution of the Canadian Arctic. Journal of Geodynamics 118, 5565.CrossRefGoogle Scholar
Matthews, WA and Guest, B (2017) A practical approach for collecting large-n detrital zircon U-Pb data sets by quadrupole LA-ICP-MS. Geostandards and Geoanalytical Research 41(2), 161180.CrossRefGoogle Scholar
Meneley, RA, Henao, D and Merritt, RK (1975) The northwest margin of the Sverdrup Basin. In Canada’s Continental Margins and Offshore Petroleum Exploration (eds Yorath, CJ), pp. 557587. Canadian Society of Petroleum Geologists, Special Publication no. 4.Google Scholar
Midwinter, D, Hadlari, T, Davis, WJ, Dewing, K and Arnott, RWC (2016) Dual provenance signatures of the Triassic northern Laurentian margin from detrital zircon U-Pb and Hf isotope analysis of Triassic-Jurassic strata in the Sverdrup Basin. Lithosphere 8(6), 668683.CrossRefGoogle Scholar
Midwinter, D, Hadlari, T and Dewing, K (2022) Upper Triassic to Lower Jurassic stratigraphy along the northeastern margin of the Sverdrup Basin, Axel Heiberg and Ellesmere islands, Nunavut: new data from measured sections. Geological Survey of Canada, Open File 8855, 17.Google Scholar
Miller, EL, Gehrels, GE, Pease, V and Sokolov, S (2010) Stratigraphy and U-Pb detrital zircon geochronology of Wrangel Island, Russia: Implications for Arctic paleogeography. American Association of Petroleum Geologists Bulletin 94(5), 665692.CrossRefGoogle Scholar
Miller, EL, Soloviev, AV, Prokopiev, AV, Toro, J, Harris, D, Kuzmichev, AB and Gehrels, GE (2013) Triassic river systems and the paleo-Pacific margin of northwestern Pangea. Gondwana Research 23, 16311645.CrossRefGoogle Scholar
Miller, EL, Toro, J, Gehrels, G, Amato, JM, Prokopiev, A, Tuchkova, MI, Akinin, VV, Dumitru, TA, Moore, TE and Cecile, MP (2006) New insights into Arctic paleogeography and tectonics from U-Pb detrital zircon geochronology. Tectonics 25(3), 119.CrossRefGoogle Scholar
Nokleberg, WJ, Parfenov, LM, Monger, JW, Norton, IO, Chanďuk, AI, Stone, DB, Scotese, CR, Scholl, DW and Fujita, K (2000) Phanerozoic tectonic evolution of the Circum-North Pacific. United States Geological Survey, Professional Paper no. 1626, 122 pp.CrossRefGoogle Scholar
Okulitch, AV, Compiler (1991) Geology of the Canadian Archipelago, Northwest Territories and North Greenland. Geological Survey of Canada, Map 1715A, scale 1:2,000,000.CrossRefGoogle Scholar
Oliebrook, HKH, Barham, M, Fitzsimons, ICW, Timms, NE, Jiang, Q, Evans, NJ and Mcdonald, BJ (2019) Tectonic controls on sediment provenance evolution in rift basins: Detrital zircon U-Pb and Hf isotope analysis from the Perth Basin, Western Australia. Gondwana Research 66, 126142.CrossRefGoogle Scholar
Omma, JE, Pease, V and Scott, RA (2011) U–Pb SIMS zircon geochronology of Triassic and Jurassic sandstones on northwestern Axel Heiberg Island, northern Sverdrup Basin, Arctic Canada. In Arctic Petroleum Geology (eds Spencer, AM, Gautier, D, Stoupakova, A, Embry, A & Sørensen, K), v. 35, pp. 559566. Geological Society of London Memoir.CrossRefGoogle Scholar
Piepjohn, K and Von Gosen, W (2018) Structural transect through Ellesmere Island (Canadian Arctic): superimposed Palaeozoic Ellesmerian and Cenozoic Eurekan deformation. Geological Society, London, Special Publications 460, 3356.CrossRefGoogle Scholar
Plafker, G and Berg, HC (1994) Overview of the geology and tectonic evolution of Alaska. In The Geology of Alaska (eds Plafker, G & Berg, HC), pp. 9891021. Geological Society of America, Geology of North America, v. G-1.CrossRefGoogle Scholar
Pointon, MA, Smyth, H., Omma, JE, Morton, AC, Schneider, S, Hulse, P, Rippington, SJ, Lopez-Mir, B, Crowley, QG, Millar, I, Whitehouse, MJ, Frei, D, Scott, RA and Flowerdew, MJ (2023) A multi-proxy provenance study of Late Carboniferous to Middle Jurassic sandstones in the Eastern Sverdrup Basin and it bearing on Arctic Paleogeographic reconstructions. Geosciences 13(10), 41.Google Scholar
Souther, JG (1963) Geological traverse across Axel Heiberg Island from Buchanan Lake to Strand Fiord, in Geology of the north central part of the Arctic Archipelago, Northwest Territories (Operation Franklin). Geological Survey of Canada Memoir 320, 426448.Google Scholar
Suneby, LB and Hills, LV (1988) Palynological zonation of the Heiberg Formation (Triassic-Jurassic) eastern Sverdrup Basin, Arctic Canada. Bulletin of Canadian Petroleum Geology 36(4), 347361 Google Scholar
Tuchkova, MI, Sokolov, SD, Khudoley, AK, Verzhbitsky, VE, Hayasaka, Y & Moiseev, AV (2011) Permian and Triassic deposits of Siberian and Chukotka passive margins: sedimentation setting and provenances. Proceedings of the International Conference on Arctic Margins VI, Fairbanks, Alaska, 61-96.Google Scholar
Tuchkova, MI, Shokalsky, SP, Petrov, OV, Sokolov, SD, Sergeev, SA and Moiseev, AV (2020) Triassic deposits of Chukotka, Wrangel Island and Mendeleev Rise, Arctic Sea: sedimentology and geodynamic implications. GFF (Journal of the Geological Society of Sweden) 142(2), 158168.Google Scholar
Withjack, MO, Schlische, R and Olsen, PE (2002) Rift-basin structure and its influence on sedimentary systems, in sedimentation in continental rifts. SEPM Special Publication 73, 5781.Google Scholar
Zhang, X, Pease, V, Skogseid, J and Wohlgemuth-Ueberwasser, CW (2016) Reconstructions of tectonic events on the northern Eurasian margin of the Arctic, from U-Pb detrital zircon provenance investigations of late Paleozoic to Mesozoic sandstones in southern Taimyr Peninsula. GSA Bulletin 128, 2946.Google Scholar
Figure 0

Figure 1. (a) Map of the Sverdrup Basin showing the location of detrital-zircon samples from the Heiberg Formation/Group. Surface and sea bottom bedrock geology is from Okulitch (1991). Detrital-zircon sample locations are from this study, Midwinter et al. (2016), Hadlari et al. (2018). See Figures 2 and 3 for stratigraphic locations of samples. (b) Map of the Arctic, depicting the basins and regions, mentioned in the text. Outline of AACM (black line) after Drachev (2011). Rotation of AACM away from the Canadian margin illustrated by black arrows (after). Red dashed line denotes location of Figure 1a. Red circle indicates the approximate location of Mendeleev rise from the Shamshur Seamount sample collected by Tuchkova et al. (2020). AA, Arctic Alaska; AACM, Arctic Alaska – Chukotka Microplate; Ch, Chukotka; M, Mendeleev rise; WI, Wrangel Island.

Figure 1

Figure 2. Triassic to Early Cretaceous stratigraphic chart of the Sverdrup Basin (Embry & Beauchamp, 2008; Hadlari et al.2016) with location and reference of detrital-zircon samples with respect to stratigraphy.

Figure 2

Figure 3. Measured sections (a) with locations of detrital-zircon data (b) presented as normalized probability density functions. Sedimentological data and exact locations of the Blue Mountains and Depot Point sections can be found in Midwinter et al. 2021. All U-Pb data from this study, except for the Romulus Member sample from Depot Point (Midwinter et al.2016).

Figure 3

Figure 4. Compiled normalized probability (a) and cumulative probability (b) plots with corresponding number of analyses. Data sources: the Devonian clastic wedge reference spectrum (Anfinson et al.2012a, 2012b); Triassic sediment of Svalbard (Bue & Andresen, 2014); Triassic sediment of Chukotka (Miller et al.2006; Tuchkova et al.2011; Amato et al.2015); Triassic sediment of Wrangel Island (Miller et al.2010); Romulus Mbr. (Midwinter et al.2016; Hadlari et al.2018; this study), Fosheim Mbr./King Christian Fm. (Midwinter et al.2016; this study). ROU: Rift onset unconformity.

Figure 4

Figure 5. The early Mesozoic position of Svalbard relative to the Sverdrup Basin, based on the palaeogeographic reconstruction of the Amerasia Basin by Grantz (2011). Detrital-zircon U-Pb age spectra from Lower Jurassic strata of the Sverdrup Basin are compared to Svalbard, note the difference in the Permian-Triassic interval (Sverdrup Basin, Midwinter et al.2016, and this study; Bue & Andresen, 2014).

Figure 5

Figure 6. Tectonic environment of northern Laurentia adapted from palaeogeographical maps of Plafker & Berg (1994), Nokleberg et al. (2000), and Midwinter et al. (2016).

Supplementary material: File

Midwinter et al. supplementary material 1

Midwinter et al. supplementary material
Download Midwinter et al. supplementary material 1(File)
File 1.2 MB
Supplementary material: File

Midwinter et al. supplementary material 2

Midwinter et al. supplementary material
Download Midwinter et al. supplementary material 2(File)
File 13.2 KB