<<

Late Paleozoic to arc magmatism north of the Sverdrup Basin in the Canadian Arctic: Evidence from detrital zircon U-Pb

D. Alonso-Torres1,*, B. Beauchamp1,*, B. Guest1,*, T. Hadlari2,*, and W. Matthews1,* 1DEPARTMENT OF GEOSCIENCE, UNIVERSITY OF CALGARY, 2500 UNIVERSITY DR. NW, T2N 1N4, CALGARY, ALBERTA, 2GEOLOGICAL SURVEY OF CANADA, 3303 33RD STREET NW, T2L 2A7, CALGARY, ALBERTA, CANADA

ABSTRACT

Paleozoic and Mesozoic tectonic reconstructions of the Arctic regions have been a subject of debate in recent . The emer- gence of a landmass north of the Sverdrup Basin in the Canadian Arctic led to the shedding of northerly derived detritus, an event that followed volcanism and basin inversion pulses that began in the late . However, the mechanisms for these events and the Paleozoic to Mesozoic paleogeography of this region remain controversial. New detrital zircon U-Pb geochronology results from Permian to Lower Triassic strata from northern and Ellesmere islands constrain the magmatic events within this northern landmass and its implications for the tectonic regime of the Sverdrup Basin and adjacent domains. Permian to lowermost Triassic strata along the northern margin of the Sverdrup Basin contain zircons derived from to rocks (420–350 Ma), Timanian-aged basement (700–500 Ma), and a Permian syndepositional source (300–250 Ma). Coeval strata in the southern margin are dominated by zircons formed during the Taconic, Scandian, and post-Scandian phases of the Appalachian and Caledonian orogenies, respectively (480–400 Ma). The detrital zircon signatures of the analyzed strata on the northern margin of the Sverdrup Basin record continuous magmatism within the northern landmass from latest (ca. 300 Ma) to at least earliest Triassic (ca. 250 Ma) time. These results are indicative of ongoing subduction and development of a magmatic arc off the northern margin of Laurentia, with the Sverdrup Basin potentially located in the backarc region of a proto-Pacific convergent margin involving parts of Arctic Alaska, Chukotka, and the Chuk- chi Shelf. The hypothesized onset of subduction in latest Carboniferous time and closure of this backarc basin in the latest Permian to earliest Triassic provides an explanation for the shift in stress regimes in the Sverdrup Basin that led to basin inversion and volcanism episodes. Therefore, the data presented here supports a backarc to retroarc setting for the Sverdrup Basin and the possibility of a con- vergent margin regime for the northern edge of Laurentia during the late Paleozoic to Triassic, contrasting with the generally accepted rift and passive margin settings.

LITHOSPHERE; v. 10; no. 3; p. 426–445; GSA Data Repository Item 2018093 | Published online 23 March 2018 https://doi.org/10.1130/L683.1

INTRODUCTION these terranes to the continental margins, aiding to unravel the geological histories of the different circum-Arctic domains. Recent detrital zircon studies in the Arctic regions have shed light The late Paleozoic evolution of the Sverdrup Basin in the Canadian on sediment provenance, paleodrainage systems, and terrane positions Arctic is marked by poorly understood basin inversion events and volca- throughout the Paleozoic and Mesozoic (e.g., Anfinson et al., 2012a, nism, coupled with the onset of northerly derived siliciclastic sediments; 2012b, 2016; Hadlari et al., 2014; Beranek et al., 2015; Midwinter et all of these potentially related to the emergence of an enigmatic land- al., 2016). However, despite recent advances in our understanding of the mass, termed “Crockerland” by Embry (1993), adjacent to the northern tectonic evolution of the Arctic region, uncertainties on the affinities, margin of the basin during the Middle Permian (Embry, 1993; Harrison, operating tectonic regimes, and geographic location of terranes during the 1995; Beauchamp et al., 2001; Embry and Beauchamp, 2008). The term Paleozoic and Mesozoic remain a major challenge for accurate paleogeo- “Crockerland” is herein used to refer to the hypothesized landmass that graphic reconstructions. Sedimentary basins throughout the Arctic provide provided a source of northwesterly derived sediments to the Sverdrup a valuable record of the tectonic setting, crustal affinity, and proximity of Basin from the Middle Permian to the , nowadays located within the continental shelf of the Chukchi Borderlands (Embry, 1993, 2009). Similarly, another hypothetical landmass comprising an amalgamation of Arctic terranes was proposed by Zonenshain (1990), recognized as *Emails: Alonso-Torres: [email protected]; Beauchamp: bbeaucha@ ucalgary​.ca; Guest: [email protected]; Hadlari: thomas.hadlari@canada​ Arctida. In this study, we use detrital zircon U-Pb geochronology on .ca; Matthews: [email protected] Carboniferous to Triassic strata along the northern and southern margins

426 © 2018 The Authors. Gold Open Access: Thiswww.gsapubs.org paper is published | underVolume the 10 terms | Number of the 3CC-BY-NC | LITHOSPHERE license.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

of the Sverdrup Basin (Fig. 1) to investigate their sediment sources and unexplained and its potential linkage to the inversion of rift structures in their links to the emergence of Crockerland in the Middle Permian, with the Sverdrup Basin not established. These basin inversion events led to the the purpose of constraining the tectonic setting of the Sverdrup Basin and partial exhumation of rift-related depocenters and generation of several adjacent domains during the late Paleozoic to Mesozoic. unconformities, two of which are angular, beneath Middle Permian and Detrital zircon U-Pb geochronology studies in the Canadian Arctic and uppermost Permian–Lower Triassic strata. The pre-Middle Permian basin Arctic Alaska have shed light into the crustal affinity of Crockerland and inversion events are collectively termed the “Melvillian Disturbance” the provenance of sub-Carboniferous and Mesozoic terrigenous sediments (Thorsteinsson and Tozer, 1970; Harrison, 1995; Beauchamp et al., 2001) (Miller et al., 2006, 2013; Beranek et al., 2010; Lemieux et al., 2011; with structural and stratigraphic evidence of compressional deformation Omma et al., 2011; Anfinson et al., 2012a, 2012b, 2013, 2016; Midwinter observed all along the southern and northern margins of the Sverdrup et al., 2016). Anfinson et al. (2012a) suggested a Timanian and Caledonian Basin. Evidence for the unnamed pre-latest Permian– event origin for Crockerland, based on the predominance of 700–500 Ma and is essentially from northern (Mayr et al., 2002), and 450–365 Ma detrital zircon sub-populations found in northerly derived possibly from the Tanquary Structural High on central Devonian strata in the Franklinian Basin, the Neoproterozoic to Devonian (Maurel, 1989; Trettin and Mayr, 1997). predecessor of the Sverdrup Basin. Anfinson et al. (2012b) determined The U-Pb age data set presented here is used to analyze variations in that the 450–365 Ma magmatism on Crockerland had a juvenile signa- sediment sources and pathways across major Permian unconformities ture, which they interpreted as evidence of ongoing subduction beneath in the Sverdrup Basin, and to determine their relationship to the Penn- Crockerland prior to its accretion to the Laurentian margin in the Late sylvanian to Permian pulses of basin inversion, volcanism, and onset Devonian. The emergence of Crockerland in the Middle Permian remains of northerly derived clastic sedimentation. These results contribute to

75ºN 80ºN 85ºN W 120º W 110ºW 100º 80ºW

Brock C-50 Middle Permian Facies Map W 120º NorthernAxel ? Heiberg Island Melville Island (SvB-01 to SvB-03 and 85ºN Isachsen J-37 SvB-06 to SvB-13) ? ?

110ºW Axel Borup Fiord Pass Heiberg (SvB-04 to SvB-05) Island

d lan e I s 180º Ellesmer 150 150º Chukotka º Arctic Legend Alaska WI 80ºN

120º Outcrop and subcrop limit of the Sverdrup Basin 0º Chukchi 12 Borderland NSI Clastic outcrop limit Amerasia Interpreted clastic deposition limit Basin Taimyr SZ 90ºW 90ºE Siliciclastic sediments

85º FJL Carbonate and chert to mudrock Svalbard Greenland NZ 80º Study Areas and sample locations 60º Barents º Sea 60 75º Well locations

70º 3 Map location (Figure 3) 0º 30º

Figure 1. Arctic region map with delineated coastlines and facies map of the Middle Permian sequence of the Sverdrup Basin with study areas and sample locations. The black square on northern Axel Heiberg Island represents the area displayed on the map of Figure 3. FJL—Franz Josef Land; NZ—; SZ—; NSI—; WI—.

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 427

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

unraveling the tectonic evolution of the Canadian Arctic and explore the depocenters and regional unconformities along both northern and southern timing of magmatism and its implications for the geodynamic significance basin margins (Stephenson et al., 1987; Beauchamp et al., 2001; Embry of Crockerland during the late Paleozoic to earliest Mesozoic. and Beauchamp, 2008). The first major unconformity resulting from these episodes is the sub- GEOLOGICAL BACKGROUND Roadian (sub-early Middle Permian) unconformity, recognized along both basin margins and in the subsurface (Thorsteinsson, 1974; Beauchamp et The late Paleozoic history of the Canadian Arctic is characterized by al., 2001). Overlying this unconformity on the northern margin is a vari- sedimentation within a NE to SW elongated depocenter: the Sverdrup Basin ably thick, northerly derived clastic wedge including the Assistance and (Fig. 1). This basin has been shown to have originated as a rift, presently Trold Fiord formations, recording the first significant clastic input from spanning over an area of 300,000 km2, preserving 13 km of Crockerland (Thorsteinsson, 1974; Embry, 1993, 2009; Beauchamp et to strata disrupted by numerous deformation events and uncon- al., 2009; Beauchamp, 2015). Similar Middle Permian deformation and formities (Balkwill, 1978; Trettin, 1991; Embry and Beauchamp, 2008). a transition from carbonate platform to siliciclastic sedimentation is also The Sverdrup Basin is the successor to the Franklinian Basin, which documented in the northern Barents Sea and Svalbard within the Tem- preserves a Neoproterozoic to Upper Devonian sedimentary record along pelfjorden Group (Worsley, 2008), suggesting that these processes were the northern edge of Laurentia (Balkwill, 1978). By the Silurian, Pearya regional in extent and not restricted to the Sverdrup Basin. had collided against the northeastern portion of the Franklinian Basin, The second major unconformity is the sub- (sub-late resulting in the development of the Clements-Markham Fold Belt and Late Permian) angular unconformity (Fig. 2). On northern Axel Heiberg the deposition of flysch units as an overlap or a syn-collisional sequence Island, folded upper Paleozoic strata are truncated and overlain by homo- (Trettin, 1987, 1998; Klaper, 1992; Hadlari et al., 2014; Beranek et al., clinal shales and of the uppermost Permian to Lower Triassic 2015). Collision led to deepening of the northeastern portion of the basin, Blind Fiord Formation (Fig. 3) (Mayr et al., 2002). On the southern margin which was filled by northerly sourced flysch units of Silurian age found in of the Sverdrup Basin, lowermost Triassic strata of the Bjorne Formation northern Ellesmere Island (Beranek et al., 2015). By the Middle Devonian record cratonic uplift that was synchronous with deformation events and Crockerland approached the Franklinian Basin, as evidenced by renewed uplifts on the northern margin (Embry and Beauchamp, 2008). deepening and widespread northwesterly derived sedimentation through- Crockerland, which remained exposed throughout the Triassic to Mid- out the basin (Trettin, 1998). Final accretion of Crockerland against the dle Jurassic, underwent rifting and ultimately separated from the Lauren- Laurentian margin led to the latest Devonian Ellesmerian Orogeny and tian margin by the , leading to the opening of the Amerasia the deposition of molasse sediments, which blanketed northern Lauren- Basin (Embry, 1993, 2009; Hadlari et al., 2016). Due to this event, the tia and are preserved as the Devonian Clastic Wedge (Thorsteinsson and landmass referred to as Crockerland lies buried beneath post-Jurassic Tozer, 1957; Embry, 1988, 1991; Patchett et al., 2004). strata, along with other terranes included within Arctic Alaska–Chukotka The Ellesmerian Orogen experienced post-collisional collapse in the Microplate (Embry, 1993, 2009; Pease, 2011). Tectonic movements earliest Carboniferous, during which compressional structures were reacti- between Greenland and North America during the caused the vated in an extensional regime (Harrison, 1995). This yet-to-be-understood , which deformed and uplifted the Sverdrup Basin and shift in stress directions produced a depocenter in which syn-rift sedimen- led to its demise as a sedimentary basin (Embry and Beauchamp, 2008). tation dominated, representing the first tectonostratigraphic phase of the Detrital zircon studies in the Canadian Arctic Islands have focused on Sverdrup Basin (Embry and Beauchamp, 2008). This first phase recorded Neoproterozoic to uppermost Devonian strata of the Franklinian Basin three major pulses of rifting, coinciding with the Viséan, Serpukhovian, (McNicoll et al., 1995; Anfinson et al., 2012a, 2012b, 2013; Hadlari et and Bashkirian sequences (Fig. 2) associated with deposition of syn- al., 2014; Beranek et al., 2015), Neoproterozoic to Silurian units on the rift clastic, carbonate, and evaporite sediments covering an increasingly Pearya terrane (Hadlari et al., 2014; Malone et al., 2014) and Carbonif- broader area with each subsequent pulse (Embry and Beauchamp, 2008). erous syn-rift formations of the Sverdrup Basin (Malone, 2012) and its Post-rift quiescence and thermal subsidence prevailed during most Mesozoic succession (Miller et al., 2006; Omma et al., 2011; Røhr et al., of the Pennsylvanian and Permian, punctuated by at least three basin 2010; Anfinson et al., 2016; Midwinter et al., 2016). Related detrital zircon inversion episodes from the latest Carboniferous to the early Permian studies that help constrain the sediment sources and tectonic evolution of (Fig. 2), collectively recognized as the Melvillian Disturbance (Thorsteins­ the rocks in the Arctic realm were conducted in northernmost Laurentia son and Tozer, 1970; Stephenson et al., 1987; Beauchamp et al., 2001). and Greenland (Kirkland et al., 2009; Beranek et al., 2010; Lemieux et The second tectonostratigraphic phase is characterized by these pulses al., 2011; Hadlari et al., 2012) Arctic Alaska (Miller et al., 2006; Amato of renewed tectonic activity with synchronous basaltic magmatism and et al., 2009; Strauss et al., 2013; Gottlieb et al., 2014; Till et al., 2014), deposition of volcaniclastic material, and followed by a marine trans- and the Barents Sea (Pettersson et al., 2009, 2010; Bue and Andresen, gression (Harrison, 1995; Beauchamp et al., 2001; Embry and Beau- 2014; Klausen et al., 2017). The Russian margin has also been the subject champ, 2008). Growth faulting along reactivated Ellesmerian structures of numerous detrital zircon studies. Franz Josef Land, the peri-Siberian in the Franklinian Basement, broad scale folding, local thrust faulting and islands, Novaya Zemlya, Taimyr Peninsula, the Verkhoyansk Margin, regional uplift, are all products of what Harrison (1995) interpreted to be Chukotka, and the Chukchi Shelf have been extensively studied using a single southeast-directed compressional episode. Multiple deformation detrital zircon geochronology in recent years (Miller et al., 2006, 2010; events from the (late Late Pennsylvanian) to the Changhsingian Lorenz et al., 2008, 2013; Pease and Scott, 2009; Tuchkova et al., 2011; (late Late Permian) were later recognized by Beauchamp et al. (2001) Soloviev and Miller, 2014; Ershova et al., 2015a, 2015b, 2015c, 2016, and Beauchamp (2015), some involving half-graben inversion and normal 2017; Amato et al., 2015; Pease et al., 2015; Soloviev et al., 2015; Brumley faulting, suggesting more complex, possibly transpressive to transten- et al., 2015; O’Brien et al., 2016). However, despite the increasing num- sive, stress conditions during these pulses (Fig. 2). As a result of these ber of detrital zircon studies in the Arctic realm, mostly focusing on the deformation episodes, passive thermal subsidence of the basin during Mesozoic and Devonian, the Carboniferous to Permian paleogeography the Carboniferous and Permian was overwhelmed by short episodes of of the northern Laurentian margin remains a knowledge gap crucial for fault-controlled subsidence, leading to the creation of numerous local understanding any subsequent plate configurations.

428 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

Southern Northern Basin Margin Margin Centre (Northern Ellesmere (Northern Axel Heiberg Island) Island)

Olenekian Blind Fiord Bjorne 250 riassi c

SvB-13 Lowe r Blind Fiord T Induan SvB-12 Changhsingian SvB-11 Black Stripe SvB-10 Lindström TF TF Degerböls 260 SvB-09 TF Degerböls Degerböls TF

Van Hauen SvB-08 TF Degerböls Degerböls TF Assistance Assistance 270 SvB-07 SvB-05 Sabine Bay Roadian v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v SvB-06 v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v Esayoov v v v v v v v v v v v v v v v v vFm v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v SvB-04 v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v v vEsayoo v v v v v v v vFm v v v v v v v

Permian Great v v v v v v v v v v v v v v v v v v GB C Bear Cape 280 Trappers Cove

Great Bear Cape Great Bear Cape

v v v v v v v v v v v v v v v v SvB-03 v v v v v v v v v v v v v v v v 290 v v ULv v vV v v v v v v v v v v v Raanes Raanes Hare Fiord SvB-02 (upper) Nansen UCM SvB-01 v v v v v v v v v v v v v v v v v v v v v v v v v v v v Belcher 300 Unnamed Channel Gzhelian

Hare Fiord Kasimovian (lower) UCM

Belcher 310 Nansen Channel Moscovian MCM ^^^^^^^^ ^ ^^^^^^^^ Pennsylvanian ^^^^^^^^^Otto ^^^^^^^^Fiord ^ ^^^^^^^^^ LCM Bashkirian ^ ^^^^^^^^ 320 ^^^^^^^^^ Audhild Unnamed v v v v v v v v v v v v v v v v vvvvvvvvvvvvvvvv Borup Fiord Borup Fiord Serpukhovian Borup Fiord 330 ^^^^^

Conglomerate ^^^^^ Evaporite Carboniferou s ^^^^^ White chert Emma Fiord Viséan v v v v v v v v v v v v v v v v v v v v 340 Limestone v v v v v v v v v v Volcanics

Non-marine shale Non-siliceous shale Mississippian

Calcareous mudrock Black, siliceous shale

Rifting event Basin inversion event (interpreted stress (interpreted stress orientations) orientations)

Figure 2. Upper Paleozoic to Lower Triassic stratigraphy and tectonic regime of the Sverdrup Basin (adapted from Embry and Beauchamp, 2008). Detrital zircon sample position indicated with the black arrowheads. Deformation events and pro- posed strain ellipsoids and orientations in map view for both basin margins from field and seismic data (Harrison, 1995; Beauchamp et al., 2001; Beauchamp, 2015). Deformation in the basin center has not been studied due to the continuous record and lithological homogeneity of the units and the associated distinction difficulties between formations in the field. GBC—Great Bear Cape; LCM—Lower Canyon Fiord member; MCM— Middle Canyon Fiord member; TF—Trold Fiord Forma- tion; UCM— Upper Canyon Fiord member; ULV—Unnamed lower volcanics (Morris, 2013).

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 429

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

ž : ž 1 ž : ž 1

$5&7,&2&($1

1SO3

(//(60(5(

,6/$1' 4 1$16(16281'

1LFH9DOOH\ 6Y% 5LYHU 6HFWLRQ 6Y% X3 )ROGV&UHHN 6Y%

%8..(1 %XNNHQ ),25' 5LYHU 6Y% &RQJORPHUDWH 6HFWLRQ 6Y%

%81'(),25' 0 *ULHVEDFK &UHHN 6Y%

ž 1 ž 1 ž : ž :

8SSHU3DOHR]RLF ,FHFRYHU X3 &DUERQLIHURXV3HUPLDQ

1HRSURWHUR]RLF/RZHU3DOHR]RLF 4 4XDWHUQDU\ 1SO3 1HRSURWHUR]RLF'HYRQLDQ

0 0HVR]RLF 6DPSOHORFDWLRQ

Figure 3. Simplified geological map of northern Axel Heiberg Island (modified from Mayr et al., 2002). The red stars mark the field localities referred to in this study, sample locations indicated.

430 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

ANALYTICAL METHODS (Mortensen and Card, 1993), NIST SRM 610 Glass (Stern and Amelin, 2003), FCT (Kuiper et al., 2008), and TRD (Ganerød et al., 2011) were Sample Preparation ablated every 20 unknowns in order to assess the accuracy and uncertainty of the data. Isotopic signal intensities were measured using an Agilent A total of 19 samples were collected during several field seasons, 7700 quadrupole ICP-MS, and the data reduction handled in Iolite v. 2.5 catalogued and archived at the Geological Survey of Canada (GSC) in (Paton et al., 2010). Subsequent data filtering and uncertainty propaga- Calgary, Alberta. The samples were processed to recover detrital zircons tion was carried out in Excel using a custom VBA macro (ARS4.0) and following procedures outlined in Matthews and Guest (2016). Samples the Isoplot plug-in of Ludwig (2012). were crushed using a BicoTM Chipmunk jaw crusher, and pulverized using Measurements were filtered using the probability of fit parameter from a BicoTM disk mill. Large samples were processed using a MD Gemini the concordia age algorithm in Isoplot (Ludwig, 1998). Measurements water table to separate dense minerals (e.g., zircon and apatite) from with <1% probability of fit were omitted from the data set. Probability less-dense minerals (e.g., quartz and feldspar). Smaller samples were density plots were constructed from Best Ages calculated using both ran- rinsed with water in a beaker to remove the clay fraction. The dense dom and systematic components of uncertainty with the cutoff between mineral fraction was further concentrated using methylene iodide and the 206Pb/238U and 207Pb/206Pb dates at 1550 Ma, following recommendations resulting dense fraction underwent magnetic separation using a Frantz™ by Spencer et al. (2016). isodynamic separator. A random population of zircons was cast in epoxy pucks, ground to expose the interior of the grains, and polished using SAMPLES AND RESULTS silicon carbide and diamond polishing films. All samples were cleaned

in diluted HNO3 prior to analysis to reduce surface Pb contamination. The 19 samples analyzed in this study are displayed as 11 samples and 2 composite samples, combined after analysis. The samples are labeled U-Pb Geochronology as SvB-# (Table 1).

U-Pb isotopic measurements were obtained using laser ablation– Pennsylvanian (Northern Margin) inductively coupled plasma–mass spectrometry (LA-ICP-MS) tech- niques at the Centre for Pure and Applied Tectonics and Thermochro- Sample SvB-01 was collected at River Section, northern Axel Heiberg nology (CPATT) at the University of Calgary following the methodology Island (Fig. 3), from red-weathering fluvial to marginal marine sand- described in Matthews and Guest (2016). Instrument settings are listed stones (Fig. 4A) within the upper Nansen Formation, just beneath an in the Table of Acquisition Parameters in the Data Repository1. Zircons unnamed intra-Nansen basalt (Figs. 2 and 4A). Fusulinaceans recovered were ablated using an ASI ResochronTM excimer laser ablation system immediately below and above this basalt suggest a Gzhelian (late Penn- and a beam diameter of 22 or 33 microns, chosen to fit the majority of the sylvanian) age for both the volcanic rocks and the sandstone unit (Fig. grains in the mount to minimize grain size bias. The well-characterized 2), (D. Baranova, personal commun., 2016). FC-1 zircon (Paces and Miller, 1993) was used as the calibration refer- Sample SvB-01 yielded 257 concordant dates dominated by 450– ence material and six validation reference materials, the 91500 zircon 400 Ma, 1050–1000 Ma, and 1650–1600 Ma sub-populations (Fig. 5). (Wiedenbeck et al., 1995, 2004), Temora 2 (Black et al., 2004), 1242 These groupings do not form isolated peaks but are part of two larger

TABLE 1. DETAIL OF THE SAMPLES ANALYZED FOR DETRITAL ZIRCON GEOCHRONOLOGY IN THIS STUDY DZ Sample # Sequence Coordinates (WGS84) Location sample # SvB-1 DA-19 Gzhelian 80.875605,–94.178196River Section (Northern Axel Heiberg Island) SvB-2 DA-20 Asselian 80.817048,–94.001547 Folds Creek (Northern Axel Heiberg Island) SvB-3 DA-21 Sakmarian 80.817117,–93.999253 SvB-4 1-9-1 Kungurian 81.038514,–81.704574 1-16-1 81.038718,–81.707039 1-49-381. 038717,–81.707413 Borup Fiord Pass (Ellesmere Island) SvB-5 1-91-7 Roadian 81.038715,–81.707574 1-93-181.038805,–81.708488 1-99-181. 038637,–81.708484 SvB-6 DA-23 Guadalupian-Lopingian 80.816934,–94.013604Folds Creek (Northern Axel Heiberg Island) SvB-7 DA-22 Roadian 80.758492,–93.336064Bukken River (Northern Axel Heiberg Island) SvB-8 96-04/97-25 Guadalupian 80.760375,–93.349407Bukken River (Northern Axel Heiberg Island) SvB-9 DA-24 Guadalupian-Lopingian 80.998341,–92.085983 Nice Valley (Northern Axel Heiberg Island) SvB-10 DA-25 Lopingian 80.725273,–93.757603Conglomerate Section (Northern Axel Heiberg Island) SvB-11 C601095 Lopingian 80.494518,–94.515296Griesbach Creek (Northern Axel Heiberg Island) SvB-12 DA-26 Induan 80.816891,–94.013094 1997-26_01 80.816335,–94.018608 Folds Creek (Northern Axel Heiberg Island) SvB-13 1997-26_02 Induan/Olenekian 80.816341,–94.018611 1997-26_03 80.816344,–94.018616

1 GSA Data Repository Item 2018093, detrital zircon U-Pb LA-ICP-MS geochronology­ data, is available at http://www.geosociety.org/datarepository/2018, or on request from [email protected].

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 431

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

A B CPV PL PTF CPN SvB-01 PA SvB-05 Ki SB C P PE PGB

C D SvB-13 TBF SvB-05 SvB-12 ss PSB c SvB-06

SvB-04 GBC PE P

E bF TBF PL

rc TF SvB-07 SvB-08 P PA PGBC PGBC

G H

SvB-10 PGBC ?

un ? SvB-5 11

PTBF

432 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

Figure 4 (on previous page). Field photographs of the sample locations and their stratigraphic relations (CPN—Carboniferous-Permian Nansen Forma­

tion; CPv—Carboniferous-Permian volcanics; PGBC—Permian Great Bear Cape Formation; PE—Permian Esayoo Formation; PSB—Permian Sabine Bay

Formation; PA—Permian Assistance Formation; PTF—Permian Trold Fiord Formation; PL—Permian Lindström Formation; PTBF—Permian-Triassic Blind

Fiord Formation; TBF—Triassic Blind Fiord Formation). Arrows indicate the location and sample number (Table 1). (A) Red sandstones interfingering with the Carboniferous–Permian volcanics at the River Section locality, northern Axel Heiberg Island. (B) Borup Fiord Pass section, the yellow square represents the area shown in detail in C. Note how the unconformity at the base of the Sabine Bay Formation cuts deeper stratigraphically toward the northeast (right). Ki—Cretaceous intrusion. (C) Detail of the unconformable contact between the Esayoo Formation and the Sabine Bay Formation on top. (D) The Permian-Triassic Blind Fiord Formation overlies folded strata of the Great Bear Cape Formation through an unconformable contact at the Folds Creek locality. ss—red weathering, bivalve-rich sandstone; c—boulder-sized chert conglomerate. (E) Detail of the red Roadian (lower Middle Permian) conglomerates at the Bukken River section in northern Axel Heiberg Island. rc—Red weathering conglomerate. (F) Bukken River section, northern Axel Heiberg Island. (G) Detail of the conglomerates above the Great Bear Cape Formation at the Conglomerate Section locality on northern Axel Heiberg Island. (H) Griesbach Creek section, where the Permian–Triassic Blind Fiord lies unconformably on an unnamed Permian unit (un—Maroon conglomerates and sandstones).

 6Y% 

 0LGGOH3HUPLDQ   6DELQH%D\$VVLVWDQFH V   Q    6Y%

   /RZHU3HUPLDQ   (VD\RR)P V 

 Q  %RUXS)LRUG3DVV 6RXWKHUQPDUJLQ   6Y%     /RZHU3HUPLDQ *UHDW%HDU&DSH)P  V  Figure 5. Probability density plots and histo- Q  grams of Pennsylvanian to lower Permian strata  of the northern margin of the basin, and lower to Middle Permian strata of the southern mar-  gin. n—number of accepted analyses; s—number of samples. 5HODWLYH3UREDELOLW\ 

 6Y%

  /RZHU3HUPLDQ

  5DDQHV)P

 V   Q     6Y%

 8SSHU&DUERQLIHURXV 1RUWKHUQ0DUJLQ 1RUWKHU$[HO+HLEHUJ,VODQGQ

 

 8QQDPHGYROFDQLFV  V  Q      %HVW$JH 0D

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 433

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

550–400 Ma and 2000–900 Ma age distributions, with an accompanying Kungurian sample (SvB-04), with two dominant sub-populations from 2900–2400 Ma group. 450 to 400 Ma and 2100 to 900 Ma, and two smaller ones at 3200 to 2500 Ma and 700 to 500 Ma. Lower Permian (Northern Margin) Middle Permian (Northern Margin) Two Lower Permian samples were collected in the Folds Creek area on northern Axel Heiberg Island (Fig. 3). Sample SvB-02 is from a thin Wordian and Capitanian sedimentation occurred at a time of maximum conglomerate bed that lies unconformably on the Nansen Formation and base level with deposition of the -rich glauconitic sandstones beneath the Sakmarian (Early Permian) Raanes Formation, a stratigraphic of the Trold Fiord Formation along both basin margins (Thorsteinsson, position that suggests a late Asselian (early Early Permian) or early Sak- 1974; Embry and Beauchamp, 2008). marian age (Fig. 2). Sample SvB-03 is from a sandy conglomerate that Sample SvB-06 was collected at the Folds Creek section from a con- rests unconformably on carbonate rocks of the Raanes Formation and glomerate of uncertain age that lies between spiculitic chert of the Lower beneath spiculitic chert of the Great Bear Cape Formation, a stratigraphic Permian Great Bear Cape Formation and a thin bivalve-rich red sandstone position indicative of a late Sakmarian to early Artinskian (middle Early unit at the base of the Triassic Blind Fiord Formation. The conglomerate is Permian) age (Fig. 2). locally more than 2 m thick but pinches out laterally over short distances Sample SvB-02 yielded 236 concordant measurements with prominent (Fig. 4D). Pebble to cobble clasts in the conglomerate are composed groupings between 450 and 400 Ma, 1100 and 1000 Ma, and 1700 and almost exclusively of chert from the Great Bear Cape Formation. The 1600 Ma. Sample SvB-03 yielded a total of 252 concordant dates with conglomerate is likely the erosional product of Early to Middle Permian dominant sub-populations between 450 and 400 Ma, 650 and 550 Ma, uplift and erosion, preserved within incised valleys and associated with 1200 and 950 Ma, 1450 and 1350 Ma, and 1750 and 1550 Ma. The three the last pulse of the Melvillian Disturbance on northern Axel Heiberg grains younger than the stratigraphically constrained age of the sample Island (Fig. 2). A Late Permian age is less likely as it contains no clasts are considered to be a product of laboratory contamination or Pb loss. derived from the erosion of the Assistance or Trold Fiord formations. The 152 measurements from sample SvB-06 exhibit a dominant Lower Permian (Southern Margin) 700–350 Ma sub-population and less abundant 2000–900 Ma and 3300– 2500 Ma sub-populations than older strata in the basin (Fig. 6). A single Sample SvB-04 was collected at Borup Fiord Pass (Figs. 1 and 2) Cretaceous grain is regarded as either laboratory contamination, or to from a shale unit that lies on top of the Kungurian (latest Early Permian) have suffered significant Pb loss based on the stratigraphic constraints Esayoo basalt. In this section, Kungurian rocks unconformably overlie on the sample. Artinskian (middle Early Permian) carbonate rocks of the Great Bear Sample SvB-07 was collected from a thin, red-weathering chert pebble Cape Formation and are unconformably overlain by the Middle Permian conglomerate at the Bukken Fiord section that lies between spiculitic Sabine Bay and Assistance formations (Figs. 4B and 4C). The Esayoo chert of the Great Bear Cape Formation and maroon-weathering shale and Volcanics at Borup Fiord Pass comprise 75 m of basalts that are overlain siltstone of the Assistance Formation (Figs. 3 and 4E). The conglomerate by 3 m of shale, recording the maximum flooding surface that followed forms a thin yet laterally extensive unit that is likely Roadian and is thus extrusion of the lava (Morris, 2013). either of the same age or slightly younger than sample SvB-06 (Fig. 2). Sample SvB-04 yielded a total of 201 concordant dates, with two Sample SvB-07 yielded 241 concordant dates and shows a prominent dominant detrital zircon sub-populations ranging from 450 to 400 Ma 300–260 Ma sub-population, with the mode at 270 Ma, accompanied and from 2050 to 930 Ma, and two minor ones from 3250 to 2330 Ma by a 450–370 Ma sub-population and a 670–490 Ma sub-population and from 650 to 500 Ma. The relative abundance of these detrital zircon (Fig. 6). Smaller sub-populations yield dates between 2000 and 900 Ma sub-populations is similar to Pennsylvanian to Lower Permian strata from and 3500–2300 Ma. the northern margin of the basin (SvB-01, SvB-02). Sample SvB-08 was collected from the Trold Fiord Formation at the Bukken River section, where it lies above the Assistance Formation (Fig. Middle Permian (Southern Margin) 4F). At this section the Trold Fiord Formation comprises bioturbated brachiopod-rich, greenish gray glauconitic sandstone of shallow marine A total of five samples consisting of white- to tan-colored Roadian origin. The Trold Fiord Formation at Bukken River is either Wordian or (Middle Permian) sandstone were collected at the Borup Fiord Pass sec- Capitanian in age (Fig. 2). tion (Fig. 1), where unfossiliferous strata with minor amounts of chert Sample SvB-08 yielded 236 concordant dates that define a detrital zir- and carbonate rock fragments (Sabine Bay Formation) pass upward into con signature similar to that of SvB-07. Sample SvB-08 is dominated by a increasingly fossiliferous calcareous sandstones (Assistance Formation). 700–360 Ma sub-population and contains a less prominent 2000–900 Ma This succession rests upon the Esayoo Formation basalt and shale (Figs. and 3100–2700 Ma sub-population (Fig. 6). The Permian sub-population 4A and 4B) in a low-angle angular unconformity, marking the culmina- (300–260 Ma) described in sample SvB-07 is also present and centered at tion of the Early Permian deformation episodes in the southern margin 265 Ma, although it is less abundant in sample SvB-08 than in SvB-07. collectively known as the Melvillian Disturbance (Thorsteinsson, 1974; Sample SvB-09 was collected from the Trold Fiord Formation in an Beauchamp et al., 2001). During the Roadian (early Middle Permian) area called Nice Valley (Fig. 3), where greenish-gray glauconitic sand- the northern margin of the basin began receiving terrigenous clastic sedi- stones lie with mild angular unconformity directly on top of the Great ments from northern sources, first recorded as the correlative Assistance Bear Cape Formation. Another low-angle angular unconformity separates Formation on northern Axel Heiberg Island (Fig. 2). the Trold Fiord Formation from the overlying Blind Fiord Formation, Five Roadian samples from Borup Fiord Pass yielded nearly identical attesting for a latest Permian–earliest Triassic episode of uplift and ero- detrital zircon signatures and as such their results were combined into sion on northern Axel Heiberg Island (Fig. 2). The sandstones of the Trold composite sample SvB-05 (Fig. 5). In total, composite sample SvB-05 Fiord Formation contain abundant , indicative of a Wordian yielded 967 concordant dates, exhibiting a similar signature to that of the or Capitanian age (Beauchamp et al., 2009).

434 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

 6Y%  8SSHU3HUPLDQ   0DURRQVDQGVWRQH  V  Q  



 6Y%  8SSHU3HUPLDQ 6XE7ULDVVLFFRQJORPHUDWH  V  Q 

 

6Y%    0LGGOH3HUPLDQ 7UROG)LRUG)P  V   Q   

Figure 6. Probability density plots and histo- grams from the Middle to Upper Permian samples 6Y%  from northern Axel Heiberg Island analyzed in  0LGGOH3HUPLDQ this study. n—number of accepted analyses; s— 7UROG)LRUG)P number of samples.

   V 

 Q   5HODWLYH3UREDELOLW\



 6Y%

 0LGGOH3HUPLDQ "  5HGFRQJORPHUDWH V  Q 

  



 6Y%  0LGGOH3HUPLDQ " 6XE7ULDVVLFFRQJORPHUDWH   V  Q   

 %HVW$JH 0D

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 435

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

 6Y%     /RZHU7ULDVVLF %OLQG)LRUG)P ,QGXDQ  V  Q     Figure 7. Probability density plots and histograms  from the uppermost Permian Lower Triassic sam-  6Y% ples from northern Axel Heiberg Island analyzed /RZHU7ULDVVLF in this study. n—number of accepted analyses; %OLQG)LRUG)P EDVH  s—number of samples.  V  5HODWLYH3UREDELOLW\   

 Q    

 %HVW$JH 0D

Sample SvB-09 yielded 242 concordant dates with a dominant sub- between 700 and 500 Ma, 2000 and 900 Ma, and 3200 and 2500 Ma are population yielding dates between 650 and 350 Ma, and smaller sub- less common in sample SvB-11 than in other middle to upper Permian populations from 295 to 250 Ma and 2100 to 800 Ma. Archean grains are samples on the northern margin of the basin. The two Jurassic grains in scarce in the sample and appear distributed between 3000 and 2500 Ma. the samples are likely to be due to laboratory contamination or Pb loss. Sample SvB-12 was taken from a thin, red-weathering, bivalve-rich Upper Permian (Northern Margin) sandstone unit that is part of the Blind Fiord Formation, resting uncon- formably upon the Upper Permian conglomerate from which sample SvB- The Upper Permian sedimentary record on northern Axel Heiberg 06 was collected. This sandstone is a thin transgressive unit of inferred Island is scarce due to erosion beneath the sub-Changhsingian unconfor- latest Permian age that lies within a shale just below the maximum flood- mity (Mayr et al., 2002). Chert and sandstone rocks of the Degerböls and ing surface in the lower Blind Fiord Formation. Trold Fiord formations are irregularly exposed in the area, but are locally Sample SvB-12 yielded 263 concordant dates with prominent sub- absent or very thin in the north-central part of northern Axel Heiberg Island populations between 430 and 345 Ma (with a mode at 384 Ma) and 1950 (Fig. 3). Maroon conglomeratic and sandstone bodies of limited lateral and 900 Ma, and two minor sub-populations from 750 to 500 and 2850 to extent can be found on northern Axel Heiberg Island above the Great 2400 Ma (Fig. 7), similar to the upper Permian samples from the northern Bear Cape Formation and below the Blind Fiord Formation (Fig. 4D). margin of the basin, though with fewer zircons of Permian age. Sample SvB-10 was collected from a gray cobble conglomerate imme- diately beneath the sub-Changhsingian unconformity at the Conglomer- Lower Triassic (Northern Margin) ate Section (Figs. 3 and 4G), where it sits on spiculitic chert of the Great Bear Cape Formation. The conglomerate interfingers with sandstone The Triassic succession in the Sverdrup Basin begins with Bjorne but pinches out laterally over short distances, suggesting it is filling a Formation strata along the southern margin and Blind Fiord Formation topographic low on an unconformity surface. These conglomerates are strata along the northern margin (Embry and Beauchamp, 2008). At the assumed here to be Late Permian in age based on the youngest dates pres- Folds Creek locality on northern Axel Heiberg Island, these deposits ent in its detrital zircon signature, with a 2σ maximum depositional age overlie, in homoclinal structure, folded Pennsylvanian to Permian strata of 252.4 ± 3.3 Ma, based on six grains (Dickinson and Gehrels, 2009). across the sub-Changhsingian unconformity (Fig. 4D) (Embry and Beau- Sample SvB-10 yielded 276 concordant dates, dominated by a 400–340 champ, 2008). The Blind Fiord Formation comprises conglomerate, sand- Ma sub-population. This unique signature incorporates a small 295–250 stone, and shale that were deposited in shallow to deep marine settings Ma sub-population and very few grains yielding dates. (Thorsteinsson, 1974; Embry, 1993; Embry and Beauchamp, 2008). A Sample SvB-11 was collected from a maroon sandstone unit uncon- major deepening and widening event of the Sverdrup Basin and associ- formably lying on the Great Bear Cape Formation and beneath the upper- ated increase in sediment supply coming from northern sources are thus most Permian (Griesbachian) to Lower Triassic Blind Fiord Formation recorded in the Induan–Olenekian second-order sequence of the Sverdrup with a conformable contact at Griesbach Creek on northern Axel Heiberg Basin (Fig. 2), (Embry and Beauchamp, 2008; Embry, 2009). Island (Figs. 3 and 4H). This sandstone is assumed Late Permian in age The Triassic samples, combined into composite sample SvB-13, were based on its stratigraphic position. collected at the Folds Creek locality from sandstones at the top of the Sample SvB-11 yielded 268 concordant dates. The dominant zircon first of three late Induan third-order sequences (Fig. 4D) (Embry, 1988). sub-populations are between 290 and 250 Ma and 450 and 345 Ma, with Embry (2009) provided evidence that the abundant siliciclastic material in significant modes at 266 Ma and 399 Ma. Zircon grains yielding dates the Lower Triassic succession along the northern margin of the Sverdrup

436 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

Basin was derived from Crockerland to the north, and attest to a time when 5RDGLDQ drainage systems were better integrated. 6Y% Composite sample SvB-13 yielded a total of 314 measurements, with V  Q  few grains older than 800 Ma and dominated by the sub-populations between 295 and 250 Ma, 380 and 340 Ma, and 700 and 500 Ma (Fig. 7). The 700–500 Ma sub-population is less abundant than in the older sam- ples, and the 295–250 and 400–350 Ma sub-populations are also in lesser .XQJXULDQ abundance. The 480–400 Ma and 2000–900 Ma grains are scarce in the 6Y% V  sample and widely distributed along these intervals. Q 

DISCUSSION

Provenance Implications 5HODWLYH3UREDELOLW\

Late Carboniferous to Early Permian Terrigenous Sediment Sources The similarities between the detrital zircon signatures of upper Penn- sylvanian to Lower Permian strata of the northern and southern margins of the Sverdrup Basin are indicative of shared sediment sources. Detrital        zircon populations in these samples are similar to those from the Silurian %HVW$JH 0D (Pridoli) Danish River Formation on Ellesmere Island, both containing 'DQLVK5LYHU)P 3ULGROL  Q  %HUDQHNHWDO 2000–900 Ma sub-populations, with the modes at ca. 1750, 1420, and 'HYRQLDQ&ODVWLF:HGJH Q  $Q¿QVRQHWDOD 1020 Ma, and from Devonian Clastic Wedge strata, containing a sub- population with a prominent ca. 2000 Ma mode (Fig. 8) (Anfinson et al., Figure 8. Comparison of the probability density plots of Lower to Middle 2012a; Beranek et al., 2015). A recycled Silurian sedimentary source with Permian strata of the southern margin of the basin analyzed in this study with other data sets. n—number of accepted analyses; S—number of minor contributions from recycled Devonian strata is also supported by samples. the relatively small proportion of 400–350 Ma and 700–500 Ma grains in these samples, contrasting with their prominence in Devonian Clastic Wedge strata (Fig. 8) (Beranek et al., 2010; Anfinson et al., 2012a, 2012b). It is also noteworthy that sample SvB-01, despite being from a sandstone margin of the Sverdrup Basin show a similar detrital zircon signature to interfingering with basalts and volcaniclastic rocks, does not contain grains that of the Lower Permian sediments on both basin margins. The major yielding dates near the known depositional age. difference between the Kungurian and Roadian sequences is the addi- These results indicate that during the late Pennsylvanian to Early Perm- tion of a 1700–1600 Ma sub-population, which is not as prominent in ian sediments that accumulated along both the northern and southern the Kungurian sample as in the Roadian one (Fig. 8). Magmatic rocks of margins of the Sverdrup Basin were either derived from recycling of the this age interval are found in Grenvillian-aged basement of the Scandina- Danish River Formation and other Franklinian Basin strata, or from the vian Caledonides and southern Greenland (Kerr et al., 1996; Bingen and same ultimate sources, that is the East Greenland Caledonides, Pearya, Solli, 2009; Pettersson et al., 2009, 2010; Gasser and Andresen, 2013). and the Laurentian Shield (Anfinson et al., 2012a; Hadlari et al., 2014; The prominence of the 1700–1600 Ma zircon sub-population can there- Beranek et al., 2015). On the northern margin, the sampled units exhibit fore be explained by derivation from more distal sources, such as the petrological characteristics typical of proximal derivation, such as a red Sveconorwegian Orogen in the East Greenland Caledonides, reworking weathering color, a narrow lateral extent of the red sandstone within the of strata from northern Greenland, and reworking of Devonian volcanic units (sample SvB-01), cobble-sized clasts and limited areal Clastic Wedge strata with abundant grains of this age interval (Fig. 8) extent of the units on top of the Nansen and Raanes formations (SvB- (Kirkland et al., 2009; Anfinson et al., 2012a). The distal source interpreta- 02 and SvB-03). On the southern margin, Middle Permian sediments tion is consistent with the inferred tectonic setting: a quiescent, thermally were deposited unconformably on the maximum flooding surface of the subsiding basin in which marine sediments accumulated shortly after underlying Kungurian (uppermost Lower Permian) sequence, filling local peneplanation of inverted grabens created during the Melvillian Distur- depocenters created by an episode of basin inversion with small drainage bance (Harrison, 1995; Beauchamp et al., 2001; Embry and Beauchamp, areas and local reworking of sediment (Harrison, 1995; Embry and Beau- 2008). However, some reworking of older strata is required to explain the champ, 2008; Morris, 2013). Hence, based on the detrital zircon signature, small 700–500 Ma sub-population, which is present in the Roadian and facies, and stratigraphy of upper Pennsylvanian to Lower Permian strata of Kungurian samples and proposed to have been originally derived from both basin margins it is likely that reworking of older units, particularly of Timanian-aged basement in Crockerland and shed into the Franklinian Silurian age, was the main sediment source for these small depocenters. Basin during the Devonian (Anfinson et al., 2012a). The absence of zircons yielding dates younger than 365 Ma in all of In contrast, detrital zircon signatures from Middle Permian to Lower the upper Pennsylvanian and Lower Permian samples indicates that syn- Triassic sediments on northern Axel Heiberg Island point to a different depositional volcanism in the area did not yield a significant number of source region, with some shared sub-populations with Devonian Clastic zircons or that these zircons were not incorporated into the sampled units. Wedge strata. The 700–500 Ma and 450–370 Ma sub-populations, abun- dant in the Middle Permian to Lower Triassic samples, are also dominant Middle Permian to Early Triassic Terrigenous Sediment Sources in strata of the Devonian Clastic Wedge, and have previously been inter- Roadian (early Middle Permian) sediments that overlie the Kungu- preted to characterize the signature of sediments derived from Crockerland rian (latest Early Permian) shale at Borup Fiord Pass along the southern in the Devonian (Beranek et al., 2010; Anfinson et al., 2012a). Franklinian

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 437

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

Basin sediments derived from Pearya contain numerous grains in the (Thorsteinsson, 1974; Mayr et al., 2002), but it is likely younger than its 480–455 Ma interval (Hadlari et al., 2014; Beranek et al., 2015), which southern margin counterpart, as it lies above the unconformity at the top are not abundant in the Middle Permian to Lower Triassic samples ana- of the Great Bear Cape Formation. It is thus probably late Kungurian lyzed, thus ruling out Pearya as a significant source for these. The 400–350 (ca. 272 Ma) in age. The geochemical signatures of these basalts point to Ma sub-populations are interpreted to have been sourced from intrusions an intraplate origin for the magma from which they originated that is most related to arc magmatism within Crockerland prior to its accretion to the likely related to the thinning of the lithosphere, elevated heat flow, partial Laurentian margin and to syn- to post-tectonic magmatism related to melting, and possible underplating associated with the rifting processes the Ellesmerian Orogeny, based on the evolved Hf isotopic signature of that led to the Sverdrup Basin following the collapse of the Ellesmerian zircons yielding dates within this age bracket (Anfinson et al., 2012b). Orogen (Cameron and Muecke, 1996; Embry and Beauchamp, 2008; The Middle Permian to Lower Triassic samples from northern Axel Morris, 2013). Heiberg Island also exhibit a spread of Permian dates between 295 and These various volcanic units did not provide any zircons to the samples 250 Ma, a sub-population similar to that previously found in Lower Tri- of the same age analyzed in this study, despite having been collected from assic samples from northern Axel Heiberg Island (Omma et al., 2011). Pennsylvanian sandstones interfingering with basalts (SvB-01) and from It is important to note that relatively thick Permian to Triassic sandstone Lower to Middle Permian shales and sandstones immediately overly- samples collected some 300–500 km to the southwest in the Isachsen J-37 ing eroded Esayoo basalts at the southern margin (SvB-04 and SvB-05, (Ellef Island) and Brock C-50 () wells (Fig. 1) do not respectively). Outpouring of this volcanic material was episodic, short- contain Permian zircons, but are instead dominated by the 420–350 Ma lived and of limited stratigraphic and geographic extent on both northern sub-population (Anfinson et al., 2016). This population, in addition to the Axel Heiberg Island and on northwestern Ellesmere Island. In contrast, absence of Permian grains, suggests the clastic material in these Perm- the spectrum of zircons that is repeatedly found along the northern mar- ian and Triassic deposits comprise sediments recycled from the Devo- gin of the basin, while lacking Carboniferous zircons, spans the entire nian Clastic Wedge (Anfinson et al., 2016). This is turn suggests that the Permian interval (295–250 Ma) and extends into the Early Triassic. This Crockerland watershed that funneled sediments to the area now occupied sub-population is abundant, often dominant, in most Middle and Upper by northern Axel Heiberg Island during the Middle Permian to Early Tri- Permian and Lower Triassic samples, requiring a high number of zircons assic was not the same drainage system that brought contemporaneous to be sourced from these basalts, which is unlikely due to the low zircon clastic sediments in the area now occupied by Ellef Ringnes to Brock productivity of this lithology. The chronological, lithological, and areal islands to the southwest. constraints of upper Paleozoic volcanic units in the Sverdrup Basin, in Three hypotheses have been proposed to address the provenance addition to the lack of zircons of near-depositional age in samples imme- of 295–250 Ma grains found in Triassic strata of the Sverdrup Basin: diately overlying or interfingering with these volcanic rocks, suggests a (1) Late Paleozoic volcanism within the Sverdrup Basin (Omma et al., different magmatic source for the 295–250 Ma zircons in Permian and 2011; Anfinson et al., 2016); (2) derivation from granitoids on Taimyr Triassic strata of the Sverdrup Basin. Peninsula and the Polar Urals (Omma et al., 2011); and (3) unexposed crust of uncertain nature beneath the (submerged areas of Uralian and Siberian Provenance Chukotka and the Chukchi Borderlands) (Omma et al., 2011), potentially A far-traveled Uralian or Siberian provenance for the near-depositional related to arc magmatism (Midwinter et al., 2016). These three hypotheses zircons found in Middle Permian, Upper Permian, and Lower Triassic are addressed below. units of the northern Sverdrup Basin, as proposed for Mesozoic strata by Anfinson et al. (2016), is problematic. Transportation of sediments Late Paleozoic Volcanism within the Sverdrup Basin from the Urals into the Sverdrup Basin could only be explained by long Permian volcanic units are known from the Sverdrup Basin, Alaska’s distance airborne and/or fluvial transport. Airborne transport can be ruled North Slope, the Uralian Orogen, and the Siberian Traps (Beauchamp, out based on the abundance, continual stratigraphic range, and limited 1995; Cameron and Muecke, 1996; Vernikovsky et al., 1995, 2003; Pease, geographical extent of the 295–250 Ma zircons in the Sverdrup Basin. 2001; Embry and Beauchamp, 2008; Reichow et al., 2009; Omma et al., Far-traveled ash-fall deposits would have blanketed much larger areas 2011). In the Sverdrup Basin, Carboniferous and Permian volcanic rocks extending over both the northern and southern margins of the basin. Such are alkaline basalts that form relatively thin units of short duration of erup- deposits would also be unlikely to be stratigraphically persistent in one tion comprising a small number of individual flows restricted to two small small area for more than 40 m.y. areas. The first area is along the basin’s southern margin, west and south of Rivers carrying sediment from the Urals into the Sverdrup Basin, as Borup Fiord Pass on Ellesmere Island. Here a thin unnamed unit of lowest proposed by Anfinson et al. (2016) require Crockerland to be a large, Artinskian (ca. 290 Ma) basalts and volcaniclastic rocks outcrops in the low-lying landmass to allow for sediment bypass. Such an inferred physi- vicinity of Oobloyah Bay (Morris, 2013). A thicker and more widespread ography flies against the strong evidence of pre-Early Triassic uplifts on unit of lower Kungurian (ca. 278 Ma) basalts, the Esayoo Volcanics, occurs northern Axel Heiberg Island, most likely extending to the northwest at Mount Leith and in the Krieger Mountains (Thorsteinsson, 1974). The on Crockerland. Accordingly, erosion related to the development of the second area is along the basin’s northern margin from northern Axel sub-Changhsingian (sub-late Late Permian) unconformity cuts deeply Heiberg Island to northwestern Ellesmere Island (Kleybolte Peninsula). down-section toward the northwest, eventually removing the entire upper Here Lower Carboniferous (upper Serphukhovian; ca. 323 Ma) basalts, Paleozoic succession, such as west of Aurland Fiord on northwestern the unit named the Audhild volcanics, contemporaneous with the second Axel Heiberg Island where the Lower Triassic Blind Fiord Formation rifting pulse of the Sverdrup Basin (Embry and Beauchamp, 2008), lie in rests directly on Franklinian basement strata (Fig. 3) (Mayr et al., 2002). the upper part of the Borup Fiord Formation (Trettin, 1988; Thorsteins- Clearly, Crockerland was everything but a low-relief flat land during the son, 1974). An unnamed unit of uppermost Carboniferous (Gzhelian) Middle and Late Permian. Furthermore, the Uralian fluvial hypothesis basalts occur within the upper Nansen Formation on northwestern Axel requires the existence of ~1000-km-long land bridge linking the northern- Heiberg Island (Mayr et al., 2002). The Lower Permian Esayoo Volcanics most portion of the Uralian Orogen with Crockerland, such as portrayed was also mapped in two discrete areas on northern Axel Heiberg Island by Anfinson et al. (2016) for the . Such a bridge is nowhere

438 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

to be seen in any of the large numbers of Middle to Late Permian plate zone ca. 305–300 Ma. Considering the detrital zircon signature of Middle reconstructions found in the literature, including that of Anfinson et al. Permian to Lower Triassic strata of the northern margin of the Sverdrup (2016) for the Middle Permian, for there is simply no evidence that such a Basin, which is dominated by near-depositional dates, the existence of a paleogeographic feature existed at that time. In contrast, the area between Permian-Triassic magmatic arc is consistent with the characteristic sig- the northern Urals and Crockerland, which is represented by the north- nature of sediments deposited in the convergent margin setting signature ern extension of the Barents Sea, was a broad open marine seaway that as compiled by Cawood et al. (2012). connected the Paleo-Pacific Ocean to the west with the Barents Shelf, Previous work by Embry and Anfinson (2013) also referred to the pres- Finnmark Platform and Timan-Pechora Basin to the east (e.g., Anfinson ence of near-depositional zircons in Upper Triassic strata of the Sverdrup et al., 2016). Basin and proposed the existence of magmatic activity in Crockerland Additional evidence against the Uralian fluvial hypothesis includes related to the paleo-Pacific margin. Furthermore, Hf isotopic measure- the fact that a large fluvial system from the Urals would transport sedi- ments carried out on 350–210 Ma zircons from Upper Triassic units in ments sourced from a large drainage area including a large portion of the northern Axel Heiberg Island display a broad range of values without any orogen, as well as eroded foreland basin sediments. The consistent strati- noticeable differentiation trends, suggesting derivation from a mixture of graphically narrow 295–250 Ma zircon sub-population found in samples evolved, intermediate and juvenile lithosphere, interpreted as a continen- SvB-07 to SvB-13 contrasts with the abundance of grains yielding dates tal arc signature (Midwinter et al., 2016). These data, combined with the between 340 and 260 Ma in various Triassic units of and the Bar- presence of near-depositional zircons and newly described bentonites in ents Sea, interpreted to have received sediments from the Uralian Orogen Triassic strata of the Sverdrup Basin, are consistent with a convergent (Miller et al., 2006; Soloviev et al., 2015; Ershova et al., 2015a; Zhang margin setting along the northwestern edge of Laurentia (Midwinter et et al., 2016; Klausen et al., 2017; Ershova et al., 2017). The far-traveled, al., 2016; Hadlari et al., 2017). The detrital zircon signatures of Upper fluvial transport interpretation is thus at odds with the 340–295 Ma gap Triassic and younger strata of the Sverdrup Basin contain grains yielding a in the detrital zircon spectra of Middle Permian to Lower Triassic strata spread of dates between 340 and 210 Ma (Anfinson et al., 2016; Midwinter of the northern margin of the Sverdrup Basin. This implies that the drain- et al., 2016), contrasting with the isolated 295–250 Ma sub-population age areas for these sediments did not incorporate zircon-bearing rocks found in Middle Permian to Lower Triassic strata in the Sverdrup Basin. or sediments containing zircons of these age interval for ~40 m.y. These Therefore, the source of near-depositional grains in Middle Permian to strata are lithologically heterogeneous, ranging from pebble-sized chert Lower Triassic units must differ from the sources of detrital zircons in conglomerate to fine-grained sandstone, which, added to the euhedral, Upper Triassic and younger strata. relatively unabraded morphology of the 295–250 Ma grains found by Relevant to this discussion is the fact that Triassic strata in the Barents Omma et al. (2011) in lowermost Triassic strata on the northern margin Sea contain a young near-depositional 250–210 Ma sub-population, inter- of the Sverdrup Basin, imply minimal transport. preted by Klausen et al. (2017) as derived from inferred magmatic activity Additionally, Upper Triassic strata of the Sverdrup Basin contain zir- associated with the development of the Novaya Zemlya fold and thrust cons as young as ca. 210 Ma (Anfinson et al., 2016; Midwinter et al., belt. Cobbles yielding Permian dates are also present in Jurassic strata 2016). The youngest igneous bodies documented in the Urals are syenite- of Franz Josef Land, derived from basement uplifts containing plutonic granitic stocks (249–241 Ma) and mafic intrusions (229–227 Ma) related rocks of this age, although it is unclear where they were derived from and to magmatism associated with the Siberian Traps (Vernikovsky et al., the mechanisms responsible for this magmatism (Ershova et al., 2017). 2003; Walderhaug et al., 2005), and are thus not a plausible source for the Triassic strata in Chukotka and the Lisburne Hills in Alaska also contain a near-depositional zircons in Upper Triassic units of the Sverdrup Basin. young, near-depositional zircon sub-population of uncertain origin yield- ing Carboniferous to Triassic dates (Miller et al., 2006; Tuchkova et al., Unexposed Crust beneath the Arctic Ocean 2011). Miller et al. (2006) and Tuchkova et al. (2011) suggested a Uralian The remaining explanation for the source of 295–250 Ma zircons and Siberian provenance for these zircons in Chukotka and the Lisburne in Middle Permian to Lower Triassic strata requires magmatic activity Hills, although Tuchkova et al. (2011) noted that the Late Triassic shelf occurring north of the Sverdrup Basin spanning the Permian and Early depositional environment of the Lisburne Hills was incompatible with a Triassic. Continuous influx of near-depositional zircons into the northern Uralian or Siberian source. Thus, the possibility of two or more unidenti- part of the basin for over 40 m.y. requires long-lasting, proximal, localized fied sources of Permian to Triassic zircons in the Arctic region, and their magmatic activity in a landmass located outboards of the northern portion implications for paleogeographic reconstructions cannot be discarded. of the Sverdrup Basin. That landmass (Crockerland) became emergent by the Roadian (early Middle Permian), coinciding with the first major Evidence for the Magmatic Arc clastic influx from the north. The accompanying Paleozoic and Precam- brian zircon sub-populations present in the samples derived from this Zircon grains yielding Permian and earliest Triassic dates in Middle landmass point at the existence of evolved crust containing Timanian, Permian to Lower Triassic strata of the northern margin of the Sverdrup Caledonian, and Ellesmerian aged basement or sedimentary cover, as well Basin, accompanied by the evidence of local derivation of these rocks, as Precambrian rocks and sediments with age distributions similar to the calls for the existence of a magmatic arc located north of the Sverdrup Permian samples on the southern margin and the upper Pennsylvanian Basin during the Permian to Triassic (Fig. 9). This magmatic arc would to Lower Permian samples of the northern margin. Magmatism in this comprise Timanian-aged crust involved in the Ellesmerian Orogeny and landmass began in the earliest Permian (ca. 300 Ma), coincident with the minor abundance of Caledonian-aged rocks. It would have experienced change in stress fields in the Sverdrup Basin that caused a transition from continuous magmatism throughout the Permian distinct from the mag- post-rift passive subsidence to episodic basin inversion (Beauchamp et matism that shed young zircons onto the Siberian margin and Chukotka. al., 2001; Embry and Beauchamp, 2008). The tectonic setting that best Development of this hypothetical magmatic arc to the north of the explains the detrital zircon data set and the stratigraphic, sedimentologi- Sverdrup Basin would have followed the initiation of a southeast-dipping cal, and structural observations is that of a convergent margin in which subduction zone ca. 305 Ma and would have remained magmatically active a magmatic arc developed as a response to the initiation of a subduction from the earliest Permian (ca. 300 Ma) to at least the earliest Triassic (ca.

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 439

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

Lower Triassic paleogeography during the late Paleozoic and early Mesozoic. Beginning s=4 in the Late Devonian, the crustal block on which Permian arc magmatism n=225 took place (namely Crockerland), was located within the Arctic Alaska– Chukotka Microplate and was involved in the Ellesmerian Orogeny, as indicated by Ellesmerian-aged zircons (390–355 Ma) derived from this landmass and its inferred paleogeographic position (Embry, 1993; Miller et al., 2006; Beranek et al., 2010). Detrital zircon and stratigraphic evidence from Chukotka, Wrangel Middle and upper Island, and the Kara Block (Severnaya Zemlya and northern Taimyr) Permian s=6 place these elements adjacent to the Laurentian and Baltican margin, n=743 respectively, during the Devonian (Kos’ko et al., 1993; Miller et al., 2010; Ershova et al., 2015a, 2015b, 2015c, 2016; Pease et al., 2015). Between the Devonian and Carboniferous, the Kara Block detached from the Baltica- Laurentia margin and collided against the Siberian margin, as interpreted from a change in sediment provenance from Baltican to Uralian sources Lower Permian between these periods (Pease et al. 2015; Ershova et al., 2015a). As such, s=3 post-collisional rifting associated with the initiation of subsidence in the n=95 Sverdrup Basin may have had a northeastern continuation through which the Kara Block rifted apart and became disconnected from the Baltican margin (Ershova et al., 2016). The northern edge of the Sverdrup Basin was dominated by carbonate platform sedimentation during much of the Pennsylvanian and Early Permian, with little to no clastic detritus derived from northern sources due to the paucity of emergent land to the north 100 200 300 400 500 600 700 of the basin (Beauchamp, 2015). Most current models of Arctic tectonic Best Age (Ma) evolution place Crockerland adjacent to the Sverdrup Basin up until the Permian Arc (295 - 250 Ma) opening of the Amerasia Basin by the Cretaceous (Fig. 10), with frag- Post-Caledonian and Ellesmerian (400 - 350 Ma) ments of Crockerland now included in the Chukchi Borderlands based on Caledonides and Pearya (480 - 400 Ma) its position after restoration of the opening of the Arctic Ocean, (Miller et Timanides (700 - 500 Ma) al., 2006; Lawver et al., 2011; Laverov et al., 2013; Gottlieb et al., 2014; Amato et al., 2015; Hadlari et al., 2016; O’Brien et al., 2016). Triassic Wrangel Island (n=198) (Miller et al., 2010) The convergent margin required to explain the detrital zircon signature Triassic Chukotka (n=199) (Miller et al., 2006) of Middle Permian to Lower Triassic strata from the northern margin Norian Sverdrup Basin (n=252) (Anfinson et al., 2016) of the Sverdrup Basin conflicts with most reconstructions of the Arctic Alaska–Chukotka Microplate, which consider the Pacific side of this entity Figure 9. Comparison of the probability density plots of Middle Permian to as a passive margin (e.g., Miller et al., 2006, 2010; Laverov et al., 2013; Lower Triassic strata of the southern margin of the basin analyzed in this Anfinson et al., 2016). However, Midwinter et al. (2016) challenged the study with other data sets. n—number of accepted analyses; s—number of samples. passive margin hypothesis based on the predominance of Carboniferous to near-depositional zircons in Upper Triassic strata from the Sverdrup Basin with Hf isotopic signatures of a continental arc (Midwinter et al., 2016), as well as the presence of similar near-depositional zircons in Tri- 250) with a major pulse in the Middle Permian (ca. 265 Ma). This tim- assic units in the Lisburne Hills, Chukotka, and Wrangel Island (Miller ing is roughly coincident with the change in stress fields and resulting et al., 2006, 2010; Tuchkova et al., 2011; Amato et al., 2015), too young transpressional deformation in the Sverdrup Basin starting in the latest to be sourced from any known magmatic rocks in the Urals and . Carboniferous (ca. 305 Ma), and the development of the sub-Roadian Permian to Triassic northerly derived strata in Chukotka incorporate near- and sub-Changhsingian locally angular unconformities (ca. 272 and 251 depositional zircons, volcaniclastic debris (tuffs, lapilli and volcanic lithic Ma, respectively), both associated with basin-wide uplift and deformation fragments) and geochemical characteristics (Th/U ratio and Nd isotopic and followed by major transgressions. All of these events have so far not signatures) that point at a contribution from juvenile, undifferentiated received a satisfactory geodynamic explanation. arc sources (Tuchkova et al., 2011), questioning the passive interpretation In this context, the onset of siliciclastic input from northern sources for this margin. into the basin could correspond to a period of unroofing of the proposed A convergent margin setting for Chukotka in the late Paleozoic to arc starting in the Middle Permian and continuing throughout at least the Mesozoic is compatible with geological observations, given that large- Early Triassic. Similarly, the basin inversion events that started in the lat- scale correlations are hampered by overprinting due to Mesozoic oro- est Carboniferous (ca. 305 Ma) can be attributed to a change in the stress genesis, and that the lack of documentation of a Permo-Triassic subduc- regime from extensional to compressional (or transpressional) of this sub- tion complex can be explained by poor preservation due to underplating duction complex, forcing a transition from backarc to retroarc dynamics. beneath the arcs accreted in the Jurassic (see Amato et al., 2015, and ref- erences therein), and the remoteness and lack of detailed mapping of this Implications for Arctic Paleogeography region. No evidence of a late Paleozoic to early Mesozoic active margin, or Permo-Triassic magmatic activity (except for the volcanics encountered Understanding the role of this magmatic arc in relation to all other in the Tunalik-1 well in the Chukchi Sea; Beauchamp, 1995), has been tectonic elements in the Arctic realm is crucial for reconstructing its documented within Arctic Alaska, probably due to similar reasons. There

440 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

Latest Pennsylvanian (ca. 300 Ma) Middle Permian (ca. 270)

WI WI

Chukotka Chukotka ? ?

(ChB) CL a

sk

Alaska Ala

Arctic Arctic ?

500 Km 500 Km

Figure 10. Simplified schematic model of proposed latest Carboniferous and Middle Permian paleogeography. Shadowed areas are interpreted as emerged land and coastal settings. Svalbard and Greenland are restored to their inferred position during each time period. Chukotka is shown rifting away from the Laurentian margin with an adjacent magmatic arc since the Devonian. Southeast-dipping subduction is shown underneath Crockerland (CL), producing its emergence in the form of a magmatic arc. Arctic Alaska is restored adjacent to the Sverdrup Basin. ChB—Chukchi Borderland; WI—Wrangel Island.

exists, however, a record of an active margin farther south in the Yukon (Slide Mountain and Angayucham oceans) documented by Colpron and during Permian to Triassic time and the opening of an oceanic backarc Nelson (2011) and Israel et al. (2014). The rapid collapse of this orogen basin, the Slide Mountain and Angayucham oceans (see Colpron and can be explained in this model by a shift in the stress regime from com- Nelson, 2011, and references therein). Thus, an active margin setting for pressional to extensional due to subduction initiation and slab-rollback the paleo-Pacific margin of Arctic Alaska–Chukotka should be consid- of the convergent margin beneath Arctic Alaska–Chukotka. This newly ered, even preferred, when explaining the tectonic setting of the Sverdrup created extensional regime in a backarc position could have induced the Basin and adjacent domains during the late Paleozoic and early Mesozoic. initiation of Sverdrup Basin subsidence, creating depocenters within the A more complex reconstruction than a single subduction zone is Laurentian margin and the terranes now incorporated within the Arctic required due to the discrepancies in the detrital zircon spectra between Alaska–Chukotka Microplate, as indicated by the distinct geochronologi- Lower Triassic strata in Chukotka–Wrangel Island and the Sverdrup Basin, cal signature of lower Paleozoic strata in various areas considered to be which represent two distinct sources of near-depositional zircons. Such part of this landmass (Miller et al., 2010; Ershova et al., 2015c, 2016). separation into two different magmatic domains arises from the prolonged Analogues of this proposed backarc-induced collapse of the Ellesmerian and reproducible absence of 340–300 Ma zircons in Permian and Lower Orogen are also recognized along the Variscan Orogen, which led Ziegler Triassic strata of the Sverdrup Basin. Zircons yielding dates within the (1988) to first propose that this mechanism was responsible for the onset 340–300 Ma interval did not appear in the Sverdrup Basin until the Norian of subsidence in the Sverdrup Basin. (Late Triassic) (Anfinson et al., 2016; Midwinter et al., 2016), that is ~50 The isolated 295–250 Ma zircon sub-population is interpreted in our m.y. after northerly derived sedimentation in the Sverdrup Basin began. An model to have originated from a magmatic arc that started developing alternative model with a single arc would require a landward migration of in the latest Carboniferous to earliest Permian outboard of the Sverdrup arc magmatism and the drainage divide, which would prevent grains aged Basin, in the hypothetical terrane that Embry (1993) termed Crockerland. 340–300 Ma from reaching the Sverdrup Basin, even from recycling of By the Middle Permian, a change in subduction dynamics underneath older strata. Such an explanation requires shallowing of the subduction Crockerland led to the emergence of the magmatic arc and onset of silici- angle and/or an underplating event causing partial melting of the crust in clastic sedimentation along the northern margin of the basin, providing which the arc developed, as well as a spatially restricted catchment area a source of Permian zircons. Upper Triassic to Lower Jurassic strata of for the basin for over 40 m.y., and is therefore deemed unlikely. Further- the Sverdrup Basin contain 340–300 Ma and younger sub-populations, more, the detrital zircons analyzed yielding Permian and Early Triassic yielding nearly identical detrital zircon signatures to Triassic strata of dates do not show evidence of inherited cores, a characteristic that would Chukotka and Wrangel Island (Miller et al., 2010; Tuchkova et al., 2011; be expected from zircons sourced from magmas originated through partial Omma et al., 2011; Anfinson et al., 2016; Midwinter et al., 2016), imply- melting of the crust due to underplating. ing that the magmatic arc that sourced Triassic strata of Wrangel Island The following speculative model for the tectonic evolution of the and Chukotka became a source of sediment for the Sverdrup Basin by Sverdrup Basin (Figs. 10 and 11), which integrates the available geo- the Late Triassic (Midwinter et al., 2016; Hadlari et al., 2017). This is chronological data and geological observations from the regions involved, proposed here to have occurred after closure of the backarc basin that begins with the development of a backarc basin between Chukotka–Wran- formed during the Carboniferous between the Laurentian margin and gel Island and the Laurentian margin immediately following the collapse Chukotka, and subsequent amalgamation of both magmatic arcs in the of the Ellesmerian Orogen in the early Mississippian. latest Permian to Early Triassic. Accretion of these land masses to the This backarc basin represents a northern extension of the chain of northern margin of the Sverdrup Basin is proposed as the mechanism backarc basins developed along the northwestern margin of Laurentia for the development of the sub-Changhsingian (sub-late Late Permian)

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 441

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

NW SE Late Devonian (ca. 365 Ma) Ellesmerian Orogen Chukotka WI Franklinian Basin Sea level

Laurentia 390-350 Ma Figure 11. Schematic cross sec- Pennsylvanian (ca. 310 Ma) 420 - 350 Ma tion of the proposed model of evolution of the Sverdrup Basin and former adjacent domains. The time intervals refer to the age of Sea level Chukotka & WI Chukchi Shelf Sverdrup Basin the magmatism interpreted to have sourced the detrital zircons Laurentia in strata of the various domains shown. East-directed subduction 350 - 300 Ma beneath the Laurentian margin in the Devonian to earliest Carbon- iferous led to backarc extension and collapse of the Ellesmerian Middle Permian (ca. 270 Ma) Orogen. WI—Wrangel Island. Sverdrup Pre-Roadian Sea level Chukotka & WI Chukchi Shelf Basin strata

Laurentia

300 - 210 Ma 300 - 250 Ma

unconformity and the deepening of the basin in the earliest Triassic, as Subduction initiation and accretion of pericratonic terranes along the well as the major influx of northerly derived sediments in the Late Triassic margins of Pangea have been proposed to occur in response to a large (Heiberg Formation) as the product of the erosion of the highlands to the scale change in the stress regime after closure of the Uralian Trough and north, with the accompanying shift towards a juvenile Nd isotopic signa- culmination of the Uralian Orogen in Late Permian to earliest Triassic ture (Patchett et al., 2004; Embry and Beauchamp, 2008; Embry, 2009). time (Cawood and Buchan, 2007). It is worth considering that the main However, none of the Permian and Lower Triassic samples that Patchett hypothesized subduction zone beneath Arctic Alaska–Chukotka represents et al. (2004) analyzed for Nd isotopes were collected from northerly the northern continuation of the convergent margin in which the Uralian derived units, and thus it is likely that this juvenile shift occurred during Orogen developed, with more complexities than we are able to elucidate, the Middle Permian in the northern margin of the basin. The proposed potentially incorporating elements similar to the tectonic configuration of opening and closure of the backarc basin and the juxtaposition of the two the present-day southwestern Pacific. If this was the case in the Canadian arcs with the Laurentian margin, is similar in time and configuration to Arctic, Late Permian deformation events in the Sverdrup Basin could be the evolution of the Yukon-Tanana terrane and the Slide Mountain Ocean linked to resulting changes in the subduction dynamics of these magmatic farther south, which closure culminated in the Late Permian Klondike arcs and their backarc regions, although the specific operating mechanisms Orogeny (Beranek and Mortensen, 2011; Colpron and Nelson, 2011; remain a matter of debate. Israel et al., 2014). The initiation of a subduction zone beneath Crockerland may have CONCLUSIONS occurred in response to the changes in plate rotation of Euramerica (Lau- russia) after its collision with the Kazakhstan plate near the Carbonifer- Using detrital zircon U-Pb geochronology of upper Paleozoic to Lower ous–Permian boundary (Nikishin et al., 1996; Torsvik et al., 2001). Lower Triassic strata of the Sverdrup Basin three main sediment sources are rec- to units in Svalbard do not contain Permian to Triassic ognized. Lower Permian sediments on the southern and northern margins zircons (Bue and Andresen, 2014), and in the absence of detrital zircon were derived from erosion of adjacent highs resulting from episodes of studies of Permian to Triassic units in northern Greenland, the easternmost basin inversion, with no zircons derived from the volcanic units in the extent of this subduction zone is placed offshore of Ellesmere Island in the area (Esayoo Formation and Carboniferous-Permian volcanics). After Canadian . Permian to Triassic strata in the subsurface peneplanation of the pre-Roadian relief, Middle Permian sediments on near and further south do not contain the 295–250 the southern margin of the Sverdrup Basin were derived from sources Ma detrital zircon sub-population (Anfinson et al., 2016), which, added to in the Canadian and Greenland shields, with some reworking of grains the absence of volcanic deposits and post-Early Permian deformation in from underlying units. the western portion of the basin, limits the western extent of the proposed Middle Permian to lowermost Triassic successions along the north- subduction zone to offshore of Ellef Ringnes Island (Fig. 10). ern margin of the Sverdrup Basin contain detrital zircons of Permian to

442 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

earliest Triassic age (295–250 Ma) forming a sub-population isolated and tectonics: Journal of the Geological Society, v. 172, no. 2, p. 201–212, https://doi​ ​.org​ /10​.1144​/jgs2014​-027. from other Paleozoic sub-populations, interpreted to derive from a newly Bingen, B., and Solli, A., 2009, Geochronology of magmatism in the Caledonian and Sveconor- defined magmatic arc that experienced magmatism throughout this time wegian belts of Baltica: Synopsis for detrital zircon provenance studies: Norsk Geologisk interval and different from any other magmatic events within the Arctic Tidsskrift, v. 89, p. 267–290. Black, L.P., Kamo, S.L., Allen, C.M., Davis, D.W., Aleinikoff, J.N., Valley, J.W., Mundil, R., Camp- region recognized to this date. This convergent system originated in the bell, I.H., Korsch, R.J., Williams, I.S., and Foudoulis, C., 2004, Improved 206Pb/238U micro- backarc region of a convergent margin near Chukotka and Wrangel Island, probe geochronology by the monitoring of a trace-element-related matrix effect; SHRIMP, resulting in a second subduction zone underneath Crockerland beginning ID-TIMS, ELA-ICP-MS and oxygen isotope documentation for a of zircon standards: Chemical Geology, v. 205, no. 1, p. 115–140, https://doi​ .org​ /10​ .1016​ /j​ .chemgeo​ .2004​ .01​ .003.​ near the Carboniferous-Permian boundary, with a southeastward dip and Brumley, K., Miller, E.L., Konstantinou, A., Grove, M., Meisling, K.E., and Mayer, L.A., 2015, located adjacent to the northeastern portion of the Sverdrup Basin. Contin- First bedrock samples dredged from submarine outcrops in the Chukchi Borderland, Arctic Ocean: Geosphere, v. 11, p. 76–92, https://​doi​.org​/10​.1130​/GES01044​.1. ued subduction led to the closure of the backarc basin and the amalgama- Bue, E.P., and Andresen, A., 2014, Constraining depositional models in the Barents Sea re- tion of the two magmatic arcs against the margin of Laurentia, affecting gion using detrital zircon U-Pb data from Mesozoic sediments in Svalbard, in Scott, R.A., the Sverdrup Basin evolution during the late Paleozoic to early Mesozoic. Smyth, H.R., Morton, A.C., and Richardson, N., eds., Sediment Provenance Studies in Hydrocarbon Exploration and Production: Geological Society Special (London) Publica- tion 386, p. 261–279, https://​doi​.org​/10​.1144​/SP386​.14. ACKNOWLEDGEMENTS Cameron, B.I., and Muecke, G.K., 1996, Permian alkaline basalts associated with formation of Funding for this research was provided by the GeoMapping for Energy and Mineral (GEM) the Sverdrup Basin, Canadian Arctic: Canadian Journal of Earth Sciences, v. 33, no. 10, Program of Natural Resources Canada (NRCan), the Polar Continental Shelf Program (PCSP), p. 1462–1473, https://​doi​.org​/10​.1139​/e96​-110. and the Natural Sciences and Engineering Research Council of Canada (NSERC) through Cawood, P.A., and Buchan, C., 2007, Linking accretionary orogenesis with supercontinent as- various discovery grants. Detrital zircon analyses were conducted at the Centre for Pure and sembly: Earth-Science Reviews, v. 82, no. 3, p. 217–256, https://doi​ .org​ /10​ .1016​ /j​ .earscirev​ ​ Applied Thermochronology and Tectonics (CPATT) at the University of Calgary in the LA-ICP- .2007​.03​.003. MS laboratory funded by the Canadian Foundation for Innovation (CFI project 30696). We Cawood, P.A., Hawkesworth, C.J., and Dhuime, B., 2012, Detrital zircon record and tectonic are thankful to Dr. Luke Beranek, Dr. Jaime Toro, and an anonymous reviewer for providing setting: Geology, v. 40, p. 875–878, https://​doi​.org​/10​.1130​/G32945​.1. valuable feedback on a previous version of this manuscript. Anirudh Bhargava provided Colpron, M., and Nelson, J.L., 2011, A Palaeozoic NW Passage and the Timanian, Caledonian valuable support in the field, and discussions with Dr. Ashton Embry further refined the and Uralian connections of some exotic terranes in the North American Cordillera, in ideas here presented. Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, A.V., and Sørensen, K., eds., Arc- tic Petroleum Geology: Geological Society (London) Memoir 35, p. 463–484, https://​doi​ .org​/10​.1144​/M35​.31. REFERENCES CITED Dickinson, W.R., and Gehrels, G.E., 2009, Use of U–Pb ages of detrital zircons to infer maxi- Amato, J.M., Toro, J., Miller, E.L., Gehrels, G.E., Farmer, G.L., Gottlieb, E.S., and Till, A.B., mum depositional ages of strata: a test against a Colorado Plateau Mesozoic database: 2009, Late Proterozoic–Paleozoic evolution of the Arctic Alaska–Chukotka terrane based on Earth and Planetary Science Letters, v. 288, no. 1, p. 115–125, https://​doi​.org​/10​.1016​/j​ U-Pb igneous and detrital-zircon ages: Implications for Neoproterozoic paleogeographic .epsl​.2009​.09​.013. reconstructions: Geological Society of America Bulletin, v. 121, no. 9-10, p. 1219–1235, Embry, A.F., 1988, Middle-Upper Devonian sedimentation in the Canadian Arctic Islands and https://​doi​.org​/10​.1130​/B26510​.1. the Ellesmerian Orogeny, in MacMillan, N.J., Embry, A.F., and Glass, D.J., eds., Devonian Amato, J.M., Toro, J., Akinin, V.V., Hampton, B.A., Salnikov, A.S., and Tuchkova, M.I., 2015, of the World: Canadian Society of Petroleum Geologists, v. 2, p. 15–28. Tectonic evolution of the Mesozoic South Anyui suture zone, eastern Russia: A critical Embry, A.F., 1991, Middle Upper Devonian clastic wedge of the Arctic Islands, in Trettin, H.P., component of paleogeographic reconstructions of the Arctic region: Geosphere, v. 11, ed., Geology of the Innuitian Orogen and Arctic Platform of Canada and Greenland: no. 5, p. 1530–1564, https://​doi​.org​/10​.1130​/GES01165​.1. Ottawa, Geological Survey of Canada, Geology of Canada, no. 3, p. 261–280, https://doi​ ​ Anfinson, O.A., Leier, A.L., Embry, A.F., and Dewing, K., 2012a, Detrital zircon geochronology .org​/10​.4095​/133959, and Boulder, Colorado, Geological Society of America, Geology of and provenance of the Neoproterozoic to Late Devonian Franklinian basin, Canadian North America, v​. E, https://​doi​.org​/10​.1130​/DNAG​-GNA​-E​.261. Arctic Islands: Geological Society of America Bulletin, v. 124, no. 3–4, p. 415–430, https://​ Embry, A.F., 1993, Crockerland—the northwest source area for the Sverdrup Basin. Canadian doi​.org​/10​.1130​/B30503​.1. Arctic Islands, in Vorren, T.O., ed., Arctic Geology and Petroleum Potential: Norwegian Anfinson, O.A., Leier, A.L., Gaschnig, R., Embry, A.F., and Dewing, K., 2012b, U-Pb and Hf iso- Petroleum Society Special Publication 2, p. 205–216, https://​doi​.org​/10​.1016​/B978​-0​-444​ topic data from Franklinian Basin strata: Insights into the nature of Crockerland and the -88943​-0​.50018​-6. timing of accretion, Canadian Arctic Islands: Canadian Journal of Earth Sciences, v. 49, Embry, A.F., 2009, Crockerland—the source area for the Triassic to strata of no. 11, p. 1316–1328, https://​doi​.org​/10​.1139​/e2012​-067. northern Axel Heiberg Island, Canadian Arctic Islands: Bulletin of Canadian Petroleum Anfinson, O.A., Leier, A.L., Dewing, K., Guest, B., Stockli, D.F., and Embry, A.F., 2013, Insights Geology, v. 57, p. 129–140, https://​doi​.org​/10​.2113​/gscpgbull​.57​.2​.129. into the Phanerozoic tectonic evolution of the northern Laurentian margin: detrital apa- Embry, A.F., and Anfinson, O.A., 2013, Crockerland—The little Arctic terrane that could [ex- tite and zircon (U–Th)/He ages from Devonian strata of the Franklinian Basin, Canadian tended abs.]: GeoConvention 2013, Canadian Society of Petroleum Geologists: Integration, Arctic Islands: Canadian Journal of Earth Sciences, v. 50, no. 7, p. 761–768, https://​doi​ https://www.geoconvention.com/archives/2013/333_GC2013_A_History_of​_Crockerland​ .org​/10​.1139​/cjes​-2012​-0177. .pdf (accessed January 2018). Anfinson, O.A., Embry, A.F., and Stockli, D.F., 2016, Geochronologic constraints on the Perm- Embry, A.F., and Beauchamp, B., 2008, Sverdrup Basin, in Miall, A., ed., Sedimentary Basins ian–Triassic northern source region of the Sverdrup Basin, Canadian Arctic Islands: Tec- of the World, v. 5: The Sedimentary Basins of the United States and Canada: Amster- tonophysics, v. 691, p. 206–219, https://​doi​.org​/10​.1016​/j​.tecto​.2016​.02​.041. dam, Elsevier, p. 451–472. Balkwill, H.R., 1978, Evolution of Sverdrup Basin: American Association of Petroleum Geolo- Ershova, V.B., Prokopiev, A.V., Khudoley, A.K., Sobolev, N.N., and Petrov, E.O., 2015a, U/Pb gists Bulletin, v. 62, p. 1004–1028. dating of detrital zircons from late Palaeozoic deposits of Bel’kovsky Island (New Siberian Beauchamp, B., 1995, Permian history of North America, in Scholle, P.A., Peryt, T.M., and Islands): Critical testing of Arctic tectonic models: International Geology Review, v. 57, Ulmer-Scholle, O.S., eds., The Permian of Northern Pangea, v. 1, Paleogeography, Pa- no. 2, p. 199–210, https://​doi​.org​/10​.1080​/00206814​.2014​.999358. leoclimates and Stratigraphy: Heidelberg, New York, Springer Verlag, p. 3–22, https://doi​ ​ Ershova, V.B., Prokopiev, A.V., Khudoley, A.K., Sobolev, N.N., and Petrov, E.O., 2015b, Detrital .org​/10​.1007​/978​-3​-642​-78590​-0_1. zircon ages and provenance of the Upper Paleozoic successions of Kotel’ny Island (New Beauchamp, B., 2015, The Middle Permian emergence of Crockerland [abs.]: AAPG 3P Arc- Siberian Islands archipelago): Lithosphere, v. 7, p. 40–45, https://​doi​.org​/10​.1130​/L387​.1. tic, The Polar Petroleum Potential, Stavanger, , 29 September–2 October 2015. Ershova, V.B., Prokopiev, A.V., Nikishin, V.A., Khudoley, A.K., Malyshev, N.A., and Nikishin, Beauchamp, B., Harrison, J.C., Utting, J., Brent, T.A., and Pinard, S., 2001, Carboniferous and A.M., 2015c, New data on Upper Carboniferous-Lower Permian deposits of Bol’shevik Permian subsurface stratigraphy, , , Canadian Island, Severnaya Zemlya Archipelago: Polar Research, v. 34, https://doi​ ​.org​/10​.3402​ Arctic, Geological Survey of Canada, Bulletin 565, 93 p., https://​doi​.org​/10​.4095​/212595. /polar​.v34​.24558. Beauchamp, B., Henderson, C.M., Grasby, S.E., Gates, L.T., Beatty, T.W., Utting, J., and James, Ershova, V.B., Lorenz, H., Prokopiev, A.V., Sobolev, N.N., Khudoley, A.K., Petrov, E.O., and N.P., 2009, Late Permian sedimentation in the Sverdrup Basin, Canadian Arctic: The Lind- Thomsen, T.B., 2016, The De Long Islands: A missing link in unraveling the Paleozoic ström and Black Stripe Formations: Bulletin of Canadian Petroleum Geology, v. 57, no. 2, paleogeography of the Arctic: Gondwana Research, v. 35, p. 305–322, https://​doi​.org​ p. 167–191, https://​doi​.org​/10​.2113​/gscpgbull​.57​.2​.167. /10.1016​/j​.gr​.2015​.05​.016. Beranek, L.P., and Mortensen, J.K., 2011, The timing and provenance record of the Late Perm- Ershova, V.B., Prokopiev, A.V., Sobolev, N.N., Petrov, E.O., Khudoley, A.K., Faleide, J.I., Gaina, ian Klondike orogeny in northwestern Canada and arc‐continent collision along western C., and Belyakova, R.V., 2017, New data on the basement of Franz Josef Land, Arctic re- North America: Tectonics, v. 30, no. 5, TC5017, https://​doi​.org​/10​.1029​/2010TC002849. gion: Geotectonics, v. 51, no. 2, p. 121–130, https://​doi​.org​/10​.1134​/S0016852117020030. Beranek, L.P., Mortensen, J.K., Lane, L., Allen, T., Fraser, T., Hadlari, T., and Zantvoort, W., Ganerød, M., Chew, D.M., Smethurst, M.A., Troll, V.R., Corfu, F., Meade, F., and Prestvik, T., 2010, Detrital zircon geochronology of the western Ellesmerian clastic wedge, north- 2011, Geochronology of the Tardree Rhyolite Complex, Northern : Implications for western Canada: Insights on Arctic tectonics and the evolution of the northern Cordil- zircon fission track studies, the North Atlantic Igneous Province and the age of the Fish leran miogeocline: Geological Society of America Bulletin, v. 122, p. 1899–1911, https://​ Canyon sanidine standard: Chemical Geology, v. 286, no. 3, p. 222–228, https://​doi​.org​ doi​.org​/10​.1130​/B30120​.1. /10​.1016​/j​.chemgeo​.2011​.05​.007. Beranek, L.P., Pease, V., Hadlari, T., and Dewing, K., 2015, Silurian flysch successions of Elles- Gasser, D., and Andresen, A., 2013, Caledonian terrane amalgamation of Svalbard: Detri- mere Island, Arctic Canada, and their significance to northern Caledonian palaeogeography tal zircon provenance of Mesoproterozoic to Carboniferous strata from Oscar II Land,

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 443

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 ALONSO-TORRES ET AL.

western : Geological Magazine, v. 150, no. 6, p. 1103–1126, https://doi​ ​.org​ McNicoll, V.J., Harrison, J.C., Trettin, H.P., and Thorsteinsson, R., 1995, Provenance of the /10.1017​/S0016756813000174. Devonian clastic wedge of Arctic Canada: evidence provided by detrital zircon ages, in Gottlieb, E.S., Meisling, K.E., Miller, E.L., and Charles, G., 2014, Closing the Canada Basin: Dorobek, S.L., and Ross, G.M., eds., Stratigraphic Evolution of Foreland Basins: SEPM Detrital zircon geochronology relationships between the North Slope of Arctic Alaska (Society of Sedimentary Geology) Special Publication 52, p. 77–93, https://​doi​.org​/10​ and the Franklinian mobile belt of Arctic Canada: Geosphere, v. 10, p. 1366–1384, https://​ .2110​/pec​.95​.52​.0077. doi​.org​/10​.1130​/GES01027​.1. Midwinter, D., Hadlari, T., Davis, W.J., Dewing, K., and Arnott, R.W.C., 2016, Dual provenance Hadlari, T., Davis, W.J., Dewing, K., Heaman, L.M., Lemieux, Y., Ootes, L., Pratt, B.R., and Pyle, signatures of the Triassic northern Laurentian margin from detrital-zircon U-Pb and Hf- L.J., 2012, Two detrital zircon signatures for the Cambrian passive margin of northern isotope analysis of Triassic–Jurassic strata in the Sverdrup Basin: Lithosphere, v. 8, p. 668– Laurentia highlighted by new U-Pb results from northern Canada: Geological Society of 683, https://​doi​.org​/10​.1130​/L517​.1. America Bulletin, v. 124, p. 1155–1168, https://​doi​.org​/10​.1130​/B30530​.1. Miller, E.L., Toro, J., Gehrels, G., Amato, J.M., Prokopiev, A., Tuchkova, M.I., Akinin, V.V., Du- Hadlari, T., Davis, W.J., and Dewing, K., 2014, A pericratonic model for the Pearya terrane as mitru, T.A., Moore, T.E., and Cecile, M.P., 2006, New insights into Arctic paleogeography an extension of the Franklinian margin of Laurentia, Canadian Arctic: Geological Society and tectonics from U-Pb detrital zircon geochronology: Tectonics, v. 25, https://​doi​.org​ of America Bulletin, v. 126, p. 182–200, https://​doi​.org​/10​.1130​/B30843​.1. /10​.1029​/2005TC001830. Hadlari, T., Midwinter, D., Galloway, J.M., Dewing, K., and Durbano, A.M., 2016, Mesozoic Miller, E.L., Gehrels, G.E., Pease, V., and Sokolov, S., 2010, Stratigraphy and U-Pb detrital zircon rift to post-rift tectonostratigraphy of the Sverdrup Basin, Canadian Arctic: Marine and geochronology of Wrangel Island, Russia: Implications for Arctic paleogeography: The Petroleum Geology, v. 76, p. 148–158, https://​doi​.org​/10​.1016​/j​.marpetgeo​.2016​.05​.008. American Association of Petroleum Geologists Bulletin, v. 94, no. 5, p. 665–692, https://​ Hadlari, T., Midwinter, D., Poulton, T.P., and Matthews, W.A., 2017, A Pangean rim of fire: Reviewing doi​.org​/10​.1306​/10200909036. the Triassic of western Laurentia: Lithosphere, v. 9, p. 579–582, https://​doi.org​ /10​ .1130​ /L643​ .1.​ Miller, E.L., Soloviev, A.V., Prokopiev, A.V., Toro, J., Harris, D., Kuzmichev, A.B., and Gehrels, Harrison, J.C., 1995, Melville Island’s salt-based fold belt, Arctic Canada: Geological Survey G.E., 2013, Triassic river systems and the paleo-Pacific margin of northwestern Pangea: of Canada, Bulletin 472, 331 p, https://​doi​.org​/10​.4095​/203576. Gondwana Research, v. 23, no. 4, p. 1631–1645, https://​doi​.org​/10​.1016​/j​.gr​.2012​.08​.015. Israel, S., Beranek, L., Friedman, R.M., and Crowley, J.L., 2014, New ties between the Alex- Morris, N., 2013, Stratigraphy and Geochemistry of Lower Permian volcanics in the Sverdrup ander terrane and Wrangellia and implications for North America Cordilleran evolution: Basin, Northwest Ellesmere Island, [M.Sc. thesis]: University of Calgary, 222 p. Lithosphere, v. 6, p. 270–276, https://​doi​.org​/10​.1130​/L364​.1. Mortensen, J.K., and Card, K.D., 1993, U–Pb age constraints for the magmatic and tectonic Kerr, A., Ryan, A.B., Gower, C.F., and Wardle, R.J., 1996, The Makkovik Province: Extension of evolution of the Pontiac Subprovince, Quebec: Canadian Journal of Earth Sciences, v. 30, the Ketilidian mobile belt in mainland North America, in Brewer, T.S., ed., Precambrian no. 9, p. 1970–1980, https://​doi​.org​/10​.1139​/e93​-173. Crustal Evolution in the North Atlantic Region: Geological Society of London Special Nikishin, A.M., Ziegler, P.A., Stephenson, R.A., Cloetingh, S.A.P.L., Furne, A.V., Fokin, P.A., Ershov, Publication 112, p. 155–177, https://​doi​.org​/10​.1144​/GSL​.SP​.1996​.112​.01​.09. A.V., Bolotov, S.N., Korotaev, M.V., Alekseev, A.S., and Gorbachev, V.I., 1996, Late Precam- Kirkland, C.L., Pease, V., Whitehouse, M.J., and Ineson, J.R., 2009, Provenance record from brian to Triassic history of the East European Craton: Dynamics of sedimentary basin evolu- Mesoproterozoic–Cambrian sediments of Peary Land, North Greenland: Implications for tion: Tectonophysics, v. 268, no. 1, p. 23–63, https://doi​ .org​ /10​ .1016​ /S0040​ -1951​ (96)00228​ -4.​ the ice-covered Greenland Shield and Laurentian paleogeography: Precambrian Research, O’Brien, T.M., Miller, E.L., Benowitz, J.P., Meisling, K.E., and Dumitru, T.A., 2016, Dredge sam- v. 170, p. 43–60, https://​doi​.org​/10​.1016​/j​.precamres​.2008​.11​.006. ples from the Chukchi Borderland: Implications for paleogeographic reconstruction and Klaper, E.M., 1992, The Paleozoic tectonic evolution of the northern edge of North America: tectonic evolution of the Amerasia Basin of the Arctic: American Journal of Science, A structural study of northern Ellesmere Island, Canadian Arctic Archipelago: Tectonics, v. 316, no. 9, p. 873–924, https://​doi​.org​/10​.2475​/09​.2016​.03. v. 11, no. 4, p. 854–870, https://​doi​.org​/10​.1029​/92TC00277. Omma, J.E., Pease, V., and Scott, R.A., 2011, U-Pb SIMS zircon geochronology of Triassic Klausen, T.G., Müller, R., Slama, J., and Helland-Hansen, W., 2017, Evidence for Late Triassic and Jurassic sandstones on northwestern Axel Heiberg Island, northern Sverdrup Basin, provenance areas and sediment supply turnover in the Barents Sea Basin Arctic Canada, in Spencer, A.M., Gautier, D., Stoupakova, A., Embry, A., and Sørensen, of northern Pangea: Lithosphere, v. 9, p. 14–28, https://​doi​.org​/10​.1130​/L556​.1. K., eds., Arctic Petroleum Geology: Geological Society (London) Memoir 35, p. 559–566, Kos’ko, M., Cecile, M., Harrison, J., Ganelin, V., Kandoshko, N., and Lopatin, B., 1993, Geol- https://​doi​.org​/10​.1144​/M35​.37. ogy of Wrangel Island, between Chukchi and Eastern Siberian Seas, Northeastern Russia: Paces, J.B., and Miller, J.D., 1993, Precise U‐Pb ages of Duluth complex and related mafic Geological Survey of Canada, Bulletin, v. 461, 101 p. intrusions, northeastern Minnesota: Geochronological insights to physical, petrogenetic, Kuiper, K.F., Deino, A., Hilgen, F.J., Krijgsman, W., Renne, P.R., and Wijbrans, J.R., 2008, Syn- paleomagnetic, and tectonomagmatic processes associated with the 1.1 Ga midcontinent chronizing rock clocks of Earth history: Science, v. 320, no. 5875, p. 500–504, https://doi​ ​ rift system: Journal of Geophysical Research, Solid Earth, v. 98, no. B8, p. 13997–14013, .org​/10​.1126​/science​.1154339. https://​doi​.org​/10​.1029​/93JB01159. Laverov, N.P., Lobkovsky, L.I., Kononov, M.V., Dobretsov, N.L., Vernikovsky, V.A., Sokolov, S.D., Patchett, P.J., Embry, A.F., Ross, G.M., Beauchamp, B., Harrison, J.C., Mayr, U., Isachsen, and Shipilov, E.V., 2013, A geodynamic model of the evolution of the Arctic basin and C.E., Rosenberg, E.J., and Spence, G.O., 2004, Sedimentary cover of the Canadian shield adjacent territories in the Mesozoic and Cenozoic and the outer limit of the Russian Conti- through Mesozoic time reflected by Nd isotopic and geochemical results for the Sver- nental Shelf: Geotectonics, v. 47, no. 1, p. 1–30, https://​doi.org​ /10​ .1134​ /S0016852113010044.​ drup Basin, Arctic Canada: The Journal of Geology, v. 112, no. 1, p. 39–57, https://​doi​.org​ Lawver, L.A., Gahagan, L.M., and Norton, I., 2011, Palaeogeographic and tectonic evolution /10​.1086​/379691. of the Arctic region during the Palaeozoic, in Spencer, A.M., Embry, A.F., Gautier, D.L., Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., and Maas, R., 2010, Im- Stoupakova, A.V., and Sørenson, K., eds., Arctic Petroleum Geology: Geological Society proved laser ablation U-Pb zircon geochronology through robust downhole fraction- (London) Memoir 35, p. 61–77, https://​doi​.org​/10​.1144​/M35​.5. ation correction: Geochemistry Geophysics Geosystems, v. 11, no. 3, https://​doi​.org​/10​ Lemieux, Y., Hadlari, T., and Simonetti, A., 2011, Detrital zircon geochronology and prove- .1029​/2009GC002618. nance of Devono-Mississippian strata in the northern Canadian Cordilleran miogeocline: Pease, V., 2001, East European Craton Margin source for the allochthonous northern terrane Canadian Journal of Earth Sciences, v. 48, p. 515–541, https://​doi​.org​/10​.1139​/E10​-056. of Taimyr, Arctic Siberia: Eos, v. 82, p. F1181. Lorenz, H., Gee, D.G., and Simonetti, A., 2008, Detrital zircon ages and provenance of the Pease, V., 2011, Eurasian orogens and Arctic tectonics: An overview, in Spencer, A.M., Em- Late Neoproterozoic and Palaeozoic successions on Severnaya Zemlya, Kara Shelf: A tie bry, A.F., Gautier, D.L., Stoupakova, A.V., and Sørensen, K., eds., Arctic Petroleum Geol- to Baltica: Norsk Geologisk Tidsskrift, v. 88, no. 4, p. 235–258. ogy: Geological Society (London) Memoir 35, p. 311–324, https://doi​ ​.org​/10​.1144​/M35​.20. Lorenz, H., Gee, D.G., Korago, E., Kovaleva, G., McClelland, W.C., Gilotti, J.A., and Frei, D., Pease, V., and Scott, R.A., 2009, Crustal affinities in the Arctic Uralides, northern Russia: sig- 2013, Detrital zircon geochronology of Palaeozoic Novaya Zemlya–a key to understand- nificance of detrital zircon ages from Neoproterozoic and Palaeozoic sediments in Novaya ing the basement of the Barents Shelf: Terra Nova, v. 25, no. 6, p. 496–503, https://doi​ ​ Zemlya and Taimyr: Journal of the Geological Society, v. 166, no. 3, p. 517–527, https://​ .org​/10​.1111​/ter​.12064. doi​.org​/10​.1144​/0016​-76492008​-093. Ludwig, K.R., 1998, On the treatment of concordant uranium-lead ages: Geochimica et Cos- Pease, V., Kuzmichev, A.B., and Danukalova, M.K., 2015, The New Siberian Islands and evi- mochimica Acta, v. 62, no. 4, p. 665–676, https://​doi​.org​/10​.1016​/S0016​-7037​(98)00059​-3. dence for the continuation of the Uralides, Arctic Russia: Journal of the Geological So- Ludwig, K.R., 2012, Isoplot v. 3.75—A geochronological toolkit for Microsoft Excel: Berkeley ciety, v. 172, no. 1, p. 1–4, https://​doi​.org​/10​.1144​/jgs2014​-064. Geochronology Center, Special Publication 5, p. 1–75. Pettersson, C.H., Pease, V., and Frei, D., 2009, U/Pb zircon provenance of metasedimen- Malone, S.J., 2012, Tectonic evolution of northern Ellesmere Island: Insights from the Pearya tary basement of the Northwestern Terrane, Svalbard: Implications for the Grenvillian- Terrane, Ellesmerian Clastic Wedge and Sverdrup Basin [Ph.D. thesis]: University of Iowa, Sveconorwegian orogeny and development of Rodinia: Precambrian Research, v. 175, 276 p., https://ir.uiowa.edu/etd/3496/ (accessed January 2018). p. 206–220, https://​doi​.org​/10​.1016​/j​.precamres​.2009​.09​.010. Malone, S.J., McClelland, W.C., von Gosen, W., and Piepjohn, K., 2014, Proterozoic evolu- Pettersson, C.H., Pease, V., and Frei, D., 2010, Detrital zircon U–Pb ages of Silurian–Devonian tion of the North Atlantic–Arctic Caledonides: Insights from detrital zircon analysis of sediments from NW Svalbard: a fragment of Avalonia and Laurentia?: Journal of the metasedimentary rocks from the Pearya terrane, Canadian High Arctic: The Journal of Geological Society, v. 167, no. 5, p. 1019–1032, https://​doi​.org​/10​.1144​/0016​-76492010​-062. Geology, v. 122, no. 6, p. 623–647, https://​doi​.org​/10​.1086​/677902. Reichow, M.K., Pringle, M.S., Al’Mukhamedov, A.I., Allen, M.B., Andreichev, V.L., Buslov, M.M., Matthews, W.A., and Guest, B., 2016, A Practical Approach for Collecting Large-n Detrital Zir- Davies, C.E., Fedoseev, G.S., Fitton, J.G., Inger, S., and Medvedev, A.Y., 2009, The timing con U-Pb Data sets by Quadrupole LA-ICP-MS: Geostandards and Geoanalytical Research, and extent of the eruption of the Siberian Traps large igneous province: Implications for v. 41, no. 2, p. 161–180, https://​doi​.org​/10​.1111​/ggr​.12146. the end-Permian environmental crisis: Earth and Planetary Science Letters, v. 277, no. 1, Maurel, L.E., 1989, Geometry and evolution of the Tanquary Structural High and its effects on p. 9–20, https://​doi​.org​/10​.1016​/j​.epsl​.2008​.09​.030. the paleogeography of the Sverdrup basin, northern Ellesmere Island, Canadian Arctic: Røhr, T.S., Andersen, T., Dypvik, H., and Embry, A.F., 2010, Detrital zircon characteristics of Geological Survey of Canada, Paper 89-1G, p. 177–189. the Lower Cretaceous Isachsen Formation, Sverdrup Basin: Source constraints from age Mayr, U., Beauchamp, B., Embry, A.F., and Harrison, J.C., 2002, Geology of Upper Paleozoic and Hf isotope data: Canadian Journal of Earth Sciences, v. 47, no. 3, p. 255–271, https://​ and Mesozoic Formations of Northern Axel Heiberg Island, Parts of NTS 560D, Cape doi​.org​/10​.1139​/E10​-006. Stallworthy and 560A, Bukken Fiord, Nunavut: Geological Survey of Canada, 2 sheets, Soloviev, A.V., and Miller, E.L., 2014, The U-Pb age dating of detrital zircons from Upper scale 1:125,000, https://​doi​.org​/10​.4095​/213034. Jurassic-Lower Cretaceous deposits of Stolbovoy Island (New Siberian Islands):

444 www.gsapubs.org | Volume 10 | Number 3 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021 Late Paleozoic to Triassic Arc magmatism north of the Sverdrup Basin | RESEARCH

Stratigraphy and Geological Correlation, v. 22, no. 5, p. 507–517, https://doi​ ​.org​/10​.1134​ Colorado, Geological Society of America, Geology of North America, v​. E, https://doi​ ​.org​ /S0869593814050086. /10​.1130​/DNAG​-GNA​-E​.295. Soloviev, A.V., Zaionchek, A.V., Suprunenko, O.I., Brekke, H., Faleide, J.I., Rozhkova, D.V., Trettin, H.P., 1998, Pre-Carboniferous geology of the northern part of the Arctic islands: Geo- Khisamutdinova, A.I., Stolbov, N.M., and Hourigan, J.K., 2015, Evolution of the prove- logical Survey of Canada Bulletin 425, p. 1–401. nances of Triassic rocks in Franz Josef Land: U/Pb LA-ICP-MS dating of the detrital zircon Trettin, H.P., and Mayr, U., 1997, Geological Maps and correlation chart of northern Ellesmere from Well Severnaya: Lithology and Mineral Resources, v. 50, no. 2, p. 102–116, https://​ and Axel : Geological Survey of Canada, Maps 1881A to 1888A, scale doi​.org​/10​.1134​/S0024490215020054. 1:250,000, 17 sheets. Spencer, C.J., Kirkland, C.L., and Taylor, R.J., 2016, Strategies towards statistically robust Tuchkova, M.I., Sokolov, S.D., Khudoley, A.K., Verzhbitsky, V.E., Hayasaka, Y., and Moiseev, A.V., interpretations of in situ U–Pb zircon geochronology: Geoscience Frontiers, v. 7, no. 4, 2011, Permian and Triassic deposits of Siberian and Chukotka passive margins: Sedimen- p. 581–589, https://​doi​.org​/10​.1016​/j​.gsf​.2015​.11​.006. tation setting and provenances. In ICAM VI: Proceedings of the International Conference Stern, R.A., and Amelin, Y., 2003, Assessment of errors in SIMS zircon U–Pb geochronology on Arctic Margins VI, 31 May–2 June 2010, Fairbanks, Alaska, p. 61–96. using a natural zircon standard and NIST SRM 610 glass: Chemical Geology, v. 197, no. 1, Vernikovsky, V.A., Neimark, L.A., Ponomarchuk, V.A., Vernikovskaya, A.E., Kireev, A.D., and Kuz- p. 111–142, https://​doi​.org​/10​.1016​/S0009​-2541​(02)00320​-0. min, D.S., 1995, Geochemistry and age of collision granitoids and metamorphites of the Stephenson, R.A., Embry, A.F., Nakiboglu, S.M., and Hastaoglu, M.A., 1987, Rift-initiated Perm- Kara microcontinent (northern Taimyr): Russian Geology and Geophysics, v. 36, p. 46–60. ian to subsidence of the Sverdrup basin, in Beaumont, C., and Tankard, Vernikovsky, V.A., Pease, V.L., Vernikovskaya, A.E., Romanov, A.P., Gee, D.G., and Travin, A.V., A.J., eds., Sedimentary Basins and Basin-Forming Mechanisms: Canadian Society of 2003, First report of early Triassic A-type granite and syenite intrusions from Taimyr: Prod- Petroleum Geologists, Memoir, v. 12, p. 213–231. uct of the northern Eurasian superplume?: Lithos, v. 66, no. 1, p. 23–36. Strauss, J.V., Macdonald, F.A., Taylor, J.F., Repetski, J.E., and McClelland, W.C., 2013, Lauren- Walderhaug, H.J., Eide, E.A., Scott, R.A., Inger, S., and Golionko, E.G., 2005, Palaeomagnetism tian origin for the North Slope of Alaska: Implications for the tectonic evolution of the and 40Ar/39Ar geochronology from the South Taimyr igneous complex, Arctic Russia: a Mid- Arctic: Lithosphere, v. 5, no. 5, p. 477–482, https://​doi​.org​/10​.1130​/L284​.1. dle-Late Triassic magmatic pulse after Siberian flood-basalt volcanism: Geophysical Jour- Thorsteinsson, R., 1974, Carboniferous and Permian stratigraphy of Axel Heiberg Island and nal International, v. 163, no. 2, p. 501–517, https://doi​ .org​ /10​ .1111​ /j​ .1365​ -246X​ .2005​ .02741​ .x.​ western Ellesmere Island, Canadian Arctic Archipelago: Geological Survey of Canada Wiedenbeck, M.A.P.C., Alle, P., Corfu, F., Griffin, W.L., Meier, M., Oberli, F., Quadt, A.V., Rod- Bulletin, v. 224, 115 p, https://​doi​.org​/10​.4095​/103460. dick, J.C., and Spiegel, W., 1995, Three natural zircon standards for U-Th-Pb, Lu-Hf, trace Thorsteinsson, R., and Tozer, E.T., 1957, Geological Investigations in Ellesmere and Axel Hei- element and REE analyses: Geostandards Newsletter, v. 19, no. 1, p. 1–23, https://doi​ ​.org​ berg Islands, 1956: Arctic, v. 10, p. 2–31, https://​doi​.org​/10​.14430​/arctic3752. /10​.1111​/j​.1751​-908X​.1995​.tb00147​.x. Thorsteinsson, R., and Tozer, E.T., 1970, Geology of the Arctic Archipelago, in Douglas, R.J.W, Wiedenbeck, M., Hanchar, J.M., Peck, W.H., Sylvester, P., Valley, J., Whitehouse, M., Kronz, ed., Geology and Economic Minerals of Canada: Geological Survey of Canada Economic A., Morishita, Y., Nasdala, L., Fiebig, J., and Franchi, I., 2004, Further characterisation of Geology Report 1, p. 547–590. the 91500 zircon crystal: Geostandards and Geoanalytical Research, v. 28, no. 1, p. 9–39, Till, A.B., Amato, J.M., Aleinikoff, J.N., and Bleick, H.A., 2014, U-Pb detrital zircon geochronol- https://​doi​.org​/10​.1111​/j​.1751​-908X​.2004​.tb01041​.x. ogy as evidence for the origin of the Nome Complex, northern Alaska, and implications for Worsley, D., 2008, The post-Caledonian development of Svalbard and the western Barents regional and trans-Arctic correlations, in Dumoulin, J.A., and Till, A.B., eds., Reconstruc- Sea: Polar Research, v. 27, p. 298–317, https://​doi​.org​/10​.1111​/j​.1751​-8369​.2008​.00085​.x. tion of a Late Proterozoic to Devonian Continental Margin Sequence, Northern Alaska, Its Zhang, X., Pease, V., Skogseid, J., and Wohlgemuth-Ueberwasser, C., 2016, Reconstruction Paleogeographic Significance, and Contained Base-Metal Sulfide Deposits: Geological of tectonic events on the northern Eurasia margin of the Arctic, from U-Pb detrital zir- Society of America Special Paper 506, p. 111–131, https://​doi​.org​/10​.1130​/2014​.2506​(04). con provenance investigations of late Paleozoic to Mesozoic sandstones in southern Torsvik, T.H., Van der Voo, R., Meert, J.G., Mosar, J., and Walderhaug, H.J., 2001, Reconstruc- Taimyr Peninsula: Geological Society of America Bulletin, v. 128, p. 29–46, https://doi​ ​ tions of the continents around the North Atlantic at about the 60th parallel: Earth and Plan- .org​/10​.1130​/B31241​.1. etary Science Letters, v. 187, no. 1, p. 55–69, https://doi​ .org​ /10​ .1016​ /S0012​ -821X​ (01)00284​ -9.​ Ziegler, P.A., 1988, Late Carboniferous to Permian Evolution of Arctic Basins, the Norwegian- Trettin, H.P., 1987, Pearya: a Composite Terrane with Caledonian Affinities in Northern Elles- Greenland Sea Rift and Permian Basins of West and Central Europe, in Evolution of the mere Island: Canadian Journal of Earth Sciences, v. 24, no. 2, p. 224–245, https://​doi​.org​ Arctic–North Atlantic and the Western Tethy: American Association of Petroleum Geolo- /10​.1139​/e87​-025. gists (AAPG) Memoir 43, p. 33–42. Trettin, H.P., 1988, Early Namurian (or older) alkali basalt in the Borup Fiord Formation, northern Zonenshain, L.P., 1990, Geology of the USSR: A plate-tectonic synthesis. Geodynamics se- Axel Heiberg Island, Arctic Canada: Geological Survey of Canada, Paper 88-ID, p. 21–26. ries, v. 21, p. 120. Trettin, H.P., 1991, Introduction (Silurian-Early Carboniferous deformational phases and as- sociated metamorphism and plutonism, Arctic Islands), in Trettin, H.P., ed., Geology of MANUSCRIPT RECEIVED 25 JUNE 2017 the Innuitian Orogen and Arctic Platform of Canada and Greenland: Geological Survey of REVISED MANUSCRIPT RECEIVED 27 DECEMBER 2017 Canada, Geology of Canada, v. 3, p. 295–344, https://doi​ .org​ /10​ .4095​ /133959,​ and Boulder, MANUSCRIPT ACCEPTED 6 FEBRUARY 2018

LITHOSPHERE | Volume 10 | Number 3 | www.gsapubs.org 445

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/10/3/426/4180110/426.pdf by guest on 25 September 2021