<<

Dual provenance signatures of the Triassic northern Laurentian margin from detrital-zircon U‑Pb and Hf‑isotope analysis of Triassic–Jurassic strata in the Sverdrup Basin

Derrick Midwinter1, Thomas Hadlari2, W.J. Davis3, Keith Dewing2, and R.W.C. Arnott1 1DEPARTMENT OF AND ENVIRONMENTAL SCIENCES, UNIVERSITY OF OTTAWA, 120 UNIVERSITY, OTTAWA, ONTARIO, K1N 6N5, CANADA 2GEOLOGICAL SURVEY OF CANADA, 3303-33RD STREET NW, CALGARY, ALBERTA, T2L 2A7, CANADA 3GEOLOGICAL SURVEY OF CANADA, 601 BOOTH STREET, OTTAWA, ONTARIO, K1S 2V1, CANADA

ABSTRACT

The tectonic setting of northern Laurentia prior to the opening of the Arctic Ocean is the subject of numerous tectonic models. By better understanding the provenance of detrital zircon in the Canadian Arctic prior to rifting, both the prerift tectonic setting and timing of rifting can be better elucidated. In the Sverdrup Basin, two distinct provenance assemblages are identified from new detrital-zircon U-Pb data from Lower Triassic to Lower Jurassic strata in combination with previously published detrital-zircon data. The first assemblage comprises an age spectrum identical to that of the Devonian clastic wedge in the Canadian Arctic and is termed the recycled source. In contrast, the second assemblage is dominated by a broad spectrum of near syndepositional Permian–Triassic ages derived from north of the basin and is termed the active margin source. Triassic strata of Yukon and Arctic Alaska exhibit a similar dual provenance signature, whereas in northeastern Russia, Chukotka contains only the active margin source. Complementary hafnium isotopic data on Permian–Triassic zircon have eHf values that are consistent with the common evolved crustal signature of the Devonian clastic wedge detrital-zircon grains and Neoproterozoic– basement rocks in the Arctic Alaska–Chukotka microcontinent. Furthermore, newly identified volcanic ash beds throughout the Triassic section from the northern part of the Sverdrup Basin, along with abundant Permian–Triassic detrital zircon, suggest a protracted his- tory of magmatism to the north of the basin. We interpret that these zircons were sourced from a magmatically active region to the north of the Sverdrup Basin, and in the context of a rotational model for opening of Amerasia Basin, this was probably part of a convergent margin fringing northern Laurentia from the northern Cordillera along the outboard edge of Arctic Alaska and Chukotka terranes. In strata, Permian–Triassic zircons decrease substantially, implying the diminution of the active margin as a sediment source as initial rifting isolated the Permian–Triassic source from the Sverdrup Basin.

LITHOSPHERE; v. 8; no. 6; p. 668–683; GSA Data Repository Item 2016151 | Published online 20 May 2016 doi:10.1130/L517.1

INTRODUCTION in Triassic–Jurassic strata in the Sverdrup Basin Carboniferous to the Paleogene (Embry and and compared their age spectra to those from Beauchamp, 2008). The basin is underlain by an Siliciclastic sedimentary successions can other Triassic strata in the circum-Arctic. Those up to 10-km-thick sedimentary pile of Devonian provide an important record of the tectonic set- previous detrital-zircon studies (Miller et al., clastic wedge strata that were deformed during ting and tectonic evolution of a basin through 2006; Omma et al., 2011) in the Sverdrup Basin the Late Devonian–Early Carboniferous Elles- stratigraphic and detrital-zircon patterns. Pre- lacked data from the –Early Juras- merian orogeny (Embry, 1991). Strata equiva- vious efforts to interpret the tectonic evolu- sic Heiberg Formation. This paper provides new lent to the Devonian clastic wedge are widely tion of the Sverdrup Basin have been made detrital-zircon data from this interval, which distributed, including the northern Cordillera (e.g., Balkwill, 1978; Embry and Beauchamp, serves to constrain the provenance of the Sver- of North America, Arctic Alaska, and north- 2008), but important gaps in knowledge remain. drup Basin during the Triassic–Jurassic. Also, ern Russia (Amato et al., 2009; Beranek et al., Incipient rifting of the proto–Amerasia Basin in U-Pb detrital-zircon ages are augmented with 2010a; Drachev, 2011; Lemieux et al., 2011). the Jurassic–Cretaceous (Embry, 1990, 1991; eHf isotopic data for Permian–Triassic zircon The Ellesmerian orogeny was succeeded by Houseknecht and Bird, 2011) was followed by grains. These data provide important insight into initial rifting of the Sverdrup Basin that began opening of the Amerasian ocean basin, which the of the source terrane and magmatism in the Early Carboniferous and ended in the separates Arctic Canada, Alaska, and northeast- (cf. Vervoort and Patchett, 1996). Permian (Embry and Beauchamp, 2008). Cur- ern Russia (e.g., Grantz et al., 1979) (Fig. 1). rent models suggest that following the Permian, Previous detrital-zircon studies in the Sver- GEOLOGIC SETTING the Sverdrup Basin was tectonically quiescent drup Basin (Miller et al., 2006; Omma et al., and underwent thermal subsidence until rifting 2011) have identified a detrital-zircon signa- Sverdrup Basin recommenced in the Jurassic (e.g., Embry and ture in Triassic–Jurassic strata that resembles Beauchamp, 2008). that of the underlying Devonian clastic wedge The Sverdrup Basin is located in the Cana- The Triassic stratigraphy of the Sverdrup (e.g., Anfinson et al., 2012a). Those studies also dian (Fig. 2) and records Basin (Fig. 3) is controlled by repetitive trans- identified several different zircon assemblages near continuous sedimentation from the gressive-regressive events (e.g., Embry, 1988;

668 © 2016 Geological Society of Americawww.gsapubs.org | For permission | toVolume copy, contact8 | Number [email protected] 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

equivalent to the Heiberg Group that comprises five formations in the western part of the basin (Fig. 3; Embry, 1983b). The formations of the Pacific Heiberg Group consist of the sandstone-rich Ocean deltaic Skybattle Formation overlain uncon- formably by marine mudstone of the Grosve- nor Island Formation with the Triassic–Jurassic boundary occurring in its upper part (Embry and Suneby, 1994). These strata then coarsen upward into sandstone-rich strata of the deltaic Maclean Strait Formation with the upper part containing W the base-Sinemurian boundary. Farther upward, AACM these strata are overlain by marine shale of the

Arctic NSI Formation capped by the sand- Ocean e stone-dominant King Christian Formation. During the Triassic there were two princi- pal sources of sediment into the basin deter- Lomonosov Ridg mined by the general facies distributions shown in Figure 4 (Embry, 2009). Sediment transport was directed into the basin from the southern and eastern margins, indicating a southern and eastern sediment source area. U-Pb zircon data (Miller et al., 2006; Anfinson et al., 2012a) and Sm-Nd isotopic data (Patchett et al., 2004) are consistent with recycling from the Devonian Atlantic clastic wedge and older north Laurentian strata Ocean (e.g., Hadlari et al., 2012, 2014). The facies indi- cate that another Triassic sediment source was derived from north of the basin (Embry, 2009), which is consistent with sandstone samples with detrital-zircon age spectra that are different from those on the south side of the basin (Miller et Figure 1. Circum-Arctic orogens, terranes, and locations modified from Colpron and Nelson (2011) al., 2006; Omma et al., 2011). and Pease et al. (2014). AA—Arctic Alaska; AACM—Arctic Alaska–Chukotka microplate; CH—Chu- In the Jurassic–Cretaceous, a narrow paleo- kotka; NSI—New Siberian Islands; NZ—Novaya Zemlya; PE—Pearya; SAS—South Anyui suture zone; high, the Sverdrup Rim (Fig. 2), separated the SV—Svalbard; WI—Wrangel Island; YTT—Yukon Tanana terrane. Black circles represent approximate Sverdrup Basin from the rift grabens of the detrital-zircon sample locations from AA (Miller et al., 2006; Gottlieb et al., 2014); CH (Miller et al., proto–Amerasia Basin (Meneley et al., 1975; 2006; Tuchkova et al., 2011; Amato et al., 2015); WI (Miller et al., 2010); and YTT (Beranek et al., 2010b; Beranek and Mortensen, 2011). The outlines of the SAS and AACM are from Drachev (2011). Red Embry, 1993). The northern source region was circles represent approximate location of eNd or eHf values from AA (Amato et al., 2009); Arctic fully separated from northern Laurentia by the Canada (Anfinson et al., 2012b; Morris, 2013); New Siberian Islands (Akinin et al., 2015); and opening of the Amerasia Basin in the Cretaceous (Malitch et al., 2010). (Embry, 2009).

Paleogeographic Restoration Embry and Beauchamp, 2008). regression (Embry, 1991). Earliest Carnian trans- units of the Sverdrup Basin comprise the Blind gression deposited the Hoyle Bay Formation The tectonic interpretation of the Arctic Fiord Formation, which consists of shale and above the Roche Point Formation. Prograda- region prior to the opening of the Amerasia siltstone representing mid-outer shelf, slope, tion of sandstone-rich, shallow marine deposits Basin is complex and is the subject of much and deeper basin-floor deposits, and the Bjorne (Pat Bay Formation) extended across the basin debate (e.g., Grantz et al., 1979; Embry, 1990; Formation, which consists mostly of sandstone during the late Carnian (Embry, 1993) and was Lawver and Scotese, 1990; Lane, 1997; Lawver confined to the basin margins interpreted to rep- terminated by a major transgression in the latest et al., 2002; Grantz et al., 2011; Pease, 2011; resent deltaic deposits (Embry, 1986). These Carnian–early Norian that deposited prodelta Pease et al., 2014). The region that is central to two siliciclastic units mark the first major clas- mud and silt of the Barrow Formation (Embry, reconstruction is typically referred to as the Arc- tic influx into the basin with the accumulation 1991). These strata progressively coarsen upward tic Alaska–Chukotka microcontinent (AACM; of 2000 m of strata in the basin center (Embry, into marginal marine to nonmarine sandstones of Fig. 1), a lithospheric block occupying north- 1991). The is marked by a major the Heiberg Formation (Embry, 1988). eastern Russia, Arctic Alaska, and their respec- transgression that deposited bituminous source The Heiberg Formation in the central and tive offshore shelves (Pease et al., 2014). The rocks of the Murray Harbour Formation (Embry eastern parts of the basin is subdivided into three AACM was separated from the Siberian craton and Beauchamp, 2008). These strata are overlain predominantly sandstone-rich members (Embry, by the Anyui Ocean, which closed when the by clastic and/or carbonate rocks of the Roche 1983a)—the Romulus, Fosheim, and Remus Amerasia Basin and Arctic Ocean opened. The Point Formation, which represents a short-lived (Fig. 3). These members are stratigraphically timing and mechanism of this opening is the

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 669

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

140°W 130° 120° 110° 100° 90° 80°W

Legend Previous dZr sample location New dZr sample location Devonian clastic wedge dZr sample location Amerasia Basin εHf samples (Anfinson et al., 2012b) Ellesmere Island Volcanic ash beds Bedrock geology Figure 2. Sverdrup Basin map with Lower Triassic strata Cenozoic detrital-zircon (dZr) sample and Sample Name Mesozoic SP volcanic ash bed locations. Detri- 78°N Paleozoic PB tal-zircon data references from Is.- Isachsen Fm. m BF Is. DB- Deer Bay Fm. Isachsen Formation (Røhr et al., DB Axel Heiberg 2010); Sandy Point and Deer Bay SP- Sandy Point Fm. B Sverdrup RiEllef Island formations (Omma et al., 2011); Early Jurassic LH KC- King Christian Fm. Ringnes Pat Bay Formation (Miller et al., 2006; Omma et al., 2011—both Late Triassic KC Island LH- Lower Heiberg Fm. from approximately the same loca- Is. PB- Pat Bay Fm. tion); Bjorne Formation (Miller et Early Triassic al., 2006—location approximately 76° B- Bjorne Fm. Is. BF- Blind Fiord Fm. the same as this study); Blind Fiord Sverdrup Formation (Omma et al., 2011). Prince Basin U-Pb data from Devonian clastic Patrick wedge (Anfinson et al., 2012a);e Hf Island data from Devonian Clastic Wedge e (Anfinson et al., 2012b). Lower Triassic strata: Blind Fiord Forma- Sverdrup Outlin 74° tion (brown); Bjorne Formation (orange). General southern outline of the Sverdrup Basin from Embry Banks and Beauchamp (2008). Surface Island Melville Island and sea bottom bedrock geology is from Okulitch (1991). The Sver- drup Rim was first described by North Meneley et al. (1975). 72° America

0 125 250 500 km

subject of several different models (Grantz et al., Islands and Chukotka (Pease et al., 2014, and therefore is part of the AACM, or if it extends 1979; Embry, 1990; Lane, 1997; Nokleberg et references therein). north of the NSI, and therefore is unrelated to al., 2000; Miller et al., 2006; Kuzmichev, 2009; In a restored position, the outboard margin of the AACM (see discussion in Kuzmichev, 2009; Grantz et al., 2011; Lawver et al., 2011). the AACM is typically interpreted to have been Pease, 2011; Pease et al., 2015). Detrital zircons This paper tests the rotational opening model a passive margin to the Anyui and Angayucham from the Triassic Burustas Formation, exposed of Grantz et al. (1979, 2011) and Embry (1990) oceans in the Triassic (Nokleberg et al., 2000; on the NSI, have a strong fraction of ca. 252 that places the AACM against the Canadian Miller et al., 2006; Sokolov et al., 2009; Tuch- Ma zircon that, in addition to geochemical and Arctic Islands margin prior to the Cretaceous. kova et al., 2009; Miller et al., 2010; Tuchkova et petrographical characteristics of the zircon, are Restoration of the AACM to its location prior al., 2011; Miller et al., 2013; Amato et al., 2015). consistent with zircon ages from Siberian Trap to separation and rotation from the northern This was followed in the Cretaceous by collision magmatism (Miller et al., 2006), possibly indi- Laurentian margin places the North Slope of with the Siberian craton–Verkhoyansk margin, cating that the NSI were connected to Southern Alaska adjacent to Banks and Prince Patrick known as the South Anyui suture (SAS; Fig. 1), Taimyr and Siberia in the Early Mesozoic (e.g., islands. As a result, the provenance of the Sver- after closure of Anyui Ocean and concomitant Kuzmichev and Pease, 2007). Recent studies drup Basin during the Triassic should resemble opening of the Amerasia Basin (e.g., Drachev, (Ershova et al., 2015; Pease et al., 2015) show a that of the North Slope of Arctic Alaska and 2011; Houseknecht and Bird, 2011; Laverov et Baltican detrital-zircon character in the Carbon- Chukotka, since the Sverdrup Basin, Hanna al., 2013; Amato et al., 2015). The SAS contains iferous that changes to Uralian in the Permian. Trough, and Arctic Alaska Basin would have remnants of a Jurassic–Early Cretaceous conver- formed a continuous sedimentary basin from gent system, made up of island arc, continental Summary of Published U-Pb and Sm-Nd the Carboniferous to Jurassic (e.g., Gottlieb et terrane and oceanic basin rocks (Kuzmichev, Provenance Data al., 2014). Seismic interpretation of an extinct 2009; Drachev, 2011; Amato et al., 2015). spreading ridge has a trend that is parallel to the It is not clear whether the SAS extends The overall detrital-zircon spectrum for the shelves offshore of the western Canadian Arctic south of the New Siberian Islands (NSI), and Devonian clastic wedge in the Canadian Arctic

670 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

Basin MESOZOIC Centre Margins εNd (T) Age Age Epoch Age/Stage -30 -20 -1010 0 (Ma) (Ma) Formations Period Legend Maastrichtian 70 Sandstone 72.1 Expedition Mudstone-Siltstone Campanian 80 Evaporite Late 83.6 Santonian 86.3 Kanguk Coniacian Non-marine shale 90 89.8 Turonian Limestone Dominant Juvenile 93.9 Strand Hassel Cenomanian Fiord Volcanic shift 100 100.5 Chert Albian Christopher 110 Mixed clastic-carbonate

Cretaceous 113.0 εNd values 120 Aptian Early Location 126.3 Isachsen NE Ellesmere Island Barremian 130 130.8 S. Ellesmere / Axel Heiberg Island Hauterivian 133.9 Melville Island Valanginian 140 139.4 Berriasian 145.0 Deer Bay Tithonian 150 152.1 Late Kimmeridgian Awingak 157.3 Ringnes 160 Oxfordian 163.5 Callovian 166.1 Hiccles Cove Middle Bathonian 168.3 McConnell Island 170 Bajocian 170.3 Sandy Point Aalenian 174.1

Jurassi c Toarcian 180 Jameson Bay 182.7 Pliens- Heiberg Group / Formation Early bachian Remus (Mbr) 190 190.8 King Christian Sinemurian Fosheim (Mbr) Lougheed Island 199.3 200 Hettangian 201.3 Maclean Strait Rhaetian Figure 3. Initial eNd values from Juvenile Grosvenor Island sedimentary rocks show positive 210 209.5 shift Skybattle / Romulu shift to more juvenile provenance s (Mb r) in the Late Triassic and Late Cre- Norian 220 Late Barrow taceous, and increased negative scatter in eNd values in the Early 228.4 Cretaceous is from a greater con-

230 riassi c Hoyle Bay Pat Bay T Carnian tribution from Shield (Patchett et 237.0 al., 2004). Stratigraphic chart of 240 Ladinian the Sverdrup Basin modified from Middle 241.5 Murray Harbour

Anisian Roche Point Embry and Beauchamp (2008); 247.1 Paleozoic volcanics: Audhild (Tret- 250 Early Olenekian 250.0 Bjorne Induan 252.2 Blind Fiord tin, 1988); Nansen (Mayr et al., Changhsingian 254.2 Late Wuchiapingian Lindstrom 2002); Unnamed Lower volcanics 259.8 Capitanian Degerböls (ULV) and Esayoo (Morris, 2013). Middle 265.1 Trold Fiord Wordian 268.8 van Hauen Assistance Roadian 272.3 275 Sabine Bay Kungurian Esayoo volcanics 279.3

Permia n Early Artinskian Trappers Cove Great Bear Cape Unnamed Lower Volcanics 290.1 Sakmarian Raanes 295.5 Nansen Volcanics Asselian 298.9 300 Gzhelian Hare Fiord Canyon Fiord L 303.7 Kasimovian 307.0 M Moscovian Nansen 315.2 E Bashkirian Otto Fiord Canyon Fiord 323.2 325 Audhild Volcanics L Serpukhovian Borup Fiord 330.9 M Visean Emma Fiord Carboniferous 346.7 350 E Tournaisian

358.9

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 671

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

has a characteristic signature of 700–360 Ma Early Triassic (Olenekian) ages and a broad spectrum of Proterozoic ages

Land N. Ellesmere (Fig. 5; Anfinson et al., 2012a, 2012b). Studies Fluvial Amerasia Basin Island of strata tectonostratigraphically equivalent to Shallow marine the Devonian clastic wedge confirm a consistent Deep marine shelf/slope Basin floor Ellef Axel and diagnostic detrital-zircon signature of this Ringnes Heiberg clastic package across northwestern Laurentia Island Island and the AACM (Beranek et al., 2010a, 2010b; Drachev, 2011; Lemieux et al., 2011; Anfinson Prince et al., 2012a, 2012b), which could have been Patrick Island ultimately derived from Paleozoic arc rocks Sverdrup of the AACM (e.g., Amato et al., 2009, 2015; Basin S. Ellesmere Lemieux et al., 2011; Hadlari et al., 2014). Anal- Island yses of detrital zircon from Triassic strata on the North Slope of Alaska (Gottlieb et al., 2014) d and Triassic and Jurassic strata (Bjorne and Melville Islan Banks Pat Bay formations: Miller et al., 2006; Sandy Island 0 125 250 500 km A Point Formation: Omma et al., 2011) from the Sverdrup Basin have a similar signature as the Devonian clastic wedge (Fig. 5). Sedimentary early Late Triassic (Carnian) recycling of detrital zircon is a common way

Shallow marine N. Ellesmere that younger strata can mirror the detrital-zircon Deep marine shelf/slope Island spectra of older strata (Hadlari et al., 2015), and so the signature of the Devonian clastic wedge Amerasia Basin provides a useful reference curve for circum- Ellef Axel Ringnes Heiberg Island Arctic Mesozoic spectra (Fig. 5). Island Triassic strata within the Sverdrup Basin (Blind Fiord and Pat Bay formations: Omma Prince Patrick et al., 2011) have a detrital-zircon signature Island that has been attributed to a source in western Sverdrup Basin Siberia (e.g., Taimyr, Urals, and Siberian Traps, S. Ellesmere Island Fig. 1; Miller et al., 2006, 2013). The diagnos- tic age fraction for the northern Sverdrup Basin source is a nearly continuous spectrum of near- nd a syndepositional Permian–Triassic ages (Fig. 5). Melville Isl Banks Embry (1993, 2009) postulated a northwestern 0 125 250 500 km Island B provenance for Lower and Upper Triassic strata of the Sverdrup Basin based on lithofacies dis- tribution (Fig. 4) and that this sediment source latest Triassic (Rhaetian) remained active until the lower Middle Jurassic. smere N. Elle Fluvial-delta plain Island The northern source of Triassic sediment in the Shallow marine-delta plain Sverdrup Basin, with its characteristic Permian– Deep marine shelf Triassic zircon, is observed elsewhere in the Arc- Axel Amerasia Basin Ellef tic (Fig. 5), specifically in Chukotka (Miller et al., Ringnes Heiberg Island Island 2006; Tuchkova et al., 2011; Amato et al., 2015), Wrangel Island (Miller et al., 2006, 2010), and Prince Lisburne Hills, Alaska (Miller et al., 2006). The Patrick Western Interior Platform adjacent to the Yukon Island Sverdrup Tanana terrane has a similar dual detrital-zircon Basin e signature with one source displaying strong sim- S. Ellesmer Island ilarities to the Devonian clastic wedge, whereas the other source has abundant near-syndeposi- d tional Permian–Triassic ages derived from arc Melville Islan rocks of the Yukon Tanana terrane (Beranek et Banks 0 125 250 500 km Island al., 2010b; Beranek and Mortensen, 2011). C Sm-Nd isotopic data from Carboniferous to Cretaceous sedimentary rock samples from the Figure 4. General facies distribution during the (A) Early Triassic (modified from figure 36.13 in Embry, 2011); (B) Late Triassic (modified from figure 36.20 in Embry, 2011); Sverdrup Basin have a nearly uniform eNd sig- and (C) latest Triassic (modified from figure 14 in Embry and Beauchamp, 2008). nature throughout the basin’s history (Fig, 3), Arrows represent general sediment transport direction, based on Embry (1991, 2009). which is interpreted to represent the progressive

672 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

Devonian Clastic stage (Mayr et al., 2002; Embry and Beauchamp, Wedge 2008). Early Permian volcanic rocks are termed n=44 Middle Jurassic the Unnamed Lower volcanics (ULV) at the Sandy Point Fm base of the Great Bear Cape Formation, and the n=32 Esayoo volcanics in the Sabine Bay Formation

iassi c (Morris, 2013). The Esayoo volcanics have been Tr interpreted as intra-plate basalts with alkaline to transitional affinities (Cameron and Muecke, Late Triassic 1996). Although there are no previously reported

Permian- Pat Bay Fm (Miller) records of Triassic to Jurassic volcanic rocks in n=47 the Sverdrup Basin, magmatism was active dur- ing the Cretaceous (e.g., Evenchick et al., 2015). Late Triassic n Pat Bay Fm (Omma) ANALYTICAL PROCEDURES n=74 U-Pb Geochronology y Early Triassic Bjorne Fm Samples of the Bjorne Formation and Romu- n=29 lus Member were collected from Ellesmere Sverdrup Basi Island by the authors in 2011, and the King Christian Formation sample was collected by Age Probabilit Carol Evenchick in 2010. U-Pb ages of detrital Early Triassic zircon were analyzed by secondary ion micro- Blind Fiord Fm probe, and performed using the sensitive high- n=914 resolution ion microprobe (SHRIMP) at the Triassic, Chuktoka Geological Survey of Canada (GSC), Ottawa. n=320 Triassic, Wrangel Island Russi a SHRIMP analytical procedures followed those n=267 described by Stern (1997), with standards and Triassic, North Slope U-Pb calibration methods following Stern n=200 and Amelin (2003). Briefly, zircon were cast Alaska Triassic, Lisburne Hills in 2.5-cm-diameter epoxy mounts along with n=519 fragments of the GSC laboratory standard zir- con (z6266, with 206Pb/238U age = 559 Ma). The

ko n Triassic, YTT Platform midsections of the zircon were exposed using

Yu n=1108 Triassic, YTT Overlap 9, 6, and 1 µm diamond compound, and the n=1464 internal features of the zircon (such as zoning, Dev. Clastic Wedge structures, alteration, etc.) were characterized 0 200 400 600 800 1000 1200 1400 1600 1800 2000 in backscattered electron mode utilizing a Zeiss Age (Ma) Evo 50 scanning electron microscope. Mount Figure 5. Triassic to Middle Jurassic published detrital-zircon age spectra. Sverdrup Basin sources are: surfaces were evaporatively coated with 10 nm Sandy Point Formation (Omma et al., 2011); Pat Bay Formation (Miller et al., 2006; Omma et al., 2011); of high-purity Au. Analyses were conducted Bjorne Formation (Miller et al., 2006); Blind Fiord Formation (Omma et al., 2011); Triassic of Chukotka using a 16O- primary beam, projected onto the (Miller et al., 2006; Tuchkova et al., 2011; Amato et al., 2015); Triassic of Wrangel Island (Miller et al., zircon at 10 kV. The sputtered area used for anal- 2010); Triassic of North Slope, Alaska (Gottlieb et al., 2014); Triassic of Lisburne Hills, Alaska (Miller ysis was ~18 µm in diameter with a beam cur- et al., 2006); Triassic of the Yukon Tanana terrane (YTT) platform (Beranek et al., 2010b); Triassic of the YTT Overlap (Beranek and Mortensen, 2011); Devonian clastic wedge (Anfinson et al., 2012a). See rent of ~8–9 nA. The count rates at ten masses Table DR1 (see text footnote 1) for U-Pb detrital-zircon sample location, number, age, and reference. including background were measured over five scans with a single electron multiplier and a pulse-counting system with dead time of 23 ns. recycling of Devonian clastic wedge strata dur- Arctic Large Igneous Province (e.g., Buchan and Offline data processing used SQUID 2.5 soft- ing the Mesozoic (Patchett et al., 2004). Two Ernst, 2006; Estrada and Henjes-Kunst, 2013). ware written by Ludwig (2003). The 1s external excursions to more positive eNd isotopic val- errors of 206Pb/238U ratios reported in the GSA ues occurred during the Late Triassic–earliest Igneous Record of the Sverdrup Basin data repository item1 (Table DR2) incorporate Jurassic and (Fig. 3). Patchett a ±1.0% error in calibrating the standard zircon et al. (2004) hypothesized that the Nd isoto- The earliest record of volcanism within the (see Stern and Amelin, 2003). No fractionation pic shift in the Late Triassic–earliest Jurassic Sverdrup Basin is the Lower Carboniferous resulted from minor volcanic contribution to Audhild volcanics (Trettin, 1988; Embry and 1 GSA Data Repository Item 2016151, which contains the basin, although no evidence of volcanism Beauchamp, 2008), which coincided with early Table DR1: Location, sample number, and references had been identified at that time. The shift in the rifting. Mafic volcanic rocks within Early Perm- for detrital-zircon data used in manuscript; Table DR2: Sverdrup Basin U-Pb data; and Table DR3: Sverdrup Late Cretaceous is consistent with sedimentary ian carbonates of the Nansen Formation have Basin Hf data, is available at www.geosociety​ .org​ /pubs​ ​ input from juvenile volcanic rocks of the High been interpreted to mark the end of the Asselian /ft2016.htm, or on request from [email protected].

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 673

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

correction was applied to the Pb-isotope data; c Upper Heiberg Gp common Pb correction utilized the Pb composi- King Christian Fm 6 tion of the surface blank (Stern, 1997). Isoplot n=45/66, 95-105% conc. v. 3.00 (Ludwig, 2003) was used to generate 4 concordia plots and calculate weighted means. 2 Probability density function plots use ages based on the concordia age calculation method Early Jurassi (Ludwig, 1998) as outlined in Nemchin and Lower Heiberg Fm Cawood (2005) to avoid using different isoto- Romulus Mbr 8 pic ratios for age interpretations based on an y n=48/74, 95-105% conc. arbitrary age cutoff and to be able to evaluate iassi c 6 Tr probability of concordance for zir- 4 Frequenc cons for which the traditional discordance mea- surement is subject to very large errors. Rather Late 2 than picking a subjective cutoff between using 206 238 207 206 Age Probabilit Pb / U and Pb/ Pb age, which is com- y Bjorne Fm 8 mon practice in the literature (e.g., 1200 Ma for Cape O’Brien Mbr Omma et al., 2011; 1000 Ma for Anfinson et al., n=97/113, 95-105% conc. 6 iassi c

2012a), the concordia function calculates a single Tr 4 age based on the relative errors of the measured 206Pb /238U and 207Pb/206Pb ratios (Ludwig, 1998). 200 Ma 2 The approach outlined by Nemchin and Cawood Early (2005) involves a minimum degree of decision y 1.0 making and the use of probability of concordance 0.9 as a screening parameter to reduce bias. In this 0.8 study, a sample is not included if either the prob- 0.7 King Christian Fm ability of concordance is <0.01 or if discordance 0.6 Lower Heiberg Fm measurements are <-5 or >5 (Table DR2). 0.5 Bjorne Fm When comparing multiple samples, the 0.4 probability density function (PDF) is assumed 0.3 to represent all the possible ages in the samples, 0.2 and all the zircon ages must fall within the PDF 0.1 curve. Accordingly, the area below the curve for Cumulative Probabilit each sample is comparable. Another compara- 0 5001000 1500 2000 2500 3000 tive technique uses the cumulative distribution Age (Ma) function (CDF). Although similar to the PDF, Figure 6. Relative probability distribution and cumulative probability plots for three new the CDF sums the probabilities with increasing detrital-zircon samples from this study. Kernel density estimation represented by solid age but requires equivalence in the population gray line. See Table DR2 (see text footnote 1) for complete isotopic analyses. size. This step function does not account for uncertainties in the measured values. In spite of the differences, a PDF and CDF display the U-Pb zircon ages with a total of 24 analyses 0.000309 (Blichert-Toft, 2008). Measured val- same information, but each has its own strength: conducted. Hf analyses were conducted with a ues were ~15% higher than accepted values with PDFs are easier to use when evaluating the pres- Photon excimer laser and a Nu Plasma multi- a reproducibility of ~23% at 95% confidence. ence or absence of specific ages in age distri- collector inductively coupled plasma mass spec- Reproducibility of the 176Lu/176Hf value was bet- butions, whereas CDFs are more useful when trometer in time-resolved analyses mode at the ter than 5% for the 6266 standard. assessing similarities or differences within a set GSC, Ottawa. Data were acquired using either Accuracy and reproducibility were moni- of age distributions (Guynn and Gehrels, 2010). a 30, 40, or 50 mm beam size selected based on tored by replicate analyses of four zircon stan- Kernel density estimation (KDE) is displayed grain size of the target. dards (91500, Temora 2, 6266, and Mud Tank), with PDF in Figure 6. An advantage of KDE is Chondritic uniform reservoir (CHUR) values each showing excellent agreement with pub- that the bandwidth is adaptive; therefore, with of 176Lu/177Hf = 0.0336 and 176Hf/177Hf = 0.282785 lished data (Woodhead and Hergt, 2005; Wu et sparse data density, the density estimate becomes are from Bouvier et al. (2008). Depleted man- al., 2006; Blichert-Toft, 2008). Based on inter- increasingly smooth. Kernel plots were produced tle model based on 176Lu/177Hf = 0.03902 and nal precision, each of the standards exhibits using Density Plotter software (Vermeesch, 176Hf/177Hf = 0.28327 and new crust model age excess scatter as indicated by high mean square 2012). Results are presented in Figure 6 with calculations utilized values of Dhuime et al. of weighted deviates (MSWD), and external full U-Pb analytical data compiled in Table DR2. (2011) (176Lu/177Hf = 0.03781 and 176Hf/177Hf = errors in the range from 0.9 to 1.2 epsilon units 0.28316). 176Lu decay constant of 1.867 × 10-11 at 2s are required. A minimum external error Hafnium-Isotope Methods yr-1 was used (Söderlund et al., 2004). of 1.2e (2s) is assumed for the data. Elemental fractionation of Lu and Hf was Hafnium data are presented on Hf-evolu- Two samples were analyzed for Hf isotopes monitored and corrected based on the 91500 tion diagrams that show eHf values at the time with grains selected specifically for their young standard and an accepted 176Lu/176Hf value of of crystallization. In order to make use of the

674 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

greater number of Nd isotopic data reported for Romulus Member of the Heiberg Depleted Mantle potential source rocks, published neodymium Formation 15 isotopic data are converted to “equivalent” Hf- On central Axel Heiberg Island (Depot isotope values based on the high degree of cor- Point), a sample was collected from the Romu- JUVENILE 10 relation established between the two systems lus Member in the lower part of the Heiberg (Vervoort et al., 1999). The equation, derived by Formation. The Romulus Member is interpreted

s INTERMEDIATE Vervoort et al. (1999), was utilized to calculate to be part of a delta front and is given a Norian 5

equivalent eHf values. age (Embry, 1983a, and references therein). alue Detrital-zircon ages can be divided into three 0 CHUR eHf = 1.36 eNd + 2.95; r2 = 0.86 (1) distinct fractions—ca. 2580–1690 Ma, 920–750 Ma, and 560–215 Ma—with a notable absence

This conversion allows comparison of more of Mesoproterozoic ages (Fig. 6). The Zircon εHf V -5 geographically widespread eNd data on the Hf- to Triassic zircon fraction has 30 grains (~63%

evolution plot. of the sample population) with peaks at 425 and VED For the calculation of hafnium-depleted 270 Ma. The Neoproterozoic ages, of which -10 EVOL model ages (TDM), crustal evolution trajectories there are five grains (~10% of the sample), has 176 177 minimum 2σ assume present-day Lu/ Hf ratio of 0.0093 one peak at ca. 850 Ma. The Paleoproterozoic -15 uncertainty (Vervoort and Patchett, 1996; Amelin et al., ages have a broad spectrum with three peaks, 2000; Gehrels and Pecha, 2014). These model 2500, 2150, and 1750 Ma (~27% of the sample). 200 300 400 ages or crustal residence ages for zircon provide Age (Ma) a qualitative estimate of the time of separation King Christian Formation of the Heiberg of the source rocks, or their precursors, from a Group Figure 7. Plot shows eHf values for sam- hypothetical depleted mantle reservoir (Bahl- The Sinemurian to Pliensbachian King ple data from the Heiberg Formation. burg et al., 2011). While these model ages do Christian Formation sample is from western Only grains younger than 340 Ma were analyzed. Solid lines isolate depleted not necessarily provide real age information, it , where the formation mantle (DM) and chondritic uniform is useful for comparative analyses. When com- is the thickest (180 m) in the basin, and here mantle (CHUR). See Table DR3 (see text

paring the model ages (TDM) of this study with the unit is interpreted to be a deltaic deposit footnote 1) for error measurements. other published isotopic studies, no conversion (Embry, 1983b). Zircon ages have two distinct Dashed lines separate fields described was applied to the T if it was originally calcu- age ranges—specifically Paleozoic and Precam- as juvenile (0–5 epsilon units below DM DM), intermediate (5–12 epsilon units lated from eNd values. Full Hf analytical results brian age fractions. The Paleozoic ages range below DM), and evolved (>12 epsilon can be found in Table DR3. from ca. 441 to 262 Ma with peaks at 425 and units below DM) following Bahlburg et 275 Ma (~18% of the sample). The al. (2011) and Gehrels and Pecha (2014). RESULTS can be subdivided into a -age range and a Paleo- to Mesoproterozoic range. The Tonian U-Pb Analyses zircon, specifically ca. 1020–910 Ma, makes up to the proposed northern sediment source. At 8% of the sample. The predominant age range Bunde Fiord, ochreous horizons were com- Bjorne Formation is ca. 2100–880 Ma with peaks at 2000, 1650, monly observed in marine mudstones of the The Bjorne Formation is more than 1000 m 1450, and 1200 Ma (73% of the sample), which Blind Fiord, Murray Harbour, Hoyle Bay, and thick along the southern and eastern margins of is similar to zircon age distribution in the Bjorne Barrow formations (Fig. 8). The ochre layers the Sverdrup Basin and, along with the Blind Formation (Fig. 6). are typically less than 10 cm thick and later- Fiord Formation, represents the first significant ally continuous, and they are interpreted to be episode of siliciclastic deposition in the basin. Hf-Isotope Analyses volcanic ash beds due to their field properties It has been subdivided into the predominantly (color, moisture, and plasticity). X-ray powder sandstone Cape Butler, Pell Point, and Cape Twenty-four zircon grains were analyzed diffraction (XRD) using CuKa radiation with O’Brien members (Embry, 1986). A single to provide further information on the source a scanning speed of 1°2q/min show that these sample was analyzed for detrital-zircon geo- of Permian–Triassic zircon from the Lower ash beds consist of quartz, hydrotalcite, illite, chronology from the Cape O’Brien Member Heiberg and King Christian formations. The jarosite, with lesser zircon and halloysite (Fig. (Olenekian) on central Ellesmere Island. Zircon eHf values range from +16 to -17 with no dis- 8A). Halloysite is derived from the dissolution ages fall into three modes—a tight late Silu- cernible eHf groupings within the data (Fig. 7). of volcanic glass or weathered volcanic ash rian peak (ca. 460–420 Ma); a broad range of There is no trend (R2 = 0.06) for eHf values with a crystalline structure similar to kaolinite Paleo- to Mesoproterozoic (ca. 2100–950 Ma); becoming more depleted as the U-Pb age of (Joussein et al., 2005). Similarly, hydrotalcite

and (ca. 3000–2500 Ma) ages (Fig. 6). the zircon decreases. TDM values show a broad is formed from volcanic glass, and it has been The robust, Caledonian-aged peak has 10 grains spectrum, but the majority of ages are Meso- to produced experimentally by a reaction between (~10% of the sample population), with the major Neoproterozoic. basaltic glass and seawater (Crovisier et al., peak at 430 Ma. The Mesoproterozoic range, 1982). Hall and Stamatakis (2000) observed the primary age fraction, has 74 grains (~76% Volcanic Ash Beds hydrotalcite infilling the molds left by the dis- of the sample) with several peaks at 1650, 1450, solution of volcanic glass shards. Collectively, and 1100 Ma. The Archean range has 12 grains The study location along northern Axel Hei- the presence of volcanically derived minerals (~13% of the sample) with two minor peaks at berg Island (Fig. 2) represents a geographically and macroscopic textural characteristics confirm 2760 and 2530 Ma. proximal location within the Sverdrup Basin that the ochreous layers are volcanic ash beds.

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 675

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

20000 1300 Heiberg Formation Hydrotalcite Hydrotalcite Late A B Barrow Formation

Quartz Tr 1200

Gypsum Quartz iassic (Norian) Zircon Sill Jarosite 1100 Jarosite Sill Carbonate 15000 1000 Sandstone Barrow Formation Ash bed occurrence Sill Mudstone/siltstone 900 Halloysite Halloysite Quartz Quartz Illite Quartz 800 Zircon Late Poor exposure Hoyle Bay (Carnian)

10000 Formation Tr 700

iassi c Sill 600

Intensity (counts) Pat Bay Formation T

Middle 500 riassi c Murray Harbour Hoyle Bay Formation Formation 400 5000 Early 300 Roche Point Formation Blind Fiord Tr

Formation iassi c 200 Murray Harbour Formation

100 0 Blind Fiord Formation 10.000 20.000 30.000 40.000 50.000 60.000 0 2θ (deg.) m Permian

Figure 8. (A) X-ray diffraction traces from volcanic ash beds; and (B) stratigraphic log from Bunde Fiord on Axel Heiberg Island, highlighting stratigraphic location of volcanic ash beds. Ash beds not to scale.

DISCUSSION associated with the Uralian orogeny in the Tai- Late Triassic Detrital-Zircon Provenance myr region in central Russia (e.g., Vernikovsky Early Triassic Detrital-Zircon Provenance et al., 1995; Zhang et al., 2013). The majority Pat Bay Formation of Permian volcanic rocks are observed within Detrital-zircon U-Pb age data of the Pat The detrital-zircon signature from the Bjorne the northern part of the Sverdrup Basin, which Bay Formation (Miller et al., 2006; Omma et Formation is remarkably similar to samples suggests there were probably volcanic equiva- al., 2011) vary considerably between samples from the Devonian clastic wedge (Anfinson et lents to the north of the basin. The Lower Perm- indicative of two different sources during depo- al., 2012a) and is consistent with sample AE1 ian volcanic rocks in the Sverdrup Basin (the sition (e.g., Embry, 2009). Sample AE2 from reported by Miller et al. (2006). The detrital-zir- Esayoo, the Unnamed Lower volcanics [ULV], Miller et al. (2006) has an assemblage of ages con signature of the Bjorne Formation is similar or equivalents) could provide the limited age that overlap within the Devonian clastic wedge to Triassic strata from the North Slope of Alaska range of Permian zircon within the Blind Fiord spectrum, including a single 376 Ma age, an age (Gottlieb et al., 2014) and similar aged strata Formation; basalts may be able to supply zir- fraction of ca. 620–505 Ma, and a broad range from the northwestern Cordillera (Fig. 9A), the con even though they typically have poor zircon of Paleo- to Mesoproterozoic ages. Miller et al. latter being recycled from the strata of north- potential (Rioux et al., 2012; Candan et al., 2015; (2006) hypothesized the sample represented sed- western Laurentia (Beranek et al., 2010b). Col- Iles et al., 2015). The converted eNd values and iment derived from north of the Sverdrup Basin

lectively the geographically dispersed areas sug- the TDM from the Esayoo volcanics (ca. 276 Ma) because those ages were unknown in northern gest a common, areally expansive source, which are comparable to Hf isotope data from simi- Canada at the time. Subsequent studies (e.g., most probably was the Devonian clastic wedge lar aged zircon from this study (Fig. 10). Long Lemieux et al., 2011; Anfinson et al., 2012a) and equivalents (Fig. 9B), and so provenance is distance transport from the Urals is unlikely to report abundant ca. 700–500 Ma zircon ages from a recycled source. produce the tight range of Permian ages within from rocks of the Late Devonian clastic wedge A sample from the Blind Fiord Formation has the Blind Fiord Formation, particularly because as well as the Silurian flysch (Beranek et al., a wholly different signature interpreted to repre- Uralide granitoids formed at an almost constant 2015). A prominent ca. 700–500 Ma age-frac- sent the northern source to the basin (Omma et rate from 370 Ma to 250 Ma, older in the south tion coupled with an absence of Permian–Tri- al., 2011). Notable is the occurrence of a suite of and younger in the north (Vernikovsky et al., assic zircon ages would suggest a source much Permian ages (ca. 290–265 Ma) and the absence 1995; Bea et al., 2002). Granitoids in the north- like Silurian and Devonian strata in the Franklin- of Caledonian and Ellesmerian orogen ages that ern Urals have ages from ca. 300 to 280 Ma, ian Basin; these strata were probably ultimately typify the Devonian clastic wedge (ca. 700–360 and post-tectonic granitoids dated at ca. 260 Ma derived from rocks in Arctic Alaska–Chukotka Ma). Omma et al. (2011) suggested the source (Pease et al., 2015). Transport from the Urals is of Timanide age (e.g., Cecile et al., 1991; Amato of the young ages was either Early Permian possible but would likely provide a broad spec- et al., 2009, 2014). In contrast, the later work of basaltic magmatic activity within the basin (e.g., trum of U-Pb ages rather than narrow peaks as Omma et al. (2011) reported a prominent range Thorsteinsson, 1974), or mid-Permian syenites seen in the Blind Fiord Formation. of near-syndepositional ages (ca. 255–217 Ma),

676 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

Active Margin Source Recycled Source Franklinian Devonian Franklinian Devonian A Clastic Wedge Mesoproterozoic Clastic Wedge Mesoproterozoic E. Cretaceous Isachsen Fm y E. Cretaceous Deer Bay Fm M. Jurassic iassi c

Tr Sandy Point Fm Age Probabilit

Permian- E. Jurassic King Christian L. Triassic Fm Lower Heiberg Fm

Sverdrup Basin

L. Triassic L. Triassic Pat Bay Fm Pat Bay Fm iassi c Tr

E. Triassic E. Triassic Blind Fiord Fm Permian- Bjorne Fm

Triassic, Chuktoka

Triassic, Wrangel Russia Triassic Triassic Alaska Lisburne Hills North Slope

Triassic Yukon Triassic YTT Overlap YTT Platform

Dev. Clastic Wedge

0 200 400 600 800 1000 1200 1400 1600 1800 200 400 600 800 1000 1200 1400 1600 1800 2000 Age (Ma)

1.0 B 0.9

0.8

Sverdrup Basin 0.7 Active Margin Source King Christian Fm

y Lower Heiberg Fm 0.6 Pat Bay Fm (Miller) Pat Bay Fm (Omma) Bjorne Fm 0.5 200 Ma Blind Fiord Fm Russia 0.4 Chukotka Wrangel Island

Cumulative Probabilit Alaska 0.3 North Slope Lisburne Hills Yukon Tanana Terrane 0.2 Recycled Source YTT Platform YTT Overlap 0.1 Devonian Clastic Wedge Franklinian Devonian Clastic Wedge 0 0 500 1000 1500 2000 2500 3000 Age (Ma) Figure 9. Compiled results from this study: (A) relative probability distributions and (B) cumulative probability plots. Illustration of the detrital age spectra grouped by signatures of the “Active margin source” and the “Recycled source.” Data from the Isachsen Formation (Røhr et al., 2010), from studies in the Canadian Arctic (Miller et al., 2006; Omma et al., 2011; Anfinson et al., 2012a), from Russia (Miller et al., 2006, 2010; Tuchkova et al., 2011; Amato et al., 2015), from Alaska (Miller et al., 2006; Gottlieb et al., 2014), and from Yukon (Beranek et al., 2010b; Beranek and Mortensen, 2011).

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 677

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

New Siberian Islands formations are similar to those reported from A basement rocks Triassic strata of the AACM from Lisburne Hills, Western Siberia North Slope of Alaska, Wrangel Island, and Chu- Carb.-Tr. igneous rocks kotka (Fig. 4). Samples from the Upper Triassic

Sverdrup Basin Otuk Formation from the Lisburne Hills have Esayoo Volcanics an age fraction from ca. 275–220 Ma, in addi- tion to Carboniferous, Lower Paleozoic ages, Arctic Alaska Seward Peninsula and minor Mesoproterozoic ages (Phanerozoic peaks at 420, 355, 315, 255, and 220 Ma) (Miller Franklinian Basin et al., 2006). The detrital-zircon signature from Triassic strata along the North Slope is remark- ably similar to the Devonian clastic wedge with This study a strong age fraction of ca. 700–360 Ma and a broad spectrum of Proterozoic ages (Gottlieb et 0 500 1000 1500 2000 2500 al., 2014). When comparing detrital-zircon data

TDM (Ma) from the Lisburne Hills and the North Slope, the B two regions have notably different age distribu- tions, which strongly resemble the pattern of the 15 Depleted Mantle two different provenance signatures within the Sverdrup Basin (Fig. 9). Similar to Lisburne Hills, Triassic samples 10 from Wrangel Island and Chukotka document a

JUVENILE general assemblage of near-syndepositional zir- 5 con ages and a minor representation of Mesopro- terozoic ages (Miller et al., 2006, 2010; Tuch- kova et al., 2011; Amato et al., 2015). More CHUR INTERMEDIATE 0 specifically, Wrangel Island samples have nearly continuous ages from ca. 480 to 205 Ma (Pha- εHf Value s n nerozoic peaks at 440, 350, 305, 250, and 230 -5 Ma) (Miller et al., 2010), and samples from

EVO L Chukotka (Miller et al., 2006; Tuchkova et al., 2011; Amato et al., 2015) have a range of ca. -10 VED 405–216 Ma (robust Phanerozoic peaks at 500, average crustal evolutio T 300, and 250 Ma). Both regions have minor con- DM tributions from the Paleoproterozoic to Silurian. -15 In summary, both the Sverdrup Basin and the broader AACM share the active margin source, 200 400 600 800 1000 1200 1400 1600 1800 2000 which is consistent with the rotational model for Age (Ma) the opening of the Arctic Ocean. Alternatively,

Figure 10. (A) Hafnium-depleted model ages (TDM) calculated from either eNd or eHf. (B) eHf values for Zhang et al. (2015) presented a reconstruction in sample data and relevant Hf-isotope data comparisons. CHUR—Chondritic uniform reservoir. Data which littoral currents allow for the redistribu- sets are as follows: detrital zircon from Neoproterozoic to Upper Devonian strata of the Franklinian tion of detrital material from the Polar Urals and Basin (Anfinson et al., 2012b); Carboniferous to Triassic zircon from igneous rocks of the northwest- ern Siberian Craton (Malitch et al., 2010); New Siberian Islands basement rocks (Akinin et al., 2015); Taimyr during the Triassic due to the sublitho- Neoproterozoic to Devonian igneous rocks of Seward Peninsula, Arctic Alaska, converted from eNd spheric spreading associated with the Siberian values (Amato et al., 2009); Lower Permian Esayoo volcanics from eastern Sverdrup Basin ca. 276 mantle plume at ca. 250 Ma, which does not Ma, converted from eNd values (Morris, 2013). Gray box and black arrow represent interpreted aver- explain the post–Siberian Trap range of detrital- age crustal evolution trajectories assuming present-day 176Lu/177Hf = 0.0093 (Vervoort and Patchett, zircon ages of 240–210 Ma. 1996; Bahlburg et al., 2011; Gehrels and Pecha, 2014). Yukon Tanana Terrane

in addition to a broad Paleozoic spectrum (ca. and Blind Fiord formations, there is a promi- A similar pattern of dual detrital-zircon spec- 490–295 Ma), which is indicative of derivation nent near-syndepositional–age fraction in the tra is observed along the northwestern margin from the active margin source region. Lower Heiberg Formation with a continuous of Laurentia in the Yukon. Sediment recycled spectrum of ca. 300–215 Ma ages. In contrast from western interior basins of Laurentia has a Lower Heiberg Formation to the Pat Bay and Blind Fiord formations, there detrital-zircon signature consistent with deriva- The Lower Heiberg Formation (Romulus is a notable absence of Mesoproterozoic ages tion from a Devonian clastic wedge-type source Member) sample has a similar detrital-zircon (Fig. 9A) that are typically present in Laurentian (Beranek et al., 2010b)—namely broad Phanero- signature to the Pat Bay sample analyzed by strata (e.g., Hadlari et al., 2012). zoic peaks at 450, 420, and 365 Ma—and a wide Omma et al. (2011). As with the two previous The consistent zircon signatures from spectrum of Proterozoic and Lower Paleozoic northerly-derived samples from the Pat Bay the Blind Fiord, Pat Bay, and Lower Heiberg ages (YTT platform, Fig. 9). This signature

678 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

remains relatively stable through Early to Late reincorporated during the Permian–Triassic 2011; Anfinson et al., 2012b) derived from an Triassic strata. magmatism in the AACM. Arctic Alaska–Chukotka arc such as at Seward The other provenance signature is from sedi- The other proposed source for Permian–Tri- Peninsula (Amato et al., 2009); it is logical that a ment derived from the west, along the active assic zircon in the Sverdrup Basin is from the similar arc in the Triassic (Fig. 11) would incor-

margin of the Yukon-Tanana terrane (YTT). The Siberian region (e.g., Urals, Taimyr, and NSI: porate a similar evolved eHf signature and TDM. YTT is a pericratonic terrane that comprises a Fig. 1; Omma et al., 2011; Miller et al., 2013, of deformed arc, backarc, and continental and references therein). Miller et al. (2013) Early Jurassic Detrital-Zircon Provenance margin assemblages (Colpron et al., 2006; Nel- and Zhang et al. (2015) recently outlined pos- son et al., 2006). During the Middle Permian sible sources of the younger zircon ages, which The Lower Jurassic King Christian Forma- it was separated from ancestral North America included the Carboniferous to Early Permian tion sample has a similar cumulative trend to by a backarc basin, the Slide Mountain Ocean plutonic belts from the northernmost Urals; sediment that would be sourced from erosion of (Plafker and Berg, 1994); however, the closure granites as young as ca. 250 Ma from the south- the Devonian clastic wedge. There is a peak at of the backarc basin in the Late Permian and ern Urals; mafic magmatism (ca. 252 Ma) from 420 Ma and a broad range of Proterozoic ages shortening across the YTT and the western mar- the initiation of the Siberian Traps; earliest Tri- that are similar to the detrital-zircon signature of gin of Laurentia (Klondike orogeny) formed a assic felsic and mafic magmatism in the Taimyr the Bjorne Formation (Fig. 9A). The difference Triassic foreland to backarc basin to the east region; Permian–Triassic to Triassic rift-related between the King Christian Formation from the of the YTT (Beranek and Mortensen, 2011). magmatism in the Kara Sea; and granites and southerly-sourced Bjorne Formation is the pres- Detrital-zircon signatures from Triassic strata syenites (ca. 249–241 Ma) and basalts from ence of three young zircon grains (315, 279, and sourced from the convergent, western margin southern and central Taimyr. The eHf values 262 Ma), which could have been cannibalized of the YTT exhibit a suite of near-syndeposi- from Carboniferous to Triassic magmatic rocks from Triassic strata. Zircon assemblages in the tional Permian–Triassic zircon ages (Beranek along the northwestern part of the Siberian Cra- Middle Jurassic to Early Cretaceous (Sandy and Mortensen, 2011). ton are substantially more juvenile (Malitch et Point, Deer Bay, and Isachsen formations) are The two provenance signatures from Triassic al., 2010) than similar-aged zircon within the consistent with the provenance signature of the

strata of the YTT, specifically an active mar- Sverdrup Basin. The TDM of the Siberian zir- Devonian clastic wedge with minor recycled

gin source and a recycled source, share a close con are younger than the TDM from this study, Permian–Triassic zircon (Røhr et al., 2010; resemblance to the dual provenance signature of which suggests that the Siberian region was not Omma et al., 2011; Fig. 9A). Triassic strata of the Sverdrup Basin and Arctic a source region for the Sverdrup Basin during Previous authors (Embry, 1993; Embry and Alaska (Fig. 9B). This suggests that the tectonic the Triassic. The eHf data from basement rocks Beauchamp, 2008; Embry, 2009) proposed that setting for YTT basin could have been similar of the NSI are within the spectrum of eHf and Jurassic extension dismembered the northern

to the tectonic setting of the Sverdrup Basin TDM values of Arctic Alaska and Devonian strata sediment source region and prevented com- during the Triassic. from northern Canada, and these data inter- munication with the Sverdrup Basin by trap- preted to represent juvenile magmatic addition ping sediment in extensional basins such as the Provenance of Permian–Triassic Zircon during the Neoproterozoic (Akinin et al., 2015). proto–Amerasia Basin. The new detrital-zircon The new observations of volcanic ash beds data from the Heiberg Group suggest that the Identifying source(s) for the consistently throughout the Triassic section on Axel Heiberg provenance change occurred below the King near-syndepositional–aged zircon in Trias- Island are noteworthy because they are a record Christian Formation and above the Romulus sic strata throughout the Sverdrup Basin and of ash air fall on the northern side of the basin. Member, which is equivalent to the Skybattle AACM is essential to understanding the tectonic To date there have been no observations of volca- Formation (Fig. 3; ca. 210–190 Ma). After depo- setting of the Sverdrup Basin during the early nic ash beds in Triassic strata from the southern sition of the King Christian Formation in the Mesozoic. The broad range of eHf values from Sverdrup Basin, implying that ash transport was Early Jurassic, there was no longer a supply of this study of U-Pb zircon ages between 330 confined to the northern part of the basin and not near-syndepositional zircon into the basin. The and 200 Ma is interpreted to record a mixed carried a great distance; otherwise they would Early Jurassic onset of rifting is supported by source with a combination of juvenile, inter- be widely distributed throughout the basin. A the presence of upper Heiberg Group strata at mediate, and evolved crust (Fig. 7). Other eHf stratigraphic section of Triassic strata from ~200 the base of half-grabens on Prince Patrick Island or eNd values from the circum-Arctic display km south of Bunde Fiord had no observable vol- (Harrison and Brent, 2005). Deposition of the

a variety of isotopic compositions and TDM. canic ash beds (Embry, 1983b); therefore, Trias- Isachsen Formation occurred after the breakup The ca. 710–380 Ma zircon population from sic zircon being sourced from the north of the unconformity at ca. 130 Ma (Embry and Beau- Devonian strata of the Franklinian Basin pro- basin supports the sediment provenance direc- champ, 2008). Detrital-zircon ages from the vides eHf values (Anfinson et al., 2012b) that tion suggested by Embry (2009), rather than the Lower Cretaceous Isachsen sandstone exhibit are congruent with a juvenile to intermediate near-syndepositional zircon originating from a only a single grain (ca. 244 Ma) younger than source (Fig. 10B) in the AACM. Neoproterozoic great distance east of the basin (e.g., Urals, Tai- 400 Ma (Røhr et al., 2010). to Devonian igneous rocks from Arctic Alaska myr; Omma et al., 2011; Miller et al., 2013). If have converted eNd values that are slightly less the active margin outboard of the YTT along the Tectonic Setting of the Basin juvenile than the Devonian strata (Amato et al., western margin of Laurentia was generating the

2009). The TDM of both studies overlap with the Permian–Triassic zircon observed within Trias- The abundance of Permian–Triassic detrital

TDM of Permian–Triassic zircon of the Lower sic strata, then the same tectonic process could zircon and confirmation of Triassic volcanic ash Heiberg Formation with predominantly Meso- generate similar-aged zircon along the outboard beds in the Sverdrup Basin suggests that the to Neoproterozoic model ages (Fig. 10A). It is margin of the AACM. This is consistent within region to the north was tectonically active dur- plausible that the lithospheric source involved the framework of the rotational model, with the ing the Triassic, and likely part of the Permian during the formation of the older zircon was Devonian clastic wedge (e.g., Lemieux et al., as well. We interpret that the combination of

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 679

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

Triassic

NSC (TA, KT)

SOUTH ANYUI CH Figure 11. Evolution of the northern margin of Laurentia OCEAN adapted from paleographical maps of Plafker and Berg COLL (1994) and Nokleberg et al. (2000). Outline of Yukon Tanana Hanna SVERDRUP NSV, Tr BASIN terrane (YTT; orange polygon) is from Nelson et al. (2006). Ar . KN ctic Axis of Hanna Trough and Arctic Alaska Basin from Gottlieb Bas Alaska

et al. (2014). Proposed tectonic model: (A) in the Triassic, the in Sverdrup basin occupied a retroarc position to Chukotka ANGAYUCHAM NAC along strike with the Triassic basin adjacent to the Yukon OCEAN Tanana terrane arc; (B) Early Jurassic extension to form the NAM proto–Amerasia Basin cutoff the Sverdrup Basin from the Arctic Alaska–Chukotka microplate (AACM); and (C) in the BASIN Early Cretaceous during the postrift stage of Sverdrup Basin, the Amerasia and Canada basins opened as the South Anyui Ocean closed. YTT NAM

CACHE CREEK COLL A OCEAN Early Jurassic INITIATION OF RIFT GRABENS Legend CH OPENING OF PROTO- AMERASIA Triassic dZr locations & CANADA BASINS SVERDRUP Ar BASIN SOUTH ctic Provenance direction Basin ANYUI Alas

OCEAN k Active margin sediment a AA NAC Recycled sediment

Tectonic features and contacts NAM ANGAYUCHAM OCEAN Subduction zone. Barbs extend YTT towards subducting margin NAM COLL Backarc spreading, barbs face spreading basin FARALLON B OCEAN Accreted terranes and extinct terranes Early Cretaceous Continental-margin arc, island arc, or turbidite basin OPENING OF PROTO- AMERASIA BASIN CH COLLAGE OF RIFTED Accretionary wedge complex TERRANES: NR,CK,AS,AP DEVELOPMENT OF SVERDRUP RIM Acronyms: AA—Arctic Alaska; CH—Chukotka; OPENING OF PROTO- COLL—Collage of accreted terranes; NAC—North CANADA BASIN American Craton; NAM—North American Margin; AA YTT—Yukon Tanana Terrane NAC Early Cretaceous: AP—Artis Plateau; CK—Chilliwack River; CS—Chukchi Spur; NR—Northwind Ridge NAM

Triassic: KN—Kular Nera; KT—Kotel’nyi Terrane; YTT

NSC—North Siberia Craton; NSV—Verkhoyansk NAM Foldbelt; TA—Taimyr FARALLON C OCEAN COLL

680 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

volcanic ash beds and identification of remark- Triassic—an active margin source to the north Amato, J.M., Toro, J., Miller, E.L., Gehrels, G.E., Farmer, G.L., Gottlieb, E.S., and Till, A.B., 2009, Late Proterozoic–Pa- ably similar dual detrital-zircon provenance pat- and a recycled source to the south and east of leozoic evolution of the Arctic Alaska–Chukotka terrane terns from the Sverdrup Basin, YTT, and AACM the basin. These two signatures are similar to the based on U-Pb igneous and detrital-zircon ages: Implica- allows for an interpretation of the outboard mar- two distinct sources observed in Arctic Alaska tions for Neoproterozoic paleogeographic reconstruc- tions: Geological Society of America Bulletin, v. 121, gin of the AACM as convergent, as shown in and the YTT. A hypothetical northwestern con- no. 9–10, p. 1219–1235, doi:​10​.1130​/B26510​.1. Figure 11 (adapted from Plafker and Berg, 1994; vergent margin of Laurentia would explain the Amato, J.M., Aleinikoff, J.N., Akinin, V.V., McClelland, W.C., and Toro, J., 2014, Age, chemistry, and correlations of Nokleberg et al., 2000), thereby providing a geo- remarkable similarity in the source age profiles Neoproterozoic–Devonian igneous rocks of the Arctic dynamic model for rifting in the Jurassic con- of the Sverdrup Basin and the overlap assem- Alaska–Chukotka terrane: An overview with new U-Pb current with closure of the South Anyui ocean blage between YTT and the North American cra- ages, in Till, A.B., and Dumoulin, J.A., eds., Reconstruc- tion of a Late Proterozoic to Devonian Continental Mar- basin. The Nokleberg et al. (2000) reconstruc- ton. The presence of volcanic ash beds through- gin Succession, Northern Alaska, Its Paleogeographic tion is modified by indicating that rifting of the out Triassic strata, apparently restricted to the Significance, and Contained Base-Metal Sulfide Depos- its: Boulder, Colorado, Geological Society of America proto–Amerasia Basin developed as a retroarc northern margin of the basin, indicates that Special Paper 506, p. 29–58, doi:​10​.1130​/2015​.2506​(02). system driven by a subducting slab, which in volcanic activity that lasted more than ~50 m.y Amelin, Y., Lee, D.-C., and Halliday, A.N., 2000, Early-middle turn caused AACM to rift from northern Lau- was relatively proximal to this northern margin, Archaean crustal evolution deduced from Lu-Hf and U-Pb isotopic studies of single zircon grains: Geochi- rentia (as proposed by Kuzmichev, 2009). Dur- which is consistent with a convergent outboard mica et Cosmochimica Acta, v. 64, p. 4205–4225, doi:​10​ ing slab rollback, retroarc extension can occur margin of the AACM during the Triassic. Previ- .1016​/S0016​-7037​(00)00493​-2. far inboard from the convergent plate bound- ous interpretations attributed the source of the Anfinson, O.A., Leier, A.L., Embry, A.F., and Dewing, K., 2012a, Detrital zircon geochronology and provenance of the ary (Lawton and McMillan, 1999; Hadlari and Permian–Triassic zircon to be in Russia, specifi- Neoproterozoic to Late Devonian Franklinian basin, Ca- Rainbird, 2011), which is how we interpret the cally the Siberian Traps, Uralian granites, and nadian Arctic Islands: Geological Society of America Bulletin, v. 124, no. 3–4, p. 415–430. Jurassic to Early Cretaceous rifting of the Sver- Taimyr magmatism, none of which make for an Anfinson, O.A., Leier, A.L., Gaschnig, R., Embry, A.F., and Dew- drup Basin. Our detrital-zircon data support the ideal candidate for the spectrum of ages, par- ing, K., 2012b, U-Pb and Hf isotopic data from Franklin- hypothesis of Embry (2009) that crustal exten- ticularly for the Middle to Late Triassic zircon ian Basin strata: Insights into the nature of Crockerland and the timing of accretion, Canadian Arctic Islands: Ca- sion dismembered the northern sediment source within sandstones of the Sverdrup Basin. The nadian Journal of Earth Sciences, v. 49, no. 11, p. 1316– from the Sverdrup Basin and show that this tran- more negative eHf values of Permian–Triassic 1328, doi:​10​.1139​/e2012​-067. zircon compared to juvenile values of Permian– Bahlburg, H., Vervoort, J.D., and DuFrane, S.A., 2011, Plate sition occurred in the Early Jurassic (Fig. 11), in tectonic significance of Middle Cambrian and Ordovi- contrast to Arctic Alaska, which continued to Triassic igneous rocks in Siberia support that cian siliciclastic rocks of the Bavarian Facies, Armorican receive sediment from Chukotka in the Jurassic the provenance was from the AACM instead Terrane Assemblage, Germany—U-Pb and Hf isotope evidence from detrital zircons: Research, and into the Albian–Aptian (Moore et al., 2015). of western Siberia. The contribution of U-Pb v. 17, no. 2, p. 223–235. The lack of near-syndepositional zircon in detrital-zircon geochronology, eHf isotope Balkwill, H.R., 1978, Evolution of Sverdrup Basin: The Ameri- Jurassic–Cretaceous strata is important to under- values, and newly described volcanic ash beds can Association of Petroleum Geologists Bulletin, v. 62, p. 1004–1028. standing the nature of the basin because during from Triassic strata of the Sverdrup Basin help Bea, F., Fershtater, G., and Montero, P., 2002, Granitoids of the the synrift phase (ca. 190–130 Ma), there is a lim- to develop an argument that active magmatism Uralides: Implications for the evolution of the orogeny, in Brown, D., Juhlin, C., and Puchkov, V., eds., Moun- ited record of magmatism other than the single north of the basin supplied Permian–Triassic tain Building in the Uralides: Pangea to the Present: 158 Ma zircon grain within the Deer Bay Forma- zircon. Furthermore, the provenance signature American Geophysical Union Geophysical Monograph, tion. The breakup unconformity (130 Ma; see from Jurassic strata suggests initial rifting iso- v. 132, p. 211–232. Beranek, L.P., and Mortensen, J.K., 2011, The timing and prov- Embry and Beauchamp, 2008) coincides with lated the Sverdrup Basin from the active margin enance record of the Late Permian Klondike orogeny in the onset of extensive magmatism in the Sver- source by the Early Jurassic. northwestern Canada and arc-continent collision along drup Basin (e.g., Evenchick et al., 2015). Postrift- western North America: Tectonics, v. 30, no. 5, TC5017, doi:​10​.1029​/2010TC002849. phase magmatism persisted for more than 30 m.y. ACKNOWLEDGMENTS Beranek, L.P., Mortensen, J.K., Lane, L.S., Allen, T.L., Fraser, The South China Sea provides an analog for This work was supported by the Geological Survey of Can- T.A., Hadlari, T., and Zantvoort, W.G., 2010a, Detrital zir- ada and the University of Ottawa. Fieldwork was assisted con geochronology of the western Ellesmerian clastic the Sverdrup and Amerasia basins. The South by the Polar Continental Shelf Program. Dr. Carol Evenchick wedge, northwestern Canada: Insights on Arctic tecton- China Sea formed in a retroarc setting and had provided a sample of the King Christian Formation, and the ics and the evolution of the northern Cordilleran mio- authors are grateful to Dr. Ashton Embry for guidance in the limited synrift magmatism and persistent mag- geocline: Geological Society of America Bulletin, v. 122, field in 2011. We are grateful for Tara Kell’s assistance with the no. 11–12, p. 1899–1911, doi:​10​.1130​/B30120​.1. matism in the postrift phase (Franke, 2013). Slab XRD analysis. Ray Chung, Ron Christie, and Tom Pestaj are Beranek, L.P., Mortensen, J.K., Orchard, M.J., and Ullrich, T., roll back is thought to have been the driving thanked for technical assistance preparing mineral separates 2010b, Provenance of North American Triassic strata and maintaining the SHRIMP ion probe at the Geochronology from west-central and southeastern Yukon: Correlations mechanism in South China Sea causing backarc Laboratory in Ottawa. We thank Victoria Pease, Elizabeth L. with coeval strata in the Western Canada Sedimentary extension (Zhou and Li, 2000; Doust and Sumner, Miller, Jeff Amato, and associate editor Damian Nance for Basin and Canadian Arctic Islands: Canadian Journal of their thoughtful comments that helped improve the manu- Earth Sciences, v. 47, no. 1, p. 53–73, doi:10​ .1139​ /E09​ -065.​ 2007). Rifting began at 110 Ma (Li et al., 2008), script. This is a Circum-Arctic Lithosphere Evolution (CALE) Beranek, L.P., Pease, V., Hadlari, T., and Dewing, K., 2015, Si- contribution. This is GSC Contribution #20160017. and generation of oceanic crust likely began ca. lurian flysch successions of Ellesmere Island, Arctic 40–30 Ma (Franke, 2013, and references therein), Canada, and their significance to northern Caledonian which indicates a protracted period of extension REFERENCES CITED palaeogeography and tectonics: Journal of the Geo- Akinin, V.V., Gottlieb, E.S., Miller, E.L., Polzunenkov, G.O., logical Society, London, v. 172, no. 2, p. 201–212, doi:​ with no associated magmatic activity followed Stolbov, N.M., and Sobolev, N.N., 2015, Age and com- 10​.1144​/jgs2014​-027. by extensive postrift magmatism, similar to the position of basement beneath the De Long archipelago, Blichert-Toft, J., 2008, The Hf isotopic composition of zircon Arctic Russia, based on zircon U-Pb geochronology and reference material 91500: Chemical Geology, v. 253, Cretaceous record of the Sverdrup Basin and O-Hf isotopic systematic from crustal xenoliths in ba- no. 3–4, p. 252–257, doi:​10​.1016​/j​.chemgeo​.2008​.05​.014. possibly also Amerasia Basin. salts of Zhokov Island: Arktos, v. 1, p. 1–10, doi:​10.1007​ ​ Bouvier, A., Vervoort, J.D., and Patchett, P.J., 2008, The Lu-Hf /s41063​-015​-0016​-6. and Sm-Nd isotopic composition of CHUR: Constraints Amato, J.M., Toro, J., Akinin, V.V., Hampton, B.A., Salnikov, from unequilibrated chondrites and implications for the CONCLUSION A.S., and Tuchkova, M.I., 2015, Tectonic evolution of the bulk composition of terrestrial planets: Earth and Plan- Mesozoic South Anyui suture zone, eastern Russia: A etary Science Letters, v. 273, no. 1, p. 48–57, doi:10​ .1016​ ​ critical component of paleogeographic reconstructions /j​.epsl​.2008​.06​.010. The Sverdrup Basin has two identifiable of the Arctic region: Geosphere, v. 11, no. 5, p. 1530– Buchan, K.L., and Ernst, R.E., 2006, Giant dyke swarms detrital-zircon provenance signatures during the 1564, doi:​10​.1130​/GES01165​.1. and the reconstruction of the Canadian Arctic islands,

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 681

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 MIDWINTER ET AL.

Greenland, Svalbard and Franz Josef Land, in Hanski, E., Embry, A.F., 2009, Crockerland—The Source Area for the Trias- PLoS ONE, v. 10, no. 12, p. e0144727, doi:10​ .1371​ /journal​ ​ Mertanen, S., Rämö, T., and Vuollo, J., eds., Dyke sic to Middle Jurassic Strata of Northern Axel Heiberg .pone​.0144727. Swarms: Time Markers of Crustal Evolution: Nether- Island, Canadian Arctic Islands: Bulletin of Canadian Hall, A., and Stamatakis, M.G., 2000, Hydrotalcite and an lands, Balkema Publishers/Taylor and Francis, p. 27–48. Petroleum Geology, v. 57, no. 2, p. 129–140, doi:​10.2113​ ​ amorphous clay mineral in high-magnesium mud- Cameron, B.I., and Muecke, G.K., 1996, Permian alkaline ba- /gscpgbull​.57​.2​.129. stones from the Kozani Basin, Greece: Journal of Sedi- salts associated with formation of the Sverdrup Basin, Embry, A.F., 2011, Petroleum prospectivity of the Triassic–Ju- mentary Research, v. 70, no. 3, p. 549–558, doi:10​ .1306​ ​ Canadian Arctic: Canadian Journal of Earth Sciences, rassic succession of Sverdrup Basin, Canadian Arctic /2DC40927​-0E47​-11D7​-8643000102C1865D. v. 33, no. 10, p. 1462–1473. Archipelago, in Spencer, A.M., Gautier, D., Stoupakova, Harrison, J.C., and Brent, T.A., 2005, Basins and fold belts Candan, O., Koralay, O.E., Topuz, G., Oberhänsli, R., Fritz, H., A., Embry, A., and Sørensen, K., eds., Arctic Petroleum of Prince Patrick Island and adjacent areas, Canadian Collins, A.S., and Chen, F., 2015, Late Neoproterozoic gab- Geology: Geological Society, London, Memoir, v. 35, Arctic Islands: Geological Survey of Canada Bulletin, bro emplacement followed by early Cambrian eclogite- p. 545–558. v. 560, 197 p. facies metamorphism in the Menderes Massif (W. Tur- Embry, A.F., and Beauchamp, B., 2008, Sverdrup Basin, in Houseknecht, D.W., and Bird, K.J., 2011, Geology and petro- key): Implications on the final assembly of Gond­wana: Miall, A.D., ed., The Sedimentary Basins of the United leum potential of the rifted margins of the Canada Basin: Gondwana Research, p. 1–16, doi:10​ .1016​ /j​ .gr​ .2015​ .02​ .015.​ States and Canada: Amsterdam, Elsevier, p. 451–471. Geological Society, London, Memoir, v. 35, p. 509–526. Cecile, M.P., Harrison, J.C., Kos’ko, M.K., and Parrish, R.R., Embry, A.F., and Suneby, L.B., 1994, The Triassic–Jurassic Iles, K.A., Hergt, J.M., Sircombe, K.N., Woodhead, J.D., Bodor- 1991, Precambrian U-Pb ages of igneous rocks, Wrangel boundary in the Sverdrup Basin: Arctic Canada: Ca- kos, S., and Williams, I.S., 2015, Portrait of a reference Complex, Wrangel Island, USSR: Canadian Journal of nadian Society of Petroleum Geologists Memoir, v. 17, material: Zircon production in the Middledale Gabbroic Earth Sciences, v. 28, p. 1340–1348, doi:​10.1139​ /e91​ ​-117. p. 857–868. Diorite, Australia, and its implications for the TEMORA Colpron, M., and Nelson, J.L., 2011, A Palaeozoic NW Passage Ershova, V.B., Prokopiev, A.V., Khudoley, A.K., Sobolev, N.N., standard: Chemical Geology, v. 402, p. 140–152, doi:​10​ and the Timanian, Caledonian, and Uralian connections and Petrov, E.O., 2015, Detrital zircon ages and prov- .1016​/j​.chemgeo​.2015​.02​.036. of some exotic terranes in the North American Cordillera, enance of the Upper Paleozoic successions of Kotel’ny Joussein, E., Petit, S., Churchman, J., Theng, B., Righi, D., in Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, Island (New Siberian Islands archipelago): Lithosphere, and Delvaux, B., 2005, Halloysite clay minerals—A re- A.V., and Sørensen, K., eds., Arctic Petroleum Geology: v. 7, p. 40–45. view: Clay Minerals, v. 40, no. 4, p. 383–426, doi:10​ .1180​ ​ Geological Society, London, Memoir, v. 35, p. 463–484. Estrada, S., and Henjes-Kunst, F., 2013, 40Ar- 39Ar and U-Pb /0009855054040180. Colpron, M., Nelson, J.L., and Murphy, D.C., 2006, A tec- dating of Cretaceous continental rift-related magmatism Kuzmichev, A.B., 2009, Where does the South Anyui suture tonostratigraphic framework for the pericratonic on the northeast Canadian Arctic margin: Zeitschrift der go in the New Siberian islands and Laptev Sea?: Impli- terranes of the northern Canadian Cordillera, in Col- Deutschen Gesellschaft für Geowissenschaften, v. 164, cations for the Amerasia basin origin: Tectonophysics, pron, M., and Nelson, J. L., eds., Paleozoic evolution p. 107–130, doi:​10​.1127​/1860​-1804​/2013​/0005. v. 463, no. 1, p. 86–108, doi:​10​.1016​/j​.tecto​.2008​.09​.017. and metallogeny of pericratonic terranes at the ancient Evenchick, C.A., Davis, W.J., Bedard, J.H., Hayward, N., and Kuzmichev, A.B., and Pease, V.L., 2007, Siberian trap magma- Pacific margin of North America, Canadian and Alaskan Friedman, R.M., 2015, Evidence for protracted High Arc- tism on the New Siberian Islands: constraints for Arctic cordillera: Geological Association of Canada Special tic large igneous province magmatism in the central Mesozoic plate tectonic reconstructions: Journal of the Paper, v. 45, p. 1–23. Sverdrup Basin from stratigraphy, geochronology, and Geological Society, London, v. 164, no. 5, p. 959–968, Crovisier, J.L., Eberhart, J.P., Thomassin, J.H., Juteau, T., paleodepths of saucer-shaped sills: Geological Society doi:​10​.1144​/0016​-76492006​-090. Touray, J.C., and Ehret, G., 1982, Interaction eau de of America Bulletin, v. 127, no. 9–10, p. 1366–1390, doi:​ Lane, L.S., 1997, Canada Basin, Arctic Ocean: Evidence against mer-verre basaltique à 50 C°, Formation d’un hydroxy- 10​.1130​/B31190​.1. a rotational origin: Tectonics, v. 16, p. 363–387, doi:​10​ carbonate et de produits silicatés amorphes (Al, Mg) Franke, D., 2013, Rifting, lithosphere breakup and volcanism: .1029​/97TC00432. et mal cristallisés (Al, Fe, Mg). Étude en microscopie comparison of magma-poor and volcanic rifted mar- Laverov, N.P., Lobkovsky, L.I., Kononov, M.V., Dobretsov, électronique et par spectrométrie des photoélectrons gins: Marine and Petroleum Geology, v. 43, p. 63–87, N.L., Vernikovsky, V.A., Sokolov, S.D., and Shipilov, E.V., (E.S.C.A): Academie des Sciences Paris, Comptes Ren- doi:​10​.1016​/j​.marpetgeo​.2012​.11​.003. 2013, A geodynamic model of the evolution of the Arc- dus, series II, v. 294, p. 989–994. Gehrels, G., and Pecha, M., 2014, Detrital zircon U-Pb geo- tic basin and adjacent territories in the Mesozoic and Dhuime, B., Hawkesworth, C., and Cawood, P., 2011, When chronology and Hf isotope geochemistry of Paleozoic Cenozoic and the outer limit of the Russian Continental continents formed: Science, v. 331, no. 6014, p. 154–155, and Triassic passive margin strata of western North Shelf: Geotectonics, v. 47, no. 1, p. 1–30, doi:​10​.1134​ doi:​10​.1126​/science​.1201245. America: Geosphere, v. 10, no. 1, p. 49–65, doi:10​ .1130​ ​ /S0016852113010044. Doust, H., and Sumner, H.S., 2007, Petroleum systems in rift /GES00889​.1. Lawton, T.F., and McMillan, N.J., 1999, Arc abandonment as basins a collective approach in Southeast Asian basins: Gottlieb, E.S., Meisling, K.E., Miller, E.L., and Charles, G., a cause for passive continental rifting: Comparison of Petroleum Geoscience, v. 13, p. 127–144, doi:10​ ​.1144​ 2014, Closing the Canada Basin: Detrital zircon geo- the Jurassic Mexican Borderland rift and the Cenozoic /1354​-079307​-746. chronology relationships between the North Slope of Rio Grande rift: Geology, v. 27, no. 9, p. 779–782, doi:​ Drachev, S.S., 2011, Tectonic setting, structure and petroleum Arctic Alaska and the Franklinian mobile belt of Arctic 10​.1130​/0091​-7613​(1999)027​<0779:​AAAACF>2​.3​.CO;2. geology of the Siberian Arctic offshore sedimentary Canada: Geosphere, v. 10, no. 6, p. 1366–1384, doi:10​ ​ Lawver, L.A., and Scotese, C.R., 1990, A review of tectonic basins, in Spencer, A.M., Gautier, D., Stoupakova, A., .1130​/GES01027​.1. models for the evolution of the Canada Basin, in Grantz, Embry, A., and Sørensen, K., eds., Arctic Petroleum Grantz, A., Eittreim, S., and Dinter, D.A., 1979, Geology and A., Johnson, L., and Sweeney, J.F., eds., The Arctic Geology: Geological Society, London, Memoir, v. 35, tectonic development of the continental margin north Ocean Region: Boulder, Colorado, Geological Society p. 369–394. of Alaska: Tectonophysics, v. 59, no. 1, p. 263–291, doi:​ of America, , v. L, p. 593–618. Embry, A.F., 1983a, Stratigraphic subdivision of the Heiberg 10​.1016​/0040​-1951​(79)90050​-7. Lawver, L.A., Grantz, A., and Gahagan, L.M., 2002, Plate ki- Formation, eastern and central Sverdrup Basin, Arctic Grantz, A., Hart, P.E., and Childers, V.A., 2011, Geology and tec- nematic evolution of the present Arctic region since the Islands: Geological Survey of Canada Current Research, tonic development of the Amerasia and Canada Basins, Ordovician, in Miller, E.L., Grantz, A.W., and Kemperer, v. 83-1B, p. 205–213. Arctic Ocean, in Spencer, A.M., Gautier, D., Stoupakova, S.L., eds., Geology and Tectonic Development of the Embry, A.F., 1983b, The Heiberg Group, western Sverdrup A., Embry, A., and Sørensen, K., eds., Arctic Petroleum Bering and Chukchi Shelves and Adjacent Arctic Mar- Basin, Arctic Islands: Geological Survey of Canada Cur- Geology: Geological Society, London, Memoir, v. 35, gins: Boulder, Colorado, Geological Society of America rent Research, v. 83-1B, p. 381–389. p. 771–799. Special Paper 360, p. 333–358. Embry, A.F., 1986, Stratigraphic subdivision of the Blind Fiord Guynn, J., and Gehrels, G., 2010, Comparison of detrital zir- Lawver, L.A., Gahagan, L.M., and Norton, I., 2011, Palaeo- and Bjorne Formations (Lower Triassic), Sverdrup Basin, con age distributions using the K-S test: Tucson, Univer- geographic and tectonic evolution of the Arctic re- Arctic Island: Geological Survey of Canada Current Re- sity of Arizona, 16 p., www.laserchron.org/laserchron. gion during the Palaeozoic, in Spencer, A.M., Gautier, search, v. 86-1B, p. 329–340. Hadlari, T., and Rainbird, R.H., 2011, Retro-arc extension and D., Stoupakova, A., Embry, A., and Sørensen, K., eds., Embry, A.F., 1988, Triassic sea level changes: evidence from continental rifting: A model for the Paleoproterozoic Arctic Petroleum Geology: Geological Society, London, the Canadian Arctic Archipelago, in Wilgus, C., ed., Sea Baker Lake Basin, Nunavut: Canadian Journal of Earth Memoir, v. 35, p. 61–77. Level Changes—An Integrated Approach: Society of Sciences, v. 48, no. 8, p. 1232–1258, doi:10​ .1139​ /e11​ -002.​ Lemieux, Y., Hadlari, T., and Simonetti, A., 2011, Detrital zir- Economic Paleontologists and Mineralogists Special Hadlari, T., Davis, W.J., Dewing, K., Heaman, L.M., Lemieux, con geochronology and provenance of Devono-Mis- Publication, v. 42, p. 249–259. Y., Ootes, L., Pratt, B.R., and Pyle, L.J., 2012, Two detrital sissippian strata in the northern Canadian Cordilleran Embry, A.F., 1990, Geological and geophysical evidence in zircon signatures for the Cambrian passive margin of miogeocline: Canadian Journal of Earth Sciences, v. 48, support of the hypothesis of anticlockwise rotation of northern Laurentia highlighted by new U-Pb results from no. 2, p. 515–541. northern Alaska: Marine Geology, v. 93, p. 317–329, doi:​ northern Canada: Geological Society of America Bul- Li, C.F., Zhou, Z., Hao, H., Chen, H., Wang, J., Chen, B., and Wu, 10​.1016​/0025​-3227​(90)​90090​-7. letin, v. 124, no. 7–8, p. 1155–1168, doi:10​ .1130​ /B30530​ ​.1. J., 2008, Late Mesozoic tectonic structure and evolution Embry, A.F., 1991, Mesozoic history of the Arctic Islands, in Hadlari, T., Davis, W.J., and Dewing, K., 2014, A pericratonic along the present-day northeastern South China Sea Trettin, H., ed., Geology of the Innuitian Orogen and model for the Pearya terrane as an extension of the continental margin: Journal of Asian Earth Sciences, Arctic Platform of Canada and Greenland: Geological Franklinian margin of Laurentia, Canadian Arctic: Geo- v. 31, p. 546–561, doi:​10​.1016​/j​.jseaes​.2007​.09​.004. Survey of Canada, Geology of Canada, v. 3, p. 369–434. logical Society of America Bulletin, v. 126, no. 1–2, Ludwig, K.R., 1998, On the treatment of concordant uranium- Embry, A.F., 1993, Crockerland—The northwest source area p. 182–200, doi:​10​.1130​/B30843​.1. lead ages: Geochemica et Cosmochimica Acta, v. 62, for the Sverdrup Basin, Canadian Arctic Islands: Arctic Hadlari, T., Swindles, G.T., Galloway, J.M., Bell, K.M., Sul- p. 665–676. Geology and Petroleum Potential: Norwegian Petroleum phur, K.C., Heaman, L.M., Beranek, L.P., and Fallas, Ludwig, K.R., 2003, User’s Manual for Isoplot 3.00, A Geochro- Society Special Publication, v. 2, p. 205–216, doi:​10.1016​ ​ K.M., 2015, 1.8 billion of detrital zircon recycling nological Toolkit for Microsoft Excel: Berkeley Geochro- /B978​-0​-444​-88943​-0​.50018​-6. calibrates a refractory part of Earth’s sedimentary cycle: nology Center Special Publication No. 4, 70 p.

682 www.gsapubs.org | Volume 8 | Number 6 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021 Tectonic evolution of the Triassic northern Laurentian margin | RESEARCH

Malitch, K.N., Belousova, E.A., Griffin, W.L., Badanina, I.Y., on northwestern Axel Heiberg Island, northern Sver- Tuchkova, M.I., Sokolov, S., and Kravchenko-Berezhnoy, I.R., Pearson, N.J., Presnyakov, S.L., and Tuganova, E.V., drup Basin, Arctic Canada, in Spencer, A.M., Gautier, 2009, Provenance analysis and tectonic setting of the 2010, Magmatic evolution of the ultramafic-mafic -Kh D., Stoupakova, A., Embry, A., and Sørensen, K., eds., Triassic clastic deposits in western Chukotka, northeast araelakh intrusion (Siberian Craton, Russia): Insights Arctic Petroleum Geology: Geological Society, London, Russia, in Stone, D.B., Fujita, K., Layer, P.W., Miller, E.L., from trace-element, U-Pb and Hf-isotope data on zircon: Memoir, v. 35, p. 559–566. Prokopiev, A.V., and Toro, J., eds., Geology, Geophys- Contributions to Mineralogy and Petrology, v. 159, no. 6, Patchett, P.J., Embry, A.F., Ross, G.M., Beauchamp, B., Har- ics and Tectonics of Northeastern Russia: A Tribute to p. 753–768, doi:​10​.1007​/s00410​-009​-0452​-z. rison, J.C., Mayr, U., Isachsen, C., Rosenberg, E., and Leonid Parfenov: Stephan Mueller Special Publication, Mayr, U., Beauchamp, B., Embry, A.F., and Harrison, J.C., 2002, Spence, G., 2004, Sedimentary cover of the Canadian v. 4, p. 177–200. Geology of Upper Paleozoic and Mesozoic formations of Shield through Mesozoic time reflected by Nd isotopic Tuchkova, M.I., Sokolov, S.D., Khudoley, A.K., Verzhbitsky, V.E., northern Axel Heiberg Island, parts of NTS 560D, Cape and geochemical results for the Sverdrup Basin, Arctic Hayasaka, Y., and Moiseev, A.V., 2011, Permian and Tri- Stallworthy and 560A, Bukken Fiord, Nunavut scale Canada: The Journal of Geology, v. 112, no. 1, p. 39–57, assic deposits of Siberian and Chukotka passive mar- 1:125,000: Geological Survey of Canada, Open-File 3513. doi:​10​.1086​/379691. gins: Sedimentation setting and provenances: Fairbanks, Meneley, R.A., Henao, D., and Merritt, R.K., 1975, The north- Pease, V., 2011, Eurasian orogens and Arctic tectonics: An Alaska, Proceedings of the International Conference on west margin of the Sverdrup Basin, in Yorath, C.J., overview, in Spencer, A.M., Embry, A.F., Gautier, D.L., Arctic Margins VI, p. 61–96. Parker, E.R., and Glass, D.J., eds., Canada’s continental Stoupakova, A.V., and Sørensen, K., eds., Arctic Petro- Vermeesch, P., 2012, On the visualisation of detrital age dis- margins and offshore petroleum exploration: Canadian leum Geology: Geological Society, London, Memoir, tributions: Chemical Geology, v. 312, p. 190–194, doi:10​ ​ Society of Petroleum Geologists Memoir, v. 4, p. 557–587. v. 35, p. 311–324. .1016​/j​.chemgeo​.2012​.04​.021. Miller, E.L., Toro, J., Gehrels, G., Amato, J.M., Prokopiev, A., Pease, V., Drachev, S., Stephenson, R., and Zhang, X., 2014, Vernikovsky, V.A., Neimark, L.A., Ponomarchuk, V.A., and Tuchkova, M.I., Akinin, V.V., Dumitru, T.A., Moore, T.E., Arctic lithosphere—A review: Tectonophysics, v. 628, Vernikovskaya, A.E., 1995, Geochemistry and age of and Cecile, M.P., 2006, New insights into Arctic paleoge- p. 1–25, doi:​10​.1016​/j​.tecto​.2014​.05​.033. collisional granitoids and metamorphites of the Kara ography and tectonics from U-Pb detrital zircon geochro- Pease, V., Kuzmichev, A.B., and Danukalova, M.K., 2015, The microcontinent (Northern Taimyr): Russian Geology and nology: Tectonics, v. 25, no. 3, doi:10​ .1029​ /2005TC001830.​ New Siberian Islands and evidence for the continuation Geophysics, v. 36, no. 12, p. 50–64. Miller, E.L., Gehrels, G.E., Pease, V., and Sokolov, S., 2010, of the Uralides, Arctic Russia: Journal of the Geologi- Vervoort, J.D., and Patchett, P.J., 1996, Behavior of hafnium Stratigraphy and U-Pb detrital zircon geochronology cal Society, London, v. 172, no. 1, p. 1–4, doi:​10​.1144​ and neodymium isotopes in the crust: Constraints from of Wrangel Island, Russia: Implications for Arctic pa- /jgs2014​-064. crustally derived granites: Geochimica et Cosmochi- leogeography: The American Association of Petroleum Plafker, G., and Berg, H.C., 1994, Overview of the geology mica Acta, v. 60, p. 3717–3733, doi:​10​.1016​/0016​-7037​ Geologists Bulletin, v. 94, no. 5, p. 665–692, doi:​10.1306​ ​ and tectonic evolution of Alaska, in Plafker, G., and (96)00201​-3. /10200909036. Berg, H.C., eds., The Geology of Alaska: Boulder, Colo- Vervoort, J.D., Patchett, P.J., Blichert-Toft, J., and Albarede, F., Miller, E.L., Soloviev, A.V., Prokopiev, A.V., Toro, J., Harris, D., rado, Geological Society of America, Geology of North 1999, Relationships between Lu-Hf and Sm-Nd isotopic Kuzmichev, A.B., and Gehrels, G.E., 2013, Triassic river America, v. G-1, p. 989–1021. systems in the global sedimentary system: Earth and systems and the paleo-Pacific margin of northwestern Rioux, M., Bowring, S., Kelemen, P., Gordon, S., Dudás, F., Planetary Science Letters, v. 168, p. 79–99, doi:10​ .1016​ ​ Pangea: Gondwana Research, v. 23, no. 4, p. 1631–1645, and Miller, R., 2012, Rapid crustal accretion and magma /S0012​-821X​(99)00047​-3. doi:​10​.1016​/j​.gr​.2012​.08​.015. assimilation in the Oman-UAE ophiolite: High precision Woodhead, J.D., and Hergt, J.M., 2005, A preliminary ap- Moore, T.E., O’Sullivan, P.B., Potter, C.J., and Donelick, R.A., U-Pb zircon geochronology of the gabbroic crust: Jour- praisal of seven natural zircon reference materials for 2015, Provenance and detrital zircon geochronologic nal of Geophysical Research, Solid Earth, v. 117, no. B7. in situ Hf isotope determination: Geostandards and evolution of lower Brookian foreland basin deposits Røhr, T.S., Andersen, T., Dypvik, H., and Embry, A.F., 2010, Geoanalytical Research, v. 29, no. 2, p. 183–195, doi:​10​ of the western Brooks Range, Alaska, and implications Detrital zircon characteristics of the Lower Cretaceous .1111​/j​.1751​-908X​.2005​.tb00891​.x. for early Brookian tectonism: Geosphere, v. 11, no. 1, Isachsen Formation, Sverdrup Basin: Source constraints Wu, F.Y., Yang, Y.H., Xie, L.W., Yang, J.H., and Xu, P., 2006, Hf p. 93–122, doi:​10​.1130​/GES01043​.1. from age and Hf isotope data: Canadian Journal of Earth isotopic compositions of the standard zircons and bad- Morris, N., 2013, Stratigraphy and Geochemistry of Lower Sciences, v. 47, no. 3, p. 255–271, doi:10​ .1139​ /E10​ -006.​ deleyites used in U-Pb geochronology: Chemical Geol- Permian volcanics in the Sverdrup Basin, Northwest Söderlund, U., Patchett, P.J., Vervoort, J.D., and Isachsen, ogy, v. 234, no. 1, p. 105–126, doi:​10​.1016​/j​.chemgeo​ Ellesmere Island, Nunavut [M.S. thesis]: University of C.E., 2004, The 176Lu decay constant determined by Lu- .2006​.05​.003. Calgary, 222 p. Hf and U-Pb isotope systematics of Precambrian mafic Zhang, X., Omma, J., Pease, V., and Scott, R., 2013, Prov- Nelson, J.L., Colpron, M., Piercey, S.J., Dusel-Bacon, C., Mur- intrusions: Earth and Planetary Science Letters, v. 219, enance study of late Paleozoic–Mesozoic sandstones phy, D.C., and Roots, C.F., 2006, Paleozoic tectonic and no. 3–4, p. 311–324, doi:10​ .1016​ /S0012​ -821X​ (04)00012​ -3.​ from the Taimyr Peninsula, Arctic Russia, in Schmidt, metallogenic evolution of pericratonic terranes in Yu- Sokolov, S.D., Bondarenko, G.Y., Layer, P.W., and Kravchenko- J., ed., Sedimentary Basins and Orogenic Belts: Geosci- kon, northern British Columbia and eastern Alaska, in Berezhnoy, I.R., 2009, South Anyui suture: Tectono-stra- ences, v. 3, p. 502–527. Colpron, M., and Nelson, J.L., eds., Paleozoic evolution tigraphy, deformations, and principal tectonic events: Zhang, X., Pease, V., Skogseid, J., and Wohlgemuth-Ueber- and metallogeny of pericratonic terranes at the ancient Stephan Mueller Special Publication Series, v. 4, p. 201– wasser, C., 2015, Reconstruction of tectonic events on Pacific margin of North America, Canadian and Alaskan 221, doi:​10​.5194​/smsps​-4​-201​-2009. the northern Eurasia margin of the Arctic, from U-Pb cordillera: Geological Association of Canada, Special Stern, R.A., 1997, The GSC sensitive high resolution ion mi- detrital zircon provenance investigations of late Paleo- Paper, v. 45, p. 323–360. croprobe (SHRIMP): Analytical techniques of zircon zoic to Mesozoic sandstones in southern Taimyr Pen- Nemchin, A.A., and Cawood, P.A., 2005, Discordance of the U-Th-Pb age determinations and performance evalua- insula: Geological Society of America Bulletin, v. 128, U-Pb system in detrital zircons: Implication for prov- tion: Geological Survey of Canada Radiogenic Age and no. 1–2, p. 29–46. enance studies of sedimentary rocks: Sedimentary Ge- Isotope Studies, Report 10, Current Research, p. 1–31. Zhou, X.M., and Li, W.X., 2000, Origin of Late Mesozoic igne- ology, v. 182, no. 1, p. 143–162, doi:​10​.1016​/j​.sedgeo​ Stern, R.A., and Amelin, Y., 2003, Assessment of errors in ous rocks in Southeastern China: Implications for litho- .2005​.07​.011. SIMS zircon U-Pb geochronology using a natural zircon sphere subduction and underplating of mafic magmas: Nokleberg, W.J., Parfenov, L.M., Monger, J.W.H., Norton, I.O., standard and NIST SRM 610 glass: Chemical Geology, Tectonophysics, v. 326, p. 269–287, doi:​10​.1016​/S0040​ Khanchuk, A.I., Stone, D.B., Scotese, C.R., Scholl, D.W., v. 197, p. 111–142, doi:​10​.1016​/S0009​-2541​(02)00320​-0. -1951​(00)00120​-7. and Fujita, K., 2000, Phanerozoic tectonic evolution of Thorsteinsson, R., 1974, Carboniferous and Permian stratig- the Circum-North Pacific: U.S. Geological Survey Pro- raphy of Axel Heiberg Island and western Ellesmere fessional Paper 1626, 122 p. Island, Canadian Arctic Archipelago: Geological Survey Okulitch, A.V., compiler, 1991, Geology of the Canadian Archi- of Canada Bulletin, v. 224, 115 p. MANUSCRIPT RECEIVED 26 JANUARY 2016 pelago, Northwest Territories and North Greenland: Geo- Trettin, H.P., 1988, Early Namurian (or older) alkali basalt in the REVISED MANUSCRIPT RECEIVED 8 APRIL 2016 logical Survey of Canada Map 1715A, scale 1:2,000,000. Borup Fiord Formation, northern Axel Heiberg Island, MANUSCRIPT ACCEPTED 26 APRIL 2016 Omma, J.E., Pease, V., and Scott, R.A., 2011, U-Pb SIMS zir- Arctic Canada: Geological Survey of Canada Paper, v. con geochronology of Triassic and Jurassic sandstones 88-1D, p. 21–26. Printed in the USA

LITHOSPHERE | Volume 8 | Number 6 | www.gsapubs.org 683

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/8/6/668/3051096/668.pdf by guest on 26 September 2021