1. Trang chủ
  2. » Ngoại Ngữ

New Insights into Arctic Paleogeography and Tectonics from U-Pb Detrital Zircon Geochronology

24 0 0

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

THÔNG TIN TÀI LIỆU

p.1 New Insights into Arctic Paleogeography and Tectonics from U-Pb Detrital Zircon Geochronology Elizabeth L Miller1, Jaime Toro2, George Gehrels3, Jeffrey M Amato4, Andrei Prokopiev5, Marianna Tuchkova6, Vyacheslav V Akinin7, Trevor A Dumitru1, Thomas E Moore8, Ashton Embry9 and Michael P Cecile9 Department of Geological and Environmental Sciences, Stanford Univ., Stanford, CA, USA Department of Geology and Geography, West Virginia Univ., Morgantown, WV, USA Department of Geosciences, Univ of Arizona, Tucson, AZ, USA Department of Geological Sciences, New Mexico State Univ., MSC 3AB, PO Box 30001, Las Cruces, NM, USA Diamond and Precious Metal Geology Institute, Russian Academy of Sciences, Yakutsk, Russian Federation Russian Academy of Sciences, Moscow, Russian Federation Russian Academy of Sciences, Magadan, Russian Federation U.S Geological Survey, Menlo Park, CA, USA Geological Survey of Canada, Calgary, Alberta, Canada p.2 ABSTRACT In order to test existing models for the formation of the Amerasian Basin of the Arctic,  detrital zircon suites from 12 samples of Triassic sandstone from the circum­Arctic region  were dated by laser ablation­ICPMS. Triassic successions represent the last major phase of  regional deposition in the Arctic prior to the formation of the Amerasian Basin. Detrital  zircon samples from the northern Verkhoyansk fold­and­thrust belt (Siberia), from Chukotka  and Wrangel Island (NE Russia), and from northwestern Alaska have similar age  distributions distinguished by Carboniferous and Permo­Triassic peaks. In contrast, samples  from further east in Alaska and Canada (Sadlerochit Mts. and Sverdrup Basin) are dominated  by Cambro­Ordovician, or Proterozoic zircons. The samples from Chukotka most closely  resemble those from the Verkhoyansk. The current most popular reconstruction of the Amerasian Basin involves counter-clockwise rotation of Arctic Alaska and Chukotka away from the Canadian Arctic margin Although this satisfies many stratigraphic and geophysical constraints for the Alaska portion of the reconstruction, it places Chukotka a great distance away from Siberia and it leads to considerable overlap of continental crust if Chukotka is treated as a rigid block To resolve these discrepancies we propose an alternative model for formation of the Amerasian Basin where Chukotka originates closer to Siberia, but after rotation with Alaska, moves further away from Siberia by extension and right-lateral strikeslip related to the subsequent opening of the Makarov Basin portion of the Amerasian Basin INDEX TERMS: 1165 Geochronology: Sedimentary geochronology; 8155 Tectonophysics: Plate motions (3040); 8178 Tectonics and magmatism; 9315 Geographic Location: Arctic region (0718, 4207) KEYWORDS: Detrital Zircons, Canada Basin, Amerasian Basin, Chukotka, Triassic p.3 Introduction Despite an increased interest in the exploration of the Arctic Ocean for its role in controlling the Tertiary climate history of Earth, we know remarkably little about its formation and changing paleogeography through time Sea floor spreading anomalies in the Eurasian Basin (the northern continuation of the mid-Atlantic ridge system) constrain the history of formation of the Arctic Ocean by rifting back to about 55 Ma, but the earlier history of the adjacent Amerasian Basin (Figure 1) remains an unsolved plate tectonic puzzle The wide variety of proposed solutions to this puzzle was chronicled in Lawver et al [1990] Since then, increasingly sophisticated geophysical data sets and seismic based geologic relations has lent support to the “rotational opening model”, often referred to as the “windshield wiper” model (Figures and 2) This model proposes that Early Cretaceous rifting translated a small continental plate, known as the Arctic Alaska-Chukotka microplate, southward from the Arctic margin of Canada The plate rotated about a pole located in the McKenzie Delta region, opening the Amerasian Basin by rifting while closing the Angayucham and Anyui ocean basins to the south [e.g Grantz et al., 1990] (Figure 2) Data in support of this model include correlation of the upper Paleozoic-Mesozoic stratigraphy of the North Slope of Alaska to the Sverdrup Basin of Arctic Canada [Grantz et al., 1990; Embry et al., 1990; Toro et al., 2004], and magnetic and gravity anomalies that identify a paleospreading center in part of the Amerasian Basin [Laxon and McAdoo, 1996; Brozena et al., 2002] Although the rotational model best satisfies geological and geophysical data from the Canadian and Alaskan portion of the circum-Arctic, reconstructing the Russian portion of the Arctic Alaska-Chukotka plate by this model creates a space problem Geological constraints suggest that the Arctic Alaska-Chukotka microplate extends as far west as the New Siberian Islands and includes most of the immense and poorly known East Siberian continental shelf (Figure 1) Restoring this piece of real estate against the Canadian Arctic and Barents margins p.4 with the rotation model produces significant overlap of continental crust (Figure 2) [e.g Drachev , 2004; Natal’in, 2004] Understanding the development of the Amerasian Basin is required if we are to understand the origin and evolution of the vast Russian continental shelves and basins as well as the origin of several enigmatic bathymetric features in the Arctic Ocean such as the continental Lomonosov Ridge, the Alpha-Mendeleev Ridge, and the Chukchi Borderland (Figure 1) At present, land-based geologic efforts provide a straightforward and costeffective means of addressing this problem In particular, data from NE Russia has the potential for making significant contributions to our understanding For instance, the reconstruction shown in Figure suggests Chukotka, Russia, should have depositional ties with Arctic Canada and not Siberia Here we present U-Pb dating of single grains from detrital zircon populations from Triassic sandstones of the circum-Arctic, carried out in order to better constrain the position of the different parts of the Arctic Alaska-Chukotka microplate prior to the opening of the Amerasian basin Recent advances in laser ablation-mass spectrometry and ion microprobe technology have made the dating of large populations of detrital zircon feasible and thus is a highly desired tool for provenance studies [e.g Gehrels et al., ] The utility of such studies for plate tectonic reconstructions and for our understanding of frontier regions is just beginning to be demonstrated The Arctic Alaska-Chukotka Microplate The key element in the plate tectonic evolution of the Amerasian Basin is the Arctic Alaska-Chukotka microplate Figure portrays the microplate as a single elongate continental sliver Its northern boundary (present-day coordinates) is the Arctic Alaskan and Russian outer shelf edges Its southern boundary is defined by a belt of ophiolitic rocks that p.5 includes the Angayucham terrane of the southern Brooks Range [e.g Moore et al., 1994] and the South Anyui zone of western Chukotka [e.g Sokolov et al., 2002] Arctic Alaska-Chukotka is believed to be a unified continental fragment because of : 1) similar-age Neoproterozoic volcanic and plutonic basement rocks [Moore et al.,1994; Amato et al., 2003; Kos’ko et al.,1993]; 2) An early Paleozoic carbonate succession with Siberian faunal affinities that unconformably overlies this basement [Dumoulin et al., 2002; Natalin et al., 1998; Kos’ko et al.,1993]; 3) A suite of Devonian plutons found in the Brooks Range that have counterparts on the Seward Peninsula and along the north coast of Chukotka [Moore et al., 1994; Toro et al., 2002; Kos’ko et al., 1993]; 4) Similarities in the nature and age of Middle Jurassic-Early Cretaceous structural trends including ophiolite belts developed by collision-related deformation along the southern side of the plate These shorteningrelated structures appear to extend continuously from the Brooks Range to offshore Wrangel Island and into the Chukotka fold belt (where they remain poorly dated) (Figure 1a) Important along-strike differences in the stratigraphy of the microplate include the varying nature of Triassic sedimentary successions Northern Alaska is characterized by a relatively thin, clastic passive margin succession, which includes chert and other pelagic deposits in the south and more proximal thin platform sandstones in the north In Chukotka, the Triassic consists of thick turbidite sequences intruded by gabbroic dikes and sills at their base [Gelman, 1963; Ivanov and Milov, 1975] Thus, in contrast to Alaska, Chukotka was paleogeographically linked to a major clastic source and probably experienced a Triassic rifting event that formed the deep-water basins [Tuchkova et al., 2004] Analytical Methods We separated detrital zircons from twelve samples of Triassic sandstones (Table 1) Zircons were mounted in epoxy and polished to expose the interior of the grains Isotopic p.6 analyses were performed with a Micromass Isoprobe multicollector ICPMS with a laser ablation system Laser beam diameter was ~30 m and yielded ablation pits ~20 m deep Inter-element fractionation was monitored by analyzing fragments of a large concordant zircon with a known (ID-TIMS) age of 564 ± Ma (2 σ error) This standard was analyzed once for every four unknowns Grains were selected randomly from all sizes and morphologies present, except for avoidance of grains with fractures or inclusion The ablated material is carried in argon gas into the plasma source of a Micromass Isoprobe, which is equipped with a flight tube of sufficient width that U, Th, and Pb isotopes are measured simultaneously All measurements are made in static mode, using Faraday detectors for 238U, 232 Th, 208-206Pb, and an ion-counting channel for 204Pb Ion yields are ~1 mv per ppm Each analysis consists of one 20-second integration on peaks with the laser off (for backgrounds), 20 one-second integrations with the laser firing, and a 30 second delay to purge the previous sample and prepare for the next analysis Common Pb correction is made by using the measured 204Pb and assuming an initial Pb composition from Stacey and Kramers [1975] (with uncertainties of 1.0 for 206Pb/204Pb and 0.3 for 207Pb/204Pb) Our measurement of 204Pb is unaffected by the presence of 204Hg because backgrounds are measured on peaks (thereby subtracting any background 204Hg and 204Pb), and because very little Hg is present in the argon gas Inter-element fractionation of Pb/U is generally 1.2 Ga 206 Pb/207Pb ages Age interpretations for 1.2 Ga, 206Pb/207Pb ages are generally more precise than 206Pb/238U ages, and are accordingly used in age interpretations For the older grains, you can assess discordance because the 207-206 ages are also precise However, it should be noted that 206Pb/207Pb ages underestimate the true age for discordant grains given that isotopic disturbance occurred in the geologic past To reduce the effect of discordance, only analyses that are less than 15% discordant are included in age interpretations (applicable only to older grains) The interpreted ages are presented in Table The U-Pb dates are plotted on a normalized relative-probability distribution diagrams [Ludwig, 2003] The relative height of the peaks corresponds to the significance of that population of ages (Figure 3) Some analyses were not plotted because of one of two reasons: 1) for ages 10%; 2) for ages >1000 Ma, the 207Pb/206Pb uncertainty was >10% or the discordance between the 206Pb/238U and 207Pb/206Pb ages was > 20% Insert reference to Supplementary Data Repository here p.8 Geological Setting of Samples and Interpretation of Detrital Data Triassic sandstones were selected for our study because they represent the youngest stage of widespread marine deposition prior to the onset of rifting that led to the dispersal of terranes during formation of the Amerasian Basin Rifting started in the Late Jurassic along the Alaskan margin [Grantz et al., 1990], but the major episode of sea floor spreading appears to have occurred in the Early Cretaceous [Grantz et al., 1990, but see Drachev, 2004] In addition, compressional deformation also affected parts of Arctic Alaska starting as early as the mid Jurassic when ophiolitic sequences were emplaced in Alaska and the Kolyma region [Moore et al., 1998; Oxman et al., 1995] Detrital age data from Triassic sedimentary rocks along the length of the Arctic Alaska-Chukotka microplate were collected to test the integrity and continuity of this crustal fragment and constrain its paleogeographic links to source terranes Three additional samples from the northern Verkhoyansk fold-and-thrust belt were used to characterize a “Siberian source”, and two additional samples were dated from the Sverdrup Basin to characterize a “Canadian source” (Figure 1) Nearly all of the Triassic sandstones analyzed were derived from the erosion of Paleozoic and early Mesozoic source terranes Precambrian zircons are present in all samples but only in limited amounts, with the exception of one sample from the southern Sverdrup basin (Figure 3) Our discussion below focuses on the interpretation of the Phanerozoic ages in the zircon populations because these are most easily interpreted with respect to possible source areas making them useful for paleogeographic reconstructions The samples are discussed below from west to east The shallow marine Triassic succession of the Verkhoyansk (Figure 1) is part of an immense passive-margin siliciclastic wedge that ranges in age from Carboniferous to Early Jurassic [Khudoley and Guriev, 1994] A major fluvial system connected the tectonically p.9 active Baikal Mountain region to the Verkhoyansk Siberian margin via the long-lived Vilyui graben system (Figure 1) and other distributary systems [e.g Khudoley and Guriev, 1994] Earliest Triassic strata of the northern Verkhoyansk contain basalts and tuffs that are believed to be coeval with Siberian Trap volcanism [Andrianova and Andrianov, 1970] Triassic sandstones have petrographic characteristics of “recycled orogen” sources, compatible with their inferred provenance The age distribution of detrital zircon suites from one Middle Triassic and two Upper Triassic sandstone samples are very similar and are grouped in Figure (a total of 281 ages) Post-Vendian peaks in the probability distribution occur at 288, 482, 517 and 597 Ma These ages are compatible with their inferred source in the Baikal Mountain region along the southern margin of the North Asia craton, known to have been an active Andean margin throughout the Paleozoic and early Mesozoic Major batholiths were intruded in the Carboniferous; Ordovican and Cambrian plutonic rocks are present as well [Wickam et al., 1995; Yarmolyuk et al., 1997; Bukharov et al., 1992] The Triassic of Chukotka is represented by up to km of mostly distal turbidite sequences that range upwards into shelf deposits of Upper Triassic (Norian) age [Tuchkova et al., 2004] Sandstones are mostly fine-grained with compositions varying from “continental block” to “recycled orogen” [Tuchkova et al., 2004] We analysed one Middle Triassic (Carnian) and two Upper Triassic (Norian) samples Zircon populations from these three samples are very similar and are grouped in Figure (285 grains) Zircon ages show peaks at 247, 299, 370, 440, and 494 Ma with older peaks at 575 and 650 Ma It is notable that the Paleozoic peak ages in these samples are very similar to those found in the Verkhoyansk samples The Permo-Triassic peak at 247 Ma lies within error of the age of well-dated Siberian Trap magmatism (248-253 Ma) [e.g Campbell et al., 1992; Venkatesan et al., 1997; Renne and Basu, 1996] Although Siberian Trap magmatism was dominantly basaltic, lesser silicic intrusions and tuffs have been described In the Taymir Peninsula there are 249-241 p.10 Ma high level syenite-granite stocks that have been linked to the mantle plume which also generated the Siberian Traps [Vernikovsky et al., 2003] Tuffs and tuffaceous sediments are also present in the Late Permian and earliest Triassic deposits of both the Verkhoyansk and Chukotka [Andriavova and Andrianov, 1970; Chasovitin and Shpetnyi,1964; Belik and Sosunov,1969; Sosunov and Tilíman, 1960] suggesting that magmatism related to the Siberian Traps was the most likely origin for zircons of this age Wrangel Island lies between the sample sites in Chukotka and those in western Alaska (Figure 1) The Triassic strata of Wrangel Island consist of proximal siliciclastic turbidites [Kos’ko et al., 1993] A single small Triassic sample from Wrangel Island (provided by Mike Cecile of the Geological Survey of Canada) yielded 47 datable zircons (Figure 3) Phanerozoic zircon age peaks at 253, 280, and 449 Ma are similar to the peaks seen in the samples from Chukotka (to which Wrangel Island is linked, geographically, stratigraphically and structurally) and the Verkhoyansk (which is quite distant) (Figure 3) A single grain yielded an age younger than 200 Ma, but this is not considered to be statistically significant given that concordance cannot be demonstrated (Figure 3) (Supplementary Data) The westernmost sample site in Alaska is in the Lisburne Hills fold-and-thrust belt which consists of lower Paleozoic to Mesozoic sedimentary rocks [e.g Moore et al., 2002] Two samples were collected from an unusual sandstone member of the Triassic Otuk Formation which lies stratigraphically between the shale and chert members described by Blome et al [1988] The sandstone is very fine to fine-grained, massive, and about m thick Monotis fossils found in the sandstone suggest a Late Triassic age The sandstone is lithic, consisting of about 40 percent monocrystalline quartz, 20 percent altered feldspar, and 40 percent lithic fragments (argillite, chert, siltstone, granitic fragments and carbonate) The Otuk is thought to represent condensed sedimentation in an outer shelf environment [Moore and others, 1994] Although a predominantly sedimentary provenance is inferred for this p.11 sandstone, the granitic grains suggest that it was shed from a source region that lay to the NW in the offshore Chukchi platform region [K.W Sherwood, 2000, written communication as cited in Moore and others, 1994] Because the age spectra of the two samples from the Lisburne Hills are similar, the data are grouped in Figure (187 zircons) Similarities and differences in the ages of the detrital zircons from the Lisburne Hills and NE Russia are evident A Permo-Triassic age peak at 255 Ma is present in the Lisburne Hills Triassic which is also seen in samples from Wrangel Island and Chukotka Paleozoic peaks at 321, 362 and 431 Ma are similar to the Russian samples Differences include a younger Triassic (222 Ma) peak in the Lisburne Hills sandstones Its closeness in age to the depositional age of the sandstone suggests the presence of an active magmatic source The Lisburne Hills samples also lack the strong Cambrian and Late Proterozoic peaks seen in the Chukotka samples A sample from the Early Triassic Ledge Member of the Ivishak Formation of the Sadlerochit Mountains in northeastern Alaska (Figures 1, 2) shows the greatest difference in provenance ages from the samples previously discussed Sedimentary structures and burrows indicate a shallow marine deltaic origin for the Ledge Member [Mariani, 1987] This unit is also notable because it is the main oil reservoir of the Prudhoe Bay field Zircons from the Ivishak Sandstone exhibit age peaks at 466, 531 and 564 Ma Permo-Triassic and Late Carboniferous to earliest Permian ages are conspicuously absent (Figure 3) Despite their geographic proximity, two samples from the Triassic successions of the Sverdrup Basin in Arctic Canada (Figure 1) have very different zircon age populations from each other Sample AE2 from Axel Heiberg Island on the northern flank of the Sverdrup Basin has a zircon age distribution virtually identical to that of the Sadlerochit Mountains of northeastern Alaska (Figure 3) This lends support to the rotational hypothesis for the opening of the Canada basin [Grantz et al., 1990], which would place these two samples in p.12 close proximity prior to rifting (Figure 2) In contrast, sample AE1, from the southern flank of the Sverdrup Basin on Ellesmere Island, is dominated by Proterozoic zircons with major peaks at 1.2 and 1.78 Ga, plus a few zircons of Paleozoic age (Figure 3) An in depth provenance study utilizing Sm-Nd isotopic data concluded that most of the sediments in the Sverdrup basin were sourced from recycled Franklinian-Caledonian (Early Paleozoic) sources [Patchett et al., 2004] This is consistent with the detrital ages of sample AE2 and the detrital zircon ages from the Triassic of the Sadlerochit Mountains However, the zircons dated in sample AE1 appear to represent the erosion of older crystalline rocks to the south and east of the Sverdrup Basin, possibly related to uplift of Early Proterozoic and Greenvillian age basement during the early stages of Atlantic rifting Discussion Despite the differences in the stratigraphic successions of Chukotka and Alaska, Triassic sandstones in the Lisburne Hills have similar zircon populations to those of Chukotka, Wrangel Island, and NE Russia, arguing for no major paleogeographic barriers between the Russian and westernmost Alaskan parts of the Arctic Alaska-Chukotka microplate in the Triassic The greatest difference in source regions across Arctic AlaskaChukotka is found by a comparison of zircon suites from Triassic sandstones from western versus eastern Alaska, even though this part of the microplate has well-documented structural and stratigraphic continuity [e.g Moore et al., 1994] It appears that Caledonian-Franklinian sources predominated in northeastern Alaska and northern Canada, while North Asian, or NE Russian sources shed debris onto the western part of the microplate Triassic strata of Chukotka have the greatest similarities in ages of source regions with coeval rocks of the Verkhoyansk belt of Siberia (Figure 3) However, Chukotka now lies a great distance from the northern Verkhoyansk and also from the Siberian Traps from which it may have received p.13 its Permo-Triassic zircons (Figure 1) The preferred model for opening of the Amerasian Basin (Figure 2) separates Chukotka from potential Siberian sources by a 1000-km wide ocean Furthermore, the rotational model restores Chukotka to a position adjacent to the Sverdrup Basin, whose source regions for Triassic sandstones are the most dissimilar to those of Chukotka As a means of addressing these questions, a hypothetical reconstruction of Arctic Alaska-Chukotka that places Chukotka closer to Siberia, is presented in Figure 4a The model accepts the data in support of a rotational opening model for the Canada Basin [Grantz, 1990] but restores Chukotka to a position closer to the northern Verkhoyansk, the Taimyr Peninsula, and the Siberian Traps region by invoking a series of additional rotations and internal deformation of the microplate Any attempt to maintain the integrity of the microplate from Alaska to Chukotka (as evidenced by geologic similarities) and also to connect the Chukotka part of the plate to the Siberian margin results in a space problem that is the opposite of that obtained by the rotation model—there is too much space between the microplate and the Lomonosov Ridge In our speculative model, this space problem is solved by proposing that the Alpha-Mendeleev Ridge is underlain in part by thinned, stretched continental crust that once helped fill this gap, and by proposing a younger opening of the Makarov Basin (Figures 4b and c) Closing the Makarov Basin by extension sub-parallel to the Lomonosov Ridge (as originally proposed by Taylor et al [1981] and Vogt et al [1982]) may be the main way to account for this gap To so would involve a change in the pole of rotation of the Arctic Alaska-Chukotka plate relative to North America from a position in the Mackenzie Delta region to one near the Canada/Greenland end of the Lomonosov Ridge during the opening of the Makarov Basin Rift opening of the Makarov Basin is supported by the parallel systems of horsts and grabens along its border with the Lomonosov Ridge (Figure 1) but expected parallel magnetic anomalies are difficult to distinguish in the aeromagnetic p.14 data [e.g Brozena et al., 2002] During this inferred opening, it is also hypothesized that extension continued in continental crust beneath the East Siberia Shelf, increasing the region in size, thus moving Chukotka further away from the Siberian margin This motion might have been bound by a transform-like feature that represents the western continuation of the South Anyui Zone in the subsurface, characterized by a very strong set of linear positive magnetic anomalies [Racey et al., 1996] This motion may also be the possible cause of the right-lateral sense bend of the trends of the northern Verkhoyansk (Figure 4c) The opening of the Eurasian Basin in the Eocene is not shown as a step in Figure 4, but together with the documented ~ 50% extension of continental crust beneath the Laptev Sea shelf [Drachev, 2004], Chukotka moves even further away from its initial position to its present location (compare Figures 4, 1) In addition to addressing the Siberian source region for Triassic sandstones in Chukotka, Wrangel Island and westernmost Alaska, the reconstruction shown in Fig 4a places Chukotka near the region of Siberia that was affected by a significant and widespread Triassic rifting event [Aplotov,1988] This helps explain the differences in its Triassic and Jurassic stratigraphy compared to that of the adjacent Alaska portion of the Arctic Alaska-Chukotka microplate Acknowledgements This work builds on fieldwork and sampling in Arctic Russia and Alaska funded by NSF (EAR-), Exxon-Mobil, Stanford University’s School of Earth Science McGuee funds, and the Department of Geological Sciences at Stanford Funding for the analyses presented here was provided by the Exxon Mobil Corporation Special thanks to Steve Creaney, Mike Sullivan, Bob Ferderer, Ian Norton, and Pinar Ilmaz of ExxonMobil for their interest in Arctic tectonics p.15 References Aplonov, S (1988), An aborted Triassic Ocean in West Siberia, Tectonics, 7, 6, 1103-1122 Amato, J M., E L Miller, and G E.Gehrels (2003), Lower Paleozoic through Archean detrital zircon ages from metasedimentary rocks of the Nome Group, Seward Peninsula, Alaska, Eos Trans AGU, 84(46) Fall Meet Suppl., Abstract T31F-0891 Belik, G Ya., and G M Sosunov (1969), Geologic map of USSR: North-Eastern Geological Directorate, Anyuisko- Chaunskaya series, Map R-58-XXIX, XXX, scale 1:200,000 Blome, C D., K M Reed, and I L Tailleur (1988), Radiolarian biostratigraphy of the Otuk Formation in and near the National Petroleum Reserve in Alaska, in Geology and exploration of the National Petroleum Reserve in Alaska, 1974 to 1982, edited by G Gryc, pp 725–776, U.S Geol Surv Prof Pap 1399 Brozena, J M., L A Lawver, L C Kovacs, and V A Childers (2002), Aerogeophysical evidence for the rotational opening of the Canada Basin, abstracts, AAPG Pacific Section and SPE Western Region conference, Am Assoc Pet Geol Bull., 86, 1138 Bukharov, A A., V A Khalilov, T M Strakhova, and V.V Chernikov (1992), Geology of the Baikal-Patom Upland according to new data of U-Pb dating, Russ Geol and Geoph., 33, 24-33 Campbell, I H., G K Czamanske, V A Fedorenko, R I Hill, and V Stepanov (1992), Synchronism of the Siberian traps and the Permian-Triassic boundary, Science, 258, 1760-1763 Chasovitin, M D., and A P Shpetnyi (1964), Geologic map of USSR: North-Eastern Geological Directorate, Anyuisko- Chaunskaya series, Map R-58-XXXIII, XXXIV, scale 1:200,000 Dickinson and Gehrels, 2003 ? could not find p.16 Drachev, S S (2004), Siberian Arctic Continental Margin: Constraints and Uncertainties of Plate Tectonic Models, Eos Trans AGU, 85(47) Fall Meet Suppl., Abstract GP43C07 Gelman, M L (1963), Triassic diabase association in the Aniuy fold zone (Chukotka) (in Russian), Geologiya I Geofizica, 2, 127-134 Grantz, A., S D May, and P E Hart (1990), Geology of Arctic continental margin of Alaska, in The Geology of North America, v L, The Arctic Ocean region., edited by Grantz, A., L Johnson and J F Sweeney, pp 257-288, Geol Soc Am., Boulder, CO Grantz, A., S D May, P.T Taylor, and L.A Lawver (1990), Canada Basin, in The Geology of North America, v L, The Arctic Ocean region., edited by Grantz, A., L Johnson and J.F Sweeney, pp 379-402, Geol Soc Am., Boulder, CO Ivanov, O N and A P Milov (1975), The diabase association in the Chukchi fold system and its relation to the basic magmatiism in the northern portion of the Pacific Mobile Belt (in Russian) in Magmatism of North East Asia, part 2, edited by E T Shatalov, pp 274-278 Khudoley, A K and G A Guriev (1994), The formation and development of a late Paleozoic sedimentary basin on the passive margin of the Siberian paleocontinent, in Pangea:Global Environments and Resources, edited by A F Embry, B Beauchamp, and D J Glass, Can Soc of Pet Geol., Mem 17, 131-143 Kos'ko, M K., M P Cecile, J C Harrison, V G Ganelin, N V Khandoshko, and B G Lopatin (1993), Geology of Wrangel Island, between Chukchi and East Siberian seas, northeastern Russia, Geol Surv Can Bull., Report 461, 101 pp Kuzmichev, A (2004), Where does the South Anyui Suture go to In the New Siberian Islands?: towards verification of the rotational hypothesis for the Amerasian Basin opening, unpublished poster, Geol Inst Russ Acad Sci., Moscow p.17 Lawver, L A., C R Scotese (1990), A review of tectonic models for the evolution of the Canada Basin in The geology of North America, v L., The Arctic Ocean region, edited by A Grantz, L Johnson, and J.F Sweeney, Boulder, CO, Geol Soc Am., pp 593618 Lawver, L A., A Grantz, and L M Gahagan (2002), Plate Kinematic evolution of the present Arctic region since the Ordovician, in Tectonic Evolution of the Bering ShelfChukchi Sea-Arctic Margin and Adjacent Landmasses edited by E L Miller, A Grantz, and S L Klemperer, Geol Soc Am Sp Pap 360, 333-358 Laxon, S and D McAdoo (1994), Arctic Ocean gravity field derived from ERS-1 satellite altimetry, Science, 265, 621-624 Ludwig, K R (2003), User’s manual for Isoplot 3.0: a geochronological toolkit for Microsoft Excel, Berkeley Geochronology Center, Sp Pub 4, 71 p Mariani, R K (1987), Petrography and diagenesis of the Ledge Sandstone member of the Triassic Ivishak Formation, in Petroleum geology of the northern part of the Arctic National Wildlife Refuge, northeastern Alaska, edited by K J Bird and L B Magoon, U.S Geol Surv Bull 1778, 101–115 Moore, T E., T A Dumitru, K E Adams, S N Witebsky, and A G Harris (2002), Origin of the Lisburne Hills-Herald Arch structural belt: Stratigraphic, structural and fission track evidence from the Cape Lisburne area, northwestern Alaska, in Tectonic Evolution of the Bering Shelf-Chukchi Sea-Arctic Margin and Adjacent Landmasses edited by E L Miller, A Grantz, and S L Klemperer, Geol Soc Am Sp Pap 360, 77-109 Moore, T E., W K Wallace, K J Bird, S M Karl, C G Mull, and J T Dillon (1994), Geology of northern Alaska, in The Geology of North America, v G-1, The Geology of Alaska edited by Plafker, G and H C Berg, Geol Soc Am, Boulder, CO, 49-140 p.18 Natal’in, B A (2004), Phanerozoic Tectonic Evolution of the Chukotka-Arctic Alaska Block: Problems of the Rotational Model, Eos Trans AGU, 85(47) Fall Meet Suppl., Abstract GP43C-04 Natal'in, B A., J M Amato, J Toro, and J E Wright (1999), Paleozoic rocks of northern Chukotka Peninsula, Russian Far East; implications for the tectonics of the Arctic region, Tectonics, 18, 977-1003 Oxman, V S., L M Parfenov, A V Prokopiev, V F Timofeev, F F Tretyakov, Y D Nedosekin, P W Layer, and K Fujita (1995), The Chersky Range ophiolite belt, Northeast Russia, J of Geol., 103, 539-557 Patchett, P J., A F Embry, G M Ross, B Beauchamp, J C Harrison, U Mayr, C E Isachsen, E J Rosenberg,, and G O Spence (2004), Sedimentary cover of the Canadian Shield through Mesozoic time reflected by Nd isotopic and geochemical results for the Sverdrup Basin, Arctic Canada, J of Geol., 112, 39-57 Racey, S.D., S.J McLean, W.M Davis, R.W Buhmann, and A.M Hittelman (1996), Magnetic anomaly data of the Former Soviet Union, version 1.0: NOAA, National Geophysical Data Center, Boulder, Colorado, CD-ROM Renne, P R., and A.R Basu (1991), Rapid eruption of the Siberian traps flood basalts at the Permo-Triassic boundary, Science, 253, 176-179 Sosunov, G.M., and S.M Tilíman, (1960), Geologic map of USSR, North-Eastern Geological Directorate, Anyuisko- Chaunskaya series, Map R-58-XXXV, XXXVI, scale 1:200,000 Stacey, J and J Kramers (1975), Approximation of terrestrial lead isotope evolution by a two-stage model, Earth Planet Sci Lett 26, 207– 221 p.19 Taylor, P T., L C Kovacs, P R Vogt, and G L Johnson (1981), Detailed aeromagnetic investigation of the Arctic Basin, Jour Geoph Res B, 86, 6323-6333 Toro, J., F Toro, K J Bird, C Harrison, (2004), The Arctic Alaska-Canada connection revisited, Geol Soc Am Abstr Prog., 36, 22 Tuchkova, M I., G Y Bondarenko, E L Miller, S M Katkov (2004), Triassic terrigeneous deposits of Western Chukotka: sedimentation, mineral composition, deformations (NE Russia), Eos Trans AGU, 85(47) Fall Meet Suppl., Abstract GP41A-0816 Vernikovsky, V A., V L Pease, A E Vernikovskaya, A P Romanov, D G Gee, A V Travin (2003), First report of early Triassic A-type granite and syenite intrusions from Taimyr: product of the northern Eurasian superplume? Lithos, 66, 23-36 Venkatesan, T.R., A Kumar, K Gopolan, and A I Al’mukhamedov (1997), 40Ar/39Ar age of Siberian basaltic volcanism, Chem Geol., 138, 303-310 Vogt, P R., P T Taylor, L C Kovacs, and G L Johnson (1982), The Canada Basin: aeromagnetic constraints on structure and evolution, Tectonoph., 89, 295-336 Yarmolyuk, V.V., V.I Kovalenko, A B Kotov, E B Sal'nikova (1997), The Angara-Vitim Batholith; on the problem of batholith geodynamics in the Central Asia foldbelt, Geotectonics, 31, 359-373 Figure Captions Figure Circum-Arctic map showing the location of Triassic detrital zircon and main tectonic features of the Arctic Basin Bathymetry is from IBCAO [2002] Sample locations are shown as numbered circles See Table for sample data, and Figure for zircon ages The Arctic Alaska-Chukotka microplate is bounded by a heavy dotted line The fossil spreading center for the Canada Basin is shown by a double line [Laxon and McAdoo, 1994] p.20 flanked by magnetic anomalies (positive-thin solid line, negative-thin dashed line) Abbreviations: AHI, Axel Heiberg Island, AN- Angayucham belt, AR- Alpha Ridge, BRBrooks Range, CB- Canada Basin, CH- Chukotka, CP- Chukchi Plateau, Cr- Chersky Range, EI- Ellesmere Island, LR- Lomonosov Ridge, MB- Makarov Basin, MR- Mendeleben Ridge, NSI- New Siberian Islands, SAZ- South Anyui Zone, SB- Sverdrup Basin, VG- Vilyui Graben, WHA, Wrangel-Herald Arch, WI- Wrangel Island Figure Early Cretaceous plate reconstruction of the Circum-Arctic region prior to opening of the Amerasian Basin by simple rotation of Arctic Alaska-Chukotka (AAC) modified from Lawver et al [2002] The area of overlap of continental crust that results if AAC is rotated as a rigid plate is shown in black The exposed area of the South Anyui Zone (SAZ) is highlighted by a rulled pattern Numbered circles are detrital zircon sample locations Abbreviations: AN- Angayucham belt, CR- Chersky range, KO- Kolyma-Omolon superterrane, KY- Koyukuk arc Figure Relative probability distribution diagrams for detrital zircon U-Pb ages from Triassic sandstones in the Arctic Sample locations shown in Figure Height of peak is proportional to the probability that these ages are present in the sample, with the area under the curve = On the left, spectra include all of the ages from each sample On the right, spectra include only ages

Ngày đăng: 18/10/2022, 16:05

Xem thêm:

w