Concerning tectonics and the tectonic evolution of the Arctic

Transcription

Concerning tectonics and the tectonic evolution of the Arctic
Available online at www.sciencedirect.com
ScienceDirect
Russian Geology and Geophysics 54 (2013) 838–858
www.elsevier.com/locate/rgg
Concerning tectonics and the tectonic evolution of the Arctic 1
V.A. Vernikovsky *, N.L. Dobretsov, D.V. Metelkin, N.Yu. Matushkin, I.Yu. Koulakov
A.A. Trofimuk Institute of Petroleum Geology and Geophysics, Siberian Branch of the Russian Academy of Sciences,
pr. Akademika Koptyuga 3, Novosibirsk, 630090, Russia
Novosibirsk State University, ul. Pirogova 2, Novosibirsk, 630090, Russia
Received 13 March 2013; accepted 10 April 2013
Abstract
The particularities of the current tectonic structure of the Russian part of the Arctic region are discussed with the division into the
Barents–Kara and Laptev–Chukchi continental margins. We demonstrate new geological data for the key structures of the Arctic, which are
analyzed with consideration of new geophysical data (gravitational and magnetic), including first seismic tomography models for the Arctic.
Special attention is given to the New Siberian Islands block, which includes the De Long Islands, where field work took place in 2011. Based
on the analysis of the tectonic structure of key units, of new geological and geophysical information and our paleomagnetic data for these
units, we considered a series of paleogeodynamic reconstructions for the arctic structures from Late Precambrian to Late Paleozoic. This paper
develops the ideas of L.P. Zonenshain and L.M. Natapov on the Precambrian Arctida paleocontinent. We consider its evolution during the
Late Precambrian and the entire Paleozoic and conclude that the blocks that parted in the Late Precambrian (Svalbard, Kara, New Siberian,
etc.) formed a Late Paleozoic subcontinent, Arctida II, which again “sutured” the continental masses of Laurentia, Siberia, and Baltica, this
time, within Pangea.
© 2013, V.S. Sobolev IGM, Siberian Branch of the RAS. Published by Elsevier B.V. All rights reserved.
Keywords: tectonics; geodynamics; rifting; subduction; accretionary-collisional belt; paleomagnetism; paleoreconstruction; seismic tomography; Arctic
Introduction
The youngest ocean on the planet—the Arctic Ocean is of
great interest from a scientific as well as from a practical point
of view. Investigations of Russian and foreign authors show
that knowledge of the geologic structure, formation and
evolution of this ocean, can help in solving many important
regional issues, including those concerning the formation of
oil and gas sedimentary basin. These attainments also allow
us to understand the relationship between continental margins,
including the shelves, and the various units, characterized by
a continental structure of the crust, but at the moment located
rather far from the continents. The answers to these questions
are undoubtedly necessary in determining the location of the
outer border of the continental shelf for Russia as well as other
states with an Arctic coast (Laverov et al., 2013).
The modern structure of the Arctic Ocean is determined by
the position of two of the largest basins: the Amerasian
(Canada) basin, for which the time and character of spreading
1
The article was translated by the authors.
* Corresponding author.
E-mail address: [email protected] (V.A. Vernikovsky)
processes are under discussion and the Eurasian basin with a
modern spreading center in the form of the middle-oceanic
Gakkel Ridge (Fig. 1). The formation of the former is linked
to the detachment of the Chukchi–Alaska block from the
North-American margin at the Jurassic–Cretaceous boundary,
and the latter—with the detachment of the Lomonosov Ridge
and possibly the Alpha–Mendeleev Ridge from the Barents–
Kara margin. It was these tectonic motions, as well as the
relative position of the Eurasian and North American continents that determined its structure in the last 150 m.y. (Fig. 2).
Fig. 1 shows the structure of the oceanic basins and their
framing as well as the structure of the shelf and continental
parts. On the other hand, Fig. 2 shows a tectonic map of the
Arctic depicting the main structural elements of fold belts and
continental blocks and their division by age, crust type and
basement age. This allows us to determine the Timan–Pechora,
Svalbard, Kara, Novaya Zemlya microplates, the New Siberian
island and De Long archipelagos continental block, the
Chukchi microcontinent, as well as continental terranes in the
ocean, represented by the Lomonosov, Mendeleev Ridges and
other blocks. During the determination and representation of
the entire structure of the East Arctic the following tectonic
1068-7971/$ - see front matter D 201 3, V.S. So bolev IGM, Siberian Branch of the RAS. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.rgg.201
+
3.07.006
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
839
Fig. 1. The main geomorphologic structures of the Arctic Ocean (http://www.ngdc.noaa.gov/mgg/bathymetry/arctic/currentmap.html). The framed area represents the
region of Fig. 2. 1, faults and lineaments, definite and inferred; 2, thrusts; 3, mid-oceanic ridge axis; 4, Okhotsk–Chukchi volcanic-plutonic belt.
maps were used: (Bognanov and Khain, 1996, 1998; Khain et
al., 2011).
Therefore, this paper discusses the particularities of the
modern tectonic structure of the Russian part of the Arctic
region with division into the Barents–Kara and Laptev–Chukchi continental margins; also new geophysical data (Gaina et
al., 2010, 2011; Glebovsky et al., 2013), including the first
seismic tomography models for the Arctic (Jakovlev et al.,
2012). Based on the analysis of the tectonic structure of key
units, our own paleomagnetic data for these units and new
geologic and geophysical information we consider a series of
paleogeodynamic reconstructions for the arctic structures from
the Neoproterozoic to the end of the Paleozoic. Based on these
reconstructions we consider important and complicated issues
of the tectonic structure and geodynamics of the Central and
Eastern Arctic, related to the evolution of the Arctic Ocean
840
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
841
Fig. 2. The map of the main tectonic and geomorphologic elements of the Arctic. The map was composed using publications: (Bogdanov and Khain, 1996, 1998;
Grantz et al., 2009; Khain et al., 2011; Shipilov and Vernikovsky, 2010). The frame shows the contour of Fig. 3. 1, young plates cover; 2, modern continental slope;
3, riftogenic and syn-strike-slip basins Mz–Cz; 4, fragments of Caledonian orogenic belts; 5, troughs and slopes with intermediate type crust; 6, accretionary-collisional complexes; 7, oceanic crust; 8, blocks and ridges with continental crust; 9–11, Siberian platform: 9, basement metamorphic complexes (AR–PP); 10, cover
(undeformed); 11, cover (deformed in the Mesozoic); 12–15, lithospheric plates: 12, with Grenvillian basement; 13, with Late Precambrian basement; 14, with Late
Precambrian basement that was subjected to Hercynian tectonic deformations, 15, with Late Precambrian basement that was subjected to Cimmerian tectonic
deformations; 16, basalts and bimodal volcanic associations of intraplate type (K1–N); 17, Okhotsk–Chukchi volcanic-plutonic belt (K1–2); 18, Alpha–Mendeleev
large igneous province (alkaline basalts, 120–90 Ma); 19, active spreading axis; 20, sutures; 21, thrusts and upthrows; 22, normal faults; 23, detachments;
24, strike-slips; 25, faults with unclear kinematics.
from the Late Precambrian to modern times. At the same time
the last stages of the evolution of the Arctic in the Mesozoic–
Cenozoic are considered in detail in another paper of this issue
of the journal (Koulakov et al., 2013).
Barents–Kara continental margin
The Barents–Kara margin is represented by two plates—the
Kara plate or the Kara microcontinent (Vernikovsky et al.,
1995) in the east and the Svalbard plate, or Barentsia
(Bogatsky et al., 1996) in the west, in which the Franz Josef
Land and Spitsbergen archipelagos come above the sea level
(ref. Fig. 2). The particularities of the structure of the
Barents–Kara continental margins can be characterized by the
sizes of the Spitsbergen, Franz Josef Land, Novaya Zemlya,
Severnaya Zemlya archipelagos and by comparing them with
Northern Taimyr by using drilling data of single wells and
geologic-geophysical materials on the offshore areas. Here we
will mention only some aspects, which are important for the
reconstruction of the tectonic history of the region.
According to current investigations the Svalbard plate
occupies nearly all the territory of the Barents Sea margin and
includes the territories of the Spitsbergen, Franz Josef Land
and Novaya Zemlya archipelagos (ref. Fig. 2). The basement
of the Svalbard plate is represented by Grenvillian complexes,
known on Spitsbergen and Novaya Zemlya (Korago et al.,
1992; Korago and Timofeeva, 2005). This leads to the
supposition that the formation of the structure of the Svalbard
basement took place as a result of collisional events during
the amalgamation of Rodinia. Based on paleomagnetic data
the Baltica Subcontinent is usually placed in such a way that
the Grenvillian Sveconorwegian structures serve as the northern (in modern coordinates) “ending” of the Grenvillian
structures of the western margin of Laurentia, while the
Meso-Neoproterozoic fold belts of Amazonia are linearly
oriented along the Grenvillian margin of Laurentia (Cawood
and Pisarevsky, 2006). In this context it is logical to assume
that the structures of the Svalbard orogen could occupy the
northernmost position relative to Baltica and Laurentia, prolonging the Grenvillian belt and marking the suture between
these cratons during the formation of Rodinia.
As a result of the collision of the Timan edge of Baltica
with the structures of Svalbard the Late Precambrian–Early
Paleozoic Timan–Pechora orogen was formed. Its existence is
confirmed by a deep cutting out of the Late Precambrian
complexes of the basement of the Timan–Pechora plate and
by a clearly defined unconformity in the base of the Paleozoic
formations of its cover, as well as by the emplacement of
I-type (695–515 Ma) and A-type (564–516 Ma) granitoid
plutons (Kuznetsov et al., 2005). At the same time geologic
and structural data indicate that the collision mode could have
been “oblique” (Kuznetsov, 2008).
In the far west and southwest, in close proximity to the
Caledonian thrust in Western Spitsbergen and Northern Scandinavia, the reworking of the basement was related to the
Caledonian motions. Also there are some clues to that the
basement of the Franz Josef Land and Spitsbergen region was
subjected to tectonic-thermal alteration (metamorphism?)
420 Ma during the main stage of Caledonian metamorphism
in Northern Scandinavia (Gramberg, 1988), which could be
related to the occurrence of a Devonian orogen in the north
of Svalbard plate, probably as a result of collision between
Arctida and Euro-America or Arct-Europa with America
(Laurentia) (Zonenshain et al., 1990).
Within the Kara plate the basement surface is characterized by a complex and rugged geography. The maximum
elevations of the most submerged basement zones reach values
of 12–14 km (in St. Anna trough, as well as in other basins
of the Kara plate), and the minimum ones—1–2 km. Analysis
of the geologic structure of the coastal and island framing
together with data on potential geophysical fields allows to
recognize the structures of the basement of the Severnaya
Zemlya and North Taimyr areas to a significant part of the
North Kara shelf. As a result this region can be regarded as
a single plate mainly with a Precambrian basement. The
Precambrian complexes, with which the foundation of the
Kara plate is built, are raised above sea level on the Severnaya
Zemlya archipelago and in the northern part of Taimyr
peninsula (Bogdanov et al., 1998; Kaban’kov et al., 1982;
Lorenz et al., 2008; Pogrebitsky, 1971; Proskurnin, 1995;
Vernikovsky, 1996). The main elements of the structure of the
southern part of the Kara plate are oriented in accordance with
the strike of the largest dextral strike-slip zones of its
boundary, which indicates a significant role of strike-slip
tectonics during the formation of the Taimyr–Severnaya
Zemlya orogenic belt. The time of its formation is registered
in Late Paleozoic regional metamorphism and granitoid magmatism (306–260 Ma) (Vernikovsky et al., 1995). Here Late
Carboniferous syncollisional and Late Permian postcollisional
granitoids are distinctly determined. Their composition varies
from S-type granites to the intermediate S-I-type and A-type.
Petrologic and isotopic-geochemical data of the collisional
event are agreeing well with paleomagnetic data (Metelkin et
al., 2005) and underlining its kinematic characteristics. During
the Late Paleozoic collision of the Kara plate (Kara microcon-
842
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
tinent) and the Siberian craton the Late Neoproterozoic Central
Taimyr accretionary belt located between them and containing
cratonic terranes, paleoisland arcs and ophiolites fragments
was deformed (Khain et al., 1997; Pease et al., 2001; Uflyand
et al., 1991; Vernikovsky and Vernikovskaya, 2001; Vernikovsky et al., 1996, 2011).
It should be noted that in the structure of geophysical fields
(gravitational and magnetic) (Gaina et al., 2010) the Kara plate
is noticeably different from the Svalbard plate and adjacent
areas of the Siberian craton, therefore it appears to be an
independent block with characteristic individual particularities
of its inner structure. This must be accounted for when dealing
with the issue of relationships and evolution of the Svalbard
and Kara plates.
In the geography of the Barents–Kara continental margin
there are clear deep shut-ins, tectonically corresponding to
graben-trenches (or troughs), of which the largest were
Franz-Victoria, St. Anna, Voronin (ref. Figs. 2, 3). From the
west and north the elements of the platform are cut by
flexure/strike-slip belts of continental slopes of the Norwegian–Greenland and Eurasian oceanic basins, respectively. The
Precambrian basement blocks of the marginal-continental
platform are separated by multiple riftogenic troughs, which
had polycyclic developments (Aplonov et al., 1996; Shipilov
and Senin, 1988; Shipilov and Vernikovsky, 2010; Verba et
al., 2005), the largest among them is the East Barents trough.
The East Barents and St. Anna troughs form a combined
structure with a triple junction (ref. Fig. 3) and the Voronin
trough is probably a more ancient structure. Taking into
account the maximum thickness and sedimentation rate,
among the riftogenic troughs and basins of the Kara and
Svalbard plates one can determine: (1) Late Caledonian–Hercynian (sedimentation maximum in S–D1, unconformity in the
base of C2), of which the Voronin trough can be an example;
(2) with a maximum in P2–T1 (St. Anna trough and North
Barents basin) and (3) Cretaceous (triple junction center in the
intersection of St. Anna and East Barents troughs) (Shipilov,
2008; Shipilov and Vernikovsky, 2010).
Laptev–Chukchi continental margin
The eastern part of the Arctic (ref. Figs. 1, 2, and 4) is
represented by the combination of marginal and oceanic
structures. In the oceanic part the Eurasian basin with the
Gakkel Ridge and the Canada basin are identified, between
which the Lomonosov Ridge, the Alpha–Mendeleev Rise are
located and they in turn are separated by the Makarov basin.
In the continental part, in the framing of the Siberian platform,
accretionary-collisional structures are located. They are represented by a fan of ridges—from the Verkhoyansk to Chersky
and further to the east with a series of arc-shaped ridges in
Koryakiya, Chukotka, and Alaska, where they characterize the
folded structures of the Pacific border.
The seismic zones (ref. Fig. 4) of the continental part form
a fan of structures, widening from the Laptev Sea and the
mouth of Lena River towards the Okhotsk seashore and the
Kamchatka neck, where the distance across seismogenic
structures exceeds 1000 km. This structure is interpreted in a
series of papers as a “diffuse” margin of the North American
and Eurasian plates (Bogdanov, 1998; Stein and Sella, 2002).
However a collage of microplates has long since been
identified here, including the Amur, Okhotsk Sea, and Bering
Sea (Korago et al., 2010; Zonenshain et al., 1990). On the
opposite side of the East Arctic sector, in Alaska and the
adjacent part of the Rocky Mountains we also observe a
fan-like complex of seismogenic structures. They are connected through the seismogenic structure of the Aleutian island
arc and the Chukchi–Koryakian seismic belt (ref. Fig. 4). This
seismicity was in a simplified way interpreted as the interaction between the North American and Eurasian plates, which
generates the opening of the North Atlantic which continues
the Central Arctic through the Gakkel Ridge and the complex
interaction between the Siberian and Canada platforms and
the Pacific plate, which leads to the formation of the
microplates collage.
We will consider this interaction with taking into account
a new seismic tomography model (Jakovlev et al., 2012) and
new data on Cenozoic volcanism in the East Arctic sector.
Fig. 4 shows a scheme of current geodynamics of the East
Arctic, on which we demonstrate the seismic zones close to
those described in (Avetisov, 2009; Poselov et al., 2012;
Jakovlev et al., 2012), as well as the South Anyui suture and
its continuation in South Chukotka (Miller and Verzhbitsky,
2009; Sokolov et al., 2001, 2009, 2010), the Okhotsk–Chukchi
volcanic belt (Late Albian–Campanian (Akinin, 2012)), the
Kolyma–Omolon superterrane (microcontinent). The fan of
seismogenic faults spreading from the mouth of the Lena River
is located between the Verkhoyansk Range and the southwestern margin of the Kolyma–Omolon massif (Imaev et al.,
1998). The scheme shows a projection of a cold high-velocity
lens in the transitional zone of the upper mantle at depth
640 km (ST2) that was identified on the horizontal seismic
tomography section on Fig. 5. This margin is located to the
south of the Chukchi prolongation of the South Anyui suture
and represents a relict of the South Anyui oceanic plate that
subducted in the Early Cretaceous. Fig. 4 also shows the
margin of the areas of plume alkaline and moderate alkaline
volcanism on Chukotka and Alaska (the Anyui group, Enmelen volcanoes, Devik, Imuruk, St. Michael, Besep, etc. (ref.
Fig. 1 in (Dobretsov et al., 2013) of this issue)), close to the
supposed margin of the Bering Sea block (Bogdanov, 1998).
These areas are also well determined on the horizontal seismic
tomography section for depth 220 km (Fig. 5).
Thus the Bering Sea block is framed in the south by a slab,
subducting to the north in the Alaska island arc zone to a
depth of 350–400 km (Fig. 6) (Koulakov et al., 2011; Jakovlev
et al., 2012), and in the north—by a cold lens in the
intermediate zone of the upper mantle at a depth of 400–
650 km, a relict of the initially south-subducting slab of the
Anyui plate (Lobkovsky et al., 2011). On the surface the
younger (Late Cretaceous) subduction zone can be manifested
in the Chukchi segment of the Okhotsk–Chukchi volcanic
zone (ref. Fig. 4).
Fig. 3. Tectonic scheme of the Barents–Kara region and adjacent areas, according to (Bogdanov and Khain, 1996, 1998; Shipilov and Vernikovsky, 2010). 1, Precambrian metamorphic complexes of the
basement; 2, cover of the Siberian craton: a, undeformed; b, subjected to tectonic deformations in the Mesozoic; 3, lithospheric plates: a, with Grenvillian basement; b, with Late Precambrian basement;
c, subjected to Hercynian tectonic deformations; 4, Neoproterozoic Taimyr accretionary belt; 5, Hercynian and Early Cimmerian fold belts and tectonic deformations zones; 6, folded belts of Mesozoides;
7, cover of young plates; 8, troughs with suboceanic type crust; 9, continental Mesozoic rifts; 10, areas with oceanic crust; 11, middle-oceanic rift zone; 12, largest thrusts; 13, largest strike-slips and transform
faults; 14, normal faults; 15, undetermined normal faults and strike-slips; 16, edge of the continental slope.
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
843
844
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
Fig. 4. Scheme of modern-day geodynamics and Neogene–Quaternary volcanism of the Northeast of Asia and Alaska (Imaev et al., 1998; Korago et al., 2010) with
our additions. 1, active faults, strike-slips, thrusts; 2, margins of modern subduction zones (troughs); 3, spreading zone (Gakkel Ridge); 4, passive faults; 5, Cenozoic
depressions; 6, directions and speed (cm/year) of the plates and blocks drift; 7, South Anyui suture; 8, Okhotsk–Chukchi volcanic belt; 9, Omolon superterrane
(paleomicrocontinent); 10, projections of the earthquakes centers with more than 7, 6.9–6.0 and less than 6 intensity; 11, manifestations of Neogene–Quaternary
volcanism of alkaline and moderate-alkaline composition; 12, rotation poles of plates and blocks (NA, North American plate; BS, Bering Sea plate; OS, Okhotsk Sea
plate; BR, Brooks Ridge); 13, isotherms of the heat flow, mW/m2; 14, basalts and bimodal volcanic associations of intraplate type (K1–N); 15, contours of ranges
(ST1) of plume Late Cenozoic volcanism of the: A, continental (northeastern) and B, Bering Sea provinces, from (Anikin, 2012), taken from the seismic tomographic
horizontal section (220 km), ref. Fig. 5; 16, border of the cold high-velocity lens in the transition zone in the upper mantle at depth 640 km (ST2), taken from the
seismic tomographic horizontal section (220 km), ref. Fig. 5.
Indeed, the Chukchi and Okhotsk segments demonstrate
significant differences in composition and age of igneous rocks
(Khanchuk, 2006); however this question must be studied
more. The formation of a lens in the C layer approximately
1000 km wide (ref. Fig. 5 and 6) would take at least 30 m.y.,
if compared to the South Kuril segment of the Kuril arc
(Koulakov et al., 2011).
The Chukchi–Alaska block has a key significance in the
structure of the Eastern Arctic. In a series of publications
(Grantz et al., 1998; Khain et al., 2009; Koulakov et al., 2013)
it is supposed that the Precambrian structures of the Chukchi–
Alaska block, known in the Brooks Ridge and in the East
Chukchi Rise, overbuilt the Early Precambrian structures of
the Greenland–Ellesmere block and occupied a position close
to the northeastern margin of Laurentia. The “tear” of the
Chukchi–Alaska tectonic unit from the North American craton
took place only in the Jurassic during the opening of the
Canada basin (Grantz et al., 1998; Khain et al., 2009). The
collision of this subcontinent with the Verkhoyansk–Kolyma
margin of Siberia is structurally manifested in the South Anyui
zone (ref. Figs. 2, 4), for which suture-identifying Late
Jurassic–Early Cretaceous oceanic and island arc associations
are typical (Oxman, 2003; Sokolov et al., 2010). The location
of this boundary on the shelf is a matter of discussion
(Kuzmichev, 2009). Together with (Golonka et al., 2003; Li
et al., 2008) we suppose that from the moment of the
formation of Rodinia and during the entire Neoproterozoic and
Paleozoic the named blocks did not change their position
relatively to the North American craton, forming the margin
of Laurentia. N.I. Filatova and V.E. Khain believe that a series
of geologic facts indicate a possible terrane history for the
Chukchi–Alaska unit in the Late Precambrian–Early Paleozoic
(Filatova and Khain, 2007; Khain et al., 2009). The most
prominent fact is the occurrence of abyssal facies of the Early
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
845
Fig. 5. Anomalies of seismic velocity of P-waves on horizontal sections, obtained as a result of an inversion of actual data for the Arctic region (Jakovlev et al., 2012,
with additions). ST1, contours of Late Cenozoic plume volcanism at depth 220 km, whose projections are shown on Fig. 4. ST2, contour of the cold high-velocity lens
in the transition zone of the upper mantle at depth 640 km, whose projection is shown on Fig. 4.
Paleozoic at the margin of the Greenland–Ellesmere (including
Ellesmere Isl. with the Peary ophiolitic allochton (Klemperer
et al., 2002)) block, as well as the widespread manifestation
of the structural unconformity surface, as well as deformations,
metamorphism, and granitoid magmatism, due to the Elles-
mere orogenesis. In a published model (Lawver et al., 2011)
it is supposed that before the Devonian the Chukchi–Alaska
block formed the northeastern (modern coordinates) framing
of the Siberian continent and the Ellesmere orogenesis reflects
its collision with Laurentia. The rifting that followed it led to
846
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
Fig. 6. Vertical sections of the model of seismic anomalies of P-waves velocity along three vertical sections, whose locations are shown on Fig. 5. The main
geographic elements are indicated above the traverses. CA, Chukchi anomaly, whose origin is discussed in the text.
the formation of the Angayucham oceanic basin (South Anyui
ocean) and by the end of the Devonian—in the Carboniferous
Siberia was separated from the Greenland–Alaska–Chukchi
margin of Laurentia by an already wide space (Lawver et al.,
2011).
The New Siberian block is another key structure of the
Arctic (ref. Fig. 2). It is included in the Russian Arctic shelf
and, according to many researchers, is a continental terrane
with a Precambrian basement and a Paleozoic, mostly carbonaceous cover (Drachev et al., 1998; Kos’ko, 1977; Natal’in
et al., 1999). However, many questions concerning its structure
and margins are under intense discussion.
The traditional views regard this tectonic unit that includes
the territory of the New Siberian Islands and the surrounding
Laptev Sea shelf (ref. Fig. 2) in the structure of the New
Siberian–Chukchi–Brooks fold and thrust system (Filatova and
Khain, 2007). Consequently in nearly all reconstructions it is
included in the Chukchi–Alaska terrane. However the position
of the western margin of this terrane, where the New Siberian
block is located is still discussed (Drachev, 2011; Kuzmichev,
2009). Among other things it is supposed that this all-important boundary could divide the New Siberian block in two by
passing between Kotelny and Faddey islands or between them
and New Siberia Island, or between the Anzhu and De Long
Islands. The latter is explained by the existence of anomalies
of the gravitational and magnetic fields (Gaina et al., 2010,
2011), as well as by weak seismicity (Poselov et al., 2012),
which allows the supposition of a riftogenic structure (Drachev
et al., 2010). However, as we show next, the first ever
paleomagnetic data for the New Siberian islands allows us to
say that the rocks of the Anzhu and De Long archipelagos
rocks formed within a unified New Siberian terrane at least
from the Early Ordovician, that is to say on the same basement
(Vernikovsky et al., 2013). In this case the New Siberian block
becomes comparable in size to the Kara block and by analogy
with the latter can be regarded as the New Siberian plate or
the New Siberian microcontinent. The riftogenic structure
assumed above that supposedly divides the New Siberian as
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
such and the De Long Islands can be very young, and its
formation has not yet led to the breakup of the united block
and the displacement of its parts.
The possibility of the appurtenance of the New Siberian
block to the marginal structures of Siberia or the Kara block
is discussed (Kuzmichev and Pease, 2007). The Paleozoic part
of the section, building the New Siberian Islands, is described
as a typical shelf of a passive continental margin. The most
complete section is known on Kotelny Island, where it has
mainly a carbonaceous composition. More distal facies with
the predominance of turbidites is typical for the Lower
Paleozoic complex on Bennett Island (Vernikovsky et al.,
2013). These sections have a similarity with the deposits
composing the series of carbonate blocks of the Omulevkatype along the Polousny–Kolyma suture that form the Kolyma
structural loop (Oxman, 2003). New biostratigraphic data on
the Ordovician–Silurian indirectly confirm the relative nearness of the paleobasin and the Siberian margin (Vernikovsky
et al., 2013).
Geologic and geophysical data used in geodynamic
reconstructions
The data available in the 60–80s of the last century on the
existence of Precambrian metamorphic and igneous complexes
amongst the main structures of the Arctic framing the ocean
allowed many authors to assume the existence of an ancient
continent between Laurentia, Siberia and Baltica that was
variously called the Hyperborean platform, Barentsia, Arctida
or the Arctic continent (Khain, 1979; Shatsky, 1963–1965;
Zonenshain and Natapov, 1987; etc.). This ancient continent,
which is now more and more often called Arctida thanks to
L.P. Zonenshain and L.M. Natapov, was broken up as a result
of rifting and its various continental blocks either were
overlain with continental margin sediments or were included
in fold belts on the periphery of the ocean (Fig. 7). In the
classic representation Arctida consisted of several blocks of
sialic crust, whose relicts are now located in the Arctic sector:
the Kara block, the New Siberian block (New Siberian Islands
and adjacent shelves), the North Alaska block (Brooks Ridge,
Seward Peninsula), and the Chukchi block, as well as small
fragments of the Inuit fold belt in the north of Greenland
(Peary Land, the northern part of Ellesmere and Axel Heiberg
islands) and the block of the submerged Lomonosov Ridge
(Zonenshain and Natapov, 1987). These authors did not
include the Barents–Spitsbergen block into Arctida, although
they considered the Timan belt as a suture marking the
collision of the lesser Barentsia continent with Eastern Europe
(Baltica) in the Late Precambrian.
Later a number of researchers developing these ideas
included Barentsia in the structure of Arctida (Arctic continent) in their reconstructions, noting like the authors mentioned above, the Late Precambrian (Vernikovsky, 1996) or
Vendian–Cambrian age of its collision with Baltica (Borisova
et al., 2001; Kuznetsov, 2008). At the same time T.P. Borisova
et al. named the newly constructed continent “Arct-Europe”,
847
which included Arctida and Baltica, and which later combined
with the North American continent—Laurentia with the
formation of a Caledonian fold belt and a new continent—
Arct-Laurussia or Euro-America. In the opinion of Zonenshain
and co-authors the collision of Eastern Europe with North
America took place before the Devonian, and the collision of
Euro-America and Arctida—in the Devonian (Zonenshain and
Natapov, 1987; Zonenshain et al., 1990). By the Late Carboniferous–Early Permian Euro-America including Arctida approached Siberia, which led to the closure of the Ural
paleoocean (Dobretsov, 2003; Zonenshain et al., 1990), the
formation of the Kara (Taimyr–Severnaya Zemlya) orogenic
belt (Vernikovsky, 1996) and the South Anyui oceanic bay
(Zonenshain et al., 1990).
The evolution of the continental blocks of the Arctic was
considered by many foreign authors. Among them in the 80s
of the last century based on the analysis of seismic data some
suppositions were made on the continental nature of the rises
of Lomonosov and Mendeleev (Churkin and Trexler, 1981;
Forsyth and Mair, 1984; Jackson and Johnson, 1984; etc.),
which today find more and more confirmations (Alvey et al.,
2008; Grantz et al., 2001; Langinen et al., 2009). According
to materials of complex geologic and geophysical investigations conducted in recent years on Lomonosov Ridge and in
the zone of its junction with the Laptev–Chukchi continental
margin, a distinct genetic link between it and the structures
of the shelf regions has been determined (Lebedeva-Ivanova,
2006; Poselov et al., 2012; Rekant and Gusev, 2012). It has
been shown that the tectonic structure of the ridge cannot be
considered separately from adjacent continental structural
ensembles. The ridge is an ancient continental crust block
submerged to bathial depths that evidently is a fragment of
the corresponding masses and formed during the formation of
the current structure of the Eurasian basin.
A reliable substantiation for the relative configuration of
the blocks composing Arctida in space and time could be done
with paleomagnetic data. Unfortunately, a severe lack of
quality paleomagnetic data for the Late Precambrian and
Paleozoic on the Arctic territory does not allow us to
compose a “mature” palinspastic basis for paleotectonic analysis (Vernikovsky et al., 2010). The International paleomagnetic
database (IAGA GPMDB) has no more than 30 determinations, mainly for the Late Paleozoic (younger than the
Devonian)–Early Mesozoic for the territory of Barentsia and
the Greenland–Ellesmere region. For the territory of Chukotka
and Northern Alaska—the largest fragment of the classic
Arctida—no paleomagnetic determinations exist. Reliable paleomagnetic data exist for the Kara block. They include three
paleomagnetic poles for 500, 450 and 420 Ma that allowed
justifying the apparent polar wander path (APWP) for this
plate (Metelkin et al., 2005). The character of the apparent
wander and the latitudes are similar to those of the Siberian
poles (Cocks and Torsvik, 2007) and the APWP of Baltica
(Torsvik and Cocks, 2005), but it deviates significantly to the
east. The latter is interpreted as the validation of the tectonic
unity of the Baltica and Kara blocks in the structure of the
hypothetical Arct-Europe (Gee et al., 2006; Kuznetsov, 2008),
Fig. 7. The main geologic structures of the Arctic (a) and a reconstruction for the Early Jurassic, showing the ancient arctic masses, assembled in the Arctida continent, attached to Laurasia (b) (Zonenshain
and Natapov, 1987). 1, oceanic basins deeper that 2000 m; 2, isobaths for 2000 and 3000 m; 3, active spreading center (a, definite, b, diffusive or indefinite); 4, shields; 5, platform; 6, orogenic belts;
7, ancient masses, remains of the Arctida continent; 8, Mesozoic and Cenozoic sedimentary basins; 9, folding fronts; 10, dead spreading axis; 11, continental contours; 12, blocks contours; 13, the Canada
basin shelf contour; 14, areas of blocks overlapping in the reconstruction.
848
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
although in fact it means only that the relative position of the
continental masses in the Early Paleozoic was similar to their
current spatial configuration. A detailed consideration reveals
that the angular space between the poles of the Baltica and
Kara blocks for the Cambrian–Ordovician boundary is 10º–
40°, and for the Silurian it is 20º–30°, and only the Late
Ordovician poles are not statistically different. Data synthesis
and analysis of the reconstructed kinematics of the Kara plate,
Siberia and Baltica lead to the conclusion that the Kara
continental mass had a terrane history in the Paleozoic
(Metelkin et al., 2005, 2012). From the Ordovician to the Late
Silurian the terrane experienced a drift from 40° S to 10° N
with a speed of approximately 5 cm/year and a counter
clockwise rotation of 1°/m.y. A similar drift, however with a
faster plate rotation is reconstructed for Baltica (Torsvik
and Cocks, 2005). The main difference in the kinematics of
the Siberian plate is clockwise rotation (Cocks and Torsvik,
2007). The oppositely directed rotation of the blocks should
have established the formation of transforms between them
and aided the strike-slip motion of the Kara plate from Baltica
(or Laurentia) towards Siberia. It is probable that the relicts
of these transforms serve as controlling structures in the
formation of the St. Anna trough and North Siberian step
(Fig. 2).
Thus, based on the listed geologic and geophysical data we
suppose that in the structure of Arctida (during the formation
of Rodinia) the Kara block was located between the Greenland–Ellesmere block as a continuation of its structures and
Svalbard. Its Paleozoic history corresponds to a terrane stage,
and therefore consisted of “tearing” from the Canada margin
of Laurentia and drifting towards the Taimyr margin of
Siberia. Moreover, during this drift, large transforms should
have played a determining role, including those functioning
between the Kara microcontinent and Svalbard.
Our first investigations of the Early Paleozoic of the New
Siberian Islands let us determine the coordinates of paleomagnetic poles for the Early, Middle Ordovician and the Late
Ordovician–Early Silurian and establish the following patterns
(Vernikovsky et al., 2013). First, the coincidence of paleomagnetic directions determined in sedimentary rocks of Kotelny
Isl. and Bennett Isl. tell us that the rocks of the Anzhu and
De Long archipelagos formed within one single New Siberian
terrane (microcontinent) at least since the Early Ordovician,
which means on the same basement. The South Anyui suture
we discussed earlier could not cross it even during the opening
of the Amerasian basin in the Early Cretaceous. Secondly, the
coordinates of the paleomagnetic poles we obtained differ
significantly from the corresponding interval of the apparent
polar wander paths for Siberia, Laurentia and the Kara
microcontinent.
The most favourable interpretation of the paleomagnetic
data is the version in which the New Siberian block was an
independent terrane in the Early Paleozoic. Its paleogeographic position at the Ordovician–Silurian boundary was
30°–40° of supposedly northern latitudes. Considering the
existence of mafic subvolcanic intrusions on Bel’kovsky Isl.,
which can be compared to the episode of trap magmatism in
849
Siberia (Kuzmichev and Goldyrev, 2007; Kuzmichev and
Pease, 2007), we should consider the possibility of a preMesozoic collision of the New Siberian block with the
Siberian margin. In this case the territory of the New Siberian
archipelago could have found itself in the zone of influence
of the Siberian plume when supposedly occupying a marginal
position in the structure of the Siberian trap province (Kusmichev and Goldyrev, 2007).
These paleomagnetic data for the territory of the Kara and
New Siberian microplates, together with the extensive paleomagnetic database for the Laurentia, Baltica, Siberia and
Gondwana cratons (Li et al., 2008; Metelkin et al., 2010, 2012;
McElhinny and MacFadden, 2000; Pechersky and Didenko,
1995; Pisarevsky et al., 2008; Smethurst et al., 1998; Torsvik
and Van der Voo, 2002; Torsvik et al., 1996; Wingate and
Giddings, 2000) became the basis for the paleotectonic
reconstructions we present here. The paleogeographic position
of the cratons is corrected (within the margins of error of
paleomagnetic poles) in accordance with the general model
and available global reconstructions that include the structures
of the Arctic sector (Cocks and Torsvik, 2002, 2007; Golonka
et al., 2003; Lawver et al., 2011; Scotese, 1997). The position
of the Arctida blocks for time intervals that lack paleomagnetic
data is reconstructed based on geologic considerations.
Paleotectonic reconstructions
750 Ma. This time is associated with the breakup of
Rodinia—the Grenvillian supercontinent that formed ~1 Ga.
We assume (Li et al., 2008) that the Arctida structures within
Rodinia, including Svalbard, the Kara block, the Greenland–
Ellesmere, Alaska–Chukchi and New Siberian blocks occupied
a band between the Canadian margin of Laurentia, the
southwestern margin of Siberia and the northeastern margin
of Baltica. According to current views the breakup of Rodinia
began as soon as ~950 Ma and went on for a very long time,
up to the Ediacaran. Strike-slips had an important role in this
process, by actually determining the “tectonic demolition” of
the supercontinent (Metelkin et al., 2012). By 750 Ma the
Siberian craton had already detached and strike-slip systems
dominated along the Grenvillian orogenic belt. The Paleogeographic position of the Arctic subcontinent was in subequatorial latitudes (Fig. 8). The most ancient ophiolites that
mark the paleooceanic margins at 1000–900 Ma were identified at the Taimyr (Vernikovsky et al., 2011).
650 Ma. There are many examples showing that the
breakup of the supercontinent was accompanied by the
destruction of the margins of plates into independent terranes,
microcontinents, small and average plates. By the end of the
Cryogenian, during the intense formation of oceanic crust of
the Paleo-Asian Ocean that divided Siberia and Laurentia, the
destruction of the Laurentian margin involved most of the
blocks of ancient Arctida (ref. Fig. 8). The New Siberian block
located in the northern periphery, was one of the first to
detach. In the context of the presented kinematic setting the
block’s furthermost northern position predetermined its indi-
850
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
Fig. 8. Plate tectonic reconstructions showing the position of the continents and blocks of Arctida in the Neoproterozoic–Early Paleozoic. 1, continental masses;
2, continental blocks of Arctida; 3, oceanic basins; 4, inferred position of spreading zones; 5, active continental margins; 6, general strike of the transform/strike-slip
zones with indicated strike-slip kinematics.
vidual tectonic history in the Northern Hemisphere. The Kara
and Svalbard plates, which occupied a southern position in
the structure of ancient Arctida, were detached somewhat later
as a result of rifting that spread southward and in the end
conditioned the isolation of Baltica and its sharp clockwise
rotation. Such a rotation was supported by the evolution of
the active continental margin located close to the Timan–Ural
margin of Baltica. Oblique subduction on one hand and
extension on the other caused the transform character of its
drift and interaction with the Kara microplate. In the Late
Neoproterozoic these microplates together with their large
neighbours were moved southward and occupied the subtropical latitudes of the Southern Hemisphere.
540 Ma. In the evolution of the Svalbard plate the
Precambrian–Cambrian boundary was marked by a collisional
event that resulted in the formation of the Timan–Pechora
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
orogen, which sutured the plate with the timanian margin of
Baltica (Kuznetsov, 2008; Kuznetsov et al., 2005). From this
moment this territory of Barentsia was a part of the East-European paleocontinent (Nikishin et al., 1996). The Kara block
continued to experience a mostly transform drift relatively to
Barentsia. The strike-slip mode characterized the entire northeastern (in modern coordinates) margin of the Siberian plate
and its interactions with the Laurentia and Baltica plates
(Fig. 8).
The southwestern margin of Siberia of this time was in a
setting of active continental margin of pacific type experiencing oblique subduction, which contributed to the clockwise
rotation typical for Siberia starting from the Cambrian (Metelkin, 2013; Metelkin et al., 2012). It is probable that from
this time the New Siberian block started its approach from the
plates of the Paleo-Asian Ocean towards the Siberian continental margin. The motions had a transform kinematic. The
orientation of the transforms that controlled the drift of the
New Siberian block was practically parallel to the sublatitudinal strike of the subduction system in the south of Siberia.
This not only led to the mostly strike-slip motion of the New
Siberian plate along the Siberian margin, but also caused this
block to remain in relatively high northern latitudes.
Another important event on the Precambrian–Cambrian
boundary was the birth of the Iapetus Ocean on the margins
of Laurentia and Baltica, which determined the subsequent
drift of Laurentia (Cocks and Torsvik, 2002).
500 Ma. By the Cambrian–Ordovician boundary the
Iapetus oceanic space was characterized by an active spreading
regime between Laurentia and Baltica (Cocks and Torsvik,
2002). This time was marked by processes of continental crust
breakup along the eastern (modern coordinates) margin of
Baltica with the formation of the oceanic floor of the Ural
paleoocean (Dobretsov, 2011; Puchkov, 2003). Thus the
East-European continent was nearly on the entirety of its
periphery (except the north) surrounded by young oceanic
basins, whose growth dynamic determined the active counter
clockwise rotation of the continental plate. The northern
margin of Baltica, including the margin of Barentsia, had a
transform development mode. Large scale strike-slips connected the northern margins of Baltica and Siberia, providing
a gradual drift for the Kara block towards Siberia (Metelkin
et al., 2000, 2005). The latter after an intense accretionary
event in the southwest from the side of the Paleo-Asian Ocean
continued to experience clockwise rotation, which led to
lamination of the newly formed continental crust and intense
strike-slip deformations (Metelkin, 2013; Metelkin et al.,
2012). The ongoing development of this strike-slips system in
the southeastern margin of Siberia (in modern coordinates)
connected and controlled the on-coming drift of the New
Siberian block, which now found itself in the influence zone
of the plates of the Paleo-Pacific.
450 Ma. The end of the Ordovician was marked by the
beginning of the closure of the Iapetus oceanic space (Fig. 9).
Active subduction processes manifested everywhere on its
continental margins (Cocks and Torsvik, 2005). The Barentsia
margin of Baltica and the Kara terrane on its periphery
851
experienced a significant rapprochement with the Kara margin
of Siberia mainly because of their rotation in opposite
directions. The whole system continued its general northward
drift to the equator. The position of the New Siberian block
on this reconstruction close to 30° N corresponds to paleomagnetic data (Vernikovsky et al., 2013). The New Siberian
terrane also experienced a drift conformal to that of Siberia,
which was controlled by sinistral strike-slips at the craton’s
periphery, according to models already published (Metelkin et
al., 2012).
420 Ma. The Silurian–Devonian boundary was marked by
the closure of the Iapetus Ocean and the collisional event
between Laurentia and Baltica and their unification into
Laurussia (Cocks and Torsvik, 2002). At the same time as the
formation of the Scandinavian orogen, the Caledonian orogeny
manifested itself in the northwestern part of Spitsbergen and
the northeastern margin of Greenland, continuing to spread
along the Greenland–Ellesmere area of Laurentia and reaching
Alaska in the Late Devonian–Early Carboniferous (Filatova
and Khain, 2007; Khain et al., 2009). By the end of the
Silurian the Ellesmere–Alaska margin of Laurentia was an
active margin, where relicts of the Iapetus oceanic crust were
subducted. The Kara microcontinent already was located
directly near the Taimyr margin of Siberia (ref. Fig. 9). The
early stages of its collision consisted in sliding along a
transform, which does not exclude the existence of oceanic
crust fragments between Siberia and the Kara microplate
(Metelkin et al., 2005). Apparently a thin “gap” of oceanic
crust also existed between the Svalbard plate and the Kara
microcontinent. The Ural margin of Baltica and the southwestern margin of Siberia were characterized by intense subduction
magmatism, which indicates the closure mode of the Ural and
Paleo-Asian oceanic basins (Dobretsov, 2011). The active
rapprochement of the Siberian and East-European masses was
due to the domination of subduction and strike-slip modes
within their current arctic margins (Metelkin et al., 2012). As
a result, by the end of the Devonian, most of the arctic blocks
except the New Siberian terrane found themselves near each
other again in a subequatorial position (ref. Fig. 9).
380 Ma. According to modern views (Golonka et al., 2003)
when the Devonian began the blocks of Arctida composed a
continental bridge between Siberia and Laurussia, joining the
structures of the supercontinent. In the Middle Devonian a
spreading zone and a narrow oceanic basin are assumed to
have formed in this region separating the Chukchi–Alaska
margin of Laurentia and the Lena–Anabar margin of Siberia
(Golonka et al., 2003). According to our reconstructions and
the interpolation of paleomagnetic data for the Early Paleozoic
of the Kara terrane (Metelkin et al., 2005) we assume that the
Siberian margin in the Silurian–Devonian did not have any
common boundaries with Laurussia. This peculiar bay of the
Paleo-Pacific separated these continental blocks during the
entire Late Paleozoic. They came closest to each other in the
end of the Silurian when an active continental margin could
have been evolving at the Chukchi–Alaska periphery. In the
Devonian this margin was turned into a transform (ref. Fig. 9).
The transform mode was dominant along all the continental
852
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
Fig. 9. Plate tectonic reconstructions showing the position of the continents and blocks of Arctida in the Middle Paleozoic. See legend keys on Fig. 8.
margins of Arctida in the Middle Paleozoic. Strike-slips
provided the sliding of the Siberian and Kara continental plates
systems to the right along the northwestern margin of
Laurussia, which led to the further transformation of the bay
mentioned above into a wide, marginal sea basin. This basin
opened towards the Paleo-Pacific and united the Chukchi–
Alaska, Svalbard, Kara, and Priverkhoyan shelves. We assume
that a sinistral strike-slip system existed along the latter,
causing the drift of the New Siberian block to its current
position.
This kinematic agrees well with the predominant tectonic
processes in the other oceans and their margins. Among these
the Devonian interval is associated with a subduction setting
on the Siberian margin of the Paleo-Asian ocean (Saraev et
al., 2012; Wilhem et al., 2012) and the Baltica margin of the
Ural paleoocean (Puchkov, 2000). The dominating subduction
contributed to the withdrawal of the Siberian paleocontinent
and the Arctic blocks of its margin from the Chukchi–Alaska
margin of Laurentia.
Available paleomagnetic data for 380 Ma lead us to assume
that the narrow area of the southeastern Laurussia margin was
in contact with the Amazonian margin of Gondwana (Torsvik
and Van der Voo, 2002). The oceanic space between Laurentia
and the African margin of Gondwana, known as the Rheic
Ocean (Linnemann et al., 2008; Murphy et al., 2006) was in
active closure, whereas at this time the Paleo-Tethys could
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
853
Fig. 10. Plate tectonic reconstructions showing the position of the continents and blocks of Arctida in the Late Paleozoic. See legend keys on Fig. 8.
have been growing. Thus the presented kinematic assumes a
strike-slip development mode for the margin of Laurussia and
Gondwana (Fig. 9).
355 Ma. The global tectonic regime did not undergo any
significant alterations in the Early Carboniferous. The reconstructed arctic blocks and the related continents moved a little
to the north. The New Siberian terrane drew significantly
closer to the main group of Arctida landmasses. A relatively
calm tectonic regime characterized this entire shelf of the
Paleo-Pacific, which was opening due to the withdrawal of
Siberia. It is probable that the main cause for this strike-slip
was spreading inside the Paleo-Pacific, causing the corre-
sponding direction of movement of the oceanic plates. The
closure of the Paleo-Asian and Ural oceans system, which
were located on the opposite side of the Siberian paleocontinent, aided the transform development regime for this paleoshelf. Active subduction processes manifested also in the
other paleooceans that divided the blocks of the Gondwanian
and Laurentian–Siberian continental groups. From this time a
tendency to amalgamation of continental masses into a single
continent is clearly seen (Fig. 9).
330 Ma. During the Carboniferous all continents continued
to drift northwards, getting nearer to each other (Fig. 10). The
Rheic Ocean was almost closed even at the beginning of this
854
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
period (Linnemann et al., 2008; Murphy et al., 2006). Active
subduction processes continued on the Ural margin of Baltica
(Filippova et al., 2001; Puchkov, 2000; Windley et al., 2007).
A large subduction zone appeared on the other margin of the
Ural Ocean dipping under Kazakhstan and marked by the
Valerianov volcanic belt (Filippova et al., 2001; Wilhem et
al., 2012). As a result of two-sided subduction the space of
the Ural Ocean decreased rapidly and we suppose that its
width was less than 500 km. The last basins of the Paleo-Asian
Ocean closed in a setting of oblique subduction (the Ob’–
Zaisan Ocean) (Korobkin and Buslov, 2011; Windley et al.,
2007). Thus active processes associated with the closure of
paleooceans captured the entire southeastern (in ancient
coordinates) margin of the continental masses of Laurentia,
Baltica, and Siberia. At the same time the opposite northwestern shelf kept a relatively calm tectonic mode. The arctic
blocks were located in a chain one after another along ~30°
N from west to east: Chukchi–Alaska, Svalbard, Kara, New
Siberian, and probably experienced insignificant strike-slip
motions relative to each other (Fig. 10).
305 Ma. The Carboniferous–Permian time of the Earth’s
history is characterized by the age of the main continental
collisions during the formation of Pangea (Dobretsov, 2003;
Golonka and Ford, 2000; Zonenshain et al., 1990). The
collision is noted on the Ural, where also a large scale
strike-slip system is clearly manifested (Puchkov, 2000). We
suppose that the strike-slips deformed in the strongest way the
fold and thrust structure of the Ural orogen in the Late
Permian–Triassic, however their emplacement was logically
linked to the pre-Late Paleozoic kinematic of the interacting
continental masses of Laurasia that we described earlier. The
start of the collisional stage is marked between Siberia and
Kazakhstan (Wilhem et al., 2012); however the reconstruction
shows a small space between these continents with an active
subduction zone (Fig. 10). A specific collision setting characterized the Taimyr margin of Siberia (Metelkin et al., 2005;
Vernikovsky, 1996). Here the “continent–microcontinent”
collision took place as a soft interaction of sialic masses in
conditions of oblique encounter and their rotation relative to
each other. The early stages of this collision were accompanied
by the sliding of sialic masses along a transform, which does
not exclude the existence of oceanic crust fragments between
Siberia and the Kara microplate. Geologic data indicate that
the conclusion of the collisional processes took place in the
Late Carboniferous–Permian with the formation of the Taimyr–Severnaya Zemlya fold belt (Pogrebitsky, 1971; Vernikovsky, 1996). This is confirmed by isotopic dates of three
generations of granitoids—Late Carboniferous, Early- and
Late Permian, among which the last is considered as postcollisional (Vernikovsky et al., 1995) and by paleomagnetic data
(Metelkin et al., 2005). The lithosphere, well heated by
collisional processes contributed on the Taimyr to the formation of granitoids and syenites even in the Early Triassic,
which are linked to the manifestations of the Siberian
superplume (Dobretsov and Vernikovsky, 2001; Vernikovsky
et al., 2001, 2003).
280 Ma. In the Early Permian the blocks of Laurasia and
Gondwana were already united into the Pangea supercontinent
(Fig. 10). None the less the deformations caused by the
orogenic events within the Laurasian part and now related
mostly to strike-slips went on. In accordance with paleomagnetic data the intraplate strike-slip displacements of large and
rigid tectonic units of the Eurasian continent (Siberian and
East European cratons) went on until the Late Mesozoic
(Metelkin et al., 2010, 2012).
255 Ma. By the end of the Permian as a result of a mainly
transform rapprochement of the Kara–New Siberian and
Chukchi–Svalbard continental margins the unification of the
arctic masses in the structure of a single shelf was over. Thus
the Permian–Triassic boundary can be called the time of the
second formation of Arctida, this time within Pangea. The
arctic continental masses were located in the northern margin
of Pangaea in the region of the 60th latitude, occupying
moderate and subpolar areas of the Northern Hemisphere (ref.
Fig. 10).
Conclusions
The analysis of the particularities of the tectonic structure
of key units of the Russian part of the Arctic region and our
new geologic-geophysical data presented in this paper allow
the possibility of building geodynamic reconstructions for
individual continental blocks of the Arctic region, as well as
for their ensembles in the form of paleocontinents or subcontinents. No doubt, some of our conclusions need to be
specified or verified by additional investigations. Unfortunately there are still many white spots in the field of tectonic
structure of the Arctic, and all the more so, in the issues of
the evolution of the units of the Arctic region. However, the
obtaining of new geologic-geophysical data, including paleomagnetic, for each Arctic structure lets us to better understand
its formation history, as well as the evolution of the Arctic
region as a whole.
This paper develops the ideas of L.P. Zonenshain and
L.M. Natapov on the Precambrian Arctida paleocontinent.
This Precambrian paleocontinent was in the structure of
Rodinia and located between Laurentia, Siberia and Baltica
approximately one billion years ago. The processes of Rodinia’s breakup led to the breakup of the Arctic paleocontinent
into independent microcontinents and terranes, among which
the Kara, Svalbard, New Siberian blocks must be included
among others. After considering the evolution of these blocks
during the Late Precambrian and the entire Paleozoic we
conclude that the units separated in the Late Precambrian
formed a Late Paleozoic subcontinent—Arctida-II, which
again “sutured” the continental masses of Laurentia, Siberia
and Baltica, this time within Pangea.
We demonstrate the further Meso-Cenozoic tectonic history
of the structures of the Arctic sector in (Koulakov et al., 2013;
Laverov et al., 2013; Lobkovsky et al., 2011). It is related to
the detachment of the Chukchi–Alaska fragment and Laurentia
as a whole, accretionary-collisional events on the Verkhoyansk
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
margin with the formation of the Verkhoyansk–Chukchi
fold-and-thrust region, rifting that led to the formation of the
Canada, Eurasian and other arctic basins, which finally led to
the emplacement of the current structure of the Arctic.
This work was prepared within the framework of the
project on the State contract of Rosnedra—RAS “Plotting of
plate tectonic reconstructions and stress conditions models of
the lithosphere of the Arctic region in relation to the problem
of widening the outer margin of the continental shelf of the
Russian Federation”, with support from the ONZ-1 program
of the Earth sciences division of the RAS and the RFBR grant
No. 10-0500128.
References
Akinin, V.V., 2012. Late Mesozoic and Cenozoic Magmatism and Alteration
of the Lower Crust in the Northern Framing of the Pacific. Doctorate
Dissertation Thesis. IGEM RAS, Moscow.
Alvey, A., Gaina, C., Kusznir, N.J., Torsvik, T.H., 2008. Integrated crustal
thickness mapping and plate reconstructions for the high Arctic. Earth
Planet. Sci. Lett. 274, 310–321.
Aplonov, S.V., Shmelev, G.B., Krasnov, D.K., 1996. Geodynamics of the
Barents-Kara shelf: geophysical evidence. Geotectonics (Geotektonika) 30
(4), 309–326 (58–76).
Avetisov, G.P., 2009. Seismological investigations of NIIGA-VNIIOkeangeologiya in the Arctic (history, accomplishments, prospects). Problemy
Arktiki i Antarktiki 82 (2), 27–41.
Bogatsky, V.I., Bogdanov, N.A., Kostuchenko, S.L., Senin, B.V., Sobolev, S.F., Shipilov, E.V., Khain, V.E., 1996. Explanatory note to the
Tectonic map of the Barents Sea and the Northern part of European Russia.
Scale 1:2,500,000 [in Russian]. Inst. Lithosphere RAS, Moscow.
Bogdanov, N.A., 1998. Boundary between the Eurasian and North American
Plates in the eastern Arctic regions. Geotectonics (Geotektonika) 32 (2),
70–80 (4–16).
Bogdanov, N.A., Khain, V.E. (Eds.), 1996. Tectonic Map of the Barents Sea
and the Northern Part of European Russia. Scale 1:2,500,000 [in Russian].
ISLAN, PKO “Kartografiya”, Moscow.
Bogdanov, N.A., Khain, V.E., Rosen, O.M., Shipilov, E.V., Vernikovsky, V.A., Drachev, S.S., Kostyuchenko, S.H., 1998. Explanatory note to
Tectonic Map of the Kara and Laptev Seas and North Siberia (scale
1:2,500,000) [in Russian]. Inst. Lithosphere of Marginal and Inner Seas
RAS, Moscow.
Borisova, T.P., Gertseva, M.V., Yegorov, A.Yu., Kononov, M.V., Kuznetsov,
N.B., Shipunov, S.V., Zykin, V.S., 2001. The Arct-Europe supercontinent
and its meaning for global plate tectonic reconstructions, in: Paleomagnetism and Magnetism of Rocks: Theory, Practice, Experiment [in
Russian]. Geos, Moscow, pp. 93–96.
Cawood, P.A., Pisarevsky, S.A., 2006. Was Baltica right-way-up or upsidedown in the Neoproterozoic? J. Geol. Soc. London 163, 753–759.
Cocks, L.R.M., Torsvik, T.H., 2002. Earth geography from 500 to 400 million
years ago: a faunal and paleomagnetic review. J. Geol. Soc. London 159,
631–644.
Cocks, L.R.M., Torsvik T.H., 2007. Siberia, the wandering northern terrane,
and its changing geography through the Paleozoic. Earth Sci. Rev. 82,
29–74.
Churkin, M., Trexler, J.H., 1981. Continental plates and accreted oceanic
terranes in the Arctic, in: The Ocean Basins and Margins, Vol. 5: The
Arctic Ocean. Plenum Press, New York, pp. 1–20.
Dobretsov, N.L., 2003. Evolution of structures of the Urals, Kazakhstan, Tien
Shan, and the Altai–Sayan region within the Ural-Mongolian fold belt
(Paleoasian Ocean). Geologiya i Geofizika (Russian Geology and Geophysics) 44 (1–2), 5–27 (3–26).
855
Dobretsov, N.L., 2011. Early Paleozoic tectonics and geodynamics of Central
Asia: role of mantle plumes. Russian Geology and Geophysics (Geologiya
i Geofizika) 52 (12), 1539–1552 (1957–1973).
Dobretsov, N.L., Vernikovsky, V.A., 2001. Mantle plumes and their geological manifestations. Int. Geol. Rev. 43, 771–787.
Dobretsov, N.L., Vernikovsky, V.A., Karyakin, Yu.V., Korago, E.A., Simonov, V.A., 2013. Mesozoic–Cenozoic volcanism and geodynamic
events in the Central and Eastern Arctic. Russian Geology and Geophysics
(Geologiya i Geofizika) 54 (8), 874–887 (1126–1144).
Drachev, S.S., 2011. Tectonic setting, structure and petroleum geology of the
Siberian Arctic offshore sedimentary basins, in: Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, A.V., Sørensen, K. (Eds.), Arctic
Petroleum Geology. Geol. Soc. London, Memoirs 35, pp. 369–394.
Drachev, S.S., Savostin, L.A., Groshev, V.G., Bruni, I.E., 1998. Structure and
geology of the continental shelf of the Laptev Sea, Eastern Russian Arctic.
Tectonophysics 298, 357–393.
Drachev, S.S., Malyshev, N.A., Nikishin, A.M., 2010. Tectonic history and
petroleum geology of the Russian Arctic Shelves: an overview. Petroleum
Geology Conference Series, Geological Society, London, Vol. 7, pp. 591–
619.
Filatova, N.I., Khain, V.E., 2007. Tectonics of the Eastern Arctic Region.
Geotectonics (Geotektonika) 41 (3), 171–194 (3–29), doi: 10.1134/
S0016852107030016.
Filippova, I.B., Bush, V.A., Didenko, A.N., 2001. Middle Paleozoic subduction belts: The leading factor in the formation of the Central Asian
fold-and-thrust belt. Russ. J. Earth Sci. 3 (6), 405–426.
Forsyth, D.A., Mair, J.A., 1984. Crustal structure of the Lomonosov Ridge
and the Fram and Makarov basins near the North Pole. J. Geophys. Res.
89, 473–481.
Gaina, C., Saltus, R., Aaro, S., Damaske, D., Ebbing, J., Forsberg, R.,
Glebovsky, V., Johnson, K., Jonberger, J., Koren, T., Korhonen, J.,
Litvinova, T., Oakey, G., Olesen, O., Pilkington, M., Rasmussen, T.,
Saltus, R., Schreckenberger, B., Smelror, M., Harrison, C., St-Onge, M.,
Tella, S., Paul, D., Lynds, T., Petrov, O., Strelnikov, S., Shokalsky, S.,
Lopatin, B., Wilson, F., Hults, C., Bergman, S., Solli, A., Jepsen, H.F.,
2010. Circum-Arctic Mapping Project: New Magnetic Anomaly Map
Linked to the Geology of the Arctic. Scale 1:5,000,000. Stereographic
North Pole Projection. Standard Parallel 70° N Coordinate System WGS
1984 [in Russian].
Gaina, C., Werner, S.C., Saltus, R., Maus, S., 2011. Circum-Arctic mapping
project: new magnetic and gravity anomaly maps of the Arctic, in:
Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, A.V., Sørensen, K. (Eds.), Arctic Petroleum Geology Book Series. Geol. Soc.
Memoirs 35, pp. 39–48, doi: 10.1144/M35.3.
Gee, D.G., Bogolepova, O.K., Lorenz, H., 2006. The Timanide, Caledonide
and Uralide orogens in the Eurasian high Arctic, and relationships to the
paleo-continents Laurentia, Baltica and Siberia. Geol. Soc. London,
Memoirs 32, pp. 507–520.
Glebovsky, V.Yu., Astafurova, E.G., Chernykh, A.A., Korneva, M.A.,
Kaminsky, V.D., Poselov, V.A., 2013. Thickness of the Earth’s crust in
the deep Arctic Ocean: Results of a 3D gravity modeling. Russian Geology
and Geophysics (Geologiya i Geofizika) 54 (3), 247–262 (327–344).
Golonka, J., Bocharova, N.Y., Ford, D., Edrich, M.E., Bednarczyk, J.,
Wildharber, J., 2003. Paleogeographic reconstructions and basins development of the Arctic. Mar. Pet. Geol. 20 (3–4), 211–248, doi:10.1016/
S0264-8172(03)00043-6.
Golonka, J., Ford, D., 2000. Pangean (Late Carboniferous–Middle Jurassic)
paleoenvironment and lithofacies. Palaeogeogr. Palaeoclimatol. Palaeoecol. 161 (1–2), 1–34.
Gramberg, I.S. (Ed.), 1988. Barents Shelf Plate [in Russian]. Nedra,
Leningrad.
Grantz, A., Clark, D.L., Phillips, R.L., Srivastava, S.P., Blome, C.D.,
Gray, L.B., Haga, H., Mamet, B.L., McIntyre, D.J., McNeil, D.H.,
Mickey, M.B., Mullen, M.W., Murchey, B.I., Ross, C.A., Stevens, C.H.,
Silberling, N.J., Wall, J.H., Willard, D.A., 1998. Phanerozoic stratigraphy
of Northwind Ridge, magnetic anomalies in the Canada basin, and the
geometry and timing of rifting in the Amerasia basin, Arctic Ocean. Geol.
Soc. Am. Bull. 110 (6), 801–820.
856
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
Grantz, A., Pease, V.L., Willard, D.A., Phillips, R.L., Clark, D.L., 2001.
Neogene paleoceanography of Lomonosov Ridge and a tie to the Barents
shelf Bedrock. Geol. Soc. Am. Bull. 113, 1272–1281.
Grantz, A., Scott, R.A., Drachev, S.S., Moore, T.E., 2009. Map showing the
sedimentary successions of the region (58º–64º to 90º N) that be
prospective for hydrocarbons. GIS-UDRIL,s Open Files.
Imaev, V.S., Imaeva, L.S., Koz’min, B.M., Mackey, K., Fujita, K., 1998.
Seismic and tectonic processes on the boundary of lithospheric plates of
North East Asia and Alaska. Tikhookeanskaya Geologiya 17 (2), 3–17.
Jackson, G.P., Johnson, G.L., 1984. Structure and history of the Amerasian
Basin, in: Geology of the Arctic (Proc. 27 I.G.C.; Coll. 04). Moscow,
pp. 105–118.
Jakovlev, A.V., Bushenkova, N.A., Koulakov, I.Yu., Dobretsov, N.L., 2012.
Structure of the upper mantle in the Circum-Arctic region from regional
seismic tomography. Russian Geology and Geophysics (Geologiya i
Geofizika) 53 (10), 963–971 (1261–1272).
Kaban’kov, V.Ya., Rogozov Yu.G., Makar’ev A.A., 1982. Stratigraphy of
Upper Proterozoic deposits of Bolshevik Island (Severnaya Zemlya
archipelago), in: Geology of the Severnaya Zemlya Archipelago [in
Russian]. Sevmorgeo, Leningrad, pp. 5–22.
Khain, V.E., 1979. Regional Geotectonics: Out-Alpine Asia and Australia [in
Russian]. Nedra, Moscow.
Khain, V.E., Gusev, G.S., Khain, E.V., Vernikovsky, V.A., Volobuyev, M.I.,
1997. Circum-Siberian Neoproterozoic Ophiolite Belt. Ofioliti 22 (2),
195–200.
Khain, V.E., Polyakova, I.D., Filatova, N.I., 2009. Tectonics and petroleum
potential of the East Arctic province. Russian Geology and Geophysics
(Geologiya i Geofizika) 50 (4), 334–345 (443–460).
Khain, V.E., Filatova, N.I., Polyakova, I.D., Chekhovich, V.D. (Eds.), 2011.
Tectonic Map of the Eastern Arctic Region. Scale 1:2,500,000 [in
Russian]. IPGG SB RAS, Novosibirsk.
Khanchuk, A.I. (Ed.), 2006. Geodynamics, Magmatism and Metallogeny of
the East of Russia [in Russian], in 2 vols. Dal’nauka, Vladivostok.
Klemperer, S.L., Miller, E.L., Scholl, D.W., 2002. Crustal structure of the
Bering and Chukchi shelves: deep seismic reflection profiles across the
North American continent between Alaska and Russia, in: Miller, E.L.,
Grantz, A., Klemperer, S.L. (Eds.), Tectonic Evolution of the Bering
Shelf–Chukchi Sea–Arctic Margin and Adjacent Landmasses. Geol. Soc.
Amer., Spec. Pap. 360, Boulder, USA, pp. 1–24.
Korago, E.A., Timofeeva, T.N., 2005. Magmatism of Novaya Zemlya (in the
Context of the Geologic History of the Barents–North Kara Region)
(Trans. VNIIOkeangeologiya, Vol. 209) [in Russian]. VNIIOkeangeologiya, St. Petersburg.
Korago, E.A., Kovaleva, G.N., Ilyin, V.F., Pavlov, L.G., 1992. Tectonics and
Metallogeny of the Early Cimmeride of Novaya Zemlya [in Russian].
Nedra, St. Petersburg.
Korago, E.A., Evdokimov, A.N., Stolbov, N.M., 2010. Late Mesozoic and
Cenozoic Basic Magmatism of the Northwest of Eurasia’s Continental
Margin (Trans. VNIIOkeangeologiya, Vol. 215) [in Russian]. VNIIOkeangeologiya, St. Petersburg.
Korobkin, V.V., Buslov, M.M., 2011. Tectonics and geodynamics of the
western Central Asian Fold Belt (Kazakhstan Paleozoides). Russian
Geology and Geophysics (Geologiya i Geofizika) 52 (12), 1600–1618
(2032–2055).
Kos’ko, M.K., 1977. Folded basement of shelf sedimentary basins, in:
Tectonics of the Arctic [in Russian]. NIIGA, Leningrad, pp. 55–85.
Koulakov, I.Yu., Dobretsov, N.L., Bushenkova N.A., Jakovlev, A.V., 2011.
Slab shape in subduction zones beneath the Kurile–Kamchatka and
Aleutian arcs based on regional tomography results. Russian Geology and
Geophysics (Geologiya i Geofizika) 52 (6), 650–667 (830–851).
Koulakov, I.Yu., Gaina, C., Dobretsov, N.L., Vasilevsky, A.N., Bushenkova, N.A., 2013. Plate reconstructions in the Arctic region based on joint
analysis of gravity, magnetic, and seismic anomalies. Russian Geology
and Geophysics (Geologiya i Geofizika) 54 (8), 859–873 (1108–1125).
Kuzmichev, A.B., 2009. Where does the South Anyui suture go in the New
Siberian islands and Laptev Sea?: Implications for the Amerasia basin
origin. Tectonophysics 463, 86–108.
Kuzmichev, A.B., Goldyrev, A.E., 2007. Permian–Triassic trap magmatism
in Bel’kov Island (New Siberian Islands). Russian Geology and Geophysics (Geologiya i Geofizika) 48 (2), 167–176 (216–228).
Kuzmichev, A.B., Pease, V.L., 2007. Siberian trap magmatism on the New
Siberian Islands: constraints for Arctic Mesozoic plate tectonic reconstructions. J. Geol. Soc. London, 164, 959–968.
Kuznetsov, N.B., 2008. The Cambrian Pre-Uralide–Timanide orogen: structural evidence for its collisional origin. Dokl. Earth Sci. 423A (9),
1383–1387, doi: 10.1134/S1028334X08090122.
Kuznetsov, N.B., Soboleva, A.A., Udoratina, O.V., Gertseva, M.V., 2005.
Pre-Ordovician Granitoids of the Timan-Ural Region and Evolution of the
Protouralides–Timanides [in Russian]. Geoprint, Syktyvkar.
Langinen, A.E., Lebedeva-Ivanova, N.N., Gee, D.G., Zamansky, Yu.Ya.,
2009. Correlations between the Lomonosov Ridge, Marvin Spur and
adjacent basins of the Arctic Ocean based on seismic data. Tectonophysics
472, 309–322.
Laverov, N.P., Lobkovsky, L.I., Kononov, M.V., Dobretsov, N.L., Vernikovsky, V.A., Sokolov, S.D., Shipilov, E.V., 2013. A geodynamic model of
the evolution of the Arctic basin and adjacent territories in the Mesozoic
and Cenozoic and the outer limit of the Russian Continental Shelf.
Geotectonics (Geotektonika) 47 (1), 1–30 (3–35), doi: 10.1134/S0016852
113010044.
Lawver, L.A., Ganagan, L.M., Norton, I., 2011. Paleogeographic and tectonic
evolution of the Arctic region during the Paleozoic, in: Spencer, A.M.,
Embry, A.F., Gautier, D.L., Stoupakova, A.V., Sørensen, K. (Eds.), Arctic
Petroleum Geology. Geol. Soc. London Memoirs 35, pp. 61–77.
Lebedeva-Ivanova, N.N., Zamansky, Y.Y., Langinen, A.E., Sorokin, M.Y.,
2006. Seismic profiling across the Mendeleev Ridge at 82° N: Evidence
of continental crust. Geophys. J. Int. 165, 527–544.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., Waele, B.De., Ernst,
R.E., Fitzsimons, I.C.W., Fuck, R.A., Gladkochub, D.P., Jacobs, J.,
Karlstrom, K.E., Lul, S., Natapov, L.M., Pease, V., Pisarevsky, S.A.,
Thrane, K., Vernikovsky, V., 2008. Assembly, configuration, and break-up
history of Rodinia: A synthesis. Precambrian Res. 160, 179–210.
Linnemann, U., Pereira, F., Jeffries, T.E., Drost, K., Gerdes, A., 2008. The
Cadomian Orogeny and the opening of the Rheic Ocean: The diacrony of
geotectonic processes constrained by LA-ICP-MS U–Pb zircon dating
(Ossa-Morena and Saxo-Thuringian Zones, Iberian and Bohemian Massifs). Tectonophysics 461, 21–43.
Lobkovsky, L.I., Verzhbitsky, V.E., Kononov, M.V., Shreider, A.A., Garagash, I.A., Sokolov, S.D., Tuchkova, M.I., Kotelkin, V.D., Vernikovsky, V.A., 2011. A geodynamic model for the evolution of the Arctic
region in the Late Mesozoic-Cenozoic and the issues of the outer boundary
of Russia’s continental shelf. Arktika. Ecologiya i Ekonomika 1 (1),
104–115.
Lorenz, H., Männik, P., Gee, D., Proskurnin, V., 2008. Geology of the
Severnaya Zemlya Archipelago and the North Kara Terrane in the Russian
high Arctic. Int. J. Earth Sci. (Geol. Rundsch.) 97 (3), 519–547, doi:
10.1007/s00531-007-0182-2.
McElhinny, M.W., MacFadden, P.L., 2000. Paleomagnetism: Continents and
Oceans. Academic Press, San Diego.
Metelkin, D.V., 2013. Kinematic reconstruction of the Early Caledonian
accretion in the southwest of the Siberian paleocontinent based on
paleomagnetic results. Russian Geology and Geophysics (Geologiya i
Geofizika) 54 (4), 381–398 (500–522).
Metelkin, D.V., Kazansky, A.Yu., Vernikovsky, V.A., Gee D., Torsvik T.,
2000. First paleomagnetic data on the Early Paleozoic rocks from the
Severnaya Zemlya archipelago and their geodynamic interpretation.
Geologiya i Geofizika (Russian Geology and Geophysics) 41 (12),
1816–1820 (1767–1772).
Metelkin, D.V., Vernikovsky, V.A., Kazansky, A.Yu., Bogolepova, O.K.,
Gubanov, A.P., 2005. Paleozoic history of the Kara microcontinent and
its relation to Siberia and Baltica: paleomagnetism, paleogeography and
tectonics. Tectonophysics 398, 225–243.
Metelkin, D.V., Vernikovsky, V.A., Kazansky, A.Yu., Wingate, M.T.D., 2010.
Late Mesozoic tectonics of Central Asia based on paleomagnetic evidence.
Gondwana Res. 18 (2–3), 400–419.
Metelkin, D.V., Vernikovsky, V.A., Kazansky, A.Yu., 2012. Tectonic evolution of the Siberian paleocontinent from the Neoproterozoic to the Late
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
Mesozoic: paleomagnetic record and reconstructions. Russian Geology and
Geophysics (Geologiya i Geofizika) 53 (7), 675–688 (883–899).
Miller, E.L., Verzhbitsky, V.E., 2009. Structural studies near Pevek, Russia:
Implications for formation of the East Siberian Shelf and Makarov Basin
of the Arctic Ocean, in: Stone, D.B., Fujita, K., Layer, P.W., Miller, E.L.,
Prokopiev, A.V., Toro, J., Geology, Geophysics and Tectonics of Northeastern Russia: A Tribute to Leonid Parfenov. EGU Stephan Mueller
Publication Series, Vol. 4, pp. 223–241.
Murphy, J.B., Gutierrez-Alonso, G., Nance, R.D., Fernandez-Suarez, J.,
Keppie, J.D., Quesada, C., Strachan, R.A., Dostal, J., 2006. Origin of the
Rheic Ocean: Rifting along a Neoproterozoic suture? Geology 34,
325–328.
Natal’in, B.A., Amato, J.M., Toro, J., Wright, J.E., 1999. Paleozoic rocks of
northern Chukotka Peninsula, Russian Far East: Implications for the
tectonics of the Arctic region. Tectonics 18, 977–1003.
Nikishin, A.M., Ziegler, P.A., Stephenson, R.A., Cloetingh, S.A.P.L.,
Furne, A.V., Fokin, P.A., Ershov, A.V., Bolotov, S.N., Korotaev, M.V.,
Alekseev, A.S., Gorbachev, V.I., Shipilov, E.V., Lankreijer, A., Bembinova, E.Yu., Shalimov, I.V., 1996. Late Precambrian to Triassic history
of the East European Craton: dynamics of sedimentary basin evolution.
Tectonophysics 268, 23–63.
Oxman, V.S., 2003. Tectonic evolution of the Mesozoic Verkhoyansk-Kolyma
belt (NE Asia). Tectonophysics 365, 45–76.
Pease, V., Gee, D.G., Vernikovsky, V., Vernikovskaya, A.E., Kireev, S., 2001.
Geochronological evidence for Late-Grenvillian magmatic and metamorphic events in Central Taimyr, Northern Siberia. Terra Nova 13, 270–280.
Pechersky, D.M., Didenko, A.N., 1995. The Paleo-Asian Ocean: Petromagnetic and Paleomagnetic Information on its Lithosphere [in Russian]. JIPE
RAS, Moscow.
Pisarevsky, S.A., Natapov, L.M., Donskaya, T.V., Gladkochub, D.P., Vernikovsky, V.A., 2008. Proterozoic Siberia: a promontory of Rodinia.
Precambrian Res. 160, 66–76.
Pogrebitsky, Yu.E., 1971. Paleotectonic Analysis of the Taimyr Fold System
[in Russian]. Nedra, Leningrad.
Poselov, V.A., Avetisov, G.P., Butsenko, V.V., Zholondz, S.M., Kaminsky, V.D., Pavlov S.P., 2012. The Lomonosov Ridge as a natural
extension of the Eurasian continental margin into the Arctic Basin.
Russian Geology and Geophysics (Geologiya i Geofizika) 53, 1276–1290
(1662–1680).
Proskurnin, V.F., 1995. A new volcanic-plutonic association in Severnaya
Zemlya and particularities of its ore potential, in: Subsoil of Taimyr [in
Russian]. Norilsk, Issue 1, pp. 93–100.
Puchkov, V.N., 2000. Paleogeodynamics of the Southern and Middle Ural [in
Russian]. Dauriya, Ufa.
Puchkov, V.N., 2003. Uralides and Timanides: their structural relationship
and position in the geologic history of the Ural-Mongolian fold belt.
Geologiya i Geofizika (Russian Geology and Geophysics) 44 (1–2), 28–39
(27–38).
Rekant, P.V., Gusev, E.A., 2012. Seismic geologic structure model for the
sedimentary cover of the Laptev Sea part of the Lomonosov Ridge and
adjacent parts of the Amundsen Plain and Podvodnikov Basin. Russian
Geology and Geophysics (Geologiya i Geofizika) 53 (11), 1150–1162
(1497–1512).
Saraev, S.V., Baturina, T.P., Bakharev, N.K., Izokh, N.G., Sennikov, N.V.,
2012. Middle–Late Devonian island-arc volcanosedimentary complexes in
northwestern Rudny Altai. Russian Geology and Geophysics (Geologiya
i Geofizika) 53 (10), 982–996 (1285–1303).
Scotese, C.R., 1997. Paleogeographic Atlas / PALEOMAP Progress Report
90-0497, Department of Geology, University of Texas at Arlington,
Arlington, TX.
Shatsky, N.S., 1963–1965. Selected Works. Vols. I–IV [in Russian]. Nauka,
Moscow.
Shipilov, E.V., 2008. Generations of spreading basins and stages of breakdown
of Wegener’s Pangea in the geodynamic evolution of the Arctic Ocean.
Geotectonics (Geotektonika) 42 (2), 105–124 (32–54), doi: 10.1134/
S0016852108020027.
Shipilov, E.V., Senin, B.V., 1988. Deep structure of the Barents Sea floor.
Geotectonika 76, 96–100.
857
Shipilov, E.V., Vernikovsky, V.A., 2010. The Svalbard–Kara plates junction:
structure and geodynamic history. Russian Geology and Geophysics
(Geologiya i Geofizika) 51 (1), 58–71 (75–92).
Smethurst, M.A., Khramov, A.N., Torsvik, T.H., 1998. The Neoproterozoic
and Paleozoic paleomagnetic data for the Siberian Platform: from Rodinia
to Pangaea. Earth Sci. Rev. 43, 1–24.
Sokolov, S.D., Bondarenko, G.E., Morozov, O.L., Ganelin, A.V., Podgorny I.I., 2001. Nappe-type tectonics of the South-Anyui suture in the
Western Chukot Peninsula. Dokl. Earth Sci. 376 (1), 7–11.
Sokolov, S.D., Bondarenko, G.Ye., Layer, P.W., Kravchenko-Berezhnoy, I.R.,
2009. South Anyui suture: tectono-stratigraphy, deformations, and principal tectonic events, in: Stone, D.B., Fujita, K., Layer, P.W., Miller, E.L.,
Prokopiev, A.V., Toro, J. (Eds.), Geology, Geophysics and Tectonics of
Northeastern Russia: A Tribute to Leonid Parfenov. EGU Stephan Mueller
Publication Series, Vol. 4, pp. 201–221.
Sokolov, S.D., Tuchkova, M.I., Bondarenko, G.E., 2010. A tectonic model of
the South Anyui suture and its role in the formation in the structures of
the East Arctic, in: Structure and Evolution History of the Lithosphere [in
Russian]. Paulsen, Moscow, pp. 204–227.
Stein, S., Sella, F., 2002. Plate boundary zones: concept and approaches, in: S.
Stein, S., Freymueller, J. (Eds.), Plate Boundary Zones. Geodynamics
Series 30. AGU, Washington, D.C., pp. 1–26.
Torsvik, T.H., Cocks, L.R.M., 2005. Norway in space and time: A Centennial
cavalcade. Norwegian J. Geol. 85, 73–86.
Torsvik, T.H., Van der Voo, R., 2002. Refining Gondwana and Pangea
paleogeography: estimates of Phanerozoic non-dipole (octupole) fields.
Geophys. J. Int. 151 (3), 771–794.
Torsvik, T.H., Smethurst, M.A., Meert, J.G., Van der Voo, R., McKerrow, W.S., Brasier, M.D., Sturt, B.A., Walderhaug, H.J., 1996. Continental
break-up and collision in the Neoproterozoic—a tale of Baltica and
Laurentia. Earth Sci. Rev. 40, 229–258.
Uflyand, A.K., Natapov, L.M., Lopatin, V.M., Chernov, D.V., 1991. On the
tectonic nature of Taimyr. Geotectonika 25 (6), 76–93.
Verba, M.L., Roslov, Yu.V., Sakulina, T.S., 2005. Novaya Zemlya in the
structure of the Ob’-Barents sedimentary megabasin from results on the
supporting profile 2-AP. Razvedka i Okhrana Nedr, No. 1, 6–9.
Vernikovsky, V.A., 1996. Geodynamic Evolution of Taimyr Folded Area [in
Russian]. Publishing House SB RAS, SPC UIGGM, Novosibirsk.
Vernikovsky, V.A., Vernikovskaya, A.E., 2001. Central Taimyr accretionary
belt (Arctic Asia): Meso-Neoproterozoic tectonic evolution and Rodinia
breakup. Precambrian Res. 110, 127–141.
Vernikovsky, V.A., Neimark, L.A., Ponomarchuk, V.A., Vernikovskaya, A.E.,
1995. Geochemistry and age of collisional granitoides and metamorphites
of the Kata microcontinent (Northern Taimyr). Geologiya i Geofizika
(Russian Geology and Geophysics) 36 (12), 50–64 (46–60).
Vernikovsky, V.A., Vernikovskaya, A.E., Chernykh, A.I., Mel’gunov M.S.,
1996. Petrology and geochemistry of Taimyr Riphean ophiolites. Geologiya i Geofizika (Russian Geology and Geophysics) 37 (1), 113–129
(103–121).
Vernikovsky, V.A., Pease, V.L., Vernikovskaya, A.E., Romanov, A.P.,
Gee, D.J., Travin, A.V., 2001. Early Triassic A-granites in the Taimyr
region as a result of a North Asian superplume. Dokl. Earth Sci. 380,
758–763.
Vernikovsky, V.A., Pease V., Vernikovskaya, A.E., Romanov, A., Gee, D.G.,
Travin, A., 2003. First report of early Triassic A-type granite and syenite
intrusions from Taimyr: product of the northern Eurasian superplume?
Lithos 66 (1–2), 23–36.
Vernikovsky, V.A., Dobretsov, N.L., Kaminsky, V.D., Lobkovsky, L.I.,
Shipilov, E.V., 2010. Geodynamics of the Central and Eastern Arctic, in:
Proceedings of the Meeting on the Coordination of the RAS on the Study
of the Arctic and the Antarctic [in Russian]. Arkhangelsk–Yekaterinburg,
pp. 41–58.
Vernikovsky, V.A., Metelkin, D.V., Vernikovskaya, A.E., Sal’nikova, E.B.,
Kovach, V.P., Kotov, A.B., 2011. The oldest island arc complex of
Taimyr: Concerning the issue of the Central-Taimyr accretionary belt
formation and paleogeodynamic reconstructions in the arctic. Dokl. Earth
Sci. 436 (2), 186–192, doi: 10.1134/S1028334X1102019X.
Vernikovsky, V.A., Metelkin, D.V., Tolmacheva, T.Yu., Malyshev, N.A.,
Petrov, O.V., Sobolev, N.N., Matushkin, N.Yu., 2013. Concerning the
858
V.A. Vernikovsky et al. / Russian Geology and Geophysics 54 (2013) 838–858
issue of paleotectonic reconstructions in the Arctic and of the tectonic
unity of the New Siberian Islands terrane: new paleomagnetic and
paleontological data. Dokl. Earth Sci. 451 (4).
Wilhem, C., Windley, B.F., Stampfli, G.M., 2012. The Altaids of Central
Asia: A tectonic and evolutionary innovative review. Earth Sci. Rev. 113,
303–341.
Windley, B.F., Alexeiev, D., Xiao, W.J., Kroner, A., Badarch, G., 2007.
Tectonic models for accretion of the Central Asian orogenic belt. J. Geol.
Soc. London 164, 31–47.
Wingate, M.T.D., Giddings, J.W., 2000. Age and paleomagnetism of the
Mundine Well dyke swarm, Western Australia implications for an
Australia–Laurentia connection at 755 Ma. Precambrian Res. 100, 335–
357.
Zonenshain, L.P., Natapov, L.M., 1987. Tectonic history of the Arctic, in:
Current Problems of the Tectonics of Oceans and Continents [in Russian].
Nauka, Moscow, pp. 31–57.
Zonenshain, L.P., Kuzmin, M.I., Natapov, L.M., 1990. Plate Tectonics of the
USSR [in Russian], in 2 vols. Nedra, Moscow, Vol. 1.