Provenance and detrital zircon geochronologic evolution of lower ...

27 downloads 1047 Views 4MB Size Report
Jan 14, 2015 - zoic (2000–1750 Ma) zircon populations. The zircon populations display fission-track evi- dence of cooling during the Brookian event.
Evolution of lower Brookian foreland basin deposits, northern Alaska

Provenance and detrital zircon geochronologic evolution of lower Brookian foreland basin deposits of the western Brooks Range, Alaska, and implications for early Brookian tectonism Thomas E. Moore1, Paul B. O’Sullivan2,*, Christopher J. Potter 3, and Raymond A. Donelick2,† 1

U.S. Geological Survey, 345 Middlefield Road, MS969, Menlo Park, California 94025-3561, USA Apatite to Zircon, Inc., 1075 Matson Road, Viola, Idaho 83872, USA 3 U.S. Geological Survey, 610 Taylor Road, Piscataway, New Jersey 08854, USA 2

ABSTRACT The Upper Jurassic and Lower Cretaceous part of the Brookian sequence of northern Alaska consists of syntectonic deposits shed from the north-directed, early Brookian orogenic belt. We employ sandstone petrography, detrital zircon U-Pb age analysis, and zircon fission-track double-dating methods to investigate these deposits in a succession of thin regional thrust sheets in the western Brooks Range and in the adjacent Colville foreland basin to determine sediment provenance, sedimentary dispersal patterns, and to reconstruct the evolution of the Brookian orogen. The oldest and structurally highest deposits are allochthonous Upper Jurassic volcanic arc–derived sandstones that rest on accreted ophiolitic and/or subduction assemblage mafic igneous rocks. These strata contain a nearly unimodal Late Jurassic zircon population and are interpreted to be a fragment of a forearc basin that was emplaced onto the Brooks Range during arc-continent collision. Synorogenic deposits found at structurally lower levels contain decreasing amounts of ophiolite and arc debris, Jurassic zircons, and increasing amounts of continentally derived sedimentary detritus accompanied by broadly distributed late Paleozoic and Triassic (359–200 Ma), early Paleozoic (542–359 Ma), and Paleoproterozoic (2000–1750 Ma) zircon populations. The zircon populations display fission-track evidence of cooling during the Brookian event and evidence of an earlier episode of cooling in the late Paleozoic and Triassic. Surprisingly, there is little evidence for erosion of the continental basement of Arctic Alaska, its Paleozoic sedimentary cover, or its hinterland metamorphic rocks in early foreland *Present address: GeoSep Services, 1521 Pine Cone Road, Moscow, Idaho 83843, USA. † Present address: Expert to Machine LLC, 1075 Matson Road, Viola, Idaho 83872, USA.

basin strata at any structural and/or stratigraphic level in the western Brooks Range. Detritus from exhumation of these sources did not arrive in the foreland basin until the middle or late Albian in the central part of the Colville Basin. These observations indicate that two primary provenance areas provided detritus to the early Brookian foreland basin of the western Brooks Range: (1) local sources in the oceanic Angayucham terrane, which forms the upper plate of the orogen, and (2) a sedimentary source region outside of northern Alaska. Pre-Jurassic zircons and continental grain types suggest the latter detritus was derived from a thick succession of Triassic turbidites in the Russian Far East that were originally shed from source areas in the Uralian-Taimyr orogen and deposited in the South Anyui Ocean, interpreted here as an early Mesozoic remnant basin. Structural thickening and northward emplacement onto the continental margin of Chukotka during the Brookian structural event are proposed to have led to development of a highland source area located in eastern Chukotka, Wrangel Island, and Herald Arch region. The abundance of detritus from this source area in most of the samples argues that the Colville Basin and ancestral foreland basins were supplied by longitudinal sediment dispersal systems that extended eastward along the Brooks Range orogen and were tectonically recycled into the active foredeep as the thrust front propagated toward the foreland. Movement of clastic sedimentary material from eastern Chukotka, Wrangel Island, and Herald Arch into Brookian foreland basins in northern Alaska confirms the interpretations of previous workers that the Brookian deformational belt extends into the Russian Far East and demonstrates that the Arctic Alaska– Chukotka microplate was a unified geologic entity by the Early Cretaceous.

INTRODUCTION Foreland basins are the receptacles for detritus shed from adjacent thrust belts and provide a stratigraphic record of their evolution (e.g., Graham et al., 1976). In addition to the welldocumented linkages between foreland basins and their adjacent thrust belts, recent work has shown that the detrital record of some basins also contains sediment derived from distant parts of the same orogenic belt and from other geologic provinces in the autochthon or distant hinterland (e.g., Lawton et al., 2003, 2009; Dickinson and Gehrels, 2008; Saylor et al., 2011; Dickinson et al., 2012). Such sediments commonly are transported long distances down the axis of foreland basins by longitudinal dispersal systems and mix with detritus derived from local source areas by short-headed and transverse dispersal systems. A full understanding of the contributions from local and distant source areas can provide a regional context of the orogenic activity as well as provide a record of deformation in the adjacent thrust belt. Investigations of the interplay between transverse and longitudinal dispersal systems in foreland basins can be approached through the study of the compositional changes in a foreland basin through time. The early history of foreland basins is commonly not well preserved, however, because sedimentary successions deposited near the thrust front get incorporated into the advancing thrust belt and are at high risk for imbrication, uplift, and erosion during later stages of orogenic activity. The parts of foreland basins most likely to be preserved in front of the orogen typically record distal up-dip deposition relative to the basin axis late in the basin’s development (e.g., DeCelles, 2012). Thus, the oldest sediments sampled in many intact foreland basins are starved or thin sedimentary sequences deposited near the end of tectonism. For this reason, study of the synorogenic deposits preserved in the thrust belt are important for unraveling the relative con-

Geosphere; February 2015; v. 11; no. 1; p. 93–122; doi:10.1130/GES01043.1; 18 figures; 6 tables; 1 supplemental file. Received 13 February 2014 ♦ Revision received 7 August 2014 ♦ Accepted 13 November 2014 ♦ Published online 14 January 2015 For permission to copy, contact [email protected] Geosphere, February 2015 © 2015 Geological Society of America

93

Moore et al. tributions from transverse and longitudinal dispersal systems in foreland basins through time. The latest Jurassic and Early Cretaceous Brooks Range fold-thrust belt of northern Alaska is a peripheral orogen with a stack of north-directed thrust sheets, each capped by synorogenic deposits. A foreland basin to the north, the Colville Basin, is filled with a 10-kmthick succession of lithic marine, shelfal, and nonmarine sedimentary rocks deposited on the autochthon. These deposits are largely younger than the synorogenic deposits associated with the thrust sheets in the Brooks Range, indicating that the Colville Basin represents only the last stage of a forward-migrating coupled deformational belt and foreland basin system that existed throughout the Early Cretaceous. The synorogenic deposits of the western Brooks Range provide a good record of the early history of sedimentation derived from the developing orogenic belt. Additional motivation for this study derives from the fact that the Brooks Range is one of the major orogenic systems bordering the Arctic Ocean basin, and its foreland basin may contain a sedimentary record of some of the events that occurred in the Arctic region before and during construction of the western Arctic Ocean basin by rifting in the Early Cretaceous. The tectonic history of this older part of the Arctic basin (Amerasia Basin) has long been controversial and difficult to resolve (e.g., Lawver et al., 2002; Grantz et al., 2011b; Pease, 2011). This study examines the early foreland basin deposits of the Brooks Range using a combined approach that unites sandstone petrography, detrital zircon U-Pb geochronology, zircon fission-track U-Pb double dating of individual zircons, and field studies to understand the origin of the depositional systems that dispersed sediment into the basin as the orogen evolved. The results have implications for the nature and tectonic significance of the upper plate of the Brooks Range orogen, for the collisional processes that were involved in the deformation and how they changed along strike, and for the timing of exhumation. The results also provide clues about the regional paleogeography and tectonic configuration of the Arctic region during the critical time that the Amerasia Basin was formed and the dispersal pathways for sediment that were derived from the Uralian-Taimyr orogen and deposited on the opposite side of the Arctic Ocean in the synorogenic deposits of the Brooks Range. GEOLOGIC FRAMEWORK The Brooks Range fold-thrust belt extends E–W across northern Alaska from the Canadian border to the Chukchi Sea and probably

94

continues farther west beneath the Chukchi Sea as the Herald and Wrangel arches to perhaps as far west as the New Siberian Islands, ~750 km northwest of Figure 1 (Drachev et al., 2010; Grantz et al., 2011a). The Brooks Range and adjacent Colville Basin in northern Alaska are thought to be the products of north-directed Jurassic and Early Cretaceous arc-continent collision between the south-facing (present coordinates) passive continental margin of the Arctic Alaska terrane and the Jurassic Koyukuk arc terrane of interior Alaska (e.g., Moore et al., 1994). The collision resulted in closing of the Angayucham Ocean and northward emplacement of the Angayucham terrane, an oceanic assemblage composed of a Jurassic ophiolite and subduction complex, and the Koyukuk arc terrane onto Arctic Alaska (Fig. 2A). A similar history is recognized in the Russian Far East, where the closing of the South Anyui Ocean basin culminated in collision of Carboniferous to Jurassic arc terranes and Jurassic and Early Cretaceous ophiolite and subduction complexes of the Alazeya-Oloy fold belt and South Anyui terrane onto the Chukotka terrane to the north (Sokolov et al., 2009; Sokolov, 2010). In western Chukotka, the closure of the South Anyui Ocean may have culminated in collision with the outer margin of Siberia (Miller et al., 2009). The arc-continent collision along the entire length of the orogen was approximately concurrent with rifting of the Arctic Alaska and Chukotka terranes away from their original positions along the northern margin of North America, probably by counterclockwise rotation during formation of the Amerasia Basin in the Jurassic and Early Cretaceous (e.g., Lawver et al., 2002; Grantz et al., 2011b), although other scenarios have been proposed (e.g., Lane, 1997; Miller et al., 2006, 2009; Amato et al., 2009; Dickinson, 2009). Models for the late Mesozoic collisional events in Alaska and the Russian Far East portray these areas as parts of a complex region of arcs and microcontinental fragments along the northern margin of North America in the ancestral northern Pacific Ocean (e.g., Rubin et al., 1995; Nokleberg et al., 2000). These models correlate the Angayucham and South Anyui oceans as parts of the same late Paleozoic to Mesozoic oceanic basin and the Arctic Alaska and Chukotka terranes as eastern and western parts, respectively, of the same continental margin. Similarities in Neoproterozoic and lower Paleozoic stratigraphy, plutonic belts, and Mesozoic deformational and metamorphic events support correlation of the terranes (e.g., Natal’in et al., 1999; Dumoulin et al., 2002). In this paper, “Arctic Alaska–Chukotka microplate” is used to designate the conjoined Arctic Alaska and Chu-

Geosphere, February 2015

kotka terranes, but we refer to “Arctic Alaska” and “Chukotka,” respectively, to specify features limited to only those parts of the microplate. In the western Brooks Range, the forward (northern) part of the orogen is composed of a series of stratigraphically thin thrust sheets consisting of Upper Devonian to Jurassic continental margin sediments, each depositionally overlain by a succession of deep-marine synorogenic clastic deposits (Martin, 1970; Mayfield et al., 1988) (Fig. 3). From bottom up, these allochthons consist of the Endicott Mountains allochthon (EMA), the Picnic Creek allochthon (PCA), Kelly River allochthon (KRA), Ipnavik River allochthon (IRA), and Nuka Ridge allochthon (NRA) (Fig. 2B). A proximal autochthonous succession underlies the allochthons and constitutes basement for the orogen-derived deposits in the foreland. The upper Paleozoic and lower Mesozoic deposits in both the allochthons and the autochthon consist of generally similar, texturally and compositionally mature strata, including Devonian and Lower Mississippian quartz- and chert-rich siliciclastic strata (Endicott Group), Mississippian and Pennsylvanian carbonate platform deposits (Lisburne Group), and Permian and Triassic fluvial to outer continental margin clastic, shale, and siliceous deposits (Sadlerochit Group, Shublik Formation, and Etivluk Group) (Fig. 4). In the Nuka Ridge allochthon, the compositionally mature Mississippian siliciclastic deposits are replaced by coeval arkosic deposits that have yielded abundant zircons with ca. 2.1–2.0 Ga U-Pb ages and microclines with 1.2–1.1 Ga K-Ar ages (Moore et al., 1997). Granitic rocks of this age and cooling history are not known in the Arctic Alaska– Chukotka microplate but have been reported locally in continental terranes in southwestern Alaska and Siberia (Moll-Stalcup et al., 1996; Bradley et al., 2007; MacLean et al., 2009). Stratigraphic and structural relations indicate that the western Brooks Range constitutes a latest Jurassic and earliest Cretaceous forelandyounging thrust stack (Mayfield et al., 1988). Paleontologic data indicate that synorogenic deposits are as old as latest Jurassic in the higher allochthons but are probably no older than Valanginian (Early Cretaceous) in the Endicott Mountains allochthon (Fig. 4). This suggests that each allochthon was covered by ancestral foreland basin successions derived from the orogenic belt at the time it was located near the front of the orogen. As thrusting propagated northward, foreland deposits were successively detached with their underlying platform successions as allochthons and incorporated into the thrust belt. Estimates of total shortening represented by the Brookian orogen range from 250 to 500 km or more (Moore

Evolution of lower Brookian foreland basin deposits, northern Alaska 70°N

160°E

170°E

S. A

W ra

East Siberian Sea

nyu i su ture SA

el-

170°W

He

12d 12b 12c

a Bas

ld

thr

9,10

us

in

?

Eastern Chukotka

OM

3 2 1

North Slope Colville Basin 13b

7,8 AM

Brooks

4

e

ng

Ra

13a

AM

Hope Basin

? 5f ia ss Ru .S. Seward U Peninsula

? Bering Sea

Overlap and adjacent units

?

5a-d

t

RT

KY

KK

?

Tintina faul

KB

VM

KK

140°W

Beaufort Sea

Lisburne Pen. Corwin Bluff

S. Chukchi Basin

KK

t

Chukchi Sea

Western Chukotka

OM

150°W

Canada Basin ra

6

Long

160°W

ada Can S. U.

e

11

ng

Wrangel Island

12a

zon

AL-OL

180°

YTT?

YTT

KB

5e

lt.

gf

lta

Ka

FW

PAW CG

KY

Arctic Alaska-Chukotka Microplate (AACM)

Surficial deposits (Quaternary)

Colville (foreland) Basin (Cret.)

Successor basins (Tertiary)

Hinterland basins (Cret.)

Okhotsk-Chukotka Volcanic Belt (OCVB) (Late Cret.)

Syntectonic deposits (Late Jur. and Early Cret.)

Accreted terranes of Pacific margin (Jurassic to Cen.)

Clastic turbidite unit (Trias.)

Accreted continental terranes (Prot.-Paleoz.)

Parautochthon (Prot. to Mes.)

Accreted terranes of South Anyui & Angayucham oceans (Dev. to Early Cret.) Autochthonous North America (Prot.-Mes.)

Allochthons (Dev. to Early Cret.) Semischists, phyllites (Prot.-Paleoz.) Penetratively deformed schists, gneisses (Prot.-Paleoz.)

Detrital zircon U-Pb age locations

Contact Basin axis and depositional plunge Cenozoic fault; teeth on upper plate of thrust, barbs on downthrown block of normal fault Mesozoic fault; teeth on upper plate of thrust, barbs on downthrown block of normal fault

?

Boundary of the AACM, dashed where approximate, querried where location poorly constrained

1-3 This study 4-13 Previous studies

Figure 1. Map showing geology and selected locations and structural features of the Arctic Alaska–Chukotka microplate (AACM) and locations of detrital zircon U-Pb samples discussed in this paper. Intrusive rocks not shown. Box in western Brooks Range shows location of Figure 3. Edge of continental shelf in Arctic Ocean depicted at 200 m isobath. Map modified from Moore et al. (2000); sources: Grantz et al. (2011a), Kos’ko et al. (1993), Nokleberg et al. (2000), Patton et al. (2009), Sokolov (2010), Sokolov et al. (2002, 2009), and Tikhomirov et al. (2008). Abbreviations: Accreted terranes of Pacific margin: CG—Chugach; KK—terranes of Koryak-Kamchatka region; PAW— Peninsula, Alexander, and Wrangellia. Accreted terranes of South Anyui–Angayucham oceans: AL-OL—terranes of the Alezeya-Oloy fold belt; AM—Angayucham; KY—Koyukuk; SA—South Anyui; VM—Vel’may. Accreted continental terranes: FW—Farewell; OM— Omolon; RT—Ruby; YTT—Yukon-Tanana. Hinterland basins: KB—Koyukuk Basin. Detrital zircon U-Pb age sample localities for westernmost samples collected for this study: 1—TM-01; 2—TM-02; 3—94TM-47 (see Table 1; additional sample localities shown on Fig. 3). Sources for detrital zircon U-Pb analyses from Paleozoic sandstones of the AACM: 4—Dumoulin et al. (2013); 5a–5f—Amato et al. (2009); 6–7—Miller et al. (2010). Triassic sample localities: 8—Miller et al. (2010); 9–10—Miller et al. (2006, 2010); 11—Miller et al. (2006). Jurassic and Cretaceous sample localities: 12a–12d—Miller et al. (2009) and 13a (Fortress Mountain Formation) and 13b (Nanushuk Formation)—Wartes (2008).

Geosphere, February 2015

95

Moore et al.

96

South

North Koyukuk arc terrane

Angayucham ocean basin Angayucham terrane

Arctic-Alaska continental margin

(NRA)

uPz-lMz sed. cover (IRA) (KRA) (PCA) (EMA)

Km. 0

(CPA) (MMA)

10

pre-Miss. basement

t rus

B S

20

nc

ea

Oc

30

Brooks Range fold-thrust belt

Colville (foreland) basin

N

Misheguk Mountain

An allochthon (MMA) g te ayu rra ch n e a m Copter Peak allochthon (CPA)

Ar

cti

cA las k

Foredeep

Wedgetop

deposits deposits Ko Nuka Ridge allochthon (NRA) ? ? Ko Ipnavik River allochthon (IRA) ? Nanushuk Formation Ko Fortress Kelly River allochthon (KRA) ? Ko Mountain Picnic Creek allochthon (PCA) Torok Shale Ko Formation Endicott Mountains allochthon (EMA)

a

te

rra

Kingak Shale

ne

Au

Ellesmerian sequence

to

ch

Pre-Mississippian metamorphic rocks Upper allochthons Lower allochthons Autochthon Platform deposits of Arctic-Alaska terrane

Ko

Okpikruak Formation

C

Ma SW

th

on

Depositional contact Thrust Suture

NE

100

Nanushuk Fm.

Albian

Foredeep deposits 113

Early Cretaceous

et al., 1986; Oldow et al., 1987; Mayfield et al., 1988; Young, 2004). The structurally highest allochthons in the Brooks Range are klippen of the Copter Peak allochthon (CPA) and the overlying Misheguk Mountain allochthon (MMA), which together form the oceanic Angayucham terrane (Moore et al., 1994) (Fig. 2B). The Copter Peak allochthon consists of imbricates of mafic igneous rocks and covering units of chert that are thought to compose an early Mesozoic subduction assemblage, whereas the Misheguk Mountain allochthon consists principally of Middle Jurassic ophiolite (Harris, 2004) that probably represents forearc basement to the Koyukuk arc located farther south in interior Alaska (Fig. 2A). The hinterland of the Brookian orogen consists of a continental assemblage of Devonian and older quartz schist, marble, and orthogneiss exposed in the southern Brooks Range. Large 395–380 Ma orthogneiss batholiths intrude these rocks in addition to sparse, smaller orthogneiss plutons dated at 750–700 Ma and 971 Ma (e.g., Karl and Aleinikoff, 1990; Aleinikoff et al., 1993). Most of these rocks underwent high-pressure/low-temperature metamorphism between ca. 160 and 120 Ma (Late Jurassic to Early Cretaceous) (Christiansen and Snee, 1994) followed by regional greenschist-facies metamorphism at the time of exhumation, ca. 120–90 Ma (Albian–Cenomanian) (e.g., Miller and Hudson, 1991). Exhumation was accompanied by south-side-down normal faulting along the southern flank of the Brooks Range (e.g., Little et al., 1994) (Fig. 1). These events produced a regional northward tilt that extends

A

Torok Shale Mt. Kelly 1 Graywacke

Aptian

Wedgetop deposits

Mt. Kelly Graywacke

126

L. Brookian shale

Barremian Hauterivian Valanginian

131 134 139

Berriasian

Late Jur.

Figure 2. Conceptual models for stratigraphic a nd t e c t o ni c f r a m ew o r k o f the Brooks Range fold-thrust belt and the Colville Basin. (A) Arc-continent collisional model showing closure of Angayucham Ocean basin by subduction of continental margin of Arctic Alaska beneath Koyukuk arc in the Late Jurassic and Early Cretaceous. (B) Structural stacking of allochthons in the Brooks Range and the relation of ancestral (Okpikruak Formation), wedgetop (Fortress Mountain Formation), and foredeep (Torok and Nanushuk formations) foreland basin deposits to allochthons and underlying autochthon. (C) Chronostratigraphic diagram for Brookian foredeep, wedgetop, and allochthonous deposits. Restored relative positions of allochthons prior to collision are shown in parentheses in diagram A. Diagrams shown in presentday coordinates.

Okpikruak Formation 2

Hue Shale

Kingak Shale

145

Unit Jw

Tithonian

Allochthonous deposits

Tingmerkpuk Ss.

152

THRUST BELT

COLVILLE BASIN

across the southern Brooks Range and the frontal part of the orogen to the axis of the Colville Basin (Vogl et al., 2002). Rocks similar to those in the southern Brooks Range are also present in the Seward Peninsula and the Ruby terrane south of the Brooks Range (e.g., Roeske et al.,

Geosphere, February 2015

Autochthonous deposits 1

Tongue of Fortress Mountain Formation

2

Composite time span of Okpikruak in all allochthons

1995; Amato et al., 2009) suggesting that the hinterland of the Brooks Range orogen probably extends southward into those areas (Fig. 1). Deformed and metamorphosed Paleozoic and older rocks also constitute most of basement of Chukotka (Bering Strait Geologic Field Party,

163° 69°

162° 30’

162°

A

161° 30’

0

Tupikchak

10

20

Creek

Kilometers U-Pb sample locality

ise

Cre

Tupikchak Mtn. x

Poko Mtn.

ek

x

4

Anticline x

NATIONAL PETROLEUM RESERVE—

Ri

ve

r

Igloo Mtn.

N

L. Jur. and E. Cret.

rpr

Riv

Su

Contact

er

Surprise Ck. anticline

ALASKA

Eagle Creek well Creek g Tin

Torok Fm.

Mt. Kelly Tongue, Fortress Mountain Formation

Okpikruak Fm.

rk me

puk

Unit Jw Jur.

Thrust fault, (map), teeth on upper plate

Nanushuk Fm.

Jur.

gl e

Surficial deposits

lik

Ea

Koko

Syncline

Barr.- Aptian- Albian- Cenozoic Aptian Cenom. Cenom.

Evolution of lower Brookian foreland basin deposits, northern Alaska

Misheguk Mtn. alloc.

Copter Peak alloc.

10 Hill

Inac

iv lD nta ine t n Co

9

ces

sibl

12

ge

e

ide

13

“Wulik Knot” xAmphitheatre

Sheep Mtn.

Mtn. AVAN

HILLS

lt

7 Wren

ch

ek

Red Dog Mine ok

Rive

Mtn.

rur

x Deadlock

Kelly River alloc.

Red Dog Key Ck. plate plate

2 1

Wolverine plate Cre

gu

Sivukat Mtn. fault

. fau

Ku

8

h Ck

Avan

Wre nc

Kelly

RED DOG MINERAL DISTRICT

Miss.E. Cret.

River

Riv

er

Cenozoic normal fault (cross section)

Ipnavik River alloc.

Picnic Creek alloc. L. Dev- Miss.E. Miss. E. Cret.

x

Nuka Ridge alloc.

Miss.- Miss.E. Cret. E. Cret.

x Chevron

Rid

Endicott Mountains allochthon

11

Jurassic and Early Cret. thrust fault (cross section)

68°

r15

14

A′

Parautochthon

un ta in

5

o M

Paleogene thrust fault (cross section)

Miss.- Miss.E. Cret. E. Cret.

Tingmerkpuk Mtn.

Jur. E. Cret.

Fr

x

Pre-Miss. Miss.Trias

on

t

Dev.Jur.

6

Normal fault (map), teeth on hanging wall

Beaufortian Seq.

Ellesmerian Seq.

Metasedimentary rocks

Figure 3 (on this and following page). Geologic map and structural section of the western Brooks Range and Colville Basin (modified from Moore and Potter, 2008) showing the locations of rocks sampled for detrital zircon U-Pb age analysis. Map and cross section depict the structural configuration of the western Brooks Range as the product of multiple tectonic events including thin-skinned thrusting in the Late Jurassic and Early Cretaceous (black faults), deeper seated thrusting in the Paleogene (red faults), and extensional faulting in the Cenozoic (blue faults). Localities sampled for U-Pb age dating are shown with numbered stars as follows: 4—98TM-365; 5—04TM-141; 6—04TM-143; 7—10TM-105A; 8—03CP-10; 9—06TM-29A; 10—10TM-78D; 11—10TM-82D; 12—04CP-30; 13—03CP-10; 14—06TM-08; 15—08CP-07. Sample numbers and structural-stratigraphic positions are shown in Table 1.

Geosphere, February 2015

97

6 8 10 10 km. No vertical exaggeration Ellesmerian sequence

Mt. Kelly Graywacke EMA2 EMA1

Kelly River Avan Hills MMA CPA

10

8

4 CPA 6

2

0

Unit Jw

EMA1

KRA IRA Kugururok Km. River

BROOKIAN SEQUENCE

Figure 3 (continued).

Continental Inaccessible Divide Ridge Okpikruak Tingmerkpuk Okpikruak Mountain Fm. PCA Fm. KRA

Bend

Bend

Western Brooks Range A′ 98

1997), although prograde high-pressure metamorphic conditions have not been confirmed in these rocks and a covering succession of allochthons composed of imbricated platform and continental margin sedimentary strata like those in the Brooks Range have not been identified. Deformed and metamorphosed Triassic turbidite clastic rocks (Ustieva unit of Sokolov et al., 2009) instead overlie the Paleozoic rocks in Chukotka. These rocks are overlain by allochthons of ophiolite (the South Anyui terrane) and Upper Jurassic and Lower Cretaceous orogenic strata similar to the Copter Peak allochthon and Misheguk Mountain allochthon in the structurally highest parts of the Brooks Range (Sokolov, 2010).

Pre-Mississippian basement

4 Torok Fm.

0 2 Nanushuk Fm.

Beaufortian sequence

Km. Tupichak syncline Mountain front

Eagle Creek well

Colville Basin

Surprise Creek anticline

A

Moore et al.

Ancestral foreland basin strata in the Brooks Range and depositional fill of the Colville foreland basin are composed of litharenites derived from the Brooks Range (Mull, 1985; Bird and Molenaar, 1992). These deposits compose the older part of the upper Mesozoic and Cenozoic Brookian sequence, which is distinguished by its lithic composition and southern provenance from the underlying quartz-rich, northerly derived platform and rift-shoulder deposits of the upper Paleozoic and lower Mesozoic Ellesmerian and Beaufortian sequences (Lerand, 1973; Hubbard et al., 1987). In the Colville Basin, the early Brookian clastic sediments prograded longitudinally along the basin from west to east during the Aptian and Albian (late Early Cretaceous), although paleocurrent data near the southern margin of the basin indicate significant detritus also came from the Brooks Range to the south (Molenaar, 1988; Bird and Molenaar, 1992; Houseknecht et al., 2009). A major deltaic depocenter in the western part of the basin composed mainly of westerly derived sediments is termed the Corwin Delta, whereas a large depocenter located in the central North Slope composed of southderived sediment is termed the Umiat Delta (Fisher et al., 1969; Ahlbrandt et al., 1979; Huffman et al., 1988) (Fig. 5). East-west seismic reflection profiles along the axial part of the basin reveal a spectacular clinoform geometry that has driven the Cretaceous stratigraphic nomenclature employed in the basin (Molenaar, 1988; Bird and Molenaar, 1992). Bottomset seismic facies mark deposits of condensed deep marine mudstone and radioactive shale that are assigned to the Barremian to Campanian Hue Shale (Fig. 2C). Stratigraphically higher foreset seismic facies correspond to deep marine basinal, slope, and outer-shelf mudstone and sandstone turbidites

Geosphere, February 2015

of the Aptian to Cenomanian Torok Formation. Topset facies characterize shallow-marine to non-marine coal-bearing sandstone, mudstone, and conglomerates of the Albian to Cenomanian Nanushuk Formation (Fig. 6A), which are >6 km thick in the western Brooks Range (Mull et al., 2003). Along the southern margin of the Colville Basin, the Hue Shale is absent and the basal strata consist of coarse-grained shallow-marine deposits assigned to the Fortress Mountain Formation (Molenaar, 1988; Houseknecht and Wartes, 2013) (Figs. 2B and 2C). In the western part of the basin, these strata consist of up to ~3 km of sand-rich deposits that are characteristically rich in detrital carbonate and white mica (Fig. 6C). This part of the Fortress Mountain Formation is called the Mount Kelly Graywacke Tongue of the Fortress Mountain Formation (Mull, 1985), which for simplicity is referred to as the Mount Kelly Graywacke in this paper. Locally, the Mount Kelly Graywacke is underlain by more than 250 m of thin-bedded, finegrained sandstone and mudstone that was informally named the lower Brookian shale by Mull (1995). These strata have yielded Hauterivian microfossils and are interpreted to represent the earliest phase of Colville Basin sedimentation in this part of the basin. The Mount Kelly Graywacke is probably Aptian (Mull et al., 2000), but stratigraphic relations allow it to be as old as Hauterivian or Barremian. It is overlain by slope mudstone deposits of the Torok Formation. The well-developed progradational patterns of the axial part of the Colville Basin depositional fill suggest that its provenance lies in the Lisburne Peninsula, Chukotka Peninsula, and now subsided highlands such as the Herald Arch (Molenaar, 1988). Based on compositional observations, Mull (1985) concluded that the Mount Kelly Graywacke was derived from sources in both the metamorphic hinterland and the allochthons that comprise the fold-thrust belt of the Brooks Range. Ancestral foreland basin deposits associated with the allochthonous sequences of the Brooks Range are largely assigned to the Okpikruak Formation (Figs. 2B and 2C), which consists of deep-marine micaceous sandstone, shale, polymict conglomerate, and pebbly mudstone (Moore et al., 1986; Young, 2004). In the Red Dog Mining district where it is best described, the Okpikruak Formation comprises successions of dark mudstone and thin-bedded turbidites more than 375 m thick (Young, 2004). Polymict conglomerate is locally common and contains angular to well-rounded clasts of carbonate rock and chert derived from the Lisburne and Etivluk groups, igneous clasts including basalt, granite, gabbro, and ultramafic rocks

Evolution of lower Brookian foreland basin deposits, northern Alaska

Ma 90

MMA

CPA

NRA

IRA

KRA

PCA

North Slope autochthon, Colville Basin

EMA

145

Misheguk igneous complex

134

139

unit Jw 160

?

?

?

2 Foreland basin

120 131

Okpikruak Formation

1

Kingak Shale

?

1 (0.15)

3+

200

0.02

245

0.11

0.09

Shublik & Sag River Formations

Etivluk Gp.

0.12

0.3 (0.07)

Sadlerochit Group

0.2

Basaltchert accret. complex

300

deposits

0.6 (0.11)

Uncf Nuka Fm.

~4

0.1

0.6

0.07

0.07

0.6

Uncf 0.24

Lisburne Group

0.13

Kayak Shale

4.4

Endicott Group

0.3

350

1.2 (1.7)

0.35

0.1

?

? 360

0.15

Baird Gp.

0.3

385

0.4 (0.3)

Uncf

0.5 (0.6)

0.7

Deformed Neoprot. to M. Dev. strata

Explanation Suprasubduction ophiolite

Lithic clastic deposits

Chert & shale

Mafic igneous rocks

Chert

Marine shale

Inner shelf carbonate deposits Outer shelf carbonate deposits

Arkosic carbonate deposits

Deltaic quartz-rich clastic deposits

Shelf quartz-rich clastic deposits

Deformed strata

Figure 4. Summary stratigraphic columns for allochthonous and autochthonous units of the Brooks Range and North Slope. Colors denote stratigraphic units as labeled; patterns show rock types. Red numbers show maximum stratigraphic thicknesses of pre-Cretaceous strata in kilometers in allochthons of the western Brooks Range (Curtis et al., 1984, 1990; Ellersieck et al., 1990; Kelley and Jennings, 2004) and for autochthon in eastern North Slope and Brooks Range (Moore et al., 1994) and Lisburne Peninsula (in parentheses) (Campbell, 1967; Moore et al., 2002). Thicknesses for Misheguk Mountain allochthon (MMA) and Copter Peak allochthon (CPA) estimated from Figure 3 (this study). Recessed units on stratigraphic column denote mechanically incompetent units that are subject to structural detachment. Age scale is variable. See Figure 2B for allochthon abbreviations. Foreland basin deposits on North Slope autochthon subdivided into 1—lower Brookian shale and Fortress Mountain Formation, and 2—Torok and Nanushuk formations.

that may have been derived from the ophiolite of the Misheguk Mountain allochthon and Copter Peak allochthon, and schist, orthogneiss, and other rock types that are not found in any of the allochthons (Young, 2004; De Vera, 2005) (Fig. 6E). Olistostromes are locally common, containing meter- to kilometer-scale blocks of these same lithologies, with blocks of the Etivluk and Lisburne groups being particularly common (Fig. 6F). The olistostromes have both mudrich and sand-rich matrices and are interpreted to record deposition at sites proximal to locations of active thrusting (De Vera, 2005). Silicacemented quartz-chert sandstones such as those of the Endicott Group in Endicott Mountains allochthon and Picnic Creek allochthon and in the lower Lisburne Group in Kelly River allochthon evidently are poorly represented or absent

as clasts in the conglomerates and olistostromes. Outside of the Red Dog Mining district, the Okpikruak Formation mainly consists of thinto medium-bedded, fine- to medium-grained sandstone turbidites with local conglomerate bodies and mudstone sequences in all of the continental allochthons (e.g., Curtis et al., 1984, 1990; Mull, 1985) (Fig. 6D). Stratigraphic thicknesses up to 3 km are reported (Mayfield et al., 1988), although these sections have few marker beds and may have been duplicated by unmapped faults. The oldest known Brookian deposits in the western Brooks Range form a thin-bedded sequence of volcanogenic sandstone and mudstone turbidites that is more than 150 m thick exposed in stream cuts along the Kugururok River (Figs. 2C and 6B). Curtis et al. (1984)

Geosphere, February 2015

reported the presence of Late Jurassic Buchia pelecypods and mapped these rocks as “unit Jw” because of uncertainties about their stratigraphic and structural position. Despite local faulting, field relations of these strata suggest to us that unit Jw turbidites most likely rest on mafic rocks belonging to either the Copter Peak allochthon and/or the Misheguk Mountain allochthon. METHODS We collected 17 samples of fine- to mediumgrained sandstone from representative parts of the Brookian sequence mostly along a northwest-trending transect extending from the structurally highest parts of the fold-thrust belt in the western Brooks Range to the axial part of the Colville Basin along the Chukchi Sea

99

Moore et al. 71N°

100 km

A

100 mi

97 Ma 85 Ma

Chukchi Se a 120 Ma

115 Ma

110 Ma

70°

105 Ma

Corwin He ral

69°

3

delta

d Arc

2 1

h

13b

Umiat

68°

Fig. 3

Hope Basin 166W°

164°

162°

Paleogeographic model (mid-Albian): Alluvial facies Delta-plain facies Shelf facies Slope and basin facies

BROOKS 160°

13a

delta

RANGE

158°

156°

154°

Paleoflow direction from clinoforms in seismic data Paleoflow direction from outcrop studies Detrital glaucophane (Till, 1992) U-Pb location, this study U-Pb location (Wartes, 2008) Well

152°

150°

Shelf-margin positions and age: 120 Ma Estimated 97 Ma Mapped

Figure 5. Map showing paleographic model for the Colville Basin in mid-Albian time (Huffman et al., 1988, with modifications from Houseknecht et al., 2009). Southern and western boundaries of the basin are postdepositional thrust faults that bring up older rocks. Yellow lines show estimates of shelf-margin positions at specified times between 120 and 85 Ma (Aptian, Albian, and Late Cretaceous; Schenk et al., 2012). Also shown are paleoflow data (Bird and Andrews, 1979; Huffman et al., 1988; Houseknecht et al., 2009), detrital glaucophane localities (Till, 1992), detrital zircon U-Pb localities of Wartes (2008), and the westernmost zircon U-Pb samples of this study.

for sedimentary petrography, detrital zircon U-Pb geochronology, and zircon fission-track dating. We also dated two clasts, a quartz-rich sandstone and a granitic rock, from a conglomeratic debris flow in the Okpikruak Formation of the Picnic Creek allochthon (Figs. 1 and 3 and Table 1). Sample description information is provided in the Supplemental File1. Thin sections of these sandstone samples and two additional samples from closely related sites were stained for potassium feldspar and point counted (n > 400) using the Gazzi-Dickinson counting technique to minimize com1 Supplemental File. The Supplemental File contains sample locality information and descriptions, discussions of the U-Pb and zircon fission-track analytical techniques employed in the study, tables reporting the raw sandstone point-count data, U-Pb zircon age data and zircon fission-track age data, and figures showing relative probability plots for U-Pb detrital zircon samples, graphical evidence for Pb loss in sample 06TM-08, and modeling of the zircon fission-track data. If you are viewing the PDF of this paper or reading it offline, please visit http://dx.doi .org/10.1130/GES01043.S1 or the full-text article on www.gsapubs.org to view the Supplemental File.

100

positional dependence on grain size (Ingersoll et al., 1984; Zuffa, 1985) (Tables 2 and 3; Supplemental Table SF1 [see footnote 1]). Grains affected by low-grade diagenesis were counted as framework grains if their origin was identifiable; otherwise they were recorded as diagenetic minerals. Volcanic grains were identified by the presence of volcanic texture; possible volcanic grains that lack volcanic texture were counted as chert if silicic in composition or as argillite if mafic and altered to chlorite. Detrital carbonate grains were included as framework grains on ternary diagrams (Ingersoll et al., 1987), but accessory minerals and fossils were excluded. Point count framework grain types and recalculated modal data are presented in Tables 2 and 3 and raw data are shown in Supplemental Table SF1 (see footnote 1). Zircon ages in 15 sandstone samples and two conglomerate clasts were determined using a laser ablation–inductively coupled plasma mass spectrometer (LA-ICPMS) at the Washington State University Geoanalytical Laboratory in Pullman, Washington. The U-Pb ages of 60–110 zircon grains were determined for each sample.

Geosphere, February 2015

The U-Pb ages of an additional 30 grains in four of the same sandstones and 60 grains from another sandstone were determined using the Agilent 7700× quadrapole of Donelick Properties in Viola, Idaho. Zircon fission-track ages of the same 30 zircons from the four sandstones were also determined, such that a total of 120 grains were double dated by U-Pb and zircon fission-track methods. Analytical techniques for the U-Pb geochronology and zircon fission-track dating are presented in the Supplemental File and Supplemental Tables SF2, SF3, and SF5 (see footnote 1), and the complete U-Pb geochronologic results are reported in Excel format in Moore (2014). For the purpose of discussion, zircon U-Pb ages determined on the two instruments were combined without bias, resulting in 80–129 zircons being dated from each sample and a total of 1888 zircons dated in this study. Histograms of U-Pb ages, relative probability curves, and weighted mean U-Pb ages were calculated using Isoplot 3.0 (Ludwig, 2005); cumulative probability curves and Kolmogorov-Smirnoff (K-S) statistics (Press et al., 1986; Guynn and Gehrels, 2010) were determined using the plotting program of Gehrels (2010). The geologic time scale of Walker et al. (2013) is used in this paper. We have organized the samples into five groups based on structural and stratigraphic position for the purpose of interpretation (Table 1). The two samples from unit Jw are grouped together because of their apparent depositional position on the allochthonous oceanic rocks of the Angayucham terrane. The samples from the Ipnavik River allochthon, Kelly River allochthon, and Picnic Creek allochthon are combined together as a “highallochthon” group, whereas the samples from the Endicott Mountains allochthon and Endicott Mountains allochthon(?) are grouped together as a “low-allochthon” group. For the samples from the Colville Basin, we have followed the classification scheme of DeCelles and Giles (1996) for foreland basin deposits and grouped together the samples from the lower Brookian shale and the Mount Kelly Graywacke on the southern flank of the basin as a “wedgetop” group, whereas the three samples from the Nanushuk Formation are grouped together as a “foredeep” group. SANDSTONE PETROGRAPHY Brookian sandstones range in composition from litharenite to feldspathic litharenites, sublitharenites, and calclitharenites. The quartzrich sandstone clast from the debris flow in the Okpikruak Formation, in contrast, is a subarkose. Framework grains identified and

Evolution of lower Brookian foreland basin deposits, northern Alaska

A

C

E

B

D

F

Figure 6. Photos of lower Brookian deposits. (A) Shallow-marine foredeep sandstone (Nanushuk Formation) at Tupikchak Mountain; (B) convolute lamination and ripple cross lamination in unit Jw, sample location 06TM-08, Kugururok River; (C) thick-bedded wedgetop sandstone of Mount Kelly Graywacke, near Eagle Creek well; (D) folded turbidites of the Okpikruak Formation, Picnic Creek allochthon (PCA), western Inaccessible Ridge; (E) conglomeratic debris-flow deposits in Okpikruak Formation, PCA, sample location 06TM-29, headwaters of Kukpowruk River; (F) carbonate blocks in olistostrome of Okpikruak Formation, PCA, headwaters of Kukpowruk River.

Geosphere, February 2015

101

Moore et al. TABLE 1. INFORMATION FOR SAMPLES ANALYZED IN THIS PAPER Sample number

Map location

Stratigraphic unit

Allochthon

Lithology

Age

Latitude* (N)

Longitude* (W)

Types of analysis

Foredeep samples TM-01 1, Fig. 1 Nanushuk Formation Sandstone Albian 68.88686° 164.93878° PC, U-Pb 68.86528° 165.23053° PC, U-Pb, ZFT TM-02 2, Fig. 1 Nanushuk Formation Sandstone Albian 98TM-365 4, Fig. 2 Nanushuk Formation Sandstone Albian 68.86894° 161.79036° PC, U-Pb Wedgetop samples 96TM-23B 5, Fig. 2 Lower Brookian shale unit† Sandstone Haut.-Bar. 68.55694° 162.40800° PC Sandstone Haut.-Bar. 68.54850° 162.43731° PC, U-Pb, ZFT 04TM-141 5, Fig. 2 Lower Brookian shale unit† § Sandstone Aptian? 68.60111° 162.61233° PC, U-Pb 04TM-143 6, Fig. 2 Mount Kelly Graywacke Low-allochthon samples 94TM-47 3, Fig. 1 Igrarok Hills unit# EMA Sandstone Berr.-Val. 68.86831° 165.97419° PC, U-Pb 10TM-105A 7, Fig. 2 Okpikruak Formation EMA Sandstone Berr.-Val. 68.19008° 162.36967° PC, U-Pb 03CP-10 8, Fig. 2 Okpikruak Formation EMA Sandstone Berr.-Val. 68.10972° 163.00706° PC, U-Pb High-allochthon samples 06TM-29A 9, Fig. 2 Okpikruak Formation PCA Sandstone Berr.-Val. 68.38608° 162.72783° PC, U-Pb 06TM-29B 9, Fig. 2 Okpikruak Formation PCA Clast Berr.-Val. 68.38608° 162.72783° PC, U-Pb 06TM-29D 9, Fig. 2 Okpikruak Formation PCA Clast Berr.-Val. 68.38608° 162.72783° U-Pb 10TM-78D 10, Fig. 2 Okpikruak Formation PCA Sandstone Berr.-Val. 68.50125° 162.29475° PC, U-Pb, ZFT 10AD-5K 10, Fig. 2 Okpikruak Formation PCA Sandstone Berr.-Val. 68.50178° 162.29461° PC 10TM-82B 11, Fig. 2 Okpikruak Formation KRA Sandstone Berr.-Val. 68.51069° 161.89486° PC, U-Pb 04CP-30 12, Fig. 2 Okpikruak Formation KRA Sandstone Berr.-Val. 68.51167° 161.57647° PC, U-Pb 03CP-25 13, Fig. 2 Okpikruak Formation IRA Sandstone Berr.-Val. 68.33881° 161.60581° PC, U-Pb, ZFT 03JD33 13, Fig. 2 Okpikruak Formation IRA Sandstone Berr.-Val. 68.33833° 161.60833° PC Unit Jw samples 06TM-08 14, Fig. 2 Unit Jw CPA or MMA Sandstone Tithonian 68.10100° 161.54469° PC, U-Pb 06CP-07 15, Fig. 2 Unit Jw CPA or MMA Sandstone Tithonian 68.09789° 161.56997° PC, U-Pb Note: Allochthon abbreviations: CPA—Copter Peak allochthon; EMA—Endicott Mountains allochthon; MMA—Misheguk Mountain allochthon; PCA—Picnic Creek allochthon; KRA—Kelly River allochthon; IRA—Ipnavik River allochthon. Age abbreviations: Berr.-Val.—Berriasian to Valanginian; Haut.-Bar.—Hauterivian to Barremian. Abbreviations for types of analysis: PC—sandstone point count; U-Pb—U-Pb zircon dating; ZFT—zircon fission-track double dating. *Datum: NAD27 Alaska. Nomenclature: †of Mull (1995); §Tongue of Fortress Mountain Formation; #of Moore et al. (2002).

Symbol Qm Qpq Qc P K Lsa Lac Lss Lsc Lm Lv Ld Lsp Lgr wm ch bt px am ep op hm ofg Ls Lvt L Qp Lt

102

TABLE 2. POINT-COUNTING PARAMETERS Definition Monocrystalline quartz Polycrystalline quartz Chert Plagioclase feldspar Potassium feldspar Mudstone, shale, and argillite grains Carbonaceous shale grains Siltstone grains Detrital carbonate grains Phyllite and schist grains Volcanic grains Diabase and microgabbro grains Serpentinite grains Granitic grains White mica Chlorite Biotite Pyroxene Amphibole Epidote Opaque minerals Other heavy minerals Grains of uncertain composition Total sedimentary grains (Lsa + Lac + Lss + Lsc) Total volcanic grains (= Lv + Ld) Total unstable lithic grains (= Lm + Lvt + Ls) Total polycrystalline quartz (+Qpq + Qc) Total lithic grains (= L + Qp) Recalculated parameters QtFL%Q = 100Qt/(Qt + F + L) QtFL%F = 100F/(Qt + F + L) QtFL%L = 100L/(Qt + F + L) QmFLt%Qm = 100Qm/(Qm + F + Lt) QmFLt%F = 100F/(Qm + F + Lt) QmFLt%Lt = 100Lt/(Qm + F + Lt) LmLvLs%Lm = 100Lm/(Lm + Lvt + Ls) LmLvLs%P = 100Lv/(Lm + Lvt + Ls) LmLvLs%K = 100Ls/(Lm + Lvt + Ls) QmPK%Qm = 100Qm/(Qm + P + K) QmPK%P = 100P/(Qm + P + K) QmPK%K = 100K/(Qm + P + K)

Geosphere, February 2015

counted in these sandstones include monocrystalline quartz, polycrystalline quartz, chert, plagioclase, potassium feldspar (perthite and microcline), shale, siltstone, extrabasinal carbonate, mafic and intermediate volcanic, metamorphic, serpentinite, diabasic, and granitic lithic grains (Table 2). Interstitial material in most samples is almost entirely pseudomatrix, although zeolite cement is present in samples 06CP-07 and 06TM-08. Quartz overgrowths and cement are common in sample 06TM-29B. Sandstones show evidence of diagenetic alteration including creation of one or more of the following: calcite, clay minerals, quartz, zeolites, and alteration of plagioclase to albite. Sandstone textures and grain compositions are nonetheless typically identifiable, although small patches of samples 06CP-07 and 06TM-08 are overprinted by laumontite to such an extent that grains could not be identified with certainty. These areas were avoided during point counting. Point-count data presented in Table 3 reveal that lithic grains in Colville foredeep samples are dominantly siliciclastic sedimentary, chert, and detrital carbonate grains, whereas the wedgetop group is dominated by detrital carbonate grains. Coarse white mica and metamorphic lithic grains are also common in the wedgetop deposits. Lithic grains in samples from the Okpikruak Formation in the lower allochthons include abundant siliciclastic sedimentary and chert grains and lesser proportions of detrital

Evolution of lower Brookian foreland basin deposits, northern Alaska TABLE 3. RECALCULATED MODAL POINT-COUNT DATA FOR LOWER BROOKIAN SANDSTONES, WESTERN BROOKS RANGE QtFL QmFLt LmLvtLs QmKP (%) (%) (%) (%) Qt F L Qm F Lt Lm Lvt Ls Qm K P Sample P/F Foredeep sandstones TM-01 46 16 39 31 16 53 2 0.0 98 66 0 34 1.0 TM-02 54 18 29 35 18 47 1 0.0 99 66 4 30 0.9 98TM-365 54 10 36 34 10 56 2 8 90 78 0 22 1.0 Wedgetop sandstones 96TM-23B 36 2 62 31 2 67 5 1 94 94 0 5 1.0 04TM-141 44 13 43 34 13 53 3 1 96 73 0 27 1.0 04TM-143 31 7 62 20 6 74 9 2 89 75 0 25 1.0 Low-allochthon sandstones 94TM-47 20 2 78 12 2 86 4 0 96 84 0 16 1.0 10TM-105A 51 4 45 37 4 59 21 3 76 91 0 9 1.0 03CP-10 54 23 24 41 23 36 5 0 95 64 0 35 1.0 High-allochthon sandstones 06TM-29A 24 10 65 7 10 83 5 37 58 41 7 52 0.9 10TM-78D 53 22 25 40 22 38 7 3 90 64 0 36 1.0 10AD-5K 27 4 69 21 4 75 5 7 88 84 0 16 1.0 10TM-82B 21 11 68 8 11 80 11 23 66 43 0 57 1.0 04CP-30 28 23 49 13 24 63 14 31 55 36 0 64 1.0 03CP-25 25 29 46 22 29 49 3 41 56 43 4 53 0.9 03JD33 34 28 38 27 28 45 1 55 44 50 1 49 1.0 High-allochthon clast 06TM-29B 81 13 6 79 13 9 0 0 100 86 11 3 0.2 Unit Jw sandstones 06TM-08 34 34 32 28 34 38 1 85 14 45 1 54 1.0 06CP-07 22 25 53 13 25 62 0 87 13 35 2 63 1.0 Note: See Table 2 for definition of recalculated parameters.

carbonate and metamorphic grains, whereas the Okpikruak in higher allochthons is marked by highly variable proportions of chert, siliciclastic sedimentary, detrital carbonate, and volcanic grains. Serpentinite, gabbro, pyroxene, and epidote are common in several of the samples from the Kelly River allochthon and Picnic Creek allochthon, indicating contributions from a nearby ophiolitic source. Lithic grains in unit Jw are almost entirely volcanic fragments. The quartz-rich sandstone clast from the olistostrome in the Okpikruak Formation contains few lithic fragments and is dominated by a mature assemblage of quartz and potassium feldspar. Figure 7 shows the sandstone petrographic data plotted on ternary diagrams after Dickinson (1985) and Ingersoll and Suczek (1979). We include for comparison data published by previous workers, which were produced using methodologies different from the Gazzi-Dickinson method employed in the present study and are likely to be biased toward somewhat higher proportions of rock fragments. On the total quartz-feldspar-lithic fragment (QtFL) ternary plot (Fig. 7A), most samples plot along two linear trends, one consisting mainly of the foredeep, wedgetop, and low-allochthon samples having a relatively higher proportion of siliceous grains (line A in Fig. 7A), and the other consisting of the Jw and most high-allochthon samples at relatively higher proportions of feldspar and lithic grains (line B in Fig. 7A). The samples in trend A all plot in the recycled orogenic field, whereas those in trend B plot in

the undissected to dissected arc fields. This suggests that two source areas may be represented in the data, one consisting of recycled detritus of sedimentary or metasedimentary origin (trend A) and the other consisting of debris from a volcanic arc complex (trend B). The sandstone clasts in the Okpikruak Formation plot in the craton interior and basement uplift fields in stark contrast to all of the other samples in the study. On the monocrystalline quartz-feldspar-total lithic fragment (QmFLt) diagram (Fig. 7B), the samples fall along similar trends emanating from the Lt pole. Most of the samples from the low allochthons and foredeep and wedgetop deposits of the Colville Basin plot in the lithic recycled, transitional recycled, and mixed recycled-dissected arc fields along a trend of decreasing lithic grains (trend A), whereas those from most of the higher allochthon and the unit Jw samples plot across the transitional arc and dissected arc fields (trend B). As in the QtFL diagram, the clasts from the Okpikruak conglomerates plot separately in the craton interior and basement uplift fields. On the LvLmLs plot, the samples from Colville Basin and the lower allochthons fall in the “rifted continental margins” field near the sedimentary (Ls) pole on the lithic grain diagram (Fig. 7C), whereas those from the higher allochthons and Jw plot mostly at higher proportions of volcanic component (Lv) in or near the “mixed magmatic arcs and rifted continental margins” field. The samples from unit Jw plot very close to the “magmatic arcs” field, indicat-

Geosphere, February 2015

Lvt/L 0.0 0.0 0.1 0.0 0.0 0.0 0.0 0.0 0.0 0.3 0.0 0.0 0.2 0.2 0.4 0.6 0.0 0.7 0.8

ing that they have volcanic compositions close to those of magmatic arcs. The monocrystalline quartz-plagioclase-feldspar (PQmK) ternary diagram (Fig. 7D) demonstrates that most Colville Basin and Okpikruak Formation samples have little to no potassium feldspar, with only three samples containing more than 1%. These diagrams show that most samples from the Okpikruak Formation in the higher allochthons and unit Jw are relatively enriched in volcanic and felsic components, indicating strong contributions from the Angayucham and Koyukuk terranes, which together form the upper plate of the Brookian orogen. Samples from the lower allochthons and most of those from the Colville Basin, on the other hand, contain a much lower proportion of arc material and consist instead mainly of mixtures of continental and recycled grain types. The overall pattern suggests that the Late Jurassic to Albian part of the Brookian sequence of the western Brooks Range represents an unroofing sequence. The oldest deposits, consisting of the syntectonic deposits of the Okpikruak Formation and unit Jw in the higher allochthons, represent erosion primarily of the Angayucham and Koyukuk terranes in the structurally higher parts of the Brookian orogen. The Okpikruak sandstones in the lower allochthons and most debris in the Colville Basin, on the other hand, consist mostly of debris derived from a sedimentary or metasedimentary orogenic area, which is consistent with

103

Moore et al. Figure 7. Detrital modes of Qt Qm A B Craton Brookian sandstones and conCraton interior glomerate clasts from the westinterior Quartzose ern Brooks Range. Sandstones recycled Transitional Transitional are grouped by structural posicontinental continental Recycled tion. Provenance fields of QtFL A and QmFLt ternary plots are Mixed orogenic after Dickinson (1985); proveA Transitional recycled nance fields indicated by dashed Dissected Basement Basement lines on LvLmLs ternary plot Dissected arc arc uplift uplift are after Ingersoll and Suczek Lithic B B recycled (1979). Yellow symbols indiTransitional arc Transitional arc Undissected cate analyses conducted as part arc of this study (n = numbers of F L Lt F Undissected new analyses) and reported in arc s die Table 3 and Supplemental Table y stu tud ous Qm Lm s SF1 (see footnote 1); smaller is vi Th Pre C D gray and black symbols show Unit Jw (n = 2) modes from published studies, Lower allochthons (n = 3) including Bartsch-Winkler and Higher allochthons (n = 7) Wedgetop deposits (n = 8) Huffman (1988) from Corwin Suture belts TM Bluff (CB, n = 13) and TupikForedeep deposits (n = 34) CB chak Mountain (TM, n = 16), Clasts (n = 2) Molenaar et al. (1988) (Mount Kelly Graywacke, n = 3), and Rifted Magmatic continental Wartes and Reifenstuhl (1998) Mixed magmatic arcs & arcs margins subduction complexes (Mount Kelly Graywacke, n = 2; lower Brookian shale, Mixed magmatic arcs & n = 1; Nanushuk Formation, rifted continental margins n = 1; clast from Okpikruak Lv P K Ls Formation, n = 1). Only those analyses that report plagioclase and potassium feldspar contents are shown on PQmK plot. Red lines in 7A and 7B show compositional trends of older Brookian deposits at structurally high positions in the Brooks Range (line B) and in younger Brookian deposits at structurally lower positions in the orogen and in the Colville Basin (line A).

derivation from the Paleozoic and older rocks such as those in the Arctic Alaska–Chukotka microplate. DETRITAL ZIRCON U-PB GEOCHRONOLOGY Brookian Sandstones Zircons dated in the Brookian sandstones range in age from 3116 Ma to 111 Ma (Table 4). Plotted graphically (Fig. 8), the zircon ages fall into three major age populations, one ranging from 180 to 140 Ma (Jurassic and earliest Cretaceous), another from 542 to 200 Ma (Paleozoic and Triassic), and the third from 2000 to 1750 Ma (Paleoproterozoic). Modest, broadly distributed populations of Neoproterozoic and Archean zircons (900–600 Ma and 2800–2400 Ma, respectively) can also be recognized in the data set. The Paleozoic–Triassic population is the largest and displays a broad asymmetric distribution with peaks at 329 Ma and 237 Ma. Because of the regional sub-Mississippian unconformity

104

in the Arctic Alaska–Chukotka microplate, we divided the Paleozoic–Triassic population into an early Paleozoic (542–359 Ma) subpopulation and a late Paleozoic and Triassic (359–200 Ma) subpopulation. This division also separates the younger, most voluminous part of the Paleozoic–Triassic population from its older, long tail composed of zircons that decrease in abundance with increasing age. Ages of the four resulting population groupings are shown as color bands on Figures 8–12. Age distributions of zircons in individual samples (Table 4 and Supplemental Fig. SF1 [see footnote 1]) reveal that all 15 sandstone samples consist dominantly of some or all of the population groups defined above (Fig. 9). Most samples contain few or no zircons that approximate their stratigraphic age. Exceptions to this include unit Jw, which contains abundant Late Jurassic grains as young as 148 Ma (Tithonian) at the Tithonian Buchia fossil locality reported by Curtis et al. (1984) (sample 06CP-07). In addition, samples from the Nanushuk Formation at Corwin Bluff contain sparse zircons

Geosphere, February 2015

that may approximate the stratigraphic age of this unit, including a Barremian zircon (128 ± 4 Ma) in sample TM-01 and three Aptian zircons (weighted mean average = 115 ± 4 Ma) in sample TM-02. These ages are consistent with the 120 Ma age for the Nanushuk at Corwin Bluff estimated from seismic data by Schenk et al. (2012) (Fig. 5), but are significantly older than the mainly middle and late Albian age estimated from plant fossils in the section (Smiley, 1969; Spicer and Herman, 2010) (see the Supplemental File [see footnote 1] for details about the stratigraphic positions of these samples). Other exceptions are the lower Brookian shale from the base of the southern flank of the Colville Basin, which contains Barremian zircons (sample 04TM-141, five grains, weighted mean average = 128 ± 4 Ma) and allochthonous Brookian strata from the Lisburne Peninsula, which contain a Valanginian zircon (sample 94TM-47, 135 ± 3 Ma). The Barremian zircons in the lower Brookian shale are younger than the Hauterivian age reported by Mull (1995) from paleontologic data. With the obvious exception

Evolution of lower Brookian foreland basin deposits, northern Alaska TABLE 4. FRACTION (IN PERCENT) OF DETRITAL ZIRCON GRAIN-AGE POPULATIONS IN LOWER BROOKIAN SANDSTONES FROM THE WESTERN BROOKS RANGE Mississippian Cambrian to Cretaceous Jurassic to Triassic Devonian Neoproterozoic Mesoproterozoic (2.5 Ga)

1 2 5 3

41 34 24 33

16 17 6 13

6 5 13 13

5 6 10 13

23 30 35 30

4 3 6 4

6 1 4

23 19 22

15 22 18

6 19 11

5 12 8

32 24 29

7 4 6

2 6 0 3

27 42 24 31

27 12 19 20

16 8 21 15

11 4 5 7

15 26 29 23

1 2 2 2

7 2 25 34 0 13

20 26 18 18 46 26

31 18 22 19 27 23

13 13 12 7 13 12

6 4 2 1 2 3

18 33 19 21 11 21

5 4 1 1 1 2

90 88 89

0 1 0

1 1 1

3 3 4

0 0 0

6 6 5

0 0 0

18

24

17

10

5

22

3

middle Neoproterozoic (800–600 Ma), and early and middle Paleoproterozoic (2100–1750 Ma) zircons, but only the Paleoproterozoic zircons fit one of the zircon population groups defined in Figure 8. The high-allochthon group also contains zircons of the Jurassic and earliest Cretaceous population (13%), but zircons of this age decrease in abundance in the low allochthons

and in the wedgetop and foredeep groups (both 4%–5%, Table 4 and Fig. 11). A notable exception to this trend is sample 03CP-25 from the Ipnavik River allochthon of the high-allochthon sample group, which yielded no Jurassic zircons. Volcanic detritus in thin sections from this locality suggest this sample was sourced mainly from zircon-poor mafic volcanic rocks and chert of 60

180

All Brookian Deposits (except unit Jw) n = 1364

160 140

154 329

80

Relative Probability

of unit Jw, we conclude that lower Brookian strata received only small contributions of zircons from contemporaneous arc volcanic sources until at least the Aptian. Aside from a single spike in zircon ages at 155 Ma (i.e., the Jurassic and earliest Cretaceous population), there are few other peaks in zircon ages that are present in most or all samples. This could result from the small statistical size (~100 grains) of zircon populations analyzed from Brookian deposits. To increase the statistical size of the various zircon populations and in an attempt to reveal regional trends by age and structural-stratigraphic position, we grouped samples into the five structural-stratigraphic sample groups distinguished earlier in this paper (Figs. 10 and 11). Samples from unit Jw consist almost exclusively of Jurassic and earliest Cretaceous zircons (89%, Table 4) that range in age from 180 to 140 Ma with a maximum probability of 155 Ma (Fig. 10). Sample 06TM-08 (but not sample 06CP-07) also contains a significant number of younger zircons with a peak at 129 Ma (Barremian) (not shown in Fig. 8). Nearly all of these grains have high U contents (>3000 ppm) and therefore their ages are judged to be spurious due to Pb loss. These zircon ages consequently are omitted from further consideration in this report (see Supplemental File [see footnote 1] for a discussion of these grains). Pre-Jurassic zircons in the unit Jw sandstones include Silurian to Carboniferous (435–300 Ma), early and

Number of Grains

Number of Sample grains Foredeep sandstones TM-01 79 4 TM-02 129 2 98TM-365 96 0 Total 304 2 Wedgetop sandstones 04TM-141 124 5 04TM-143 85 0 Total 209 3 Low-allochthon sandstones 94TM-47 107 1 10TM-105A 93 0 03CP-10 100 0 Total 300 0 High-allochthon sandstones 06TM-29A 95 0 10TM-78D 134 0 04CP-30 99 0 10TM-82B 101 0 03CP-25 122 0 Total 551 0 Unit Jw sandstones 06TM-08 120 1 06CP-07 99 1 Total 219 1 All analyses of Brookian sandstones Total 1583 1

237

120 100

40

80 1822

60

20

40 20 0

0 0

500

1000

1500

2000

2500

Detrital Zircon Ages (Ma)

3000

0

200

400

600

800

1000

Detrital Zircon Ages (Ma)

Figure 8. Relative age probability curves with histograms for all zircon ages from the Brookian sequence except for unit Jw. Left diagram shows all zircon ages less than 3000 Ma (one other grain at 3116 Ma); right diagram shows expanded younger part (2.5 Ga) and Paleoproterozoic (1.6–2.5 Ga) zircons, on the other hand, increase in abundance from 23% in the high allochthon sample groups to ~35% of the wedgetop and foredeep groups (Table 4). The maximum probability in the Paleoproterozoic zircon population occurs at ca. 1.8 Ga in the allochthon and wedgetop groups but is somewhat older (1.9 Ga) in the foredeep group (Fig. 11). Well-defined Paleoproterozoic and Archean zircon age peaks occur at 2.46 Ga in the low allochthons and 2.68 Ga in the foredeep sample group, but such a peak is not present in the high allochthons. Although they are recognized in some samples (e.g., 10TM-82B, 10TM-78D, and TM-01, Supplemental Figs. SF1–SF4 [see footnote 1]), zircons of 2.0–2.1 Ga age are not common in any of the structural-stratigraphic groups, indicating that the distinctive arkosic middle Paleozoic sandstones of the Nuka Ridge allochthon did not form a significant source for Brookian deposits in the western Brooks Range. In contrast to the other zircon population groups, the Paleozoic–Triassic population displays few obvious trends by structural and stratigraphic position. Early Paleozoic zircons decrease in abundance from the high-allochthon sample group (23%) to the foredeep sample group (13%), with peaks in maximum probability occurring at ca. 425–415 Ma in the wedge-

Geosphere, February 2015

top and foredeep sample groups and smaller peaks at 480–460 Ma in the allochthon sample group. Triassic and late Paleozoic zircons have significant peaks at ca. 310–345 Ma in all of the sample groups and ca. 230–240 Ma in all but the high-allochthon sample group. The high-allochthon sample group is the one sample group (aside from unit Jw) in which Triassic zircons are not common. In order to evaluate the similarity of the zircon populations, probability (P) values were calculated from K-S statistics for the individual samples from the allochthon, wedgetop, and foredeep sample groups. K-S statistics test the hypothesis that two detrital zircon age populations were not derived from the same parent population (Guynn and Gehrels, 2010). For cases in which there is a greater than 95% probability that two populations were not derived from the same parent population or share the same sediment source P < 0.05, whereas for two populations having statistically identical populations P = 1. The K-S statistics for all of the sandstone samples we dated are shown in Table 5. Zircons that are Jurassic or younger have been omitted because K-S statistics are sensitive to variations in the abundances of zircons of the same age and the previously noted variations in the abundance of the Jurassic and Cretaceous zircons could mask other trends that might be present. For Triassic and older zircons, 47 of the sample pairs have P values >0.05. Only sample 03CP-25 from the Ipnavik River allochthon has P < 0.05 for most sample pairs, suggesting that this sample is dissimilar to the others and might have a somewhat different provenance. This might be explained by its lack of Triassic zircons and a peak in the Paleoproterozoic at 1.66 Ga (Supplemental Fig. SF4 [see footnote 1]), an age younger than the range of the Paleoproterozoic zircon population group defined

Evolution of lower Brookian foreland basin deposits, northern Alaska

Number of Grains

A

154 40

60

30

40

20 1812

0

0 238

Low Allochthons Okpikruik Formation n = 300

40

20

30

15

20

10 1800

309 203 273 150 344

459 476 538

618

5

2462

759

0

0

C 20

Wedgetop Deposits Mt. Kelly Graywacke & lower Brookian shale unit n = 209

15

229 344 424 265 379

8

1818

168 132

6

10

Relative Probability

Number of Grains

463

10

20

10

4 547 2462

5

2

2585

0

D 50

Foredeep Deposits Nanushuk Formation n = 304

40

256 238 291

20

212 10

1906

10

2682

0 0

500

1000

1500

2000

2500

3000

Detrital Zircon Ages (Ma)

with time. The two samples from unit Jw, on the other hand, clearly reflect a different source area because of the dominance of Jurassic and earliest Cretaceous zircons in these rocks (Fig. 12A). Conglomerate Clasts A cobble of quartz-rich subarkosic sandstone and a boulder of a granodiorite were collected for U-Pb age dating from a conglomeratic debris-flow deposit at location 06TM-29 in the Picnic Creek allochthon. In the field, these clasts were hypothesized to have come from basement rocks and overlying lower Mississippian clastic rocks (Endicott Group) of the Arctic Alaska terrane in the structural core of the Brooks Range.

Geosphere, February 2015

800

1000

352

141 114 185

5 0

794

335 415

15 30 20

Relative Probability

Number of Grains

288 360 257

0

above. If this sample is omitted from consideration, 65% of the sample pairs have P > 0.05 and 8% have P > 0.9, indicating a significant to very high probability that most of the samples were derived from the same Triassic parent population or the same Triassic sediment source area. Similarities in the zircon populations are evident in the cumulative probability plots shown in Figure 12. These plots also show that most of the variation between samples is due to variation in the abundance of Phanerozoic and Paleoproterozoic zircon populations and that Neoproterozoic and Mesoproterozoic zircons are relatively minor components. Exceptions include samples 94TM-47 (Endicott Mountains allochthon) and 04TM-143 (wedgetop), which display steeper slopes between ca. 800–1500 Ma. Although having the unusual characteristics described above, sample 03CP-25 (Ipnavik River allochthon) displays a cumulative probability curve that is generally similar to those of the other samples from the Okpikruak Formation, suggesting that its origin may reflect local disparities in the geology of the source region rather than an entirely different provenance area. A general shift to higher proportions of Paleoproterozoic zircons in the samples from the Colville Basin relative to the samples from the Okpikruak Formation (Fig. 12B) could indicate erosion in the source area toward older rocks

329

Relative Probability

Number of Grains

B

High Allochthons Okpikruik Formation n = 551

80

Relative Probability

Figure 11. Relative age probability plots with histograms for zircons in Brookian strata deposited on the Arctic Alaska–Chukotka microplate (AACM) (i.e., from Okpikruak, Fortress Mountain, and Nanushuk formations) grouped by structural position. Diagrams on the left show age distribution of zircons less than 3000 Ma; diagrams on right show only zircons younger than 1000 Ma. (A) Structurally high allochthons, Ipnavik River allochthon, Kelly River allochthon, and Picnic Creek allochthon (samples 03CP-25, 04CP-30, 10TM-82B, 10TM-78D, and 06TM-29A). (B) Structurally low allochthons, including the Endicott Mountains allochthon (samples 10TM105A, 03CP-10, and 94TM-47). Note that one zircon from this group having an age of 3116 Ma (sample 04CP-30) is not shown. (C) Wedgetop deposits including lower Brookian unit of Mull (1995) and Mount Kelly Graywacke (samples 04TM-141 and 04TM-143). (D) Foredeep deposits of Nanushuk Formation (samples 98TM-365, TM-01, and TM-02). Numbers indicate prominent age peaks in relative probability in Ma. Color bands show the same age populations as in Figure 8.

0

200

400

600

Detrital Zircon Ages (Ma)

Detrital zircon U-Pb ages from the quartzrich sandstone clast (sample 06TM-29B) reveal a dominant population of zircons distributed between 415 and 360 Ma with a peak in probability at 412 Ma and a bimodal subordinate Paleoproterozoic population having peaks at 1771 and 1938 Ma (Figs. 13A and 13B). The sample contains only one zircon younger than 360 Ma (280 Ma), whose age we suspect has been altered due to Pb loss prior to deposition because of its high U content (>4500 ppm). For this reason, the youngest zircon age is disregarded and the depositional age of the rock is considered to be Mississippian or younger (200 MA) ZIRCONS IN SAMPLES FROM LOWER CRETACEOUS BROOKIAN DEPOSITS AND TRIASSIC DEPOSITS IN THE CENTRAL AND WESTERN BROOKS RANGE AND THE RUSSIAN FAR EAST A B C D E F G H I J K L M N O P Q Early Brookian deposits in the western Brooks Range (this study) 0.13 0.74 0.07 0.29 0.05 A—98TM-365 X 0.00 0.00 — 0.00 0.00 0.00 — 0.00 — — 0.00 0.86 0.08 0.09 0.98 0.06 0.05 0.11 0.05 0.13 B—TM-01 0.00 X 0.00 0.00 0.00 0.05 — 0.00 0.86 0.41 0.42 0.14 0.11 0.50 0.12 0.35 0.48 0.29 C—TM-02 0.00 X 0.00 0.00 — — 0.00 0.13 0.18 0.24 0.99 0.46 0.25 0.51 0.55 0.08 0.11 0.11 D—04TM-143 0.00 0.00 X 0.00 — 0.00 0.74 0.41 0.18 0.31 0.06 0.93 0.08 0.27 E—04TM-141 0.08 X 0.00 0.00 — 0.00 — — 0.00 0.24 0.10 0.17 0.74 0.59 0.18 0.15 0.97 0.6 F—94TM-47 — 0.09 0.00 0.00 X 0.00 0.00 — 0.98 0.42 0.09 G—03CP-10 0.00 0.00 0.00 X 0.00 0.01 0.00 0.00 0.00 0.00 0.00 — 0.00 0.00 0.65 0.14 0.99 0.31 0.17 0.59 0.38 0.56 0.63 0.26 0.12 0.23 H—10TM-105A 0.00 0.00 X — 0.00 0.11 0.46 0.06 0.74 0.59 0.18 1.00 0.99 0.19 0.75 0.63 I—06TM-29A 0.00 0.00 0.00 X 0.00 — 0.29 0.06 0.50 0.25 0.93 0.38 0.18 0.31 0.70 J—10TM-78D 0.00 0.00 X — 0.00 — — 0.00 0.58 0.61 0.05 0.12 0.51 0.08 0.59 0.56 1.00 0.31 0.99 0.15 K—04CP-30 0.00 0.00 X 0.00 — 0.05 0.11 0.35 0.55 0.27 0.18 0.63 0.99 0.70 0.99 0.53 0.17 0.82 L—10TM-82B 0.00 X 0.00 — M—03CP-25 — 0.05 — — — 0.00 0.00 — 0.00 — 0.00 0.00 X — — — 0.00

— 0.24 0.00 0.00 — 0.00 0.13 0.00 0.25 0.00 0.20 0.10 0.32

Lower Brookian deposits in the central Brooks Range (Wartes, 2008) 0.48 0.08 0.15 N—Wedgetop 0.00 0.05 0.00 O—Foredeep — — — 0.00 — —

0.00 —

0.00 —

0.26 0.00

0.19 —

0.00 —

0.15 —

0.53 —

— —

X —

— X

0.00 —

0.47 —

R

Triassic deposits (Miller et al., 2006, 2010) 0.11 0.97 0.12 0.75 0.58 0.17 0.22 P—W. Chukotka — 0.00 0.00 — 0.00 — — 0.00 — X — 0.13 0.29 0.11 0.60 0.23 0.63 0.61 0.82 0.47 0.22 Q—Wrangel Is. 0.00 0.00 0.00 0.00 0.00 — X 0.00 0.24 0.13 0.25 0.21 0.10 0.32 R—Lisburne Pen. — 0.00 0.00 — 0.00 0.00 0.00 0.00 — — 0.00 X Note: P—probability values calculated from Kolmogorov-Smirnoff (K-S) statistics using error in cumulative distribution function (CDF) (Guynn and Gehrels, 2010). P values are rounded to the nearest 0.01. Where P is shown as 0.00, 0 < P ≤ 0.05; dashes indicate P = 0.000. Bold type denotes P > 0.05. Pre-Jurassic zircon ages from Brookian deposits in the central Brooks Range are taken from three samples from Fortress Mountain Formation (wedgetop) and two samples from the Nanushuk Formation (foredeep). Zircons from Triassic deposits are taken from three samples from western Chukotka (W. Chukotka), four samples from Wrangel Island (Wrangel Is.), and two samples from the northern Lisburne Peninsula (Lisburne Pen.). Numbers of analyses (n) for each sample or group of samples by letter designator: A—91; B—74; C—123; D—84; E—111; F—105; G—87; H—100; I—88; J—131; K—74; L—67; M—122; N—299; O—200; P—286; Q—320; R—187.

older zircons in the sandstone (sample 06TM29A) collected from the same locality (Table 6; Supplemental Fig. SF1 [see footnote 1]). This suggests that the quartz-rich sandstone clast may be representative of the provenance for the Devonian and older zircons in most of the Brookian sandstones and that those zircons probably were recycled from older sedimentary rocks. The zircon ages from the granitic clast (sample 06TM-29D) fall between 170 and 150 Ma and yield a weighted mean age of 159.8 ± 0.6 Ma (Late Jurassic) (Figs. 13C and 13D). This age is coincident with many of the zircon ages of unit Jw and indicates that the granodiorite was likely derived from a granitic body in the Angayucham or Koyukuk terranes rather than the basement rocks of the Arctic Alaska– Chukotka microplate. ZIRCON FISSION-TRACK AND U-PB DOUBLE DATING The thermochronologic history of a sandstone in the source region sometimes can be constrained by dating the same zircon grains using both U-Pb and zircon fission-track (ZFT) methods (i.e., “double dating”). The U-Pb isotopic system closes at temperatures of >700 °C in most natural igneous and metamorphic rocks (Hanchar and Watson, 2003), whereas the ZFT system closes at temperatures of ~200–250 °C (Lee et al., 1997; Bernet, 2009). If U-Pb and ZFT ages are plotted against each other, zircon grains that cooled rapidly at the time of crys-

108

tallization should have nearly identical (i.e., concordant) U-Pb and ZFT ages within analytical uncertainty (commonly substantial for ZFT ages) and thus fall on a line with a slope equal to 1 (Fig. 14). At the other extreme, zircons that reside at depths with temperatures >250 °C for a significant period after crystallization and later cool below the closure temperature for ZFT will fall on lines having slopes less than 1. If the zircon grains of a sample have a variety of crystallization ages (as in many sandstones) and the sample cooled through the ZFT closure temperature at a significantly later time, then the zircons will all plot at the same ZFT age regardless of the U-Pb age of the individual zircons (i.e., at a slope of zero). In that situation, if all of the zircons have ZFT ages that are greater than the depositional age of the sandstone, then the cooling must have occurred in the source area; whereas, if the ZFT age is younger than the depositional age, then the sandstone itself cooled through the closure temperature for ZFT following deposition. Thirty zircon grains from each of four of our samples were dated using both the U-Pb and ZFT methods (Supplemental Table SF5 [see footnote 1]) and the ages plotted against each other in Figure 14. These include two samples from the upper allochthons (samples 03CP-25 from the Ipnavik River allochthon and 10TM78D from the Picnic Creek allochthon, respectively), one sample from the wedgetop deposits (sample 04TM-141), and one from the foredeep deposits (sample TM-02).

Geosphere, February 2015

All four samples display generally similar distributions on ZFT versus U-Pb cross plots, indicating that the ages do not vary by the amount of stratigraphic and/or tectonic burial. In addition, all of the samples contain many grains that have ZFT ages older than the depositional age, suggesting that some ZFT ages are provenance ages. In all samples, the zircons with Paleoproterozoic U-Pb ages yield ZFT ages that are significantly younger than their U-Pb crystallization ages, with most of the ZFT ages falling between 100 and 250 Ma and all being younger than 450 Ma. This suggests that the Paleoproterozoic grains cooled through the closure temperature for ZFT and were mostly to fully annealed prior to deposition, many probably in the Mesozoic at the time of the Brookian orogeny, shown by a green band at ZFT cooling ages of 160 to ca. 100 Ma, although an older period(s) of cooling such as during the UralianTaimyr orogeny (orange band) is also permissible due to the U-Pb crystallization ages and the large uncertainties of ZFT ages. To better resolve the timing of cooling and to verify that the source area of the Brookian sandstones experienced Brookian cooling, we plotted the ZFT and U-Pb data on relative probability plots (Fig. 14). Only double-dated zircons having Triassic and older (>200 Ma) U-Pb ages are included in these plots because younger grains could not show evidence of pre-Brookian cooling events. The plots clearly show that the time of maximum probability of cooling is considerably younger than the U-Pb crystallization ages of the

Evolution of lower Brookian foreland basin deposits, northern Alaska

A

1.0

Colville Basin

0.8 0.7

upper part

0.6

lower part

EMA? EMA EMA PCA PCA KRA KRA IRA Unit Jw Unit Jw

Low allochs.

0.5 0.4

High allocs.

Cumulative Probability

0.9

0.3 0.2 Lower Brookian deposits, western Brooks Range

0.1 0

98TM-365 (n=96) TM-01 (n=79) TM-02 (n=129) 04TM-143 (n=85) 04TM-141 (n=124) 94TM-47 (n=107) 03CP-10 (n=93) 10TM-105A (n=100) 06TM-29A (n=95) 10TM-78D (n=134) 04CP-30 (n=99) 10TM-82B (n=101) 03CP-25 (n=122) 06CP-07 (n=99) 06TM-08 (n=120)

B 0.9 0.8

C

0.7 0.6 Jr3K1 units Brookian Sequence

Cumulative Probability

Figure 12. Cumulative age probability plots for Brookian deposits and rocks from possible source areas. (A) Lower Brookian samples from the western Brooks Range reported in this paper. (B) Lower Brookian sandstones in this study shown by structural-stratigraphic group and compared against Brookian sandstones from the central Brooks Range (Wartes, 2008) and Upper Jurassic and Lower Cretaceous orogenic sandstones from western Chukotka (Miller et al., 2009). (C) Devonian and older zircons from Brookian sandstones and quartz-rich sandstone clast from the central and western Brooks Range compared to possible source strata of the Arctic Alaska–Chukotka microplate (AACM) (Miller et al., 2010; Dumoulin et al., 2013) and deformed and metamorphosed AACM rocks in the Seward Peninsula (Amato et al., 2009). Light shaded area shows composition of Brookian strata in low and high allochthons, wedgetop and foredeep groups; dark shading shows composition of samples from AACM. (D) Triassic and older zircons from Brookian sandstones of the western Brooks Range compared to data from Triassic turbidite unit of Chukotka and Wrangel Island and Triassic shelf sandstone from the Lisburne Peninsula (Miller et al., 2006, 2010).

0.5 0.4 0.3 0.2 Foreland basin deposits of the AACM

0.1 0 0.9

West Brooks Range

Unit Jw (n=219) High allocs. (n=551) Low allocs. (n=300) Wedgetop (n=209) Foredeep (n=304)

Central Brooks Range

Wedgetop (n=299) Foredeep (n=200)

Russian Far East

W. Chukotka (n=401) South Anyui (n=106)

Brookian (Devonian and older zircons only) and AACM units 359 Ma

0.7

AACM units (Pz)

0.6 0.5 0.4 Brookian units (Mesozoic)

0.3 0.2

D

Wedgetop & foredeep (n=340) Allochthons (except Jw) (n=557)

West

PCA congl. clast Angayucham terrane

0.1

Lisburne Pen. (n=203) Wrangel Is. (n=505) Kuna Fm. (n=82) Seward Pen. (n=992)

Foredeep (n=185) Wedgetop (n=209)

Central

0

06TM-29B (n=95) Unit Jw (n=21)

Lower Brookian (Triassic and older zircons only) and Triassic turbidites of Chukotka

0.9 0.8

Cumulative Probability

zircons and occurred ~20 m.y. or less before the time of deposition of the samples. For three of the four samples, the time of peak cooling corresponds with the time of the Brookian deformational event. All of the samples, however, have asymmetric ZFT cooling curves with longer tails toward older ages and central (mean) cooling ages that are displaced to older ages from their respective peaks of maximum probability. This suggests that the ZFT data may reflect a mixing of a subordinate component of unannealed or partly annealed zircons having late Paleozoic and Triassic ZFT ages with the dominant population of Brookian ZFT ages. The somewhat older peak cooling age in sample 03CP-25 indicates that a relatively larger number of Paleozoic and Triassic cooling ages are present in this sample, causing the time of maximum probability of cooling to reflect an intermediate age. Analysis of the ZFT data using a deconvolution program supports the interpretation that the samples cooled initially in the Pennsylvanian, Permian, and Triassic at the time of the Uralian-Taimyr orogeny and later during the Late Jurassic and Early Cretaceous at the Brookian orogeny (Fig. 14). The zircon thermochronologic data indicate that the Brookian deposits were derived after

Cumulative Probability

0.8

200 Ma

0.7 0.6 Triassic sandstones

0.5 0.4 0.3

Brookian Sequence

0.2 0.1

West Chukotka (n=285) Wrangel Island (n=371) Lisburne Pen. (n=187) Foredeep (n= 289) Wedgetop (n=195) Low allocs. (n=291) High allocs. (n=482) Unit Jw (n=21)

0 0

500

1000

Geosphere, February 2015

1500

Age (Ma)

2000

2500

3000

109

Moore et al.

A

20

Quartzite clast sample 06TM-29B n = 98

412

14 12 10

15

8

10

482

6

1938 1771

4

5 0

B

16

721

2 0

500

1000

1500

2000

2500

3000

0

0

200

Detrital Zircon Ages (Ma)

20

C

180

Granitic clast sample 06TM-29D n = 110

Relative Probability

25

15 10 5 0 100

140

160

180

600

800

1000

D

Mean age = 159.8 ± 0.6 Ma MSWD = 1.2

170

160

150

140 120

400

Detrital Zircon Ages (Ma)

Age (Ma)

30

Number of Grains

Relative Probability

Number of Grains

25

Filters: < 5% uncertainty 359 Ma) ZIRCONS IN SAMPLES FROM LOWER CRETACEOUS BROOKIAN DEPOSITS AND PALEOZOIC AND OLDER DEPOSITS IN THE ARCTIC ALASKA–CHUKOTKA MICROPLATE (AACM) A B C D E F G H I J Brookian deposits (composite samples) 0.65 A—Foredeep, WBR X 0.00 — — 0.00 — — — — 0.65 0.13 0.29 B—Wedgetop, WBR X 0.00 — — — — — 0.13 0.20 0.59 C—Low allochthon, WBR 0.00 X 0.00 0.00 — 0.00 — 0.20 0.08 D—High allochthon, WBR — 0.00 X — 0.00 — — — 0.21 0.49 E—Foredeep, CBR — — — — X — — — 0.29 0.59 F—Wedgetop, CBR 0.00 0.00 — X 0.00 — 0.00 —

— — 0.00 — 0.06 —

Conglomerate clast in Brookian deposits G—08TM-29B —

0.00



0.00

0.08



0.00

X







K

Possible Paleozoic and older source areas in the AACM (composite samples) 0.21 0.75 H—Wrangel Island — — — — — — X — — 0.49 0.75 I—Lisburne Peninsula — — 0.00 — 0.00 — X — 0.00 0.11 J—Seward Peninsula — — — — — — — — — X 0.06 0.11 K—Kuna Formation — — 0.00 — — 0.00 — 0.00 X Note: P = probability values calculated from Kolmogorov-Smirnoff (K-S) statistics using error in cumulative distribution function (CDF) (Guynn and Gehrels, 2010). P values are rounded to the nearest 0.01. Where P is shown as 0.00, 0 < P ≤ 0.05; dashes indicate P = 0.000. Bold type denotes P > 0.05. For Brookian deposits from the western Brooks Range (WBR), pre-Mississippian zircon ages from foredeep deposits are taken from three samples and wedgetop deposits from two samples, whereas zircon ages from low allochthons are taken from three samples and high allochthons from five samples (this study). For Brookian deposits from the central Brooks Range (CBR), zircon ages from foredeep deposits are taken from two samples and wedgetop deposits from three samples (Wartes, 2008). Ages from conglomerate clast are from the Picnic Creek allochthon in the WBR (this study). Zircon ages from possible Paleozoic and older source areas in the autochthonous parts of the Arctic Alaska–Chukotka microplate (AACM) are taken from five samples from Wrangel Island and two samples from the Lisburne Peninsula (Miller et al., 2010); ages from allochthonous parts of the AACM are taken from 13 samples from the Seward Peninsula (Amato et al., 2009) and one sample from the Kuna Formation in the western Brooks Range (Dumoulin et al., 2013). Numbers of analyses (n) for each sample or group of samples by letter designator: A—191; B—151; C—200; D—338; E—187; F—200; G—96; H—506; I—204; J—993; K—83.

Range (Dumoulin et al., 2013), and (3) middle Paleozoic and older rocks from the metamorphic hinterland of the orogen in the Seward Peninsula (Amato et al., 2009). Detrital zircon ages of the Paleozoic rocks sampled on Wrangel Island are representative of parautochthonous basement rocks present across a wide area of Chukotka (Gottlieb et al., 2012), and detrital zircon ages of Mississippian clastic rocks from the Endicott Mountains allochthon are representative of the zircon ages of upper Paleozoic strata in the other allochthons of the western Brooks Range (Moore, 2010). Detrital zircon ages from the Seward Peninsula samples are similar to those in the metamorphic core of the orogen in the southern Brooks Range (Amato et al., 2009). Also included in Figure 12C are zircon data from the quartz-rich sandstone clast in the Okpikruak Formation of the Picnic Creek allochthon (sample 06TM-29B) and the detrital zircon data of Wartes (2008) from the Brookian foredeep and wedgetop deposits in the central Brooks Range. The light-shaded field in Figure 12C shows the distribution of the cumulative curves for Okpikruak and Colville Basin samples. The plot affirms that these samples are enriched in early Paleozoic and Paleoproterozoic zircons and relatively depleted in late Neoproterozoic and Mesoproterozoic zircons. In contrast, zircons from the samples from the Paleozoic and older rocks of the Arctic Alaska–Chukotka microplate define a field (dark shaded in Fig. 12C) that is enriched in Neoproterozoic and Mesoproterozoic zircons and relatively depleted in Paleoproterozoic zircons. The contrast between these fields suggests that Devonian and

older zircons from early Brookian sandstones in the western Brooks Range were not derived from lower Paleozoic and older source areas in the Arctic Alaska–Chukotka microplate. This conclusion is substantiated by K-S statistics for pre-Mississippian zircons in the Okpikruak Formation and Colville Basin and the possible source areas of similar or older age in the Arctic Alaska–Chukotka microplate in Alaska and Chukotka (Table 6), which reveal a high probability of dissimilarity (i.e., P < 0.05) in most Brookian and Arctic Alaska–Chukotka microplate comparisons. Samples 94TM-47 from the Lisburne Peninsula and to a lesser extent 04TM-143 from the Mount Kelly Graywacke have transitional characteristics that suggest they could include some sediment derived from the Arctic Alaska–Chukotka microplate. The quartz-rich sandstone clast from the Picnic

Creek allochthon displays a zircon population that is dissimilar to those of the Paleozoic and older Arctic Alaska–Chukotka microplate samples (Fig. 12C) and is instead more similar to those of the lower Brookian sandstones (e.g., the high-allochthon group in Table 6), suggesting that it might share the same provenance. Detrital zircon data of Wartes (2008) provide an opportunity to examine whether the Devonian and older zircon populations in the western Brooks Range extend into the central Brooks Range, ~500 km to the east. Wartes (2008) reported detrital zircon analyses from two locations in the Nanushuk Formation and three locations in the Fortress Mountain Formation, representing foredeep and wedgetop stratigraphic positions, respectively. Nanushuk deposits in the central Brooks Range are younger than those from the western Brooks Range because

Figure 14 (on following page). Zircon fission-track (ZFT) ages with uncertainties plotted against zircon U-Pb ages for the same zircon grains in representative samples of the Brookian sequence (left) and relative probability plots of ZFT and U-Pb data for zircons with pre-Jurassic U-Pb ages (i.e., >200 Ma) (right). (A) Sample 03CP-25 from the Ipnavik River allochthon; (B) sample 10TM-78D from the Picnic Creek allochthon; (C) sample 04TM-141 from wedgetop deposits of the Colville Basin; (D) sample TM-02 from foredeep deposits of the Colville Basin; (E) composited data from all four samples. In all plots, duration of the Brookian tectonic event shown by green band and Uralian-Taimyr event by orange band. Also shown are the approximate depositional age of the sample (dotted line), minimum U-Pb age of zircons included in the plot (yellow line, >200 Ma), central (mean) age for the ZFT data (dashed black line), and ZFT model ages for two-component mixing models of the ZFT data (solid black lines). ZFT model ages calculated using RadialPlotter (Vermeesch, 2009) (Supplemental Fig. SF3 [see footnote 1]). Note that relative probability curves for ZFT data are smoother than those for U-Pb data because of the higher uncertainties associated with ZFT data.

Geosphere, February 2015

111

ce an ord nc Co

ges

164±11 213±17

Brookian orogenic cooling

500 400 300

Uralian-Taimyr orogenic cooling Brookian orogenic cooling

Co

600 500

Sample 04TM-141 n = 22

Sample 04TM-141 Hauterivian lower Brookian deposits Wedgetop deposits n = 30

300

ZFT a

400 Uralian-Taimyr orogenic cooling

200 Brookian orogenic cooling

100

196±21

nc

ord

an

ce

C

700

127±13

0 800

es

U-Pb ages

Only grains with U-Pb ages > 200 Ma plotted. 11 U-Pb ages plot offscale because >500 Ma

ges

100

ag

9 U-Pb ages plot offscale because >500 Ma

190±16

200

Pb

Only grains with U-Pb ages > 200 Ma plotted.

300±28

an ord nc Co

600

Sample 10TM-78D n = 26

Sample 10TM-78D Berriasian(?) Picnic Ck. alloc. High allochthons n = 30

155±10

ce

B

700

es

0 800

ZFT ag

100

U-

288±17

Uralian-Taimyr orogenic cooling

200

290±29

300

U-Pb ages

ZFT a

400

U-

116±12

Uralian-Taimyr orogenic cooling

200 Brookian orogenic cooling

100

an nc

600

s age 500

1000

1500

2500 0

Figure 14.

Geosphere, February 2015

100

200

296±24

153±7 2000

U-Pb Age (Ma)

300

es

0

ag

Brookian orogenic cooling

100

189±8

Pb

ZFT

Uralian-Taimyr orogenic cooling

200

es

30 U-Pb ages plot offscale because >500 Ma

500

300

ag

Only grains with U-Pb ages > 200 Ma plotted.

All data n = 106

400

Pb

400

Relative Probability

All data Brookian Sequence n = 120

ord

E

700

ce

0 800

205±14

300

2 U-Pb ages plot offscale because >500 Ma 162±11

an ord nc Co

500

Relative Probability

600

Only grains with U-Pb ages > 200 Ma plotted.

Sample TM-02 n = 28

Sample TM-02 Aptian(?) Nanushuk Formation Foredeep deposits n = 30

ce

D

700

ges

0 800

0

112

6 U-Pb ages plot offscale because >500 Ma

ZF Ta

400

Co

Zircon Fission-track Age (Ma)

500

Relative Probability

Zircon Fission-track Age (Ma)

600

Only grains with U-Pb ages > 200 Ma plotted.

Sample 03CP-25 n = 30

Sample 03CP-25 Berriasian(?) Ipnavik River. alloc. High allochthons n = 30

Relative Probability

Zircon Fission-track Age (Ma)

A

700

U-

Zircon Fission-track Age (Ma)

800

Relative Probability

Zircon Fission-track Age (Ma)

Moore et al.

500

Evolution of lower Brookian foreland basin deposits, northern Alaska deposition in the Nanushuk was strongly progradational from west to east (Fig. 5). The Fortress Mountain in the central Brooks Range is clearly stratigraphically older than the Nanushuk in the same area (Wartes, 2008), but the ages of these deposits relative to the lower Brookian samples in the western Brooks Range are uncertain. A composite cumulative probability curve for the Fortress Mountain Formation (wedgetop) samples from the central Brooks Range (Fig. 12B) shows that they are quite similar to the Brookian samples from the western Brooks Range, whereas the composite curve for the Nanushuk Formation foredeep samples in the central Brooks Range is less segmented and quite different from the western Brooks Range samples due to differences in the relative abundances of their Paleozoic and Proterozoic zircon populations. The cumulative probability curves for Devonian and older zircons (Fig. 12C) show that the Fortress Mountain Formation (wedgetop) samples from the central Brooks Range plot near the average of the Brookian sandstones from the western Brooks Range. This suggests that the Fortress Mountain Formation of the central Brooks Range had the same provenance as the Brookian deposits of the western Brooks Range or could have been recycled from the erosion of older Brookian deposits (i.e., the Okpikruak Formation). The Nanushuk Formation foredeep deposits of the central Brooks Range, in contrast, display a steep slope that plots in the field of the Arctic Alaska–Chukotka microplate samples in Figure 12C. This suggests that Nanushuk deposits in the central Brooks Range may have been derived from erosion of the Paleozoic and older rocks of the Arctic Alaska–Chukotka microplate. These interpretations are supported by K-S statistics. Table 6 compares pre-Mississippian zircons in the samples of Wartes (2008) from the central Brooks Range with samples from various parts of the Arctic Alaska–Chukotka microplate as well as the lower Brookian deposits of the western Brooks Range and to samples from possible source areas in the Arctic Alaska–Chukotka microplate. The K-S statistics suggest similarities between the foredeep samples (i.e., Nanushuk Formation) of the central Brooks Range and possible source areas in Arctic Alaska–Chukotka microplate basement and Paleozoic cover rocks in Wrangel Island, the Lisburne Peninsula, and in the thrust sheets of the Brooks Range (P = 0.21, 0.49, and 0.6, respectively) but are consistently dissimilar to Devonian and older zircons in the lower Brookian deposits of the western Brooks Range (P < 0.05). The pre-Mississippian zircons from the wedgetop deposits (Fortress Mountain For-

mation) in the central Brooks Range, in contrast, are dissimilar to any of the possible source localities in the basement and Paleozoic cover of the Arctic Alaska–Chukotka microplate (P < 0.05). Instead, they have similarities to the preMississippian zircons in the Okpikruak Formation of the lower allochthons (P = 0.59) and wedgetop deposits (P = 0.29) of the western Brooks Range. The foredeep samples of Wartes (2008) from the central Brooks Range were collected from proximal parts of the Umiat Delta, which is highly enriched in phyllite and schist fragments (Bartsch-Winkler and Huffman, 1988). In contrast, Bartsch-Winkler and Huffman (1988) noted that sandstones of the Corwin Delta contain few of these grain types and are instead sedimentary litharenites. Till (1992) showed that sandstones having metamorphic lithic compositions similar to those from the Umiat Delta also yield detrital glaucophane in heavy-mineral separates, whereas glaucophane was not reported in heavy-mineral separates from Corwin Delta sandstones. She suggested that the first appearance of detrital glaucophane in Brookian sandstones marks the time of onset of erosion of the core of the Brooks Range and that this change in heavy-mineral compositions occurred during the mid-Albian. Schenk et al. (2012) estimated from seismic data that deposition of strata in the central Colville Basin near the location of the Umiat Delta occurred at ca. 110–105 Ma (late early to early late Albian; Ogg and Ogg, 2008), whereas two zircons in the samples dated by Wartes (2008) yield ages of 90–89 Ma (Turonian), suggesting that the change in provenance may have occurred as late as Late Cretaceous time. These observations suggest that the Arctic Alaska– Chukotka microplate did not begin contributing detritus into Brookian sandstones before the mid-Albian. This interpretation is supported by Ar-Ar cooling ages from metamorphic rocks in the presentday southern Brooks Range, which indicate that, following high-pressure metamorphism at depths of 10–40 km in the Early Cretaceous, these rocks passed through closure temperatures for hornblende (~500 °C) at 105–103 Ma and white mica (350–420 °C) at 100–90 Ma (Vogl et al., 2002). These cooling ages indicate that the onset of uplift-related cooling in the metamorphic core of the range approximately coincided with the Albian time of appearance of zircons of Arctic Alaska–Chukotka microplate affinity in the Colville Basin foredeep. Cooling in the metamorphic rocks of the southern Brooks Range at this time has been ascribed to postcollisional extensional exhumation (e.g., Miller and Hudson, 1991; Vogl et al., 2002).

Geosphere, February 2015

PROVENANCE OF EARLY BROOKIAN SANDSTONES The abundance of Jurassic zircons in the lower Brookian strata and ophiolitic detritus (clinopyroxene, epidote, serpentinite, and amphibole) in the higher allochthons (Kelly River allochthon and Picnic Creek allochthon) strongly support the interpretation that the lower Brookian strata were at least partially sourced from the Angayucham terrane. Dikes and stocks of plagiogranite and biotite tonalite are locally present in the ophiolitic sequences in the Brooks Range, and similar rock types could have been the source for the boulder of 160 Ma granodiorite (sample 06TM-29D) in the Picnic Creek allochthon. Ophiolitic and arc strata of similar age reportedly form klippen at structurally high levels in eastern Chukotka (Vel’may terrane) and are present along the South Anyui zone (Fig. 1) at the southern margin of Chukotka (Sokolov, 2010), providing evidence that thrust sheets of arc-related Jurassic oceanic rocks may have once extended across much or all of the Arctic Alaska–Chukotka microplate. Provenance of the pre-Jurassic zircons in the lower Brookian sandstones is more difficult to determine. In addition to the detrital zircons in these sandstones, there are several other attributes of the lower Brookian deposits that are relevant to understanding the provenance of this detritus. These are: (1) lower Brookian deposits contain sparse olistostromes and conglomeratic intervals that include chert, carbonate, mafic igneous, granitic, quartz-bearing schist, and quartzite clasts ranging up to hill size (e.g., Chapman and Sable, 1960; Mull, 1985) and that point to nearby source areas; (2) wedgetop deposits in the central Brooks Range region contain locally abundant tonalite and silicic volcanic cobbles that yield U-Pb ages of 274–253 Ma and suggest a proximal source of Permian magmatic rocks (Wartes et al., 2006); (3) a sandstone from the Okpikruak Formation in the central Brooks Range contains detrital white mica that yields Pennsylvanian 40Ar/39Ar ages (Toro et al. 1998) and crossite that is likely to have been derived from a non–Arctic Alaska–Chukotka microplate source area (Till, 1992); and (4) the quartz-rich sandstone clast from a debris-flow deposit in the Picnic Creek allochthon (sample 06TM29B), which suggests that many Devonian and older zircons in the Brookian sequence may have been derived from recycled middle Paleozoic deposits having a different provenance than those of the Arctic Alaska–Chukotka microplate. Finally, the ZFT thermochronology reported here coincides with the times of both Uralian-Taimyr and Brookian deformation.

113

Moore et al.

Triassic Turbidite Sequence of the Russian Far East A sequence of fine-grained, thin-bedded Triassic turbidites of Triassic age was deposited across Carboniferous and Permian carbonate platform sequences in the Arctic Alaska–Chukotka microplate in the Russian Far East (Sokolov et al., 2002; Tuchkova et al., 2009; Sokolov, 2010) (Fig. 1). These deposits have low sandstoneshale ratios and consist of prodelta sediments, distal turbidites, and contourites (Tuchkova et al., 2009). Sedimentation in the Early Triassic was accompanied by intrusions of mafic dikes and sills that suggest extensional tectonics (Miller et al., 2006) followed later in the Triassic by deposition of passive margin strata (Sokolov et al., 2009). Sandstones in the sequence are micaceous, lithic-rich, feldspathic quartz arenites that were shed from metamorphic and volcanic source areas (Tuchkova et al., 2009). These rocks are generally weakly metamorphosed, tightly folded, and record as many as three periods of deformation during the Jurassic and Early Cretaceous (Sokolov et al., 2009). Paleoflow is thought to be from the north (Sokolov, 2010), although some uncertainty exists due to the complex deformation (S.D. Sokolov, 2013, written commun.). Reported thicknesses for the turbidite sequence range from 0.8 to 1.5 km, and structural thicknesses of up to 5 km are estimated for the unit in central Chukotka (Kos’ko et al., 1993; Miller et al., 2006, 2010; Tuchkova et al., 2009). Detrital zircon U-Pb analyses from sandstones of the Triassic turbidite unit in western Chukotka, Wrangel Island, and a thin unit of compositionally similar shelf sandstones in the Lisburne Peninsula (Moore et al., 2002) reveal generally similar distributions of zircon ages that feature significant populations at 350–230 Ma, 500–400 Ma, and a smaller, variable population at 2.0–1.7 Ga (Miller et al., 2006, 2010). On the basis of these ages, Miller et al. (2006) suggested that the source area for the Triassic turbidites was in the Ural Mountains, Taimyr, and the Siberian traps volcanic field and that the sediment was transported through fluvial drainages into deep-marine depocenters along the southern margin of the Arctic Alaska–Chukotka microplate prior to opening of the Arctic Basin.

114

Figure 12D shows a cumulative probability plot that compares the ages of Triassic and older zircons from the lower Brookian sandstones of the western Brooks Range to those in the Triassic turbidites in western Chukotka, Wrangel Island, and the Lisburne Peninsula. The diagram shows that ages of the Triassic and older zircons from lower Brookian sandstones are quite similar to those in the Triassic turbidites, varying mostly in the relative proportions of Phanerozoic and Paleoproterozoic zircons. The detrital zircon ages from the Okpikruak Formation in the Brooks Range are particularly similar to those in the Triassic turbidites from Wrangel Island. Although representing similar zircon populations, the sandstones from foredeep and wedgetop deposits in the western Colville Basin are richer in Paleoproterozoic zircons than the Triassic turbidites, indicating that these sandstones have a similar but somewhat different provenance. Table 5 presents K-S statistics for zircon populations from Triassic turbidites in Chukotka, Wrangel Island, and the Lisburne Peninsula compared to our data from Triassic and older zircons from the Colville Basin and Okpikruak Formation in the western Brooks Range. For all Triassic deposits, the comparison yields values of P > 0.05 for 49% of the comparisons against the Brookian samples. For just the Wrangel Island data, P > 0.05 for 62% of the comparisons with the Brookian samples. (Table 5). These

results support the interpretation that many or most zircons in the lower Brookian deposits could have been derived from the Triassic turbidite unit, especially in the Wrangel Island area. If the Brookian deposits in the Brooks Range were sourced from the Triassic turbidite unit and that unit is composed of detritus shed from earlier erosion of the Uralian-Taimyr orogenic belt as proposed by Miller et al. (2006) and supported by our ZFT thermochronologic data, then it is not possible that the dominant, Jurassic and Early Cretaceous, episode of cooling occurred in the Uralian-Taimyr source area. Instead, detritus from the Uralian-Taimyr source area must have cooled and been uplifted by the Early Triassic when it was eroded and deposited into the Triassic turbidite basin. Following deposition, a second period of cooling occurred during the Late Jurassic and Early Cretaceous Brookian orogeny. This would indicate that the late Paleozoic and Triassic zircon grains in the Brookian deposits of the western Brooks Range are second-cycle sedimentary detritus (Fig. 15). Accreted Continental, Arc, and Oceanic Terranes of the Angayucham and South Anyui Oceans The record of the closure of the South Anyui and Angayucham oceans is preserved in the South Anyui zone in the Russian Far East,

A

B

C

Erosion of Uralian and Taimyr orogenic belts and Siberian traps Tr

South Anyui ocean basin

Brookian deformation and erosion in northern Alaska and Chukotka

Triassic and older zircons with partially to fully annealed ZFT ages

Pz

PAZ-Zr

Early Brookian deposition in northern Alaska Jurassic zircons

Angayucham terrane

Chukotka fold belt

Triassic turbidites

PAZ-Zr Prot. Collisional belt

D

Remnant ocean basin

zircons ssic ra Ju Unannealed p to fullyannealed ZFT ages

Colville Basin and Okpikruak Fm.

re -

Rocks in three source areas could explain some or all of these observations. These source areas include: (1) the Triassic turbidite sequence that rests on the Arctic Alaska–Chukotka microplate in Chukotka, (2) the accreted continental, arc, and oceanic terranes of the South AnyuiAngayucham oceans, and (3) the synorogenic deposits of western Chukotka. Evidence for and against these scenarios is discussed below.

Arc-continent collisional zone

Early Brookian foreland basins

Figure 15. Two-stage model for origin of zircon fission-track (ZFT) ages present in lower Brookian sandstones. (A) Geologic section in the Uralian-Taimyr orogenic belt consisting of Paleoproterozoic to Triassic deposits with a zircon fission-track partial annealing zone (PAZ-Zr) in late Neoproterozoic and/or lower Paleozoic strata. (B) Erosion of section in A supplies fully annealed Proterozoic zircons and partially annealed to unannealed Paleozoic to Triassic zircons to the Triassic turbidite basin in the South Anyui Ocean. (C) Turbidite section in B is thrust onto margin of Chukotka during the Late Jurassic and Early Cretaceous Brookian orogenic event and new partial annealing zone established in the resulting thickened structural section. (D) Erosion of structural culmination in C in early Brookian time supplied Jurassic zircons from structurally high thrust sheets of the Angayucham terrane and pre-Jurassic zircons, including fully annealed Proterozoic zircons and fully annealed to unannealed Paleozoic and Mesozoic zircons, into Brookian foreland basin deposits.

Geosphere, February 2015

Evolution of lower Brookian foreland basin deposits, northern Alaska where previously assembled late Paleozoic and Triassic ophiolite and volcanic arc and continental terranes of Siberian affinity (e.g., the Omolon terrane) form the basement for a Late Jurassic and Early Cretaceous arc (the composite Alazeya-Oloy arc terrane) that was thrust northward onto the Arctic Alaska–Chukotka microplate during the Late Jurassic and earliest Cretaceous (Sokolov, 2010). The thrust sheets of the upper plate of the orogen hold a variety of late Paleozoic and early Mesozoic igneous rocks that could have provided the source for the zircon populations in the Brookian deposits. The early Paleozoic and older zircons in the Brookian deposits may have been derived from erosion of the continental terranes that form the basement of some of the arc terranes, whereas ophiolitic and volcanic clastic detritus in the sandstones and conglomerates in the Picnic Creek allochthon, Kelly River allochthon, and Ipnavik River allochthon could have been derived from arc and forearc terranes in the upper plate. If the primary provenance for lower Brookian deposits lay entirely in the upper plate rocks, however, then the composition of the resulting Brookian deposits would be expected to vary spatially and temporally depending on the volume and extent of the specific terranes that were available to erosion. Instead, the generally constant and broad distributions of sandstone composition and detrital zircons suggest recycling of a largely sedimentary provenance area. In addition, a continental terrane large enough to contribute most or all of the Paleozoic and Precambrian zircons with the observed mix of ages has not been recognized in the orogenic belt. Finally, it seems unlikely that terranes in the upper plate of the Brookian orogen could have undergone sufficient burial during the Jurassic and Early Cretaceous to set ZFT ages because of their high structural positions in the orogen. Synorogenic Deposits of Western Chukotka Upper Jurassic and Lower Cretaceous synorogenic strata rest unconformably on the finegrained Triassic turbidite deposits throughout Chukotka. The Jurassic and Cretaceous deposits consist of thin- to thick-bedded turbidites, tuffaceous turbidites, polymictic conglomerates, and olistostromes that were shed northward from the South Anyui collisional zone along the southern margin of Chukotka (Sokolov et al., 2002, 2009; Bondarenko et al., 2003). Their lithologic character, age, and involvement in thrusting suggest that these deposits are correlative with the allochthonous lower Brookian predecessor foreland basin deposits of the western Brooks Range (i.e., the Okpikruak Formation).

Although broadly similar to the major zircon age populations of the lower Brookian deposits, the detrital zircon ages from syntectonic deposits in western Chukotka differ in that they are dominated by Paleoproterozoic zircons and contain a much lower proportion of lower Paleozoic, Neoproterozoic, and Mesoproterozoic zircons. As a result, the western Chukotka deposits contain a distinctly different age distribution (Fig. 12B) that suggests a source terrane enriched in Precambrian crystalline rocks. These data support the interpretation that the Jurassic and Early Cretaceous deformation of western Chukotka may have involved parts of the continental margin of Siberia or blocks that had previously rifted away from the Siberian margin such as the Omolon terrane (Miller et al., 2009). The large difference between the detrital zircon populations of the Triassic turbidite unit and the covering Upper Jurassic and pre-Aptian syntectonic strata in western Chukotka (Figs. 12B versus 12D) shows that the Triassic deposits in this area probably were not a primary source area for the overlying synorogenic deposits. For similar reasons, the western Chukotka synorogenic deposits seem unlikely to constitute an up-dip part of the foreland basin depositional system in western Alaska. It remains possible, however, that by the Aptian and Albian, some sediment from western Chukotka may have been funneled eastward down the foreland basin into the Colville Basin. If present as a small fraction, the addition of Siberian affinity zircons could have increased the proportion of Paleoproterozoic zircons in Colville Basin sandstones without otherwise altering the detrital zircon age distribution in the sediment. DISCUSSION Compositional and detrital zircon data from the lower Brookian deposits of the western Brooks Range document changes in the provenance of the sediment shed from the Brookian orogen. The oldest Brookian deposits are the Upper Jurassic thin-bedded volcaniclastic turbidites of unit Jw, which were deposited on volcanic rocks of either the Copter Peak allochthon or the Misheguk Mountain allochthon, which compose the Angayucham terrane. Unit Jw deposits occupy magmatic arc fields on sandstone compositional diagrams (Figure 7) and contain an almost unimodal population of Late Jurassic zircons that together point to derivation from the unroofing of a Late Jurassic volcanic arc. A small proportion of early Paleozoic, Neoproterozoic, and Paleoproterozoic zircons suggest that the arc may have formed on continental rocks or was close to a source of continental sediment. This suggests that unit Jw originated

Geosphere, February 2015

as a forearc basin succession that was deposited either directly on ophiolite basement (if underlain by the Misheguk Mountain allochthon) or on the hindward part of the subduction complex (if underlain by the Copter Peak allochthon). Although unit Jw strata are included in the Brookian sequence because of their lithic composition and derivation from a southern source area, our new data suggest that the unit is best viewed as a lithologic component of the Angayucham terrane. Brookian deposits from allochthons of the Arctic Alaska–Chukotka microplate in the western Brooks Range have sandstone compositions that range from magmatic arc to recycled orogen composition (Fig. 7), probably reflecting a decline over time of the amount of detritus derived from the Angayucham and/or Koyukuk terranes relative to that derived from the Triassic turbidite unit that overlies the Arctic Alaska– Chukotka microplate in Chukotka. Relatively small contributions of sediment from continental sources rich in zircon such as the Triassic turbidite unit, however, can have outsized importance in detrital zircon populations relative to those of oceanic and volcanic arc sources. The quartz-rich clasts (e.g., sample 06TM-29B) in conglomerates and olistostromes otherwise dominated by mafic and intermediate clasts suggest that continental rocks might have been present in the basement of the Angayucham and/or Koyukuk terrane. Alternatively, the clasts might have been derived from coarse-grained channel deposits in the Triassic turbidite unit that are no longer present because of subsequent erosion. Brookian deposits of the low allochthons contain locally numerous olistostromes that contain large blocks and clasts of Permian and Triassic chert from the Etivluk Group and limestone from the Lisburne Group of the Arctic Alaska–Chukotka microplate (Martin, 1970; De Vera, 2005). Significant populations of zircons with ages indicating erosion of Endicott Group strata during the Early Cretaceous are not present in our samples. This suggests that active thrusts during this time were rooted in the shalerich detachments that overlie the clastic rocks of the Endicott Group (e.g., Wallace et al., 1997) (Fig. 4). The provenance of the wedgetop deposits in the western Colville Basin (i.e., the Mount Kelly Graywacke) is uncertain. Although their composition is rich in detrital carbonate and subordinate metamorphic grains and white mica that seems to point to a source area in the metamorphic rocks of the southern Brooks Range, the detrital zircon U-Pb ages suggest these deposits were more likely derived principally from ancestral foreland basin deposits and the carbonate rocks of the Lisburne Group in the allochthons

115

Moore et al. of the Arctic Alaska–Chukotka microplate, most likely the Kelly River allochthon. The lower Brookian sandstones from foredeep deposits of the Colville Basin in the western Brooks Range contain Paleoproterozoic zircons that are generally more abundant than those in the Triassic turbidite unit of central Chukotka (Fig. 12D). This suggests that a secondary source area enriched in Paleoproterozoic zircons may have contributed sediment to the Colville Basin. Longitudinal transport of the late synorogenic sediments from western Chukotka, where sandstone with higher abundances of Paleoproterozoic zircons are present (Miller et al., 2009), could explain this observation. Paleoproterozoic zircon populations, however, are also locally abundant in the Triassic turbidite unit from Wrangel Island (e.g., sample ELM06WR518 of Miller et al., 2010), suggesting that Brookian samples with abundant Paleoproterozoic grains may simply reflect compositional variations within the Triassic turbidite unit source area. We prefer this interpretation because the introduction of significant amounts of sediment derived from Siberian sources probably also would have increased the proportion of K-feldspar in the Colville Basin deposits, an increase not supported by our sandstone compositional data. Following progradation of the Corwin Delta eastward along the axis of the Colville Basin into the central North Slope, the Umiat Delta became the major input point for sediment into the Colville Basin foredeep. The sediment of the Umiat Delta reflects a major change to a source in the metamorphic hinterland of the Brooks Range in the Arctic Alaska–Chukotka microplate. It seems likely that a similar change of provenance may have occurred in the western Brooks Range as well, but its stratigraphic record may be missing due to subsequent erosion across the western North Slope (Houseknecht et al., 2011). If arrival of sediment from the hinterland of the Arctic Alaska–Chukotka microplate was synchronous throughout the foreland region, it would indicate denudation was caused by regional tectonic changes. Alternatively, the arrival of sediment from the hinterland of the Arctic Alaska–Chukotka microplate may have varied along strike due to varying amounts of erosion along the orogenic belt, or because it was sequestered in intermontane basins until erosion provided sedimentary pathways into the foreland basin in the Albian. New data presented here support the earlier interpretation that most lower Brookian sediment in the western Brooks Range was sourced from areas underlain by the Triassic turbidite unit in the Herald Arch, eastern Chukotka, and present-day Chukchi Sea (Fig. 1). Despite hav-

116

ing comparable zircon populations, the presentday thickness and distribution of the Triassic turbidite unit in Chukotka seem inadequate to explain the voluminous lower Brookian deposits in northern Alaska. This suggests that substantial thicknesses of the Triassic strata must have once been present but subsequently were removed by erosion in the source area. This is supported by the deformation and sub–greenschist-facies metamorphism of the Triassic turbidite unit and the ZFT data that indicate burial to depths of as much as 7–10 km, providing evidence for a highland region that consisted of structurally thickened parts of the same unit. The highlands probably extended southward from the HeraldWrangel thrust across eastern Chukotka to the South Anyui zone and perhaps southward to the Seward Peninsula, although subsequent extensional deformation during the mid-Cretaceous in the Brooks Range and in the Tertiary in the Hope Basin (Tolson, 1987; Dumitru et al., 1995; Klemperer et al., 2002; Miller and Akinin, 2008) may have substantially broadened the modern north-south extent of the former highland. Based on relations in the South Anyui zone along the southwestern margin of Chukotka, Sokolov et al. (2009) reported that the Triassic turbidite unit occupies a structural position above the parautochthonous basement and platform sedimentary cover of the Arctic Alaska–Chukotka microplate and below the upper Paleozoic and lower Mesozoic ophiolitic, island-arc, and sedimentary rocks of the Russian Far East. This stacking succession suggests that the Triassic turbidite unit occupies a structural level that is equivalent to that of the allochthonous sedimentary cover sequences of Arctic Alaska–Chukotka microplate in the Brooks Range (i.e., the Endicott Mountains allochthon, Picnic Creek allochthon, Kelly River allochthon, and Ipnavik River allochthon), with the arc and ophiolite thrust sheets of the South Anyui terrane being correlative with Angayucham terrane (i.e., the Copter Peak allochthon and Misheguk Mountain allochthon) in Alaska. The Alazeya-Oloi terrane, for example, is thought to comprise remnants of the late Paleozoic to Early Cretaceous arc and subduction complex much like the Koyukuk arc terrane, Misheguk Mountain allochthon, and Copter Peak allochthon represent in Alaska, with both arc-forearc systems having been thrust onto the Arctic Alaska–Chukotka microplate during the Brookian structural event (Sokolov, 2010). The lateral change in facies from starved distal shelf deposits in the Brooks Range (Fig. 4) to a deep-water sedimentary prism composed of Triassic thin-bedded prodelta turbidites in the Chukchi Sea and Chukotka may have influenced the formation of the highland area that

Geosphere, February 2015

was the main source of lower Brookian sediments. Because the turbidite succession in the Chukchi Sea and Chukotka was significantly thicker than the distal shelf deposits in Alaska, equal amounts of shortening applied to both areas would have resulted in a significantly thicker structural stack in Chukotka than in Alaska (Fig. 16). Erosion of the resulting culmination would have begun in the structurally higher oceanic rocks followed by increasing amounts of erosion of the underlying turbidite unit as the culmination grew and was dissected. The deep-marine sedimentary environment of the Triassic turbidite unit, coupled with mafic dikes and sills in the older part of the sequence and the deposition of the turbidite section on the older extensionally deformed platform deposits and Neoproterozoic crystalline basement of the Arctic Alaska–Chukotka microplate, suggest that the Triassic turbidite unit was deposited on tectonically thinned continental crust along the rifted southern margin of the Arctic Alaska– Chukotka microplate. One interpretation is that the turbidite unit formed part of a large deltaic, prodeltaic, and basinal turbidite-fan complex that emanated from a fluvial distribution system shed southward across the Arctic region prior to formation of the Jurassic and Cretaceous Amerasia and Cenozoic Eurasia ocean basins (Miller et al., 2013). This interpretation is also consistent with south-directed paleocurrent flow of the turbidites. The Arctic restorations of Miller et al. (2006, 2010) and Amato et al. (2009) placed the Uralian and Taimyr orogens northeast of the Arctic Alaska–Chukotka microplate (in presentday coordinates) in the Triassic (Fig. 17A). An alternative interpretation favored here is that the South Anyui–Angayucham oceans included the northern and eastern part of the Uralian Ocean, which separated Baltica and Siberia in the late Paleozoic but was closed by the Uralian-Taimyr orogenic event in the Late Permian and Triassic (e.g., Sokolov et al., 2009; Miller et al., 2011, their fig. 4). We suggest that by the middle Mesozoic, the South Anyui Ocean was a remnant ocean basin bordered by highly extended and transitional crust of the Arctic Alaska–Chukotka microplate on the northeast and an active margin of late Paleozoic and early Mesozoic island arcs along the margin of Siberia on the southwest (Fig. 17B). Reconstructions of the Uralian orogen indicate that this orogen closed from south to north during the Carboniferous to the Triassic, with the youngest part of the closure having taken place in Taimyr in the Triassic (Torsvik and Andersen, 2002). Sediment, perhaps including coarse conglomeratic material, was shed from the orogen westward into the Uralian-Taimyr foreland basin and then down the basin axis into deltaic and prodeltaic

Evolution of lower Brookian foreland basin deposits, northern Alaska

A S

N S

Km 0

Brooks Range allochthons

Sea level

Angayucham Ocean

MMA MMA

Colville Basin

CPA

CPA

Arctic-Alaska

Arctic-Alaska

Late Jurassic

30

B S Km 0

Early Cretaceous

Coarse-grained channel facies a South Anyui Ocean

N S c

South Anyui suture

Structural culmination b

Sea level

Chukotka

30

N

Sea level

Wrangel Is. d

N

Chukotka

Early Cretaceous

Late Jurassic Explanation Arcs & ophiolites (MMA)

Continental crust

Triassic turbidite deposits

Basalt-chert accretionary assemblage (CPA) Ocean crust

Mantle

Carboniferous & Permian clastic & carbonate rocks

Foreland basin deposits Ellesmerian Sequence

Figure 16. Schematic tectonic models for Brookian north-vergent, arc-continent collision in (A) Arctic Alaska and (B) Chukotka. Diagrams at left show relations in Late Jurassic; diagrams at right in Early Cretaceous. In Arctic Alaska, Devonian to Triassic Ellesmerian sequence passive-margin deposits are detached from thin continental crust and imbricated into an antiformal stack of allochthons during Early Cretaceous Brookian deformation. In Chukotka, Carboniferous to Permian strata were deposited on low-standing, highly attenuated continental crust and were covered by a thick, fine-grained Triassic clastic sequence (Ustieva unit of Sokolov et al., 2009). The Triassic clastic sequence was detached at its base during Brookian deformation and thrust into a thick, brittle to ductily deformed structural culmination, which became the provenance for Brookian foreland basin deposits. The depositional setting of the Triassic clastic sequence is interpreted here as a remnant basin in the South Anyui Ocean. Following imbrication, channel facies in the axial part of the basin (a) would be emplaced into structurally high positions (b) where clasts could have been eroded and shed into Brookian foreland basin deposits. If, on the other hand, the Triassic sequence represents a deltaic succession that was shed southward from the Chukotka continental margin (c), the coarse-grained channelized facies would be expected to be found in structurally low positions such as at Wrangel Island (d), where they would be the last parts of the sequence to be eroded. CPA— Copter Creek allochthon; MMA—Misheguk Mountain allochthon.

turbidite fans constructed on distal turbidite deposits in the nearby northwestern apex of the South Anyui Ocean basin. Thrusting of the Siberian margin onto the Arctic Alaska–Chukotka microplate margin during closure of the South Anyui Ocean in the Jurassic and Early Cretaceous would have resulted in collapse of the remnant basin and accretion of the Triassic turbidite unit onto the Arctic Alaska–Chukotka microplate beneath the subducting arc systems that existed outboard of the margin of Siberia. This interpretation explains the presence of the Triassic turbidite unit along the length of the Arctic Alaska–Chukotka microplate in Chukotka and is consistent with the models of Sokolov et al. (2002) and Nokleberg et al. (2000)

for the evolution of the South Anyui and Angayucham ocean basins. It also agrees with the models for opening of the Amerasian Basin of Lawver et al. (2002) and Grantz et al. (2011b), which feature counterclockwise rotation of the Arctic Alaska–Chukotka microplate away from the Canadian margin in the Jurassic and Early Cretaceous and restore the western end of the Arctic Alaska–Chukotka microplate against Eurasia in the vicinity of the Kara Sea prior to opening of the basin. This study provides the first direct evidence for a connection between the Alaskan and Russian parts of the Arctic Alaska–Chukotka microplate. Previous correlations suggested such a connection based on similarities in lower

Geosphere, February 2015

Paleozoic stratigraphy, and tectonic history. The depositional transfer of detritus from the Triassic turbidites in Chukotka to the Brookian foreland basin system in Arctic Alaska validates the correlation back to the earliest Cretaceous and confirms that the Late Jurassic and Early Cretaceous deformational events of Arctic Alaska and Chukotka are parts of the same orogenic event. Our study further suggests that foreland basin deposition throughout the early evolution of the Brooks Range was increasingly dominated by longitudinal sediment dispersal systems that emanated from the eastern Chukotka–Herald Arch–Chukchi Sea region (Fig. 18). Although there is clear evidence of contributions of sediment from local, possibly submarine source

117

Moore et al. 60°N

Pac. O.

60°N

Alaska SP

A

ic

Arct

North America

ka Alas

CB

Chu

kotk SAZ OM a Chukchi Sea WI East Siberian ? Sea CM NSB

dg

or

on m Lo

Eurasia

og

Basin

en

I.

Greenland

Barents Shelf

60°W

B

yr

Ellesmere

NP

im Ta

os

ov

MB

Ri

Basin

0° 60°N

Urals 60°E

60°W

Paleo-Pacific Ocean ESS Siberia

SP AA

Wrangel Is.

yr

LR

or og en

CMMR DB?

im Ta

Laurentia

CH

CONCLUSIONS

of U ralia

re I. Ellesme

n fo

C.



60°W

60°N

ls

ep

Greenland

Ura

rede

Barents Shelf

30°W 30°N

Paleo-Pacific Ocean OM Remnant ocean basin

SP

Siberia

AA LR

yr

CH

or

WI CM

im Ta

Laurentia

alia

f Ur

re I.

Ellesme

o Axis

ESS

en og

?

n fo

ls



Triassic turbidite unit of Chukotka Paleoflow direction Pole of AACM rotation

areas in the Brooks Range, the abundance of Paleozoic and Triassic zircons in lower Brookian sandstones extending at least as far east as the central Brooks Range suggests that longitudinal systems may have deposited sediment along the orogen for 750 km or more. In addition, the

Ura

ep

30°W 30°N

Barents Shelf

rede

Greenland

118

Siberia

e

Amerasia

Axis

Figure 17. Paleogeographic models for deposition of the Triassic turbidite unit of Chukotka from likely source areas in the Taimyr and Uralian orogens (subsequent orogenic belts unrestored). (A) Presentday configuration of the Arctic Ocean region (Lawver et al., 2002). (B) Triassic reconstruction of the Arctic region from Amato et al. (2009) showing turbidite unit as continental margin sedimentary prism. (C) Triassic reconstruction of the Arctic region of Lawver et al. (2002) projected onto same base showing the turbidite unit as the sedimentary fill of a remnant basin formed by Uralian orogenic event. Abbreviations: AA—Arctic Alaska; AACM— Arctic Alaska–Chukotka microplate; CH—Chukotka; CM— Chukchi microplate; ESS—East Siberian Sea; LR—Lomonosov Ridge; MR—Mendelev Ridge; NP—North Pole; NSB—New Siberian Islands; OM—Omolon terrane; SP—Seward Peninsula; WI—Wrangel Island; DB?—crust of uncertain affinity. Lawver et al. (2002) interpreted MR and DB as Cretaceous oceanic units and thus, they are not shown in model C. Scale of A differs from that of B and C.

the foreland. As more sediment accumulated and became involved in the thrusting, larger volumes of syntectonic sediment were contributed into the foreland basin by recycling, thus reducing the proportion of the sediment contributed from preorogenic rocks. This process drove sandstone compositions to more sedimentary lithic compositions and to detrital zircon age distributions like those of the primary source area in Chukotka as the foreland basin evolved. The dominance of longitudinal sediment dispersal systems and local syntectonic recycling did not change until a major transverse sediment dispersal system, the Umiat Delta, developed from a source area in the metamorphic hinterland of the orogen in the Albian. This system shifted the sandstone compositions to metamorphic lithic compositions and to detrital zircon age distributions of the Paleozoic and older rocks of the Arctic Alaska– Chukotka microplate. The change to a transverse sediment dispersal system may mark a major reorganization in the orogenic belt caused by the shift from collisional to extensional tectonics in the Brooks Range in the Albian.

Continent-ocean transition Boundaries between major tectonic elements Thrust front of Brooks Range Study area

abundance of these deposits at distal positions suggests that early deposited foreland basin strata were structurally incorporated into the thrust belt where they became important contributors of recycled sediment for younger foreland basin deposits as the thrust system propagated toward

Geosphere, February 2015

(1) The latest Jurassic to late Early Cretaceous western Brooks Range is a thin-skinned fold-and-thrust belt with a well-developed foreland basin and remnants of ancestral foreland basin deposits involved as syntectonic deposits at various structural levels. (2) Sandstones from the highest structural level (unit Jw) consist almost entirely of volcanic arc detritus with a near unimodal population of detrital zircons having ages of 150– 170 Ma (Late and Middle Jurassic). These deposits are interpreted as forearc basin deposits in the Angayucham terrane, which in the western Brooks Range were deposited on ophiolite and/or subduction complex rocks that form the upper plate of the Brooks Range orogen. (3) Detritus and zircons shed from local sources in the Angayucham terrane are abundant in sandstones, conglomerates, and olistostromes in synorogenic sandstones in the immediately underlying thrust sheets, (i.e., the Ipnavik River, Kelly River, and Picnic Creek allochthons) but are less common in the later-emplaced and now lower thrust sheets of the Endicott Mountains allochthon and in the Colville Basin. (4) Quartz, chert, clastic sedimentary, carbonate, and to a lesser extent metamorphic lithic grains are common in synorogenic sandstones of the high allochthons and dominate the lower Brookian sandstones of the Endicott Mountains allochthon and Colville Basin. These sandstones yield detrital zircon ages that feature large populations at 359–200 Ma (Triassic and

Evolution of lower Brookian foreland basin deposits, northern Alaska

Structural culmination composed of Triassic turbidite unit

Eastern Chukotka Wrangel Island

South Anyui zone

Angayucham terrane

Herald arch

Arctic Alaska W

S

Explanation Endicott Gp. (Dev.-Miss.) Basement (pre-Miss.)

Angayucham terrane (Dev.-Jur.) Platform sequence (Miss.-Jur.)

Turbidite unit (Triassic) Okpikruak Fm. (Early Cret.)

N E Shelf foredeep deposits Slope & basin foredeep deposits

Olistostrome Alluvial & delta-plain foredeep deposits Fan-delta wedgetop deposits

Foredeep axis

Figure 18. Block diagram showing conceptual model of deposition for lower Brookian foreland basin deposits in the western Brooks Range. Deposition is dominated by longitudinal eastward sedimentary transport from a major highland centered in Chukotka and the present-day Chukchi Sea and short-headed transport mainly in the form of olistostromes and debris-flow deposits from the Brooks Range thrust belt to the south. Local fan-delta sedimentary systems also contributed sediment to the foreland basin by Aptian time, but their role in earlier Cretaceous time is uncertain. The apparent absence of such deposits in Valanginian and older foreland basin deposits suggests that the Brooks Range thrust belt may have been mostly submarine in earliest Cretaceous time.

late Paleozoic), subordinate populations at 542– 359 (early Paleozoic), and lesser populations at 2.0–1.75 Ga (Paleoproterozoic). (5) Although source areas for lower Brookian deposits have often been assumed to lie in the Paleozoic and older basement rocks of the metamorphic hinterland of the Brookian orogen, the Arctic Alaska-Chukotka microplate in Alaska contains virtually no rocks that could provide the Triassic and late Paleozoic zircons found in the lower Brookian deposits of the western Brooks Range. In addition, the Paleozoic and older rocks of the orogen contain distinctly different proportions of early Paleozoic and older zircons than those found in the lower Brookian strata. The first appearance of detritus having zircon age populations characteristic of the Arctic Alaska– Chukotka microplate in the Brookian sequence occurs in the mid-Albian metamorphic lithic sandstones in the Umiat Delta in the central Brooks Range.

(6) The dominant Paleozoic to Triassic zircon populations in the Brookian sandstones are very similar to those in the Triassic turbidite unit in Chukotka and Wrangel Island and, along with facies and paleoflow data from the Colville Basin, support earlier interpretations of a significant highland area located in eastern Chukotka, the Herald Arch, and the present-day southern Chukchi Sea. (7) Zircon fission-track double dating supports the interpretation that the Triassic turbidite unit in Chukotka was derived from the exhumation and erosion of upper Paleozoic and Triassic rocks in the Uralian-Taimyr orogenic belt and provides evidence that these Triassic strata were later buried to depths of as much as 7–10 km and exhumed as part of the Brookian orogenic belt in the Jurassic and Early Cretaceous. (8) The Triassic turbidite unit is proposed to have been deposited in a marginal basin setting in the South Anyui Ocean and thrust onto the

Geosphere, February 2015

Chukotka part of the Arctic Alaska–Chukotka microplate during the Late Jurassic and Early Cretaceous Brookian orogenic event. The turbidite unit was structurally thickened by the deformation, producing a regional culmination in the area of eastern Chukotka, Herald Arch, and southern Chukchi Sea that became the source area for much of the detritus in the foreland basin deposits of the western Brooks Range. (9) The presence of Triassic and Paleozoic zircon populations in lower Brookian deposits at most structural levels and throughout the development of the collisional phase of the Brookian orogen requires that sediment dispersal occurred mainly by axial (longitudinal) flow over distances in excess of 750 km along the Brookian foreland basin. Recycling of sediment due to structural imbrication of ancestral foreland basin deposits with underlying deformed platform deposits at the thrust front produced olistostromes and local sedimentary systems

119

Moore et al. that also contributed sediment to the foreland basin system. (10) The transfer of clastic detritus from source areas in the Triassic turbidite unit of eastern Chukotka, the Herald Arch, and the southern Chukchi Sea into the Brookian foreland basin system in northern Alaska confirms previous correlations of the Brookian orogen from northern Alaska across the Chukchi Sea and into the Russian Far East. ACKNOWLEDGMENTS

This study benefited from discussions with Ken Bird, Dwight Bradley, Julie Dumoulin, Elizabeth Miller, David Houseknecht, Chad Hults, Karen Kelley, Richard Lease, Gil Mull, Alison Till, Marwan Wartes, and Joe Wooden. We thank Margaret Donelick and Jim McMillan for sample preparation and Charles Knaack for technical assistance with LA-ICP-MS data collection procedures. The manuscript was sharpened and significantly improved by the thoughtful comments and advice of U.S. Geological Survey reviewers Kenneth J. Bird and Richard Lease and Geosphere reviewers Jaime Toro and Tim Lawton. Art Grantz, Paul Stone, Sarah Nagorsen, and Mike Diggles kindly read and edited the manuscript. The work was funded by the Energy Resources and Mineral Resources programs of the U.S. Geological Survey. We thank Teck, the owner and operator of the Red Dog Mine, for graciously allowing us to use their facilities for our research. Any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the U.S. Government. REFERENCES CITED Ahlbrandt, T.S., Huffman, A.C., Fox, J.E., and Pasternack, I., 1979, Depositional framework and reservoir-quality studies of selected Nanushuk Group outcrops, North Slope, Alaska, in Ahlbrandt, T.S., ed., Preliminary Geologic, Petrologic, and Paleontologic Results of the Study of Nanushuk Group Rocks, North Slope, Alaska: U.S. Geological Survey Circular 794, p. 14–31. Aleinikoff, J.N., Moore, T.E., Walter, M., and Nokleberg, W.J., 1993, U-Pb ages of zircon, monazite, and sphene from Devonian metagranites and metafelsites, central Brooks Range, Alaska, in Dusel-Bacon, C., and Till, A.B., eds., Geologic Studies in Alaska by the U.S. Geological Survey, 1992: U.S. Geological Survey Bulletin 2068, p. 59–70. Amato, J.M., Toro, J., Miller, E.L., Gehrels, G.E., Farmer, G.L., Gottlieb, E.S., and Till, A.B., 2009, Late Proterozoic–Paleozoic evolution of the Arctic Alaska–Chukotka terrane based on U-Pb igneous and detrital zircon ages: Implications for Neoproterozoic paleogeographic reconstructions: Geological Society of America Bulletin, v. 121, p. 1219–1235, doi:10.1130/B26510.1. Bartsch-Winkler, S., and Huffman, A.C., Jr., 1988, Sandstone petrography of the Nanushuk Group and Torok Formation, in Gryc, G., ed., Geology and Exploration of the National Petroleum Reserve in Alaska, 1974 to 1982: U.S. Geological Survey Professional Paper 1399, p. 801–831. Bering Strait Geologic Field Party, 1997, Koolen metamorphic complex, NE Russia: Implications for the tectonic evolution of the Bering Strait region: Tectonics, v. 16, p. 713–729, doi:10.1029/97TC01170. Bernet, M., 2009, A field-based estimate of zircon fissiontrack closure temperature: Chemical Geology, v. 259, p. 181–189, doi:10.1016/j.chemgeo.2008.10.043. Bird, K.J., and Andrews, J., 1979, Subsurface studies of the Nanushuk Group, North Slope, Alaska, in Ahlbrandt, T.S., ed., Preliminary Geologic, Petrologic, and Paleontologic Results of the Study of Nanushuk Group

120

Rocks, North Slope, Alaska: U.S. Geological Survey Circular 794, p. 32–41. Bird, K.J., and Molenaar, C.M., 1992, The North Slope foreland basin, Alaska, in Macqueen, R.W., and Leckie, D.A., eds., Foreland Basins and Fold Belts: American Association of Petroleum Geologists Memoir 55, p. 363–393. Bondarenko, G.E., Soloviev, A.V., Tuchkova, M.I., Garver, J.I., and Podgornyi, I.I., 2003, Age of detrital zircons from sandstones of the Mesozoic flysch formation in the South Anyui suture zone (western Chukotka): Lithology and Mineral Resources, v. 38, p. 162–176, doi:10.1023/A:1023456126348. Bradley, D.C., McClelland, W.C., Wooden, J.L., Till, A.B., Roeske, S.M., Miller, M.L., Karl, S.M., and Abbott, J.G., 2007, Detrital zircon geochronology of some Neoproterozoic to Triassic rocks in interior Alaska: Geological Society of America Special Paper 431, p. 155–189. Campbell, R.H., 1967, Areal geology in the vicinity of the Chariot site, Lisburne Peninsula, northwestern Alaska: U.S. Geological Survey Professional Paper 395, 71 p. Chapman, R.M., and Sable, E.G., 1960, Geology of the Utukok-Corwin region, northwestern Alaska: U.S. Geological Survey Professional Paper 303-C, p. 47–167. Christiansen, P.P., and Snee, L.W., 1994, Structure, metamorphism, and geochronology of the Cosmos Hills and Ruby Ridge, Brooks Range schist belt, Alaska: Tectonics, v. 13, p. 193–213, doi:10.1029/93TC01545. Crane, R.C., 1987, Cretaceous olistostrome model, Brooks Range, Alaska, in Tailleur, I., and Weimer, P., eds., Alaskan North Slope Geology: Bakersfield, California, Pacific Section of Society of Economic Paleontologists and Mineralogists and the Alaska Geological Society, Book 50, p. 433–440. Curtis, S.M., Ellersieck, I., Mayfield, C.F., and Tailleur, I.L., 1984, Reconnaissance geologic map of southwestern Misheguk Mountain quadrangle, Alaska: U.S. Geological Survey Miscellaneous Investigations series Map I-1502, scale 1:63,360. Curtis, S.M., Ellersieck, I., Mayfield, C.F., and Tailleur, I.L., 1990, Reconnaissance geologic map of the De Long Mountains A1 and B1 quadrangles and part of the C2 quadrangle, Alaska: U.S. Geological Survey Miscellaneous Investigations Series Map I-1930, scale 1:63,360. DeCelles, P.G., 2012, Foreland basin systems revisited: Variations in response to tectonic settings, in Busby, C., and Azor, A., eds., Tectonics of Sedimentary Basins: Recent Advances: Blackwell Publishing Ltd., p. 405–426. DeCelles, P.G., and Giles, K.A., 1996, Foreland basin systems: Basin Research, v. 8, p. 105–123, doi:10.1046/j .1365-2117.1996.01491.x. De Vera, J., 2005, Structure of the Red Dog District, western Brooks Range, Alaska [Ph.D. thesis]: United Kingdom, Royal Holloway, University of London, 664 p. Dickinson, W.R., 1985, Interpreting provenance relations from detrital modes of sandstones, in Zuffa, G.G., ed., Provenance of Arenites: Boston, D. Reidel Publishing Company, p. 333–361. Dickinson, W.R., 2009, Anatomy and global context of the North American Cordillera, in Kay, S.M., Ramos, V.A., and Dickinson, W.R., eds., Backbone of the Americas: Shallow Subduction, Plateau Uplift, and Ridge and Terrane Collision: Geological Society of America Memoir 204, p. 1–29. Dickinson, W.R., and Gehrels, G.E., 2008, Sediment delivery to the Cordilleran foreland basin: Insights from U-Pb ages of detrital zircons in Upper Jurassic and Cretaceous strata of the Colorado Plateau: American Journal of Science, v. 308, p. 1041–1082. Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., Gehrels, G.E., and Young, R.A., 2012, Provenance of the Paleogene Colton Formation (Uinta Basin) and Cretaceous–Paleogene provenance evolution in the Utah foreland: Evidence from U-Pb ages of detrital zircons, paleocurrent trends, and sandstone petrofacies: Geosphere, v. 8, p. 854–880, doi:10.1130/GES00763.1. Drachev, S.S., Malyshev, N.A., and Nikishin, A.M., 2010, Tectonic history and petroleum geology of the Russian Arctic Shelves: An overview, in Vining, B.A., and Pickering, S.C., eds., Petroleum Geology: From Mature Basins to New Frontiers: Proceedings of the

Geosphere, February 2015

7th Petroleum Geology Conference, Geological Society of London, p. 591–619. doi:10.1144/0070591. Dumitru, T.A., Miller, E.L., O’Sullivan, P.B., Amato, J.M., Hannula, K.A., Calvert, A.T., and Gans, P.B., 1995, Cretaceous to Recent extension in the Bering Strait region, Alaska: Tectonics, v. 14, p. 549–563, doi:10 .1029/95TC00206. Dumoulin, J.A., Harris, A.G., Gagiev, M., Bradley, D.C., and Repetski, J.E., 2002, Lithostratigraphic, conodont, and other faunal links between lower Paleozoic strata in northern and central Alaska and northeastern Russia, in Miller, E.L, Grantz, A., and Klemperer, S.L., eds., Tectonic Evolution of the Bering Shelf–Chukchi Sea– Arctic Margin and Adjacent Landmasses: Geological Society of America Special Paper 360, p. 291–312. Dumoulin, J., Johnson, C., Slack, J., Bird, K., Whalen, M., Moore, T., Harris, A., and O’Sullivan, P., 2013, Carbonate margin, slope, and basin facies of the Lisburne Group (Carboniferous–Permian) in northern Alaska, in Verwer, K., Playton, T., and Harris, P., eds., Deposits, Architecture, and Controls of Carbonate Margin, Slope, and Basin Systems: SEPM (Society for Sedimentary Geology) Special Publication No. 105, p. 211–236. doi:10.2110/sepmsp.105.02 Ellersieck, I., Curtis, S.M., Mayfield, C.F., and Tailleur, I.L., 1990, Reconnaissance geologic map of the De Long Mountains A2 and B2 quadrangles and part of the C2 quadrangle, Alaska: U.S. Geological Survey Miscellaneous Investigations Series Map I-1931, scale 1:63,360. Fisher, W.I., Brown, L.F., Jr., Scott, A.J., and McGowen, J.H., 1969, Delta systems in the exploration for oil and gas, a research colloquium: Austin, Texas, Bureau of Economic Geology, University of Texas, 78 p. Gehrels, G., 2010, Analysis Tools, http://www.geo.arizona .edu/alc/. Gottlieb, E.S., Miller, E.L., and Akinin, V.V., 2012, Crustal architecture of Arctic Chukotka revealed by field mapping and zircon U-Pb geochronology: San Francisco, California, American Geophysical Union Fall Meeting, 3–7 December, 2012, abstract T34A-03. Graham, S.A., Ingersoll, R.V., and Dickinson, W.R., 1976, Common provenance for lithic grains in carboniferous sandstones from Ouachita Mountains and Black Warrior Basin: Journal of Sedimentary Petrology, v. 46, p. 620–632. Grantz, A., Scott, R.A., Drachev, S.S., Moore, T.E., and Valin, Z.C., 2011a, Sedimentary successions of the Arctic region (58–64° to 90°N) that may be prospective for hydrocarbons, in Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, A.V., and Sorensen, K., eds., Arctic Petroleum Geology: Geological Society, London, Memoir 35, p. 17–37. doi:10.1144/M35.2. Grantz, A., Hart, P.E., and Childers, V.A., 2011b, Geology and tectonic development of the Amerasia and Canada Basins, Arctic Ocean, in Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, A.V., and Sorensen, K., eds., Arctic Petroleum Geology: Geological Society, London, Memoir 35, p. 771–800. Guynn, J., and Gehrels, G., 2010, Comparison of detrital zircon age distributions using the K-S test: Tucson, University of Arizona, 16 p., http://www.geo.arizona .edu/agc/labs.html (accessed 15 May 2012). Hanchar, J.M., and Watson, E.B., 2003, Zircon saturation thermometry, in Hanchar, J.M., and Hosin, P.W.O., eds., Zircon: Reviews in Mineralogy and Geochemistry, v. 53, p. 89–112. Harris, R., 2004, Tectonic evolution of the Brooks Range ophiolite, northern Alaska: Tectonophysics, v. 392, p. 143–163. Houseknecht, D.W., and Wartes, M.A., 2013, Clinoform deposition across a boundary between orogenic front and foredeep—An example from the Lower Cretaceous in Arctic Alaska: Terra Nova, v. 25, p. 206–211, doi:10.1111/ter.12024. Houseknecht, D.W., Bird, K.J., and Schenk, C.J., 2009, Seismic analysis of clinoform depositional sequences and shelf-margin trajectories in Lower Cretaceous (Albian) strata, Alaska North Slope: Basin Research, v. 21, p. 644–654, doi:10.1111/j.1365-2117.2008.00392.x. Houseknecht, D.W., Bird, K.J., and O’Sullivan, P., 2011, Constraining the age and magnitude of uplift in the northern National Petroleum Reserve in Alaska

Evolution of lower Brookian foreland basin deposits, northern Alaska (NPRA)—Apatite fission-track analysis of samples from three wells, in Dumoulin, J.A., and Dusel-Bacon, C., eds., Studies by the U.S. Geological Survey in Alaska, 2010: U.S. Geological Survey Professional Paper 1784-A, 21 p., 1 plate. Hubbard, R.J., Edrich, S.P., and Rattey, R.P., 1987, Geologic evolution and hydrocarbon habitat of the ‘Arctic Alaska microplate’: Marine and Petroleum Geology, v. 4, p. 2–34, doi:10.1016/0264-8172(87)90019-5. Huffman, A.C., Jr., Ahlbrandt, T.S., and Bartsch-Winkler, S., 1988, Sedimentology of the Nanushuk Group, North Slope, in Gryc, G., ed., Geology and Exploration of the National Petroleum Reserve in Alaska, 1974 to 1982: U.S. Geological Survey Professional Paper 1399, p. 281–298. Ingersoll, R.V., and Suczek, C.A., 1979, Petrology and provenance of the Neogene sand from Nicobar and Bengal fans, DSDP Sites 211 and 218: Journal of Sedimentary Petrology, v. 49, p. 1217–1228. Ingersoll, R.V., Bullard, T.F., Ford, T.L., Grimm, J.P., and Pickle, J.D., 1984, The effect of grain size on detrital modes: A test of the Gazzi-Dickinson point counting method: Journal of Sedimentary Petrology, v. 54, p. 103–116. Ingersoll, R.V., Cavazza, W., Graham, S.A., and Indiana University Graduate Field Seminar Participants, 1987, Provenance of impure calclithites in the Laramide foreland of southwestern Montana: Journal of Sedimentary Petrology, v. 57, p. 995–1003. Karl, S.M., and Aleinikoff, J.N., 1990, Proterozoic U-Pb zircon age of granite in the Kallarichuk Hills, western Brooks Range Alaska: Evidence for Precambrian basement in the Schist Belt, in Dover, J.H., and Galloway, J.P., eds., Geologic Studies in Alaska by the U.S. Geological Survey, 1988: U.S. Geological Survey Bulletin 1903, p. 10–19. Kelley, K.D., and Jennings, S., 2004, A special issue devoted to barite and Zn-Pb-Ag deposits in the Red Dog district, western Brooks Range, Alaska: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 99, p. 1267–1280, doi:10.2113/gsecongeo.99 .7.1267. Klemperer, S.L., Miller, E.L., Grantz, A., and Scholl, D.W., and the Bering-Chukchi Working Group, 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., and Grantz, A., eds., Tectonic Evolution of the Bering Shelf–Chukchi Sea–Arctic Margin and Adjacent Landmasses: Geological Society of America Special Paper 360, p. 1–24. Kos’ko, M.K., Cecile, M.P., Harrison, J.C., Banelin, V.G., Khandoshko, N.V., and Lopatin, B.G., 1993, Geology of Wrangel Island, between Chukchi and East Siberian seas, northeastern Russia: Geological Survey Canada Bulletin, v. 461, 101 p. Lane, L.S., 1997, Canada Basin, Arctic Ocean: Evidence against a rotational origin: Tectonics, v. 16, p. 363– 387, doi:10.1029/97TC00432. Lawton, T.F., Pollock, S.L., and Robinson, R.A.J., 2003, Integrating sandstone petrology and nonmarine sequence stratigraphy: Application to the Late Cretaceous fluvial systems of southwestern Utah, U.S.A.: Journal of Sedimentary Research, v. 73, p. 389–406, doi:10.1306 /100702730389. Lawton, T.F., Bradford, I.A., Vega, F.J., Gehrels, G.E., and Amato, J.M., 2009, Provenance of Upper Cretaceous– Paleogene sandstones in the foreland basin system of the Sierra Madre Oriental, northeastern Mexico, and its bearing on fluvial dispersal systems of the Mexican Laramide Province: Geological Society of America Bulletin, v. 121, p. 820–836. Lawver, L.A., Grantz, A., and Gahagan, L.M., 2002, Plate kinematic evolution of the present Arctic region since the Ordovician, in Miller, E.L, Grantz, A., and Klemperer, S.L., eds., Tectonic Evolution of the Bering Shelf–Chukchi Sea–Arctic Margin and Adjacent Landmasses: Geological Society of America Special Paper 360, p. 333–358. Lee, J.K.W., Williams, I.S., and Ellis, D.J., 1997, Pb, U and Th diffusion in natural zircon: Nature, v. 390, p. 159– 162, doi:10.1038/36554.

Lerand, M., 1973, Beaufort Sea, in McCrossam, R.G., ed., The Future Petroleum Provinces of Canada—Their Geology and Potential: Canadian Society of Petroleum Geology Memoir 1, p. 315–386. Little, T.A., Miller, E.L., Lee, J., and Law, R.D., 1994, Extensional origin of ductile fabrics in the Schist Belt, central Brooks Range, Alaska—I. Geologic and structural studies: Journal of Structural Geology, v. 16, p. 899–918, doi:10.1016/0191-8141(94)90075-2. Ludwig, K.R., 2005, Isoplot version 3.0: http://www.bgc.org /isoplot_etc/software.html (accessed October 2005). MacLean, J.S., Sears, J.W., Chamberlain, K.R., Khudoley, A.K., Prokopiev, A.V., Kropachev, A.P., and Serkina, G.G., 2009, Detrital zircon geochronologic tests of the SE Siberia-SW Laurentia paleocontinental connection: European Geophysical Union Stephan Mueller Special Publication Series, v. 4, p. 111–116, doi:10.5194 /smsps-4-111-2009. Martin, A.J., 1970, Structure and tectonic history of the western Brooks Range, De Long Mountains, and Lisburne Hills, northern Alaska: Geological Society of America Bulletin, v. 81, p. 3605–3622, doi:10.1130 /0016-7606(1970)81[3605:SATHOT]2.0.CO;2. Mayfield, C.F., Tailleur, I.L., and Ellersieck, I., 1988, Stratigraphy, structure, and palinspastic synthesis of the western Brooks Range, northwestern Alaska, in Gryc, George, ed., Geology and Exploration of the National Petroleum Reserve in Alaska, 1974 to 1982: U.S. Geological Survey Professional Paper 1399, p. 143–186. Miller, E.L., and Akinin, V.V., 2008, Geology of the Bering Shelf region of Alaska-Russia: Implications for extensional processes in continental crust, in Spencer, J.E., and Titley, S.R., eds., Ores and Orogenesis: Arizona Geological Society Digest 22, p. 203–212. Miller, E.L., and Hudson, T.L., 1991, Mid-Cretaceous extensional fragmentation of a Jurassic–Early Cretaceous compressional orogen, Alaska: Tectonics, v. 10, p. 781–796, doi:10.1029/91TC00044. Miller, E.L., Toro, J., Gehrels, G., Amato, J.M., Prokopiev, A., Tuchkova, M.I., Akinin, V.V., Dumitru, T.A., Moore, T.E., and Cecile, M.P., 2006, New insights into Arctic paleogeography and tectonics from U-Pb detrital zircon geochronology: Tectonics, v. 25, TC3013, doi:10.1029/2005TC001830. Miller, E.L., Soloviev, A., Kuzichev, A., Gehrels, G., Toro, J., and Tuchkova, M., 2009, Jurassic and Cretaceous foreland basin deposits of the Russian Arctic: Separated by birth of the Makaro Basin?: Norwegian Journal of Geology, v. 88, p. 201–226. Miller, E.L., Gehrels, G.E., Pease, V., and Sokolov, S., 2010, Paleozoic and Mesozoic stratigraphy and U-Pb detrital zircon geochronology of Wrangel Island, Russia: Constraints on paleogeography and paleocontinental reconstructions of the Arctic: American Association of Petroleum Geologists Bulletin, v. 94, p. 665–692, doi: 10.1306/10200909036. Miller, E.L., Kuznitsov, N., Soboleva, A., Udoratina, O., Grove, M.J., and Gehrels, G., 2011, Baltica in the Cordillera?: Geology, v. 39, p. 791–794, doi:10.1130 /G31910.1. Miller, E.L., Soloviev, A., Prokopiev, A.V., Toro, J., Harris, D., Kuzmichev, A.B., and Gehrels, G.E., 2013, Triassic river systems and the paleo-Pacific margin of northwestern Pangea: Gondwana Research, v. 23, p. 1631– 1645, doi:10.1016/j.gr.2012.08.015. Molenaar, C.M., 1988, Depositional history and seismic stratigraphy of Lower Cretaceous rock in the National Petroleum Reserve in Alaska and adjacent areas, in Gryc, G., ed., Geology and Exploration of the National Petroleum Reserve in Alaska 1974–1982: U.S. Geological Survey Professional Paper 1399, p. 593–621. Molenaar, C.M., Egbert, R.M., and Krystinik, L.F., 1988, Depositional facies, petrography, and reservoir potential of the Fortress Mountain Formation (Lower Cretaceous), central North Slope, Alaska, in Gryc, George, ed., Geology and Exploration of the National Petroleum Reserve in Alaska, 1974 to 1982: U.S. Geological Survey Professional Paper 1399, p. 257–280. Moll-Stalcup, E., Wooden, J.L., Bradshaw, J., and Aleinikoff, J., 1996, Elemental and isotopic evidence for 2.1-Ga arc magmatism in the Kilbuck terrane, southwestern Alaska: U.S. Geological Survey Bulletin 2152, p. 111–130.

Geosphere, February 2015

Moore, D.W., Young, L.E., Modene, J.S., and Plahuta, J.T., 1986, Geologic setting and genesis of the Red Dog zinc-lead-silver deposit, western Brooks Range, Alaska: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 81, p. 1696–1727, doi:10.2113/gsecongeo.81.7.1696. Moore, G.W., Drummond, K.J., Teraoka, Y., Okulitch, A.V., Gramberg, L.S., Pogrebitsky, Y.E., Musatov, E.E., Parfenov, L.M., Natal’in, B.A., Johannesson, H., McCoy, F.W., Simkin, T., and Siebert, L., 2000, Geologic map of the Circum-Pacific region, Arctic Sheet: Circum-Pacific Council for Energy and Mineral Resources, U.S. Geological Survey Circum-Pacific Map Series Map CP-48, scale 1:10,000,000. Moore, T.E., 2010, Detrital zircon U-Pb age populations in time and space in the Arctic Alaska terrane: San Francisco, California, American Geophysical Union Fall Meeting, 13–17 December, 2010, abstract T31A-2146. Moore, T.E., 2014, U-Pb zircon age data for selected sedimentary, metasedimentary, and igneous rocks from northern and central Alaska: U.S. Geological Survey Data Series 899, 4 p., http://dx.doi.org/10.3133/ds899. Moore, T.E., and Potter, C.J., 2008, Preliminary retrodeformable regional cross section, western Brooks Range, in Kelley, K.F., ed., Regional Fluid Flow and Basin Modeling in Northern Alaska: U.S. Geological Survey Circular 1319, p. 28–31. Moore, T.E., Wallace, W.K., Bird, K.J., Karl, S.M., Mull, C.G., and Dillon, T.T., 1994, Geology of northern Alaska, in Plafker, G., and Berg, H.C., eds., The Geology of North America, v. G-1, The Geology of Alaska: Boulder, Colorado, Geological Society of America, p. 49–140. Moore, T.E., Hemming, S., and Sharp, W.D., 1997, Provenance of the Carboniferous Nuka Formation, Brooks Range, Alaska: A multicomponent isotope provenance study with implications for age of cryptic crystalline basement, in Dumoulin, J.A., and Gray, J.E., eds., Geological Studies in Alaska by the U.S. Geological Survey, 1995: U.S. Geological Survey Professional Paper 1574, p. 173–194. Moore, T.E., Dumitru, T.A., Adams, K.E., Witebsky, S.N., and Harris, A.G., 2002, Origin of the Lisburne Hills– Herald arch structural belt: Stratigraphic, structural, and fission-track evidence from the Cape Lisburne area, northwestern Alaska, in Miller, E.L., and Grantz, A., eds., Tectonic Evolution of the Bering Shelf–Chukchi Sea–Arctic Margin and adjacent landmasses: Geological Society of America Special Paper 360, p. 77–109. Mull, C.G., 1985, Cretaceous tectonics, depositional cycles, and the Nanushuk Group, Brooks Range and Arctic Slope, Alaska, in Huffman, A.C., ed., Geology of the Nanushuk Group and Related Rocks, North Slope, Alaska: U.S. Geological Survey Bulletin 1614, p. 7–36. Mull, C.G., 1995, Preliminary evaluation of the hydrocarbon source rock potential of the Tingmerkpuk Sandstone (Neocomian) and related rocks, northwestern De Long Mountains, Brooks Range, Alaska: Division of Geological and Geophysical Surveys Public-Data File 95-30, 20 p. Mull, C.G., Harris, E.E., Reifenstuhl, R.R., and Moore, T.E., 2000, Geologic map of the Coke basin—Kukpowruk River area, DeLong Mountains D-2 and D-3 Quadrangles, northwestern Alaska: Alaska Division Geological and Geophysical Surveys Report of Investigations 2000-2, 1 sheet, scale 1:63,360. Mull, C.G., Houseknecht, D.W., and Bird, K.J., 2003, Revised Cretaceous and Tertiary stratigraphic nomenclature in the east-central Colville basin, northern Alaska: U.S. Geological Survey Professional Paper 1673, 51 p., http://pubs.usgs.gov/pp/p1673/text.html (accessed April 19, 2012). Natal’in, B.A., Amato, J.M., Toro, J., and Wright, J.E., 1999, Paleozoic rocks of northern Chukotka Peninsula, Russian Far East: Implications for the tectonics of the Arctic region: Tectonics, v. 18, p. 977–1003, doi:10.1029 /1999TC900044. Nokleberg, W.J., Parfenov, L.M., Monger, J.W.H., Norton, I.O., Khanchuk, A.I., Stone, D.B., Scotese, C.R., Scholl, D.W., and Fujita, K., 2000, Phanerozoic tectonic evolution of the Circum-North Pacific: U.S. Geological Survey Professional Paper 1626, 122 p.

121

Moore et al. Ogg, J.G., and Ogg, G., 2008, Early Cretaceous (103– 138 Ma time-slice) update: https://engineering.purdue .edu /Stratigraphy /charts /Timeslices /4 _Mid -Cret .pdf (accessed 24 January 2014). Oldow, J.S., Seidensticker, C.M., Phelps, J.C., Julian, F.E., Gottschalk, R.R., Boler, K.W., Handschy, J.W., and Ave Lallemant, H.G., 1987, Balanced cross sections through the central Brooks Range and North Slope, Arctic Alaska: American Association of Petroleum Geologists publication, 19 p., 8 plates, scale 1:200,000. Patton, W.W., Jr., Wilson, F.H., Labay, K.A., and Shew, N., 2009, Geologic map of the Yukon-Koyukuk basin, Alaska: U.S. Geological Survey Scientific Investigations Map 2909, scale 1:500,000. Pease, V., 2011, Eurasian orogens and Arctic tectonics: an overview, in Spencer, A.M., Embry, A.F., Gautier, D.L., Stoupakova, A.V., and Sorensen, K., eds., Arctic Petroleum Geology: Geological Society, London, Memoir 35, p. 311–324. doi:10.1144/M35.2. Press, W.H., Flannery, B.P., Tenkolsky, S.A., and Vetterling, W.T., 1986, Numerical recipes: Cambridge, Cambridge University Press, 818 p. Roeske, S.M., Dusel-Bacon, C., Aleinikoff, J.N., Snee, L.W., and Lanphere, M.A., 1995, Metamorphic and structural history of continental crust at a Mesozoic collisional margin, the Ruby terrane, central Alaska: Journal of Metamorphic Geology, v. 13, p. 25–40. Rubin, C.M., Miller, E.L., and Toro, J., 1995, Deformation of the northern circum-Pacific margin: Variations in tectonic style and plate-tectonic implications: Geology, v. 23, p. 897–900, doi:10.1130/0091-7613(1995)023 2.3.CO;2. Saylor, J.E., Horton, B.K., Nie, J., Corredor, J., and Mora, A., 2011, Evaluating foreland basin partitioning in the northern Andes using Cenozoic fill of the Floresta basin, Eastern Cordillera, Columbia: Basin Research, v. 23, p. 377–402. Schenk, O., Magoon, L.B., Bird, K.J., and Peters, K.E., 2012, Petroleum system modeling of northern Alaska, in Peters, K.E., Durry, D.J., and Kacewicz, M., eds., Basin Modeling: New Horizons in Research and Applications: American Association of Petroleum Geologists Hedberg Series, no. 4, p. 317–338. Smiley, C.J., 1969, Floral zones and correlations of Cretaceous Kukpowruk and Corwin Formations, northwestern Alaska: American Association of Petroleum Geologists Bulletin, v. 53, p. 2079–2093. Sokolov, S.D., 2010, Tectonics of northeast Asia: An overview: Geotectonics, v. 44, no. 6, p. 493–509, doi:10 .1134/S001685211006004X.

122

Sokolov, S.D., Bondarenko, G.Y., Morozov, O.L., Shekhovtsov, V.A., Glotov, S.P., Ganelin, A.V., and KravchenkoBerezhnoy, I.R., 2002, South Anyui suture, northeast Arctic Russia: Facts and problems, in Miller, E.L., and Grantz, A., eds., Tectonic Evolution of the Bering Shelf–Chukchi Sea–Arctic Margin and Adjacent Landmasses: Geological Society of America Special Paper 360, p. 209–224. Sokolov, S.D., Bondarenko, G.Y., Layer, P.W., and Kravchenko-Berezhnoy, I.R., 2009, South Anyui suture: Tectono-stratigraphy, deformations, and principal tectonic events: European Geophysical Union Stephan Mueller Special Publication Series, v. 4, p. 201–221, doi:10.5194/smsps-4-201-2009. Spicer, R.A., and Herman, A.B., 2010, The Late Cretaceous environment of the Arctic: A quantitative reassessment based on plant fossils: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 295, p. 423–442, doi:10.1016 /j.palaeo.2010.02.025. Tikhomirov, P.L., Kalinina, E.A., Kobayashi, K., and Nakamura, E., 2008, Late Mesozoic silicic magmatism of the North Chukotka area (NE Russia): Age, magma sources, and geodynamic implications: Lithos, v. 105, p. 329–346, doi:10.1016/j.lithos.2008.05.005. Till, A.B., 1992, Detrital blueschist-facies metamorphic mineral assemblages in Early Cretaceous sediments of the foreland basin of the Brooks Range, Alaska, and implications for orogenic evolution: Tectonics, v. 11, p. 1207–1223, doi:10.1029/92TC01104. Tolson, R.B., 1987, Structure and stratigraphy of the Hope Basin, southern Chukchi Sea, Alaska, in Scholl, D.W., Grantz, A., and Vedder, J.G., eds., Geology and Resource Potential of the Continental Margin of Western North America and Adjacent Ocean Basins—Beaufort Sea to Baja California: Circum-Pacific Council for Energy and Mineral Resources Earth Science Series, v. 6, p. 59–71. Toro, J., Cole, F., and Meier, J.M., 1998, 40Ar/39Ar ages of detrital minerals in Lower Cretaceous rocks of the Okpikruak Formation: Evidence for Upper Paleozoic metamorphic rocks in the Koyukuk arc, in Gray, J.E., and Riehle, J.R., eds., Geologic Studies in Alaska by the U.S. Geological Survey, 1996: U.S. Geological Survey Professional Paper 1595, p. 169–182. Torsvik, T.H., and Andersen, T.B., 2002, The Taimry fold belt, Arctic Siberia: Timing of prefold demagnetization and regional tectonics: Tectonophysics, v. 352, p. 335– 348, doi:10.1016/S0040-1951(02)00274-3. Tuchkova, M.I., Sokolov, S., and Kravchenko-Berezhnoy, I.R., 2009, Provenance analysis and tectonic setting of the Triassic clastic deposits in western Chukotka,

Geosphere, February 2015

northeast Russia: European Geophysical Union Stephan Mueller Special Publication Series, v. 4, p. 177–200, doi:10.5194/smsps-4-177-2009. Vermeesch, P., 2009, RadialPlotter: A Java application for fission track, luminescence and other radial plots: Radiation Measurements, v. 44, p. 409–410, doi:10 .1016/j.radmeas.2009.05.003. Vogl, J.J., Calvert, A.T., and Gans, P.B., 2002, Mechanisms and timing of exhumation of collision-related metamorphic rocks, southern Brooks Range, Alaska: Insights from 40Ar/39Ar thermochronology: Tectonics, v. 21, no. 3, doi:10.1029/2000TC001270. Walker, J.D., Geissman, J.W., Bowring, S.A., and Babcock, L.E., 2013, The Geological Society of America Geologic Time Scale: Geological Society of America Bulletin, v. 125, p. 259–272, doi:10.1130/B30712.1. Wallace, W.K., Moore, T.E., and Plafker, G., 1997, Multistory duplexes with forward dipping roofs, north central Brooks Range, Alaska: Journal of Geophysical Research, v. 102, p. 20,773–20,796, doi:10.1029 /96JB02682. Wartes, M.A., 2008, Evaluation of stratigraphic continuity between the Fortress Mountain and Nanushuk formation in the central Brooks Range foothills—Are they partly correlative?, in Wartes, M.A., and Decker, P.L., eds., Preliminary Results of Recent Geologic Field Investigations in the Brooks Range Foothills and North Slope, Alaska: Alaska Division of Geological and Geophysical Surveys Preliminary Interpretive Report 2008-1, p. 25–39. Wartes, M.A., and Reifenstuhl, R.R., 1998, Preliminary petrography and provenance of six Lower Cretaceous sandstones, northwestern Brooks Range, Alaska, in Clough, J.G., and Larson, F., eds., Short Notes on Alaska Geology 1997: Alaska Division Geological and Geophysical Surveys Professional Report 118, p. 131–140. Wartes, M.A., O’Sullivan, P.B., and Carroll, A.R., 2006, Enigmatic Permian U-Pb zircon ages from igneous clasts found in mid-Cretaceous deposits of the Colville foreland basin, northern Alaska: Geological Society of America Abstracts with Programs, v. 38, no. 5, p. 89. Young, L.E., 2004, A geologic framework for mineralization in the western Brooks Range, Alaska: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 99, p. 1281–1306, doi:10.2113/gsecongeo.99 .7.1281. Zuffa, G.G., 1985, Optical analyses of arenites: Influence of methodology on compositional results, in Zurffa, G.G., ed., Provenance of Arenites: Boston, D. Reidel Publishing Company, p. 165–189.