Download Upper plate proxies for flat-slab subduction processes in southern

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Provenance (geology) wikipedia , lookup

History of geology wikipedia , lookup

Geology wikipedia , lookup

Plate tectonics wikipedia , lookup

Large igneous province wikipedia , lookup

Oceanic trench wikipedia , lookup

Transcript
EPSL-10760; No of Pages 13
Earth and Planetary Science Letters xxx (2011) xxx–xxx
Contents lists available at ScienceDirect
Earth and Planetary Science Letters
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / e p s l
Upper plate proxies for flat-slab subduction processes in southern Alaska
Emily S. Finzel a,⁎, Jeffrey M. Trop b, Kenneth D. Ridgway a, Eva Enkelmann c
a
b
c
Dept. of Earth and Atmospheric Sciences, Purdue University, West Lafayette, IN 47907, USA
Dept. of Geology, Bucknell University, Moore Avenue, Lewisburg, PA 17837, USA
Universität Tübingen-Institut für Geowissenschaften, Wilhelmstr. 56, 72074, Tübingen, Germany
a r t i c l e
i n f o
Article history:
Received 19 June 2010
Received in revised form 12 January 2011
Accepted 13 January 2011
Available online xxxx
Editor: T.M. Harrison
Keywords:
Alaska
flat-slab subduction
geochronology
thermochronology
provenance
sedimentary basin
a b s t r a c t
The timing of initiation of flat-slab subduction beneath southern Alaska and the upper plate record of this
process are not well understood. We explore the record of flat-slab subduction in southern Alaska by
integrating stratigraphic, provenance, geochronologic, and thermochronologic data from the region directly
above and around the perimeter of ongoing flat-slab subduction. These datasets document a change from
regional Paleocene–Oligocene subduction-related magmatism and basin development to an absence of
magmatism and initiation of rock exhumation that continues to today. We infer that initiation of flat-slab
subduction prompted crustal shortening, exhumation, inversion of sedimentary basins, and cessation of
magmatism above and around the area of ongoing flat-slab subduction. Surface uplift and erosion above the
flat slab resulted in deposition of thick, clastic wedges in sedimentary basins located along the western and
northern perimeters of the flat-slab region. Along the eastern perimeter, northwestward-propagating
Oligocene–Quaternary slab-edge volcanism and transtensional basin development along dextral strike-slip
faults record progressive northwestward insertion of a shallow slab against the curved continental margin of
eastern Alaska. Collectively, these geologic data indicate that flat-slab subduction was shaping southern
Alaska by late Eocene–early Oligocene time, much earlier than previous models infer. Upper plate processes
related to subduction of a flat slab in Alaska are similar to those documented in other modern flat-slab regions.
These processes include: 1) shortening and exhumation of the upper plate several hundred kilometers
inboard from the plate margin, 2) cessation of subduction-related magmatism within ten million years of
the onset of shallow subduction, 3) shoaling or inversion of sedimentary basins above the flat-slab region,
4) deformation and/or erosion of the accretionary prism during flat-slab subduction, and 5) in some settings,
reestablishment of sedimentation in the accretionary prism after the shallow slab has migrated laterally.
Unlike other flat-slab margins, strike-slip-related volcanism and basin development characterizes one edge of
the flat-slab region in Alaska, a consequence of oblique insertion of the flat slab into the corner of a curved
continental margin.
© 2011 Elsevier B.V. All rights reserved.
1. Introduction
The upper plate geologic record of flat-slab subduction is a topic of
considerable discussion for both modern and ancient convergent
margins (e.g., Dickinson and Snyder, 1979; Espurt et al., 2008;
Hampel, 2002; Jordan and Allmendinger, 1986). Studies based on
seismicity and associated crustal deformation suggest that flat-slab
subduction results in zones of broad, diffuse deformation at
convergent plate margins as it increases the transfer of compressive
stresses inboard and upward to the overriding plate (Gutscher et al.,
2000; Lallemand et al., 2005). The southern margin of Alaska is
characterized by an east-to-west transition from transform tectonics
to flat-slab subduction to normal subduction (Figs. 1 and 2). On the
upper plate, the eastern region is marked by active volcanism in the
⁎ Corresponding author. Now at ExxonMobil Exploration Co., Houston, TX 77060,
USA. Tel.: + 1 765 490 9055; fax: +1 765 496 1210.
E-mail address: emily.s.fi[email protected] (E.S. Finzel).
Wrangell volcanic belt, active dextral displacement along the Denali–
Totschunda fault system, and shallow seismicity (b50 km; Figs. 1 and
3A). The central region is distinguished by relatively high topography,
a lack of active volcanism, and a shallowly-dipping Wadati–Benioff
zone produced by the Yakutat microplate that extends nearly
horizontally for ~ 250 km northwestward beneath Alaska before
reaching a depth of 150 km more than ~ 600 km inboard of the
Aleutian trench (Figs. 2 and 3B). In the western region, subduction of
the Pacific oceanic plate produces a more steeply dipping Wadati–
Benioff zone that reaches depths of 100–150 km within ~400 km of
the trench (Figs. 2 and 3C) and active volcanism in the Aleutian arc
(AVA on Figs. 1 and 3C).
Cenozoic deformation in southern Alaska associated with subduction of the Yakutat microplate was first proposed by Plafker et al. (1978)
and expanded on by Plafker (1987) and Plafker et al. (1994). In these
previous models, normal oceanic crust of the northwestern part of the
Yakutat microplate was subducted beneath Alaska from early Oligocene
(ca. 30 Ma) to middle Miocene time; the onset of subduction was based
0012-821X/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2011.01.014
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
2
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
Fig. 1. Tectonic framework of southern Alaska, including faults with known or suspected Neogene and younger displacement (modified from Plafker et al., 1994), Holocene volcanoes
(red triangles; modified from Cameron, 2005), plate motion vectors relative to North America (DeMets et al., 1994; Leonard et al., 2007), exposed (light yellow) and interpreted
subducted extent of the Yakutat microplate (region within bold dashed line), locations of exhumation studies discussed in Section 2.1 (yellow circles) and Section 4.2 (orange
circles), and locations of U-Pb detrital samples discussed in Section 3.1 (blue rectangles). The flat-slab region is the area within the solid red line. Edges of the subducted Yakutat
microplate are loosely constrained (modified from Eberhart-Philips et al., 2006 and Fuis et al., 2008). Width of the Neogene accretionary prism shown in gray shaded polygon
(modified from vonHuene et al., 1999). Sedimentary basins: CI, Cook Inlet; TB—Tanana, CB—Colorado Creek, MB—Matanuska, SB—Susitna, CR—Copper River. AVA—Alaska PeninsulaAleutian volcanic arc; MD-Mount Drum; MC-Mount Churchill; KP—Kenai Peninsula; PWS—Prince William Sound; WAR—western Alaska Range; CAR—central Alaska Range; EAR—
eastern Alaska Range; BM, Buzzard Creek maar; WVB—Wrangell volcanic belt; CMF—Castle Mountain fault; CSEF—Chugach-St. Elias fault; DRF-Duke River fault; QC-FF.—QueenCharlotte-Fairweather fault; TF—Totschunda fault.
on the age of the oldest lavas in the Wrangell volcanic belt (WVB on
Fig. 1). These previous models infer that beginning in middle Miocene
time, more buoyant continentalized crust of the southeastern part of the
Yakutat microplate was subducted at a shallow angle beneath the
continental margin of south-central Alaska based on the Neogene age of
a N5-km-thick succession of siliciclastic strata (i.e., the Yakataga
Formation) deposited upon the unsubducted part of the Yakutat
microplate (Plafker et al., 1994).
Recent geophysical studies, however, have shown that the entire
Yakutat microplate (both the subducted and unsubducted parts)
consists of thick, buoyant crust (Christeson et al., 2010; EberhartPhilips et al., 2006; Ferris et al., 2003; Lowe et al., 2008). The
subducted Yakutat slab is imaged as a low-velocity zone with a high
ratio of P-wave to S-wave velocities, consistent with the typical
seismic character of thick oceanic crust. Compilation of these recent
studies shows that the crust of the subducted northwestern portion of
the Yakutat microplate is 11–22-km-thick and the crust of the
unsubducted southeastern portion is 20–25 km thick. Contemporary
studies have focused on the timing for initiation of subduction of the
thicker southeastern portion of the microplate. These studies support
a middle–late Miocene timing for initiation of flat-slab subduction
based on the creation of high topographic relief in the modern
Chugach, St. Elias, and Talkeetna Mountains (Hoffman and Armstrong,
2006; O'Sullivan and Currie, 1996; Parry et al., 2001), and the imaged
length of the subducted portion of the Yakutat microplate compared
to present-day plate motions (Koons et al., 2010). In all these models,
flat-slab subduction in southern Alaska represents a relatively young
event that initiated since middle Miocene time, an interpretation that
has been propagated through the literature (i.e., Gulick et al., 2007;
Pavlis et al., 2004; Rondenay et al., 2010).
In light of the new geophysical findings indicating that the entire
Yakutat slab has thick crust, our study reviews previously published
geologic data and presents new provenance data from the upper plate
of southern Alaska. All these datasets are evaluated in the context of
flat-slab processes. In particular, our documentation of magmatism,
exhumation, basin development, and provenance provides an
integrated record of upper plate responses to flat-slab subduction
and evidence that flat-slab subduction began as early as late Eocene–
early Oligocene time.
2. Processes above flat-slab region
The area above the present-day flat-slab subduction in southern
Alaska is characterized by large-magnitude earthquakes, Oligocene or
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
3
Fig. 2. Map of southern Alaska illustrating slab seismicity (N 50 km depth; Alaska Earthquake Information Center catalog) and locations of transects shown in Fig. 3. Additional
symbols are the same as in Fig. 1. Note the northeastward increase in the gap between slab seismicity and the trench as well as the paucity of seismicity deeper than 50 km along the
northeastern edge of the slab.
older bedrock exposures, no Quaternary volcanoes, and transpressive
deformation evidenced by folds, reverse faults, and dextral strike-slip
faults. The flat-slab region is defined as the area of the upper plate
presently above the shallow part of the subducted Yakutat slab
(b50 km deep; region within solid red line on Fig. 1). Integration of
Fig. 3. Cross-sections showing changes in seismicity (within ~50 km of each transect)
between eastern, central, and western transects across southern Alaska (Alaska
Earthquake Information Center catalog). Locations of transects shown on Fig. 2. Note
that seismicity from all depths is shown and transects are aligned parallel with presentday plate motions. DF—Denali fault; TR—Transition fault; see Fig. 1 for additional
abbreviations. Default depths of 10 km and 33 km are assigned for events with poorly
constrained depths in oceanic and continental areas, respectively.
new and recently published geochronologic, thermochronologic,
and stratigraphic data help to constrain the timing of magmatism,
exhumation, and basin development in the flat-slab region.
2.1. Geochronology
2.1.1. Data
Although previous workers document the presence of a modern
volcanic gap above the subducting flat-slab (Nye, 1999; Fuis et al., 2008),
the timing of regional cessation of magmatism has not been reported in
the literature. Our new compilation of all published Cenozoic crystallization ages from plutonic and volcanic rocks exposed in south-central
Alaska indicates cessation of subduction-related magmatism in the flatslab region during late Paleogene time (Fig. 4). The publications from
which the data were collected range from 1966 to 2007 and include data
between 140°W–156°W longitude and 60°W–63°W latitude (region
within dotted box on Fig. 1). The compilation includes 505 ages from
bedrock samples analyzed using various techniques and phases (see
Supplementary material for details). Ages that the original authors
deemed questionable or reset and dates on air-fall material were excluded.
Geochronologic data were not recalibrated using modern standards.
Subduction-related magmatism in southern Alaska is attributed to
various episodes of volcanic arc activity, as well as subduction of a
spreading ridge between ca. 61 and 50 Ma that created a slab-window
that allowed for mantle upwelling to the upper plate (see References
in Supplementary material). Magmatism was regionally extensive
across south-central Alaska during early Paleogene time (ca. 65.5–
44 Ma), but ceased east of ~148°W by ca. 43 Ma and west of ~ 148°W
by ca. 32 Ma based on the published age data.
2.1.2. Interpretation
The cessation of magmatism east of 148° was concurrent with
major plate reorganization when relative plate convergence between
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
4
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
Fig. 4. Plot comparing published igneous ages of plutonic and volcanic rocks versus longitude in south-central Alaska in area outlined by dotted box on Fig. 1. Dashed line denotes
present region above flat-slab subduction. Each circle represents a calculated age that does not depict errors. Note cessation of magmatism throughout the flat-slab region east of
~148°W by ca. 43 Ma and west of ~148°W by ca. 32 Ma. See Data Repository for age data and references.
the subducting oceanic plate and North America shifted from N–NE to
more NW-directed motion ca. 46–43 Ma (Doubrovine and Tarduno,
2008; Engebretson et al., 1985; Stock and Molnar, 1988). This shift in
plate motion would be expected to prevent subduction-related
magmatism east of ~ 148°W because that region would have been
located inboard (east) of the new transform boundary developed as a
consequence of the change in plate motions and therefore would not
have a slab actively subducting beneath that part of the plate margin.
Plafker (1987) proposed that the transform boundary originally
coincided with the Transition fault (Fig. 1) and consequently shifted
inboard (east) ca. 30 Ma, presumably to its present-day position along
the Queen Charlotte–Fairweather fault system (Fig. 1), prior to the
Yakutat microplate moving northward and subducting beneath
Alaska. We agree that the paleo-transform boundary likely coincided
with the present-day Transition fault that separates the Pacific and
Yakutat plates, but we infer that its initiation occurred sometime
between ca. 46 and 32 Ma.
We interpret cessation of magmatism west of ~148°W as the
product of shallow northwestward subduction of the Yakutat slab and
associated compressive crustal stresses in the flat-slab region that
would have inhibited the rise and eruption of magmas generated from
heating and dehydration of the subducted slab (e.g., McNamara and
Pasyanos, 2002; Nye et al., 2002; Qi et al., 2007). Subduction of oceanic
lithosphere with thick crust (aseismic ridges, oceanic plateaus or
seamount chains) typically coincides spatially and temporally with the
absence of arc volcanism (e.g., McGeary et al., 1985; van Hunen et al.,
2002). Localized conduits did permit isolated small-volume eruption of
magmas along the leading Yakutat slab edge. For example, small
Holocene maars in the foothills north of the Alaska Range (BM on Fig. 1)
record eruption of subduction-related magmas far inboard along the
northern edge of the subducted Yakutat slab (Nye et al., 2002).
2.2. Thermochronology
2.2.1. Data
Compilation of all published low-temperature thermochronologic
data from above the subducting Yakutat slab provides a record of the
onset of rock exhumation in the upper plate. The publications from
which the data were collected range from 1989 to 2010 and include all
available thermochronologic data younger than 43 Ma from within
the area above the subducted Yakutat slab (black dashed line on
Fig. 1). Our analysis focuses on thermochronologic data that record
exhumation after 43 Ma.
On the upper plate in the northern St. Elias Mountains, apatite and
zircon fission track (FT) analyses record middle Eocene (ca. 43 Ma)
initiation of rapid exhumation with distinct pulses of exhumation
during late Eocene through Pliocene time (ca. 40–5 Ma), and peak
exhumation ca. 35–20 Ma (1a on Fig. 1; Fig. 5; Enkelmann et al., 2008,
2010; O'Sullivan and Currie, 1996). In the Prince William Sound region
to the west, apatite FT and (U-Th)/He ages reflect exhumation between
ca. 35 and 3 Ma (2 on Fig. 1; Arkle et al., 2009). Farther northwest
(~300 km) in the northern Chugach Mountains, detrital zircon FT data
from near the Knik, Matanuska, and Tazlina glaciers record significant
exhumation events at 40, 36, and 28 and 3.1 Ma, respectively (3a, 3b,
and 3c on Fig. 1; Sendziak et al., 2009). In the same region, apatite FT
ages suggest exhumation between ca. 24 and 17 Ma (4 on Fig. 1; Little
and Naeser, 1989). North of the Chugach Mountains (U-Th)/He ages
Fig. 5. Histogram of cooling ages from bedrock and detrital samples collected in the
St. Elias Mountains (1a, 1b on Fig. 1) in the southern part of the flat-slab region. Fission
track ages are from detrital zircons collected in modern rivers that drain source terranes
in the upper plate as well as accreted lower plate material. The bulk of zircon fission
track (FT) ages that range from ca. 40 to 5 Ma (late Eocene–early Pliocene) are from
rivers with source terranes on the upper plate. These ages suggest exhumation began in
the southern part of the flat-slab region by late Eocene time. The Pliocene (b5 Ma)
peaks in the bedrock apatite (U-Th)/He and detrital zircon FT cooling ages are from
samples located on the south side of the St. Elias Mountains (1b on Fig. 1), where
accretion of the Yakutat sedimentary cover is ongoing. Zircon FT cooling ages older than
ca. 90 Ma are from samples located north of location 1a on Fig. 1. This study provides a
new interpretation for these age data; see Berger and Spotila (2008), Berger et al.
(2008), Enkelmann et al. (2008, 2009, 2010), and Spotila et al. (2004) for detailed
explanations of sample collection and analyses.
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
record exhumation in the Talkeetna Mountains (5 on Fig. 1) ca. 20–
15 Ma (Hoffman and Armstrong, 2006). In the eastern Alaska Range,
~50 km north of the flat-slab region (6 on Fig. 1), apatite FT ages range
from 15 Ma to 1 Ma and (U-Th)/He cooling ages range from 4 Ma to
1 Ma (Benowitz et al., in press). Combined with higher-temperature
thermochronologic ages from K-feldspar and biotite, the available data
indicate rapid exhumation in the eastern Alaska Range from 22 Ma to
present (Benowitz et al., in press). Apatite FT ages from Mt. McKinley in
the central Alaska Range, located even farther north of the flat-slab
region (~200 km; 7 on Fig. 1), indicate a sharp increase in exhumation
ca. 6–5 Ma (Fitzgerald et al., 1995).
The youngest apatite and zircon cooling ages found in Alaska today are
located in the southern and central St. Elias Mountains in lower plate
material that has been accreted to the upper plate. In this region, ongoing
Yakutat subduction causes rapid rock uplift of the accreted material that is
then efficiently eroded by massive glaciers (1b on Fig. 1; Fig. 5; Berger and
Spotila, 2008; Berger et al., 2008; Enkelmann et al., 2009, 2010).
2.2.2. Interpretation
The distinct episodes of exhumation in the upper plate described
above were attributed by the original authors to changes in plate
motions (ca. 43, 20, and 5 Ma), underplating of subducted Yakutat
material (ca. 15 Ma), and collision of the thicker, southeastern part of
the Yakutat microplate with southern Alaska (between ca. 25 and
10 Ma). While more low-temperature thermochronologic data are
needed in southern Alaska to fully understand the spatial pattern of
Cenozoic rock uplift, integration of the existing datasets clearly shows
significant upper plate exhumation beginning in late Eocene time and
continuing through Neogene time, as well as a potential general
northwestward younging pattern of rock uplift and exhumation.
Exhumation in the upper plate of the flat-slab region broadly
overlapped with cessation of magmatism recorded by geochronometry (Fig. 4).
2.3. Stratigraphy
2.3.1. Data
Stratigraphic data document a well-defined depositional hiatus
across most of the flat-slab region during Miocene–Holocene time.
Remnant sedimentary basins in the flat-slab region stopped accumulating clastic strata and were deformed by reverse faults and folds. For
example, folded middle to late(?) Oligocene and older sedimentary
basin strata in the Matanuska Valley region are unconformably
overlain by Quaternary surficial deposits (Figs. 1, 6A and 7; Trop et al.,
2003). North of the flat-slab region (~130 km), early Oligocene and
older sedimentary strata along the south flank of the Alaska Range are
unconformably overlain by Quaternary surficial deposits (CB on Figs. 1
and 7; Fig. 6B; Trop et al., 2004).
2.3.2. Interpretation
We interpret this regional stratigraphic hiatus to represent late
Oligocene and younger basin inversion associated with flat-slab
subduction beneath Alaska. Folds and reverse faults regionally deform
these Oligocene and older sedimentary strata (Kortyna et al., 2009;
Pavlis and Roeske, 2007; Trop et al., 2004), consistent with Neogene
contractile deformation coeval with basin inversion. Neogene shortening and inversion of basinal strata in the upper plate of the flat-slab
region began shortly after cessation of magmatism recorded by
geochronometry (Fig. 4) and rock uplift and erosion recorded by
thermochronometry (Fig. 5).
3. Processes around flat-slab perimeter
Sedimentary basins located along the western and northern
perimeter of the flat-slab region record accumulation of thick, clastic
wedges coeval with exhumation and basin inversion above the flat slab.
5
Integration of published stratigraphic and compositional data with new
provenance data help link the depositional history of the perimeter
basins to exhumation of the upper plate in the flat-slab region. Along the
eastern perimeter, volcanic strata are the product of transtensional
tectonics and partial melting of the eastern edge of the flat slab.
3.1. Western perimeter
The Cook Inlet basin is located along the western perimeter of the
flat-slab region (Fig. 1), where Cenozoic nonmarine depositional
environments deposited up to 10 km of strata in parts of the basin
(Flores et al., 2004). Paleogene strata (an unnamed unit, West
Foreland and Hemlock Formations) reach a maximum combined
thickness of ~ 2.5 km, whereas mainly Neogene strata (Tyonek,
Beluga, and Sterling Formations) are much thicker (up to 7.5 km)
and contain coal seams up to 15 m thick (Figs. 7 and 8A–B). Sediment
accumulation rates increased beginning in late Oligocene time
continuing through Pliocene time (Fig. 9).
Interpretation of new provenance data suggests that Cook Inlet
detritus was derived predominantly from source terranes located
northeast of the basin above the flat-slab region and areas to the north
in the central Alaska Range during Eocene through Pliocene time. The
U-Pb ages of 590 detrital zircons from Eocene through Pliocene strata
of the Cook Inlet basin contain significant middle Cretaceous (ca. 94–
111 Ma) and Late Cretaceous–Paleogene (ca. 47–72 Ma) age populations (Fig. 10). The low ratios of uranium–thorium (b10) in most of
the grains (~ 98%; Fig. 10) suggest that they were derived from
igneous rather than metamorphic source areas. The Late Cretaceous–
Paleogene detrital ages match the age of plutons exposed in the
western and central Alaska Range and the Talkeetna Mountains
(Wilson et al., 1998; 2009). The greater regional thickness of the late
Oligocene–early Miocene and late Miocene–Pliocene strata relative to
the rest of the Cenozoic units, as well as the documentation of coeval
exhumation events in the western and central Alaska Range, has
previously been inferred to reflect regional exhumation and increased
sediment influx from the central and western Alaska Range during
these times (Fitzgerald et al., 1995; Haeussler, 2008; Kirschner and
Lyon, 1973; Plafker et al., 1992). Comprising a conspicuously small
population in our detrital data, however, are 30–45 Ma detrital zircon
ages (gray bar on Fig. 10) that would match the ages of plutons that
are common in the western Alaska Range (Moll-Stalcup, 1994; Wilson
et al., 2009), precluding that region as a dominant sediment source.
The dominant peak of Middle Cretaceous detrital zircon ages (Fig. 10)
closely matches the ages of plutons that are widespread in the central
Alaska Range and the Talkeetna Mountains, both located above the
flat-slab region; similar age plutons are very sparsely scattered in the
western Alaska Range (Wilson et al., 2009). The detrital populations
also contain several distinct ages (such as ca. 210, 194, 156–176, 104–
143, 80, and 66–52 Ma in Fig. 10) that are interpreted to best reflect
sediment input from source rocks in the flat-slab region. Key potential
source areas for these populations in the flat-slab region include the
Talkeetna arc and Alaska–Talkeetna Range belt in the Talkeetna
Mountains, as well as the Chitina and Chisana arcs. Plutons representing
the Mesozoic Chitina and Chisana arc systems are located in the eastern
Alaska Range and eastern Chugach Mountains in the eastern part of the
flat-slab region (Fig. 1; Miller, 1994; Nokleberg et al., 1994; Plafker et al.,
1989; Roeske et al., 2003; Short et al., 2005; Snyder and Hart, 2005,
2007). Additional possible sources located above the flat-slab region
that have ages that overlap with age populations in our samples include
accretionary prism strata of the Valdez Group in the western Chugach
Mountains (Bradley et al., 2009), and exhumed Cretaceous–Paleogene
strata in the Matanuska Valley (Fig. 1; Kortyna et al., 2010; Trop, 2008).
For more details on the ages and distribution of potential sources for
Cenozoic sediment of the Cook Inlet basin see Finzel (2010).
We suggest the existence of two major sediment pathways derived
from above the flat-slab region based on our provenance data. One
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
6
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
Fig. 6. Photos of Eocene–Oligocene sedimentary strata that record basin inversion and crustal shortening above the flat-slab region during Miocene–Holocene time. Black tadpole
symbols denote bedding. A) Eocene fluvial-lacustrine strata (Arkose Ridge Formation) deformed by asymmetric folds and reverse faults in the southern Talkeetna Mountains in the
remnant Matanuska basin (MB on Fig. 1). Height of outcrop is N150 m. B) Oligocene fluvial-lacustrine strata (Oc) exposed in a footwall syncline adjacent to a reverse fault along the
south flank of the central Alaska Range in the remnant Colorado Creek basin (CB on Fig. 1). Jurassic–Cretaceous marine strata (KJk) are exposed in the hanging wall of the reverse
fault. Height of outcrop with Oligocene strata is N 200 m. Snow covered peak in upper left side of photo is Mt. McKinley (elevation 20, 320 ft.); the Yakutat slab is located ~ 150 km
beneath this mountain based on geophysical imaging.
pathway predominantly drained source areas exposed north and
northeast of the basin, including the central Alaska Range (Fig. 1).
Another key pathway derived sediment from eastern and northeastern
source regions, including the Talkeetna Mountains (Fig. 1). We infer that
exhumation of the flat-slab region resulted in relative subsidence in the
Cook Inlet basin, created prolific sediment sources to the east and
Fig. 7. Late Eocene–Quaternary stratigraphy of sedimentary basins in south-central Alaska (1—Flores et al., 2004; 2–4—Trop and Ridgway, 2007; 5—Richter et al., 1990 and Trop et al.,
2007). Note lack of sedimentary strata deposited after Oligocene time in the region above flat-slab subduction and increase in thicknesses of post-Oligocene sedimentary strata in
basins located along western and northern perimeter of flat-slab subduction. Also note northwestward younging of volcanism and sedimentation in the Wrangell volcanic belt
derived from abundant geochronologic and palynological data (black vertical lines). Quaternary strata in the Cook Inlet and Tanana basins are unnamed. Time scale of Walker and
Geissman (2009).
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
7
Fig. 8. Photos of Neogene sedimentary strata that record enhanced sediment accumulation along the perimeter of the flat-slab region. A) and B) Strata deposited in the Cook Inlet
basin located along the western perimeter of the flat-slab region (CB on Fig. 1). A) Miocene fluvial sandstone, mudstone, and coal (Beluga Formation). Person circled for scale.
B) Pliocene fluvial sandstone and mudstone (Sterling Formation). Note the thick channel complex between the white dashed lines. Arrow points to person for scale. C), D), and
E) Strata deposited in the Tanana transpressional foreland basin located along the northern margin of the flat-slab region (TB on Fig. 1). C) Miocene coal, sandstone, and mudstone
(Usibelli Group) deposited in fluvial channels and floodplain mires. Coal beds are up to 20 m thick. Spruce forest at bottom of photo for scale. D) Miocene lacustrine mudstone with
seasonal varves (Usibelli Group) deposited in floodplain lakes. Person (lower right) for scale. E) Pliocene conglomerate and sandstone (Nenana Gravel) deposited in alluvial fans.
Person (lower right) for scale. F) Miocene strata deposited in transtensional basins along the eastern margin of the flat-slab region in the Wrangell volcanic belt (WVB on Fig. 1).
Miocene alluvial-lacustrine conglomerate, mudstone, tuff, and lava flows (Mf-Frederika Formation) are overlain by lava flows and volcanic breccia (Mw-Wrangell Lava). Height of
outcrop is N 700 m. Black tadpole symbols denote bedding. G) Mount Wrangell, a 4317-m (14,163 ft)-high andesite shield volcano in the Wrangell Mountains with documented
historical activity. View is to the northeast. Photograph by B. Cella, U.S. National Park Service, 1987.
northeast of the basin, and led to southwestward progradation of a large
clastic wedge away from the region presently above flat-slab subduction
and into the basin (Finzel, 2010; Finzel et al., 2007).
An interpreted northern extension of the Cook Inlet basin, the
Susitna basin (Fig. 1; Haeussler, 1998; Haeussler et al., 2002), also
appears to contain a thick section of Cenozoic strata. Very little is
known about the age and stratigraphy of the basin, but based on three
drill holes and gravity data there may be as much as 4 km of
Paleocene–Miocene nonmarine clastic strata (Merritt, 1986; Meyer
and Boggess, 2003). Much more data is needed, but we tentatively
interpret the Paleocene–Miocene strata of the Susitna basin to be part
of the large clastic wedge along the western perimeter of flat-slab
subduction.
3.2. Northern perimeter
The Tanana basin is positioned directly north of the flat-slab region
(Fig. 1). Along the southern margin of the basin, Cenozoic strata are
2–3 km thick and have been exhumed by thrust faults that form the
foothills along the north side of the Alaska Range (Ridgway et al.,
2002, 2007). Recent studies clearly show that the foothills are part of
an active northward-propagating thrust belt that is deforming
Neogene strata of the Tanana foreland basin (Bemis and Wallace,
2007; Lesh and Ridgway, 2007). The Usibelli Group consists of 970 m
of mainly middle and upper Miocene strata that were deposited in
fluvial, lacustrine, and peat bog environments (Fig. 7). In one part of
the Tanana basin, the Rex Creek area, the oldest part of the Usibelli
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
8
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
at least Late Eocene time and continues to the present (e.g., EberhartPhilips et al., 2006; Ridgway et al., 1995, 2002; Ruppert et al., 2008;
Trop et al., 2004).
3.3. Eastern perimeter
Fig. 9. Maximum stratigraphic thickness (Flores et al., 2004) versus time plot for
Paleocene to Pliocene strata in Cook Inlet basin. Note that accumulation rates increased
significantly during Oligocene time and continued to increase through Pliocene time.
All formation ages are constrained by biostratigraphy, including palynology and leaf
macrofossils (Kirschner and Lyon, 1973; Magoon et al., 1976; Wolfe and Tanai, 1980),
with additional age control in the Beluga and Sterling formations derived from
geochronology of interbedded tuffs (Dallegge and Layer, 2004; Reinink-Smith, 1990;
Triplehorn et al., 1977; Turner et al., 1980). Time scale of Walker and Geissman (2009).
Group may extend into Late Eocene time (Wolfe and Tanai, 1980), but
at the type section and in most studied parts of the basin, the group is
entirely Miocene in age (Leopold and Liu, 1994; Ridgway et al., 2007).
Thick successions of lacustrine mudstone and coal seams N20-m-thick
imply regional basin subsidence during sediment accumulation
(Fig. 8C–D). The Nenana Gravel consists of 1200 m of Pliocene strata
that were deposited in northward-prograding alluvial-fan and
braidplain environments that record erosion of high topography in
the central Alaska Range (Fig. 8E). The available stratigraphic data
suggest that the Tanana basin formed in flexural response to
northward propagation of shortening along the southern margin of
the basin (Ridgway et al., 2007). We interpret the Tanana basin as a
transpressional foreland basin because its development has been
coeval with dextral strike-slip displacement on the Denali fault
system. Cenozoic displacement on the Denali fault system started by
3.3.1. Wrangell volcanic belt
The eastern perimeter of flat-slab subduction is bounded by the
Wrangell volcanic belt (WVB on Fig. 1) and active strike-slip faults. The
WVB consists of lava flows, lava domes, pyroclastic strata, and
nonmarine sedimentary strata with maximum preserved thicknesses
N3000 m (Fig. 8F–G; Richter et al., 1990; Skulski et al., 1992). Strata crop
out near regional northwest-striking strike-slip faults (Denali, Duke
River, and Totschunda faults on Fig. 1) and are locally cut by northstriking normal faults (Eisbacher and Hopkins, 1977; MacKevett, 1978;
Richter et al., 2006). Extensive geochronologic and palynologic data
document northwestward younging of WVB strata (Fig. 7; Richter et al.,
1990; Ridgway et al., 1995; Skulski et al., 1992). The southeastern WVB
consists of Eocene–Upper Miocene alluvial-lacustrine sedimentary
strata and eruptive centers with mainly transitional and minor alkaline
and calc-alkaline geochemical compositions (Cole and Ridgway, 1993;
Ridgway and DeCelles, 1993; Skulski et al., 1991). The northwestern
WVB is characterized by Middle Miocene–Holocene alluvial-lacustrine
sedimentary strata and lavas with transitional to calc-alkaline geochemical compositions typical of subduction-related volcanic suites
(Preece and Hart, 2004; Trop et al., 2007).
We interpret the WVB stratigraphic record as the product of
transtensional tectonics and partial melting along the eastern edge of
the shallow Yakutat slab. We suggest that northwestward insertion of
a shallow slab beneath Alaska prompted northwestward-migrating
eruptive centers and narrow (b20 km wide) transtensional basins
along northwest-striking strike-slip faults (Totshunda and Duke
River faults on Fig. 1). Stratigraphic evidence for lacustrine deposition and coal formation indicate active subsidence within basins
during deposition. The orientation of north-striking normal faults is
consistent with east–west extensional subsidence related to dextral
displacement along adjacent strike-slip faults (Trop et al., 2007). The
lack of geophysical evidence for a present-day subducted slab deeper
Fig. 10. Age probability plot showing the zircon age distributions for seven sandstone samples from middle Eocene–Pliocene strata in Cook Inlet basin and age ranges for potential
source terranes in southern Alaska. Histograms are in 10 m.y. intervals. A total of 634 grains were analyzed via laser ablation ICP mass spectrometry at the University of Arizona
LaserChron Center. All b 250 Ma ages are plotted (n = 590). See Finzel (2010) for methods and raw age data. Inset is a plot of U/Th versus age for all 590 grains. Note that ~ 98% of the
grains have U/Th values N10. See Fig. 1 for sample locations and general geographic locations of source terranes described below. Potential source terranes located west of the flatslab region include the Western Alaska Range (WAR on Fig. 1), where the Aleutian Arc (AA; 0–10 Ma) and the Western Alaska Range igneous belt (WAR; 30–45 Ma) crop out.
Potential igneous source terranes located above and east of the flat-slab region include the Caribou Creek-Central Talkeetna Mountains volcanics (CTM; 35–60 Ma), the Central
Alaska Range belt (CAR; 70–120 Ma, CAR on Fig. 1), and the Chisana (CS; 105–140 Ma) and Chitina (CT; 135–175 Ma) arcs (both located between the St. Elias Mountains and the
Denali fault on Fig. 1). The Alaska Range–Talkeetna Mountains batholith (ATM; 55–74 Ma) and the Talkeetna Arc (TA; 153–201 Ma) occur in both areas, spanning the northwest side
of Cook Inlet to the Talkeetna Mountains. References for source terrane age data are discussed in Section 3.1.
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
than 50 km beneath the WVB (Figs. 2 and 3A; Eberhart-Philips et al.,
2006; Qi et al., 2007) together with the presence of adakite compositions in volcanic rocks from Mounts Drum and Churchill in the
WVB (Fig. 1; Preece and Hart, 2004) indicate partial melting along the
northeastern edge of the Yakutat slab. Adakite melts commonly form
by slab melting when the downgoing slab is young (b20–30 Ma) and
at shallow depths (b85 km), the dip of the downgoing plate is very
shallow, or the plate margin is characterized by elevated shear
stresses, highly oblique subduction, or slow subduction (Drummond
et al., 1996; Gutscher et al., 2000; Yogodzinski et al., 1995, 2001).
Spatial variations in WVB geochemical compositions imply magmatism mainly along extensional leaky strike-slip faults in the southeast
(Skulski et al., 1991, 1992) and increased flux of slab-derived
constituents into the mantle source in the northwest (Preece and
Hart, 2004), consistent with northwestward-directed subduction of
the flat slab along the curved continental margin of eastern Alaska.
3.3.2. Copper River basin
The Copper River basin lies between the Talkeetna Mountains and
the WVB along the eastern perimeter of the flat-slab region (Fig. 1).
Exploratory wells encountered 580 m of Cenozoic strata and 1715 m
of Upper Jurassic–Upper Cretaceous strata (Alaska Geological Society,
1970a,b). Subsurface data, including seismic profiles and aeromagnetic and gravity surveys, indicate a maximum thickness of 1200 m
for the Cenozoic strata and 7000 m for the Mesozoic strata (Fuis and
Plafker, 1989; Meyer and Boggess, 2003; Meyer and Saltus, 1995). The
specific ages of the Cenozoic strata in this basin are poorly known
making it difficult to determine any possible relationship between
timing of deposition and possible flat-slab processes.
4. Discussion
4.1. Upper plate record of flat-slab subduction in Alaska
We propose that a flat subduction style initiated in southern
Alaska ca. 20 m.y. earlier than previously inferred and that the Yakutat
microplate has been subducting at a shallow angle since the initiation
of its subduction beneath Alaska in late Eocene–early Oligocene time.
This premise is supported by recently published geophysical data that
document thick oceanic crust for the entire Yakutat slab, together
with our integrated geochronologic, thermochronologic, stratigraphic, and provenance data from the region presently above and around
flat-slab subduction in southern Alaska. Our data provide constraints
on the timing of subduction-related magmatism, enhanced rock uplift
and erosion, and development of sedimentary basins. In the flat-slab
region, exhumation initiated near the southern margin ca. 43 Ma
(Eocene) and continues through Neogene time throughout the entire
flat-slab region, magmatism ceased at ca. 32 Ma (Oligocene), and
deposition in sedimentary basins ended by ca. 23 Ma (Miocene).
Sedimentary basins positioned along the western and northern
perimeter of the flat-slab region record enhanced sediment accumulation rates and sediment delivery from bedrock sources exhumed
above the flat-slab region beginning in late Oligocene and middle
Miocene time respectively. Subduction-related volcanism and basin
development along strike-slip faults that bound the eastern perimeter
of the flat-slab region initiated ca. 26 Ma (Oligocene) and migrated
northwestward during Miocene–Holocene time. Collectively, we
interpret these age constraints, which were derived from different
proxies, to represent late Eocene–early Oligocene initiation and
continued shallow subduction of the Yakutat microplate. All these
lines of evidence are consistent with flat-slab processes being
important from Oligocene to present time along the southern margin
of Alaska.
An alternative hypothesis to flat-slab subduction of the Yakutat
microplate beginning during Oligocene time is that shallow subduction may have been a lingering product of subduction of a spreading
9
ridge beneath southern Alaska ca. 61–50 Ma. Oriented subparallel to
the paleo-trench, the spreading ridge subducted obliquely from west
to east beneath southern Alaska judging by the diachronous ages of
near-trench plutons related to slab-window magmatism (e.g.,
Haeussler et al., 2003). Adjacent to the spreading ridge, there would
have been very young oceanic crust subducting at the trench along the
southern margin of Alaska that presumably would have been
relatively warm and buoyant. An argument against the ridge
subduction hypothesis comes from our compilation of geochronologic
data that indicates that subduction-related magmatism resumed in
southern Alaska after the spreading ridge passed sometime after
50 Ma (Fig. 4). Therefore, whatever the age of the crust subducting
beneath southern Alaska between ca. 50 and 43 Ma, it was subducting
at a steep enough angle to produce a volcanic arc. The presence of arc
magmatism after early Paleogene ridge subduction suggests that ridge
subduction is a less likely explanation for the upper plate processes
documented in our databases.
4.2. A larger Yakutat microplate?
Comparison of geochronologic, thermochronologic, and stratigraphic data from above and to the west of the present-day flat-slab
region allows for a reevaluation of models on the kinematic history of
the Yakutat microplate. Here, we hypothesize that the original leading
edge of the Yakutat microplate may have been wider and extended
farther west along strike compared to the current geometry. Similar to
the present-day flat-slab region, published reconnaissance thermochronologic data from areas to the west also record a possible general
northward younging of the onset of rock exhumation from the Kenai
Peninsula towards the western Alaska Range northwest of Cook
Inlet (Fig. 1). In the Chugach Mountains on the Kenai Peninsula,
apatite (U-Th)/He ages record exhumation of the accretionary prism
between ca. 40 and 12 Ma (8 on Fig. 1; Buscher et al., 2008). Farther
inboard (~200 km), apatite FT ages from Paleogene granite in the
western Alaska Range record rapid exhumation at ca. 23 and 6 Ma, with
one apparently anomalous sample recording rapid cooling at ca. 35 Ma
(9 on Fig. 1; Haeussler et al., 2008). Geochronologic data from igneous
rocks indicate that subduction-related magmatism ceased in the
western Alaska Range ca. 30–23 Ma (Fig. 4; Moll-Stalcup, 1994) and
did not resume on the Alaska Peninsula until ca. 10 Ma (Reinink-Smith,
1990). Detrital data from Middle Miocene–Pliocene strata in the forearc
basin also demonstrate a lack of Early–Middle Miocene aged igneous
sources (Fig. 10). Farther to the west, Aleutian arc magmatism
synchronously waned ca. 30 Ma and resumed ca. 16–11 Ma (Jicha
et al., 2006). If the same tectonic process is responsible for upper plate
processes above and inboard of the present-day flat-slab region as
well as for these areas to the west, namely flat-slab subduction, then
it is possible that shallow subduction of the Yakutat microplate
impacted a larger extent of southern Alaska in the past than what is
evident today.
An alternative travel history for the Yakutat microplate has been
previously proposed based on plate reconstructions and seismic
imaging of accretionary prism strata near the Kenai Peninsula (Fruehn
et al., 1999). In that model, the Yakutat microplate is assumed to be
coupled with the Pacific plate since at least Pliocene time. Using
published plate motion models, these authors propose that extraction
of the Yakutat microplate from beneath Alaska would cause its
southwestern trailing edge, coincident with the Transition fault, to
pass beneath the accretionary prism outboard of the Kenai Peninsula
from ~ 3.5 to 2 Ma. Evidence for passage of a buoyant Yakutat slab
through this region includes truncation and erosion of pre-Neogene
accretionary prism strata and expansion of the Neogene accretionary
prism southward away from the present-day southwestern boundary
of the Yakutat microplate (Fig. 1). This model differs from the Plafker
(1987) model in that the original western extent of North America–
Yakutat interaction would have extended much farther west, at least
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
10
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
to the southern end of the Kenai Peninsula, the flat Yakutat slab would
have potentially passed beneath much of the Cook Inlet forearc basin
during middle to late Cenozoic time, and the Yakutat microplate
would not necessarily have been located adjacent to the western
North American margin while it traveled north toward Alaska.
Recent studies support the proposal by Plafker (1987) that the
southeastern unsubducted portion of the Yakutat microplate was
positioned along the western continental margin of North America
since Eocene time. For example, stratigraphic and detrital provenance
studies from Eocene and younger sedimentary strata exposed on the
microplate document deltaic-shallow marine deposystems that were
sourced from continental margin source terranes in southeastern
Alaska or western Canada (Landis, 2007; Perry et al., 2009). Similarly,
age data and geochemical compositions from igneous rocks in both
the microplate and adjacent portions of Alaska are attributable to
slab-window magmatism associated with subduction of a spreading
ridge beneath the continental margin during Eocene time, requiring
juxtaposition of the microplate with southeastern Alaska or western
Canada (Bradley et al., 2003; Davis and Plafker, 1986; Haeussler et al.,
2003).
We suggest a scenario that accounts for uplift and truncation of the
accretionary prism strata in the Kenai Peninsula region while keeping
the eastern, unsubducted portion of the slab adjacent to the North
American continental margin. We postulate that the original width of
the subducting slab may have been greater than the width presently
indicated by the unsubducted portion or the geophysically imaged
subducted slab. Published geophysical images have only loosely
constrained the nature of the boundary between the Pacific and
Yakutat plates at shallow depths (Eberhart-Philips et al., 2006). Thus,
the subducted portion of the slab could have been wider and therefore
interacted with a broader extent of the southern Alaska margin in the
past. Alternatively, the Yakutat slab may not have been wider but
instead stresses transmitted inboard from the flat-slab region may
have contributed to the deformation to the west. The eastward
decrease in the width of the accretionary prism towards the presentday flat-slab region (Fig. 1), however, suggests that whatever perturbed
the depositional system outside the present-day flat-slab region
traveled from west to east along the southern margin of Alaska.
Presumably, as the disturbance migrated to the northeast, sedimentation resumed in the accretionary prism, leading to the present-day
geometry of the prism as described above. In summary, the possibility
exists that the original width of the Yakutat microplate may have been
wider than its presently imaged dimensions, but much more data is
needed to test this hypothesis.
4.3. Comparison with other modern flat-slab margins
The widely distributed, diffuse deformation and the transpressional tectonic regime documented in southern Alaska are common
characteristics of regions undergoing oblique subduction of shallow
slabs. Insertion of a shallow slab into a subduction margin commonly
causes regional shortening and exhumation of the upper plate as a
result of the inboard propagation of plate boundary compressive
stresses and isostatic adjustments (Espurt et al., 2007, 2008; Gutscher,
2002). Earthquake focal mechanisms demonstrate that transpressive
stress regimes dominate many modern flat-slab margins, including
southern Alaska, New Guinea, northern Columbia, Japan, and
segments of the Andean margin (Gutscher et al., 2000; Haeussler,
2008). The transpressive nature and broad distribution of deformation
several hundred kilometers inboard from the plate margin in these
regions is attributable to coupling of the subducted flat slab to the
upper plate and oblique convergence between the slab and the upper
plate. This is consistent with enhanced contact between the two
plates and the cooler temperature due to lack of an asthenospheric
wedge (e.g., Gutscher, 2002). Transpressive shortening of the upper
plate prompts regional exhumation, a process that may be accom-
modated by discrete pre-existing structures or zones of weakness
(Espurt et al., 2007, 2008; Gutscher, 2002). Strong interplate coupling
in regions experiencing both flat-slab subduction and oblique
convergence may also lead to strain partitioning far inboard of the
subduction zone (Chemenda et al., 2000; Pabellier and Cobbold, 1996;
Pinet and Cobbold, 1992). For example, along the northwestern
Andean margin and in southwestern Japan, major strike-slip faults
occur ~300–400 km inboard of the trench. In southern Alaska, the
Castle Mountain fault accommodates ~ 3 mm/yr of dextral-oblique
strike-slip displacement ~250 km inboard of the trench (Willis et al.,
2007). Even farther inboard, the Denali fault accommodates ~10 mm/yr
of dextral-oblique strike-slip motion ~400 km inboard of the plate
margin (Fig. 1; Haeussler, 2008).
Flat-slab subduction occurs when relatively buoyant oceanic crust
enters a subduction zone and causes the slab to progressively flatten
(Espurt et al., 2008; Gutscher et al., 2000). Buoyant oceanic crust
results from two main characteristics: a moderate to young age of the
subducting crust (b50 m.y.) in combination with crust thicker than
~15 km. Thick oceanic crust often characterizes aseismic ridges and
oceanic plateaus. Two well-documented modern examples of thick
oceanic crust subducting beneath continental margins are the
aseismic Nazca Ridge in South America and the Ontong Java Plateau
in the south Pacific (Cowley et al., 2004; Espurt et al., 2007, 2008;
Hampel, 2002; Knesel et al., 2008; Mann and Taira, 2004; Phinney
et al., 2004; Taira et al., 2004; vonHuene et al., 1999). In these regions,
as well as in southern Alaska, flat-slab subduction of thick oceanic
crust has resulted in upper plate shortening and exhumation above
the flattened slab, cessation of magmatism within a few million years
of the onset of shallow subduction, and erosion and subsequent reestablishment of the accretionary prism during and after lateral
migration of the shallow slab.
The trench parallel widths and crustal thicknesses of the Nazca
Ridge and Yakutat microplate are comparable, and although the Nazca
Ridge has been subducting for only ca. 11–12 Ma compared to ca.
35 Ma for the Yakutat microplate inferred from this study, a higher
convergence rate in South America has led to similar amounts of
subducted material in the two regions. Both margins are characterized
by oblique convergence, which, in South America, has resulted in the
southward migration of the Nazca Ridge along the South American
trench (Hampel, 2002). In Alaska and South America, exhumation and
deformation in the forearc region was coincident with the arrival of
thick crust at the trench, magmatism ceased above the subducting flat
slab ca. 3–8 m.y. later, strong interplate coupling led to deformation
within the upper plate that extended hundreds of kilometers inboard,
and accretionary prism sedimentation resumed along strike after the
flat slab migrated laterally (Espurt et al., 2008; Hampel, 2002;
vonHuene et al., 1999). In South America, stratigraphic data record
uplift of forearc depocenters coincident with passage of the Nazca
Ridge.
The Ontong Java Plateau is the world's largest oceanic plateau,
covering about 1.9 million km2, with an average crustal thickness of
~30–35 km (Knesel et al., 2008; Mann and Taira, 2004). Approximately 1000 km of the trench parallel width of the plateau is currently
subducting along the North Solomon trench. The Ontong Java Plateau
has only been subducting for ca. 4–6 Ma, resulting in a total subducted
slab length of ~ 200–300 km (Cowley et al., 2004; Mann and Taira,
2004; Phinney et al., 2004; Taira et al., 2004). Prior to arrival of the
plateau at the subduction zone, Miocene and Pliocene arc volcanism
was prevalent in the Solomon Islands region. After the plateau arrived
at the trench ca. 4–6 Ma, arc magmatism began to wane and is less
common today than it was prior to Pliocene time, major folding and
uplift occurred in the accretionary prism, and intra-arc normal faults
were reactivated as thrust faults. Presently, convergence between the
Pacific plate and Australian plate is accommodated at both the North
Solomon trench and a young northeast-dipping subduction zone that
has developed in the backarc region.
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
Several important characteristics are shared among the Alaska,
South American, and south Pacific flat-slab settings. First, shallow
subduction of buoyant oceanic crust led to increased plate coupling,
which resulted in upper plate shortening, rock uplift and exhumation
above the flat-slab region. Second, subduction-related magmatism
waned or ceased b10 m.y. after the onset of shallow subduction.
Third, shoaling or inversion of basins above the flat-slab region led to
an increase in coarser sedimentation and/or a lack of deposition.
Fourth, the accretionary prism was deformed or eroded above the flat
slab, and in the case of Alaska and South America, sedimentation
within the prism was reestablished after the shallow slab passed
laterally along the margin. The most distinct difference between
Alaska and these other regions is that the northeastern boundary of
the Yakutat microplate is a strike-slip margin. The effect on the upper
plate of this boundary has been the propagation of transtensional
tectonics along the northeastern edge of the subducting slab. This is
reflected in the northwestward younging of slab-edge volcanism and
basin development along the Denali–Duke River strike-slip fault
system in the WVB (Figs. 1 and 7).
5. Conclusions
Integration of thermochronologic, geochronologic, stratigraphic,
and provenance data from the upper plate area above and around flatslab subduction of the Yakutat microplate in southern Alaska provides
evidence that flat-slab processes have been occurring since late
Eocene–early Oligocene time. Above the flat slab, regional exhumation initiated ca. 43 Ma, magmatism ceased ca. 32 Ma, and deposition
in sedimentary basins ended ca. 23 Ma. Along the western and
northern perimeter of the flat-slab region, sedimentary basins
experienced enhanced subsidence and sediment delivery from the
flat-slab region beginning in late Oligocene and middle Miocene time,
respectively. In the Wrangell volcanic belt along the eastern margin of
the flat-slab region, volcanism and basin development along dextral
strike-slip faults initiated ca. 26 Ma (Oligocene) and migrated
northwestward during Miocene–Holocene time. We propose that
these data support shallow subduction of the buoyant Yakutat
microplate starting in late Eocene–early Oligocene time. Furthermore,
stratigraphic and thermochronologic data from the Kenai Peninsula
and western Alaska Range, west of the present-day flat-slab region,
indicate flat-slab processes may have been active in those regions
previously. Consequently, we postulate that the subducted portion of
the Yakutat microplate may have extended farther west along strike
and affected a larger region of southern Alaska than the width
presently indicated by the unsubducted portion of the Yakutat
microplate or the geophysically imaged subducted slab. Modern
convergent margins influenced by flat subduction styles are shaped by
processes similar to those observed in southern Alaska, including
regional exhumation, cessation of magmatism, shoaling or inversion
of sedimentary basins, and deformation or erosion of the accretionary
prism above the flat-slab region.
Supplementary materials related to this article can be found online
at doi:10.1016/j.epsl.2011.01.014.
Acknowledgements
Primary funding was provided by the National Science Foundation,
Donors of the Petroleum Research Fund administered by the
American Chemical Society, and the Bucknell University Program for
Undergraduate Research. Reviewers S. Roeske, P. Haeussler, and R.
Dorsey offered constructive reviews that helped us improve the
manuscript. We thank A. Till, W. Wallace, and B. McNulty for reviews
of an earlier version of the manuscript. We also thank Bucknell
University undergraduate students E. Bauer, R. Delaney, C. Kortyna, R.
Tidmore, and J. Witmer for field assistance and thesis research.
11
References
Alaska Geological Society, 1970a. Copper River basin stratigraphic correlation section;
Tawawe Lake to Moose Creek, Copper River basin, Alaska, 1 sheet.
Alaska Geological Society, 1970b. Copper River basin stratigraphic correlation section;
Eureka to Rainbow, 1 sheet.
Arkle, J.C., Armstrong, P.A., Haeussler, P.J., 2009. The western Chugach Mountains and
northern Prince William Sound (Alaska): locus of subduction-related exhumation?
Geol. Soc. Am. Abstr. Programs 41, 290.
Bemis, S.P., Wallace, W.K., 2007. Neotectonic framework of the north-central Alaska
Range foothills. In: Ridgway, K.D., Trop, J.M., Glen, J.M.G., O'Neill, J.M. (Eds.),
Tectonic Growth of a Collisional Margin: Crustal Evolution of Southern Alaska:
Geological Society of America Special Paper, 431, pp. 549–572.
Benowitz, J.A., Layer, P., Armstrong, P., Perry, S., Haeussler, P., Fitzgerald, P., in press.
VanLaningham, S. Spatial variations in focused exhumation along a continental-scale strikeslip fault: the Denali fault of the eastern Alaska Range. Geosphere. Manuscript GS589R2.
Berger, A.L., Spotila, J.A., 2008. Denudation and deformation in a glaciated orogenic
wedge: the St. Elias orogen, Alaska. Geology 36, 523–526.
Berger, A.L., Spotila, J.A., Chapman, J.B., Pavlis, T.L., Enkelmann, E., Ruppert, N.A.,
Buscher, J.T., 2008. Architecture, kinematics, and exhumation of a convergent
orogenic wedge: a thermochronological investigation of tectonic-climatic interactions within the central St. Elias orogen, Alaska. Earth Planet. Sci. Lett. 270, 13–24.
Bradley, D., Kusky, T., Haeussler, P., Goldfarb, R., Miller, M., Dumoulin, J., Nelson, S.W.,
Karl, S., 2003. Geologic signature of early Tertiary ridge subduction in Alaska. In:
Sisson, V.B., Roeske, S.M., Pavlis, T.L. (Eds.), Geology of a Transpressional Orogen
Developed During Ridge-trench Interaction Along the North Pacific Margin:
Geological Society of America Special Paper, 371, pp. 19–49.
Bradley, D.C., Haeussler, P.J., O'Sullivan, P., Friedman, R., Till, A.B., Bradley, D., Trop, J.M.,
2009. Detrital zircon geochronology of Cretaceous and Paleogene strata across the
south-central Alaskan convergent margin. U.S. Geological Survey Professional Paper
1760-F. 36 pp.
Buscher, J.T., Berger, A.L., Spotila, J.A., 2008. Exhumation in the Chugach-Kenai
Mountains belt above the Aleutian subduction zone, southern Alaska. In:
Freymueller, J.T., Haeussler, P.J., Wesson, R.J., Ekstrom, G. (Eds.), Active Tectonics
and Seismic Potential of Alaska: Geophysical Monograph Series, 179, pp. 151–166.
Cameron, C.E., 2005. Latitudes and Longitudes of Volcanoes in Alaska. Alaska Division of
Geological & Geophysical Surveys Raw Data File RDF 2005-3, 1 CD-ROM.
Chemenda, A., Lallemand, S., Bokun, A., 2000. Strain partitioning and interplate friction
in oblique subduction zones: constraints provided by experimental modeling.
J. Geophys. Res. Solid Earth 105, 5567–5581.
Christeson, G.L., Gulick, S.P.S., van Avendonk, H.J.A., Worthington, L.L., Reece, R.S., Pavlis,
T.L., 2010. The Yakutat terrane: dramatic change in crustal thickness across the
Transition fault, Alaska. Geology 38, 895–898.
Cole, R.B., Ridgway, K.D., 1993. The influence of volcanism on fluvial depositional
systems in a Cenozoic strike-slip basin, Denali fault system, Yukon Territory,
Canada. J. Sed. Petrol. 63, 152–166.
Cowley, S., Mann, P., Coffin, M.F., Shipley, T.H., 2004. Oligocene to recent tectonic
history of the Central Solomon intra-arc basin as determined from marine seismic
reflection data and compilation of onland geology. Tectonophysics 389, 267–307.
Dallegge, T.A., Layer, P.W., 2004. Revised chronostratigraphy of the Kenai Group from
40Ar/39Ar dating of low-potassium bearing minerals, Cook Inlet Basin, Alaska. Can.
J. Earth Sci. 41, 1159–1179.
Davis, A.S., Plafker, G., 1986. Eocene basalts from the Yakutat Terrane—evidence for the
origin of an accreting terrane in southern Alaska. Geology 14, 963–966.
Demets, C., Gordon, R.G., Argus, D.F., Stein, S., 1994. Effect of recent revisions to the
geomagnetic reversal time-scale on estimates of current plate motions. Geophys.
Res. Lett. 21, 2191–2194.
Dickinson, W.R., Snyder, W.S., 1979. Geometry of subducted slabs related to SanAndreas transform. J. Geol. 87, 609–627.
Doubrovine, P.V., Tarduno, J.A., 2008. A revised kinematic model for the relative
motion between Pacific oceanic plates and North America since Late Cretaceous.
J. Geophys. Res. Solid Earth 113. doi:10.1029/2008JB005585.
Drummond, M.S., Defant, M.J., Kepezhinskas, P.K., 1996. Petrogenesis of slab-derived
trondhjemite-tonalite-dacite/adakite magmas. Trans. R. Soc. Edinb. Earth Environ.
Sci. 87, 205–215.
Eberhart-Philips, D., Christensen, D.H., Brocher, T.M., Hansen, R., Ruppert, N.A.,
Haeussler, P.J., Abers, G.A., 2006. Imaging the transition from Aleutian subduction
to Yakutat collision in central Alaska, with local earthquakes and active source data.
J. Geophys. Res. Solid Earth 111. doi:10.1029/2005JB004240.
Eisbacher, G.H., Hopkins, S.L., 1977. Mid-Cenozoic paleogeomorphology and tectonic
setting of the St. Elias Mountains, Yukon Territory. Report of Activities, Geological
Survey of Canada paper 77-1B, pp. 319–335.
Engebretson, D.C., Cox, R.G., Gordon, R.G., 1985. Relative Motions Between Oceanic and
Continental Plates in the Pacific Basin. Geological Society of America Special Paper,
206. 59 pp.
Enkelmann, E., Garver, J.I., Pavlis, T.L., 2008. Rapid exhumation of ice-covered rocks of
the Chugach-St. Elias orogen, Southeast Alaska. Geology 36, 915–918.
Enkelmann, E., Zeitler, P.K., Pavlis, T.L., Garver, J.I., Ridgway, K.D., 2009. Intense
localized rock uplift and erosion in the St Elias orogen of Alaska. Nat. Geosci. 2,
360–363.
Enkelmann, E., Zeitler, P.K., Garver, J.I., Pavlis, T.L., Hooks, B.P., 2010. The thermochronological
record of tectonic and surface process interaction at the Yakutat–North American collision
zone in southeast Alaska. Am. J. Sci. 310, 231–260.
Espurt, N., Baby, P., Brusset, S., Roddaz, M., Hermoza, W., Regard, V., Antoine, P.O., SalasGismondi, R., Bolanos, R., 2007. How does the Nazca Ridge subduction influence the
modern Amazonian foreland basin? Geology 35, 515–518.
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
12
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna, C., Brusset, S.,
2008. Flat subduction dynamics and deformation of the South American plate:
insights from analog modeling. Tectonics 27. doi:10.1029/2007TC002175.
Ferris, A., Abers, G.A., Christensen, D.H., Veenstra, E., 2003. High resolution image of the
subducted Pacific (?) plate beneath central Alaska, 50–150 km depth. Earth Planet.
Sci. Lett. 214, 575–588.
Finzel, E.S., 2010. Geodynamics of flat-slab subduction, sedimentary basin development,
and hydrocarbon systems along the southern Alaska convergent plate margin. [Ph.D.]
Purdue University, 411 pp.
Finzel, E.S., Ridgway, K.D., Brennan, P.R.K., Landis, P., 2007. Miocene and Pliocene
sedimentary footprint of flat-slab subduction of the Yakutat terrane. Geol. Soc. Am.
Abstr. Programs 39, 491.
Fitzgerald, P.G., Sorkhabi, R.B., Redfield, T.F., Stump, E., 1995. Uplift and denudation of
the central Alaska Range—a case-study in the use of apatite fission-track
thermochronology to determine absolute uplift parameters. J. Geophys. Res. Solid
Earth 100, 20175–20191.
Flores, R.M., Stricker, G.D., Kinney, S.A., 2004. Alaska Coal Geology, Resources, and
Coalbed Methane Potential. U.S. Geological Survey DDS-77. 125 pp., 3 sheets.
Fruehn, J., vonHuene, R., Fisher, M.A., 1999. Accretion in the wake of terrane collision: the
Neogene accretionary wedge off Kenai Peninsula, Alaska. Tectonics 18, 263–277.
Fuis, G.S., Plafker, G., 1989. Evolution of deep structure along the Trans-Alaska Crustal
Transect, Chugach Mountains and Copper River basin, southern Alaska. J. Geophys.
Res. 96, 4229–4253.
Fuis, G.S., Moore, T.E., Plafker, G., Brocher, T.M., Fisher, M.A., Mooney, W.D., Nokleberg,
W.J., Page, R.A., Beaudoin, B.C., Christensen, N.I., Levander, A.R., Lutter, W.J., Saltus,
R.W., Ruppert, N.A., 2008. Trans-Alaska Crustal Transect and continental evolution
involving subduction underplating and synchronous foreland thrusting. Geology
36, 267–270.
Gulick, S., Lowe, L., Pavlis, T., Gardner, J., Mayer, L., 2007. Geophysical insights into the
Transition fault debate: propagating strike slip response to stalling Yakutat block
subduction in the Gulf of Alaska. Geology 35, 763–766.
Gutscher, M.A., 2002. Andean subduction styles and their effect on thermal structure
and interplate coupling. J. S. Am. Earth Sci. 15, 3–10.
Gutscher, M.A., Spakman, W., Bijwaard, H., Engdahl, E.R., 2000. Geodynamics of flat
subduction: seismicity and tomographic constraints from the Andean margin.
Tectonics 19, 814–833.
Haeussler, P.J., 1998. Surficial Geologic Map Along the Castle Mountain Fault Between
Houston and Hatcher Pass Road, Alaska. U.S. Geological Survey Open-File Report 98–480.
Haeussler, P.J., 2008. An overview of the neotectonics of interior Alaska: far-field
deformation from the Yakutat microplate collision. In: Freymueller, J.T., Haeussler,
P.J., Wesson, R.J., Ekstrom, G. (Eds.), Active Tectonics and Seismic Potential of
Alaska: Geophysical Monograph Series, 179, pp. 269–285.
Haeussler, P.J., Best, T.C., Waythomas, C.F., 2002. Paleoseismology at high latitudes:
seismic disturbance of Late Quaternary deposits along the Castle Mountain fault
near Houston, Alaska. Geol. Soc. Am. Bull. 114, 1296–1310. doi:10.1130/0016-7606
(2002) 114, 1296: PAHLSD.2.0.CO;2.
Haeussler, P.J., Bradley, D.C., Wells, R.E., Miller, M.L., 2003. Life and death of the
Resurrection plate: evidence for its existence and subduction in the northeastern
Pacific in Paleocene–Eocene time. Geol. Soc. Am. Bull. 115, 867–880.
Haeussler, P.J., O'Sullivan, P., Berger, A.L., Spotila, J.A., 2008. Neogene exhumation of the
Tordrillo Mountains, Alaska, and correlations with Denali (Mount McKinley). In:
Freymueller, J.T., Haeussler, P.J., Wesson, R.J., Ekstrom, G. (Eds.), Active Tectonics
and Seismic Potential of Alaska: Geophysical Monograph Series, 179, pp. 269–285.
Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian active
margin: a re-evaluation. Earth Planet. Sci. Lett. 203, 665–679.
Hoffman, M.D., Armstrong, P.A., 2006. Miocene exhumation of the southern Talkeetna
Mountains, south central Alaska, based on apatite (U-Th)/He thermochronology.
Geol. Soc. Am. Abstr. Programs 38, 9.
Jicha, B.R., Scholl, D.W., Singer, B.S., Yogodzinski, G.M., Kay, S.M., 2006. Revised age of Aleutian
Island Arc formation imples high rate of magma production. Geology 38, 661–664.
Jordan, T.E., Allmendinger, R.W., 1986. The Sierras Pampeanas of Argentina—a modern
analog of Rocky-Mountain foreland deformation. Am. J. Sci. 286, 737–764.
Kirschner, C.E., Lyon, C.A., 1973. Stratigraphic and tectonic development of Cook Inlet
petroleum province. AAPG Bull. Mem. 19, 396–407.
Knesel, K.M., Cohen, B.E., Vasconcelos, P.M., Thiede, D.S., 2008. Rapid change in drift of the
Australian plate records collision with Ontong Java plateau. Nature 454, 754–758.
Koons, P.O., Hooks, B.P., Pavlis, T., Upton, P., Barker, A.D., 2010. Three-dimensional
mechanics of Yakutat convergence in the southern Alaskan plate corner. Tectonics
29. doi:10.1029/2009TC002463.
Kortyna, C.D., Trop, J.M., Lecomte, A.A., Bauer, E.M., Kassab, C.M., Ridgway, K.D.,
Sunderlin, D., 2009. Sedimentology, paleontology, and structural framework of the
central Arkose Ridge Formation, Talkeetna Mountains, Alaska. Geol. Soc. Am. Abstr.
Programs 41, 304.
Kortyna, C.D., Trop, J.M., Idleman, B., Kassab, C.M., Ridgway, K.D., Gehrels, G., 2010.
Provenance signature of a forearc basin modified by spreading ridge subduction:
detrital zircon geochronology and detrital modes from the Paleogene Arkose Ridge
Formation, southern Alaska. Geol. Soc. Am. Abstr. Programs 42, 54.
Lallemand, S., Heuret, A., Boutelier, D., 2005. On the relationships between slab dip,
back-arc stress, upper plate absolute motion, and crustal nature in subduction
zones. Geochem. Geophys. Geosyst. 6. doi:10.1029/2005GC000917.
Landis, P.S., 2007. Stratigraphic framework and provenance of the Eocene–Oligocene Kulthieth
Formation, Alaska: implications for paleogeography and tectonics of the early Cenozoic
continental margin of northwestern North America. [M.S.] Purdue University, 200 pp.
Leonard, L.J., Hyndman, R.D., Mazzotti, S., Nykolaishen, L., Schmidt, M., Hippchen, S.,
2007. Current deformation in the northern Canadian Cordillera inferred from GPS
measurements. J. Geophys. Res. Solid Earth 112. doi:10.1029/2007JB005061.
Leopold, E.B., Liu, G., 1994. A long pollen sequence of Neogene age: Alaska Range.
Quatern. Int. 22 (23), 103–140.
Lesh, M.E., Ridgway, K.D., 2007. Geomorphic evidence of active transpressional
deformation in the Tanana foreland basin, south-central Alaska. In: Ridgway, K.D.,
Trop, J.M., Glen, J.M.G., O'Neill, J.M. (Eds.), Tectonic Growth of a Collisional Margin:
Crustal Evolution of Southern Alaska: Geological Society of America Special Paper, 431,
pp. 573–592.
Little, T.A., Naeser, C.W., 1989. Tertiary tectonics of the Border Ranges fault system,
Chugach Mountains, Alaska—deformation and uplift in a fore-arc setting. J. Geophys.
Res. Solid Earth Planet. 94, 4333–4359.
Lowe, L.A., Gulick, S.P., Christeson, G.L., van Avendonk, H., Reece, R., Elmore, R., Pavlis, T.,
2008. Crustal structure and deformation of the Yakutat microplate: new insights
from STEEP marine seismic reflection data. EOS Trans. Am. Geophys. Union 89, 53
Fall Meeting Supplement, Abstract T53B-1941.
MacKevett, E.M., 1978, Geologic map of the McCarthy quadrangle, Alaska. U.S. Geological
Survey Miscellaneous Investigation Series I-1032, 1:250, 000.
Magoon, L. B., Adkinson, W. L., and Egbert, R. M., 1976. Map showing geology, wildcat wells,
Tertiary plant fossil localities, K–Ar age dates, and petroleum operations, Cook Inlet area,
Alaska. U.S. Geological Survey Miscellaneous Investigations Series Map I-1019, scale
1:250,000.
Mann, P., Taira, A., 2004. Global tectonic significance of the Solomon Islands and Ontong
Java Plateau convergent zone. Tectonophysics 389, 137–190.
McGeary, S., Nur, A., Ben-Avraham, Z., 1985. Spacial gaps in arc volcanism: the effect of
collision or subduction of oceanic plateaus. Tectonophysics 119, 195–221.
McNamara, D.E., Pasyanos, M.E., 2002. Seismological evidence for a sub-volcanic arc
mantle wedge beneath the Denali volcanic gap, Alaska. Geophys. Res. Lett. 29.
doi:10.1029/2001GL014088.
Merritt, R.D., 1986. Paleoenvironmental and tectonic controls in major coal basins of
Alaska. In: Lyons, P.C., Rice, C.L. (Eds.), Paleoenvironmental and Tectonic Controls in
Coal-Forming Basins in the United States: Geological Society of America Special
Paper, 210, pp. 173–200.
Meyer, J.F., Boggess, P.L., 2003. Principle Facts for Gravity Data Collected in the Susitna
Basin Area, South Central Alaska. Alaska Division of Geological and Geophysical
Surveys, Preliminary Investigative Report 2003–3. 13 pp.
Meyer, J.F., Jr., and Saltus, R.W., 1995. Merged aeromagnetic map of interior Alaska. U.S.
Geological Survey Map GP-1014, scale 1:500, 000, 2 sheets.
Miller, T.P., 1994. Pre-Cenozoic plutonic rocks in mainland Alaska. In: Plafker, G., Berg,
H.C. (Eds.), The Geology of Alaska: Geological Society of America, The Geology of
North America, pp. 535–554.
Moll-Stalcup, E.J., 1994. Latest Cretaceous and Cenozoic magmatism in mainland
Alaska. In: Plafker, G., Berg, H.C. (Eds.), Geol. Soc. Am. The Geol. N. Am. 589–619.
Nokleberg, W.J., Plafker, G., Wilson, F.H., 1994. Geology of the south-central Alaska. In:
Plafker, G., Berg, H.C. (Eds.), The Geology of Alaska: Geological Society of America,
The Geology of North America, pp. 311–366.
Nye, C.J., 1999. Magmatism at the eastern end of the Aleutian Arc: implications for the
western Alaska Range volcanic gap. In: Severin, B., Anderson, P.M. (Eds.), Science in
the north: 50 years of change: Arctic Science Conference, 50, Program and
Abstracts, Denali National Park and Preserve, AK, Sept 19–22, 1999, pp. 178–180.
Nye, C., Wyss, M., Ratchkovski, N., Fletcher, H., 2002. Magmatism in the Denali Volcanic
Gap, southern Alaska. EOS Trans. Am. Geophys. Union 83 (47) Fall Meeting
Supplement, Abstract V12C-06.
O'Sullivan, P.B., Currie, L.D., 1996. Thermotectonic history of Mt. Logan, Yukon Territory,
Canada: implications of multiple episodes of middle to late Cenozoic denudation.
Earth Planet. Sci. Lett. 144, 251–261.
Pabellier, M., Cobbold, P.R., 1996. Analogue models for the transpressional docking of
volcanic arcs in the Western Pacific. Tectonophysics 253, 33–52.
Parry, W.T., Bunds, M.P., Bruhn, R.L., Hall, C.M., Murphy, J.M., 2001. Mineralogy, 40Ar/
39Ar dating and apatite fission track dating of rocks along the Castle Mountain
fault, Alaska. Tectonophysics 337, 149–172.
Pavlis, T.L., Roeske, S.M., 2007. The Border Ranges Fault System, southern Alaska. In:
Ridgway, K.D., Trop, J.M., Glen, J.M.G., O'Neill, J.M. (Eds.), Tectonic Growth of a
Collisional Margin: Crustal Evolution of Southern Alaska: Geologic Society of
America Special Paper, 431, pp. 95–128.
Pavlis, T.L., Picornell, C., Serpa, L., Bruhn, R.L., Plafker, G., 2004. Tectonic processes during
oblique convergence: insights from the St. Elias Orogen, northern North American
Cordillera. Tectonics 23. doi:10.1029/2003TC001557.
Perry, S.E., Garver, J.I., Ridgway, K.D., 2009. Transport of the Yakutat terrane, southern
Alaska: evidence from sediment petrology and detrital zircon fission-track and U/Pb
double dating. J. Geol. 117, 156–173.
Phinney, E.J., Mann, P., Coffin, M.F., Shipley, T.H., 2004. Sequence stratigraphy, structural
style, and age of deformation of the Malaita accretionary prism (Solomon ArcOntong Java Plateau convergent zone). Tectonophysics 389, 221–246.
Pinet, N., Cobbold, P.R., 1992. Experimental insights into the partitioning of motion
within zones of oblique subduction. Tectonophysics 206, 371–388.
Plafker, G., 1987. Regional geology and petroleum potential of the northern Gulf of Alaska
continental margin. In: Scholl, D.W., Grantz, A., 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: Earth Science Series, pp. 229–268.
Plafker, G., Hudson, T., Bruns, T., Rubin, M., 1978. Late Quaternary offsets along Fairweather
fault and crustal plate interactions in southern Alaska. Can. J. Earth Sci. 15, 805–816.
Plafker, G., Nokleberg, W.J., Lull, J.S., 1989. Bedrock geology and tectonic evolution of
the Wrangellia, Peninsular, and Chugach terranes along the Trans-Alaska Crustal
Transect in the Chugach mountains and southern Copper River Basin, Alaska. J.
Geophys. Res. Solid Earth Planet. 94, 4255–4295.
Plafker, G., Naeser, C.W., Zimmerman, R.A., Lull, J.S., Hudson, T., 1992. Cenozoic uplift
history of the Mount McKinley area in the central Alaska Range based on fission-track
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014
E.S. Finzel et al. / Earth and Planetary Science Letters xxx (2011) xxx–xxx
dating. In: Bradley, D.C., Dusel-Bacon, C. (Eds.), Geologic Studies in Alaska by the U.S.
Geological Survey, 1991: U.S. Geological Survey Bulletin, 2041, pp. 202–212.
Plafker, G., Gilpin, L.M., Lahr, J.C., 1994. Neotectonic map of Alaska, in: Plafker, H., Berg, H.C.
(Eds.), The Geology of Alaska, The Geology of North America, v. G1. Geological Society
of America, Plate 12, scale 1:2,500,000.
Preece, S.J., Hart, W.K., 2004. Geochemical variations in the b 5 Ma Wrangell Volcanic
Field, Alaska: implications for the magmatic and tectonic development of a
complex continental arc system. Tectonophysics 392, 165–191.
Qi, C., Zhao, D.P., Chen, Y., 2007. Search for deep slab segments under Alaska. Phys. Earth
Planet. Inter. 165, 68–82.
Reinink-Smith, L.M., 1990. Relative frequency of Neogene volcanic events as recorded
in coal partings from the Kenai lowland, Alaska: a comparison with deep-sea core
data. Geol. Soc. Am. Bull. 102, 830–840.
Richter, D.H., Smith, J.G., Lanphere, M.A., Dalrymple, G.B., Reed, B.L., Shew, N., 1990. Age and
progression of volcanism, Wrangell volcanic field, Alaska. Bull. Volcanol. 53, 29–44.
Richter, D.H., Preller, C.C., Labay, K.A., Shew, N.B., 2006. Geologic map of the Wrangell
St. Elias Park and Preserve, Alaska. U.S. Geological Survey Scientific Investigations,
Map 2877, 1:350, 000.
Ridgway, K.D., DeCelles, P.G., 1993. Stream-dominated alluvial-fan and lacustrine
depositional systems in Cenozoic strike-slip basins, Denali fault system, YukonTerritory, Canada. Sedimentology 40, 645–666.
Ridgway, K.D., Sweet, A.R., Cameron, A.R., 1995. Climatically induced floristic changes
across the Eocene–Oligocene transition in the northern high-latitudes, YukonTerritory, Canada. Geol. Soc. Am. Bull. 107, 676–696.
Ridgway, K.D., Trop, J.M., Nokleberg, W.J., Davidson, C.M., Eastham, K.R., 2002. Mesozoic
and Cenozoic tectonics of the eastern and central Alaska Range: progressive basin
development and deformation in a suture zone. Geol. Soc. Am. Bull. 114, 1480–1504.
Ridgway, K.D., Thoms, E.E., Layer, P.W., Lesh, M.E., White, J.M., Smith, S.V., 2007.
Neogene transpressional foreland basin development on the north side of the
central Alaska Range, Usibelli Group and Nenana Gravel, Tanana basin. In: Ridgway,
K.D., Trop, J.M., Glen, J.M.G., O'Neill, J.M. (Eds.), Tectonic Growth of a Collisional
Margin: Crustal Evolution of Southern Alaska: Geological Society of America Special
Paper, 431, pp. 507–547.
Roeske, S.M., Snee, L.W., Pavlis, T.L., 2003. Dextral-slip reactivation of an arc-forearc
boundary during Late Cretaceous–Early Eocene oblique convergence in the
northern Cordillera. In: Sisson, V.B., Roeske, S.M., Pavlis, T.L. (Eds.), Geology of a
Transpressional Orogen Developed During Ridge-trench Interaction Along the
North Pacific Margin: Geologic Society of America Special Paper, 371, pp. 141–169.
Rondenay, S., Montesi, L.G., Abers, G.A., 2010. New geophysical insight into the origin of
the Denali volcanic gap. Geophys. J. Int. 182, 613–630.
Ruppert, N.A., Ridgway, K.D., Freymueller, J.T., Cross, R.S., Hansen, R.A., 2008. Active
tectonics of interior Alaska: a synthesis of seismic, GPS and geomorphic studies. In:
Freymueller, J.T., Haeussler, P.J., Wesson, R.J., Ekstrom, G. (Eds.), Active Tectonics
and Seismic Potential of Alaska: Geophysical Monograph Series, 179, pp. 109–133.
Sendziak, K., Armstrong, P.A., Haeussler, P.J., 2009. Constraints on exhumation of the
western Chugach Mountains (Alaska) based on zircon fission-track analysis of
modern glacial outwash. Geol. Soc. Am. Abstr. Programs 41, 290.
Short, E., Snyder, D.C., Trop, J.M., Hart, W.K., Layer, P.W., 2005. New findings on Early
Cretaceous volcanism within the allochthonous Wrangellia terrane, south-central
Alaska: stratigraphic, geochronologic, and geochemical data from the Chisana
Formation, Nutzotin Mountains. Geol. Soc. Am. Abstr. Programs 37, 81.
Skulski, T., Francis, D., Ludden, J., 1991. Arc-transform magmatism in the Wrangell
volcanic belt. Geology 19, 11–14.
Skulski, T., Francis, D., Ludden, J., 1992. Volcanism in an arc-transform transition zone—
the stratigraphy of the St. Clare Creek volcanic field, Wrangell volcanic belt, Yukon,
Canada. Can. J. Earth Sci. 29, 446–461.
Snyder, D.E., Hart, W.K., 2005. Isotopic constraints on source reservoirs for Cretaceous
magmatism within the Wrangellia terrane. Geol. Soc. Am. Abstr. Programs 37, 81.
Snyder, D.C., Hart, W.K., 2007. The White Mountains Granitoid Suite: isotopic constraints
on source reservoirs for Cretaceous magmatism within the Wrangellia Terrane. In:
13
Ridgway, K.D., Trop, J.M., Glen, J.M.G., O'Neill, J.M. (Eds.), Tectonic Growth of a
Collisional Margin: Crustal Evolution of Southern Alaska: Geologic Society of America
Special Paper, 431, pp. 379–400.
Spotila, J.A., Buscher, J.T., Meigs, A.J., Reiners, P.W., 2004. Long-term glacial erosion of active
mountain belts: example of the Chugach St. Elias Range, Alaska. Geology 32, 501–504.
Stock, J., Molnar, P., 1988. Uncertainties and implications of the Late Cretaceous and
tertiary position of North America relative to the Farallon, Kula, and Pacific plate.
Tectonics 7, 1339–1384.
Taira, A., Mann, P., Rahardiawan, R., 2004. Incipient subduction of the Ontong Java
Plateau along the North Solomon trench. Tectonophysics 389, 247–266.
Triplehorn, D.M., Turner, D.L., Naeser, C.W., 1977. K–Ar and fission-track dating of ash
partings in coal beds from the Kenai Peninsula, Alaska: a revised age for the
Homerian Stage–Clamgulchian Stage boundary. Geol. Soc. Am. Bull. 88, 1156–1160.
Trop, J.M., 2008. Latest Cretaceous forearc basin development along an accretionary
convergent margin: south-central Alaska. Geol. Soc. Am. Bull. 120, 207–224.
Trop, J.M., Ridgway, K.D., 2007. Mesozoic and Cenozoic tectonic growth of southern
Alaska: a sedimentary basin perspective. In: Ridgway, K.D., Trop, J.M., Glen, J.M.
G., O'Neill, J.M. (Eds.), Tectonic Growth of a Collisional Margin: Crustal
Evolution of Southern Alaska: Geological Society of America Special Paper,
431, pp. 55–94.
Trop, J.M., Ridgway, K.D., Spell, T.L., 2003. Synorogenic sedimentation and forearc
basin development along a transpressional plate boundary, Matanuska valley–
Talkeetna Mountains, southern Alaska. In: Sisson, V.B., Roeske, S., Pavlis, T.L.
(Eds.), Geology of a Transpressional Orogen Developed During Ridge–trench
Interaction Along the North Pacific Margin: Geological Society of America Special
Paper, 371, pp. 89–118.
Trop, J.M., Ridgway, K.D., Sweet, A.R., 2004. Stratigraphy, palynology, and provenance of
the Colorado Creek basin, Alaska, USA: oligocene transpressional tectonics along
the central Denali fault system. Can. J. Earth Sci. 41, 457–480.
Trop, J.M., Snyder, D.C., Hart, W.K., Idleman, B., Delaney, M.R., 2007. Miocene intra-arc
basin development within the Wrangell volcanic field, Frederika Formation and
Lower Wrangell Lava, eastern Wrangell Mountains, Alaska. Geol. Soc. Am. Abstr.
Programs 38, 491.
Turner, D.L., Triplehorn, D.M., Naeser, C.W., Wolfe, J.A., 1980. Radiometric dating of ash
partings in Alaskan coal beds and upper Tertiary paleobotanical stages. Geology 8,
92–96.
van Hunen, J., van den Berg, A.P., Vlaar, N.J., 2002. On the role of subducting oceanic
plateaus in the development of shallow flat subduction. Tectonophysics 352, 317–333.
vonHuene, R., Klaeschen, D., Fruehn, J., 1999. Relation between the subducting plate
and seismicity associated with the Great 1964 Alaska Earthquake. Pure Appl.
Geophys. 154, 575–591.
Walker, J.D., Geissman, J.W., 2009. Geologic Time Scale. Geological Society of America.
doi:10.1130/2009.CTS004R2C.
Willis, J.B., Haeussler, P.J., Bruhn, R.L., Willis, G.C., 2007. Holocene slip rate for the
western segment of the Castle Mountain fault, Alaska. Bull. Seismol. Soc. Am. 97,
1019–1024.
Wilson, F.H., Dover, J.H., Bradley, D.C., Weber, F.R., Bundtzen, T.K., Haeussler, P.J., 1998.
Geologic Map of Central (Interior) Alaska. U.S. Geological Survey Open-File Report
98-133-A.
Wilson, F.H., Hults, C.P., Schmoll, H.R., Haeussler, P.J., Schmidt, J.M., Yehle, L.A., Labay, K.A.,
2009. Preliminary Geologic Map of the Cook Inlet Region, Alaska. U.S. Geological
Survey Open-File Report 2009-1108.
Wolfe, J.A., Tanai, T., 1980. The Miocene Seldovia Point flora from the Kenai Group,
Alaska. U.S. Geological Survey Professional Paper 1105. 52 pp.
Yogodzinski, G.M., Kay, R.W., Volynets, O.N., Klolskov, A.V., Kay, S.M., 1995. Magnesian
andesite in the western Aleutian Komandorsky region: implications for slab
melting and processes in the mantle wedge. Geol. Soc. Am. Bull. 107, 505–519.
Yogodzinski, G.M., Lees, J.M., Churikova, T.G., Dorendorf, F., Wöerner, G., Volynets, O.N.,
2001. Geochemical evidence for the melting of subducting oceanic lithosphere at
plate edges. Nature 409, 500–504.
Please cite this article as: Finzel, E.S., et al., Upper plate proxies for flat-slab subduction processes in southern Alaska, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.01.014