Flexural versus dynamic processes of subsidence in the North American

advertisement
GEOPHYSICAL RESEARCH LETTERS, VOL. 40, 1–5, doi:10.1002/grl.50831, 2013
Flexural versus dynamic processes of subsidence in the North American
Cordillera foreland basin
Clayton S. Painter1 and Barbara Carrapa1
Received 14 May 2013; revised 5 August 2013; accepted 5 August 2013.
retroarc foreland basin to a longer-wavelength foreland basin
(during the Coniacian to Santonian time, ~88.7–83.5 Ma)
and then to a basin with a depocenter more than 400 km east
of the thrust front during Campanian time [Roberts and
Kirschbaum, 1995].
[4] There is evidence that some dynamic subsidence
affected the Western Interior throughout the Cretaceous; for
example, Bond [1976] asserted that a eustatic sea level rise
was not sufficient to explain the thickness of the Cretaceous
marine deposits in the Western Interior, and Mitrovica et al.
[1989] state that dynamic subsidence, related to the
shallowly subducting Farallon Plate during the Cretaceous,
caused the transgression of the Western Interior Seaway.
Past flexural modeling and subsidence analyses have been
able to closely match the foreland basin profile caused by
the tectonic loading of the North American Cordillera during
the Early Cretaceous [Jordan, 1981; Heller et al., 1986;
Currie, 1998, 2002; Pang and Nummedal, 1995]. The wavelength of the foreland basin profile during the Late
Cretaceous lengthens; efforts to explain the cause of the
lengthening of the flexural wavelength in the Late
Cretaceous and the later migration of the depocenter to the
east during the Campanian have produced a variety of
hypotheses. Jordan [1981] reported that the early wavelength broadening was caused by the erosion and redistribution of the thrust load, whereas Pang and Nummedal
[1995] proposed that dynamic subsidence was responsible
for the “middle Cretaceous” shift in subsidence style.
Understanding whether this lengthening of the foreland
basin wavelength was caused by flexural processes or
dynamic subsidence will allow us to better test current
foreland basin and geodynamic models.
[5] The Campanian shift in subsidence has been attributed
to both isostatic and dynamic processes. Cross and Pilger
[1978] proposed that a shallowly subducting plate coupled
with the overlying continental lithosphere and caused broad
isostatic subsidence in the Sevier foreland basin. Later,
researchers attribute the Campanian depocenter shift to
dynamic subsidence caused by the cold oceanic lithosphere
being subducted at a shallow angle below the overriding
plate [Mitrovica et al., 1989; Gurnis, 1993; Liu and
Nummedal, 2004] (Figure 2). Recent advances in tomography and inverse modeling have allowed researchers to model
dynamic effects on subsidence during the Cretaceous based
on the current location of the Farallon Plate, plate motion
rates, and mantle viscosities [Spasojevic et al., 2009].
[6] The goal of this study is to more precisely constrain the
transition from a narrow flexural foreland basin to a longerwavelength foreland basin to a nonflexural basin in order to
better understand the mechanism underlying dynamic subsidence in the western U.S., test existing models, and provide
more detailed basin geometries to constrain geodynamic
[1] Dynamic subsidence related to subduction is an
important process in retroarc foreland basin systems.
Dynamic subsidence in the North American retroarc foreland
has been proposed as dominant in the Late Cretaceous;
however, questions remain about the nature of the subcrustal
load and the basin response to such processes. We present
new isopach data using 130 data points covering a large
portion of the U.S. and a revised flexural analysis of the
Sevier foreland. Higher rigidity associated with the
Wyoming craton can best model flexural subsidence directly
prior to ~81 Ma. We constrain the transition from flexural to
nonflexural subsidence to ~81 Ma and correlate this with
large Late Cretaceous progradation. The locus of subsidence
and the location of the subducted Shatsky oceanic plateau
suggest that dynamic loading from viscous flow above the
subducting plate is the mechanism driving dynamic
subsidence in the retroarc foreland basin of the North
American Cordillera. Citation: Painter, C. S., and B. Carrapa
(2013), Flexural versus dynamic processes of subsidence in the
North American Cordillera foreland basin, Geophys. Res. Lett., 40,
doi:10.1002/grl.50831.
1. Introduction
[2] The North American Cordillera extends from southern
Nevada, northeast through Utah and then north-northwest
through western Wyoming, eastern Idaho, and western
Montana, and through the Canadian Rockies, forming
together with the South American Cordillera the archetypical
retroarc orogenic and foreland basin system (Figure 1). Thrust
belt deformation initiated in the western U.S. interior in the
Middle to Late Jurassic and continued until the Paleocene
[DeCelles and Burden, 1992; Hudec, 1992; Lawton et al.,
1993; Currie, 1998; DeCelles, 2004]. Subsidence mechanisms
for Cordilleran-type orogenic systems include flexural loading
and dynamic processes [Jordan, 1981; Mitrovica et al., 1989;
Pang and Nummedal, 1995; Liu and Nummedal, 2004; Liu
et al., 2010].
[3] Roberts and Kirschbaum [1995] produced six largescale foreland basin isopach maps in the western U.S. for
the time periods spanning the Late Cretaceous, from the
Cenomanian (~98.5 Ma) to the Campanian (~72 Ma). Those
isopach maps document a transition from a narrow flexural
Additional supporting information may be found in the online version of
this article.
1
Department of Geosciences, University of Arizona, Tucson, Arizona, USA.
Corresponding author: C.S. Painter, Department of Geosciences,
University of Arizona, 1040 E. 4th St., Tucson, AZ 85721, USA.
(cpainter@email.arizona.edu)
©2013. American Geophysical Union. All Rights Reserved.
0094-8276/13/10.1002/grl.50831
1
PAINTER AND CARRAPA: FLEXURAL VERSUS DYNAMIC PROCESSES
N
Sevier fold-thrust belt
Cordilleran magmatic arc
Laramide foreland province
550
0 200
Major thrust fault
kilometers
Study area
Major basins − Abbreviations outlined in caption
Figure 1. Regional index map of the western U.S. including the Sevier fold-thrust belt, Laramide province, Cordilleran
magmatic arc, and major basins. Basins shown are the San Juan (SJB), Paradox (PB), Uinta-Piceance (U-PB), DenverJulesburg (D-JB), Green River (GRB), Wind River Basin (WRB), Big Horn (BHB), and Powder River (PRB) (modified from
DeCelles [2004]).
models. To do this, we construct four isopach maps dividing
Turonian (~93 Ma) to Campanian (~76 Ma) time (Figure 3
and supporting information). Also, we flexurally model the
early, Santonian to Campanian, long-wavelength flexural foreland basin and compare the nonflexural subsided basin of the
later Campanian and discuss it in relation to the various dynamic topography models that have been proposed [Cross
and Pilger, 1978; Mitrovica et al., 1989; Gurnis, 1993; Liu
et al., 2010].
2. Isopach Maps
[7] New isopach maps are here constructed using ~130
geophysical well logs, depending on the horizon that was
correlated. These isopachs span basins in Wyoming, Utah,
Colorado, and New Mexico. The main selection criteria for
the well logs are to at least penetrate the top of the Dakota
Sandstone, which is ~93 Myr [Molenaar et al., 2002] and
to have continuous gamma ray logs from the top of the
Dakota Sandstone into Campanian sandstones (~76 Ma).
[8] The thicknesses between horizons are calculated based
on correlations (criteria described in Text S1 in supporting
information). A well list data file that includes the measured
depths to the correlated horizons and the calculated thicknesses is included in Text S2 in supporting information. XY
grids were then produced in PETRA® (a well log management and interpretation software). Subsequently, the grids
were imported into ArcGIS where the grids were contoured
using a nearest-neighbor interpolation algorithm. Finally,
contours were smoothed by hand, while still honoring the
point data, producing isopach maps spanning ~93–89 Ma,
~89–84 Ma, ~84–81 Ma, and ~81–76 Ma, the total span being within the Upper Cretaceous from lower Turonian to
middle Campanian. In order to have a better spatial coverage,
these isopachs are extended to the north by using existing
isopachs from the literature [Weimer, 1960; Weimer and
Flexer, 1985; Roberts and Kirschbaum, 1995]. Isopachs
from Weimer and Flexer [1985] and Roberts and
Kirschbaum [1995] were used to both guide the trend of the
contour lines and the thicknesses present to the north.
Weimer [1960] does not include isopach data, but it does include a stratigraphic regional transect; thickness from this
transect was used to constrain isopach thickness to the north.
The data coverage in the northern part of the isopach maps
can be found in Figure S1 in supporting information. The
~93–89 Ma and ~89–84 Ma isopach maps show an eastward
a)
b)
Figure 2. Cartoon of two models that have been proposed
as the cause of the nonflexural subsidence observed during
the Cretaceous in the Western Interior. (a) Cross and Pilger
[1978] state that subcrustal loading from by a flat subducting
plate causes broad wavelength subduction. (b) Dynamic
effects of a shallowly subducting plate on the mantle above
has also been asserted as the cause of broad wavelength
subduction [Mitrovica et al., 1989; Spasojevic et al., 2009;
Liu et al., 2010].
2
PAINTER AND CARRAPA: FLEXURAL VERSUS DYNAMIC PROCESSES
a)
b)
c)
d)
Figure 3. Isopach maps that span ~ 93–76 Ma. Black dots represent well log control. Where there is no well log control, we
use previously published isopach maps as reference [Weimer, 1960; Weimer and Flexer, 1985; Roberts and Kirschbaum,
1995]. Maps spanning (a) ~93–89 Ma, (b) ~89–84 Ma, (c) ~84–81 Ma, and (d) ~81–76 Ma. Note the lengthening of the basin
wavelength in Figure 3c. A–A′ is the location of the profile in Figure 4. Figure 3d includes the proposed location of the
Shatsky at 80 Ma [Liu et al., 2010].
the narrow foredeep, parallel to the thrust front, is no longer
present. Instead, there is a broad depocenter in northeast
Utah, southwest Wyoming, and northwest Colorado, which
thins to the south-southwest (Figure 3c). This change in
subsidence profile has been attributed to early onset of
dynamic subsidence [Gurnis, 1993; Pang and Nummedal,
1995; DeCelles, 2004]. Yet, even though the character of
the depocenter has changed, we show that the subsidence
profile can still be approximated using reasonable flexural
parameters (Figure 4).
[11] The ~84–81 Ma isopach map pattern has an ~1400 m
thick and ~300 km wide depocenter (Figure 3c). Here the
lithosphere is treated as an infinite, unbroken elastic plate
with a rectangular load [Turcotte and Schubert, 1982;
Angevine et al., 1990]. A variety of parameters were used
in trying to most closely match the observed basin profile,
and the following parameters produce the closest match.
Typically, load dimensions of 400 km wide and 2 km high
are used to approximate the Cordilleran thrust belt [Jordan,
1981; Currie, 2002]. The period of 84–81 Ma spans a time
after the Wasatch Culmination had been formed and was still
building, indicating that a higher tectonic load can be
assumed [DeCelles, 1994]. The dimensions used in these
migrating, narrow depocenter in the west that shallows to the
east (Figures 3a and 3b). The ~84–81 Ma isopach map shows
a broad depocenter in southern Wyoming that thins to the
south, southeast, and east (Figure 3c). At ~81–76 Ma, the location of the depocenter shifts ~480 km to the east where it
trends north-south in what is now eastern Wyoming and central Colorado (Figure 3d).
[9] Foredeep to back-bulge foreland basin strata from
~93–81 Ma are dominated by marine shale deposits with
intermittent marginal marine and fluvial deposits. At
~81 Ma, the foredeep strata become coarser, with marginal
marine and fluvial deposits becoming more prevalent [e.g.,
Franczyk et al., 1992].
3. Flexural Modeling
[10] As shown by previous researchers, the observed
foreland basin system can be reproduced with flexural calculations from the Late Jurassic to early on in the Late
Cretaceous [Jordan, 1981; Yingling and Heller, 1992; Pang
and Nummedal, 1995; Currie, 1997, 2002]. However, as
documented in this study and in previously published
isopach maps [Roberts and Kirschbaum, 1995], at ~84 Ma,
3
PAINTER AND CARRAPA: FLEXURAL VERSUS DYNAMIC PROCESSES
Flexural subsidence, rectangular load
1000
compared to the hypothesized location of the Shatsky
Plateau at 80 Ma (Figure 3d). The isopach map shows
that the dynamically subsiding basin overlies the area
where the northern portion of the Shatsky Plateau was
plunging into the asthenosphere, with a similar northsouth trend (Figure 3d).
2000
5. Discussion and Conclusions
subsidence in meters
-1000
0
0
[13] This study constrains the transition from flexural
subsidence to dynamic subsidence that was previously
placed at ~83.5 Ma [Roberts and Kirschbaum, 1995;
DeCelles, 2004] at ~81 Ma. Our data show that the shape of
the dynamically subsiding basin is not east-west trending,
as shown in Roberts and Kirschbaum [1995], but rather
north-south trending, similar to what is proposed by
Weimer [1970]. Stratigraphically, this transition from
flexural subsidence to dynamic subsidence takes place at
the start of a regional and widespread progradation of coarse,
foreland basin deposits, becoming more incisional higher in
the section. A basin subsidence transition from higher
accommodation in the west and less accommodation in the
east (as is the case in a flexural basin) to higher rates of subsidence in the east and less in the west (as is the case when
this dynamic subsidence begins) would produce a rapid
progradation as observed in the Campanian [Mitrovica
et al., 1989; Pang and Nummedal, 1995; DeCelles, 2004].
[14] Flexure is driving subsidence from ~93 to 81 Ma.
However, flexural foreland basin depozones of foredeep,
forebulge, and back-bulge [DeCelles and Giles, 1996] should
not be applied to the western U.S. for times after ~81 Ma
because the foreland basin after this time is controlled by
dynamic processes.
[15] Previous researchers have observed the broadening of
the flexural wavelength in the Coniacian to Santonian
(~88.7–83.5 Ma) and have attributed this to the erosion and
redistribution of the tectonic load [Jordan, 1981] and to early
dynamic influence [Pang and Nummedal, 1995]. Based on
our new isopach maps, this broadening of the flexural
foreland takes place at ~84 Ma. Neither the erosion nor
redistribution of the tectonic load nor dynamic influence is
necessary to explain this wavelength change. Instead, the
wavelength change can be explained by an increase in
rigidity of the loaded lithosphere. Jordan [1981] assumed a
constant flexural rigidity of the lithosphere and approximated
it at 1023 N m. The resulting models fit the observed flexural
profile in the Early Cretaceous. Pang and Nummedal [1995]
recognized the fact that flexural rigidity likely changes
throughout the western U.S. Lowry and Smith [1994] used
coherence analysis of gravity and topography to calculate
the flexural rigidity of several provinces in the western U.S.
Their calculated value for the flexural rigidity of the
Archean Wyoming craton was 4.1 × 1024 N m, significantly
higher than the 1023 N m used for modeling central Utah.
DeCelles [1994] reports that at ~84–75 Ma, the Absaroka
thrust breaks out to the east, advancing the thrust belt
eastward. This would effectively load the western margin
of the Archean Wyoming craton. This also coincides with
the broadening of the flexural wavelength as seen in the
isopach map (Figure 3c). Furthermore, by 84 Ma, the development of the Wasatch Culmination effectively increased
the height of the topographic load. Therefore, a flexural rigidity of 4.1 × 1024 N m and tectonic load dimensions of 400 km
3000
4000
0
100
200
300
400
500
600
distance from thrust front (km)
Figure 4. Flexural model of the profile of A–A′ in
Figure 3c. The dotted line is the observed profile, the solid
line is the flexural profile for a rigidity of 3 × 1024 N m, and
the dashed line (the best match) is for a rigidity of
4.1 × 1024 N m. B–B′ in Figure 3c is also modeled, showing
that the tectonic load height decreases (2 km) to the north
and the lithosphere becomes less rigid (2 × 1024 N m).
calculations are a 400 km wide and 2.5 km high tectonic load.
Its density is 2500 kg/m3; the basin fill and mantle density are
2300 kg/m3 and 3300 kg/m3, respectively. All of these
parameters are similar to those used in previous flexural
studies [Jordan, 1981; Yingling and Heller, 1992; Currie,
2002]. Because of the broad wavelength of the depocenter,
here, two high flexural rigidities are used: 3 × 1024 N m
[White et al., 2002] and 4.1 × 1024 N m [Lowry and Smith,
1994]. Lowry and Smith [1994] calculated the flexural
rigidity throughout the western U.S. using coherence analysis of gravity and topography. Using the above parameters,
the foredeep is calculated to be ~3.3 km deep, at the thrust
front, and ~300 km wide (Figure 4). The flexural curve with
a rigidity of 4.1 × 1024 N m best matches the observed profile
during ~84–81 Ma, which shows a stratal thickness of
1400 m in the Green River Basin, approximately 140 km east
of the thrust front during that time (Figures 3c and 4). In order
to address possible along-strike variability in the flexural
response of the region, we calculate a second flexural profile
to the north of the A–A′ transect, B–B′ (Figure 3c). Slightly
different flexural parameters are used to match this profile,
but all are within the reasonable parameters that are
explained above. The two differences are a tectonic load
height of 2 km (instead of 2.5 km in A–A′) and a less rigid
plate of 2 × 2024 N m, producing a very close match to the
observed profile geometry (Figure 4).
4. Flat-Slab Subduction and Dynamic Subsidence
[12] Many have hypothesized that the subduction of an
oceanic plateau caused flat-slab subduction during the Late
Cretaceous [Livaccari et al., 1981; Tarduno et al., 1985;
Saleeby, 2003]. Liu et al. [2010] use plate reconstruction
models to predict where the hypothesized Shatsky Plateau
was at 80 Ma. They place the center of the plateau and the
shallowest portion near the intersection of the New Mexico,
Arizona, Utah, and Colorado borders with the plateau
trending north-south and plunging to the north. In this study,
the dynamically subsided basin during ~81–76 Ma is
4
PAINTER AND CARRAPA: FLEXURAL VERSUS DYNAMIC PROCESSES
reconstructions for the Uinta-Piceance basin study area, Colorado and
Utah, U.S. Geol. Surv. Bull., 1787, 37 pp.
Gurnis, M. (1993), Depressed continental hypsometry behind oceanic trenches:
A clue to subduction controls on sea-level change, Geology, 21(1), 29–32,
doi:10.1130/0091-7613(1993)021<0029:DCHBOT>2.3.CO;2.
Heller, P. L., S. S. Bowdler, H. P. Chambers, J. C. Coogan, E. S. Hagen,
M. W. Shuster, N. S. Winslow, and T. F. Lawton (1986), Time of initial
thrusting in the Sevier orogenic belt, Idaho–Wyoming and Utah,
Geology, 14, 388–391.
Hudec, M. R. (1992), Mesozoic structural and metamorphic history of the
central Ruby Mountains metamorphic core complex, Nevada, Geol. Soc.
Am. Bull., 104, 1086–1100.
Jordan, T. E. (1981), Thrust loads and foreland basin evolution, Cretaceous,
western United States, AAPG Bull., 65, 2506–2520.
Lawton, T. F., P. J. Talling, R. S. Hobbs, J. H. Trexler Jr., M. P. Weiss, and
D. W. Burbank (1993), Structure and stratigraphy of Upper Cretaceous
and Paleogene strata (North Horn Formation), eastern San Pitch
Mountains, Utah—Sedimentation at the front of the Sevier orogenic belt,
U.S. Geol. Surv. Bull., 1787, pp. II1–II33.
Liu, L., M. Gurnis, M. Seton, J. Saleeby, R. D. Muller, and J. M. Jackson
(2010), The role of oceanic plateau subduction in the Laramide orogeny,
Nat. Geosci., 3, 353–357.
Liu, S., and D. Nummedal (2004), Late Cretaceous subsidence in Wyoming:
Quantifying the dynamic component, Geology, 32, 397–400,
doi:10.1130/G20318.1.
Livaccari, R. F., K. Burke, and A. M. C. Şengör (1981), Was the Laramide
orogeny related to subduction of an oceanic plateau?, Nature, 289,
276–278, doi:10.1038/289276a0.
Lowry, A. R., and R. B. Smith (1994), Flexural rigidity of the Basin and
Range-Colorado Plateau-Rocky Mountain transition from coherence analysis of gravity and topography, J. Geophys. Res., 99, 20,123–20,140,
doi:10.1029/94JB00960.
Mitrovica, J. X., C. Beaumont, and G. T. Jarvis (1989), Tilting of continental interiors by the dynamical effects of subduction, Tectonics, 8,
1079–1094.
Molenaar, C. M., W. A. Cobban, E. A. Merewether, C. L. Pillmore,
D. G. Wolfe, and J. M. Holbrook (2002), Regional stratigraphic
cross sections of Cretaceous Rocks from east–central Arizona to the
Oklahoma Panhandle, U.S. Geol. Surv. Misc. Field Stud. Map MF2382.
Pang, M., and D. Nummedal (1995), Flexural subsidence and basement tectonics of the Cretaceous Western Interior basin, United States, Geology,
23, 173–176.
Roberts, L. N. R., and M. A. Kirschbaum (1995), Paleogeography of the
Late Cretaceous of the western interior of middle North America: Coal
distribution and sediment accumulation, U. S. Geol. Surv. Prof. Pap.,
1561, 115 pp.
Saleeby, J. (2003), Segmentation of the Laramide slab—Evidence from the
southern Sierra Nevada region, Geol. Soc. Am. Bull., 115, 655–668.
Spasojevic, S., L. Liu, and M. Gurnis (2009), Adjoint models of mantle convection with seismic, plate motion, and stratigraphic constraints: North
America since the Late Cretaceous, Geochem. Geophys. Geosyst., 10,
Q05W02, doi:10.1029/2008GC002345.
Tarduno, J. A., M. McWilliams, M. G. Debiche, W. V. Sliter, and
M. C. Blake Jr. (1985), Franciscan complex Calera limestones: Accreted
remnants of Farallon Plate oceanic plateaus, Nature, 317, 345–347,
doi:10.1038/317345a0.
Turcotte, D. L., and G. Schubert (1982), Geodynamics: Applications of
Continuum Physics to Geological Problems, 450 pp., John Wiley, New
York.
Weimer, R. J. (1960), Upper Cretaceous stratigraphy, Rocky Mountain area,
AAPG Bull., 44(1), 1–20.
Weimer, R. J. (1970), Rates of deltaic sedimentation and intrabasin deformation, Upper Cretaceous of Rocky Mountain Region, in Deltaic
Sedimentation: Modern and Ancient, Soc. Econ. Paleontol. Mineral.,
Spec. Publ., vol. 15, edited by J. P. Morgan and R. H. Shaver,
pp. 270–292, Tulsa.
Weimer, R. J., and A. Flexer (1985), Depositional pattern and unconformities, Upper Cretaceous, eastern Powder River basin, Wyoming, in
Thirty-Sixth Annual Field Conference, Wyoming Geological Association
Guidebook, pp. 131–147, Casper.
White, T., K. Furlong, and M. Arthur (2002), Forebulge migration in the
Cretaceous Western Interior basin of the central United States, Basin Res.,
14(1), 43–54, doi:10.1046/j.1365-2117.2002.00165.x.
Yingling, V. L., and P. L. Heller (1992), Timing and record of foreland
sedimentation during the initiation of the Sevier Orogenic belt in central
Utah, Basin Res., 4, 279–290.
wide and 2.5 km high are reasonable parameters to use for
84–81 Ma. These parameters match the flexural profile
observed in southern Wyoming during that time (Figure 4).
To the north, the load height and the plate rigidity decrease
to 2 km and 2 × 1024 N m, respectively, and closely match
the observed profile (Figure 4).
[16] The shallow subduction angle of the Farallon Plate
and its effects on the mantle have been asserted as the cause
of the dynamic subsidence at ~81–76 Ma [Pang and
Nummedal, 1995]. Another hypothesis is that a flat-slab
subducting plate can displace the hot asthenosphere, couple
with the overlying continental lithosphere, and cause
isostatic subsidence [Cross and Pilger, 1978] (Figure 2).
Whereas an oceanic plateau (such as the proposed Shatsky
Plateau) could possibly cause subcrustal loading similar to
what Cross and Pilger [1978] proposed, its location at
81 Ma is too far south to drive the observed depocenter
during that time (Figure 3d). This study instead shows that
the dynamically subsided basin during ~81–76 Ma occurs
in the location where the hypothesized Shatsky Plateau is
plunging into the asthenosphere (Figure 3d), suggesting that
dynamic loading from viscous flow above the subducting
plate is the mechanism driving subsidence in the western
U.S. at this time (Figure 2).
[17] Acknowledgments. We thank Shell Oil, ExxonMobil, and GSA
research grant for funding support, and IHS for data and software use. Kirk
Schafer, Jason Morgan, and Paul Goodman were extremely helpful in
technical conversations, managing data agreements, and lending
software expertise.
[18] The editor thanks Michael Gurnis and an anonymous reviewer for
assistance in evaluating this paper.
References
Angevine, C. L., P. L. Heller, and C. Paola (1990), Quantitative Sedimentary
Basin Modeling, AAPG Course Note Series #32, Continuing Education,
247 pp. American Association of Petroleum Geologists, Tulsa.
Bond, G. (1976), Evidence for continental subsidence in North America during the Late Cretaceous global submergence, Geology, 4, 557–560,
doi:10.1130/0091-7613(1976)4<557:EFCSIN>2.0.CO;2.
Cross, T. A., and R. H. Pilger Jr. (1978), Tectonic controls of late Cretaceous
sedimentation, western interior, USA, Nature (London, U. K.), 274(5672),
653–657.
Currie, B. S. (1997), Sequence stratigraphy of nonmarine Jurassic–
Cretaceous rocks, central Cordilleran foreland-basin system, Geol. Soc.
Am. Bull., 109(9), 1206–1222, doi:10.1130/0016-7606(1997)109<1206:
SSONJC>2.3.CO;2.
Currie, B. S. (1998), Upper Jurassic-Lower Cretaceous Morrison and Cedar
Mountain Formations, NE Utah-NW Colorado: Relationships between
nonmarine deposition and early Cordilleran foreland basin development,
J. Sediment. Res., 68(4), 632–652.
Currie, B. S. (2002), Structural configuration of the Early Cretaceous
Cordilleran foreland-basin system and Sevier thrust belt, Utah and
Colorado, J. Geol., 110(6), 697–718, doi:10.1086/342626.
DeCelles, P. G. (1994), Late Cretaceous-Paleocene synorogenic sedimentation and kinematic history of the Sevier thrust belt, northeast Utah and
southwest Wyoming, Geol. Soc. Am. Bull., 106, 32–56.
DeCelles, P. G. (2004), Late Jurassic to Eocene evolution of the Cordilleran
thrust belt and foreland basin system, western USA, Am. J. Sci., 304,
105–168.
DeCelles, P. G., and E. T. Burden (1992), Nonmarine sedimentation in the
overfilled part of the Jurassic–Cretaceous Cordilleran foreland basin:
Morrison and Cloverly Formations, central Wyoming, USA, Basin Res.,
4, 291–313.
DeCelles, P. G., and K. A. Giles (1996), Foreland basin systems, Basin Res.,
8(2), 105–123, doi:10.1046/j.1365-2117.1996.01491.x.
Franczyk, K. J., T. D. Fouch, R. C. Johnson, C. M. Molenaar, and
W. A. Cobban (1992), Cretaceous and Tertiary paleogeographic
5
Download