ArticlePDF Available

Plate tectonics and basin subsidence history

Authors:

Abstract and Figures

Tectonic setting exerts first-order control on basin formation as reflected in basin subsidence history. While our approach ignores the effects of flexural loading and eustatic sea-level change, consistency of backstripped subsidence histories (i.e., with local loading effects of sediment removed) suggests consistent tectonic driving mechanisms in each tectonic setting, with the possible exception of forearc basins. Based on published subsidence curves and open-file stratigraphic data, we show the subsidence characteristics of passive margins, strike-slip basins, intracontinental basins, foreland basins, and forearc basins. Passive margin subsidence is characterized by two stages, rapid initial, synrift subsidence and slow post-rift thermal subsidence, with increasing subsidence rates toward the adjacent ocean basin. Subsidence of intracontinental basins is similar in magnitude to that seen in passive margin settings, but the former is generally slower, longer lived, and lacks initial subsidence. Long-lived subsidence for many intracontinental basins is consistent with cooling following thermal perturbation of thick lithosphere found beneath old parts of continents. Basins associated with strike-slip faults are usually short lived with very rapid subsidence. Changes in local stress regimes as strike-slip faults evolve, and migrate over time, coupled with three-dimensional heat loss in these small basins likely explain this subsidence pattern. Foreland basin subsidence rates reflect the flexural response to episodic thrust loading. Resultant subsidence curves are punctuated by convex-up (accelerating) segments. Forearc basins have the least consistent subsidence patterns. Subsidence histories of these basins are complex and may reflect multiple driving mechanisms of subsidence in forearc settings. Second-order deviations in subsidence suggest reactivation or superimposed tectonic events in many basin settings. The effects of eustatic sea-level change may also explain some deviations in curves. For many of these settings, subsidence histories are sufficiently distinctive to be used to help determine tectonic setting of ancient basin deposits.
Content may be subject to copyright.
doi:10.1130/B26398.1
2009;121;55-64 Geological Society of America Bulletin
Xiangyang Xie and Paul L. Heller
Plate tectonics and basin subsidence history
Geological Society of America Bulletin
on 16 January 2009 gsabulletin.gsapubs.orgDownloaded from
Email alerting services
articles cite this article
to recieve free email alerts when newwww.gsapubs.org/cgi/alertsclick
Subscribe
Geological Society of America Bulletin
to subscribe towww.gsapubs.org/subscriptions/index.ac.dtlclick
Permission request
to contact GSAhttp://www.geosociety.org/pubs/copyrt.htm#gsaclick
official positions of the Society.
citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect
presentation of diverse opinions and positions by scientists worldwide, regardless of their race,
includes a reference to the article's full citation. GSA provides this and other forums for the
the abstracts only of their articles on their own or their organization's Web site providing the posting
to further education and science. This file may not be posted to any Web site, but authors may post
works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms
requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent
their employment. Individual scientists are hereby granted permission, without fees or further
Copyright not claimed on content prepared wholly by U.S. government employees within scope of
Notes
© 2009 Geological Society of America
ABSTRACT
Tectonic setting exerts fi rst-order control
on basin formation as refl ected in basin sub-
sidence history. While our approach ignores
the effects of fl exural loading and eustatic
sea-level change, consistency of backstripped
subsidence histories (i.e., with local loading
effects of sediment removed) suggests con-
sistent tectonic driving mechanisms in each
tectonic setting, with the possible exception of
forearc basins.
Based on published subsidence curves and
open-fi le stratigraphic data, we show the sub-
sidence characteristics of passive margins,
strike-slip basins, intracontinental basins,
foreland basins, and forearc basins. Pas-
sive margin subsidence is characterized by
two stages, rapid initial, synrift subsidence
and slow post-rift thermal subsidence, with
increasing subsidence rates toward the adja-
cent ocean basin. Subsidence of intracontinen-
tal basins is similar in magnitude to that seen
in passive margin settings, but the former is
generally slower, longer lived, and lacks initial
subsidence. Long-lived subsidence for many
intracontinental basins is consistent with
cooling following thermal perturbation of
thick lithosphere found beneath old parts of
continents. Basins associated with strike-slip
faults are usually short lived with very rapid
subsidence. Changes in local stress regimes
as strike-slip faults evolve, and migrate over
time, coupled with three-dimensional heat
loss in these small basins likely explain this
subsidence pattern. Foreland basin subsi-
dence rates refl ect the fl exural response to
episodic thrust loading. Resultant subsidence
curves are punctuated by convex-up (acceler-
ating) segments. Forearc basins have the least
consistent subsidence patterns. Subsidence
histories of these basins are complex and may
refl ect multiple driving mechanisms of subsi-
dence in forearc settings.
Second-order deviations in subsidence sug-
gest reactivation or superimposed tectonic
events in many basin settings. The effects of
eustatic sea-level change may also explain
some deviations in curves. For many of these
settings, subsidence histories are suffi ciently
distinctive to be used to help determine tec-
tonic setting of ancient basin deposits.
Keywords: subsidence analysis, passive mar-
gins, intracontinental basins, foreland basins,
strike-slip basins, forearc basins.
INTRODUCTION
Sedimentary basins refl ect prolonged subsi-
dence of Earth’s surface, due to large-scale tec-
tonic processes operating between and within
plates (Kusznir and Ziegler, 1992). To the degree
that tectonic processes are refl ected in subsi-
dence history, basins in similar tectonic settings
should show similar patterns of subsidence.
Subsidence histories are taken to refl ect isostatic
adjustment to lithospheric processes, such as
thermal events, thickness changes, and loading
history. Therefore, subsidence history provides
insight into basin-forming mechanisms. Differ-
ences in subsidence histories between basins
may refl ect how fundamental driving mecha-
nisms vary as well as secondary infl uences, such
as sea-level change and sediment loading. By
comparing subsidence curves between different
basins of similar tectonic setting, it is possible to
determine consistency and/or differences in the
processes that drive subsidence.
Dickinson (1976) and Angevine et al. (1990)
compiled subsidence histories of basins in order
to discriminate between subsidence styles in
various tectonic settings. Since those studies,
more data have become available that may be
used to more fully defi ne subsidence patterns as
a function of tectonic setting. In this paper, we
use some of these data to demonstrate the styles
of subsidence in various plate tectonic settings,
emphasizing those settings where the modes
of subsidence are still poorly understood. Our
objective is to show that these histories can be
used as templates allowing tectonic interpreta-
tions based on subsidence patterns.
SUBSIDENCE ANALYSIS
Subsidence analysis yields a graphic repre-
sentation of the vertical movement of a strati-
graphic horizon, with respect to a datum in a
sedimentary basin. It tracks the subsidence and
uplift history at that location since the horizon
was deposited (van Hinte, 1978). Data needed
to reconstruct subsidence history include strati-
graphic thickness, lithology, estimate of paleo-
water depths, and age control. Subsidence anal-
ysis begins with a plot of sediment accumulation
through time using the present-day thickness
of each dated stratigraphic unit. Second, the
effects of compaction are included based on the
assumption that porosity lost is mostly caused
by mechanical compaction. Third, since sea
level is used as the datum for subsidence analy-
sis, paleobathymetry corrections are needed
to correct the seafl oor position to this datum.
The resulting curve refl ects the total subsidence
history (van Hinte, 1978) including the contri-
bution of tectonic loads, sediment loads, and
sea-level changes. Of these, the local isostatic
effects of sediment loading can be removed by
“backstripping” (Steckler and Watts, 1978).
The resulting subsidence curve, referred to
as “tectonic subsidence,” shows the idealized
subsidence history of a basin that would have
existed if only water, and no sediment, fi lled
the subsiding hole. Tectonic subsidence history
refl ects basin subsidence due to factors other
than sediment deposition and attendant isostatic
adjustment and compaction. More importantly,
it provides a way of normalizing subsidence in
different basins that have undergone very dif-
ferent sedimentation histories.
It is important to be aware of some limitations
to this analysis. These come from the inaccuracy
of data used to reconstruct history and from the
assumptions built into the method. In particu-
lar, age control and water depth often hamper
For permission to copy, contact editing@geosociety.org
© 2008 Geological Society of America
55
Plate tectonics and basin subsidence history
Xiangyang Xie
†*
Paul L. Heller
Department of Geology and Geophysics, University of Wyoming, Laramie, Wyoming 82071, USA
E-mail: xiangyang@utig.ig.utexas.edu
*Present address: Institute for Geophysics, Uni-
versity of Texas at Austin, 10100 Burnet Road Build-
ing 196, Austin, Texas 78758, USA.
GSA Bulletin; January/February 2009; v. 121; no. 1/2; p. 55–64; doi: 10.1130/B26398.1; 7 fi gures.
Xie and Heller
56 Geological Society of America Bulletin, January/February 2009
subsidence analysis. Water-depth estimates are
often diffi cult to determine because of a paucity
of unique depth indicators in sedimentary rocks.
However, the impact of poorly constrained water
depth can be reduced by analyzing sections that
are mostly composed of shallow marine deposits,
thus reducing the absolute magnitude of water-
depth uncertainties. In addition, any uncertain-
ties due to water-depth assignments are reduced
by working with relatively thick stratigraphic
successions and relatively shallow water depths.
For our compilation, we have chosen to use a
simplifi ed water-depth scale (Angevine et al.,
1990), assigned based on fossil and/or lithofa-
cies evidence: nonmarine is here considered to
be 50 m above sea level, inner shelf is 50 m, outer
shelf is 150 m, and upper slope is 350 m. Non-
marine water depths are typically from coastal
plain deposits associated with shorelines. For
the most part, errors in water-depth assignment
are relatively small compared to the magnitudes
of the curves. However, signifi cant changes in
water depth occurring within the large uncer-
tainties inherent in bathyal-abyssal water depths
may leave important tectonic signals undetected
(Dickinson et al., 1987) and so are not included
here. Age control is another potential source of
inaccuracy. The number and resolution of age
assignments vary widely for different sedimen-
tary successions. We have focused on sections
that have at least four to fi ve dated stratigraphic
horizons identifi ed as points for subsidence
analysis. For this compilation we accept the
time scales used by the original authors. The
possibility of signifi cant hiatuses can also
lead to error. In order to reduce the impact of
this issue, we limit our study to those sections
where major unconformities, where identi-
ed by the original authors, are few. We com-
pensate for changes in stratigraphic thickness
through time due to compaction following the
standard approach outlined by van Hinte (1978)
and Allen and Allen (2005). For simplicity, we
used the exponential porosity versus depth rela-
tionships from Sclater and Christie (1980), and
generalized lithologies to sandstone, shale, or
limestone. Subsidence is calculated with respect
to sea level. Since no universally accepted quan-
tifi ed curve of sea level exists, we have simply
chosen to ignore sea-level changes and assume
that they cause only relatively low-magnitude
variations in our calculated subsidence. Even
the magnitude of long-term sea-level change is
poorly constrained, but is likely small (<200 m)
relative to the magnitude of subsidence seen in
these curves (Haq et al., 1987; Harrison, 1990).
As such, we assume that the sea-level datum
has not changed over time. Our approach to
backstripping utilizes the one-dimensional local
isostatic method of Steckler and Watts (1978).
However, this approach does imply that signifi -
cant variations in sediment loading near the ana-
lyzed site in the basin do not cause subsidence.
For broad basins, where basin thicknesses do
not vary greatly over short distances, this is a
reasonable assumption (Angevine et al., 1990).
This becomes more important for relatively
small basins and/or those formed over relatively
rigid lithosphere. This error is reduced by the
fact that we are looking at comparing the overall
shape of resultant curves in similar tectonic set-
tings and in basins of similar size.
All subsidence curves presented here are tec-
tonic subsidence curves. Subsidence histories
were only collected from sites where the plate
tectonic setting, as reported in the literature, is
well known (Fig. 1). It is possible that future
work will lead to a reinterpretation of tectonic
setting for some of these curves. We document
both the location and the original author used
in our compilation so that these redefi ned cases
can be clearly identifi ed. Subsidence curves are
grouped together by tectonic setting and plot-
ted at the same scale to facilitate comparisons
(Figs. 2–7). For clarity, several simplifying
approaches are used to improve our compari-
son of curves. First, no uncertainties in age and
paleobathymetry are shown on the graphs. Sec-
ond, major unconformities, if present, are shown
by fl at horizontal lines on the subsidence curves,
and small unconformities are not shown.
RESULTS
Our compilation is limited to sources that pro-
vide quantitative data that include stratigraphic
thickness, age assignments, proxy water-depth
data, and where the tectonic setting is well
established. In addition, an individual basin
may undergo different phases of subsi dence as
tectonic settings change through time. We have
tried to isolate those phases, or megasequences
(Allen and Allen, 2005), during which the spe-
cifi c tectonic setting of interest occurs. Fig-
ure 1 locates the sites and tectonic setting for
Intercontinental basins
Foreland basins
Forearc basins
7-7
7-1
7-2
2-4
2-5
2-7
6-1
2-1
2-6
2-8
4-2
3-1
3-2 3-3
3-4 3-5
3-6
3-9
4-1
4-3 4-4
4
6-2
6-4
7-3
7-5
7-6
7-9
7-8
7-4
3-8
3-8
4-5
4-7
4-8
2-3
3-7
6-5
4-6
4
6
62
6-7
6-6
66
6-3
Intercontinental basins
Foreland basins
Forearc basins
2-2
3-1
6
6
6
6
6
6
6
6
6
3-8
6-8
Passive margins
Strike-slip basins
6
6
6
6
2-
2
1
Figure 1. Locations of subsidence data by tectonic settings. Numbers refer to specifi c subsidence curves. First value is fi gure number, and
second value is curve number in that fi gure (e.g., 2–3 = Figure 2, curve 3).
Basin subsidence
Geological Society of America Bulletin, January/February 2009 57
data compiled in this study. Figures 2–7 show
the results of subsidence analysis as a function
of plate tectonic settings. On each set of subsi-
dence curves we include, for comparison, a ref-
erence curve that parallels best-fi t thermal subsi-
dence of the seafl oor assuming a semi-infi nite
half-space model from Stein and Stein (1992).
All curves are corrected for compaction and
backstripped assuming local isostasy.
Passive Margins (Fig. 2)
Subsidence following continental rifting and
breakup leads to asymmetric subsidence and
foundering of continental margins (Steckler and
Watts, 1978). As a result, the amount of subsi-
dence increases seaward of the hinge zone. All
subsidence curves show an initial phase of rapid
subsidence followed by a phase in which subsi-
dence rates are reduced (e.g., Watts and Ryan,
1976; Steckler and Watts, 1978) and mimic the
age-depth curve of the seafl oor. Some margins
demonstrate an abrupt change in subsidence
rates between these phases. However, this
abruptness may refl ect a poorly constrained his-
tory of early subsidence in some cases. Initial
subsidence deposits are often coarse nonmarine
deposits that are notoriously diffi cult to bio-
stratigraphically date. In addition, in modern
settings, the initial subsidence deposits are the
deepest and, thus, less frequently penetrated
parts of the sections. As a result there tend to
be few age constraints to delimit the early sub-
sidence history and the transition from rapid to
slower subsidence. Nonetheless, some well-con-
strained curves, such as the one shown from the
Gulf of Lion (Steckler and Watts, 1980), indi-
cate that abrupt changes can be real. Subsidence
in passive-margin settings typically continues
for more than 150 m.y. Maximum subsidence
(Fig. 2) varies up to 4 km, in part depending
on distance seaward of the hinge zone (i.e., the
landward limit of extension).
Passive margin formation and subsidence
mechanisms have been much studied follow-
ing the breakthrough work of Watts and Ryan
(1976) and Steckler and Watts (1978). Rift
basins develop early during continental breakup
followed by passive margin subsidence once
breakup is complete. Not all rifts go to comple-
tion, and many “failed rifts” can be found (cf.
Allen and Allen, 2005, their Fig. 9.11). Theoreti-
cal and analytical studies suggest that tectonic
subsidence can be divided into an initial “synrift”
phase that primarily refl ects isostatic response
to extension and thinning of continental crust,
followed by a “post-rift” phase driven by ther-
mal reequilibration as the lithosphere cools and
thickens back to equilibrium. Synrift stretching
and thinning by factors of less than 2 are com-
mon in rift basins (e.g., Hendrie et al., 1994;
Kusznir et al., 1996a, 1996b; Roberts et al., 1995;
Swift et al., 1987) and variable along individual
passive margins. In general, stretching factors
increase seaward to the point of continental rup-
ture and ocean crust formation. In addition, local
variability in subsidence can refl ect local struc-
ture and thinning as well as superimposed effects
(King and Ellis, 1990; Nadin and Kusznir, 1995).
Various mechanical models have been proposed
to explain details of subsidence curves in this
setting. Such models consider how extensional
strain is partitioned through the lithosphere (e.g.,
pure shear versus simple shear and depth-depen-
dent stretching), character (e.g., symmetric ver-
sus asymmetric and volcanic versus nonvolcanic
6b
6a
8b
3
2
1
300 200 100
Ma
0
0
2
4
kilometers
300 200 100
Ma
0
0
2
4
PASSIVE MARGINS
4 5
600 500 400
Ma
300
Seafloor subsidence
Seafloor subsidence
7
kilometers
8a
Figure 2. Tectonic subsidence curves for passive margin settings.
Locations shown on Figure 1. Solid curves correspond to time scale
at top of graph and dotted lines to time scale at bottom of graph.
Thermal decay curve (dashed) for subsi dence of cooling seafl oor
(Stein and Stein, 1992), minus (i.e., shallowed) 500 m, is shown
for comparison. 1—Paleozoic Miogeocline, southern Canadian
Rocky Mountains (Bond and Kominz, 1984); 2—Moroccan Basin
(Ellouz et al., 2003); 3—Campos Basin (Mohriak et al., 1987);
4—Gippsland Basin (Falvey and Mutter, 1981; P. Yin, 1985, per-
sonal commun.); 5—Gulf of Lion (Benedicto et al., 1996); 6—U.S.
Cordilleran Miogeocline (Bissell, 1974; Armin and Mayer, 1983;
Devlin et al., 1986; Devlin and Bond, 1988); 7—Lusitanian Basin
(Stapel et al., 1996); 8—U.S. Atlantic margin (Steckler and Watts,
1978; Swift et al., 1987).
Xie and Heller
58 Geological Society of America Bulletin, January/February 2009
margins), timing, and rate of heat loss during and
following continental breakup (e.g., Bott, 1980;
Jarvis and McKenzie, 1980; Turcotte, 1980;
Watts, 1981; Wernicke et al., 1982; Cochran,
1983; Nadon and Issler, 1997), as well as super-
imposed tectonic events (e.g., Dore and Stewart,
2002; Nielsen et al., 2002). Variations in these
factors may explain differences in magnitude of
curves shown.
Strike-Slip Basins (Fig. 3)
Basins related to strike-slip faults include a
variety of basin types that result from a com-
bination of transform fault movement that may
include either elements of crustal extension or
shortening (Christie-Blick and Biddle, 1985).
Strike-slip basins show a large variety in basin
size and geometry. However, they are typically
narrow and smaller than those produced by
regional extension or shortening. Different types
of basins can form in strike-slip settings, from
simple pull-apart basins developed along fault
oversteps to more complex basin forms in zones
of transtension and transpression (e.g., May et
al., 1993; Sutherland and Melhuish, 2000). Fault
geometry and related fault-mechanical process
are the critical controls for the development of
strike-slip basins as demonstrated by different
authors (e.g., Crowell, 1974; Mann et al., 1983;
Ingersoll, 1988). Regardless of specifi c basin
shapes and locations, all curves generated for
this setting are characterized by rapid and short-
lived (typically <10 m.y.) subsidence.
Subsidence in this setting is dictated by the
spatial confi guration of the various scales of
strike-slip fault systems within the basin, as well
as the history of displacement and attendant heat
loss (Sawyer et al., 1987; Chen and Nábelek,
1988). Depending on local patterns of defor-
mation, subsidence curves in strike-slip basins
may be episodic and end abruptly (e.g., Crow-
ell, 1974; Mann et al., 1982; May et al., 1993).
Tectonic subsidence typically exceeds 2 km
and reaches 4 km in exceptional cases (e.g., the
Los Angeles basin, Fig. 3, line 5). The magni-
tude and concave-up shape of these curves are
similar to those from passive margin settings
(Fig. 1), although the subsidence rates are much
faster. The fact that most strike-slip basins are
short lived probably refl ects the evolution and
frequent change in position of the master strike-
slip faults (Sylvester, 1988; Cloetingh et al.,
1996; Storti et al., 2003; Allen and Allen, 2005;
Waldron, 2005).
Stretching models have been applied to strike-
slip basins; however, compensation is made for
the small space and time scales associated with
these basins, such as fi nite rifting times, accen-
tuated lateral heat fl ow, and depth-dependent
extension (Cochran, 1983). Heat fl ow increases
due to lithospheric thinning. Theoretical and
eld studies suggest that heat is lost rapidly dur-
ing the extension process, in part, by lateral con-
duction (Cochran, 1983; Pitman and Andrews,
1985). The short-lived tails seen at the ends
of most curves may result from cooling of the
small remaining thermal anomaly once the fault
ceased to be active (Pitman and Andrews, 1985).
The result is that there is very little subsidence
that continues once extension stops (Nilsen and
McLaughlin, 1985). In some cases, the absence
of evidence for post-rift thermal subsidence may
also be a result of subsequent deformation of the
basin (Christie-Blick and Biddle, 1985).
Intracontinental Basins (Fig. 4)
Intracontinental, or intracratonic, basins are
large basins formed on an old continental lith-
osphere away from any known active tectonic
margin (Dickinson, 1976). These basins are
typically quite large (>150,000 km
2
in area),
have relatively slow, long-lived subsidence
(typically >200 m.y. in duration), but in most
compiled cases tectonic subsidence is less than
2 km. Cross-sectional geometries of the North
American examples are approximately symmet-
ric (Illinois, Michigan, and Williston Basins);
whereas, others are not.
Subsidence curves of intracontinental basins
in Figure 4 are approximately exponential in
shape, similar to passive margins, but most lack
a rapid initial subsidence phase. Overall, subsi-
dence curves follow the shape and magnitude of
seafl oor subsidence, but with longer decay con-
stants. Such a comparison has led some (Haxby
et al., 1976; Sleep and Sloss, 1980; Cercone,
1984; Nunn and Sleep, 1984; Nunn et al., 1984;
Howell and van der Pluijm, 1999; Kominz et al.,
2001) to suggest a thermal decay origin for at
least some of these basins.
A comparison of intracontinental subsi-
dence curves to simple thermal subsidence
models (Fig. 5) indicates broad consistency.
To model thermal subsidence, we use McKen-
zie’s (1978) simple-stretching model, but only
calculate post-rift subsidence resulting from
lithosphere reequilibration following thinning.
Stretching (thinning) factors in this case only
refl ect thinning of mantle lithosphere due to
thermal perturbation and not necessarily exten-
sion. Stretching factors ranging from 1.1 to 1.5
and equilibrium lithosphere thickness of 125
and 200 km are shown. Notice that thicker lith-
osphere has a longer decay constant to reach
thermal equilibrium. Thermal decay constants
increase as the square of lithosphere thickness.
As a result, it is not too surprising that thick
lithosphere, which tends to exist beneath the
oldest cores of continents (Chapman and Pol-
lack, 1977; Artemieva and Mooney, 2001), has
the longest subsidence histories.
Most models of thermal reequilibration are
similar in approach to that used by Haxby et al.
(1976) for the Michigan Basin. In this model a
large-scale, but undocumented, thermal event
leads to formation of a dense crustal mass that
causes subsequent subsidence as the lithosphere
cools. Flexure broadens the width of basin
defl ection (Nunn and Sleep, 1984).
Notable in the subsidence curves are the
deviations from idealized thermal subsidence
(Fig. 5). These deviations are more pronounced
than those seen in passive margins and suggest
that tectonic reactivation characterizes many
intracontinental basins. The most extreme of
these is the Ordos Basin (Fig. 4, line 6). How-
ever, recent work suggests that this basin may be
the result of constructive interaction of deforma-
tion events around the basin margin and is not
primarily driven by thermal effects (Xie, 2007).
Most of the other examples also show strong
deviations away from simple thermal equilibra-
tion, more than can be reasonably accounted for
by eustatic sea-level changes. Various authors
have suggested interacting tectonic mechanisms
impacting these basins including intraplate
stresses, multiple thermal perturbations, reac-
tivation of inherited structures, far-fi eld effects
of nearby tectonic events, or changes in litho-
sphere rheology (Nunn and Sleep, 1984; Klein
and Hsui, 1987; Bond, 1991; Kaminski and Jau-
part, 2000).
Foreland Basins (Fig. 6)
Foreland basins are asymmetric basins adja-
cent, and parallel, to an attendant contractional
orogenic belt. Foreland basins, or foredeeps, sit
atop a defl ected continental lithosphere of the
underlying plate in both continental collision
zones (peripheral foreland basins of Dickinson
[1976]) and behind volcanic arcs (retroarc fore-
land basin of Dickinson [1976]). Many studies
have demonstrated that, for the most part, these
basins form as a regional isostatic (fl exural)
response to loading by the adjacent orogenic belt
(e.g., Beaumont, 1981; Jordan, 1982; DeCelles
and Giles, 1996).
Foreland basin subsidence curves differ from
thermal subsidence curves seen in most other
basins in that the former are characterized by
their convex-up shape and frequent episodic sub-
sidence events. The convex-up profi le refl ects
accelerating subsidence as the tectonic load
migrates toward the foreland coupled with the
curved fl exural profi le of the basin. As the basin
widens due to migration of the thrust load and
associated sedimentation, the distal parts of the
Basin subsidence
Geological Society of America Bulletin, January/February 2009 59
300 200 100
Ma
0
0
2
4
kilometers
INTRACONTINENTAL BASINS
1
2
3
4
5
6
7
8
500 400
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
Figure 4. Tectonic subsidence of intracontinental basins. Locations shown in Figure 1. See thermal decay curve
(dashed) for subsidence of cooling seafl oor (Stein and Stein, 1992), minus 1500 m, is shown for comparison. 1—Illi-
nois Basin, Farley well (Bond and Kominz, 1984); 2—Michigan Basin (Bond and Kominz, 1984); 3—Williston
Basin, North Dakota (Bond and Kominz, 1984); 4—Williston Basin, Saskatchewan (Fowler and Nisbet, 1985);
5—Northeast German Basin (Scheck and Bayer, 1999); 6—Southwest Ordos Basin (Xie, 2007); 7—Paris Basin
(Prijac et al., 2000); 8—Parana Basin (Zalan et al., 1990).
STRIKE-SLIP BASINS
100
Ma
0
0
2
4
kilometers
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
45a
5b
6
Ma
100 0
0
2
4
kilometers
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
12
3
Ma
100 0
0
2
4
kilometers
7
8
9
Figure 3. Tectonic subsidence curves for strike-slip basins. Locations shown in Figure 1. Thermal decay curve
(dashed) for subsidence of cooling seafl oor (Stein and Stein, 1992), minus 500 m, is shown for comparison. 1—Chuck-
anut Basin (Johnson, 1984, 1985); 2—Ridge Basin (Crowell and Link, 1982; Karner and Dewey, 1986); 3—Death
Valley (Hunt and Mabey, 1966); 4—Salinian block (Graham, 1976); 5—Los Angeles Basin (Rumelhart and Ingersoll,
1997); 6—Gulf of California (Curray and Moore, 1984); 7—Cuyama Basin (Dickinson et al., 1987); 8—Bozhang
Depression (Hu et al., 2001); 9—Salton Trough (Kerr et al., 1979).
Xie and Heller
60 Geological Society of America Bulletin, January/February 2009
basin show time-transgressive subsidence. That
is, while the proximal foreland basin responds
immediately to adjacent thrust loads, the distal
parts of the basin may show later subsidence as
loads migrate basinward over time (Jones et al.,
2004). The result is a time lag as the tectonic
and sediment load propagates across a foreland
basin. In addition, the redistribution of sediment
and subsidence over time in this way leads to
attening of the basin geometry and reduction
in size of the attendant forebulge. Jones et al.
(2004) suggest time lags of tectonic signals
on the order of a few million years or less out
across foreland basins. The total duration of oro-
gens, as seen in subsidence histories (Fig. 6), is
typically a few tens of millions of years.
Smaller scale episodes of subsidence super-
imposed on the overall subsidence profi le pri-
marily refl ect intermittent thrust events (e.g.,
Heller et al., 1986), although not every event
signifi cantly changes the confi guration of the
thrust load. Duration and episodicity of subsi-
dence varies from basin to basin, as set by the
pace of growth of the adjacent orogen, and in
different parts of a single basin (e.g., Fig. 6,
lines 8a and 8b), as a function of local loading
history. The maximum magnitude of tectonic
subsidence seen in compiled curves is ~3 km.
Blind thrusts often propagate into the proxi-
mal foreland basin and sedimentation can con-
tinue above these structures. DeCelles and Giles
(1996) refer to this part of the foreland basin
as the “wedge top.” While these basins are not
included here, it is clear that thrust emplacement
will impact subsidence history in these parts of
the proximal foreland (e.g., Vergés et al., 1998).
Other smaller basins may form in concert with
deformation of the adjacent orogen. These
include piggyback and back-bulge basins (Ori
and Friend, 1984; DeCelles and Giles, 1996).
Subsidence history of piggyback basins tends
to be shorter lived and of less magnitude than
their associated foreland basins (e.g., Burbank
et al., 1992; Carrapa et al., 2003), and is not
considered here. Back-bulge basins, if present,
are very subdued features that lie outboard of
foreland basins, beyond the forebulge, and form
as a dampened fl exural response to loading
of an elastic plate. Magnitude of defl ection of
back-bulge basins is very small, typically a few
percent of the depth of the associated foreland
basin, and may be diffi cult to uniquely identify.
Forearc Basins (Fig. 7)
Forearc basins lie between trenches and their
associated, parallel, magmatic arcs (Dickinson,
1995). The sizes and confi gurations of both
modern and ancient forearc basins are highly
variable, but it is clear that typical forearc basins
are narrow and elongate, with thick sediment
packages confi ned to deep structural troughs. We
note that there is often a large range of paleoba-
thymetry in these settings, so that resultant sub-
sidence curves may be less well constrained.
Subsidence curves from forearc basins, as a
group, have the most diverse range of shapes
(Fig. 7). Some show very rapid, short-lived sub-
sidence similar to strike-slip basins. Others have
slower, relatively linear subsidence. Still others
show an abrupt transition from rapid subsidence
to very slow subsidence rates, similar to some
200 300 400
m.y.
0
2
kilometers
1
2
3
4
5
6
8
7
0 100
200 300
m.y.
0
2
kilometers
0 100
β = 1.1
β = 1.2
β = 1.4
β = 1.1
β = 1.2
β = 1.4
β = 1.3
β = 1.3
β = 1.5
β = 1.1
β = 1.1
β = 1.2
β = 1.3
β = 1.3
β = 1.4
β = 1.4
β = 1.5
Figure 5. Comparison of intracontinental basin subsidence curves (numbered heavy lines from Fig. 4) with post-
rift thermal subsidence curves calculated from the McKenzie (1978) stretching model. Time (m.y.) is shown since
basin formation. Thin solid lines assume lithosphere thickness of 125 km; dashed lines assume lithosphere thick-
ness of 200 km. Stretching factors (β) from 1.1 to 1.5 are shown.
Basin subsidence
Geological Society of America Bulletin, January/February 2009 61
curves from passive margins. In addition, some
forearc basins show large uplift events, such as
in the Indonesian forearc basin (Beaudry and
Moore, 1985). Other basins, such as the Chilean
forearc, show signifi cant amounts of rotation
and widening of basin fi lls over time (Coul-
bourn and Moberly, 1977). Most of the curves
show less than 2 km of tectonic subsidence. The
modern Tonga forearc is exceptionally deep
(Fig. 7, line 4).
The range of shapes of subsidence curves in
this setting indicates that a variety of factors may
contribute to forearc basin subsidence. Most
curves are relatively simple in form and imply
a monotonic driving mechanism. The curves
from the Great Valley of California (Fig. 7, line
1), exhibit an abrupt change in subsidence rate
possibly refl ective of a change in driving mecha-
nism. The Great Valley curves also show very
different timings of infl ection points indicating
that the basin is tectonically segmented into dif-
ferentially subsiding zones. Basin segmentation
is seen elsewhere (Izart et al., 1994) and may
be common in these settings. Episodic subsi-
dence and even uplift of some basins is seen in
Figure 7, although it is not clear to what extent
these may refl ect errors in bathymetric assign-
ments. Causes of segmentation include parti-
tioned strain associated with oblique subduc-
tion (Izart et al., 1994), bathymetric changes in
the underlying subducted slab that isostatically
impact the overlying plate (Kobayashi, 1995),
and collision of crustal fragments in the subduc-
tion zone (Clift and MacLeod, 1999).
Possible subsidence mechanisms in forearc
basins include growth, loading, and under-
plating of the accretionary prism, which may
drive tectonic rotation and basin widening in
some settings (Coulbourn and Moberly, 1977).
Basin growth has also been tied to an increase
in width of the arc-trench gap due to fl attening
of the underthrusted plate and resultant migra-
tion of the accretionary wedge and volcanic arc
(Dickinson, 1995). Regional isostatic effects of
changing lithospheric thickness and density due
to age and structure of the underthrusted plate
can account for segmentation, and even uplift,
of forearc basin subsidence (Moxon and Gra-
ham, 1987). Of course, compression associated
with the coupling of the upper and lower plates
across convergent margins suggests that fold-
ing and thrust loading may contribute to subsi-
dence (Fuller et al., 2006). However, extensional
faulting may contribute to subsidence in some
forearc settings (Izart et al., 1994; Unruh et
al., 2007). Thermal subsidence associated with
either cooling of the fl ank of the adjacent arc
massif (Moxon and Graham, 1987) or cooling
of an accreted warm microplate (Angevine et al.,
1990) are possible mechanisms. In fact, thermal
subsidence in forearc settings can be accelerated
due to refrigeration by the underthrusted plate
(Mikhailov et al., 2007). Clift and MacLeod
(1999) discuss the role of tectonic erosion by
the down-going slab as a cause of subsidence
and tilting of the forearc basin. Subsidence of
forearc basins is the least understood and most
poorly constrained of the tectonic settings
explored in this study.
SUMMARY
Tectonic setting exerts primary control on sed-
imentary basin subsidence history. Several basin
settings seem to have distinctive subsidence pat-
terns suggesting a limited range of driving mech-
anisms. As such, calculated subsidence history is
a potential tool for identifying tectonic setting of
ancient basins of unknown origin. Passive mar-
gins show rapid initial synrift subsidence fol-
lowed by prolonged thermal subsidence similar
to that seen for subsiding seafl oor. Strike-slip
basins all demonstrate rapid, albeit short-lived,
subsidence. Foreland basins are characterized by
segmented convex-up subsidence. Intracontinen-
tal basins studied here show long-lived gradual
subsidence. While overall the subsidence pat-
tern of intracontinental basins is consistent with
thermal subsidence of thick lithosphere, most
profi les contain large deviations from predicted
Figure 6. Tectonic subsidence of foreland basins. Locations
shown in Figure 1. Thermal decay curve (dashed) for subsi-
dence of cooling seafl oor (Stein and Stein, 1992), minus 1500 m,
is shown for comparison. 1—Eastern Avalonia, Anglo-Brabant
fold belts (van Grootel et al., 1997); 2—Southern Alberta Basin
(Gillespie and Heller, 1995); 3—San Rafael Swell, Utah (Heller
et al., 1986); 4—Pyrenean foreland basin, Gombrèn (Vergés et
al., 1998); 5—Swiss Molasse basin (Burkhard and Sommaruga,
1998) modifi ed from total subsidence using water:sediment
density contrast); 6—Hoback Basin, Wyoming (Cross, 1986);
7—Green River Basin, Wyoming (Cross, 1986; Heller et al.,
1986); 8—Magallanes Basin (Biddle et al., 1986).
300 200 100 0
0
2
4
kilometers
300 200 100
Ma
Ma
0
0
2
kilometers
FORELAND BASINS
1
3
2
4
5
6
7
8a
8b
4
S
e
a
f
l
o
o
r
s
u
b
s
i
dence
S
e
a
f
l
o
o
r
s
u
b
s
i
dence
Xie and Heller
62 Geological Society of America Bulletin, January/February 2009
thermal curves suggestive of tectonic reactiva-
tion. Forearc basins studied have various subsi-
dence profi les, suggesting there may be a range
of driving mechanisms of subsidence.
Although subsidence analysis can be a use-
ful tool in identifying tectonic setting in ancient
sequences, second-order variations in the sub-
sidence rate provides specifi c information on
important local details of driving forces and
tectonic timing. Finally, caution should be used
given the limitation of data sets available for this
compilation and because basins can span mul-
tiple tectonic settings over time and/or space.
This approach is best used in conjunction with
other structural and basin analysis techniques.
ACKNOWLEDGMENTS
We would like to thank Philip Allen, Barbara
Carrapa, Ken Dueker, William Dickinson, Steve
Graham, Andrew Hynes, Michelle Kominz, Ranie
Lynds, and Osamu Takano for reviews, discussions,
and useful references.
REFERENCES CITED
Allen, P.A., and Allen, J.R., 2005, Basin analysis: Principles
and applications: Oxford, Blackwell Scientifi c Publi-
cations, 549 p.
Angevine, C.L., Heller, P.L., and Paola, C., 1990, Quanti-
tative sedimentary basin modeling: Tulsa, Oklahoma,
American Association of Petroleum Geologists, Con-
tinuing Education Course Note Series no. 32, 256 p.
Armin, R.A., and Mayer, L., 1983, Subsidence analysis of
the Cordilleran miogeocline: Implications for timing
of late Proterozoic rifting and amount of extension:
Geology, v. 11, no. 12, p. 702–705, doi: 10.1130/0091-
7613(1983)11<702:SAOTCM>2.0.CO;2.
Artemieva, I.M., and Mooney, W.D., 2001, Thermal thick-
ness and evolution of Precambrian lithosphere: A global
study: Journal of Geophysical Research, v. 106, no. B8,
p. 16,387–16,414, doi: 10.1029/2000JB900439.
Bachman, S.B., Lewis, S.D., Hegarty, K.A., and Schmitt,
K.R., 1983, Cretaceous basin formation and subsi-
dence patterns, northern Great Valley forearc basin,
California: Eos (Transactions, American Geophysical
Union), v. 64, no. 45, p. 833.
Beaudry, D., and Moore, G.F., 1985, Seismic stratigraphy
and Cenozoic evolution of West Sumatra forearc basin:
American Association of Petroleum Geologists Bulle-
tin, v. 69, no. 5, p. 742–759.
Beaumont, C., 1981, Foreland basins: Geophysical Journal
of the Royal Astronomical Society, v. 65, p. 291–329.
Benedicto, A., Labaume, P., Séguret, M., and Séranne, M.,
1996, Low-angle crustal ramp and basin geometry in
the Gulf of Lion passive margin: Oligocene-Aquitanian
Vistrenque graben, SE France: Tectonics, v. 15, no. 6,
p. 1192–1212, doi: 10.1029/96TC01097.
Biddle, K.T., Uliana, M.A., Mitchum, R.M., Jr., Fitzgerald,
M.G., and Wright, R.C., 1986, The stratigraphic and
structural evolution of the central and eastern Magal-
lanes Basin, southern South America, in Allen, P.A.,
and Homewood, P., eds., Foreland basins: International
Association of Sedimentologists Special Publication:
Oxford, Blackwell Scientifi c Publications, p. 41–61.
Bissell, H.J., 1974, Tectonic control of late Paleozoic and
early Mesozoic sedimentation near the hinge line of the
Cordilleran Miogeosynclinal Belt: Special Publication,
Society of Economic Paleontologists and Mineralo-
gists, v. 22, p. 83–97.
Bond, G.C., 1991, Disentangling Middle Paleozoic sea
level and tectonic events in cratonic margins and
cratonic basins of North America: Journal of Geo-
physical Research, v. 96, no. B4, p. 6619–6639, doi:
10.1029/90JB01432.
Bond, G.C., and Kominz, M.A., 1984, Construction of tec-
tonic subsidence curves for the early Paleozoic mio-
geocline, southern Canadian Rocky Mountains: Impli-
cations for subsidence mechanisms, age of breakup,
and crustal thinning: Geological Society of America
Bulletin, v. 95, p. 155–173, doi: 10.1130/0016-7606
(1984)95<155:COTSCF>2.0.CO;2.
Bott, M.H.P., 1980, Mechanisms of subsidence at passive
continental margins, in Bally, A.W., Bender, P.L.,
McGetchin, T.R., and Walcott, R.I., eds., Dynamics
of plate interiors: American Geophysical Union, Geo-
dynamics Series, v. 1, p. 27–32.
Burbank, D.W., Vergés, J., Muñoz, J.A., and Bentham, P.,
1992, Coeval hindward- and forward-imbricating
thrusting in the south-central Pyrenees, Spain: Tim-
ing and rates of shortening and deposition: Geological
Society of America Bulletin, v. 104, p. 3–17, doi: 10.11
30/0016-7606(1992)104<0003:CHAFIT>2.3.CO;2.
Burkhard, M., and Sommaruga, A., 1998, Evolution of the
western Swiss Molasse basin: Structural relations with
the Alps and the Jura belt, in Mascle, A., Puigdefábre-
gas, C., Luterbacher, H.P., and Fernàndez, M., eds.,
Cenozoic foreland basins of Western Europe: Geologi-
cal Society Special Publications: The Geological Soci-
ety of London, p. 279–298.
Carrapa, B., Bertotti, G., and Krijgsman, W., 2003, Subsi-
dence, stress regime and rotation(s) of a tectonically
active sedimentary basin within the western Alpine
Orogen: The Tertiary Piedmont Basin (Alpine domain,
NW Italy): Journal of the Geological Society, v. 208,
p. 205–227.
Cercone, K.R., 1984, Thermal history of Michigan Basin:
American Association of Petroleum Geologists Bulle-
tin, v. 68, no. 2, p. 130–136.
Chapman, D.S., and Pollack, H.N., 1977, Regional geo-
therms and lithospheric thickness: Geology, v. 5, no. 5,
p. 265–268, doi: 10.1130/0091-7613(1977)5<265:RG
ALT>2.0.CO;2.
Chen, W.-P., and Nábelek, J., 1988, Seismogenic strike-
slip faulting and the development of the North China
Basin: Tectonics, v. 7, no. 5, p. 975–989, doi: 10.1029/
TC007i005p00975.
Christie-Blick, N., and Biddle, K.T., 1985, Deformation and
basin development along strike-slip faults, in Christie-
Blick, N., and Biddle, K.T., eds.: Society of Economic
Paleontologists and Mineralogists Special Publica-
tion 37, p. 1–34.
Clift, P.D., and MacLeod, C.J., 1999, Slow rates of subduc-
tion erosion estimated from subsidence and tilting of the
Tonga forearc: Geology, v. 27, no. 5, p. 411–414, doi: 10
.1130/0091-7613(1999)027<0411:SROSEE>2.3.CO;2.
Cloetingh, S., Ben-Avraham, Z., Sassi, W., and Horvath, F.,
1996, Dynamics of basin formation and strike-slip tec-
tonics: Tectonophysics, v. 266, p. 1–10, doi: 10.1016/
S0040-1951(96)00179-5.
Cochran, J.R., 1983, Effects of fi
nite rifting times on the
development of sedimentary basins: Earth and Planetary
Science Letters, v. 66, p. 289–302, doi: 10.1016/0012-
821X(83)90142-5.
Coulbourn, W.T., and Moberly, R., 1977, Structural evidence
of the evolution of fore-arc basins off South America:
Canadian Journal of Earth Sciences, v. 14, p. 102–116.
Cross, T.A., 1986, Tectonic controls of foreland basin subsi-
dence and Laramide style deformation, western United
States, in Allen, P.A., and Homewood, P., eds., Fore-
land basins: International Association of Sedimentolo-
gists Special Publication, p. 15–39.
Crowell, J.C., 1974, Origin of Late Cenozoic basins in south-
ern California, in Dickinson, W.R., ed., Tectonics and
sedimentation: Society of Economic Paleontologists
and Mineralogists Special Publication, p. 190–204.
Crowell, J.C., and Link, M.H., eds., 1982, Geologic his-
tory of Ridge Basin, southern California: Los Angeles,
Pacifi c Section, Society of Economic Paleontologists
and Mineralogists, 304 p.
Curray, J.R., and Moore, D.G., 1984, Geologic history of
the mouth of the Gulf of California, in Crouch, J.K.,
and Bachman, S.B., eds., Tectonics and sedimenta-
tion along the California margin: Field trip guidebook,
Pacifi c Section, Society of Economic Paleontologists
and Mineralogists, v. 38, p. 17–36.
DeCelles, P.G., and Giles, K.A., 1996, Foreland basin systems:
Basin Research, v. 8, p. 105–123, doi: 10.1046/j.1365-
2117.1996.01491.x.
Ma MaMa
FOREARC BASINS
100 0
0
2
4
kilometers
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
S
e
a
f
l
o
o
r
s
u
b
s
i
d
e
n
c
e
100 0
0
2
4
kilometers
100 0
0
2
4
kilometers
2
1a
1b
1c
6
7
8
9
3
4
5
Figure 7. Tectonic subsidence curves of forearc basins. Locations shown in Figure 1. Ther-
mal decay curve (dashed) for subsidence of cooling seafl oor (Stein and Stein, 1992), minus
1500 m, is shown for comparison. 1—Great Valley (Moxon and Graham, 1987); 2—Sac-
ramento Basin (Dickinson et al., 1987); 3—Peninsular Ranges (Kimbrough et al., 2001);
4—Tonga forearc (Clift and MacLeod, 1999); 5—Japan forearc, Deep Sea Drilling Project
sites 438 and 439 (von Huene, 1982; Ingle, 1992); 6—Southern Lesser Antilles, northwest
of Tabago (Ysaccis, 1997); 7—Oregon Coast Range (Angevine et al., 1990; Heller, 1983);
8—Luzon Central Valley (Bachman et al., 1983); 9—Kazusa forearc basin (Ito, 1995).
Basin subsidence
Geological Society of America Bulletin, January/February 2009 63
Devlin, W.J., and Bond, G.C., 1988, The initiation of the
early Paleozoic Cordilleran Miogeocline: Evidence
from the uppermost Proterozoic-Lower Cambrian
Hamill Group of southeastern British Columbia: Cana-
dian Journal of Earth Sciences, v. 25, no. 1, p. 1–19.
Devlin, W.J., Bond, G.C., Brueckner, H.K., and Anonymous,
1986, Initiation of Cordilleran Miogeocline of western
North America: American Association of Petroleum
Geologists Bulletin, v. 70, no. 5, p. 581.
Dickinson, W.R., 1976, Plate tectonic evolution of sedimen-
tary basins, in Dickinson, W.R., and Yarborough, H.,
eds., Plate tectonics and hydrocarbon accumulation:
Tulsa, Oklahoma, American Association of Petroleum
Geologists, Continuing Education Course Note Series,
p. 1–62.
Dickinson, W.R., 1995, Forearc basins, in Busby, C.J., and
Ingersoll, R.V., eds., Tectonics of sedimentary basins:
Blackwell Scientifi c, p. 221–261.
Dickinson, W.R., Armin, R.A., Beckvar, N., Goodlin, T.C.,
Janecke, S.U., Mark, R.A., Norris, R.D., Radel, G.,
and Wortman, A.A., 1987, Geohistory analysis of rates
of sediment accumulation and subsidence for selected
California basins, in Ingersoll, R.V., and Ernst, W.G.,
eds., Cenozoic basin development of coastal California:
Englewood Cliffs, New Jersey, Prentice-Hall, 496 p.
Dore, A.G., and Stewart, I.C., 2002, Similarities and dif-
ferences in the tectonics of two passive margins: The
Northeast Atlantic margin and the Australian North West
Shelf, in Keep, M., and Moss, S.J., eds., Sedimentary
basins of Western Australia: Proceedings of Petroleum
Exploration Society of Australia Symposium: Perth,
Petroleum Exploration Society of Australia, p. 89–117.
Ellouz, N., Patriat, M., Gaulier, J.-M., Bouatmani, R., and
Sabounji, S., 2003, From rifting to Alpine inversion:
Mesozoic and Cenozoic subsidence history of some
Moroccan basins: Sedimentary Geology, v. 156,
p. 185–212, doi: 10.1016/S0037-0738(02)00288-9.
Falvey, D.A., and Mutter, J.C., 1981, Regional plate tecton-
ics and the evolution of Austria’s passive continental
margins: Journal of Australian Geology and Geophys-
ics, v. 6, p. 1–29.
Fowler, C.M.R., and Nisbet, E.G., 1985, The subsidence of
the Williston Basin: Canadian Journal of Earth Sci-
ences, v. 22, no. 3, p. 408–415.
Fuller, C.W., Willet, S.D., and Brandon, M.T., 2006, Forma-
tion of forearc basins and their infl uence on subduction
zone earthquakes: Geology, v. 34, no. 2, p. 65–68, doi:
10.1130/G21828.1.
Gillespie, J.M., and Heller, P.L., 1995, Beginning of fore-
land subsidence in the Columbian-Sevier belts, south-
ern Canada and northwest Montana: Geology, v. 23,
no. 8, p. 723–726, doi: 10.1130/0091-7613(1995)023
<0723:BOFSIT>2.3.CO;2.
Graham, S.A., 1976, Tertiary sedimentary tectonics of the
central Salinian block of California [Ph.D. thesis]:
Stanford University, 510 p.
Haq, B.U., Hardenbol, J., and Vail, P.R., 1987, Chronology of
uctuating sea levels since the Triassic: Science, v. 235,
p. 1156–1167, doi: 10.1126/science.235.4793.1156.
Harrison, C.G.A., 1990, Long-term eustasy and epeirogeny
in continents, in National Research Council, Geophys-
ics Study Committee, ed., Sea-level change: Washing-
ton D.C., National Academy Press, p. 141–158.
Haxby, W.F., Turcotte, D.L., and Bird, J.M., 1976, Ther-
mal and mechanical evolution of the Michigan Basin:
Tectonophysics, v. 36, p. 57–75, doi: 10.1016/0040-
1951(76)90006-8.
Heller, P.L., 1983, Sedimentary response to Eocene tectonic
rotation in western Oregon [Ph.D. thesis]: Tucson, Ari-
zona, University of Arizona, 321 p.
Heller, P.L., Bowdler, S.S., Chambers, H.P., Coogan, J.C.,
Hagen, E.S., Shuster, M.W., and Winslow, N.S., 1986,
Time of initial thrusting in the Sevier orogenic belt,
Idaho-Wyoming and Utah: Geology, v. 14, no. 5,
p. 388–391, doi: 10.1130/0091-7613(1986)14<388:T
OITIT>2.0.CO;2.
Hendrie, D.B., Kusznir, N.J., Morley, C.K., and Ebinger,
C.J., 1994, Cenozoic extension in northern Kenya: A
quantitative model of rift basin development in the Tur-
kana region: Tectonophysics, v. 236, p. 409–438, doi:
10.1016/0040-1951(94)90187-2.
Howell, P.D., and van der Pluijm, B.A., 1999, Structural
sequences and styles of subsidence in the Michigan
basin: Geological Society of America Bulletin, v. 111,
no. 7, p. 974–991, doi: 10.1130/0016-7606(1999)111<
0974:SSASOS>2.3.CO;2.
Hu, S., O’Sullivan, P.B., Raza, A., and Kohn, B.P., 2001,
Thermal history and tectonic subsidence of the Bohai
Basin, northern China: A Cenozoic rifted and local
pull-apart basin: Physics of the Earth and Planetary
Interiors, v. 126, p. 221–235, doi: 10.1016/S0031-
9201(01)00257-6.
Hunt, C.B., and Mabey, D.R., 1966, Stratigraphy and struc-
ture Death Valley, California, U.S. Geological Survey
Professional Paper 494-A, 162 p.
Ingersoll, R.V., 1988, Development of the Cretaceous
forearc basin of central California, in Graham, S.A.,
and Olson, H.C., eds., Studies of the geology of the San
Joaquin basin: Los Angeles, Pacifi c Section, Society of
Economic Paleontologists and Mineralogists, Field
Trip Guide, v. 60, p. 141–155.
Ingle, J.C., Jr., 1992, Subsidence of the Japan Sea: Strati-
graphic evidence from ODP sites and onshore sections,
Proceedings of the Ocean Drilling Program, Scientifi c
Results: College Station, Texas, Texas A&M Univer-
sity, Ocean Drilling Program, p. 1197.
Ito, M., 1995, Volcanic ash layers facilitate high-resolution
sequence stratigraphy at convergent plate margins:
An example from the Plio-Pleistocene forearc basin
ll in the Boso Peninsula, Japan: Sedimentary Geol-
ogy, v. 95, no. 3–4, p. 187–206, doi: 10.1016/0037-
0738(94)00108-7.
Izart, A., Kemal, B.M., and Malod, J.A., 1994, Seismic
stratigraphy and subsidence evolution of the north-
west Sumatra fore-arc basin: Marine Geology, v. 122,
p. 109–124, doi: 10.1016/0025-3227(94)90207-0.
Jarvis, G.T., and McKenzie, D.P., 1980, Sedimentary
basin formation with fi nite extension rates: Earth
and Planetary Science Letters, v. 48, p. 42–52, doi:
10.1016/0012-821X(80)90168-5.
Johnson, S.Y., 1984, Stratigraphy, age, and paleogeography
of the Eocene Chuckanut Formation, northwest Wash-
ington: Canadian Journal of Earth Sciences, v. 21,
p. 92–106.
Johnson, S.Y., 1985, Eocene strike-slip faulting and nonma-
rine basin formation in Washington, in Biddle, K.T.,
and Christie-Blick, N., eds., Strike-slip deformation,
basin formation and sedimentation, Society of Eco-
nomic Paleontologists and Mineralogists, Special Pub-
lication, p. 283–302.
Jones, M.A., Heller, P.L., Roca, E., Garces, M., and Cabrera,
L., 2004, Time lag of syntectonic sedimentation across
an alluvial basin: Theory and example from the Ebro
Basin, Spain: Basin Research, v. 16, no. 4, p. 489–506,
doi: 10.1111/j.1365-2117.2004.00244.x.
Jordan, T.E., 1982, Tectonic loading and foreland basin
subsidence: International Congress on Sedimentology,
v. 11, p. 131.
Kaminski, E., and Jaupart, C., 2000, Lithosphere struc-
ture beneath the Phanerozoic intracratonic basins of
North America: Earth and Planetary Science Letters,
v. 178, no. 1–2, p. 139–149, doi: 10.1016/S0012-821X
(00)00067-4.
Karner, G.D., and Dewey, J.F., 1986, Rifting: Lithospheric
versus crustal extension as applied to the Ridge Basin
of southern California, in Halbouty, M.T., ed., Future
petroleum province of the world: Tulsa, American Asso-
ciation of Petroleum Geologists Memoir, p. 317–337.
Kerr, D.R., Pappajohn, S., and Peterson, G.L., 1979, Neo-
gene stratigraphic section at Split Mountain, eastern
San Diego County, California, in Crowell, J.C., and
Sylvester, A.G., eds., Tectonics of the juncture between
the San Andreas fault system and the Salton Trough,
southeastern California: San Diego, California, Geo-
logical Society of America Annual Meeting, Field Trip
Guidebook, p. 111–123.
Kimbrough, D.L., Smith, D.P., Mahoney, J.B., Moore, T.E.,
Grove, M., Gastil, R.G., Ortega-Rivera, A., and Fan-
ning, C.M., 2001, Forearc-basin sedimentary response
to rapid Late Cretaceous batholith emplacement in the
Peninsular Ranges of southern and Baja California:
Geology, v. 29, no. 6, p. 491–494, doi: 10.1130/0091-7
613(2001)029<0491:FBSRTR>2.0.CO;2.
King, G., and Ellis, M., 1990, The origin of large local uplift
in extensional regions: Nature, v. 348, p. 689–693, doi:
10.1038/348689a0.
Klein, G.D., and Hsui, A.T., 1987, Origin of cratonic basins:
Geology, v. 15, p. 1094–1098, doi: 10.1130/0091-7613
(1987)15<1094:OOCB>2.0.CO;2.
Kobayashi, K., 1995, Role of subducted lithospheric slab
in uplift and subsidence of the northwestern Pacifi c
margins: Marine Geology, v. 127, p. 119–144, doi:
10.1016/0025-3227(95)00007-L.
Kominz, M.A., Werkema, D., Barnes, D.A., Harrison, W.,
III, Kirwan, E., Malin, M., and Anonymous, 2001, The
Michigan Basin is thermal in origin: American Asso-
ciation of Petroleum Geologists Bulletin, v. 85, no. 8,
p. 1533–1534.
Kusznir, N.J., and Ziegler, P.A., 1992, The mechanics of
continental extension and sedimentary basin forma-
tion: A simple-shear/pure-shear fl exural cantilever
model: Tectonophysics, v. 215, no. 1–2, p. 117–131,
doi: 10.1016/0040-1951(92)90077-J.
Kusznir, N.J., Kovkhuto, A.M., and Stephenson, R.A.,
1996a, Syn-rift evolution of the Pripyat Trough: Con-
straints from structural and stratigraphic modelling:
Tectonophysics, v. 268, no. 1–4, p. 221–236, doi:
10.1016/S0040-1951(96)00231-4.
Kusznir, N.J., Stovba, S.M., Stephenson, R.A., and
Poplavskii, K.N., 1996b, The formation of the north-
western Dniepr-Donets Basin: 2-D forward and reverse
syn-rift and post-rift modelling: Tectonophysics, v. 268,
p. 237–255, doi: 10.1016/S0040-1951(96)00230-2.
Mann, P., Bradley, D., and Hempton, M.R., 1982, Empirical
model of pull-apart evolution: International Congress
on Sedimentology, v. 11, p. 38–39.
Mann, P., Hempton, M., Bradley, D., and Burke, K., 1983,
Development of pull-apart basins: The Journal of Geol-
ogy, v. 91, p. 529–554.
May, S.R., Ehman, K.D., Gray, G.G., and Crowell, J.C.,
1993, A new angle on the tectonic evolution of the
Ridge Basin, a “strike-slip” basin in southern Califor-
nia: Geological Society of America Bulletin, v. 105,
no. 10, p. 1357–1372, doi: 10.1130/0016-7606(1993)
105<1357:ANAOTT>2.3.CO;2.
McKenzie, D.P., 1978, Some remarks on the development of
sedimentary basins: Earth and Planetary Science Letters,
v. 40, p. 25–32, doi: 10.1016/0012-821X(78)90071-7.
Mikhailov, V.O., Parsons, T., Simpson, R.W., Timoshkina,
E.P., and Williams, C., 2007, Why the Sacramento
Delta area differs from other parts of the Great Valley:
Numerical modeling of thermal structure and thermal
subsidence of forearc basins: Izvestiya Physics of
the Solid Earth, v. 43, no. 1, p. 75–90, doi: 10.1134/
S1069351307010089.
Mohriak, W.U., Karner, G.D., and Dewey, J.F., 1987, Subsi-
dence history and tectonic evolution of Campos Basin,
offshore Brazil: American Association of Petroleum
Geologists Bulletin, v. 71, no. 5, p. 594.
Moxon, I.W., and Graham, S.A., 1987, History and controls
of subsidence in the Late Cretaceous-Tertiary Great
Valley forearc basin, California: Geology, v. 15, no. 7,
p. 626–629, doi: 10.1130/0091-7613(1987)15<626:HAC
OSI>2.0.CO;2.
Nadin, P.A., and Kusznir, N.J., 1995, Palaeocene uplift and
Eocene subsidence in the northern North Sea Basin
from 2D forward and reverse stratigraphic modelling:
Journal of the Geological Society, v. 152, p. 833–848,
doi: 10.1144/gsjgs.152.5.0833.
Nadon, G.C., and Issler, D.R., 1997, The compaction of
oodplain sediments: Timing, magnitude and implica-
tions: Geoscience Canada, v. 24, no. 1, p. 37–43.
Nielsen, S.B., Paulsen, G.E., Hansen, D.L., Gemmer, L.,
Clausen, O.R., Jacobsen, B.H., Balling, N., Huuse, M.,
and Gallagher, K., 2002, Paleocene initiation of Ceno-
zoic uplift in Norway, in Dore, A.G., Cartwright, J.A.,
Stoker, M.S., Turner, J.P., and White, N.J., eds., Exhu-
mation of the North Atlantic: Timing, mechanisms and
implications for petroleum exploration: Geological
Society Special Publications: The Geological Society
of London, p. 45–65.
Nilsen, T.H., and McLaughlin, R.J., 1985, Comparison of
tectonic framework and depositional patterns of the
Hornelen strike-slip basin of Norway and the Ridge
and Little Sulphur Creek strike-slip basins of Califor-
nia: Special Publication, Society of Economic Paleon-
tologists and Mineralogists, v. 37, p. 79–103.
Nunn, J.A., and Sleep, N.H., 1984, Thermal contraction and
exure of intracratonal basins: A three-dimensional
Xie and Heller
64 Geological Society of America Bulletin, January/February 2009
study of the Michigan basin: Geophysical Journal of
the Royal Astronomical Society, v. 76, p. 587–635.
Nunn, J.A., Sleep, N.H., and Moore, W.E., 1984, Thermal
subsidence and generation of hydrocarbons in Michi-
gan Basin: American Association of Petroleum Geolo-
gists Bulletin, v. 68, p. 296–315.
Ori, G.G., and Friend, P.F., 1984, Sedimentary basins formed
and carried piggyback on active thrust sheets: Geology,
v. 12, p. 475–478, doi: 10.1130/0091-7613(1984)12<4
75:SBFACP>2.0.CO;2.
Pitman, W.C., III, and Andrews, J.A., 1985, Subsidence
and thermal history of small pull-apart basins: Special
Publication, Society of Economic Paleontologists and
Mineralogists, v. 37, p. 45–119.
Prijac, C., Doin, M.P., Gaulier, J.M., and Guillocheau, F.,
2000, Subsidence of the Paris Basin and its bearing on
the late Variscan lithosphere evolution: A comparison
between plate and CHABLIS models: Tectonophys-
ics, v. 323, no. 1–2, p. 1–38, doi: 10.1016/S0040-
1951(00)00100-1.
Roberts, A.M., Yielding, G., Kusznir, N.J., Walker, I.M.,
and Dorn-Lopez, D., 1995, Quantitative analysis of
Triassic extension in the northern Viking Graben: The
Geological Society of London, v. 152, p. 15–26, doi:
10.1144/gsjgs.152.1.0015.
Rumelhart, P.E., and Ingersoll, R.V., 1997, Provenance of
the upper Miocene Modelo Formation and subsidence
analysis of the Los Angeles Basin, southern California:
Implications for paleotectonic and paleogeographic
reconstructions: Geological Society of America Bul-
letin, v. 109, no. 7, p. 885–899, doi: 10.1130/0016-760
6(1997)109<0885:POTUMM>2.3.CO;2.
Sawyer, D.S., Hsui, A.T., and Toksoz, M.N., 1987, Extension,
subsidence, and thermal evolution of the Los Angeles
Basin—A two-dimensional model: Tectonophysics,
v. 133, p. 15–32, doi: 10.1016/0040-1951(87)90277-0.
Scheck, M., and Bayer, U., 1999, Evolution of the Northeast
German Basin: Inferences from a 3D structural model
and subsidence analysis: Tectonophysics, v. 313, no. 1–2,
p. 145–169, doi: 10.1016/S0040-1951(99)00194-8.
Sclater, J.G., and Christie, P.A.F., 1980, Continental stretch-
ing: An explanation of the post-mid-Cretaceous subsi-
dence of the central North Sea basin: Journal of Geo-
physical Research, v. 85, p. 3711–3739, doi: 10.1029/
JB085iB07p03711.
Sleep, N.H., and Sloss, L.L., 1980, The Michigan Basin, in
Bally, A.W., Bender, P.L., McGetchin, T.R., and Walcott,
R.I., eds., Dynamics of plate interiors: American Geo-
physical Union, Geodynamics Series, v. 1, p. 93–98.
Stapel, G., Cloetingh, S., and Pronk, B., 1996, Quantitative
subsidence analysis of the Mesozoic evolution of the
Lusitanian Basin (western Iberian margin): Tectono-
physics, v. 266, no. 1–4, p. 493–507, doi: 10.1016/
S0040-1951(96)00203-X.
Steckler, M.S., and Watts, A.B., 1978, Subsidence of the
Atlantic type continental margin off New York: Earth
and Planetary Science Letters, v. 41, p. 1–13, doi:
10.1016/0012-821X(78)90036-5.
Steckler, M.S., and Watts, A.B., 1980, The Gulf of Lion:
Subsidence of a young continental margin: Nature,
v. 287, p. 425–429, doi: 10.1038/287425a0.
Stein, C.A., and Stein, S., 1992, A model for the global
variation in oceanic depth and heat fl ow with litho-
spheric origin: Nature, v. 359, no. 6391, p. 123–129,
doi: 10.1038/359123a0.
Storti, F., Holdsworth, R., and Salvini, F., 2003, Intraplate
strike-slip deformation belts, in Storti, F., Holdsworth,
R., and Salvini, F., eds., Intraplate strike-slip deforma-
tion belts: The Geological Society of London, p. 1–14.
Sutherland, R., and Melhuish, A., 2000, Formation of evolu-
tion of the Solander Basin, southwestern South Island,
New Zealand, controlled by a major fault in conti-
nental crust and upper mantle: Tectonics, v. 19, no. 1,
p. 44–61, doi: 10.1029/1999TC900048.
Swift, B.A., Sawyer, D.S., Grow, J.A., and Klitgord, K.D.,
1987, Subsidence, crustal structure, and thermal evolu-
tion of Georges Bank Basin: American Association of
Petroleum Geologists Bulletin, v. 71, no. 6, p. 702–718.
Sylvester, A.G., 1988, Strike-slip faults: Geological Society
of America Bulletin, v. 100, p. 1666–1703, doi: 10.113
0/0016-7606(1988)100<1666:SSF>2.3.CO;2.
Turcotte, D.L., 1980, Models for the evolution of sedimen-
tary basins, in Bally, A.W., Bender, P.L., McGetchin,
T.R., and Walcott, R.I., eds., Dynamics of plate inte-
riors: American Geophysical Union, Geodynamics
Series, v. 1, p. 21–26.
Unruh, J.R., Dumitru, T.A., and Sawyer, T.L., 2007, Coupling
of early Tertiary extension in the Great Valley forearc
basin with blueschist exhumation in the underlying
Franciscan accretionary wedge at Mount Diablo, Cali-
fornia: Geological Society of America Bulletin, v. 119,
no. 11–12, p. 1347–1367, doi: 10.1130/B26057.1.
van Grootel, G., Verniers, J., Geerkens, B., Laduron, D., Ver-
haeren, M., Hertogen, J., and De Vos, W., 1997, Timing
of magmatism, foreland basin development, metamor-
phism and inversion in the Anglo-Brabant fold belt:
Geological Magazine, v. 134, no. 5, p. 607–616, doi:
10.1017/S0016756897007413.
van Hinte, J.E., 1978, Geohistory analysis—Application of
micropaleontology in exploration geology: American
Association of Petroleum Geologists Bulletin, v. 62,
p. 201–222.
Vergés, J., Marzo, M., Santaeularia, T., Serra-Kiel, J., Bur-
bank, D., Muñoz, J.A., and Gimenez-Montsant, J.,
1998, Quantifi ed vertical motions and tectonic evolu-
tion of the SE Pyrenean foreland basin, in Mascle, A.,
Puigdefábregas, C., Luterbacher, H.P., and Fernandez,
M., eds., Cenozoic foreland basins of Western Europe:
The Geological Society of London, p. 107–134.
von Huene, R., 1982, Sedimentation across the Japan Trench
off northern Honshu Island, in Arthur, M.A., ed.: The
Geological Society of London Special Publication, p. 27.
Waldron, J., 2005, Extensional fault arrays in strike-slip
and transtension: Journal of Structural Geology, v. 27,
p. 23–34, doi: 10.1016/j.jsg.2004.06.015.
Watts, A.B., 1981, The U.S. Atlantic continental margin:
Subsidence history, crustal structure and thermal evo-
lution, in Bally, A.W., Watts, A.B., Grow, J.A., Mans-
peizer, W., Bernoulli, D., Schreiber, C., and Hunt, J.M.,
eds., Geology of passive continental margins: History,
structure and sedimentologic record (with special
emphasis on the Atlantic margin): American Asso-
ciation of Petroleum Geologists, Continuing Education
Course Note Series no. 19, 75 p.
Watts, A.B., and Ryan, W.B.F., 1976, Flexure of the litho-
sphere and continental margin basins: Tectonophysics,
v. 36, p. 25–44, doi: 10.1016/0040-1951(76)90004-4.
Wernicke, B., Spencer, J.E., Burchfi el, B.C., and Guth, P.L.,
1982, Magnitude of crustal extension in the southern
Great Basin: Geology, v. 10, p. 499–502, doi: 10.1130/
0091-7613(1982)10<499:MOCEIT>2.0.CO;2.
Xie, X.Y., 2007, Sedimentary record of Mesozoic intracon-
tinental deformation in the south Ordos Basin, China
[Ph.D. thesis]: University of Wyoming, 280 p.
Ysaccis, R., 1997, Tertiary evolution of the northeastern Ven-
ezuela offshore [Ph.D. thesis]: Rice University, 285 p.
Zalan, P.V., Wolff, S., Astolfi , M.A.M., Santos Viera, I.,
Conceicao, J.C.J., Appi, V.T., Neto, E.V.S., Cerqueira,
J.R., and Marques, A., 1990, The Parana Basin, Brazil:
American Association of Petroleum Geologists Mem-
oir, v. 51, p. 681–708.
MANUSCRIPT RECEIVED 31 DECEMBER 2007
R
EVISED MANUSCRIPT RECEIVED 15 JUNE 2008
M
ANUSCRIPT ACCEPTED 18 JUNE 2008
Printed in the USA
... The knowledge about the tectonic evolution of a given sedimentary basin is extremely useful in understanding the controls on the sediment accumulation and to quantify and hierarchize the different tectonic phases acting during its origin, filling and deformation. Consequently, an important tool to understand the tectono-sedimentary evolution is the subsidence analysis, which allows isolating and inferring the basinforming mechanisms (Ingersoll and Busby, 1995;Xie and Heller, 2009;Ingersoll, 2012;Allen and Allen, 2013;Lee et al., 2018;among others), in several geological basins recorded from Precambrian to Cenozoic times (Poprawa et al., 1999;Scheck and Bayer, 1999;Lee and Wagreich, 2017;Rodríguez Tribaldos and White, 2018;Scivetti and Franzese, 2019;Scivetti et al., 2021;among others). However, it is important to emphasize that some of the main problems in the analysis of Precambrian basins include uncertainties associated with tectonic plates during this time interval such as heat flow behavior, evolution and rate, estimations about erosion and sediment supply rates, and limited rock record (Eriksson et al., 2001). ...
... The paleo-water depth constitutes an important parameter for the conception of the subsidence model, and errors in water-depth determinations can be reduced by analyzing sections that are mostly composed of shallow marine deposits (Xie and Heller, 2009;Rodríguez Tribaldos and White, 2018). For each lithostratigraphic unit, the paleowater depth was estimated based on facies analysis and fossil assemblage described in the bibliography (Poiré, 1993;Poiré et al., 2018;Gómez-Peral et al., 2014;2019;. ...
... In the case of Precambrian basins, the protracted stretching (up to 50 Ma) (Armitage and Allen, 2010) at low strain rates over thick (>150 km) Precambrian lithosphere (Rodríguez Tribaldos and White, 2018) results in cooling advection corresponding to thermal diffusivity. This process develops nearly sublinear -to a gently negative exponential (Xie and Heller, 2009) -subsidence trajectories, eventually altered by secondary subsidence mechanisms (Armitage and Allen, 2010). As well as in some Paleozoic examples, their sedimentary covers are relatively thin, and the stratigraphic successions are bounded by basin-wide erosional unconformities (Sloss and Speed, 1974;Allen and Armitage 2012;Rodríguez Tribaldos and White, 2018). ...
Article
Several studies on the Neoproterozoic sedimentary cover of the Tandilia System, Argentina, have been carried out in recent decades. The integration of this knowledge and new results have allowed us to carry out an integral analysis including backstripping technique, forward models, sedimentology, and stratigraphy to provide an integral , innovative, and robust subsidence model to classify this Neoproterozoic basin. The water-loaded subsidence values obtained for the analyzed areas indicate extremely-low accommodation space creation rates. On the other hand, the forward models determined that the initial thickness of the crust and the lithosphere would have been around 35 km and 200 km, respectively, upon which a stretching factor of around 1.09 would have acted. Considering the water-loaded subsidence values, together with the stretch duration, the geometry of the subsidence mechanism, the type and characteristics of the stratigraphic succession and discontinuities, it is feasible to propose that the Neoproterozoic sedimentary units of the Tandilia Basin were accumulated in a cratonic basin that would have originated by processes associated with the fragmentation of the Rodinia Supercontinent. During its evolution, the creation of accommodation space would have been mainly controlled by background subsidence. At the same time, the eustatic oscillations, together with uplift episodes related to lithospheric deglaciation unload, would have subordinately controlled the creation of accommodation space.
... incluyéndose los cambios en el nivel del mar. Por lo tanto, los datos necesarios para reconstruir el historial de subsidencias incluyen el espesor estratigráfico, la litología, la estimación de las profundidades del paleo-agua y el control de la edad (Xie & Heller, 2006). Para el análisis de la acumulación de sedimentos a lo largo del tiempo se utiliza el espesor actual de cada unidad estratigráfica fechada, tal como ha sido estimada en los pocos pozos con esta información disponibles en la Cuenca del Salado. ...
... La subsidencia de las cuencas intracontinentales es similar en magnitud a la observada en entornos de margen pasivo, pero la primera es generalmente más lenta, de vida más larga y carece de subsidencia inicial (Dressel et al., 2017). El hundimiento de larga duración para muchas cuencas intracontinentales es consistente con el enfriamiento después de la perturbación térmica de la litosfera gruesa que se encuentra debajo de las partes antiguas de los continentes (Xie & Heller, 2006). El conocimiento de la evolución reciente de la desembocadura de la cuenca del Amazonas presenta analogías tectosedimentarias con la Cuenca del Salado que contribuyen con el análisis morfológico de la Bahía Samborombón. ...
Article
La Bahía de Samborombón está ubicada en la costa bonaranse, en elestuario del Río de la Plata. Presenta una característica morfología semicircular, deaproximadamente 100 kilómetros de longitud, entre Punta Piedras (al norte) y elCabo San Antonio (al sur). La coincidencia con la subyacente Cuenca del Saladopermite trazar relaciones espacio-temporales entre su morfología y su historiatectonosedimentaria. La exhumación andina aportó sedimentos al margen pasivoatlántico. En el Plioceno se produjo una disminución de la tasa de sedimentacióndebido al relleno de los ambientes de la Llanura Pampeana. La interposición delabanico aluvial de Córdoba contra el Alto San Guillermo generó la laguna de MarChiquita y cambios en el drenaje. Así, la tasa de sedimentación ~0,09 mm/año de laFormación Las Chilcas (Maastrichtiano-Paleoceno) se reducen a ~0,005 mm/año dela Formación Puelches (Plioceno-Pleistoceno), con un promedio de 0,042 mm/añopara las formaciones eocenas Los Cardos y Olivos. El progresivo balance negativoentre la tasa de sedimentación vs. subsidencia determina que la Cuenca del Salado esun depocentro hambriento (altamente receptivo) que ha favorecido el avance del marhacia el interior de la Bahía de Samborombón. De este modo, existen conchales,humedales, bajos y marismas donde el desarrollo de las tierras altas desde el surfavoreció la progresiva continentalización.
... Extensive/passive geodynamic contexts exhibit the greatest thicknesses and preservation rate of carbonate platforms (255-1040 m; 25-80 m/Ma), whereas compressive and intra/peri-cratonic contexts (145-275 m; 20-40 m/Ma) show relatively lower thicknesses (Figure 8). The thickness trend is directly associated with a subsidence trend: extensive systems showing greater subsidence rates and height than intra/peri-cratonic systems (see Xie and Heller, 2009). If compressive systems can show large subsidence patterns, such systems appear not to be the most favorable context for carbonate platform development, which might be related to relatively unstable paleoenvironmental conditions and large terrigenous input (Bosence, 2005;Tassy et al., 2023). ...
Article
Full-text available
Carbonate systems are influenced by a great variety of physical and biological controlling factors that operate from global to local scales. The resulting intrinsic complexity of carbonate platforms makes them difficult to predict, especially when data are limited. Predicting geologic geometries and properties based on limited sampling or uncalibrated seismic data generally relies on a priori knowledge and equivocal interpretations that are marked by geologist perception and personal experience. To overcome these uncertain interpretations of such a complex natural system, which can become critical in frontier exploration, we developed an expert system that relies on a process-based method and a standardized data set using normalized information and parameters. The main innovation relies on the realization of knowledge- and process-based synthetic carbonate stratigraphic architectures that support seismo-stratigraphic interpretations. The workflow consists of four steps: (1) bibliographic compilation of a geologic database for each case study supported by quantitative parameters (e.g., sedimentation duration and thickness) and qualitative parameters (geodynamic context, seismic architecture, and facies model); (2) statistical analyses to establish consistent geologic classes and spatiotemporal trends; (3) process-based modeling to simulate stratigraphic architectures associated with carbonate sedimentation processes in a physically constrained numerical environment and testing different geologic hypotheses; and (4) realization of a predictive palaeogeographic map representing the global distribution of carbonate stratigraphic architectures, and estimation of controlling parameters for unconstrained case studies. The expert system is based on 77 case studies of Upper Jurassic carbonate platforms, which reveal the resemblance of these carbonate systems, in response to uniform global palaeoclimatic conditions and sea level. Significant local differences in stratigraphic architectures are related to specific geodynamic contexts and subsidence trends. The thickest carbonate platforms are developed in extensive/passive geodynamic settings such as the Central Atlantic Ocean margins, while thinner platforms form in intra- and peri-cratonic settings such as those of the Arabian region.
... The Bohai Bay Basin contains seven secondary basins: the Liaohe, Liaodong Bay, Bozhong, Jizhong, Huanghua, Jiyang, and Linqing subbasins (Fig. 1b;Qiu et al., 2015;Liu et al., 2022a). The middle part of the TLFZ extends through the east of Bohai Bay Basin, which has been strongly influenced by intense normal faulting and modified by strike-slip faulting (Fig. 1b), leading to three phases of deformation since the Cenozoic: intense extension with rapid subsidence during 65-40 Ma; intense transtension during 40-23 Ma; and weak transtension from 23 Ma to the present (Xie and Heller, 2009;Wang et al., 2022). ...
... Subsequently, the thermal relaxation of the lithosphere begins through conductive cooling and thermal subsidence. The duration of both the syn-and post rift phases vary significantly, however, reaching equilibrium (steady-state) typically takes several tens to hundreds of million years Xie and Heller, 2009;Petersen et al., 2015). 25 ...
Preprint
Full-text available
The reconstruction of thermal evolution in sedimentary basins is a key input for constraining geodynamic processes and geo-energy resource potential. We present a methodology to reproduce the most important transient thermal footprints accompanying basin formation: lithosphere extension and sedimentation. The forward model is extended with data assimilation to constrain models with temperature measurements. We apply the methodology to the NW part of Hungary. Realistic past- and present-day temperature predictions for the entire lithosphere are achieved, suggesting the relatively uniform, but strong attenuation of the mantle lithosphere through extension, and relatively small variations in the present-day thermal lithosphere thickness. The new temperature model allows an improved estimation of lithosphere rheology and the interpretation of mantle xenolith origins.
... The subsidence curves obtained from backstripping (see the yellow stars in Figure 2 for location of curves) show an overall 180 convex-up shape, which is mainly characterized by rapid tectonic subsidence after ~20 Ma (Fig. 4), consistent with other examples of foreland basin subsidence curves (e.g., Xie and Heller, 2009). Flexural modeling of the Neogene Zagros tectonic and sedimentary loads (Fig. 5) demonstrates that both loads produce a shallower accommodation than we observed for the Neogene basin (base of Fatha Formation). ...
Preprint
Full-text available
Tectonic processes resulting from solid Earth dynamics control uplift and generate sediment accommodation space via subsidence. Unraveling the mechanism of basin subsidence elucidates the link between deep Earth and Surface processes. The NW Zagros fold-thrust belt results from the Cenozoic convergence and subsequent collision between the Arabian and Eurasian plates. The associated Neogene foreland basin includes ~4 km of syntectonic nonmarine clastic sediments, suggesting a strongly subsiding basin inconsistent with the adjacent topographic load. To explain such discrepancy, we assessed basin subsidence with respect to the effect of surface load and dynamic topography. The isopach map of the Fatha Formation during the middle Miocene displays a longitudinal depocenter aligned with the orogenic trend. In contrast, the maps of the Injana Formation and Mukdadiya Formation during the late Miocene illustrate a focused depocenter in the southern region of the basin. The rapid subsidence in the south during the late Miocene was coeval with the Afar plume flow northward beyond the Arabia-Eurasia suture zone in the northwestern segment of the Zagros belt. Based on isopach maps, subsidence curves, and reconstructions of flexural profiles, supported by Bouguer anomaly data and maps of dynamic topography and seismic tomography, we argue for a two-stage basin evolution. The Zagros foreland basin subsided due to the load of the surface and the subducting slab during the early-middle Miocene and was later affected by the Neothethys horizontal slab tear propagation during the late Miocene. This tear propagation was associated with a northward mantle flow above the detached segment in the NW and a focussed slab pull on the attached portion of the slab in the SE.
... Frontal fold-thrust belts are linked to foreland basin sedimentation associated with flexural subsidence induced by mountain building process (Ramos, 1999;Ramos and Kay, 2006;Xie and Heller, 2009;Ghiglione et al., 2010;Horton, 2018;VanderLeest et al., 2022;Aramendía et al., 2023). Sedimentation patterns and foreland basin depocenter distribution reveal changes in orogenesis, fold-thrust belt advance and uplift and exhumation rates, and/or periods of tectonic quiescence (DeCelles and Giles, 1996;Kraemer, 2003;Ghiglione and Ramos, 2005;Horton, 2018;Gallardo Jara et al., 2022;Goddard et al., 2023). ...
Article
Full-text available
We report uplift and shortening rates from a late Neogene–Pleistocene deformation stage of the frontal fold-thrust belt and adjacent wedge-top in the Principal Cordillera of the southern Central Andes (33-39° SL). A structural model is presented based on integration of surface field data and subsurface 2D seismic sections. Shortening, uplift, and sedimentation rates were calculated from different steps of kinematic modelling. Our structural interpretations and modelling are integrated with new detrital zircon U-Pb geochronology to define a previously overlooked Pleistocene period of orogenic shortening and syntectonic sedimentation in the Malargüe basin. This task was possible due to the dating of three samples yielding between ∼12 and 1 Ma obtained from a 900 m deep well located in the foreland. From stratigraphic correlations, our data records an active Plio-Pleistocene wedge-top depozone coeval with retreat of the volcanism, and the emplacement of retroarc basalts. Structural modelling, together with detrital zircon U-Pb provenance data register shortening producing a foredeep to wedge-top Plio-Pleistocene transition, adjusting and completing the knowledge of the frontal fold-thrust belt and foreland basin in the southern Central Andes. Supplementary material: https://doi.org/10.6084/m9.figshare.c.7033425
... These segments are typical of classic foreland basin subsidence pulses in the western part of the cross-sections, reflecting thrust-related events lasting ca. 2-4 Myr, similar to those proposed by Xie and Heller (2009). Following this period of increasing subsidence rates, there is commonly a third segment of waning subsidence, as also noted by those authors. ...
Article
Full-text available
Variability in subsidence rates within Upper Cretaceous strata of the Western Interior Basin offers crucial insights into the response of surface sedimentation styles to Sevier‐to‐Laramide tectonics and related deep mantle processes. The formation mechanisms of the Late Cretaceous Western Interior Basin in North America have long been a subject of debate. A re‐evaluation of the basin's subsidence history reveals rapid subsidence pulses lasting ca. 2 Myr within longer‐term (average 5.7 Myr) progradational or aggradational clastic wedges. The timing of these wedges, especially the widespread marine flooding resulting from subsidence, is constrained through the calibration of ammonite zonation with absolute dates. Sevier wedges exhibit a different architecture compared to the Laramide wedges. The former recorded initial rapid and widespread marine transgressions followed by long‐term coastal progradation, whereas the latter developed by initial erosional and progradational growth followed by aggradation and long‐term coastal transgression. The Sevier clastic wedges, initially accumulated within a N‐S elongated, long‐wavelength tectonic subsidence zone close to the thrust belt, gradually migrated cratonward. Starting in the early Campanian (ca. 82 Ma), the Laramide Orogeny developed along a NW‐SE trend and then migrated northeastward, roughly consistent with coeval long‐wavelength frontal basin subsidence. The spatio‐temporal variations in long‐wavelength tectonic subsidence indicate a shift in the dynamic subsidence's migration direction from eastward to northeastward, driven by changes in Farallon subduction direction and mode. Our work shows how repeated subsidence behavior in the Sevier‐to‐Laramide transition records evolving architectural responses and the trajectory of coeval dynamic topography.
Article
Full-text available
This is an open access article under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License, which permits use and distribution in any medium, provided the original work is properly cited, the use is non-commercial and no modifications or adaptations are made. Abstract In the late Permian and Triassic, the continental Mitu Group formed in exten-sional basins along the length of the Cordillera Oriental and Altiplano of present-day Peru. Given the presence of coeval arc systems only in northern Chile and southern Ecuador but not in Peru the tectonic setting of the Mitu basin has been interpreted variably as orthogonal continental rift, sinistral transtensional rift, aulacogen and back-arc basin. The Mitu Group comprises continental mass flow and alluvial fan, fluvial, aeolian and minor lacustrian facies and hosts thick piles of subalkaline and alkaline intermediate and felsic ignimbrites and mafic lavas. The age of the Mitu Group had originally been established as ranging from the Late Permian to the late Triassic on the basis of structural considerations and scarce biostratigraphic data. Recently, U-Pb zircon ages from ignimbrites and sedimentary rocks have been taken to constrain the Mitu Group to the Middle and Late Triassic. We performed a sedimentological, heavy mineral, and zircon geochronological and Lu-Hf isotope study of the Mitu Group in 14 sections mainly in southern and central Peru, and one section in northern Peru. Ten new U-Pb concordia ages on ignimbrites intercalated in the Mitu Group successions offer a new robust stratigraphic framework and constrain the stratigraphy of the Mitu Group between 260 and 205 Ma. In combination with maximum likelihood ages of deposition derived from detrital zircon, U-Pb geochronology places the deposi-tion of the Mitu Group between ca. 270 and 194 Ma (lower Guadalupian into the Sinemurian). Detrital zircon U-Pb age distributions and heavy mineral assemblages reflect a strongly recycled Precambrian Amazonian and Palaeozoic proto-Andean provenance. The Palaeozoic detrital age patterns are highly variable, and temporally and spatially random. A local provenance can generally not be identified. εHf(t) values in zircon obtained from ignimbrites and sedimentary rocks indicate variable degrees of crustal recycling. In the course of the Palaeozoic, εHf(t) values become on average progressively less negative, with a large proportion