Anda di halaman 1dari 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/282607060

Lithospheric-scale centrifuge models of pull-apart basins

Article  in  Tectonophysics · September 2015


DOI: 10.1016/j.tecto.2015.09.004

CITATIONS

10

2 authors:

Giacomo Corti Tim P Dooley


Italian National Research Council University of Texas at Austin
131 PUBLICATIONS   2,552 CITATIONS    57 PUBLICATIONS   1,386 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

MER Project View project

Volcanism distribution in extending continental crust (EARS) View project

All content following this page was uploaded by Giacomo Corti on 26 November 2015.

The user has requested enhancement of the downloaded file.


TECTO-126767; No of Pages 10
Tectonophysics xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Tectonophysics

journal homepage: www.elsevier.com/locate/tecto

Lithospheric-scale centrifuge models of pull-apart basins


Giacomo Corti a,⁎, Tim P. Dooley b
a
CNR, Consiglio Nazionale delle Ricerche, Istituto di Geoscienze e Georisorse, U.O. Firenze, Via G. La Pira, 4, 50121 Florence, Italy
b
Bureau of Economic Geology, The University of Texas at Austin, University Station, Box X, Austin, TX 78713, USA

a r t i c l e i n f o a b s t r a c t

Article history: We present here the results of the first lithospheric-scale centrifuge models of pull-apart basins. The experiments
Received 11 June 2015 simulate relative displacement of two lithospheric blocks along two offset master faults, with the presence of a
Received in revised form 20 August 2015 weak zone in the offset area localising deformation during strike–slip displacement. Reproducing the entire
Accepted 2 September 2015
lithosphere–asthenosphere system provides boundary conditions that are more realistic than the horizontal detach-
Available online xxxx
ment in traditional 1 g experiments and thus provide a better approximation of the dynamic evolution of natural
Keywords:
pull-apart basins. Model results show that local extension in the pull-apart basins is accommodated through devel-
Pull-apart basins opment of oblique–slip faulting at the basin margins and cross-basin faults obliquely cutting the rift depression. As
Analogue modelling observed in previous modelling studies, our centrifuge experiments suggest that the angle of offset between the
Centrifuge models master fault segments is one of the most important parameters controlling the architecture of pull-apart basins:
Strike–slip faulting the basins are lozenge shaped in the case of underlapping master faults, lazy-Z shaped in case of neutral offset
Lithospheric thinning and rhomboidal shaped for overlapping master faults. Model cross sections show significant along-strike variations
in basin morphology, with transition from narrow V- and U-shaped grabens to a more symmetric, boxlike geometry
passing from the basin terminations to the basin centre; a flip in the dominance of the sidewall faults from one end
of the basin to the other is observed in all models. These geometries are also typical of 1 g models and characterise
several pull-apart basins worldwide. Our models show that the complex faulting in the upper brittle layer corre-
sponds at depth to strong thinning of the ductile layer in the weak zone; a rise of the base of the lithosphere occurs
beneath the basin, and maximum lithospheric thinning roughly corresponds to the areas of maximum surface sub-
sidence (i.e., the basin depocentre).
© 2015 Elsevier B.V. All rights reserved.

1. Introduction 1987; Soula, 1984), in sand (e.g., Dooley and McClay, 1997; Dooley and
Schreurs, 2012; Dooley et al., 1999; Faugère et al., 1986; McClay and
Strike–slip tectonics is a fundamental process affecting many parts Dooley, 1995; Rahe et al., 1998; Richard et al., 1995) and in mechanically
of the Earth's lithosphere and resulting in prominent surface expres- layered systems (brittle–ductile; e.g., Basile and Brun, 1999; Dooley and
sions (Mann, 2007; Sylvester, 1988; Woodcock and Schubert, 1994). Schreurs, 2012; Mitra and Paul, 2011; Sims et al., 1999; Smit et al.,
Although strike–slip faults form linear and relatively continuous 2008a,b; Wu et al., 2009) using deformation rigs with the same basic de-
fault systems, they are typically segmented, resulting in localised sign. In most of these models the basal detachment to the pull-apart basin
extension/transtension or contraction/transpression as displacement was a horizontal shear zone underlain by a stretching rubber sheet be-
along the boundary fault system is transferred through a variety of dis- tween the rigid baseplates. In this paper we present a new series of cen-
continuities or steps (Cunningham and Mann, 2007; Mann, 2007; Mann trifuge models designed to investigate the development, evolution and
et al., 1983; Sylvester, 1988). Pull-apart basins form where bends or architecture of pull-apart basins on a lithospheric scale. Our new models
sidesteps (jogs) in the main strike–slip fault system (principal displace- adopt a set-up that is analogous to that used in the majority of 1 g models;
ment zone or PDZ) produce zones of localised extension where the however, by reproducing the entire lithosphere–asthenosphere system,
sense of step or bend in the fault system is the same as the sense of our new set-up provides boundary conditions that are more realistic
slip on the PDZ. More than 160 basins around the globe have been at- than the horizontal detachment typically used in traditional models.
tributed to strike–slip motion across segmented systems (Mann, 2007). Moreover, our new models complement previous results from classical
Physical models have greatly advanced our understanding of pull- 1 g models by providing insights into the patterns of lithospheric thin-
apart-basin genesis and evolution using models in clay (e.g., Atmaoui ning, in addition to the brittle deformation in the upper crust. To our
et al., 2006; Hempton and Neher, 1986; Mitra and Paul, 2011; Raynaud, knowledge, this is the first application of lithospheric-scale centrifuge
modelling and structural analysis to pull-apart basins and one of the
⁎ Corresponding author. Tel.: +39 0552757524; fax: +39 055290312. few experimental works in which strike–slip displacement is investigat-
E-mail address: giacomo.corti@igg.cnr.it (G. Corti). ed at a lithospheric scale.

http://dx.doi.org/10.1016/j.tecto.2015.09.004
0040-1951/© 2015 Elsevier B.V. All rights reserved.

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
2 G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx

2. Model set-up and experimental series to monitor the evolution of surface deformation. Model surface topogra-
phy was monitored by a laser scanner. Once the experiment was com-
The three models presented in this paper were performed in an arti- pleted, models were frozen before being thinly slabbed to study their
ficial gravity field of ~18 g using a large-capacity centrifuge and simulat- 3-D internal geometry.
ed strike–slip deformation of a brittle–ductile continental lithosphere
floating above a low-viscosity material representing the asthenosphere. 2.1. Rheological layering and experimental materials
The models were built inside a transparent rectangular Plexiglas® box
(with internal dimensions of 25 × 16 × 7 cm) and confined by two A vertical sequence of brittle and ductile materials was used to repro-
moveable side walls. Models consisted of two lithospheric blocks that duce the rheological multilayering that is characteristic of the continental
moved past each other to simulate symmetric, pure strike–slip defor- lithosphere (Fig. 1e). As in several previous centrifuge experiments
mation; the main strike–slip fault system (principal displacement (e.g., Agostini et al., 2009; Corti, 2012), the upper crust was simulated
zone) was characterised by an offset in the central part of the model by using a K-feldspar powder characterised by a linear increase in
in order to produce a zone of localised extension and give rise to a strength with depth to reproduce natural brittle behaviour. The lower
pull-apart basin, similar to other strike–slip modelling studies (Dooley crust was modelled by a ductile mixture of silicone (Dow Corning®
and Schreurs, 2012; Figs. 1 and 2). A weak zone was introduced be- 3179 Dilatant Compound, hereafter referred to as DC3179) and corun-
tween the two lithospheric blocks within the offset area, reproducing dum sand (100:10% in weight). The strong uppermost lithospheric man-
conditions analogous to the use of a basal rubber sheet between rigid tle was simulated with a mixture (100:20% in weight) of plasticine
plates or a localised ductile layer in 1 g models (Dooley and Schreurs, (Pongo Fantasia® modelling dough, distributed by FILA) and PDMS —
2012). Removal of rectangular blocks (spacers) at the sides of the mov- Polydimethylsiloxane (Dow Corning® SGM36). A mixture of DC3179, co-
ing walls allowed lateral motion of the two lithospheric blocks in re- rundum sand and oleic acid (100:50:1% in weight) modelled the lower
sponse to the centrifugal forces to fill the empty space (Fig. 1). lithospheric mantle. The weak zone in the offset area was reproduced
Sequential removal of spacers during successive runs in the centrifuge by introducing a weak mixture of DC3179, corundum sand and oleic
allowed control of the amount and rate of deformation. Top-view acid (100:20:5% in weight) beneath a 1 mm-thick layer of the standard
photos of the models were taken after the end of each stage in order lower crust model material (Fig. 1g). The crustal–mantle layers rested

a b c Counter
weight

Model
ω
CF
Counterweight
Model

10000
Log shear stress (Pa)

d f
1000

100
Upper lith mantle
Weak zone
10
16 cm

Lower crust
Lower lith mantle
1
0.001 0.01 0.1
Log strain rate (s-1)

g PLATE WEAK ZONE


Differential stress Differential stress
σ1-σ3 σ1-σ3
1000 2000 1000

15 cm 1220 1220

1220 1220
Depth, mm

10
1500 1500
e Upper crust 1500
Lower crust
Weak
zone

Upper lith mantle 20


Lower lith mantle 1500 1500

Fig. 1. Experimental setup. (a) Frontal view of the large-capacity centrifuge at the Tectonic Modelling Laboratory of the Institute of Geosciences and Earth Resources (National Research
Council of Italy). (b) Close-up of the internal rotor. (c) Loading conditions in the centrifuge (CFF, centrifuge force field). (d) Top-view photo of a model, illustrating the geometry of defor-
mation. (e) Model cross section illustrating the vertical rheological layering. (f) Stress–strain-rate relationships for the different viscous materials used to simulate the continental litho-
sphere (lith) plotted in a log–log graph. (g) Strength profiles of the model lithosphere in the different domains; numbers indicate the density of the different materials (in kg m−3).

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx 3

Pre-deformation (∼2.5 × 10−5 m s−1) scaled to natural values of ∼5–50 mm yr−1, depend-
ing on the assumed viscosity of ductile layers (e.g., 1023–1024 Pa s for the
strong upper lithospheric mantle). This range of values well compares
with velocities of natural strike–slip systems (e.g., ~25 mm/yr for the

Weak
zone
North Anatolian Fault, Cakir et al., 2014; 3–5 mm/yr for the Dead Sea
transform, Sadeh et al., 2012; 1–35 mm/yr for the fault segments com-
posing the San Andreas Fault, Tong et al., 2014).

2.3. Types of experiments


Post-deformation
The main parameter investigated in the current model series is the
angle of offset between the master fault segments created by control-
ling the amount of underlap (U; Fig. 2) or overlap (O; Fig. 2). This
angle has a fundamental control of the resultant pull-apart-basin geom-
etry (Dooley and Schreurs, 2012). In particular, as shown in Fig. 2, we
have investigated the three conditions of underlapping (Model 1; offset
angle A = 45°), neutral (Model 2; offset angle A = 90°) and overlapping
(Model 3; offset angle A = 135°) master faults, identical to those inves-
Underlapping master faults (Model 1) tigated by Dooley and McClay (1997) and also summarised in Dooley
and Schreurs (2012). All other parameters, including the separation
(S) between the master-fault segments and the initial width of the off-
set area/weak zone (parameters that control the length:width ratio of
1 cm Separation 1.5 cm the basins) as well as the vertical rheological layering of both the nor-
Angle (A) 45° mal lithosphere and the weak zone, were kept constant (Fig. 2).

2.4. Experimental monitoring and analysis


Neutral master faults (Model 2)
In all experiments, deformation was monitored through top-view
photos taken after each run in the centrifuge and laser scanning at sub-
1 cm millimeter resolution to reconstruct the topography of the model sur-
Separation 1.5 cm
face. At the end of each experiment, the models were soaked in water,
Angle (A) 90° frozen and cut in a set of evenly spaced, subparallel cross sections to an-
alyse the internal deformation.

Overlapping master faults (Model 3) 2.5. Simplifications and limitations of the modelling approach

As in all modelling studies, our experimental set-ups require signif-


1 cm icant simplifications of natural systems. In particular, lithospheric-scale
Separation 1.5 cm
processes involve temperature-dependent aspects that were not con-
Angle (A) 135° sidered in the current models. This means that the thermo-mechanical
response of the lithosphere to pull-apart development was simplified
to a purely mechanical response. Additionally the rheology of natural
Fig. 2. Geometrical characteristics of the model setups. (a) Overhead view illustrating pre- materials and the lithospheric strength profiles are greatly simplified
and post-deformation of the master faults and offset zone. (b) Separation and angular re-
in laboratory studies. Similarly, sedimentation was simplified with re-
lationships of the stepovers for the three models presented in this paper. The angles of the
stepovers are identical in geometry to those presented in Dooley and McClay (1997) and
spect to nature: the developing structural topography was simply filled
Dooley and Schreurs (2012). to the top of the basin shoulders, and it was performed intermittently at
imposed time intervals and not continuously as likely occurs in natural
pull-apart basins. For practical reasons (the size of the internal rotor) we
on a low-viscosity mixture made of DC3179, corundum sand and oleic have deliberately assumed a thin lower lithospheric mantle in the ex-
acid (100:80:15% in weight). The rheological characteristics (flow periments, resulting in a model lithospheres that are thinner than nor-
curves) and densities of these materials are illustrated in Fig. 1f; their mal natural continental lithospheres (e.g., Ranalli, 1995). We believe
use resulted in an increase in density, with depth and a variable strength that this setup does not greatly influence the response of our model lith-
reproducing a typical Christmas-tree strength profile of the continental osphere to strike–slip deformation and pull-apart development. This is
lithosphere (Fig. 1g; e.g., Ranalli, 1995). due to fact that the lower lithospheric mantle is a weak layer that has
Syntectonic sedimentation was simulated by filling the pull-apart a negligible contribution to the total resistance of the lithosphere,
depression (up to the top of the rift shoulders) with sieved K-feldspar which is instead mostly controlled by the upper lithospheric mantle
powder before continuing deformation within the centrifuge. and upper crust. However, this setup limits a more quantitative compar-
ison between model and nature with respect to thinning factors.
2.2. Scaling
3. Experimental results
The models were built with a geometric scale ratio of ∼3.3 × 10−7,
such that 1 cm in the experiments corresponded to ∼30 km in nature. Figs. 3–7 and Movies 1–3 illustrate the results from the three exper-
This allowed modelling ∼ 45 km of lateral displacement of ∼60-km- iments. Common to all the models is an evolution of deformation
thick continental lithospheres. Dynamic–kinematic similarity of gravita- characterised by the early development of Riedel shears above the mas-
tional, viscous and frictional stresses acting in the system (Ramberg, ter faults and oblique–slip faults accommodating localised extension in
1981) ensured that the velocity of lateral displacement in the models the offset area that bordered the developing basin (basin sidewall faults,

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
4 G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx

Model 1 - Underlapping master faults


Illumination 5 mm

Illumination 1 cm

Illumination 7 mm

Illumination 1 cm

Illumination 9 mm

Illumination 1 cm

Illumination 11 mm

Illumination 1 cm

Illumination 13 mm

Illumination 1 cm

Illumination 15 mm
SSFs
CBFs
BSFs

BSFs

SSFs

Illumination 1 cm

Fig. 3. Evolution of model with underlapping master fault (Model 1), shown as sequential top-view photos with differing illumination angles (left and central panels) and line drawings of
structures (right panels). BSFs: basin sidewall faults; CBFs: cross-basin faults; SSFs: strike–slip faults.

BSFs; Figs. 3–5). Progressive displacement led to linking of the Riedel ratio of ~ 2.5:1. Maximum subsidence occurred in the centre of the
shears to form thoroughgoing strike–slip faults above the master faults pull-apart in association with the CBFs; this corresponded at depth to
and propagation of the BSFs towards the opposite master fault. Oblique– strong thinning of the ductile layers and maximum lithospheric thin-
slip faults (cross-basin faults, CBFs) formed later, cutting the basin floor ning of 1/β ~ 0.75 (where 1/β is the ratio between the final and the ini-
and propagating to link the offset master faults with increasing tial thickness of the model crust; Figs. 6 and 7).
displacement. In Model 2 (neutral master faults, Fig. 2; Fig. 4; Movie 2), BSFs devel-
In the case of underlapping master faults (Model 1; Figs. 2 and 3; oped almost orthogonal to the master fault trend, subparallel to the
Movie 1), the BSFs acquired an en-échelon arrangement whose trend trend of the weak zone boundaries. The CBFs started forming at ~8 mm
roughly followed the trend of the weak zone boundaries at depth. of lateral displacement and were well established at ~12 mm of strike–
CBFs were well established after ~ 9 mm of lateral displacement in slip deformation. At ~14 mm of displacement the basin had acquired a
these experiments, and at ~ 15 mm of displacement the basin had ac- roughly lazy-Z shape, with a length:width ratio of ~1.7:1. Maximum sub-
quired a well-defined elongate, lozenge shape, with a length:width sidence and lithospheric thinning occurred in the centre of the pull-apart

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx 5

Model 2 - Neutral master faults


Illumination 4 mm

Illumination 1 cm

Illumination 6 mm

Illumination 1 cm

Illumination 8 mm

Illumination 1 cm

Illumination 10 mm

Illumination 1 cm

Illumination 12 mm

Illumination 1 cm

Illumination SSFs
14 mm
CBFs
BSFs

BSFs

SSFs

Illumination 1 cm

Fig. 4. Evolution of model with neutral master fault (Model 2), shown as sequential top-view photos with differing illumination angles (left and central panels) and line drawings of struc-
tures (right panels). BSFs: basin sidewall faults; CBFs: cross-basin faults; SSFs: strike–slip faults.

in correspondence to the CBFs (Figs. 6 and 7). The lithosphere was basin in correspondence with the CBFs (Figs. 6 and 7). A maximum lith-
thinned to about half its initial thickness in this model (1/β ~ 0.5 in Fig. 7). ospheric thinning of 1/β ~ 0.75 was recorded here (Fig. 7).
Model 3 investigated an overlapping master fault arrangement Cross sections through the models show that deformation is accom-
(Figs. 2 and 5; Movie 3). The arrangement of the BSFs around the modated by strong thinning of the ductile weak zone accompanied by
basin is slightly more complex than in Models 1 and 2: they formed a complex faulting of the upper crust (Fig. 6). This latter layer is cut by nu-
set of subparallel faults nearly orthogonal to the direction of strike– merous steep or subvertical faults, sometimes with a curved geometry,
slip displacement with no direct relation with the trend of the weak which may display a reverse component of motion (oversteepened),
zone boundaries. CBFs were well established at ~9 mm of strike–slip de- thus indicating a complex kinematic history. Basin morphology is
formation. At ~ 14 mm of displacement the pull-apart basin had ac- characterised by significant along-strike variations. The basin centres
quired a rhomboidal shape, with a length:width ratio of ~ 1.7:1. typically display a roughly symmetric, boxlike geometry (Fig. 6). Con-
Maximum subsidence was observed in the box-shaped centre of the versely, the areas close to the main master faults are marked by narrow

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
6 G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx

Model 3 - Overlapping master faults


Illumination 5 mm

Illumination 1 cm

Illumination 7 mm

Illumination 1 cm

Illumination 9 mm

Illumination 1 cm

Illumination 11 mm

Illumination 1 cm

Illumination 13 mm

Illumination 1 cm

Illumination SSFs 14 mm
CBFs
BSFs

BSFs
SSFs

Illumination 1 cm

Fig. 5. Evolution of model with overlapping master fault (Model 3), shown as sequential top-view photos with differing illumination angles (left and central panels) and line drawings of
structures (right panels). BSFs: basin sidewall faults; CBFs: cross-basin faults; SSFs: strike–slip faults.

V- and U-shaped grabens and negative flower structures, with a typical presence of a weak zone in the offset area that strongly localises defor-
asymmetric architecture characterised by a faulted margin on one side mation during strike–slip displacement. This set-up is analogous to that
and a more monoclonal flexure on the other that dips towards the used in the majority of 1 g models, where pull-apart development is ob-
faulted margin (Fig. 6). A dip-polarity change from one end of the tained through the stretching of a rubber sheet, with or without a weak
basin to the other is typical in all models (Fig. 6). ductile layer, between laterally moving basal rigid plates (Dooley and
Schreurs, 2012). However, by reproducing the entire lithosphere–
4. Discussion asthenosphere system, our new set-up provides boundary conditions
that are more realistic than the horizontal detachment typically used
Our new centrifuge models reproduced the development of a pull- in traditional 1 g models and thus better approximates the dynamic
apart basin at a lithospheric scale by simulating relative displacement evolution of natural pull-aparts. The close geometric matches of pull-
of two lithospheric blocks along two offset master faults, with the apart basin geometries between these centrifuge models and those

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx 7

Model 1 - Underlapping master faults Model 2 - Neutral master faults Model 3 - Overlapping master faults
A B C D E F A B C D E F A B C D E F

A’ B’ C’ D’ E’ F’ A’ B’ C’ D’ E’ F’ A’ B’ C’ D’ E’ F’

UC UC
UC
LC LC LC
ULM ULM ULM
LLM LLM LLM

Ast Ast Ast


A A’ A A’ A A’

B B’ B B’ B B’

C C’ C C’ C C’

D D’ D D’ D D’

E E’ E E’ E E’

F F’ F F’ F F’
1cm 1cm 1cm

Fig. 6. Cross sections from the underlapping (left), neutral (centre) and overlapping (right) master-fault models. Of note is the along-strike variation in basin morphology and the dip-
polarity change from one end of the basin to the other. For example, in case of underlapping master faults, the upper crust dips mainly to the right in Section C–C′ then flips to dominantly
left dipping in Section E–E′. The same occurs in the other models. Note in sections B-B′ and D-D′ of Model 2 that the thin layer of lower crust above the black weak material composing the
weakness zone has been in places completely removed by ductile thinning.

produced by traditional 1 g modelling using rigid baseplates validates comparison of the different centrifuge experiments highlight a signifi-
our new technique. Our new models also complement previous studies cant change in basin shape depending on the degree of underlap or
by offering insights into the patterns of lithospheric thinning, in addi- overlap of the master fault segments (Fig. 6), supporting the claim
tion to the brittle deformation in the upper crust. Although a few previ- that this is one of the most important parameters controlling the archi-
ous studies investigated specific aspects of strike–slip deformation in tecture of pull-apart basins (see Dooley and Schreurs, 2012, for further
the centrifuge (e.g., pluton emplacement in shear zones, Dietl et al., details). In particular, the basins are lozenge shaped in the case of
2006; reactivation of pre-existing structures, Koyi et al., 2008), to our underlapping master faults (Model 1), lazy-Z shaped in the case of neu-
knowledge this is the first application of centrifuge modelling to a sys- tral offset (Model 2) and rhomboidal shaped for overlapping master
tematic analysis of pull-apart basins. faults (Model 3; Fig. 6), as observed in previous normal-gravity, crustal
In a fashion similar to published 1 g models, our centrifuge models models (e.g., Dooley and McClay, 1997). Analysis of model cross sec-
show that local extension in the pull-apart basins is accommodated by tions shows significant along-strike variations in basin morphology,
the development of oblique–slip faulting (BSFs) at the basin margins, with transition from narrow V- and U-shaped grabens to a more sym-
whereas a portion of the lateral displacement is accommodated along metric, boxlike geometry passing from the basin terminations to the
CBFs within the depression. In agreement with previous researchers, basin centre; a flip in the dominance of the sidewall faults from one

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
8 G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx

Model 1 - Underlapping master faults Model 2 - Neutral master faults Model 3 - Overlapping master faults
15 mm 14 mm 14 mm

1 cm 1 cm 1 cm

15 mm Subsidence 14 mm Subsidence 14 mm Subsidence


0.5 0.5
0.5
1
mm 1

mm

mm
1 1.5
1.5
2
2
1.5 2.5

1 cm 1 cm 1 cm

15 mm Thinning 14 mm Thinning 14 mm Thinning


1 1 1

0.95 0.90
0.90
0.90 0.80

1/β

1/β
1/β

0.70 0.80
0.85

0.80 0.60 0.70

1 cm 1 cm 1 cm

Fig. 7. Surface deformation and lithospheric thinning in the three models, illustrated as fault pattern (top panels), amount of subsidence (central panels) and patterns of lithospheric thin-
ning (bottom panels), indicated with the 1/β factor (where 1/β is the ratio between the final and the initial thickness of the model crust).

end of the basin to the other is also observed in all the models. These ge- to the plate displacement vector. This is particularly evident in Model
ometries and along-strike structural variations are also typical of 1 g 2 (neutral master faults), which can be considered analogous to orthog-
models and have been observed in several pull-apart basins worldwide onal extension. In these conditions, rifting models show BSFs orthogo-
(e.g., Dooley and McClay, 1997; Dooley and Schreurs, 2012). nal to the direction of plate motion and parallel to the weak zone
Our new models show that complex oblique–slip faulting in the upper boundaries, whereas the CBFs instead develop highly oblique to both
brittle layer corresponds at depth to strong thinning of the ductile layer in BSFs and the direction of plate displacement. This also results in a differ-
the weak zone; a rise of the base of the lithosphere occurs beneath the ent pattern of thinning within the pull-apart basin. Maximum thinning
basin, and maximum lithospheric thinning roughly corresponds to the in rift models typically occurs beneath internal faults with an extension-
areas of maximum surface subsidence (i.e., the basin depocentre). This orthogonal axis or, in oblique rifting, with a segmented, en-échelon ar-
pattern of crustal/lithospheric thinning is also broadly similar to the typ- rangement. In pull-apart models the axis of maximum thinning follows
ical response of the continental lithosphere to extension as observed in the trend of the CBFs, being thus oblique to the plate displacement vec-
rifting models (e.g., Nestola et al., 2015), although with some significant tor. This obliquity of maximum thinning and its relationship to the CBFs
differences. Lithospheric-scale models of orthogonal and oblique rifting is well illustrated by all three models.
(Corti, 2012) show that extension is accommodated by a combination The patterns of lithospheric thinning in our models are also similar to
of marginal and axial faulting with the development of the main basin- those observed in lithospheric-scale numerical models of pull-apart basin
boundary faults and intrabasin faults that can be correlated to the BSFs development (Petrunin and Sobolev, 2006, 2008; Petrunin et al., 2012).
and CBFs of our new strike–slip models, respectively. As in our physical models lithospheric thinning in these numerical models
However, the geometry of BSFs is much more complex than that of is compensated at depth by the flow of the asthenosphere beneath the
boundary faults in purely extensional basins. Rift-margin fault systems thinned area; maximum basin subsidence is controlled by the competi-
typically form a system of major faults dipping on average 60–70° to- tion between this flow and brittle/ductile thinning of the overlying litho-
wards the rift depression (e.g., Corti, 2012; McClay et al., 2002), whereas sphere. Overall, these model results indicate that pull-apart basins are
faults bounding the pull-apart basins are normally very steep or efficient in promoting localised thinning of the crust and lithosphere, as
subvertical and may display a reverse component of motion, indicating well as pronounced basin subsidence (e.g., Xie and Heller, 2009).
complex kinematics dictated by the boundary conditions of the basin These processes and the upwelling of the asthenosphere beneath the
(Fig. 6). Moreover, while internal faults are characterised by almost thinned area may result in high heat flows and magmatism/volcanism
pure extension in rifting models, CBFs display an oblique–slip displace- within the pull-apart basins, as observed in several different natural ex-
ment in the pull-apart experiments, accommodating a portion of the amples (e.g., Aydin and Nur, 1982; Aydin et al., 1990; Bellier and Sébrier,
lateral displacement imposed by the master faults. These faults also dif- 1994; Mann et al., 1983). One spectacular example of a pull-apart sys-
fer in trend and architecture: whereas internal faults respond directly to tem with significantly enhanced heat flow and associated volcanicity
the plate displacement vector (i.e., direction of extension) – forming or- is the Coso Geothermal System in Southern California, although crustal
thogonal to this direction and, in oblique rifting, with an en-échelon ar- thinning and mantle upwelling are enhanced in the Coso area by the
rangement of fault segments – CBFs develop as a continuous fault transtensional nature of deformation (e.g., Dooley and Schreurs, 2012;
system connecting the opposite master faulting and thus form oblique Monastero et al., 2005). The efficiency of pull-apart basins in localising

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx 9

rapid, strong thinning of the lithosphere and the thermomechanical ef- experiments. Int. J. Earth Sci. 95, 225–238. http://dx.doi.org/10.1007/s00531-
005-0030-1.
fects related to magmatic processes have been suggested as promoting Aydin, A., Nur, A., 1982. Evolution of pull-apart basins and their scale independence. Tec-
rapid breakup of highly oblique continental rift systems (e.g., the Gulf of tonics 1, 91–105.
California, Umhoefer, 2011). Aydin, A., Schultz, R.A., Campagna, D., 1990. Fault normal dilation in pull-apart basins; Im-
plications for the relationship between strike–slip faults and volcanic activity.
Annales Tectonicae 4 (2), 45–52.
5. Conclusions Basile, C., Brun, J.P., 1999. Transtensional faulting patterns ranging from pull-apart basins to
transform continental margins: an experimental investigation. J. Struct. Geol. 21, 23–37.
Bellier, O., Sébrier, M., 1994. Relationship between tectonism and volcanism along the
New centrifuge models were designed to analyse the architecture Great Sumatran fault zones deduced by SPOT image analyses. Tectonophysics 233,
and evolution of pull-apart basins at a lithospheric scale. The new 215–231.
setup is analogous to that used in the majority of 1 g models; however, Cakir, Z., Ergintav, S., Akoğlu, A.M., Çakmak, R., Tatar, O., Meghraoui, M., 2014. InSAR ve-
locity field across the North Anatolian Fault (eastern Turkey): implications for the
by reproducing the development of pull-apart basins at the scale of the
loading and release of interseismic strain accumulation. J. Geophys. Res. Solid Earth
entire lithosphere–asthenosphere system, these models possess more 119, 7934–7943. http://dx.doi.org/10.1002/2014JB011360.
realistic boundary conditions than the horizontal detachment typically Corti, G., 2012. Evolution and characteristics of continental rifting: analogue modeling-
used in traditional normal-gravity models. Model results suggest the inspired view and comparison with examples from the East African Rift System.
Tectonophysics 522–523, 1–33. http://dx.doi.org/10.1016/j.tecto.2011.06.010.
following main conclusions: Cunningham, W., Mann, P., 2007. Tectonics of strike–slip restraining and releasing bends.
Geol. Soc. Lond., Spec. Publ. 290 (1), 1.
(1) Our results show that, in a fashion similar to normal-gravity Dietl, C., Koyi, H., de Wall, H., Gössmann, M., 2006. Centrifuge modelling of plutons intrud-
models, deformation in the pull-apart basins is accommodated ing shear zones: application to the Fürstenstein Intrusive Complex (Bavarian Forest,
through the development oblique–slip faulting at the basin mar- Germany). Geodin. Acta 19 (3–4), 165–184.
Dooley, T.P., McClay, K., 1997. Analog modeling of pull-apart basins. AAPG Bull. 81 (11),
gins and cross-basin faulting within the depression. The plan- 1804–1826.
view basin geometry (lozenge, lazy-Z or rhomboidal) is con- Dooley, T.P., Schreurs, G., 2012. Analogue modelling of intraplate strike–slip tectonics: a
trolled by the degree of underlap or overlap of the master fault review and new experimental results. Tectonophysics 574–575, 1–71.
Dooley, T.P., McClay, K., Bonora, M., 1999. 4D evolution of segmented strike–slip fault
segments (e.g., underlapping, overlapping or neutral master systems: applications to NW Europe. In: Fleet, A.J., Boldy, S.A.R. (Eds.), Petroleum Ge-
faults), as observed in previous normal-gravity experiments. ology of Northwest Europe: Proceedings of the 5th Conference. Geological Society,
Cross-sections morphology of the basins show significant London, pp. 215–225.
Faugère, E., Brun, J., Van Den Driessche, J., 1986. Asymmetric basins in pure extension and
along-strike variations, with a transition from narrow V- and U-
in wrenching: experimental models. 1. Bulletin du Centre de Recherches Elf Explora-
shaped grabens to a more symmetric, boxlike geometry passing tion Production, pp. 13–21.
from the basin terminations to the basin centre; a flip in the Hempton, M., Neher, K., 1986. Experimental fracture, strain and subsidence patterns over
en echelon strike–slip faults: implications for the structural evolution of pull-apart
dominance of the sidewall faults from one end of the basin to
basins. J. Struct. Geol. 8, 597–605.
the other is observed. Many of the faults bounding the basin Koyi, H., Ghasemi, A., Hessami, K., Dietl, C., 2008. The mechanical relationship between
are steep or subvertical and may display a reverse component strike–slip faults and salt diapirs in the Zagros fold-thrust belt. J. Geol. Soc. 165 (6),
of motion, indicating complex kinematics. 1031–1044. http://dx.doi.org/10.1144/0016-76492007-142.
Mann, P., 2007. Global catalogue, classification and tectonic origins of restraining and re-
(2) These new centrifuge models indicate that complex faulting in leasing bends on active and ancient strike–slip fault systems. In: Cunningham, W.D.,
the upper brittle layer corresponds at depth to strong thinning Mann, P. (Eds.), Tectonics of Strike–slip Restraining and Releasing Bends. Geological
of the ductile layer in the weak zone; a rise of the base of the lith- Society, London, Special Publications 290, pp. 13–142.
Mann, P., Hempton, M.R., Bradley, D.C., Burke, K., 1983. Development of pull-apart basins.
osphere occurs beneath the basin, and maximum lithospheric J. Geol. 91, 529–554.
thinning roughly corresponds to the areas of maximum surface McClay, K., Dooley, T., 1995. Analogue models of pull-apart basins. Geology 23, 711–714.
subsidence (i.e., the basin depocentre). This pattern of crustal/ McClay, K.R., Dooley, T., Whitehouse, P., Mills, M., 2002. 4-D evolution of rift systems: In-
sights from scaled physical models. AAPG Bull. 86 (6), 935–960.
lithospheric thinning is broadly similar to the typical response Mitra, S., Paul, D., 2011. Structural geometry and evolution of releasing and restraining
of the continental lithosphere to extension as observed in rifting bends: insights from laser-scanned experimental models. AAPG Bull. 95 (7),
models, and is also similar to what has been observed in numer- 1147–1180. http://dx.doi.org/10.1306/09271010060.
Monastero, F., Katzenstein, A., Miller, J., Unruh, J., Adams, M., Richards-Dinger, K., 2005.
ical models of lithospheric-scale of pull-apart basin develop-
The Coso geothermal field: a nascent metamorphic core complex. Geol. Soc. Am.
ment. Results indicate that pull-apart basins are efficient in Bull. 117, 1534–1553. http://dx.doi.org/10.1130/B25600.1.
promoting localised thinning and asthenosphere upwelling. Nestola, Y., Storti, F., Cavozzi, C., 2015. Strain rate-dependent lithosphere rifting and neck-
ing architectures in analog experiments. J. Geophys. Res. Solid Earth 120. http://dx.
doi.org/10.1002/2014JB011623.
Petrunin, A.G., Sobolev, S.V., 2006. What controls thickness of sediments and lithospheric
Supplementary data to this article can be found online at http://dx. deformation at a pull-apart basin? Geology 34 (5), 389–392. http://dx.doi.org/10.
doi.org/10.1016/j.tecto.2015.09.004. 1130/G22158.1.
Petrunin, A.G., Sobolev, S.V., 2008. Three-dimensional numerical models of the evolution
of pull-apart basins. Phys. Earth Planet. Inter. 171 (1–4), 387–399. http://dx.doi.org/
Acknowledgements 10.1016/j.pepi.2008.08.017.
Petrunin, A.G., Meneses Rioseco, E., Sobolev, S.V., Weber, M., 2012. Thermomechanical
We thank Giorgio Ranalli for discussions and two anonymous re- model reconciles contradictory geophysical observations at the Dead Sea Basin.
Geochem. Geophys. Geosyst. 13, Q04011. http://dx.doi.org/10.1029/2011GC003929.
viewers for comments that helped us improve the clarity of this manu-
Rahe, B., Ferill, D.A., Morri, A.P., 1998. Physical analogue modeling of pull-apart basin evo-
script. The paper was edited by Stephanie Jones. Part of this work has lution. Tectonophysics 285, 21–40.
been developed in the frame of the project “Analisi della dinamica di Ramberg, H., 1981. Gravity, deformation and the Earth's crust. Academic Press, London
apertura del Golfo della California” (Progetti di Mobilità dei Ricercatori (452 pp).
Ranalli, G., 1995. Rheology of the earth. 2nd ed. Chapman & Hall, London (413 pp).
nel P.E. Italia — Messico). TD is funded by the Applied Geodynamics Lab- Raynaud, S., 1987. Les premiers stades de la déformation dans une zone de relais entre
oratory at the Bureau of Economic Geology, the University of Texas at décrochements: exemples naturels et expérimentaux. Bull. Soc. géol. Fr.8 III (3),
Austin. 583–590.
Richard, P., Naylor, M.A., Koopman, A., 1995. Experimental models of strike–slip tectonics.
Pet. Geosci. 1, 71–80.
References Sadeh, M., Hamiel, Y., Ziv, A., Bock, Y., Fang, P., Wdowinski, S., 2012. Crustal deformation
along the Dead Sea Transform and the Carmel Fault inferred from 12 years of GPS mea-
Agostini, A., Corti, G., Zeoli, A., Mulugeta, G., 2009. Evolution, pattern and partitioning of surements. J. Geophys. Res. 117, B08410. http://dx.doi.org/10.1029/2012JB009241.
deformation during oblique continental rifting: inferences from lithospheric-scale Sims, D., Ferrill, D.A., Stamatakos, J.A., 1999. Role of a ductile décollement in the develop-
centrifuge models. Geochem. Geophys. Geosyst. 10, Q11015. http://dx.doi.org/10. ment of pull-apart basins: experimental results and natural examples. J. Struct. Geol.
1029/2009GC002676 (GCubed). 21, 533–554.
Atmaoui, N., Kukowski, N., Stöckhert, B., König, D., 2006. Initiation and development Smit, J., Brun, J.-P., Fort, X., Cloetingh, S., Ben-Avraham, Z., 2008a. Salt tectonics in pull-
of pull-apart basins with Riedel shear mechanism: insights from scaled clay apart basins with application to the Dead Sea Basin. Tectonophysics 449, 1–16.

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
10 G. Corti, T.P. Dooley / Tectonophysics xxx (2015) xxx–xxx

Smit, J., Brun, J.-P., Cloetingh, S., Ben-Avraham, Z., 2008b. Pull-apart basin formation and Umhoefer, P.J., 2011. Why did the Southern Gulf of California rupture so rapidly? Oblique
development in narrow transform zones with application to the Dead Sea Basin. Tec- divergence across hot, weak lithosphere along a tectonically active margin. GSA
tonics 27. http://dx.doi.org/10.1029/2007TC002119 (TC6018). Today 21 (11), 4–10. http://dx.doi.org/10.1130/G133A.1.
Soula, J.-C., 1984. Genèse de bassins sedimentaires en regime de cisaillement transcurrent: Woodcock, N.H., Schubert, V., 1994. Continental strike–slip tectonics. In: Hancock, P.L.
modèles expérimentaux et exemples géologiques. Bull. Soc. Belge Géol. 93 (1–2), (Ed.), Continental Deformation. Pergamon Press, Oxford, pp. 251–263.
83–104. Wu, J., McClay, K., Whitehouse, P., Dooley, T., 2009. 4D analogue modelling of transtensional
Sylvester, A.G., 1988. Strike–slip faults. Geol. Soc. Am. Bull. 84, 1293–1309. pull-apart basins. Mar. Pet. Geol. 26 (8), 1608–1623.
Tong, X., Smith-Konter, B., Sandwell, D.T., 2014. Is there a discrepancy between geological Xie, X., Heller, P.L., 2009. Plate tectonics and basin subsidence history. GSA Bull. 121,
and geodetic slip rates along the San Andreas Fault System?, J. Geophys. Res. Solid 55–64. http://dx.doi.org/10.1130/B26398.1.
Earth 119, 2518–2538. http://dx.doi.org/10.1002/2013JB010765.

Please cite this article as: Corti, G., Dooley, T.P., Lithospheric-scale centrifuge models of pull-apart basins, Tectonophysics (2015), http://dx.doi.org/
10.1016/j.tecto.2015.09.004
View publication stats

Anda mungkin juga menyukai