Sei sulla pagina 1di 20

Geological Society of Australia Special Publication 22, 6583

CHAPTER 6Three-dimensional finite-element modelling


of the tectonic stress field in continental Australia
S. ZHAO1, 2 AND R. D. MLLER1*
1 School of Geosciences, University of Sydney, NSW 2006, Australia.
2 Japan Marine Science and Technology Center, Yokosuka 237-0061, Japan.
Corresponding author: dietmar@es.usyd.edu.au
Traditionally, intraplate stress orientations have been modelled using an isotropic elastic plate. For
the Australian Plate this method has been applied successfully to model the first-order pattern of
stress orientations. However, the distribution of intraplate earthquakes and the juxtaposition of
strong, cold with hotter, younger lithosphere in many areas suggest that the spatial variation in
mechanical strength of the plate may result in substantial regional anomalies in stress orientations
and magnitudes. We explore this idea with a three-dimensional finite-element model to investigate
the regional response of the Australian continent to tectonic forces. The model covers the area of
40 to 10 (S) and 111 to 155 (E) with a spatial resolution of 90 x 90 x 50 km. The relative magnitudes of the ridge push and boundary forces, which act on the Australian continent, are estimated
through an inversion analysis of in situ stress data. The differences between modelled and
observed stress orientations are minimised in a least-squares sense. Major tectonic blocks and the
differences in their elastic strength are included in the model, and the initial estimates of the
Youngs moduli for the tectonic blocks are adapted from a published coherence analysis of gravity and topographic data. The values of the Youngs moduli are adjusted in the inversion analysis
to best fit the stress orientations observed on the Australian continent. The inversion analysis of rheological parameters is most efficient for estimating the Youngs moduli for the Northern Lachlan
Fold Belt, the New England Fold Belt, and the Southern Lachlan Fold Belt. The adjusted values for
the flexural rigidity are 0.040 x 1025 Nm for the Northern Lachlan Fold Belt, 0.037 x 1025 Nm for the
New England Fold Belt, and 0.040 x 1025 Nm for the Southern Lachlan Fold Belt, which correspond
to an effective elastic thickness of about 30 km. Based on the optimised body and boundary
forces acting on the plate, a map of maximum principal-stress distribution is constructed so that
variations of the relative magnitude of tectonic stresses can be assessed. We find a good match
between predicted zones of stress concentration and the distribution of major belts of seismicity in
Australia. The results show that while the overall pattern of stress orientations in the Australian continent is controlled by the forces which drive the Indo-Australian Plate, the maximum horizontal
stress orientations and the pattern of the stress concentration manifested by seismicity are modulated by local/regional geological structures.
KEY WORDS: Australia, crustal plates, in situ stress, numerical modelling, seismicity.

INTRODUCTION
Studies of intraplate stresses show that most intraplate
regions are characterised by compressive stress regimes
and crustal seismic activity (Zoback et al. 1989) and that
the inferred maximum horizontal principal-stress orientations are roughly parallel to the directions of ridge-push
forces. On the basis of the analysis of earthquake focal
mechanisms, in situ stress measurements and surface
deformation, Denham et al. (1979) have shown that the
Australian continent is in a state of substantial horizontal
compression. However, the maximum horizontal compression measured in situ and inferred from seismic source
mechanism solutions is approximately normal to the eastern passive margin (Denham et al. 1979), which is not in
agreement with the absolute velocity trajectories of the
Australian Plate (Richardson 1992; Zoback 1992).
Numerical modelling of the tectonic forces applied to
the Indo-Australian Plate (Figure 1) has been performed
with the finite-element method. Cloetingh and Wortel

(1986) investigated the tectonic stress field in the IndoAustralian Plate with a two-dimensional finite-element
model. Five types of tectonic forces were included in their
analysis: slab pull, ridge push, resistant force, trench suction force and drag force. The combination of the forces in
their model resulted in a concentration of the compressive
stresses of the order of 300500 MPa in some parts of the
plate (e.g. the Ninetyeast Ridge). Using a similar 2D finiteelement model, Coblentz et al. (1998) reinvestigated the
tectonic forces acting on the Indo-Australian Plate. The
stress indicators from the World Stress Map Project
(Zoback 1992) were used to constrain their numerical models. They found that: (i) the ridge-push force is likely the primary force in controlling the first-order stress pattern in the
Indo-Australian Plate; (ii) if imposing resistance along the
Himalaya, Papua New Guinea and New Zealand collisional boundaries to balance the ridge-push force, then
many of the first-order stress patterns of the observed stress
field can be explained without including either subduction
or basal-drag forces; and (iii) the observed maximum hori-

66

S. Zhao and R. D. Mller


Figure 1 Map showing the location of the Australian continent and the geometry of the Indo-Australian Plate
boundaries (inset). MOR, Mid-Ocean Ridge; J, Java Trench;
B, Banda Arc; PNG, Papua New Guinea; SM, Solomon
Trench. Arrows indicate the directions of the major tectonic
forces (not to scale). FL, the boundary force from the west;
FR, the boundary force from the east; FC, the collision and
subduction-related forces from the north; FP, the ridgepush force from the south. Inset (bold lines are plate
boundaries): EU, Eurasian Plate; PA, Pacific Plate; S,
Sumatra Trench; NH, New Hebrides; TK, TongaKermadec
Trench; NZ, New Zealand.

zontal-stress orientations and stress-regime information in


the Indo-Australian Plate can also be explained with the
models predicting low tectonic-stress magnitudes (e.g. tens
of megapascals, averaged over the thickness of the lithosphere). This implies that the large stress magnitude (hundreds of megapascals) inferred by Cloetingh and Wortel
(1986) for some parts of the Indo-Australian Plate is not
required to explain the observed stress orientations and
regime information. Coblentz et al. (1995) also discussed
the origins of the intraplate stress field in continental
Australia and suggested that stress focusing effects along
the heterogeneous convergent boundaries (implemented by
fixing the boundaries) are necessary to produce the significant compression within the continent. In addition, Zhang
et al. (1996) constructed a 2D finite-element model for part
of the eastern Australian passive margin, although lateral
stress changes could not be fully investigated because of
the nature of the 2D elastic model, and it could not be used
to interpret the stress orientations observed in the
Australian continent.
The main tectonic forces acting on the Australian continent that have been identified in previous studies are shown
in Figure 1. The ridge-push force from the mid-oceanic ridge
was inferred to be the dominant force of driving the IndoAustralian Plate northwards (Coblentz et al. 1998). In the
east, the Australian continent may be affected by the subduction of the Pacific Plate near New Zealand. The force
transferred from the subduction zone, which is over 3000
km away from the Australian continent, was considered to
have only a secondary effect (Coblentz et al. 1998). In the
north, the boundary between the Indo-Australian, Eurasian
and Pacific Plates is very complex. At the Java Trench, the
Australian Plate is subducting beneath the Eurasian Plate,
while the Eurasian Plate is subducting under the Australian

Plate at the Banda Arc. At the Solomon Trench, the


Australian Plate is subducting beneath the Pacific Plate,
while the Pacific Plate is subducting beneath the Australian
Plate at Papua New Guinea.
Although the results from the previous studies have
greatly improved our understanding of the relative magnitudes of the tectonic forces and their role in controlling the
tectonic stress field in the Australian continent, there are
several important factors that could not be addressed by
previously applied methodologies,
(1) The Australian continent is assumed to be homogeneously rigid in most of the previous models. Recent studies
on lithospheric structures from a seismic tomographic analysis of the Australian continent reveal that the lithosphere is
not homogeneous in elastic strength, but heterogeneous
(Simons et al. 1999; Simons & Van der Hilst 2002). Many
surface geological structures (e.g. cratons and basins) in the
Australian continent have extensions in the upper mantle in
terms of seismic velocity anomalies (Kennett 1997). The formation of the elastically/seismically inhomogeneous structures is closely associated with the stress-evolution process
in the lithosphere. Since some of the geological structures in
the Australian continent are quite large, up to a width of 800
km for some cratons, their effect on spatial changes in stress
orientations and magnitude is probably significant. These
elastically inhomogeneous structures were not considered in
any previous stress modelling for the Australian continent.
(2) Two-dimensional finite-element modelling was
employed in all of the previous models. In the 2D models,
vertical stresses have been ignored in the plane stress
approximation used in these studies.
(3) Crustal seismicity has occurred throughout the
Australian continent, and the origin of these intraplate
earthquakes are still puzzling (Denham 1988; Denham &

3D Modelling of Australian stress field

67

Figure 2 Map showing the main geological


blocks in continental Australia. YB, Yilgarn
Block; MUB, Musgrave Block; GB, Gawler
Block; ARB, Arunta Block; MB, Mt. Isa Block;
Hodg. F. B., Hodgkinson Fold Belt; N. LA. F.
B., Northern Lachlan Fold Belt; New Eng. F.
B., New England Fold Belt; S. LA. F. B.,
Southern Lachlan Fold Belt; Adel. F. B.,
Adelaide Fold Belt. B5, B11, B9 and B13 are
isolated blocks (see Table 1).

Windsor 1991). Because of the limitations inherent in the


previous 2D finite-element models, the spatial distribution
of the tectonic stress in the Australian continent, as well as
the origins of the intraplate seismicity, have not been investigated. In order to assess the spatial pattern of the tectonic
stresses (such as the distribution and relative magnitude) in
the Australian continent, three-dimensional modelling is
essential.
(4) A trial-and-error method was used in the previous
modelling studies and the calculated stress orientations
were visually compared with the observed stress orientations, which is not technically efficient. The studies were
mainly qualitative or semiquantitative so that a formal fit
between the observed and modelled stress orientations
could not be achieved. While recent improvement has been
made through some refined strategies in forward stress
modelling (Reynolds et al. 2003), an inverse approach for
estimating model parameters from observed stress orientations would improve our ability to find the best-fit model.
The present study is an extension of previous modelling
efforts of the tectonic stress field in continental Australia
(Coblentz et al. 1995). Our work differs from the previous
studies mainly in the following aspects: (i) it is focused on
continental Australia, and the area covered by the model is
about 45 x 31 (in longitude and latitude) with a spatial resolution of about 90 x 90 x 50 km; (ii) heterogeneous features of the Australian continent are included in our
analysis, which are associated with the differences in the
elastic strength for different geological domains, such as
major tectonic provinces and fold belts; and (iii) since a
wide range of boundary conditions can be configured to
match the observed intraplate stress field, the non-uniqueness of the problem is investigated. In this study, an inversion method is used to estimate the relative magnitudes
and directions of the tectonic forces associated with the
Australian continent from stress orientation data.
Previous studies demonstrated that the gravitational
potential energy difference across the boundary between
continental and oceanic crust may significantly affect the

regional stress field at a plate scale (Coblentz et al. 1994;


Sandiford et al. 1995). In this study, the contribution of
topography and gravity potential energy differences to a
local/regional stress field is not simulated, partly because a
quantitative simulation requires a detailed crustal/lithospheric (density) structure model, which is not presently
available. Ignoring the effect of the gravity potential energy
differences in the Australian continent will affect our
results, especially at the continental margin, and will be
discussed later.

TECTONIC BLOCKS AND FLEXURAL RIGIDITY


OF THE AUSTRALIAN CONTINENT
The Australian continent, which is geologically and tectonically complex, can be divided into several crustal blocks.
Each block has its own distinctive tectonic style and represents a significant stage in the evolution of the Australian
continent (Plumb 1979a, b). Figure 2 shows the main tectonic units in the Australian continent. The yellow areas are
the cratons, which are geologically stable; the dark green
areas are fold belts; and the blank area in the continent is
composed of basins and smaller blocks, which will be
treated indiscriminately as continental crust/lithosphere,
and assumed to have a mechanical strength less than that
of the cratons and larger than that of the fold belts for our
modelling.
The flexural rigidity of the tectonic blocks in the
Australian continent has been investigated by Zuber et al.
(1989) and Simons et al. (2000) on the basis of the analysis of Bouguer gravity and topography data. The flexural
rigidity values estimated by Zuber et al. (1989) for the
New England Fold Belt and Southern Lachlan Fold Belt
(~ 1022 N m) are about three orders of magnitude lower
than those (~ 1025 N m) of cratons. A revised estimate of
the effective elastic thickness for central Australia is about
a factor of two less than that of Zuber et al. (1989) (Simons
et al. 2000). This suggests that there are large uncertainties

68

S. Zhao and R. D. Mller

Table 1 Flexural rigidity and Youngs Modulus for the major tectonic blocks in the Australian continent.
(x 1025 Nm)

no.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17

Basins
Adelaide Fold Belt
Yilgarn Block
Pilbara Block
Block 5 (B5)
Arunta Block
Musgrave Block
Gawler Block
Block 9 (B9)
Mt. Isa Block
Block 11 (B11)
Southern Lachlan Fold Belt
Block 13
Hodgkinson Fold Belt &
Northern Lachlan Fold Belt
New England Fold Belt
Oceanic crust
Continental shelf

Estimatedc

Adjustedd

(x 1025 Nm)

2.0
2.0
2.1
0.61
0.61
0.69
2.1
2.1
0.69
0.0011
2.0
0.0016

0.95
0.95
1.0
0.29
0.29
0.33
1.0
1.0
0.33
0.00052
0.95
0.00076

3.0
0.113
5.7
5.7
6.0
1.74
1.74
1.98
6.0
6.0
1.98
0.0031
5.7
0.0046

6.656
1.970

0.113

0.113

0.040

2.330
0.690

0.040

0.040

0.0036

0.0017

0.01
5.7
0.21

0.106

0.037

a From Zuber et al. 1989.


b Obtained after dividing the flexural rigidity values by the maximum flexural rigidity value (2.1 x 1025 Nm).
c Assign a Youngs modulus value of 6 x 1010 Pa for Blocks 5, 9 and 10, which have the maximum flexural rigidity and then estimate
the Youngs modulus values of other blocks based on the scaling factor.
d Estimated from the inversion analysis.

in the estimated effective elastic thickness and the flexural


rigidity of the tectonic blocks in Australia, depending on the
method used. In this study, we are mainly concerned with
the relative values of rigidity among the tectonic blocks;
therefore uncertainties in the absolute values of the flexural
rigidity will not affect our results in terms of stress orientations. We use the relative values of the rigidity of the tectonic blocks estimated by Zuber et al. (1989) to scale the
rheological parameters associated with the elastic strength
and as initial input for our model. In addition, the values of
rheological parameters will be adjusted in the inversion
analysis of the observed stress orientations. The age and
estimated flexural rigidity for the main tectonic blocks used
here are outlined below.

Yilgarn and Pilbara Blocks


The Yilgarn and Pilbara Blocks in Western Australia formed
about 35003100 Ma. This region is geologically stable
(Plumb 1979a). The largest Moho depth here is estimated
to be ~50 km (Clitheroe et al. 2000). The flexural rigidity for
this area was estimated to be 2.0 x 1025 N m (Zuber et al.
1989).

Arunta and Musgrave Blocks


The Arunta and Musgrave Blocks are located in central
Australia, which consists of heavily faulted Proterozoic
blocks and basins. Moho offsets with amplitude variations
more than 20 km have been inferred from gravity modelling
(Lambeck 1983a) and the analysis of seismic travel time
anomalies (Lambeck 1983b; Lambeck & Penney 1984).
The flexural rigidity for this region is estimated to be 6.1 x
1024 N m (Zuber et al. 1989).

Gawler Block
The Gawler Block in South Australia formed during the
Late Archaean to Palaeoproterozoic (Plumb 1979a). The
flexural rigidity for the block is estimated to be 6.9 x 1024 N
m (Zuber et al. 1989) and the Moho depth is about 40 km
(Clitheroe et al. 2000).

Mt Isa Block and Northern Craton


To the north of the Arunta Block, the entire region (including B5 and B9 in Figure 2), containing the Mt. Isa Block, is
simply called the Northern Craton (Zuber et al. 1989;
Plumb 1979b) and consists of Palaeoproterozoic blocks
bounded by Mesoproterozoic orogenic belts. The flexural
rigidity for this region is estimated to be 2.1 x 1025 N m
(Zuber et al. 1989) and the Moho depth is estimated to be
~40 km (Clitheroe et al. 2000).

Eastern Highlands
The Eastern Highlands consist of several Palaeozoic fold
belts along the Australian coast: the Hodgkinson Fold Belt
and Northern Lachlan Fold Belt in the northeast, and the
New England Fold Belt and Southern Lachlan Fold Belt in
the southeast. Southeastern Australia is characterised by
the highest seismicity on the continent (Lambeck et al.
1984), anomalously high heat flow (Cull 1991), and high
mantle conductivity (Lilley et al. 1981). The Moho depth in
the Eastern Highlands varies from about 32 km in the north
to 52 km in the south (Clitheroe et al. 2000). The flexural
rigidity is estimated to be 1.6 x 1023 N m for the Hodgkinson
Fold Belt and the Northern Lachlan Fold Belt, 3.6 x 1022 N
m for the New England Fold Belt, and 4.4 x 1022 N m for the
Southern Lachlan Fold Belt (Zuber et al. 1989).

3D Modelling of Australian stress field

69

Figure 3 Stress orientations (solid lines


for the maximum horizontal compressive
stress directions) at 386 points in the
Australian continent, compiled from the
Australian Stress Map Project (Hillis &
Reynolds 2000; Mueller et al. 2000).
Also shown are the major tectonic blocks
in the Australian continent (see also
Figure 2). NW, northwest Australia;
BONA, Northern Bonaparte Basin; WA,
western Australia; CA, central Australia;
SA, south Australia; EA, eastern
Australia.

Figure 4 Stress orientations at the 163


sample points which are used in the
inversion analysis. The solid lines
denote the observed maximum horizontal compressive stress orientations,
which is a subset of the stress orientations shown in Figure 3. Also shown are
the major tectonic blocks in the
Australian continent (see also Figure 2).

STRESS ORIENTATION DATA ON THE


AUSTRALIAN CONTINENT
The stress data used in this study are from the World Stress
Map website (Mueller et al. 2000). The dataset is compiled
from borehole breakouts, hydraulic fracturing measurements, earthquake focal mechanisms, drilling-induced fracturing, and fault orientations through the Australian Stress
Map Project (Hillis et al. 1998, 1999; Hillis & Reynolds
2000). A total of 386 orientations (the maximum horizontal
compressive-stress directions) are plotted in Figure 3. The
stress orientations on the Australian continent do not form
a uniform direction, although there is some noticeable uni-

formity in several regions (abbreviations in round brackets


refer to Figure 3): (i) in northwest Australia there are two
dominant stress orientations: 140150N, 80-90N (NW1)
and 3050N (NW2 and BONA); (ii) in west Australia
(WA), there is a trend for the 130140N stress orientation
but eastwest orientations also appear and the scatter in
the stress directions is apparent; (iii) in central Australia,
the northsouth stress orientations are dominant (CA1),
and eastwest orientations are also present at some points
(CA2); (iv) in northeast Australia, north-northeast stress
orientations dominate (EA1); (v) in southeast Australia
(EA2), there are two dominant stress orientations,
130140N in the southernmost part and northsouth

70

S. Zhao and R. D. Mller

stress orientations are displayed at some points in the


northern part of the region, although the stress orientations
in almost every direction were observed in this area; and
(vi) in south Australia, the dominant stress orientation is
about 130140N near the coast (SA2) and further inland
near the Adelaide fold belt it is quite scattered, from
130140N to eastwest (SA1).
In this study, we use the stress orientations with quality
level AB exclusively, because of their relatively high reliability (Zoback & Zoback 1991; Zoback 1992). There are a
total of 163 stress orientations with a quality level AB
(Figure 4). Comparing Figures 3 and 4, we can see that the
main pattern of the stress orientations is still maintained in
the dataset containing only level AB stress indicators.

NUMERICAL MODELLING
Stress changes and related phenomena, such as fault activity or earthquakes, are associated with the action of various
tectonic forces, which are usually not directly observable.
Therefore, the relationship between stress observations and
tectonic forces is mainly investigated through modelling
studies. The tectonic forces acting on the Australian continent, model assumptions and strategy used in this study
are discussed below.

Major tectonic forces acting on the Australian


continent
RIDGE-PUSH FORCE FROM THE MID-OCEAN RIDGE

Mechanically, the ridge-push force results in a slide force


along the ridge-spreading direction (Lister 1975). The horizontal component of the slide force can be viewed as a
component of the gravitational force, which is subparallel
to the topographic slope. Therefore, the slide force, when
averaged over the plate thickness H, can be expressed as
(Lister 1975):
fR = g r a k T /(vH)

(1)

where g is the gravitational acceleration, r is the density, a


is the thermal expansion coefficient, k is thermal diffusivity,
and v is the average half-spreading rate of the ridge. T is the
temperature at which the mantle material becomes sufficiently weak such that the lithosphere is decoupled from
the asthenosphere. T is usually assumed to be about
9001000C for olivine rheology (Goetze & Evans 1979).
The Australian continent is located over 3000 km north
of the mid-ocean ridge (Figure 1), therefore the horizontal
component of the slide force along the ridge-spreading
direction (southnorth) can be taken as constant throughout the entire continent. The average half-spreading rate for
the mid-ocean ridge is about 30 mm/y (Mller et al. 1997).
Using typical rock property values a = 4 x 105C1, r =
3300 kg m3, k = 8 x 107 m2 s1 and T = 950C, we have
fRH = 1.03 MPa. For the interior of the Australian Plate,
which is about 3000 km from the ridge, this force integrates
to 3.03 x 1012 N/m. We use a body force (slide force) to represent the ridge push in 3D, and this value is adjusted in
the inversion analysis. As defined by a 950C isotherm, the

thickness of the oceanic lithosphere increases from nearly


zero at the crest to about 30 km at the age of ca 10 Ma. The
slide force is estimated to be FR = 51 N/m3 for oceanic
lithosphere with an average thickness of 20 km near the
ridge. This body force is considered to act uniformly
throughout the Australian continent and drives the continent northwards.
BOUNDARY FORCES

In the northern part of the Australian continent, as a result


of the interaction among the Indo-Australian, Eurasian and
Pacific Plates, the geometry of boundaries between the
plates is very complex. At the Java Trench, the IndoAustralian Plate is subducting beneath the Eurasian Plate.
At the Banda Arc and Papua New Guinea, the Eurasian
and Pacific Plates are subducting beneath the Australian
Plate, and at the Solomon Trench, the Australian Plate is
subducting beneath the Pacific Plate (Figure 1). Since the
properties of the forces acting at the northern boundary of
the Australian Plate are not clear, we assume that their
combined effect is to exert a resistant force (relative to the
ridge-push force) along the plate boundary near northern
Australia (FC in Figure 1). Along the eastern boundary of
the Australian Plate, there might be a possible boundary
force (FR in Figure 1) transferred from the (New Zealand)
subduction zone where the Pacific Plate is subducting
beneath the Australian Plate. In west Australia, it is not
clear whether there is an equivalent boundary force transmitted from the west (see later). In previous models
(Coblentz et al. 1995), the magnitude of the boundary
forces for both the northern and eastern boundaries of the
Indo-Australian Plate was taken to be ~6 x 1012 Nm1.
DRAG FORCE

Basal drag is the shear traction that the asthenosphere


applies to the base of the lithosphere. The direction of the
basal drag is usually assumed to be opposite to the direction of the absolute plate motion, but the actual direction is
difficult to assess. The magnitude of the basal drag is likely
small and of the order of 102101 MPa (Richardson
1992). Coblentz et al. (1995) estimated a torque value for
the basal drag that is about 50% of the torque of the ridgepush force for the Indo-Australian Plate. In our study, we
choose to ignore the effect of the basal drag on stress modelling because: (i) its magnitude is probably small and its
other properties (such as its distribution over the base of
the lithosphere) are unknown; and (ii) the direction of the
drag force, if resisting the plate motion of the Australian
Plate, is oriented northsouth, which is the same (opposite
in direction) as that of the ridge-push force. Therefore, it is
not possible to distinguish its presence or absence from the
stress orientation data used in this study.

Finite-element model
We used a two-layer elastic model to simulate the response
of the Australian continent to various tectonic forces
(Figure 5). The use of the elastic rheology in the stress
analysis for the Australian continent is an oversimplification, as recognised by Cloetingh and Wortel (1986) and

3D Modelling of Australian stress field

71

is indirectly included in our analysis by considering the differences in their elastic strength, as determined from the
coherence of Bouguer gravity anomalies and topography
(Zuber et al. 1989).

Boundary conditions

Figure 5 The finite-element grid used in this study.

Coblentz et al. (1995; 1998). However, an elastic rheology


is a justifiable approach for investigating the first-order tectonic stresses in continental Australia, especially when we
have insufficient data to constrain a more complicated rheology, such as a viscoelastic model, which would be more
suitable for investigating stress-relaxation processes in the
crust/lithosphere (Lambeck 1983a; Stephenson & Lambeck
1985).
The finite-element model (Figure 5) contains a total of
4185 nodes and 2640 brick elements, and its spatial resolution is about 90 x 90 x 50 km (the dimension of each element). We use the Lambert azimuth projection to transform
geographical coordinates into Cartesian coordinates, in
which the finite-element method computation is carried
out. Since the errors in delineating the boundaries of tectonic blocks is up to ~100 km, only tectonic blocks wider
than 100 km are included in the analysis, and small structures, such as faults, are ignored in our continental-scale
model. Figure 6 shows the distribution of the finite-element
nodes and the major tectonic blocks in the Australian continent. The approximation of the geometry of the tectonic
blocks is achieved by representing the irregular boundaries
with rectangles (bricks in 3D) (Figure 6). The differences in
the effective elastic thickness (and the flexural rigidity) of
the blocks reflect their differences in elastic strength, used
here to constrain the rheological parameters of our numerical model.
The geological/rheological provinces with different elastic strength are incorporated into the upper layer of a thickness of 50 km (Figure 5), which is close to the maximum
Moho depth in continental Australia (Clitheroe et al. 2000).
A bottom layer of 50 km is introduced as a reference layer
with a Youngs modulus of 14 x 1010 Pa (Turcotte &
Schubert 1982) to reflect the fact that the elastic strength of
the mantle (bottom layer) is higher than the crust (upper
layer). A change of the thickness of each layer affects the
magnitude of the stresses, but it does not affect the relative
magnitudes and pattern (including the orientations) of the
calculated stresses. Our main objective is to estimate the
relative magnitudes and the pattern of tectonic stresses,
rather than estimating the absolute magnitude of the tectonic stresses in the Australian continent. The effect of variations in the equivalent elastic thickness of the lithosphere

We assume that the nodes at the bottom (depth = 100 km)


of the tectonic block are fixed (Figure 5), which implies that
the motion/deformation of the top layer (lithosphere), if
any, is relative to the fixed bottom layer. Other boundaries
are free. In one of our models, the western boundary of the
block, corresponding to the west Australian coast, is
assumed fixed, and its effect on modelling results is examined in an inversion analysis. In all other models, boundary forces are imposed at the western, eastern and northern
boundaries. Their typical values from previous studies
(Coblentz et al. 1995) are used and adjusted in the inversion analysis.

Rheological parameters
The major geological structures considered in this study
correspond to those investigated by Zuber et al. (1989),
except for the Adelaide Fold Belt, which was not included
in their study. We assume that the rheological contrast in
the Australian continent can be represented by 17 groups
of material in terms of their differences in elastic strength
(Table 1). We adopt a constant value 0.25 for the Poisson
ratio throughout the investigated area and use different values of Youngs modulus to represent the difference in the
elastic strength for different geological structures. The estimates for the rigidity of the geological structures in the
Australian continent are taken from the flexural analysis
based on the gravity and topographic data by Zuber et al.
(1989) (Table 1). The initial values of the Youngs modulus
for the tectonic blocks are estimated on the basis of the relative magnitude of the flexural rigidity (Table 1).
The scaled values of the flexural rigidity in Table 1 are
obtained by dividing by the maximum value of the originally estimated flexural rigidity (2.1 x 1025 N m). The
Youngs modulus values of the blocks are estimated on the
basis of their scaling factors. The Youngs modulus is taken
to be 3.0 x 1010 Pa for the basins (the blank area in Figure
2) and 5.7 x 1010 Pa for the oceanic crust/lithosphere (the
area outside the continent). In Table 1, a Youngs modulus
value of 0.21 x 1010 Pa is used for the continental shelf,
which is assumed to have a lower strength than the continental crust. Inclusion of the continental shelf in the modelling analysis did not improve the quantitative analysis of
the stress orientations, implying that the stress orientations
in the continent calculated by the numerical model are not
sensitive to the rheological parameters of the continental
shelf. A possible reason is that we used a single value to
describe the strength of the continental shelf throughout
the continent. Actually, the mechanical properties of the
continental slope could be different from region to region
(e.g. from the west to south Australia), but the stress orientation data in the continent could not resolve such differences.
The flexural rigidity values of fold belts (Table 1) estimated by Zuber et al. (1989) are about three orders of mag-

72

S. Zhao and R. D. Mller

nitude lower than those of cratons. The errors in estimating


the flexural rigidity of the fold belts from the gravity and
topographic data (Zuber et al. 1989) may be large compared with those of cratons because of their relatively small
dimensions (Simons et al. 2000). These errors in the estimated flexural rigidity are transferred to the Youngs modulus values used in this study. Therefore, the values of the
Youngs modulus for the fold belts are adjusted (re-estimated) in an inversion analysis of the stress orientation
data. The values for the adjusted Youngs modulus, and the
values for refined rigidity in Table 1 are the resulting flexural rigidities for some of the tectonic blocks obtained from
the inversion analysis of the stress orientation data and will
be discussed later.

Inversion analysis
Numerical modelling with the finite-element method can be
classified into two types: forward and inverse analyses.
Forward modelling calculates the stresses (orientations) in
the continent from known tectonic forces and rheological
parameters. Inverse modelling estimates some unknown
forces and rheological parameters from observations (e.g.
orientations) and some known tectonic forces and rheological parameters. Suppose that the stress orientations (Y) in
the Australian continent are a function of the tectonic
forces (Fx) and rheological parameters (R):
Y = f(Fx, R)

(2)

where f is an operator which expresses the relationship


between Y and (Fx, R). The description of the forward problem is that we want to estimate Y, given Fx and R.
Unfortunately, we only have very limited knowledge of
the tectonic forces (Fx) and rheological properties of the
continent. In previous studies, the rheological parameter R
was taken as a constant, and only the parameter Fx was
adjusted by visually comparing calculated stress pattern
with observed stress orientations. If there are relatively
ample stress observations and we wish to use the observations as quantitative constraints, the problem has to be
considered in a reverse waythe inverse problemwhich
can be described as wanting to estimate Fx and/or R, given
Y (to estimate tectonic forces and rheological parameters
from observed stress orientations):
(Fx, R) = f1 (Y)

(3)

where f1 is an inverse operator which expresses the inverse


relationship between Y and (Fx, R).
Since the stress dataset on the Australian continent is
not (mathematically) complete (we always have only limited observations), the solutions of the associated inverse
problems are not unique. We need to introduce a priori
information into the modelling analysis. The known geometry of geological elements and basic information on the
directions and/or magnitude of tectonic forces from the previous studies are taken as a priori information in our model.
The inverse problem for estimating tectonic forces from
stress orientations, can then be expressed as:

|| Y Y (Fx ) || = min

(4)

a1 Fx a2

(5)

where Y is the observational stress orientations (vector), Y


is the modelled stress orientations (vector), Fx is the parameter vector of the tectonic forces to be estimated; a1 and
a2 are coefficients that are the lower and upper limits of the
parameter (Fx).
The rheological parameters in model (4) are assumed to
be known. In addition, the constraints for the tectonic
forces are easily obtained based on information from plate
tectonics or previous forward models. For example, for the
ridge-push force associated with the Australian continent,
its direction is approximately northward (along the Y-axis
in our model), so we have a1 >0 or Fx >0. We can also infer
a rough value (a2) for the upper limit of the ridge-push
force.
Likewise, the inverse problem of estimating the rheological parameters from stress orientations can be expressed
as:

|| Y Y ( R ) || = min
b1 R b2

( 6)
( 7)

where Y is the observational vector of the stress orientations, Y ( R ) is the modelled stress orientation (vector),
which now depends on the rheological parameters (R), and
b1 and b2 are the lower and upper limits of the rheological
parameters.
The constraints for the rheological parameters are relatively easy to obtain. For example, we know that R >0, and
from the results of laboratory experiments we can determine approximate upper limits for the rheological parameters. Methods for solving geophysical inverse problems
have been presented by Menke (1984) and Tarantola
(1987), and will not be discussed here.
The inversion analysis is conducted in a stepwise fashion. We first take the Youngs modulus scaled from the flexural rigidity analysis as initial values, and then investigate
the response of tectonic blocks to the assigned tectonic
forces. Different combinations of the tectonic forces are
examined and adjusted to fit the stress orientations in the
Australian continent. We then take the estimated tectonic
forces as known parameters and refine the estimates of the
Youngs modulus to fit the stress orientations. These procedures are repeated until the squared residuals between the
observed and modelled stress orientations reach a minimum, in a least-squares sense. The final estimates of the
tectonic forces and the Youngs moduli can be viewed as
global least-squares estimates.

RESULTS
Ridge-push force
An initial value of 6.0 N/m3 was assigned to the magnitude
of the slide force in the Australian continent. The refined
value (FP in Table 2; Figure 1) after the inverse analysis is
55.8 N/m3 (0N), which is close to the value 51 N/m3, calculated independently from Equation (1) based on parameters for the Southeast Indian Ridge. To examine the
possible eastwest component of the slide force, the tec-

3D Modelling of Australian stress field

73

Figure 6 Map showing the distribution of the finite-element nodes in plan view, and the approximation for the geometry of the major
geological structures in the Australian continent with the brick elements (50 km thick each). The areas coloured red are cratons; fold
belts are green. The yellow curve denotes the boundary of the continental shelf. The numbers represent the estimates of the effective
elastic thickness in kilometres (Zuber et al. 1989). Arrows denote the force vectors (not to scale) considered in the modelling analysis.

tonic block is divided into four sections (Figure 6), and for
each of the sections, an eastward ridge-push force component is added: the resultant changes in the stress orientations are not significant (Table 2).

Boundary forces
In the eastern boundary (Figure 1), there may be a force
(FR) resulting from the Pacific Plate, which is subducting
beneath the Australian Plate. Similar to the value
(10121013 Pa/m) used in previous studies (Coblentz et al.
1995), we assume an initial value of 2.0 x 1012 Pa/m for the
magnitude of the boundary force at the eastern boundary,
and the search for the optimal value gives an estimate of >
5.99 5.2 x 1012 Pa/m: that is, the upper limit of the magnitude is uncertain. The error in this estimate is significant
and we interpret this as implying that the dataset is not sensitive to the magnitude of boundary force from the east. For
the boundary force (FC) associated with the northern
boundary, we used the same initial value as that used for
the eastern boundary. The estimated value is 11.8 3.2 x
1012 Pa/m. The absolute values of the forces are difficult to

evaluate from the stress orientation data used in this study.


The estimated values of the forces only have relative importance, and they are model dependent. In other words, only
the relative magnitudes of the tectonic forces can be constrained by the stress-orientation data.
In order to explore the possible range of boundary
forces at the western boundary of the Australian continent,
we tested three models: (i) an eastwest boundary force of
1.0 x 1012 Pa/m; (ii) a free boundary; and (iii) a fixed
boundary. We found that none of the three models produces a significant improvement in the residuals of the
stress orientations. Since the western boundary of the
Australian continent is in the interior of the Indo-Australian
Plate (Figure 1), the tectonic deformation in western
Australia associated with plate-boundary forces is much
smaller than that in northern Australia, which supports the
use of a fixed boundary. In addition, the fixed western
boundary in the third model serves to resist any plate-tectonic forces from deflecting the western boundary, which
seems reasonable. Therefore, we adopt the fixed western
boundary assumption. A further test of the boundary force
vectors acting on four sub-segments of the western bound-

74

S. Zhao and R. D. Mller


Figure 7 Observed (black lines) and estimated (red lines, Model A1) maximum
horizontal compressive stress orientations. NW, northwest Australia; BONA,
Northern Bonaparte Basin; WA, western
Australia; CA, central Australia; SA, south
Australia; EA, eastern Australia.

Table 2 Estimated magnitude of the ridge push and boundary forces.


Ridge push (slide force) (FP)
Boundary force (FR)
Boundary force (FC)

Northward
Eastward
Northward
Eastward
Northward
Eastward

6.0 N/m3

2.00 x 1012 Pa/m


2.00 x 1012 Pa/m

ary of the block (Figure 6) did not produce any substantial


improvements on the misfit, which might suggest that the
stress observations (WA in Figure 7) discussed here cannot
be fully accounted for by the continental-scale model.
Possible mechanisms for the local stress field in west
Australia will be discussed later.
To test the effect of the possible oblique-type forces at
the northern boundary of the Australian continent, the
northern segment of the tectonic block is divided into four
segments (Figure 6). For each segment, in addition to the
northsouth component of the boundary force (along the Yaxis), an eastwest component (along the X-axis) is also
assumed to be unknown in the inversion analysis. A combination of the two components (X and Y) constitutes a
force vector. However, the inversion fails to give a significant estimate for the eastwest component of the collision
force for all segments. This suggests that detailed characteristics of the tectonic-force vectors cannot be resolved
from the available stress-orientation data. This may indicate that the variations of the boundary forces along the
northern Australian plate boundary have little effect on the
stress orientations observed within the Australian continent. This conclusion supports the interpretation that a
large amount of the energy associated with subduction
zones may be dissipated by resistance to subduction, and
therefore a surface plate may not experience substantial
slab pull (Richardson 1992; Hillis et al. 1997). However,
this does not mean that the effect of the forces acting at the

55.8 6.1 N/m3


not significant
not significant
FC >5.99 5.2 x 1012 Pa/m
11.8 3.2 x 1012 Pa/m
not significant

plate boundary on the intraplate tectonic stress field can be


dismissed. Forces originating at the plate boundary, other
than the boundary forces considered here, could have
some significant effects on the intraplate stress field, such
as the body forces transmitted from the plate boundaries,
but the properties of such forces are not very clear.
The maximum horizontal stress orientations estimated
from the inversion analysis (Model A1) are shown in Figure
7. The estimated stress orientations (red lines) are fairly consistent with the observations (black lines) in the western part
of northwest Australia (NW1), Northern Bonaparte Basin
(BONA), central part of Australia (CA1) and northeast
Australia (EA1). While the inclusion of the geological structures in our analysis has provided a reasonable fit to some of
the observed stress orientations, there are deviations
between the observed and modelled stress orientations, e.g.
in the southern part of west Australia (WA), the eastern part
of northwest Australia (NW2), central Australia (CA2), eastern Australia (EA2) and south Australia (SA2).

Refined rheological parameters


The stress orientations for the points near or inside a geological structure are associated with the material contrast
between the structure and its surrounding area. Therefore,
the rheological parameters for some of the tectonic blocks
are adjusted to obtain their optimal values by inversion
analysis of the stress-orientation data.

3D Modelling of Australian stress field

75

Figure 8 Observed (black lines) and


modelled (red lines, Model A2) maximum horizontal principal stress orientations (after adjustment of the
rheological parameters of the tectonic
blocks). NW, northwest Australia;
BONA, Northern Bonaparte Basin; WA,
western Australia; CA, central Australia;
SA, south Australia; EA, eastern
Australia.

Figure 9 Observed (black lines) and


modelled (red lines, Model A3) maximum horizontal principal stress orientations (after inclusion of the effect of
local stress fields). The shaded bars
denote the orientations of the introduced local stress fields. NW, northwest
Australia; BONA, Northern Bonaparte
Basin; WA, western Australia; CA, central Australia; SA, south Australia; EA,
eastern Australia.

INLAND BASINS

FOLD BELTS

The Youngs modulus value is taken to be 3.0 x 1010 Pa for


the basins in the Australian continent (Figure 2). A change
in the value of the Youngs modulus of the basins has little
effect on the overall residuals between the observed and
modelled stress orientations, suggesting that stress orientations are not sensitive to the change of this parameter. A
possible reason could be that we used a single value to represent the elastic strength for all of the basins, so that the
difference among the basins, which (if any) affects the local
stress distributions, could not be distinguished in the continental-scale model.

The inversion analysis indicates that the stress orientations


are sensitive to the Youngs modulus values of fold belts.
Comparing the Youngs modulus values (~1010 Nm) of the
cratons, the re-estimated values for the fold belts are about
one to two orders of magnitude lower (Table 1). The re-estimated value of the Youngs modulus is 0.113 x 1010 Pa for
the Northern Lachlan Fold Belt, 0.106 x 1010 Pa for the
New England Fold Belt and 0.113 x 1010 Pa for the
Southern Lachlan Fold Belt (Table 1). The adjusted flexural
rigidity value is 0.040 x 1025 Nm for the Northern Lachlan
Fold Belt, 0.037 x 1025 Nm for the New England Fold Belt,

76

S. Zhao and R. D. Mller

and 0.040 x 1025 Nm for the Southern Lachlan Fold Belt


(Table 1). The estimates correspond to an effective elastic
thickness of about 30 km.

COMPARING THE OBSERVED AND MODELLED


STRESS ORIENTATIONS
Figure 8 shows the observed (black lines) and estimated
(red lines) maximum horizontal stress orientations as
obtained after the adjustment of the rheological parameters
(Model A2). The standard deviation of the residuals is
45.3 and 44.5 for Models A1 and A2, respectively.
Since the standard error in the observed stress orientations
could be about 15 (Zoback 1992), the difference between
the two models is not statistically significant. However,
considering that there are many non-statistical uncertainties in the quantitative analysis of observed stress orientations as well as the associated tectonic forces, the results
obtained in this study should be viewed as semiquantitative. Comparing the observed (black lines) and modelled
(red lines) stress orientations in Figure 8, the general pattern of the observed stress orientations has been reconstructed by the numerical model, though deviations still
exist at some sites.
In eastern Australia, the fold belts are simulated as weak
zones in the numerical analysis, and substantial rotations in
the stress orientations occurred near these weak zones. The
variations in the stress orientations reflect the combined
effect of the tectonic forces and the contrast in the elastic
strength of tectonic elements on producing the intraplate
stresses. For the northern part of eastern Australia (EA1),
Model A2 (Figure 8) predicts two types of stress orientations:
northeast and north-northeast. The stress orientations modelled are generally consistent with those observed. The stress
orientations predicted for the southern part of the eastern
Australia (EA2) are also of two types, northwest and northeast, and apparent deviations exist between the predicted
north-northeast and observed northeast orientations. For the
stress indicators around the Southern Lachlan Fold Belt
(Figures 2, 8), the orientations predicted by the model are
largely of two types, northwest and north-northeast, and significant deviations exist between observations and predictions. Model A2 could not reflect the rotations of the
observed stress orientations from northeast in the north
(EA2) to northwest in the southernmost part of the region.
Stephenson and Lambeck (1985) constructed an erosion-rebound model for southeastern Australia to explain
the geomorphological and geological observations for the
uplift that occurred since Early Cenozoic time, and predicted a tensile stress field for southeastern Australia (with
northwest orientations: Stephenson & Lambeck 1985 figure
12, p. 50). Since only the regional trends of the stress orientations related to the continental-scale tectonic forces, as
well as the contrast in the elastic strength among major tectonic blocks, are simulated in our model, the local stress
changes caused by different kinematic/dynamic mechanisms, such as the erosion-rebound effect discussed by
Stephenson and Lambeck (1985), can not be directly
accounted for. However, after superimposing a local tensile
stress field (103N: green bar near EA2 in Figure 9) estimated by Stephenson and Lambeck (1985) onto the

Figure 10 Distribution of the residuals (between the observed


and modelled stress orientations) for Models A1, A2 and A3. The
vertical lines indicate the one standard deviation.

regional stress field predicted by Model A2, a hybrid model


(Model A3) is obtained, and the stress orientations predicted by the model are shown in Figure 9. There are some
significant improvements on the fit between the observations and predictions. For the points around the Southern
Lachlan Fold Belt (Figures 2, 9), the stress orientations are
now consistent with the observations.
In northwest Australia, the eastwest and northeast
stress orientations (NW1 and NW2 in Figure 8) are not fitted by the model. As mentioned before, adjusting the distribution of the collision forces at the northern boundary failed
to reduce the misfit. One possible cause for this misfit might
be the effect of some local geological structures in the
region. The borehole breakout data in this area are from the
Canning Basin, which is bounded by Fitzroy Trough in the
north. The sediments in the Fitzroy Trough are about 14 km
thick (Borissova & Symonds 1997); its length is about 700
km, but its width is merely about 100 km. The presence of
the Fitzroy Trough could have some effect on the local stress
field, but it is difficult to include this effect into our analysis
due to the narrowness of the trough, for which a model with

3D Modelling of Australian stress field

77

Figure 11 Principal-stress distribution in continental Australia (in units of 100 MPa). Also shown are the boundary of the major geological structures (yellow), the boundary of the continental shelf (green), and epicentres of the earthquakes with magnitudes of M 3.0
(triangles) and M 5.0 (stars). The blank areas represent the zones of least compression (with a compressive stress value 0 MPa).

a resolution of at least 50 km is required. In addition, a tensile stress regime has been reported in northwest Australia
(Coblentz et al. 1995), which might be related to the effect
of the continental shelf and deep basins in the region. We
therefore tentatively introduce a local tensile field with its
direction perpendicular to the coast (140N: green bar near
NW2 in Figure 9). After inclusion of the local stress field
(Model A3, Figure 9), we see that the predicted stress orientations are now consistent with the observations.
In west Australia (WA), for the eight stress indicators
(Model A3) used in the analysis, the average deviation
between the observed and modelled orientations is about
46, which is larger than the standard deviation of the
model (37.6). In addition, a test for inclusion of the
boundary-force vector with different magnitudes and orientations on four sub-segments along the western boundary failed to improve the fit (Figure 6). This suggests that
a further improvement on the fit between the observed
and modelled stress orientations with the present model is
difficult. In previous studies, two mechanisms were proposed for the rotation of the local stress field in west
Australia. Cloetingh and Wortel (1986) suggested that the

state of compression in the western part of central


Australia is induced by the action of resistant forces at the
Himalayan and Banda arc collision zones (Figure 1 inset).
However, as demonstrated by Coblentz et al. (1995,
1998), the collision forces produce stress focusing only
near the boundaries, and their effects on the orientations
of the stresses within the plate are secondary. Since most
of the stress orientations in central and northwest
Australia have been fitted by the present model (Figure 9),
a causative mechanism for the local variations of the
stress orientation in the interior of the Australian continent due to regional- or plate-scale forces is not likely. A
local mechanism for the stress changes and seismicity in
west Australia has been proposed by Lambeck et al.
(1984). They suggested that: (i) there might be a
local/regional stress field resulting from the interaction
between the Yilgarn Block (YB in Figure 2) and the nearby
Darling Fault; and (ii) the stress regime of the local stress
field could be tensile. The northsouth-oriented Darling
Fault (Borissova & Symonds 1997) is more than 800 km
long, but its width is less than 50 km. Therefore, the effect
of the fault and its interaction with the Yilgarn Block as

78

S. Zhao and R. D. Mller

Figure 12 Shear-wave speed anomalies (depth = 80 km) for the upper mantle of the Australian continent (modified from Kennett 1997,
2002). The letters mark the major zones of shear-wave speed anomaly in southern Australia (A), central Australia (C and C1), eastern
Australia (E1 and E2), northern Australia (N1) and western Australia (W1) (see text for discussion).

well as their combined effects on the local/regional stress


field could not be simulated here, given the resolution of
the present model.
Figure 10 shows the distribution of the residuals for
Models A1, A2, and A3. The numerical model has statistically fit the observed stress orientations to 37.6 (Model
A3). More than 45% of the observed stress orientations
have been fitted by our model within 25. Overall, the
numerical model provides a reasonable interpretation of
the observed stress orientations in the Australian continent.

PRINCIPAL-STRESS DISTRIBUTION AND


SEISMICITY IN CONTINENTAL AUSTRALIA
Figure 11 shows the principal-stress distribution predicted
in this study with seismicity in continental Australia superimposed. Seismicity in the Australian continent is concentrated in several zones (Figure 11).

(1) In Western Australia, earthquakes are observed


mostly in the southern part of the Yilgarn Block (also see
Figure 2) and near the North West Shelf (northwest
Australia). Fault plane solutions for two earthquakes (M =
6.8, 1968; 5.9, 1970) in Western Australia indicate thrust
faulting (Fitch et al. 1973). The source mechanisms for the
earthquakes in the North West Shelf are not well determined, as most of them occurred along the continental shelf.
(2) In South Australia, seismicity is largely confined to
the Adelaide Fold Belt and the adjacent gulf graben regions.
An average depth of about 10 km is estimated for the earthquakes recorded during 197677 by McCue and Sutton
(1979). In addition, the source mechanism solutions of the
events indicate failure by strike-slip faulting.
(3) In central Australia, seismicity is relatively diffuse. A
concentration of seismicity is observed in the Gawler Block
and near the Arunta Block (Figure 2). Source mechanism
solutions of two earthquakes (M = 6.2, 1972; 4.7, 1978) in
the Simpson Desert show failure by compression.

3D Modelling of Australian stress field

79

Figure 13 Principal-stress distribution in continental Australia (in units of 100 MPa) computed after including the rheological information from seismic tomography. Also shown are the boundary of the major geological structures (yellow), the boundary of continental shelf (green), and epicentres of the earthquakes with the magnitude of M 3.0 (triangles) and M 5.0 (stars). The blank areas
represent the zones of least compression (with a compressive principal stress value 0 MPa).

(4) In east Australia, seismicity is mostly concentrated


around the southern Lachlan Fold Belt and along the coast
adjacent to the Northern Lachlan Fold Belt and the New
England Fold Belt (Figure 2). Most of the earthquakes in
southeast Australia indicate horizontal compressive failure.
Earthquakes are indicative of where stress is concentrated so that the brittle failure limit of the crust has been
exceeded. Therefore, a correlation between seismicity and
the predicted stress distribution is expected. Comparing the
distribution of seismicity and the pattern of the stress predicted in this study, we see that such a correspondence does
exist: the seismicity in northwest and southeast Australia
falls into two bands where stress concentration is predicted.
Nevertheless, there are still several zones where the predicted stress concentration is not compatible with seismicity observed in the continent. In southwest Australia, a
zone of intense seismicity in the Yilgarn Block (Figure 2)
does not correspond to any concentration of stress predicted by our model. Along the Great Australian Bight
coastline (south Australia), a large zone of stress concentration is predicted, which is unsupported by observations.

Only the eastern part of this stress concentration zone corresponds to the seismicity near or around the Adelaide Fold
Belt. The western part of the zone does not correspond to
any recorded seismicity. In central Australia, the diffuse
seismicity is not accounted for by the stress concentration
predicted in this model.
To further interpret the seismicity in continental
Australia, it is necessary to include additional information
on the contrast in elastic strength of the tectonic elements,
such as results from seismic tomography (Kennett 1997,
2002; Simons et al. 1999) (Figure 12). Shear-velocity anomalies reveal the relative contrast in elastic strength among
the tectonic elements: seismically slow (negative anomaly
in Figure 12) indicates the material in the area is of lower
strength, and seismically fast (positive anomaly) indicates
higher strength. Seismically slowest is predicted for the
Southern and Northern Lachlan Fold Belts (marked E1 and
E2), and fastest is predicted for western (marked W1),
central (C1) and northern Australia (marked N1). In addition, relative small, but noticeable, seismically fast zones
also appear in the western part of South Australia (marked

80

S. Zhao and R. D. Mller

A) and eastern part of central Australia (marked C). These


velocity anomalies may reflect material contrasts between
the cratons, basins and fold belts. We investigate this
hypothesis by including these contrasts between lower and
higher strength lithospheric blocks based on seismic tomography in terms of differences in their Youngs moduli.
After inclusion of this additional information, a revised
map of the principal stress distribution for continental
Australia is constructed (Figure 13). A noticeable feature in
the predicted stress pattern is that the areas with the least
compression (the blank areas in the continent) are almost
seismicity free. The band of stress concentration along the
Great Australian Bight, where little seismicity is observed,
has disappeared and moved further north where it now
matches a belt of seismicity from the Musgrave Block to the
Yilgarn Block. The magnitude of the principal stresses is estimated between 10 and 40 MPa, and the deformation style is
largely compressive. The predicted area with the least compression (blank zone) in the North West Shelf corresponds
approximately to the normal faulting stress regime inferred
from the in situ stress data (Hillis 1991; Hillis & Williams
1992, 1993a, b). The stress concentration zones predicted by
the model now correspond quite well with the areas where
concentrated seismicity is observed. The improvements in
model prediction by including lateral variations in lithospheric rheology based on seismic tomography illustrate the
shortcomings of relying on surface geology and gravity/topography coherence results for estimating the spatial variation in
lithospheric rigidity. However, the inclusion of information
from seismic tomography in our model does not contribute to
further improvement of the fit between the observed and
modelled stress orientations on the Australian continent.
Features revealed by seismic tomographic analysis have a
larger length-scale and poorer resolution than those from
surface geological investigations. As discussed before, we
need to include more information on local and small-scale
stress sources into a higher resolution model in order to interpret the variations of stress orientations in some regions. The
extra information obtained from the seismic tomographic
analysis at the current resolution does not contribute significantly to a better and quantitative interpretation of the stress
orientations observed on the Australian continent, as compared to the rheological model based on the coherence of
Bouguer gravity and topography.

DISCUSSION
A three-dimensional finite-element model has been constructed and used to investigate the pattern and orientations of the tectonic stresses in continental Australia. The
model, which consists of two layers (Figure 5), provides a
spatial resolution of about 90 x 90 x 50 km. The major geological structures such as cratons and fold belts are
included in the analysis. The difference in the elastic
strength of the tectonic structures are initially estimated on
the basis of their rigidity values inferred from the coherence
of Bouguer gravity and topography (Zuber et al. 1989). The
major tectonic forces which act on the Australian continent
(such as ridge-push and plate-boundary forces) are investigated in the analysis. An inversion approach is used to estimate the relative magnitude of tectonic forces from the

observed stress orientations (equations 4 and 5). In addition, an approach for estimating the main rheological parameter (Youngs modulus) from the inversion analysis of the
observed stress orientations is also developed (equations 6
and 7) and used to estimate the values of the Youngs modulus for some of the geological structures.
Our results suggest that the slide force associated with
ridge push is the dominant force that controls the magnitude and orientations of the stress field in the Australian
continent, confirming the results of Coblentz et al. (1995,
1998). The magnitude of the slide force is estimated to be
55.8 N/m3, and the magnitude of the forces at the eastern
and northern boundaries is estimated to be >5.99 x 1012
Pa/m, and 11.8 x 1012 Pa/m, respectively (Table 2). The
estimates for the magnitude of the forces are model dependent and subject to many uncertainties (e.g. the assumed
rheological parameters and geometry of tectonic blocks).
Therefore, they may be interpreted only as semiquantitative
estimates. The boundary forces acting on the northern and
eastern boundaries of the Australian continent only have a
secondary effect on the overall stress pattern, and they do
not significantly affect the pattern of the stress in the interior of the continent.
The presence of tectonic domains with different rigidities has a significant influence on the pattern of the estimated regional and local stresses. After combining the
tectonic forces, major geological structures, and the effect
of the local stress fields in the numerical model, a reasonable fit has been achieved between the observed and modelled stress orientations (Figure 9). The in situ stress
orientations can be statistically fitted within 37.6 by the
numerical model.
The inversion analysis of rheological parameters is useful for estimating the Youngs moduli for the Northern
Lachlan Fold Belt, the New England Fold Belt, and the
Southern Lachlan Fold Belt. The adjusted values for the
flexural rigidity are 0.040 x 1025 Nm for the Northern
Lachlan Fold Belt, 0.037 x 1025 Nm for the New England
Fold Belt, and 0.040 x 1025 Nm for the Southern Lachlan
Fold Belt (Table 1), which correspond to an effective elastic
thickness of about 30 km. These estimates are about two
orders of magnitude lower than those of the cratons (~1025
Nm). The original estimates (~1022 Nm) for the fold belts
from Zuber et al. (1989), which are about three orders of
magnitude lower than those of the cratons (Table 1), may
have been underestimated (Simon et al. 2000). It appears
that the re-estimated values of the rigidity for the fold belts
from this study, which are between the maximum and minimum of the flexural rigidity estimated by Zuber et al.
(1989) and constrained by the stress-orientation data, are
more geologically plausible. Therefore, we have provided
an indirect estimate for the flexural rigidity of the fold belts
in continental Australia.
Another significant result from this study is the estimated distribution of the principal stress in the Australian
continent (Figure 13). We predict stress concentration in
northwest Australia, South Australia, and southeast
Australia. In addition, several zones with least compression
are also identified in the continent. Although the predicted
deformation style in the Australian continent by our model
is of compression and strike-slip faulting, it is plausible to
infer that normal faults are most likely to develop in the

3D Modelling of Australian stress field

areas where the least compression is predicted. It is also


noteworthy that the concentration of seismicity is not
observed inside the predicted least compression zones, but
it is mostly inside the zones of significant compression.
Therefore, the principal-stress distribution predicted here
has furnished a preliminary interpretation for the seismicity
observed in continental Australia.
Considering lateral variations in lithospheric strength in
the modelling analysis by including results from shear-wave
tomography proved to be essential to remove some firstorder artefacts from initial model, and improve the match of
modelled zones of stress concentration with observed belts
of seismicity. The results demonstrate that by combining
surface geology, lithospheric rigidity estimates from gravitytopography coherence, and seismic tomography, we
have assembled a simple rheological model for the
Australian Plate that, together with an optimised model for
plate-driving forces, accounts for the observed large-scale
patterns of intraplate seismicity in Australia.
However, like any other numerical analysis (Richardson
et al. 1979; Cloetingh & Wortel 1986; Coblentz et al. 1995,
1998), there are many limitations inherent in our model.
Although the estimated magnitude of the principal stress
between 10 and 40 MPa is compatible with the value
(~tens of megapascals, over a 100 km-thick layer) estimated by Coblentz et al. (1998), it is subject at least to the
following uncertainties: (i) since the magnitude of the
boundary forces is actually unknown, a geologically plausible value has been adapted: typically, a value of ~1012 N/m
was used; the absolute value of the forces could not be well
determined by the analysis of the stress orientation data
alone; (ii) the absolute values of the rheological parameters
of the crust/lithosphere are unknown, and a value of ~1010
Pa was used for the Youngs modulus; and (iii) the magnitude of the stresses is estimated over a layer of 50 km thickness, and the effect of the rigidity layering as well as any
other depth dependent-stress changes have been ignored,
which affects the magnitude of the calculated stresses.
What we have estimated in this study are the relative magnitude and the pattern of the tectonic stresses, rather than
the absolute magnitude of the tectonic stresses in the
Australian continent.
Our analysis shows that ignoring the effect of the gravity potential energy differences in the Australian continent
influences the modelling results. Two additional stress
fields required to fit the observed stress orientations in
northwest (NW1 and NW2) and southeast (EA2 in Figure
9) Australia may reflect the possible contribution of the
topography or gravity potential energy difference at areas
near the continental margin. One of the mechanical effects
of the gravity potential energy difference at the continental
margin is to produce a local stress field. The stress concentration reflected by seismicity near the continental margin
predicted in our model indicates that the mechanical
strength of the continental shelf is weaker than that of continental crust whose last thinning/reheating event is substantially older (see Fowler & McKenzie 1989). This
weakening effect has been incorporated in our model by
including the continental shelf as a weak zone. However,
the forces arising from the gravity potential energy difference at the continental margin are not directly simulated in
our study. Since the crustal structure at the continental

81

margin could vary from place to place, a separate analysis


of the local stress field associated with the gravity potential
energy difference or gravity instability based on a detailed
(density) structure model is required in the future.
Moreover, many small- to intermediate-scale geological
structures are not included in our study, such as basins and
faults. For some of the basins, the depth of the sediment to
the basement is more than 10 km (e.g. the Browse Basin in
the North West Shelf: Borissora & Symonds 1997), and for
some crustal-scale faults, their length scale is up to 500 km
(e.g. the Darling Fault in Western Australia). Inclusion of
geological structures into a future model with a higher spatial resolution will alter the magnitude as well as the pattern
of the calculated stress in the areas around or close to these
structures. Further, the present activity or reactivation of
faults also influences the pattern of the local/regional stress
field (Sandiford & Hand 1998). These could be the objects
of future local or regional stress analysis, which may be
designed to explore the effects of the local or regional geological structures as well as their activity on the tectonic
stress field. These factors discussed above could in part
account for the reason that about 20% of the observed
stress orientations are not well fitted by our continentalscale model.
The interaction between the lithosphere and mantle or
the upper and lower crust has not been considered in our
analysis. The stress transferred from the lower or the upper
mantle into the upper crust or lithosphere associated with
pre-existing geological structures could influence the pattern of the local or regional stress field (Kusznir & Bott
1977; Lambeck et al. 1984). However, the magnitude and
properties of the transferred stresses, which are model
dependent, are very difficult to assess. For instance, a
stress difference of 50200 MPa for eastern Australia is predicted by the erosionrebound model of Stephenson and
Lambeck (1985), which is almost at the same magnitude as
that of the predicted regional stress field (Coblentz et al.
1995, 1998). Our study has shown that a combination of
local stress-relaxation processes associated with some distinct geological structures with a continental-scale model
better accounts for the observed stresses. However, the
increasing uncertainties with adding more (speculative)
geodynamic mechanisms into any model will further
increase the ambiguity of the results. Therefore, these geodynamic processes are currently modelled separately.
The type of deformation and the stress regime inferred
from the in situ stress measurements, the seismic source
mechanisms, and the numerical model experiments are all
depth dependent. The information of the deformation type
and faulting style estimated from earthquakes in the upper
crust could be different from those from the earthquakes in
the middle crust, or different from those obtained from the
in situ stress measurements. Therefore, the available information on the stress regime from the in situ stress measurements, the earthquakes source mechanisms, and the
numerical modelling of the Australian continent is incomplete. While the dominant deformation style in the
Australian continent inferred from this study is compression, which is consistent with that from the in situ stress
measurements, caution should be taken when extrapolating the results to the state of the stress at depth because of
the inherent uncertainties stated above.

82

S. Zhao and R. D. Mller

SUMMARY

REFERENCES

The main results from the 3D stress analysis with the finiteelement method for the Australian continent are as follows.
(1) The ridge-push force is the dominant force which
controls the magnitude and pattern of the first-order
stresses in the Australian continent. The effect of the
boundary forces are secondary and they mostly influence
the pattern of the stress in areas near the boundaries. There
is no need to invoke the drag force to explain the first-order
stress pattern in the Australian continent partly because of
our poor understanding of the properties of the drag force
and the insensitivity of the stress orientation data to the
drag force. These results are consistent with those obtained
by Coblentz et al. (1995).
(2) Geological structures significantly affect the magnitude and pattern of modelled stresses. Combining spatial
variations in rigidity between major geological structures
(cratons and fold belts) and a tectonic-force model, by
simultaneously inverting for stress orientations and tectonic-force vectors, a fairly good fit has been achieved
between the observed and modelled stress orientations.
The model can explain statistically about 45% of the
observed stress orientations within 25, and about 62%
within 40.
(3) The model also provides an indirect estimate of the
flexural rigidity for the Northern Lachlan Fold Belt (0.040 x
1025 Nm), the New England Fold Belt (0.037 x 1025 Nm)
and the Southern Lachlan Fold Belt (0.040 x 1025 Nm).
These estimates correspond to an effective elastic thickness
of about 30 km.
(4) A preliminary map of principal-stress distribution
(Figure 13) is constructed for continental Australia, in
which the relative magnitude of the principal stress over
the continent can be assessed. The predicted stress-concentration zones in general correspond to the areas
where intensive seismicity is observed. In addition, the
least compression is predicted in several zones where
earthquakes are relatively sparse, and it is also inferred
that normal faults would mostly likely develop in these
zones.
(5) While the model from this study provides a reasonable interpretation for the stress orientations and seismicity
observed in the Australian continent, about 20% of the
observed stress orientations are not well-fitted by the
model. The main reason for this could be that the disturbances in the stress field associated with some local or
regional geological structures (and their present activity)
cannot be simulated in our present continental-scale
model.

BORISSOVA I. & SYMOND, P. A. 1997. Basins of Australia

ACKNOWLEDGEMENTS
We wish to thank D. Coblentz, R. Hillis and M. Sandiford
for their constructive reviews, which improved this manuscript substantially, B. L. N. Kennett for kindly providing us
his latest shear-wave model of the Australian lithosphere,
and G. Clitheroe for providing some of the data used in this
study. This research is supported by an ARC SPIRT grant
and industry sponsorship by BHP, Santos, Shell and
Woodside.

(1:6 000 000 scale map) (1st Edition). Australian Geological


Survey Organisation, Canberra.
CLITHEROE G., GUDMUNDSSON O. & KENNETT B. L. N. 2000. The crustal
thickness of Australia. Journal of Geophysical Research 105,
1369713713.
CLOETINGH S. & WORTEL R. 1986. Stress in the Indo-Australian plate.
Tectonophysics 132, 4967.
COBLENTZ D. D., RICHARDSON R. M. & SANDIFORD M. 1994. On the
gravitational potential of the Earths lithosphere. Tectonics 13,
929945.
COBLENTZ D. D., SANDIFORD M., RICHARDSON R. M., ZHOU S. & HILLIS
R. 1995. The origins of the intraplate stress field in continental
Australia. Earth and Planetary Science Letters 133, 299309.
COBLENTZ D. D., ZHOU S., HILLIS R. R., RICHARDSON R. M. & SANDIFORD
M. 1998. Topography, boundary forces, and the Indo-Australian
intraplate stress field. Journal of Geophysical Research 103,
919931.
CULL J. 1991. Heat flow and regional geophysics in Australia. In:
Cermak V. & Rybach L. eds. Terrestrial Heat Flow and the
Lithosphere Structure, pp. 486500. Springer-Verlag, New York.
DENHAM D. 1988. Australian seismicity: the puzzle of the not so stable continent. Seismological Research Letters 49, 289295.
DENHAM D., ALEXANDER L. G. & WOROTNICKI, G. 1979. Stress in the
Australian crust: evidence from earthquakes and in situ stress
measurements. BMR Journal of Australian Geology & Geophysics
4, 289295.
DENHAM D. & WINDSOR C. R. 1991. The crustal stress pattern in
Australia continent. Exploration Geophysics 22, 101105.
FOWLER S. & MCKENZIE D. 1989. Gravity studies of the Rockall and
Exmouth Plateaux using Seasat altimetry. Basin Research 2,
2734.
FITCH T. J., WORTHINGTON M. H. & EVERINGHAM I. B. 1973.
Mechanisms of Australian earthquakes and contemporary stress
in the Indian ocean plate. Earth and Planetary Science Letters 18,
345356.
GOETZE C. & EVANS B. 1979. Stress and temperature in the bending
lithosphere as constrained by experimental rock mechanics.
Geophysical Journal of the Royal Astronomical Society 59,
463478.
HILLIS R. R. 1991. AustraliaBanda collision and in situ stress in the
Vulcan sub-basin (Timor Sea) as revealed by borehole breakout
data. Exploration Geophysics 22, 189194.
HILLIS R. R., ENEVER J. R. & REYNOLDS S. D. 1999. in situ stress field of
eastern Australia. Australian Journal of Earth Sciences 46,
813825.
HILLIS R. R., MEYER J. J. & REYNOLDS S. D. 1998. The Australian stress
map. Exploration Geophysics 29, 420427.
HILLIS R. R. & REYNOLDS S. D. 2000. The Australian stress map,
Journal of the Geological Society of London 157, 915921.
HILLIS R. R., SANDIFORD M., COBLENTZ D. D. & ZHOU S. 1997.
Modelling the contemporary stress field and its implications for
hydrocarbon exploration. Exploration Geophysics 28, 8893.
HILLIS R. R. & WILLIAMS A. F. 1992. Borehole breakouts and stress
analysis in the Timor Sea. Geological Society of London Special
Publication 66, 157168.
HILLIS R. R. & WILLIAMS A. F. 1993a. The contemporary stress of the
BarrowDampier Sub-basin and its implications for horizontal
drilling. Exploration Geophysics 24, 567576.
HILLIS R. R. & WILLIAMS A. F. 1993b. The stress field of the North West
Shelf and wellbore stability. American Petroleum Association
Journal 33, 373385.
KENNETT B. L. N. 1997. The mantle beneath Australia. AGSO Journal
of Australian Geology & Geophysics 17, 4954.
KENNETT B. L. N. 2003. Seismic structure in the mantle beneath
Australia. Geological Society of Australia Special Publication 22
and Geological Society of America Special Paper xy.
KUSZNIR N. J. & BOTT M. H. P. 1977. Stress concentration in the upper
lithosphere caused by underlying visco-elastic creep.
Tectonophysics 43, 247256.
LAMBECK K. 1983a. Structure and evolution of the intracraton basins
of central Australia. Geophysical Journal of the Royal Astronomical
Society 74, 843886.
LAMBECK K. 1983b. Teleseismic travel-time anomalies and deep

3D Modelling of Australian stress field


crustal structure in central Australia. Geophysical Journal of the
Royal Astronomical Society 94, 105124.
LAMBECK K. & PENNEY C. 1984. Teleseismic travel time anomalies and
crustal structure in central Australia. Physics of the Earth and
Planetary Interiors 34, 4656.
LAMBECK K., MCQUEEN H. W. S., STEPHENSON R. A. & DENHAM D.
1984. The state of stress within the Australian continent. Annales
Geophysicae 2, 723742.
LILLEY F. E. M., WOODS D. V. & SLOANE M. N. 1981. Electrical conductivity profiles and implications for the absence or presence of
partial melting beneath central and southeast Australia. Physics of
the Earth and Planetary Interiors 25, 202209.
LISTER C. R. B. 1975. Gravitational drive on oceanic plates caused by
thermal contraction. Nature 257, 663665.
MCCUE K. F. & SUTTON D. J. 1979. South Australian earthquakes during 1976 and 1977. Journal of Geological Society of Australia 26,
231236.
MENKE W. 1984. Geophysical Data Analysis: Discrete Inverse Theory.
Academic Press, Orlando.
MLLER R. D., ROEST W. R., ROYER J-Y., GAHAGAN L. M. & SCLATER J. G.
1997. Digital isochrons of the worlds ocean floor. Journal of
Geophysical Research 102, 32113214.
MUELLER B., REINECKER J. & FUCHS K. 2000. The 2000 release of the
World Stress Map.
<http://www-wsm.physik.uni-karlsruhe.de/pub2000/>.
PLUMB K. A. 1979a. The tectonic evolution of Australia. Earth Sciences
Reviews 14, 205249.
PLUMB K. A. 1979b. Structure and tectonic style of the Precambrian
shields and platforms of northern Australia. Tectonophysics 58,
291325.
REYNOLDS S. D., COBLENTZ D. D. & HILLIS R. R. 2003. Influences of
plate-boundary forces on the regional intraplate stress field of
continental Australia. Geological Society of Australia Special
Publication 22 and Geological Society of America Special Paper xy.
RICHARDSON R. M. 1992. Ridge forces, absolute plate motions, and the
intraplate stress field. Journal of Geophysical Research 97,
1173911748.
RICHARDSON R. M., SOLOMON S. C. & SLEEP N. H. 1979. Tectonic stress
in the plates. Reviews of Geophysics 17, 9811019.
SANDIFORD M., COBLENTZ D. D. & RICHARDSON R. M. 1995. Ridge
torques and continental collision in the Indian-Australian plate.
Geology 23, 653656.
SANDIFORD M. & HAND M. 1998. Controls on the locus of intraplate

83

deformation in central Australia. Earth and Planetary Science


Letters 162, 97110.
SIMONS F. J. & VAN DER HILST R. D. 2002. Age-dependent seismic thickness and mechanical strength of the Australian lithosphere.
Geophysical Research Letters 29, 10291033.
SIMONS F. J., ZIELHUIS A. & VAN DER HILST R. D. 1999. The deep structure of the Australian continent from surface-wave tomography.
Lithos 48, 1743.
SIMONS F. J., ZUBER M. T. & KORENAGA J. 2000. Isostatic response of
the Australian lithosphere: estimation of effective elastic thickness
and anisotropy using multitaper spectral analysis. Journal of
Geophysical Research 105, 1916319184.
STEPHENSON R. & LAMBECK K. 1985. Erosion-isostatic rebound models
for uplift: an application to southeastern Australia. Geophysical
Journal of the Royal Astronomical Society 82, 3155.
TARANTOLA A. 1987. Inverse Problem Theory, Methods for Data Fitting
and Model Parameter Estimation. Elsevier, Amsterdam.
TURCOTTE D. L. & SCHUBERT G. 1982. Geodynamics: Applications of
Continuum Physics to Geological Problems. John Wiley & Sons
Inc., New York.
ZHANG Y., SCHEIBNER E., ORD A. & HOBBS B. E. 1996. Numerical modelling of crustal stresses in the eastern Australian passive margin,
Australian Journal of Earth Sciences 43, 161175.
ZOBACK M. D. & ZOBACK M. L. 1991. Tectonic stress field of North
America and relative plate motion. In: Slemmons D. L., Engdahl
E. R., Zoback M. D. & Blackwell M. L. eds. Neotectonics of North
America, pp. 339366. Geological Society of America, Boulder.
ZOBACK M. L. 1992. First- and second- order patterns of stress in the
lithosphere: the World Stress Map Project. Journal of Geophysical
Research 97, 1170311728.
ZOBACK M. L., ZOBACK M. D., ADAMS J., ASSUMPCAO M., BELL S.,
BERGMAN E. A., BLUMLING P., BRERETON N. R., DENHAM D., DING J.,
FUCHS K., GAY N., GREGERSEN S., GUPTA H. K., GVISHIANI A., JACOB
K., KLEIN R., KNOLL P., MAGEE M., MERCIER J. L., MULLER B. C.,
PAQUIN C., RAJENDRAN K., STEPHANSSON O., SUAREZ G., SUTER M.,
UDIAS A., XU Z. H. & ZHIZHIN M. 1989. Global patterns of tectonic
stress. Nature 341, 291298.
ZUBER M. T., BECHTEL T. D. & FORSYTH D. W. 1989. Effective elastic
thickness of the lithosphere and mechanisms of isostatic compensation in Australia. Journal of Geophysical Research 94,
93539367.
Received 23 July 2001; accepted 29 August 2002

Potrebbero piacerti anche