You are on page 1of 20

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

net/publication/286413187

Present-day stress orientation in the Clarence-Moreton Basin of New South


Wales, Australia: A new high density dataset reveals local stress rotations

Article in Basin Research · February 2017


DOI: 10.1111/bre.12175

CITATIONS READS

78 825

4 authors:

Mojtaba Rajabi Mark Tingay


The University of Queensland University of Adelaide
78 PUBLICATIONS 2,270 CITATIONS 123 PUBLICATIONS 6,109 CITATIONS

SEE PROFILE SEE PROFILE

Rosalind C. King Oliver Heidbach


University of Adelaide Helmholtz-Zentrum Potsdam - Deutsches GeoForschungsZentrum GFZ
108 PUBLICATIONS 1,376 CITATIONS 186 PUBLICATIONS 6,268 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Mark Tingay on 07 April 2016.

The user has requested enhancement of the downloaded file.


EAGE

Basin Research (2016) 1–19, doi: 10.1111/bre.12175

Present-day stress orientation in the Clarence-


Moreton Basin of New SouthWales, Australia: a new
high density dataset reveals local stress rotations
Mojtaba Rajabi,* Mark Tingay,* Rosalind King† and Oliver Heidbach‡
*Australian School of Petroleum, University of Adelaide, Adelaide, SA, Australia
†School of Earth and Environmental Sciences, University of Adelaide, Adelaide, SA, Australia
‡GFZ German Research Centre for Geosciences, Telegrafenberg, Potsdam, Germany

ABSTRACT
Early phases of the Australian Stress Map project revealed that plate boundary forces acting on the
Indo-Australian Plate control the long wavelength of the maximum horizontal present-day stress ori-
entation in the Australian continent. However, all numerical models of the stress field to date are
unable to predict the observed orientation of maximum horizontal stress in the northeast of New
South Wales, Australia. Recent coal seam gas exploration in the Clarence-Moreton Basin, eastern
Australia, provides an opportunity to better evaluate the state of crustal stress in this part of the con-
tinent where only limited information was available prior to this study. Herein, we conduct the first
analysis of the present-day tectonic stress in the Clarence-Moreton Basin, from drilling-induced ten-
sile fractures and borehole breakouts interpreted using 11.3 km of acoustic image logs in 27 vertical
wells. A total of 2822 drilling-induced stress indicators suggest a mean orientation of N069°E (23°)
for the maximum horizontal present-day stress in the basin which is different from that predicted by
published geomechanical-numerical models. In addition, we find significant localised perturbations
of borehole breakouts, both spatially and with depth, that are consistent with stress variations near
faults, fractures and lithological contrasts, indicating that local structures are an important source of
stress in the basin. The observation that structures can have a major control on the stresses in the
basin suggests that, while gravity and plate boundary forces have the major role in the long wave-
length (first-order) stress pattern of the continent, local perturbations are significant and can lead to
substantial changes in the orientation of the maximum horizontal present-day stress, particularly at
the basin scale. These local perturbations of stress as a result of faults and fractures have important
implications in borehole stability and permeability of coal seam gas reservoirs for safe and sustainable
extraction of methane in this area.

Australian continent shows considerable directional vari-


INTRODUCTION
ability on a 200–500 km wavelength, being broadly east-
The contemporary stress field of Australia is highly vari- west in south-western part of the continent, northeast-
able and unique as compared to other continents (Zoback, southwest to north-south in northern and north-western
1992; Hillis & Reynolds, 2003; Heidbach et al., 2010). Australia, and primarily northwest-southeast in south-
Maximum horizontal present-day stress (SHmax) orienta- eastern side of the continent (Fig. 1), (Hillis & Reynolds,
tion in other continental plates such as, North America, 2003). North-eastern Australia is characterised by a dis-
South America and Western Europe, observed to be pri- tinctive ‘horse-shoe’ shaped rotation from approximately
marily parallel to absolute plate motion, supporting the north-south in the Amadeus Basin, east-west in the
notion that the plate boundary forces that drive tectonic Cooper-Eromanga Basins and north-south in the Bowen
plates also control the intra-plate stress field (Richardson, Basin (Fig. 1), (Hillis & Reynolds, 2003). A large amount
1992; Zoback, 1992). However, the Australian continent of this complexity has been explained by combination of
displays a complex pattern of present-day stress and is not far-field plate boundary forces; however, the detail pat-
oriented (sub) parallel to the north-northeast absolute tern of stress in eastern Australia has not been satisfacto-
plate motion (Richardson, 1992; Zoback, 1992; Hillis & rily captured in these previous studies (Coblentz et al.,
Reynolds, 2003). The orientation of SHmax in the 1995, 1998; Reynolds et al., 2002, 2003; Sandiford et al.,
2004; Hillis et al., 2008).
Most of our information and understanding about
Correspondence: Mojtaba Rajabi, Australian School of Petro-
leum, University of Adelaide, Adelaide, SA 5005, Australia.
the state of stress in eastern Australia is based on stress
E-mail: mojtaba.rajabi@adelaide.edu.au observations in the Bowen and Sydney Basins (Hillis

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 1
M. Rajabi et al.

Fig. 1. Present-day stress pattern of


Australia (Hillis & Reynolds, 2003; Heid-
bach et al., 2010). Different symbols
show the orientation of maximum hori-
zontal stress (SHmax) and different col-
ours show different types of stress
indicators and the stress regimes (TF,
thrust faulting; SS, strike–slip faulting;
NF, normal faulting; black, unknown).
Length of the lines indicates the quality
of data. The SHmax orientations clearly
demonstrate the variability in stress
(compared with plate motion) particu-
larly in northeast Australia where pat-
terns show a ‘horseshoe shape’ trend.
Solid grey lines show stress trajectories of
Hillis & Reynolds (2003). Note the lack
of the present-day stress indicators in the
Clarence-Moreton Basin.

et al., 1999), which are mostly from hydraulic fracture These studies reveal that although it is clear that plate
and overcoring test measurements. These limited set of boundary forces contribute significantly to the overall
observational stress indicators have been used in numer- stress state and control the long wavelength stress pattern
ous geomechanical-numerical models to constrain and of the SHmax orientation; it is not clear to what extent local
predict the contemporary state of stress in the Australian structures, regional to local rock strength and density con-
continent (Cloetingh & Wortel, 1985, 1986; Coblentz trasts contribute to the stress state particularly in most
et al., 1995, 1998; Reynolds et al., 2002, 2003; Zhao & eastern Australia. To better understand the relative con-
M€ uller, 2003; Burbidge, 2004; Dyksterhuis et al., 2005a; tribution of these parameters, more observational stress
M€ uller et al., 2012). However, all these large scale models indicators and geomechanical-numerical models (regional
have been unable to explain the pattern of SHmax orienta- to local scales) are required to investigate the potential
tion particularly in north-eastern Australia and New influence of local stress sources (Heidbach et al., 2007).
South Wales (NSW). Furthermore, predicted stress mag- The recent exploration of coal seam gas in eastern
nitudes and thus the stress regime are also not predicted Australia provides an opportunity to make the first pre-
reasonably. For example, Cloetingh & Wortel (1985, sent-day stress map in this poorly resolved part of the
1986) by using a homogenous elastic model predicted a Australian continent where no information was available
significant east-northeast – west-southwest extension for prior to this study. This paper aims to map the present-
the most eastern parts of Australia. Reynolds et al. (2002, day stress field in the Clarence-Moreton (C-M) Basin
2003), also using a 2D elastic approach predicted a gener- using interpretation of borehole failure on acoustic
ally north-northwest – south-southeast orientation for image logs in 27 vertical coal seam gas wells. We
SHmax in the eastern part of NSW and north-northeast – examine the SHmax orientation of the C-M Basin in
south-southwest for the eastern part of Queensland. Bur- north-eastern NSW and compare the observed stress
bidge (2004) suggested an alternative anelastic model and field with previous Australian stress models. We show
predicted an east-northeast – west-southwest SHmax ori- these models do not accurately predict the observed
entation for most parts of north-eastern Australia and east-northeast – west-southwest mean SHmax orientation
NSW; but it has been unable to predict the observed in this sedimentary basin. We then investigate the
SHmax orientation and the stress regime in many other observation of localised rotations in the SHmax orienta-
basins throughout the continent. Dyksterhuis et al. tion, and describe how these appear to be caused by
(2005a, b); Dyksterhuis & M€ uller (2008) and M€ uller et al. faults and fractures perturbing the stress field. The
(2012) by 2D elastic models predict a generally north- strong influence of faults and fractures on the orienta-
west-southeast SHmax orientation for most eastern part of tion of SHmax suggests that local structures are impor-
Australia which is not in agreement with the observed tant sources of stress in the C-M Basin. Finally, we
SHmax orientation in different basins of eastern Australia. highlight the significant implications of local stress per-
Also the 3D model of Australia from Zhao & M€ uller turbations around deformational structures for coal seam
(2003) could not resolve these regional to local deviations gas exploration and production, particularly the mod-
from the general trend. elling and planning of hydraulic fracture stimulations.

© 2015 The Authors


2 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

deposition of continental sediments (Stewart & Adler,


GEOLOGICAL SETTING OF THE 1995). According to the tectonic history of the basin, pro-
CLARENCE-MORETON BASIN posed by Sommacal et al. (2008), several events con-
The intracratonic C-M Basin which is located in south- trolled the development of the C-M Basin. Sommacal
eastern Queensland and north-eastern NSW (Fig. 2) cov- et al. (2008) suggested that numerous pre-existing struc-
ers areas of approximately 26000 km2 onshore and tures, which originally formed in the early-middle Trias-
1000 km2 offshore (Jell et al., 2013). The basin was origi- sic Hunter-Bowen event, were reactivated during the
nally formed in association with extensive Late Permian compressional/transpressional Moonie event (95 Ma)
to Early Triassic plutonic rocks and silicic volcanics that where conjugate strike-slip faults accommodated crustal
were emplaced into the New England Orogen during a shortening of the C-M Basin (Sommacal et al., 2008).
period of back-arc extension and strike-slip faulting Following the Moonie event, extensive Neogene volcanic
(Shaw & Flood, 1981; O’Brien et al., 1994). The C-M activity has also been reported mainly due to the initiation
Basin and other Mesozoic basins, such as the Esk Trough of rifting and sea floor spreading along the eastern Aus-
and Ipswich Basin, were initiated on basement by major tralian margin (Sommacal et al., 2008).
strike-slip faults during this time. Oblique extension ter- The C-M Basin, which is divided into three north
minated in the early Late Triassic and was followed by trending sub-basins, namely Cecil Plains, Laidley and
thermal relaxation subsidence, resulting in significant Logan from north-west to south-east, (Martin & Saxby,

Fig. 2. Simplified geology and struc-


tural elements of the Clarence-Moreton
Basin (modified from Stewart & Adler,
1995; Wells & O’Brien, 1994). Cecil
Plains, Laidley and Logan are three
major sub-basins of the C-M Basin.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 3
M. Rajabi et al.

1982); (Fig. 2), is flanked by the D’Aguilar Block, Esk tary succession in this basin (particularly in NSW) has
Trough and Yarraman Block in the north; Beenleigh potential for both conventional and unconventional
Block in the east; Coffs Harbour, Emu Creek, Woolomin- hydrocarbon production (Fig. 3). The Walloon Coal
Texas and Silverwood blocks in the South and South- Measures has high vitrinite content and, due to numerous
west. To the west, in Queensland, Kumbarilla Ridge sep- fault structures in the basin, has excellent potential for
arates the C-M Basin from the Surat Basin (Fig. 2). coal seam gas production (Stewart & Adler, 1995). The
However, Day et al. (2008) and more recently Jell et al. Ipswich and Walloon Coal Measures of the C-M Basin
(2013) have suggested that the boundary between Surat are considered as potential conventional source rocks
and C-M basins is poorly defined and should be placed (Powell et al., 1993; Willis, 1994; Ingram et al., 1996; Jell
further eastward at the Toowoomba Strait so that the et al., 2013). For example, Powell et al. (1993) showed
Cecil Plains Sub-basin would be considered as a part of that Walloon Coal Measure and Koukandowie Formation
the Surat Basin (Fig. 2). contain type II/III kerogen and are mature for oil genera-
The Triassic-Jurassic ‘non-marine sediments’ of the C- tion in south-eastern Queensland and in the central part
M Basin (Fig. 3), which unconformably overlie older sed- of the basin in NSW, but are overmature along its eastern
iments, metasediments and igneous rocks, are similar to margin. In addition, they also proposed that the genera-
the Eromanga and Surat basins in terms of age and depo- tion of hydrocarbon took place in the period of 80–
sitional environment (Powell et al., 1993). The sedimen- 100 Ma. Some formations, mainly sandstone have good

Fig. 3. Stratigraphy and petroleum potential of the Clarence-Moreton Basin (after Stewart & Adler, 1995; Wells & O’Brien, 1994).
Most of the studied wells penetrated into the Walloon Coal Measures.

© 2015 The Authors


4 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

reservoir quality, but the lack of an appropriate seal in and acoustic) contrast of rock and fluids. Drilling-induced
many parts of the basin is the biggest challenge to com- tensile fractures appear as pairs of low-amplitude vertical
plete a typical source-to-reservoir petroleum system. fractures on acoustic images which separated by approxi-
Although, the C-M Basin has not been extensively mately 180° (Aadnoy & Bell, 1998). Borehole breakouts
explored, the NSW portion of the basin is considered to appear as a pair of elongation zones (low-amplitude zones)
be most favourable for hydrocarbon exploration (O’Brien parallel to the borehole axis and separated by approxi-
et al., 1994; Stewart & Adler, 1995; Jell et al., 2013). mately 180° (Fig. 4).
Interpretation of BOs on acoustic image logs, which
provides a fully (360°) covered picture of the borehole
wall, allows breakouts to be interpreted in various stages
DRILLING-RELATED STRESS of development, ranging from very early initial stages of
INDICATORS failure (incipient borehole breakout) right through to a
Assuming that the vertical stress (SV) is a principal stress, mature or fully developed stage. Up to now, several
the 3D state of crustal stress in the earth can be described authors have demonstrated the evolution and formation
by four components: the magnitude of three principal of BOs in laboratory samples (e.g. Haimson & Herrick,
stresses (i.e. SV, SHmax and Shmin) and the SHmax orienta- 1986; Lee & Haimson, 1993; Haimson, 2007). Haimson
tion (Bell, 1996a, b). Of these four components, the orien- (2007) examined the formation of BOs at microscopic
tation of SHmax plays an important role in different scale, and revealed two general types of breakout, termed
aspects of geosciences such as geodynamics, tectono- ‘dog-eared’ and ‘slot-like’. Regardless of different modes
physics, and petroleum geomechanics (Zoback et al., of failure for breakout formation (e.g. tensile, shear failure
1989; Richardson, 1992; Bell, 1996b). The present-day and compaction band), we attempt to show the different
SHmax orientation is commonly derived in petroleum stages in the formation of BOs using in situ acoustic image
basins from interpretations of stress-induced failure of log data from a single well.
boreholes, known as drilling-induced tensile fractures and Incipient BOs form at the initial stage in breakout
borehole breakouts (Plumb & Hickman, 1985; Zoback, development (Barton & Moos, 2010) and are pairs of frac-
1992; Bell, 1996a; Aadnoy & Bell, 1998). tures on opposite sides of the borehole wall (Aadnoy &
Borehole breakouts (BOs) are enlargements of the well- Bell, 1998; Barton & Moos, 2010); (Fig. 5a, b). These
bore due to stress concentration around the well (circum- fractures propagate and intersect horizontally and verti-
ferential or hoop stress), (Bell & Gough, 1979). The cally (Fig. 5b, c) and start to cause pieces to break off
enlargement of the wellbore is the result of compressive from the wellbore wall (Fig. 5d). The zone of conjugate
shear failure on the borehole wall which causes spalling of fractures may propagate further, causing the breakout to
pieces of the borehole wall (Bell & Gough, 1979). Hence, become more fully developed and finally form as a mature
BOs develop perpendicular to the present-day SHmax ori- breakout (Fig. 5e, f). When breakouts are very small, or
entation (Bell & Gough, 1979). Drilling-induced tensile only just partially developed (Fig. 5a, b or even c), they
fractures (DITFs) are related to tensile failure of the bore- are similar to DITFs and there is a risk of misinterpreting
hole wall and occur where the circumferential stress is less BOs as DITFs, particularly in instances where there is no
than the tensile strength of the rock (Aadnoy, 1990; Bar- mature BOs in the well.
ton & Moos, 2010). These vertical fractures form parallel
to the SHmax orientation in vertical boreholes (Aadnoy &
Bell, 1998; Barton & Moos, 2010). Figure 4 shows the THE ORIENTATION OF MAXIMUM
typical examples of these features in the studied wells in
HORIZONTAL PRESENT-DAY STRESS IN
the C-M Basin.
In this study, the present-day SHmax orientation was
THE CLARENCE-MORETON BASIN
determined from interpretation of BOs and DITFs picked We analysed 11.3 km of acoustic image logs and borehole
on acoustic borehole image logs, as described in detail in geometry tools for interpretation of BOs and DITFs in 27
the following section. Borehole geometry logging tools, vertical coal seam gas wells in the C-M Basin of NSW,
which are a type of four-arm caliper log, have also been Australia. The location and details of the studied wells are
used to interpret breakouts in two of the studied wells, summarised in Table 1 and Fig. 2. This is an area where
based on the standard procedure proposed by the World no prior information about SHmax was available in the
Stress Map (WSM) Project (Plumb & Hickman, 1985; WSM database. A total of 2805 BOs and 17 DITFs, with
Bell, 1990; Zajac & Stock, 1997; Reinecker et al., 2003). a combined length of 1606 m, were interpreted in the
studied wells (Table 1). The quality ranking system of
the WSM project (Heidbach et al., 2010) was used to
Interpretation of drilling-related stress
assess the reliability of the results in each well (Table 1
indicators in acoustic image logs
and 2).
Borehole image logs provide a ‘pseudo-picture’ of the The regional stress pattern of the Australian continent
borehole wall based on a physical (for example, resistivity for each basin has been described by the concept of ‘stress

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 5
M. Rajabi et al.

Fig. 4. (a) Orientation of borehole breakouts (BOs) and drilling-induced tensile fractures (DITFs) with respect to the circumferential
stress around the wellbore (modified from Hillis & Reynolds, 2000). (b and c) Typical examples of DITFs observed on opposite sides
of the wellbore in Corella#E07 and South Casino#07 (separated by approximately 180°). The DITFs picked here are oriented approx-
imately east-northeast – west-southwest representing the orientation of maximum horizontal present-day stress. (d) An example of
BOs in acoustic image in the Middle Creek#1A which are interpreted as a pair of low-amplitude zones on the opposite sides of the
borehole wall. These BOs have a north-south orientation and thus show an east-west orientation for maximum horizontal stress in this
depth interval.

provinces’ proposed by Hillis & Reynolds (2000, 2003). (Mardia, 1972; Davis, 2002). The Rayleigh test as an
This method aims to determine the average regional approach to assess whether there is statistical evidence of
SHmax orientation and its reliability for each basin. As the one-sidedness or directedness on the present-day stress
observations of this study are from a single sedimentary was first used by Coblentz & Richardson (1995) to analyse
basin in a particular geographical region, we consider the the global stress pattern. Hillis & Reynolds (2000, 2003)
C-M Basin as a new stress province for the Australian then applied this test for the concept of ‘stress province’
Stress Map (ASM) database. To calculate the statistical in the ASM project. To apply the test, we calculated the
parameters, all stress orientation indicators were first mean SHmax orientation, standard deviation and R-value
weighted based on the results of the WSM quality rank- (length of the mean resultant vector) for the C-M Basin.
ing system. D-quality data received a weight of ‘one’ up R-value is used to reject the null hypothesis that there is
to A-quality receiving a weight of ‘four’ (Hillis & Rey- no statistical trend in the data at a certain confidence
nolds, 2003). level.
In order to assess the significance of the mean SHmax We defined a null hypothesis that ‘stress orientations in
for the province, the Rayleigh test was used. This test is a the C-M province are random’ and assumed six types of
method to evaluate whether the population from which provinces. A type 1 stress province indicates that the null
the sample is drawn differs significantly from randomness hypothesis can be rejected at the 99.9% confidence level,

Fig. 5. A snapshot from different stages of borehole breakouts development on borehole wall. The first stage in breakouts formation
start by initiation of some vertical failures (see a) which is completely similar to drilling-induced tensile fractures (compare with Fig-
ure 4). These fractures come together and propagate (both vertically and horizontally) and this is the critical condition where rocks
start to fall off (see b, c and d). Check the cross-sections of the borehole wall (at a, b and c), which in spite of the presence of breakout,
are still circular. In time, more rocks spall off to the borehole and breakouts are more developed where the cross-sections of the bore-
hole wall are completely oval (e and f). Note that all the stages are from a single borehole.

© 2015 The Authors


6 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

(c)

(f)
(b)

(e)
(d)
(a)

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 7
8
Table 1. Orientation of maximum horizontal stress in studied wells across the Clarence-Moreton Basin of NSW. Tool run shows the depth coverage of image log data in each well.

Stress indicators

Well location Maximum BO DITF


deviation
M. Rajabi et al.

of SHmax SHmax
borehole Combined orientation Combined orientation
Well name Latitude Longitude (deg) Tool run (m) Tool Count lengths (m) (deg) SD Quality Count lengths (m) (deg) SD Quality

Bungawalbin 29.064 153.232 3.3 80.01–777.4 AST 98 42.4 66 8.8 B 0 0 N/A N/A E
Creek 01
Bungawalbin 29.210 153.132 13 77–819 AST 87 77.9 34 13.7 B 0 0 N/A N/A E
Creek 02
Bungawalbin 29.173 153.179 12 88.61–660.9 AST 45 41.6 47 12.2 B 0 0 N/A N/A E
Creek 04
Cedar Point 01 28.699 153.004 3.3 399–783.8 BGT 37 104.6 50 12 B 0 0 N/A N/A E
Corella E01 28.764 152.967 6 427.7–736.3 AST 155 66.1 87 12.7 B 3 1.17 82 6.5 D
Corella E02 28.811 152.972 6.2 86.6–737.4 AST 167 116.3 61 12 A 0 0 N/A N/A E
Corella E03 28.763 152.921 8.7 321.01–376.41 AST 2 3 67 8.5 D 0 0 N/A N/A E
Corella E04 28.805 152.939 5.4 347.6–791.7 AST 185 77.9 69 19.8 B 2 0.72 55 0.8 D
Corella E06 28.774 153.024 6.9 408.6–802.5 AST 124 46 74 22.5 C 0 0 N/A N/A E
Corella E07 28.829 153.056 5.3 311.4–754 AST 181 94 62 18.5 B 2 1 55 0.7 D
Corella E08 28.766 153.067 2.8 451–755 AST 10 4.5 63 16.4 D 0 0 N/A N/A E
Corella E09 28.734 153.020 4.3 366.5–864.5 AST 215 100.2 81 18 B 0 0 N/A N/A E
Corella E10 28.738 153.078 2.9 339.17–779.8 AST 49 40.1 44 13.6 B 0 0 N/A N/A E
Geneva 04 28.559 153.014 0.8 670.49–892.37 AST 77 82 166 8.4 B 0 0 N/A N/A E
Keerrong 1 28.725 153.270 2.3 377.5–692.5 AST 17 8.4 126 6.3 D 0 0 N/A N/A E
Middle Creek 1 28.947 152.903 6 75.6–628.8 AST 87 51.4 93 18.6 B 4 4.2 97 4.4 D
Middle Creek 1A 28.950 152.920 4 319.4–617.2 AST 120 56.6 97 18 B 0 0 N/A N/A E
Riflebird E6 28.914 153.104 6 371.7–816.2 AST 165 78 48 28.5 D 0 0 N/A N/A E
Riflebird E7 28.864 153.219 5 415.6–822.8 AST 81 40.9 38 11.8 B 0 0 N/A N/A E
South Casino 01 28.892 153.062 3.6 139.1–608.2 BGT 28 34 80 10.9 C 0 0 N/A N/A E
South Casino 03 28.894 153.062 2 88–672 AST 110 62.4 77 19.8 B 0 0 N/A N/A E
South Casino 04 28.893 153.064 2.1 83.6–675.2 AST 73 58.6 80 22.3 C 0 0 N/A N/A E
South Casino 05 28.889 153.061 2.7 447.4–648.8 AST 177 89.1 93 19.5 B 0 0 N/A N/A E
South Casino 07 28.890 153.067 1.6 179.4–651.2 AST 166 109 80 15 B 3 3 113 9.1 D
South Casino 08 28.888 153.072 1.9 179.4–487.2 AST 20 12.5 42 6.6 D 0 0 N/A N/A E
Talma 1 29.173 152.991 5.8 475–834 AST 296 86.2 75 23 C 0 0 N/A N/A E
Tullymorgan 1 29.354 153.082 1.5 269.95–716.6 AST 33 10.6 58 8.7 D 3 1.7 46 12 D

BGT, borehole geometry tool; AST, acoustic image log.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

Table 2. Quality ranking system of the World Stress Map project for the orientation of contemporary maximum horizontal stress
interpreted from four-arm caliper and acoustic image logs (Heidbach et al., 2010).

WSM quality A B C D E

Four-arm caliper tools (in a single well) Number of BOs ≥10 ≥6 ≥4 <4 No BOs or high standard
Standard deviation ≤12° ≤20° ≤25° ≤40° deviation (>40°)
Combined BO length (m) ≥300 ≥100 ≥30 <30
Acoustic image logs (in a single well) Number of BOs or DITFs ≥10 ≥6 ≥4 <4 No interpreted feature or high
Standard deviation ≤12° ≤20° ≤25° ≤40° standard deviation (>40°)
Combined BO length (m) ≥100 ≥40 ≥20 <20

Table 3. Results of circular statistics and Rayleigh test in the Clarence-Moreton Basin which is the first ‘high-quality stress province’
in the NSW.

Wells with BOs Wells with DITFs A B C D No. A–C quality Mean SHmax Azimuth SD R Conf.

27 6 1 16 4 12 21 069° 23° 0.71 99.90%

BO, borehole breakout; DITF, drilling-induced tensile fracture; R-value, length of the mean resultant vector; SD, standard deviation of stress orienta-
tion; A, B, C and D, WSM quality ranking; Conf., confidence level.

1 stress provinces according to the ASM regional ranking


scheme (Fig. 6).

DISCUSSION
Comparison of the C-M Basin SHmax with
published geomechanical-numerical models
Until recently, most of our information about the state of
stress in eastern Australian in general, and the C-M Basin
in particular, are from different published stress models
of the continent (Cloetingh & Wortel, 1985, 1986;
Coblentz et al., 1995, 1998; Reynolds et al., 2002; Hillis
& Reynolds, 2003; Reynolds et al., 2003; Zhao & M€ uller,
2003; Burbidge, 2004; Dyksterhuis et al., 2005a, b; Dyk-
sterhuis & M€ uller, 2008; M€uller et al., 2012). Almost all
Fig. 6. Plot of the Rayleigh test for the Clarence-Moreton of these models evaluated the relative importance of vari-
stress province in eastern Australia. Different lines show the ous boundary forces acting on the Indo-Australian
percentage-line according to cut-off values defined by Mardia Plate for the pattern of the SHmax orientation in the Aus-
(1972). The province is type 1 if it passes the line of 99.90%, tralian continent.
type 2 if it passes line of 99.0% up to type 5. Type 6 of stress Cloetingh & Wortel (1985, 1986) calculated the in situ
shows the province cannot pass the line of 90% due to lack of stress in the Indo-Australian Plate by using a homoge-
data or low R-value (high standard deviation). The Clarence- neous 2D elastic model. They suggested a significant east-
Moreton stress province with 21 (A–C quality) stress indicators west and northeast-southwest extension orientation for
and R-value of 0.71 passes the line of 99.9% and is defined as
north-eastern and southeastern Australia respectively.
type 1 of stress province.
Coblentz et al. (1995, 1998) investigated the origin and
first-order stress pattern on the Indo-Australian Plate by
applying different sets of boundary conditions. Their
type 2 at the 99% level, type 3 at the 97.5% level, type 4
elastic models (particularly model 4 of 1998) generally fit
at the 95% level, and type 5 at the 90% level. A type 6
the observed SHmax orientation for the western half of
stress province indicates that the null hypothesis (stress
Australia but failed to fit the observed SHmax orientation
orientations are random) cannot be rejected at the 90%
for eastern Australia. Reynolds et al. (2002, 2003) pre-
confidence level (Hillis & Reynolds, 2003). The results
sented results of 2D elastic finite element models and
show that the mean SHmax orientation of the basin is
showed that the plate boundary forces at the first-order
N069°E. In addition, the standard deviation and R-value
control the regional pattern of stress throughout the Aus-
is 23° and 0.71, respectively (Table 3), resulting in Type
tralian continent. Their best fit model demonstrated a

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 9
M. Rajabi et al.

generally northwest-southeast orientation for SHmax in The poor correlation between plate-scale models and
the eastern part of NSW including the C-M Basin. the observed stress information in the C-M Basin, like
In contrast to the previous models, which used an some major basins in northern, central and north-eastern
exclusively elastic rheology to represent the crust/ Australia (e.g. Hillis et al., 1999; Reynolds et al., 2003;
lithosphere, Burbidge (2004) applied an anelastic model Brooke-Barnett et al., 2015), raises some important issues
based on ‘thin-plate spherical finite element code’ origi- about lack of model-independent data (i.e. observed
nally developed by Bird (1999). The best fit model of Bur- SHmax orientation) and modelling strategies in the previ-
bidge (2004) predicts an east-northeast – west-southwest ous published models. Calibration of geomechanical-
SHmax orientation for most of the north-eastern part of numerical models is the key step in geomechanical-
Australia but it does not predict the SHmax orientation pat- numerical modelling where model predictions are
tern for the southern half of Western Australia, as well as adjusted to the model-independent data (Buchmann &
the Amadeus and Bowen basins in the central and the Connolly, 2007; Reiter & Heidbach, 2014). As outlined
north-eastern side of the continent. Furthermore, the pre- above, there was no stress information over the C-M
dicted stress regime by this model is not consistent with Basin prior to this study to compare with and calibrate the
observations in Western Australia (Revets et al., 2009), models. Hence, all the models were unable to predict even
Bowen and Sydney Basins of eastern Australia (Hillis & the regional trend of SHmax (east-northeast – west-south-
Reynolds, 2003). Indeed, Burbidge’s model (2004) illus- west) in this basin. Furthermore, almost all previous
trated a normal faulting stress regime for the most eastern geomechanical-numerical models of the Indo-Australian
side of Queensland and north-eastern NSW, which is in Plate simplified the 3D heterogeneities of density and
poor agreement with in situ stress data compilation of the rock properties with 2D homogenous model approaches.
ASM database where thrust and strike-slip faulting is pre- However, there are numerous studies in different regions
vailing for that region. Dyksterhuis et al. (2005a, b) and worldwide that documented the role of heterogeneity, dif-
M€ uller et al. (2012) used 2D numerical and elastic models ferent rock mechanical properties, in spatial/lateral per-
with realistic material properties for cratons, fold belts, turbation of stress (Bott & Dean, 1972; Gale et al., 1984;
basins and continental shelf and predicted generally north- Stein et al., 1989; Sonder, 1990; Zhang et al., 1994; Bell,
west-southeast SHmax orientation for the C-M Basin. 1996b; Zhang et al., 1996; Enever et al., 1999; Dykster-
The regional SHmax orientation for the C-M Basin is huis et al., 2005a; Hergert & Heidbach, 2011; King et al.,
N069°E (23°) and the analysis presented in this study 2012; Brooke-Barnett et al., 2015; Hergert et al., 2015).
highlight that none of these models can predict the orien- Localised variations have usually been explained by the
tation of the basin satisfactorily (Fig. 7). As outlined juxtaposition of different types of rocks that have differ-
above, only Burbidge’s model (2004) predicted an east- ent geomechanical properties. For example, Gale et al.
northeast – west-southwest SHmax trend on the C-M (1984) observed in situ stress variation at the Nattai Bulli
Basin but this model was unable to fit both stress orienta- colliery, NSW (Australia) and by an analytical model
tion and regime for most other parts of the continent. explained the role of basement topography on the stress

Fig. 7. The present-day maximum hori-


zontal stress (SHmax) orientation map of
the Clarence-Moreton Basin. Results of
recent geomechanical-numerical models
(Reynolds et al., 2003; M€ uller et al.,
2012) superimposed on the map (green
lines are stress trajectories from Hillis &
Reynolds, 2003). None of these models
could predict the orientation of SHmax for
this basin satisfactorily. The models pre-
dicted northwest-southeast SHmax orien-
tation for the Clarence-Moreton Basin,
while the observations show northeast-
southwest orientation for the SHmax. The
rose diagram shows the mean orientation
of SHmax based on our results in the
basin.

© 2015 The Authors


10 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

pattern of sedimentary covers. Enever et al. (1999) also including faults, fractures and igneous intrusions into
observed significant lateral perturbation of SHmax in coal sedimentary layers of the basin. Faults, fractures and
basins of eastern Australia and highlighted the impact of lithological contrasts have been highlighted as being
geological structures, particularly basement topography. important controls on the third-order SHmax orientation
More recently, Brooke-Barnett et al. (2015) investigated in several sedimentary basins worldwide (e.g. Mount &
the in situ stress pattern of the Surat Basin in southeast of Suppe, 1987; Bell et al., 1992; Zhang et al., 1994; Bell,
Queensland, approximately 350 km to the west/north- 1996a; Yale, 2003; Tingay et al., 2006; Heidbach et al.,
west of the C-M Basin, and observed a significant varia- 2007; Tingay et al., 2010) and in Australia (e.g. Gale
tion in SHmax orientation in the vicinity of basement et al., 1984; Enever et al., 1999; Hillis et al., 1999; Rey-
structures which is consistent with our results. Hence, nolds et al., 2005; Camac et al., 2006; Flottmann et al.,
modelling the Australian continent using only a simple 2014; Brooke-Barnett et al., 2015).
2D model is a possible reason that published models are The formation of the C-M and other Mesozoic basins
unable to accurately predict the stress pattern in many in this area has been described as being controlled by
eastern Australian basins, where there is significant lateral strike-slip faults located in basement rocks. Our results
variation in sediment thickness. For example in the C-M show variations in the SHmax orientation both laterally
Basin, the sediments are juxtaposed between basement (from well to well) and with depth within single wells,
rocks of the New England Orogen which have different which points to some smaller-scale sources for these vari-
geomechanical properties (Fig. 2). ations. Detailed analysis of image logs, core and drilling
Although previous published models are unable to pre- reports in most of the studied wells showed that large,
dict the regional pattern of stress within many eastern local rotations of BOs occur near faults and fractures. The
Australian basins (Hillis et al., 1999; Brooke-Barnett scale and angle of the SHmax rotations are variable and
et al., 2015 and our results), most do fit the observed depend on the density of faults and fractures. For exam-
regional SHmax throughout other sedimentary basins in ple, Fig. 8(a) shows different SHmax orientations in a
western, central and south eastern Australia, which high- 12 m interval of the borehole that are related to the pres-
lights that the long wavelengths of the Australian stress ence of faults (green sinusoids); Fig. 8(c) displays a 90°
pattern can be primarily explained in terms of the interac- rotation of SHmax orientation over only 4 m. According to
tion of plate boundary forces (Reynolds et al., 2003). our observations, the abrupt changes of SHmax are more
Besides the spatial variations of SHmax orientation, we also consistent with the presence of faults; while gradual rota-
observed numerous localised perturbations of stress ori- tions occur close to fractures and lithological contrasts
entation in most wells of the C-M Basin, particularly in such as igneous intrusions (Fig. 8d, e). However it should
the vicinity of geological features such as faults, fractures be noted that most wells studied herein show stress varia-
and lithological contrasts (discussed further in the follow- tions that are localised over short distances, such as sev-
ing section). These observations are consistent with scat- eral meters, and then return back to the regional trend.
tered SHmax orientations in the Sydney (Hillis et al., Hence, in this situation, the variability may not be cap-
1999) and Surat (Brooke-Barnett et al., 2015) basins to tured in the statistical results (Table 1).
the south and northwest respectively of the C-M Basin. In addition, Sommacal et al. (2008) interpreted several
fault maps for the basin with respect to different tectonic
events from Carboniferous up to the Late Cretaceous.
Controls on the C-M Basin stress orientation: Figure 9 shows a general map of the basin with these
the importance of faults and fractures major faults superimposed by observed SHmax orientation
The pattern of present-day stress in sedimentary basins from this study. Although the mean SHmax orientation
provides valuable information on the sources of stress throughout the C-M Basin is N069°E, the SHmax orienta-
(Bell, 1996b; Tingay et al., 2006; Heidbach et al., 2007). tions within individual wells are variable and are (sub)
Numerous studies around the world have demonstrated parallel to the faults. Fault-parallel SHmax orientations
that the SHmax pattern at small scales can be simple, sug- have been observed in several places worldwide (e.g.
gesting that the present-day stress of the basin is directly Aleksandrowski et al., 1992; Yale et al., 1994; Yale, 2003;
linked to far-field forces, or can be complex due to inter- Reynolds et al., 2005; Tingay et al., 2010). According to
action of different forces at different scales (e.g. Bell, Homberg et al. (1997) a high differential stress is neces-
1996b; Tingay et al., 2006; Heidbach et al., 2007; Rei- sary to have fault-parallel SHmax orientation. In the C-M
necker et al., 2010). Basin there is no information about the stress magnitude
As outlined above, the regional pattern of SHmax over but we observed numerous BOs in each well (Table 1).
the C-M Basin is not predicted by plate-scale published According to the Kirsch equations (Kirsch, 1898), if we
models and we discussed that the modelling strategies consider all other parameters to be equal, a high differen-
could be improved to predict the local and regional vari- tial stress is required to have large number of borehole
ability in the SHmax orientation in the basin. In addition to failures in a vertical well. In addition, high differential
this, our investigation in the C-M Basin revealed localised stresses have been reported in the Cooper-Eromanga
perturbations of SHmax orientation in most studied wells, (Reynolds et al., 2006), Gippsland (Nelson & Hillis,
particularly due to the presence of geological structures, 2005) and some part of Surat basins (Brooke-Barnett

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 11
M. Rajabi et al.

(a) (b)

(c) (d) (e)

Fig. 8. Example of borehole breakouts (BOs) rotations observed in acoustic image logs of the Clarence-Moreton Basin. In these snap-
shots, which are from different wells, BOs are outlined in blue and interpreted faults and fractures marked in green sinusoids. Rose
diagrams illustrate the orientation of maximum horizontal present-day stress (SHmax) for each stress indicators (either breakouts or
drilling-induced tensile fractures). The scale and angle of rotations are variable and are different in each case. For example (a) Several
SHmax orientations in 12 m in the vicinity of faults. (b) Abrupt rotation of BOs in a short interval. (c) 90° rotation of BOs in 4 m depth
interval. (d) and (e) Example of gradual rotations in smaller scales.

et al., 2015) in Australia, but more investigations particu- active faults that were intersected by boreholes. Barton &
larly on stress magnitude are required to assess stress ani- Moos (2010) also suggested that BO rotation might occur
sotropy in the C-M Basin. on planes of weakness, such as fractures, faults and bed-
The analysis of the present-day stress field in numerous ding planes, at or near the borehole. In contrast, it is clear
regions worldwide has demonstrated that stress rotation that fractures and faults may affect the mechanical prop-
occur near faults and fractures due to either fault slip on erties of rocks to a different degree (e.g. Nelson, 2001;
active faults or variable rock mechanical properties. Paillet Belayneh & Cosgrove, 2004; Faulkner et al., 2006; Heap
& Kim (1987), Shamir & Zoback (1992) and Barton & et al., 2010). The influence of juxtaposed stiff and weak
Zoback (1994) explained rotation of BOs due to slip on materials in perturbing the SHmax orientation has been

© 2015 The Authors


12 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

Fig. 9. The orientation of maximum


horizontal present-day stress and faults
(faults are after Sommacal et al., 2008) in
the Clarence-Moreton Basin. Most of the
SHmax orientations are consistent with
black coloured faults which according to
Sommacal et al. (2008), reactivated dur-
ing Moonie Event (95 Ma). The grey
faults are those that were present before
sedimentation of the Clarence-Moreton
Basin deposits.

(a) (b)

Fig. 10. Schematic plan view for rotation of maximum horizontal stress (SHmax) in the vicinity of geological structures (modified from
Bell, 1996b). Solid blue lines show the orientation of SHmax which is expected to swing perpendicular to mechanically stiff or hard
structures (a) and parallel to mechanically weak or soft structures (b). Structures such as cemented faults and igneous intrusions are
considered as stiff; while weak structures are open or weak faults and fractures.

shown schematically (Bell, 1996b) and from numerical


models (Pourjavad et al., 1998; Fig. 10). More recently,
IMPLICATIONS FOR HYDROCARBON
numerous studies have explained how the presence of EXPLORATION AND PRODUCTION
cracks and fractures affect elastic moduli of rocks (de- The C-M Basin has potential for both conventional and
crease Young’s modulus and increase Poisson’s ratio), and unconventional hydrocarbon resources. Implications of
hence change the stress field (e.g. Faulkner et al., 2006; SHmax orientation for conventional reservoirs have been
Heap et al., 2010; Sahara et al., 2014, 2015). Our observa- highlighted in numerous studies, particularly for wellbore
tions are consistent with Sahara et al. (2014, 2015), and stability, fluid flow in naturally fractured reservoirs and
indicate that most BO rotations occur in the vicinity of fault reactivation risk (Barton et al., 1995; Bell, 1996b;
major and minor fractures, and most likely due to changes Sibson, 1996; Tingay et al., 2005).
in rock mechanical properties. However, it should be The C-M Basin is a major coal mining region and is of
noted that we observed some BO rotations on near-hori- significant interest for coal seam gas production in eastern
zontal weak planes (Fig. 8b) which indicates shear slip- Australia. Coal seams are self-sourcing reservoirs with
page on the pre-existing planes. low matrix permeability (Ayers, 2002), and thus the

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 13
M. Rajabi et al.

natural fracture system of coal (coal cleats) plays a critical Ayers, 2002; Massarotto et al., 2003; Olsen et al., 2003;
role in hydrocarbon production (Laubach et al., 1998; Titheridge, 2014). Numerous studies have shown that
Ayers, 2002; Olsen et al., 2003; Bell, 2006). Coal cleats fractures that are aligned parallel or within 30° to SHmax
are opening-mode fractures, usually sub-vertical, and orientation enhance fluid flow (e.g. Barton et al., 1995;
form systematically in two perpendicular sets, namely Sibson, 1996; Hillis, 1997). In addition, Heffer & Lean
face cleat and butt cleat (Laubach et al., 1998; Moore, (1991) and Heffer et al. (1997) observed the strong rela-
2012). The face cleat is more continuous and connected, tionship between SHmax and preferred flow axis in flood-
while the butt cleat is discontinuous and terminates at face ing operations of petroleum reservoirs. Hence, the angle
cleats (Laubach et al., 1998; Moore, 2012). Therefore, between face cleats and orientation of SHmax is likely to be
face cleats are more significant for fluid flow, and result in an important control in the permeability of coal seam gas
permeability anisotropy in coals (Koenig, 1989, 1991; reservoirs (Titheridge, 2014). However, due to borehole

Fig. 11. Optimum orientations of hydraulically conductive cleats in relation to the maximum horizontal stress (SHmax) in different
stress regimes. Blue and grey lines indicate open and close cleats respectively. In these schematic figures we simply show two end-
members for the relationship between SHmax and face cleats (i.e. SHmax parallel and perpendicular to face cleats). The borehole stability
in different stress regimes has also highlighted to show that it is not always possible to deviate the well to intersect the majority of open
cleats (modified from Hillis & Williams (1993) and Hillis (1997)). Note that there could be more complex examples either due to differ-
ent angle between SHmax and cleats or shear failure and stress anisotropy.

© 2015 The Authors


14 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

stability problems (Hillis & Williams, 1993) it is not Canada. Their studied borehole penetrated twice into a
always possible to drill wells to intersect as many sand unit which was displaced vertically by a thrust fault.
hydraulically conductive cleats as possible. As outlined Microseismic and fracture engineering data revealed two
above, there is no detailed information about the magni- different hydraulic fracture geometries in the same sandy
tude of stresses (and stress regime) in this region and a unit at approximately 50 m depth intervals, above and
thrust or strike-slip regime can be expected based on the below the fault.
ASM database. In Fig. 11, we schematically describe Hence, localised rotation of horizontal stresses by
borehole stability and hydraulically conductive coal cleats geological features can make complex fracture geometry
in different stress regimes. which is hard to be filled with proppant. In geological
The permeability of coal seam gas reservoirs is gener- settings such as the C-M Basin, hydraulic fracture
ally low (Ayers, 2002; Olsen et al., 2007; Palmer, 2010), designs based on regional stress orientations can signifi-
and Palmer (2010) highlighted this issue as the ‘Achilles cantly affect stimulations (causing tortuosity of induced
heel’ for commercial extraction of gas from these reser- fractures; early screenouts and finally poor production
voirs. Hence, the current focus on coal seam gas reservoirs results). Therefore, it may not be appropriate to design
highlights the key challenge of finding reliable methods to fracture experiments on the SHmax orientation observed
enhance extraction of gas from these reservoirs (Olsen from just a few wells in a basin or field. Rather, it is
et al., 2003, 2007; Palmer, 2010). Hydraulic fracturing is important to undertake detailed statistical analysis of
currently the most effective method to commercially pro- in situ stress from all available data to adequately con-
duce these resources, and the propagation direction of strain the regional and local-scale perturbations in the
hydraulic fractures is primarily controlled by the present- stress field. Furthermore, current development and
day SHmax orientation (Olsen et al., 2007; Seidle, 2011). hydraulic fracture stimulations of unconventional reser-
Thus, the control of local sources of stress on the stress voirs worldwide have highlighted that stress anisotropy,
pattern in the C-M Basin, particularly the rotation of presence of natural fractures and brittleness of reservoirs
stresses near faults and fractures, has significant implica- rocks significantly affect the geometry of induced
tions for fracture stimulation and extraction of gas. For hydraulic fractures (Warpinski & Teufel, 1987; Fisher
example, Maxwell et al. (2009) showed a significant re- et al., 2002; Olsen et al., 2003; Fisher et al., 2004;
orientation of hydraulic fractures due to presence of a Cipolla et al., 2008; Fisher & Warpinski, 2012; Leem
thrust fault that was cross-cut by a borehole in western et al., 2014). Figure 12 schematically describes the

Fig. 12. Top) the pattern of induced hydraulic fractures due to presence of natural fractures and stress anisotropy (modified from
Fisher et al., 2002; Fisher & Warpinski, 2012; Leem et al., 2014). In the bottom row different pattern of hydraulic fractures are shown
in coal reservoirs due to stress anisotropy and the orientation of maximum horizontal stress (SHmax) with respect to cleats (i.e. SHmax is
parallel and perpendicular to the face cleats). (a) and (b) show planar hydraulic fractures which form parallel to the orientation of SHmax
in high stress anisotropy (i.e. high differential stress) while (c) and (d) show complex fractures due to low stress anisotropy and pres-
ence of natural fractures or cleats.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 15
M. Rajabi et al.

importance of differential stress and presence of coal REFERENCES


cleats in the geometry and pattern of hydraulic fractures
AADNOY, B. S. (1990) Inversion technique to determine the in-situ
in coal seam gas reservoirs.
stress field from fracturing data. J. Petrol. Sci. Eng., 4, 127–141.
AADNOY, B. S. & BELL, J. S. (1998) Classification of drilling-
induced fractures and their relationship to in-situ stress direc-
CONCLUSIONS tions. Log Anal., 39, 27–42.
ALEKSANDROWSKI, P.E., INDERHAUG, O.H.E. & KNAPSTAD, B.E.
This study presents the first determination of the pre- (1992) Tectonic structures and wellbore breakout orientation.
sent-day stress orientation in north-eastern of NSW. The The 33th U.S. Symposium on Rock Mechanics (USRMS).
analysis of 25 image logs and 2 oriented-caliper logs American Rock Mechanics Association, Santa Fe, New Mex-
demonstrates a mean SHmax orientation of N069°E (23°) ico, 29–37.
in the Clarence-Morton Basin, but also reveals that SHmax AYERS, W. B. (2002) Coalbed Gas systems, resources, and pro-
duction and a review of contrasting cases from the San Juan
is locally perturbed, both laterally and in depth, particu-
and powder River Basins. AAPG Bull., 86, 1853–1890.
larly in proximity to faults, fractures and geomechanical
BARTON, C. & MOOS, D. (2010) Geomechanical wellbore imag-
contrasts of rock properties. The localised perturbation of ing: key to managing the asset life cycle. In: Dipmeter and
stress is of particular importance for safe and sustainable Borehole Image Log Technology (Ed. by M. P€oppelreiter, C.
exploration and long-term production of gas in the region. Garcia-Carballido & M. Kraaijveld), 81–112. American Associ-
The east-northeast – west-southwest SHmax orientation ation of Petroleum Geologists Memoir.
observed in the Clarence-Morton Basin is not predicted BARTON, C. & ZOBACK, M. (1994) Stress perturbations associ-
by existing geomechanical-numerical models of the Aus- ated with active faults penetrated by boreholes: possible evi-
tralian stress field. The inability of numerical models to dence for near-complete stress drop and a new technique for
predict the observed stress orientations in the basin, when stress magnitude measurement. J. Geophys. Res. Solid Earth,
combined with the numerous local stress perturbations, 99(B5), 9373–9390.
BARTON, C., ZOBACK, M. & MOOS, D. (1995) Fluid flow along
suggests that second and third-order stress sources proba-
potentially active faults in crystalline rock. Geology, 23, 683–
bly play a greater role in the Australian stress pattern than
686.
previously recognised. Previously published models are BELAYNEH, M. & COSGROVE, J. W. (2004) Fracture-pattern varia-
2D (except Zhao & M€ uller (2003) which is 3D but has tions around a major fold and their implications regarding
only two element layers over 100 km thickness) and may fracture prediction using limited data: an example from the
not replicate the real lithospheric behaviour of the conti- Bristol Channel Basin. Geol. Soc. Lond. Spec. Publ., 231, 89–
nent and hence the state of stress. The stress orientation 102.
in the basin, as well as that observed in some major basins BELL, J. S. (1990) Investigating stress regimes in sedimentary
in northern, central and north-eastern Australia, indicates basins using information from oil industry wireline logs and
that, local perturbations due to presence of different geo- drilling records. Geol. Soc. Lond. Spec. Publ., 48, 305–325.
logical features are significant and can lead to substantial BELL, J. S. (1996a) Petro Geoscience 1. In situ stresses in sedi-
mentary rocks (Part 1): measurement Techniques. Geosci.
changes in the SHmax orientation, particularly at the basin
Can., 23, 85–100.
scale. Hence, to evaluate the detailed pattern of stress in
BELL, J. S. (1996b) Petro Geoscience 2. In situ stresses in sedi-
the continent both improvement in modelling approaches mentary rocks (Part 2): applications of stress measurements.
and stress observations are required. Geosci. Can., 23, 135–153.
BELL, J. S. (2006) In-situ stress and coal bed methane potential
in Western Canada. Bull. Can. Pet. Geol., 54, 197–220.
BELL, J. S. & GOUGH, D. I. (1979) Northeast-Southwest com-
ACKNOWLEDGEMENTS pressive stress in Alberta evidence from oil wells. Earth Pla-
The authors gratefully thank the New South Wales net. Sci. Lett., 45, 475–482.
Department of Resources and Energy for providing the BELL, J. S., CAILLET, G. & ADAMS, J. (1992) Attempts to detect
data of this study. This research was funded through the open fractures and non-sealing faults with dipmeter logs.
Australian Research Council (DP120103849) and the Geol. Soc. Lond. Spec. Publ., 65, 211–220.
BIRD, P. (1999) Thin-plate and thin-shell finite-element pro-
Australian Society of Exploration Geophysicists Research
grams for forward dynamic modeling of plate deformation
Foundation (RF13P02). The authors also specially thank and faulting. Comput. Geosci., 25, 383–394.
Judith Sippel, Scott Mildren and an anonymous reviewer BOTT, M. H. P. & DEAN, D. S. (1972) Stress systems at young
of this article for their time and helpful comments. Ikon continental margins. Nat. Phys. Sci., 235, 23–25.
Science of Adelaide (formerly JRS Petroleum Research) BROOKE-BARNETT, S., FLOTTMANN, T., PAUL, P. K., BUSETTI,
is thanked for the provision of the JRS suite.This paper S., HENNINGS, P., REID, R. & ROSENBAUM, G. (2015) Influence
forms TRaX record # 340. of basement structures on in situ stresses over the Surat Basin,
Southeast Queensland. Journal of Geophysical Research: Solid
Earth, 120, 4946–4965.
BUCHMANN, T. J. & CONNOLLY, P. (2007) Contemporary kine-
CONFLICT OF INTEREST matics of the upper Rhine Graben: a 3D finite element
approach. Global Planet. Change, 58, 287–309.
No conflict of interest declared.

© 2015 The Authors


16 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

BURBIDGE, D. R. (2004) Thin plate neotectonic models of the FLOTTMANN, T., BROOKE-BARNETT, S., TRUBSHAW, R., NAIDU,
Australian plate. J. Geophys. Res. Solid Earth, 109, B10405. S.-K., KIRK-BURNNAND, E., PAUL, P., BUSETTI, S. & HEN-
CAMAC, B., HUNT, S. & BOULT, P. (2006) Local rotations in NINGS, P. (2014) Influence of in-situ stresses on hydraulic
borehole breakouts-observed and modeled stress field rota- fracture stimulations in coals of the Surat Basin, Southeast
tions and their implications for the petroleum industry. Int. J. Queensland – Australia. SPE Unconventional Resources
Geomech., 6, 399–410. Conference and Exhibition-Asia Pacific, SPE-167064-MS.
CIPOLLA, C. L., WARPINSKI, N. R. & MAYERHOFER, M. J. (2008) Brisbane, Australia, 14 pages.
Hydraulic Fracture Complexity: Diagnosis, Remediation, and GALE, W. J., ENEVER, J. R., BLACKWOOD, R. L. & MCKAY, J.
Explotation. SPE Asia Pacific Oil and Gas Conference and (1984) An Investigation of the Effect of a Fault /Monocline
Exhibition, Society of Petroleum Engineers. Perth, Australia, Structure on the In-situ Stress Field and Mining Conditions
24 pages. at Nattai Bulli Colliery NSW, Australia, CSIRO Australia,
CLOETINGH, S. & WORTEL, R. (1985) Regional stress field of the Division of Geomechanics, Geomechanics of Coal Mining
Indian plate. Geophys. Res. Lett., 12, 77–80. Report No. 48.
CLOETINGH, S. & WORTEL, R. (1986) Stress in the Indo-Austra- HAIMSON, B. (2007) Micromechanisms of Borehole instability
lian plate. Tectonophysics, 132, 49–67. leading to breakouts in rocks. Int. J. Rock Mech. Min. Sci., 44,
COBLENTZ, D. D. & RICHARDSON, R. M. (1995) Statistical trends 157–173.
in the intraplate stress field. J. Geophys. Res. Solid Earth, 100, HAIMSON, B. & HERRICK, C. G. (1986) Borehole Breakouts - a
20245–20255. New Tool for Estimating in Situ Stress? ISRM International
COBLENTZ, D. D., SANDIFORD, M., RICHARDSON, R. M., ZHOU, Symposium, International Society for Rock Mechanics. Stock-
S. & HILLIS, R. R. (1995) The origins of the intraplate stress holm, Sweden.
field in continental Australia. Earth Planet. Sci. Lett., 133, HEAP, M. J., FAULKNER, D. R., MEREDITH, P. G. & VIN-
299–309. CIGUERRA, S. (2010) Elastic Moduli evolution and accompany-
COBLENTZ, D. D., ZHOU, S., HILLIS, R. R., RICHARDSON, R. M. ing stress changes with Increasing crack damage: implications
& SANDIFORD, M. (1998) Topography, boundary forces, and for stress changes around fault zones and volcanoes during
the Indo-Australian intraplate stress field. J. Geophys. Res. deformation. Geophys. J. Int., 183, 225–236.
Solid Earth, 103, 919–931. HEFFER, K. J. & LEAN, J. C. (1991) Earth stress orientation - a
DAVIS, J. C. (2002) Statistics and Data Analysis in Geology. 3rd control on, and guide to, flooding directionality in a majority
edn, Wiley, New York. p. 656. of reservoirs. In: Reservoir Characterization III (Ed. by B.
DAY, R. W., BUBENDORFER, P. J. & PINDER, B. J. (2008) Petro- Linville), pp. 799–822. PennWell Books, Tulsa.
leum Potential of the Easternmost Surat Basin in Queensland. HEFFER, K. J., FOX, R. J., MCGILL, C. A. & KOUTSABE-
Proceedings of the Petroleum Exploration Society of Aus- LOULIS, N. C. (1997) Novel techniques show links between
tralia, Eastern Australasian Basins Symposium III, Sydney, reservoir flow directionality, earth stress, fault structure
PESA, Melbourne, 191–199. and geomechanical changes in mature waterfloods. SPE J.,
DYKSTERHUIS, S. & MULLER € , R. D. (2008) Cause and evolu- 2, 91–98.
tion of intraplate orogeny in Australia. Geology, 36, 495– HEIDBACH, O., REINECKER, J., TINGAY, M., MULLER € , B., SPER-
498. NER, B., FUCHS, K. & WENZEL, F. (2007) Plate boundary
DYKSTERHUIS, S., ALBERT, R. A. & MULLER € , R. D. (2005a) forces are not enough: second- and third-order stress patterns
Finite-element modelling of contemporary and Palaeo-Intra- highlighted in the world stress map database. Tectonics, 26,
plate stress using AbaqusTM. Comput. Geosci., 31, 297–307. TC6014.

DYKSTERHUIS, S., MULLER , R. D. & ALBERT, R. A. (2005b) Pale- HEIDBACH, O., TINGAY, M., BARTH, A., REINECKER, J., KURFEß,
ostress field evolution of the Australian continent since the €
D. & MULLER , B. (2010) Global crustal stress pattern based
Eocene. J. Geophys. Res. Solid Earth, 110, B05102. on the world stress map database release 2008. Tectonophysics,
ENEVER, J. R., GALE, W. & FABJANCZVK, M. (1999) A Study of 482, 3–15.
the Variability of the Horizontal Stress Field in a Sedimentary HERGERT, T. & HEIDBACH, O. (2011) Geomechanical model of
Basin. 9th ISRM Congress, International Society for Rock the Marmara Sea Region – II. 3-D Contemporary background
Mechanics. Paris, France, 1257–1260. stress field. Geophys. J. Int., 185, 1090–1102.
FAULKNER, D. R., MITCHELL, T. M., HEALY, D. & HEAP, M. J. HERGERT, T., HEIDBACH, O., REITER, K., GIGER, S. B. &
(2006) Slip on ‘weak’ faults by the rotation of regional stress MARSCHALL, P. (2015) Stress field sensitivity analysis in a sed-
in the fracture damage zone. Nature, 444, 922–925. imentary sequence of the alpine Foreland, Northern Switzer-
FISHER, M. K. & WARPINSKI, N. R. (2012) Hydraulic-fracture- land. Solid Earth, 6, 533–552.
height growth: real data. SPE Prod. Oper., 27, 8–19. HILLIS, R. R. (1997) Does the in situ stress field control the ori-
FISHER, M. K., WRIGHT, C. A., DAVIDSON, B. M., GOODWIN, A. entation of open natural fractures in sub-surface reservoirs?
K., FIELDER, E. O., BUCKLER, W. S. & STEINSBERGER, N. P. Explor. Geophys., 28, 80–87.
(2002) Integrating Fracture Mapping Technologies to Opti- HILLIS, R. R. & REYNOLDS, S. D. (2000) The Australian stress
mize Stimulations in the Barnett Shale. SPE Annual Techni- map. J. Geol. Soc. London, 157, 915–921.
cal Conference and Exhibition, Society of Petroleum HILLIS, R. R. & REYNOLDS, S. D. (2003) In situ stress of Aus-
Engineers. San Antonio, Texas, 7 pages. tralia. Geol. Soc. Am. Spec. Pap., 372, 49–58.
FISHER, M. K., HEINZE, J. R., HARRIS, C. D., DAVIDSON, B. M., HILLIS, R. R. & WILLIAMS, A. (1993) The stress field of the
WRIGHT, C. A. & DUNN, K. P. (2004) Optimizing Horizontal North West Shelf and Wellbore stability. APPEA J., 33, 373–
Completion Techniques in the Barnett Shale Using Micro- 373.
seismic Fracture Mapping. SPE Annual Technical Confer- HILLIS, R. R., ENEVER, J. R. & REYNOLDS, S. D. (1999) In situ
ence and Exhibition, Society of Petroleum Engineers. stress field of Eastern Australia. Aust. J. Earth Sci., 46, 813–
Houston, Texas, 11 pages. 825.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 17
M. Rajabi et al.

HILLIS, R. R., SANDIFORD, M., REYNOLDS, S. D. & QUIGLEY, M. NELSON, R. A. (2001) Geologic Analysis of Naturally Fractured
C. (2008) Present-day stresses, seismicity and Neogene-to- Reservoirs, p. 352. Gulf Professional Publishing, Houston, TX.
recent tectonics of Australia’s ‘Passive’ Margins: intraplate NELSON, E. J. & HILLIS, R. R. (2005) In situ stresses of the West
deformation controlled by plate boundary forces. Geol. Soc. Tuna Area, Gippsland Basin. Aust. J. Earth Sci., 52, 299–313.
Lond. Spec. Publ., 306, 71–90. O’BRIEN, P. E., KORSCH, R. J., WELLS, A. T., SEXTON, M. J. &
HOMBERG, C., HU, J. C., ANGELIER, J., BERGERAT, F. & WAKE-DYSTER, K. D. (1994) Structure and tectonics of the
LACOMBE, O. (1997) Characterization of stress perturbations Clarence-Moreton Basin. Bull.-Bureau Min. Resour. Geol.
near major fault zones: insights from 2-D distinct-element Geophys., Australia, 232, 117–188.
numerical modelling and field studies (Jura Mountains). J. OLSEN, T. N., BRENIZE, G. & FRENZEL, T. (2003) Improvement
Struct. Geol., 19, 703–718. Processes for Coalbed Natural Gas Completion and Stimula-
INGRAM, E. T., ROBINSON, V. A. & FACER, R. A. (1996) Petro- tion. SPE Annual Technical Conference and Exhibition,
leum Prospectively of the Clarence-Morton Basin in New Society of Petroleum Engineers. Denver, Colorado, 16 pages.
South Wales. New South Wales Department of Mineral OLSEN, T. N., BRATTON, T. R., TANNER, K. V., DONALD, A. &
Resources, Petroleum Bulletin 3, Sydney, 133 pp. KOEPSELL, R. (2007) Application of Indirect Fracturing for
JELL, P. A., MCKELLER, J. L. & DRAPER, J. J. (2013) Clarence- Efficient Stimulation of Coalbed Methane. Rocky Mountain
moreton Basin. In: Geology of Queensland (Ed. by P. A. Jell), Oil & Gas Technology Symposium, Society of Petroleum Engi-
pp. 542–546. Geological Survey of Queensland, Brisbane, neers. Denver, Colorado, 10 pages.
Queensland. PAILLET, F. L. & KIM, K. (1987) Character and distribution of
KING, R., BACKE, G., TINGAY, M., HILLIS, R. & MILDREN, S. borehole breakouts and their relationship to in situ stresses in
(2012) Stress deflections around salt diapirs in the Gulf of deep Columbia River Basalts. J. Geophys. Res. Solid Earth, 92,
Mexico. Geol. Soc. Lond. Spec. Publ., 367, 141–153. 6223–6234.
KIRSCH, E. G. (1898) Die Theorie Der Elastizit€at Und Die Bed- PALMER, I. (2010) Coalbed methane completions: a world view.
€rfnisse Der Festigkeitslehre. Zeitschrift des Vereines Deutscher
u Int. J. Coal Geol., 82, 184–195.
Ingenieure, 42, 797–807. PLUMB, R. A. & HICKMAN, S. H. (1985) Stress-induced borehole
KOENIG, R. A. (1989) Hydrologic characterization of coal seams elongation: a comparison between the Four-Arm dipmeter
for optimal dewatering and methane drainage. Q. R. Methane and the borehole televiewer in the Auburn Geothermal Well.
Coal Seams Technol., 35, 30–31. J. Geophys. Res. Solid Earth, 90, 5513–5521.
KOENIG, R. A. (1991) Directional permeability in Coal. In: Gas POURJAVAD, M., BELL, J. S. & BRATLI, R. K. (1998) Stress Trajec-
in Australian Coal (Ed. W. J. Bam-berry & A. M. Depers). tories in the Neighbourhood of Fault Zones. SPE/ISRM Rock
Geological Society of Australia, Symposium Proceedings 2, 65– Mechanics in Petroleum Engineering. Society of Petroleum
75. Engineers, Trondheim, Norway.
LAUBACH, S. E., MARRETT, R. A., OLSON, J. E. & SCOTT, A. R. POWELL, T. G., O’BRIEN, P. E. & WELLS, A. T. (1993) Petroleum
(1998) Characteristics and origins of coal cleat: a review. Int. prospectivity of the Clarence-Moreton Basin, Eastern Aus-
J. Coal Geol., 35, 175–207. tralia: a geochemical perspective. Aust. J. Earth Sci., 40, 31–44.
LEE, M. & HAIMSON, B. (1993) Laboratory study of borehole REINECKER, J., TINGAY, M. & MULLER, B. (2003) Borehole
breakouts in Lac Du Bonnet Granite: a case of extensile fail- Breakout Analysis from Four-arm Caliper Logs. World Stress
ure mechanism. Int. J. Rock Mech. Min. Sci. Geomech. Map Project Guidelines. 5 pp. Available on: www.world-
Abstracts, 30, 1039–1045. stress-map.org.
LEEM, J., DAY, R., LATIMER, C., LAKANI, R. & REYNA, J. (2014) REINECKER, J., TINGAY, M., MULLER, B. & HEIDBACH, O. (2010)
Geomechanics in Optimal Multi-Stage Hydraulic Fracturing Present-day stress orientation in the Molasse Basin. Tectono-
Design for Resource Shale and Tight Reservoirs. EAGE physics, 482, 129–138.
Workshop on Geomechanics in the Oil and Gas Industry. REITER, K. & HEIDBACH, O. (2014) 3-D Geomechanical-numeri-
Dubai, UAE, 5 pages. cal model of the contemporary crustal stress State in the
MARDIA, K. V. (1972) Statistics of Directional Data / K.V. Mar- Alberta Basin (Canada). Solid Earth, 5, 1123–1149.
dia. Academic Press, London; New York. REVETS, S. A., KEEP, M. & KENNETT, B. L. N. (2009) NW Aus-
MARTIN, A. R. & SAXBY, J. D. (1982) Geology, source rocks and tralian intraplate seismicity and stress regime. J. Geophys.
hydrocarbon generation in the Clarence-Moreton Basin. Aust. Res. Solid Earth, 114, B10305.
Petrol. Prod. Exp. Assoc. J., 22, 5–16. REYNOLDS, S. D., COBLENTZ, D. D. & HILLIS, R. R. (2002) Tec-
MASSAROTTO, P., RUDOLPH, V., GOLDING, S. D. & IYER, R. tonic forces controlling the regional intraplate stress field in
(2003) The Effect of Directional Net Stresses on the Direc- continental Australia: results from new finite element model-
tional Permeability of Coal. International Coalbed Methane ing. J. Geophys. Res. Solid Earth, 107, ETG 1-1–ETG 1-15.
Symposium. Tuscaloosa, Alabama, 13 pages. REYNOLDS, S. D., COBLENTZ, D. D. & HILLIS, R. R. (2003) Influ-
MAXWELL, S. C., ZIMMER, U., GUSEK, R. W. & QUIRK, D. J. ences of plate-boundary forces on the regional intraplate
(2009) Evidence of a horizontal hydraulic fracture from stress stress field of continental Australia. Geol. Soc. Am. Spec. Pap.,
rotations across a thrust fault. SPE Prod. Oper., 24, 312–319. 372, 59–70.
MOORE, T. A. (2012) Coalbed methane: a review. Int. J. Coal REYNOLDS, S.D., MILDREN, S.D., HILLIS, R.R., MEYER, J.J. &
Geol., 101, 36–81. FLOTTMANN, T. (2005) Maximum horizontal stress orienta-
MOUNT, V. S. & SUPPE, J. (1987) State of stress near the San tions in the Cooper Basin, Australia: implications for plate-
Andreas Fault: implications for Wrench tectonics. Geology, scale tectonics and local stress sources. Geophys. J. Int., 160,
15, 1143–1146. 331–343.

MULLER , R. D., DYKSTERHUIS, S. & REY, P. (2012) Australian RICHARDSON, R. M. (1992) Ridge forces, absolute plate motions,
Paleo-stress fields and tectonic reactivation over the past and the intraplate stress field. J. Geophys. Res. Solid Earth, 97,
100 Ma. Aust. J. Earth Sci., 59, 13–28. 11739–11748.

© 2015 The Authors


18 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Localised perturbations of the present-day stress

SAHARA, D. P., SCHOENBALL, M., KOHL, T. & MULLER, B. I. R. tute of Mining and Metallurgy & Mine Managers Association of
(2014) Impact of fracture networks on borehole breakout Australia, 2014, 8–17.
heterogeneities in crystalline rock. Int. J. Rock Mech. Min. WARPINSKI, N. R. & TEUFEL, L. W. (1987) Influence of geologic
Sci., 71, 301–309. discontinuities on hydraulic fracture propagation (Includes
SAHARA, D., SCHOENBALL, M., KOHL, T. & MUELLER, B. (2015) Associated Papers 17011 and 17074). J. Petrol. Technol., 39,
Analysis of stress heterogeneities in fractured crystalline 209–220.
reservoir. Proceedings World Geothermal Congress 2015. WELLS, A. T. & O’BRIEN, P. E. (1994) Lithostratigraphic frame-
Melbourne, Australia, 9 pages. work of the Clarence-Moreton Basin. In: Geology and petro-
SANDIFORD, M., WALLACE, M. & COBLENTZ, D. D. (2004) Origin leum potential of the Clarence-Moreton Basin, New South Wales
of the in situ stress field in South-Eastern Australia. Basin and Queensland. (Ed. by A. T. Wells & O. P.E) Australian
Res., 16, 325–338. Geological Survey Organisation, Bulletin 241, 4–47.
SEIDLE, J. (2011) Fundamentals of Coalbed Methane Reservoir WILLIS, I. L. (1994) Stratigraphic implications of regional recon-
Engineering. PennWell, Tulsa, Oklahoma. naissance observations in the Southern Clarence-Moreton
SHAMIR, G. & ZOBACK, M. D. (1992) Stress orientation profile to Basin, New South Wales. In: Geology and Petroleum Potential
3.5 km depth near the San Andreas Fault at Cajon Pass, Cali- of the Clarence-Moreton Basin, New South Wales and Queens-
fornia. J. Geophys. Res. Solid Earth, 97, 5059–5080. land. (Ed. by A. T. Wells & P. E. O’Brien). Australian Geo-
SHAW, S. E. & FLOOD, R. H. (1981) The New England Batho- logical Survey Organisation, Bulletin 241, 48–71.
lith, Eastern Australia: geochemical variations in time and YALE, D. P. (2003) Fault and stress magnitude controls on varia-
space. J. Geophys. Res. Solid Earth, 86, 10530–10544. tions in the orientation of in situ stress. Geol. Soc. Lond. Spec.
SIBSON, R. H. (1996) Structural permeability of Fluid-driven Publ., 209, 55–64.
Fault-fracture Meshes. J. Struct. Geol., 18, 1031–1042. YALE, D. P., RODRIGUEZ, J. M., MERCER, Z. B. & BLAISDELL, D.
SOMMACAL, S., PRYER, L., BLEVIN, J., CHAPMAN, J. & CATHRO, W. (1994) In situ stress orientation and the effects of local
D. (2008) Clarence-Moreton Seebasetm and Structural Gis structure - Scott field, North Sea. Rock Mechanics in Petro-
Project (by Frog Tech Pty Ltd), Nsw Dpi. leum Engineering, pp. 945–952. Balkema, Rotterdam.
SONDER, L. J. (1990) Effects of density contrasts on the orienta- ZAJAC, B. J. & STOCK, J. M. (1997) Using borehole breakouts to
tion of stresses in the lithosphere: relation to principal stress constrain the complete stress tensor: results from the Sijan
directions in the transverse ranges, California. Tectonics, 9, deep drilling project and offshore Santa Maria Basin, Califor-
761–771. nia. J. Geophys. Res. Solid Earth, 102, 10083–10100.
STEIN, S., CLOETINGH, S., SLEEP, N. & WORTEL, R. (1989) Pas- ZHANG, Y.-Z., DUSSEAULT, M.B. & YASSIR, N.A. (1994) Effects
sive margin earthquakes, stresses and rheology. In: Earth- of rock anisotropy and heterogeneity on stress distributions at
quakes at North-Atlantic Passive Margins: Neotectonics and selected sites in North America. Eng. Geol., 37, 181–197.
Postglacial Rebound (Ed. by, Nato AsiSeries), 266, 231–259. ZHANG, Y., SCHEIBNER, E., ORD, A. & HOBBS, B. (1996) Numeri-
Springer, the Netherlands. cal modelling of crustal stresses in the Eastern Australian pas-
STEWART, J. R. & ADLER, J. D. (1995) New South Wales Petro- sive margin. Aust. J. Earth Sci., 43, 161–175.
leum Potential. New South Wales Department of Mineral €
ZHAO, S. & MULLER , R. D. (2003) Three-dimensional finite-ele-
Resources, Sydney, NSW. ment modelling of the tectonic stress field in continental Aus-

TINGAY, M., MULLER , B., REINECKER, J., HEIDBACH, O., WEN- tralia. Geol. Soc. Am. Spec. Pap., 372, 71–89.
ZEL, F. & FLECKENSTEIN, P. (2005) Understanding tectonic ZOBACK, M. L. (1992) First- and second-order patterns of stress
stress in the oil patch: the world stress map project. Lead. in the lithosphere: the world stress map project. J. Geophys.
Edge, 24, 1276–1282. Res. Solid Earth, 97, 11703–11728.
TINGAY, M., MULLER, B., REINECKER, J. & HEIDBACH, O. (2006) ZOBACK, M. L., ZOBACK, M. D., ADAMS, J., ASSUMPCAO, M.,
State and Origin of the Present-Day Stress Field in Sedimen- BELL, S., BERGMAN, E. A., BLUMLING, P., BRERETON, N. R.,
tary Basins: New Results from the World Stress Map Project. DENHAM, D., DING, J., FUCHS, K., GAY, N., GREGERSEN, S.,
The 41st U.S. Symposium on Rock Mechanics (USRMS), GUPTA, H. K., GVISHIANI, A., JACOB, K., KLEIN, R., KNOLL,
American Rock Mechanics Association. Golden, Colorado. P., MAGEE, M., MERCIER, J. L., MULLER, B. C., PAQUIN, C.,
TINGAY, M., MORLEY, C., HILLIS, R. R. & MEYER, J. (2010) Pre- RAJENDRAN, K., STEPHANSSON, O., SUAREZ, G., SUTER, M.,
sent-day stress orientation in Thailand’s Basins. J. Struct. UDIAS, A., XU, Z. H. & ZHIZHIN, M. (1989) Global patterns of
Geol., 32, 235–248. tectonic stress. Nature, 341, 291–298.
TITHERIDGE, D. (2014) Drilling Induced Fractures in Coal Core
from Vertical Exploration Well: A Method to Determine
Cleat Azimuth, and the Angle between Cleat and Maximum Manuscript received 13 May 2015; In revised form 3
Horizontal Stress, and its Application. 14th Coal Operators’ November 2015; Manuscript accepted 24 November 2015.
Conference, University of Wollongong, The Australasian Insti-

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 19

View publication stats

You might also like