Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

In situ stress distribution and mechanical stratigraphy in the Bowen

and Surat basins, Queensland, Australia


EMMA TAVENER1, THOMAS FLOTTMANN2* & SAM BROOKE-BARNETT2
1
Santos Ltd, Santos Place, 32 Turbot Street, Qld 4000, Australia
2
Origin Energy, 339 Coronation Drive, Qld 4064, Australia
*Correspondence: thomas.flottman@originenergy.com.au

Abstract: We present regional in situ stress analyses based on publicly available log and pressure
data from coal seam gas developments in the Permian Bowen basin, Australia. Together with
earlier data from the eastern part of the Jurassic Surat basin, our results show a broad, but system-
atic, rotation of SHmax azimuths in this part of eastern Australia as well as systematic changes in
stress state with depth. Overall, the geomechanical state of the region appears to reflect the inter-
play between basin-controlling structures and a complex far-field stress regime. At the reservoir
level, within and between Permian coal seams, this stress complexity is reflected in highly variable
stress states both vertically and laterally. Stress data, including direct pressure measurements and
observations of borehole failure in image logs, have been used to calibrate sonic-derived one-
dimensional wellbore stress models that consistently exhibit a change in tectonic stress regime
with depth. Shallow depths (,600 m) are characterized by a reverse-thrust stress regime and
deeper levels are characterized by a strike-slip regime. Changes in the stress state with depth
influence the mechanical stratigraphy of rocks with widely contrasting mechanical attributes
(coals and clastic sediments). Our results highlight the interdependency between regional tectonic,
local structural and detailed rheological influences on the well scale geomechanical conditions that
have to be taken into consideration in drilling and completion designs.

Supplementary material: Database of additional wells with image log data are available at
https://doi.org/10.6084/m9.figshare.c.3785849

The Australian continent is characterized by signifi- 2009a). Between c. 265 and 230 Ma, the Bowen
cant variability in the magnitude and orientation basin and the New England Orogen were subjected
of in situ stresses (Coblentz et al. 1995, 1998; Hillis to contractional and strike-slip deformation known
et al. 1999; Hillis & Reynolds 2000, 2003; Reynolds as the Hunter Bowen Orogeny (Holcombe et al.
et al. 2002, 2003) at both the continental and 1997b; Korsch et al. 2009b). This deformation
regional scale. This is well documented from oil- initiated basin-bounding fault systems (Fig. 2).
field data, particularly in the eastern –central The Bowen basin is broadly characterized by
interior basins (Reynolds et al. 2005; Nelson et al. two north–south-trending depocentres, the Denison
2007). The in situ stress distribution in the Austra- and Taroom troughs (Fig. 2), with internal half-
lian plate is controlled by plate boundary forces graben structures that initiated during the Early
acting on the Australian plate (Fig. 1). The key Permian. The early Permian basin-fill is dominated
plate tectonic elements bracketing the Australian by fluvio-lacustrine clastic successions.
plate include the divergent southern margin between The deposition of thick, mid –late Permian coal
Australia and Antarctica, transpressional conver- measures (2–15 m), particularly in the eastern part
gence at the southeastern plate margin, compression of the basin, occurred during thermal sag, which
along the northern and northwestern plate margin was followed by significant late Permian inversion
(particularly the Papua New Guinea fold–thrust of local half-graben. In the Bowen basin, one to
belt and the Himalayan collision zone) and subduc- three coal seams of economic interest can be devel-
tion at the northeastern margin (Indonesian Arc; oped at the well scale. Late Permian and Triassic
Reynolds et al. 2002, 2003; Sandiford et al. 2004). contraction led to complex reactivation structures
The Bowen basin (Figs 1 & 2) is interpreted as a and the deposition of late Permian–Early Triassic
Permian to Triassic back-arc basin (Holcombe et al. fluvio-marine clastics; deposition ceased during a
1997a, b; Korsch & Totterdell 2009) and is one of a mid –late Triassic contractional event. The develop-
series of rift basins that developed across eastern ment of permeability in coal seam gas fields is
Australia from the Early Permian (Korsch et al. loosely associated with structural highs.

From: Turner, J. P., Healy, D., Hillis, R. R. & Welch, M. J. (eds) Geomechanics and Geology.
Geological Society, London, Special Publications, 458, https://doi.org/10.1144/SP458.4
# 2017 The Author(s). Published by The Geological Society of London. All rights reserved.
For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics
E. TAVENER ET AL.

Fig. 1. (a) Indo-Australian plate and surrounding plate margins. Orange circles indicate earthquakes at plate
margins; the diameter of the circles corresponds to the relative earthquake magnitude. (b) Plate motion vectors,
Indo-Australian plate. Adapted from Sandiford (2016).

The southeastern part of the Bowen basin is over- Coal seam methane has been produced in SE
lain by the Surat basin (grey coloured area in Figs 2 Queensland for over 20 years. The industry saw a
& 3), which formed a broad intracontinental depres- marked acceleration of drilling and broad data
sion during the Jurassic. Surat basin sediments are acquisition in the early 2000s, when several world-
dominated by siltstones and mudstones with minor scale coal seam gas to liquefied natural gas projects
sandstones, all of which were deposited in a fluvio- commenced construction and production. Early
lacustrine depositional environment. All the clastic production was mainly from vertical wells with
lithologies contain significant volcanic components, or without hydraulic fracture completions. Ongoing
resulting in low porosity and permeability. The optimization of both drilling design and hydraulic
Surat basin contains multiple coal seams (on aver- fracture completions has fostered a wealth of diag-
age 60) with an average thickness of 0.4 m. These nostic data acquisition and some design changes
coal seams form the Walloons fairway (Fig. 2). (Flottmann et al. 2013; Kirk-Burnnand et al.
Structuring in the Surat basin is subtle and 2015). Initial work on the geomechnical framework
reflects the more intense deformation of the underly- (Brooke-Barnett et al. 2015) has shown the variabil-
ing reactivated inversion structures in the Permian ity of stress magnitudes and the potential influence
Bowen basin. Triassic structural highs set up gentle, of fundamental basement and basin structures.
low-amplitude highs in the Surat basin, which are The focus of this study is the in situ stress state in
characterized by exceptional permeabilities ranging the context of the geology and geomechanics of the
from hundreds of millidarcies (mD) to multidarcies Permian Bowen basin and the Roma Shelf region of
(e.g. Undulla Nose, Fig. 2). High permeability the Surat basin (Fig. 2). The wireline data utilized in
regions are characterized by coals with multiple this study are publicly available at QDEX (2016).
fracture orientations readily identifiable on image The results presented here complement earlier stud-
logs (i.e. no preferred fracture orientation). ies conducted in the eastern part of the Jurassic Surat
STRESS MECHANICAL STRATIGRAPHY

Fig. 2. (a) Location of Bowen and Surat basins in Australia. (b) SEEBASE map showing the structure of the top
basement underlying the Bowen basin (see text for further details); grey area shows extent of the Walloons
(production) fairway of the Surat Basin; dotted line is the location of the regional seismic line in part (c).
(c) Seismic line showing basic structural elements of Bowen and Surat basins

basin (excluding the Roma Shelf region) to the south This paper has four objectives:
of the Bowen basin (Figs 2 & 3; Flottmann et al.
2013; Brooke-Barnett et al. 2015). The data pre- (1) Documenting the significant plan view vari-
sented here show significant three-dimensional ability of the SHmax orientation in an intraconti-
complexity and granularity in the stress tensor (the nental basin setting utilizing a comprehensive
relative magnitude of the principal stresses) and and regionally extensive in situ stress dataset
their plan view orientation in an area with hitherto based on .180 wells, of which 145 wells pre-
sparse datasets; the World Stress Map (Heidbach sent new data.
et al. 2008, 2010) shows a comparatively uniform (2) Establishing systematic variations of stress
distribution of the maximum horizontal stress geometry with depth from representative
(SHmax) in eastern Queensland. Similarly, both examples of one-dimensional wellbore stress
the magnitude of differential stresses and the Ander- models based on log-derived strain-based
sonian stress state (reverse/strike-slip/normal; stress calculations.
Anderson 1951) varies significantly with depth. (3) Discussing the implications of both the lateral
The geomechanics of the Bowen and Surat basins and vertical stress variability on the bulk
are uniquely influenced by the geological setting mechanical stratigraphy which, in turn, influ-
because the stratigraphic column vertically juxta- ence completion strategies, such as hydraulic
poses lithologies with starkly contrasting rheologi- fracture stimulation and inclined and horizon-
cal properties (e.g. clastic sediments v. coals). tal drilling.
E. TAVENER ET AL.

Fig. 3. Azimuths of SHmax from breakout interpretation on image logs on the SEEBASE map (see Fig. 2b). All
orientations are from breakouts deeper than 450 m and wells with a wellbore inclination ,308. Note the variability
of SHmax azimuths in the different domains (see text for discussion). Coloured circles represent the Shmin magnitude
as derived from LOTs (below 450 m TVD RT). The map illustrates that the north and central Bowen regions appear
to be highly stressed, the Roma Shelf and Burunga Anticline are at intermediate stress and the south Bowen is in the
lowest present day stress. Open circles refer to wells presented in Figures 5 –8 (from north to south).
STRESS MECHANICAL STRATIGRAPHY

(4) Assessing the in situ stress and geomechanical 2008). The stress data are displayed on a basin struc-
implications based on data from the Permian ture map generated using the SEEBASE method
Bowen basin (presented here) with existing of integrating various datasets to generate a best
data and interpretations from the Jurassic approximate of a ‘depth-to-basement’ structure
Surat basin. image (SEEBASE 2005).
One-dimensional wellbore stress models are
used to constrain the relative magnitude of Shmin
and SHmax (in relation to SV) with depth. One-
Data, conditioning and calibration dimensional wellbore stress models are based on
The basic stratigraphic correlations are from stan- Poisson’s ratio (equation 2) and Young’s modulus
dard well logs (gamma ray, density, resistivity, (equation 3), which are derived from dipole
two-arm caliper; for stratigraphic overview, see sonic and density wireline data (dynamic data).
Cook & Jell 2013). Integrating the density log and A dynamic to static conversion of Poisson’s ratio
extrapolating the trend to the surface allows calcula- and Young’s modulus was derived regionally
tion of the vertical stress magnitude. The integral of from rock mechanics laboratory measurements
the density logs with reference to depth gives the (equations 4–7). Minimum and maximum horizon-
gradient of the vertical stress (SV), usually around tal stresses were then calculated using poroelastic
1 psi/ft (c. 19.2 ppg, c. 22.6 kPa m21) (equation 1). stress equations (Eaton 1968, 1972, 1975; Thierce-
Vertical (overburden) stress at depth z (Pa): lin & Plumb 1994; equations 8 & 9), which incorpo-
rate the static Poisson’s ratio, vertical stress,
z pore pressure, the static Young’s modulus and
Biot’s coefficient, as well as tectonic strain in the
SV = r(z)g dz (1) minimum (1min) and maximum (1max) horizontal
0 stress directions. Nominal tectonic strain values of
1max ¼ 0.0009 and 1min ¼ 0.0003 were used for
where z is the depth below ground level, r is the den- the initial calculations before calibration (Brooke-
sity in kg m23 and g is the acceleration due to grav- Barnett et al. 2015). Pore pressure was calculated
ity (assumed to be 9.81 m s22). based on a freshwater hydrostatic gradient, which
The orientation of SHmax (the maximum horizon- is commonly observed in undepleted coal reser-
tal stress) is derived from borehole breakouts voirs. Biot’s coefficient was not independently
and the drilling-induced tensile fractures (DITFs) constrained and was set as 1 to ensure consistency
observed in image logs. Breakouts form due to con- across the basin, thus true variation in poroelastic
jugate microfracturing and shear failure at the well- strain constants can be assessed.
bore wall in response to hoop stresses (Kirsch 1898). Dynamic Poisson’s ratio (ndyn) (no units):
Breakout is a product of far-field stresses interacting
with the wellbore wall and elongating the wellbore
(Vp /Vs )2 − 2
in the direction of Shmin (the minimum horizontal vdyn = (2)
stress; see Bell 1990, 1996a, b), leading to an overall 2[(Vp /Vs )2 − 1]
oval shape of the wellbore. DITFs form in the
azimuth of SHmax; they form sharp, usually linear, where Vp is the compressional sonic velocity in
fractures where the tensile hoop stresses are greater m s21 and Vs is shear sonic velocity in m s21.
than the tensile rock strength at the wellbore wall. Dynamic Young’s modulus (Edyn) (Pa):
Both breakouts and DITFs are sensitive to mud-
weight changes and the magnitude of the differential
stresses (in vertical wellbores, SHmax 2 Shmin) rVs2 (3Vp2 − 4Vs2 )
applied to the wellbore wall. For a given far-field Edyn = (3)
Vp2 − Vs2
stress state, elevated mud-weights can suppress
the initiation of borehole breakout, whereas high
mud-weights can, in turn, initiate DITFs. The where r is the density in kg m23, Vp is the compres-
wells used here are usually drilled either under sional sonic velocity in m s21and Vs is the shear
balance or slightly over balance with respect to sonic velocity in m s21.
the hydrostatic gradient of.433 psi/ft (c. 8.3 ppg; Sandstone static Poisson’s ratio (nstat) (no
9.8 kPa m21). Wireline logging is undertaken after units):
the wellbore is filled with a 3% KCl brine under
slightly overbalanced conditions. All logs are
vstat = 0.7vdyn + 0.06 (4)
referenced and corrected to true north. Stress mea-
surements were ranked according the classification
scheme of the World Stress Map (Tingay et al. where vdyn is the dynamic Poisson’s ratio.
E. TAVENER ET AL.

Siltstone static Poisson’s ratio (nstat) (no units): The initial calibration of the stress profiles was
undertaken using the stress polygon method (Moos
vstat = 0.7vdyn + 0.08 (5) & Zoback 1990; Zoback 2007). This method uses
the incidence of borehole failure (breakout and
drilling-induced tensile failure) to estimate the
where vdyn is the dynamic Poisson’s ratio. stress conditions required to induce failure within
Sandstone static Young’s modulus (Estat) the rock, using the frictional limit (defined by the
(Pa): friction angle) and compressional and tensional
strength of the rock at the point of failure occur-
Estat = 0.32Edyn (6) rence. The friction angle was calculated using the
method defined by Lal (1999; equation 10). The
where Edyn is the dynamic Young’s modulus. unconfined compressive rock strength was defined
Siltstone static Young’s modulus (Estat) (Pa): based on empirical relationships listed in Chang
et al. (2006). The equation defined by McNally
(1987), based on data from the Bowen basin, was
Estat = 0.3Edyn (7)
used for sandstones (equation 8). Where available,
fracture closure pressures derived from pressure
where Edyn is the dynamic Young’s modulus. data such as leak-off tests (LOTs), diagnostic frac-
Strain-derived SHmax (Pa): ture injection tests and modular formation dynamic
tester minifracs and pre-injection minifrac tests
vstat were used to constrain the minimum principal stress
SHmax − aPp = (SV − aPp )
s Hmax 1 − vstat (Barree et al. 2007, 2009). Table 1 gives the wells
Vertical component and the specific calibration method(s) used.
Estat Friction angle (88 ) (Lal 1999):
+  (1max + vstat 1min )
1 − v2stat  
Vp − 1000
Tectonic component
w = sin−1 (10)
vstat Vp + 1000
⇒ SHmax = (SV − aPp )
1 − vstat
where Vp is compressional sonic velocity in m s21.
Estat Sandstone compressive rock strength (MPa)
+  (1max + vstat 1min ) + aPp
1 − v2stat (McNally 1987):
(8)
UCS = 1200e(−0.036Dt) (11)
Strain-derived Shmin (Pa):
where Dt is the compressional sonic slowness in
vstat ms/ft and e is the base of natural logarithm.
Shmin − aPp = (SV − aPp )
shmin 1 − vstat
Vertical component

Estat Stress orientation


+  (1min + vstat 1max )
1 − v2stat
Tectonic component
In situ stress data covering an area of 350 × 210 km
(.70 000 km2; for comparison, an area more than
vstat
⇒ Shmin = (SV − aPp ) half the size of England) show significant variation
1 − vstat in the orientation of SHmax. Based on the results, six
Estat distinct domains can be identified: the north Bowen,
+  (1min + vstat 1max ) + aPp south Bowen and Burunga Anticline regions in the
1 − v2stat
Bowen basin and the Roma Shelf and Taroom Trough
(9) regions in the Surat basin (Fig. 3).
Regional maps of the SHmax orientation and Shmin
where sHmax is the maximum effective horizontal magnitude have been constructed using all the avail-
stress, shmin is the minimum effective horizontal able data from open file wells to December 2015.
stress, vstat is the static Poisson’s ratio, Sv is the ver- The SHmax orientation has been determined from
tical stress, a is Biot’s coefficient, Pp is the forma- observations of breakout or drilling-induced frac-
tion pressure, Estat is the static Young’s modulus, tures on image logs or from breakout observed on
1max is the strain in the maximum horizontal stress four- or six-arm caliper logs. The SHmax orientation
direction and 1min is the strain in the minimum hor- represented on the map in Figure 3 is derived
izontal stress direction. from breakout data. The DITF data give the same
STRESS MECHANICAL STRATIGRAPHY

Table 1. Wells (including offset wells) used in this study and calibration method for one-dimensional wellbore
stress models

Well name Latitude (S) Longitude (E) LOT FIT ISIP Pc IL RST

Durham Ranch 164 268 03′ 38.10′′ 1498 13′ 32.72′′ † †


FV03-15-1 258 33′ 25.31′′ 1488 59′ 51.29′′ † † † †
Arcadia Branch 5 258 17′ 29.35′′ 1488 55′ 41.88′′ † † † †
Hermitage 14 268 20′ 44.63′′ 1498 0′ 59.09′′ † † † † †
Ironbark Gully 4 258 31′ 45.57′′ 1488 55′ 58.90′′ †
Hermitage 11 268 20′ 42.63′′ 1498 2′ 55.79′′ †
Sunnyholt 2 258 17′ 37.66′′ 1488 50′ 44.84′′ † † †
Sunnyholt 3 258 17′ 17.98′′ 1488 51′ 06.91′′ † † †
Sunnyholt 4 258 17′ 47.61′′ 1488 51′ 12.36′′ † † †
Sunnyholt 11 258 17′ 50.82′′ 1488 51′ 36.55′′ †
Arcadia Valley 2 258 24′ 15.00′′ 1488 50′ 57.82′′ † † †
Mount Kingsley 1 258 13′ 55.22′′ 1488 54′ 6.84′′ † † †
FV17-35-1DW1 258 47′ 19.97′′ 1498 01′ 29.01′′ † † † † †

FIT, formation integrity test; IL, image log; ISIP, instantaneous shut-in pressure; LOT, leak-off test; Pc, closure pressure; RST, rock
strength testing.

result, but there are fewer data points and they are et al. (2015). Over the entire study area, the mean
not discussed further herein. The mean SHmax SHmax orientation is c. 428 N with a standard devia-
azimuth is represented by the straight lines given tion of c. 368 (Table 2), giving the area a type 1
in Figure 3. To avoid ambiguity, all data presented stress ranking. However, the orientation and quality
are from depths .450 m and from wells with of the in situ stresses varies significantly between
,308 wellbore inclination. the six domains that make up the whole area.
This study builds on the findings of Brooke-
Barnett et al. (2015) by applying the statistical
methodology outlined by Hillis & Reynolds (2000, (1) The northern Bowen domain is dominated by
2003) over both the Surat and Bowen basins. Conse- a NNE-trending SHmax orientation; the domi-
quently, the Rayleigh test was applied to the nant SHmax orientation here is c. 228 N with
stress orientation data to determine the confidence a standard deviation of c. 198 (Table 2).
of stress orientations over the study area (Mardia There are individual diversions from the dom-
1972; Table 2). Wells were also grouped into inant orientation in the very west of the study
regions based on the underlying SEEBASE topogra- area; in the far north individual easterly trend-
phy and SHmax orientation (Fig. 3) and the Rayleigh ing as well as one southeasterly trending out-
test was applied separately to each of these regions. lier are recorded. This region is designated
The regions were then classified into six types using as a type 2 stress region.
the following criteria: a type 1 region can reject the (2) The south-central Bowen domain of the
null hypothesis that stress orientations are random at Bowen basin forms a transitional corridor
the 99.9% confidence interval; a type 2 region can of variable SHmax orientations; both NE and
reject the null hypothesis at the 99% confidence NW trends as well as easterly trends occur,
interval; a type 3 region can reject the null hypoth- which gradually changes to an east –west to
esis at the 97.5% interval; a type 4 region can reject WNW–ESE trend of SHmax further south.
the null hypothesis at the 95% interval; a type 5 can Overall, this region exhibits a mean SHmax ori-
reject the null hypothesis at the 90% interval; and a entation of c. 708 N with a standard deviation
type 6 region suggests that the null hypothesis of c. 348 (Table 2). Despite the higher spread
cannot be rejected at the 90% interval (Hillis & in orientation in this region, the sheer amount
Reynolds 2000, 2003). As per the methodology of reliable stress indicators (21 A –C type
of Hillis & Reynolds (2000, 2003), Table 2 shows measurements) enable this region to have a
the results of the Rayleigh test as applied to the type 1 stress ranking.
mean SHmax orientations from A to C quality bore- (3) The southern Bowen domain (immediately
hole breakouts and DITF measurements. However, north of the Roma Shelf on Fig. 3) exhibits a
the mean statistics were also calculated using all very consistent east –west SHmax orientation
borehole breakouts and DITF measurements (A to with a standard deviation of c. 48 (Table 2).
E quality) as well as the average SHmax for each Similar to the north Bowen domain, the con-
well as per the methodology of Brooke-Barnett sistency of orientation between wells, despite
Table 2. Results of Raleigh analysis

Region Count A – C type Statistics Statistics Statistics well average


A–C A–E Shmax

A B C D E Total Borehole DITFs Count Mean Rn SD Conf. Type Count Mean Rn SD Count Mean Rn SD
breakouts (8N) (8) (8N) (8) (8N) (8)

E. TAVENER ET AL.
North Bowen 1 3 1 32 21 58 5 0 5 12.1 0.9 10.7 .99 2.0 58.0 25.2 0.7 22.1 29.0 22.3 0.8 18.9
South-central Bowen 2 8 11 55 18 94 5 16 21 88.2 0.6 26.7 .99.9 1.0 94.0 67.8 0.4 37.5 47.0 69.5 0.5 34.2
Southern Bowen 0 1 4 3 2 10 4 1 5 93.3 1.0 3.8 .99 2.0 10.0 87.4 1.0 9.1 5.0 90.7 1.0 4.4
Roma Shelf 0 0 1 55 22 78 0 1 1 36.0 1.0 0.0 ,90 – 78.0 17.8 0.9 13.0 39.0 18.0 0.9 13.8
Burunga Anticline 2 8 7 9 12 38 15 2 17 53.0 0.9 14.3 .99.9 1.0 38.0 50.9 0.8 17.2 19.0 48.7 0.9 13.8
Taroom Trough 1 1 2 41 21 66 4 0 4 60.2 0.6 27.8 ,90 6.0 66.0 82.4 0.3 41.6 33.0 77.9 0.3 42.9
New England Orogen 0 0 1 10 7 18 1 0 1 23.9 1.0 0.0 ,90 – 18.0 56.4 0.7 22.8 9.0 52.1 0.8 18.9
Total 6 21 27 205 103 362 34 20 54 66.2 0.5 33.1 .99.9 1.0 362.0 43.8 0.4 38.1 181.0 41.5 0.4 36.3

The Rayleigh test was applied to the entire area and the sub-regions defined in this paper based on A to C quality stress indicators (Statistics A –C). Mean statistics have also been calculated on A–E quality
stress indicators (Statistics A –E) and the mean stress orientation for each well (Statistics well total SHmax) based on input well data. Mean is the mean SHmax calculated for a region, Rn is the length of the vector
resulting from the sum of all SHmax orientations within a region (Mardia 1972), SD is the standard deviation of the calculated mean SHmax, Conf. is the confidence level at which the null hypothesis that the stress
orientation is random can be rejected and the type is the stress province type as per the methodology of Hillis & Reynolds (2002, 2003). DITFs, drilling-induced tensile fractures.
STRESS MECHANICAL STRATIGRAPHY

the small sample size, means this domain also and rock properties. The strainless one-dimensional
has a type 2 ranking. wellbore stress models are validated and refined
(4) The Roma Shelf domain, in which most stress with known data, including leak-off data, DFIT or
measurements are from wells in the Jurassic mini-frac closure pressure data, which gives an
Walloons sequence, is again dominated by estimate of the minimum horizontal stress at a par-
NNE- (c. 188) trending SHmax orientations. ticular depth. Alternatively, laboratory-based rock
Note that although the orientations between strength testing or frictional limits based on stress
wells are relatively consistent (standard devi- indicators from image log analyses were used to cal-
ation of c. 148) the quality of stress indicators ibrate the horizontal stress magnitudes (for method
in this region is low, meaning there are insuf- applied, see figure captions). One-dimensional well-
ficient data to assign a ranking to this area bore stress models show systematic variations
(Table 2). of Andersonian stress geometry with depth in the
(5) The Burunga Anticline on the eastern margin Bowen and Surat basins (see Flottmann et al.
of the Taroom Trough displays a consistent 2013; Brooke-Barnett et al. 2015).
NE trend of SHmax (c. 518) with a standard At depths shallower than around 500– 600 m
deviation of c. 148 (Table 2). The low standard TVD (total vertical depth, i.e. the depth below the
deviation and high quality of stress indicators surface measured from the drill rig floor), the stress
give this region a type 1 stress ranking. state is characterized by a reverse stress regime
(6) The Taroom Trough domain of the Surat basin (SV , Shmin , SHmax, where SV ¼ vertical stress,
(which overlies the depocentre of the Permian Shmin ¼ minimum horizontal stress and SHmax ¼
Bowen basin; Figs 2 & 3) to the east of the maximum horizontal stress). Image log data in
Roma Shelf shows a complex SHmax orienta- reverse stress regimes are dominated by borehole
tion. In the basin centre the stress azimuths breakouts (DITFs are largely absent) and breakout
are dominated by broadly easterly trends orientations indicate significant variability in the
(both ENE and ESE), but show a number azimuth of SHmax (Fig. 4). The one-dimensional
of significant deviations from a clear overall wellbore stress models presented in Figures 5 –8
trend. At the flanks of the depocentre the indicate low horizontal differential stresses (i.e.
SHmax orientation swings into parallelism the difference between SHmax and Shmin), in particu-
with the depocentre boundaries (a fault system lar at depths where reverse stress regimes are
in the east and a ramp in west, Brooke-Barnett dominant. The low differential stress is a likely
et al. 2015). Other datasets (Brooke-Barnett cause of the scatter in the breakout orientations at
et al. 2015) show a NE trend to the east, shallow depths. Below 500–600 m TVD the stress
where the Walloons depocentre is underlain geometry is typically of a strike-slip stress regime
by a basement high (New England Orogen (Shmin , SV , SHmax). Image log data in strike-slip
Region; Table 2). Despite the inclusion of stress regimes show both borehole breakouts and
additional data, this region retains the type 6 DITFs (Fig. 4a), both of which occur dominantly
designation from Brooke-Barnett et al. (2015). in shale/siltstone units.
Figure 4b and c show the wellbore stress condi-
tions at 400 and 800 m, respectively, using the
One-dimensional wellbore stress models stress and rock strength data given in Figure 6 for
(mechanical Earth models) those depths. The stress/rock strength conditions
at 400 m allow for broad compressive failure
A one-dimensional wellbore stress model (also where the maximum horizontal stress exceeds the
called a mechanical Earth model) is a numerical compressive failure. This condition is represented
representation of the geomechanical state of the sub- by the occurrence of scattered borehole breakouts.
surface over a given interval and combines known The stress conditions do not reach tensile failure,
pore pressures, the stress state (vertical, minimum resulting in the sparse development of DITFs at
and maximum horizontal) and rock mechanical this depth (Fig. 4a). At 800 m depth both the differ-
properties (uniaxial compressive strength, Young’s ential stresses and the uniaxial compressive strength
modulus and Poisson’s ratio). The one-dimensional are higher than at 400 m (Fig. 6). This results in the
wellbore stress models herein have been generated preferential development of DITFs (as the stress
using RokDoc623 software. conditions exceed the tensile rock strength). Both
The one-dimensional wellbore stress models DITFs and borehole breakouts occur in a well-
have been compiled for four example wells within defined narrow band at this depth. The data compi-
the Bowen basin (circled well locations, Fig. 3). lation in Figure 4a shows the dominance of DITFs at
The one-dimensional wellbore stress models are depths .1000 m; the dominance of DITFs at greater
created using shear and p-wave sonic velocities depths appears to be related to increasing differen-
and elastic models to produce estimates of stress tial stress with depth in this part of the Bowen basin.
E. TAVENER ET AL.

Fig. 4. (a) Compilation of wellbore breakout data (grey circles) and DITFs (black circles) from the central Bowen
basin region. Note the variability of breakout azimuth at shallow depth and the progressive domination of DITFs at
greater depth. (b) Circumferential stress distribution at 400 m depth using rock strength and stress data given in
Figure 6; note that the minimum stress does not allow the generation of tensile fractures (i.e. no DITFs), but the
maximum stress exceeds the compressive strength, resulting in the wide-ranging development of borehole breakouts
(BO). (c) Same as part (b), but at 800 m; note that the minimum stress exceeds the tensile strength, resulting the
preferential development of DITFs at greater depths. (d) Rose plot of combined azimuth of SHmax data from
wellbore analysis presented in part (a).

The transition between reverse stress regimes and 2013). The dataset presented here shows a similar
strike-slip stress regimes is also well documented in stress regime transition at the same depth range,
tiltmeter data acquired during hydraulic stimula- but in the Permian Bowen basin. The co-occurrence
tions in the Jurassic Surat basin (Flottmann et al. of the transition from a reverse to a strike-slip stress
STRESS MECHANICAL STRATIGRAPHY

Fig. 5. Logs and one-dimensional wellbore stress model for the Arcadia Branch 5 well (see Fig. 3, circled wells
show well locations of Figs 5–8 from north to south). Left-hand side: input log data used to calculate Shmin and
SHmax on right-hand side. Shmin generally smaller than SV, but SHmax greater than SV, indicating a strike-slip stress
regime. Colour scheme: Vs and Vp logs, blue colours slow (e.g. coals, carbonaceous shales), red colours fast (hard
sand and shales). Formation tops: BAND denotes Bandanna Formation, above Rewan Formation (for stratigraphic
detail, see Cook & Jell 2013).

state at a similar depth range in two different basins constrained by rock strength testing from offset
suggests that the transition in stress state is con- wells in both the interburden and the coals. Shmin
trolled by the present day depth rather being con- has been constrained using DFIT, minifrac and leak-
trolled by geological or stratigraphic parameters. off data. Results from one-dimensional wellbore
A second transition from a strike-slip to a normal stress models display some key contrasts in mechan-
stress regime (Shmin , SHmax , SV) has been docu- ical stratigraphy. In principle, (non-coal) interburden
mented at c. 650 –800 m in some areas in the Surat rocks have a higher rock strength than coals, which
basin. This transition does not occur in the Bowen display a consistently low rock strength based on a
basin. The magnitude of differential stresses in the high Poisson’s ratio and low Young’s modulus.
Surat basin also show significant variability. This Coals are dominated by a normal stress regime,
appears to be attributed to both the nature of the regardless of whether the surrounding rocks are in
basement and/or the thickness of the sedimentary a reverse, strike-slip or normal overall stress regime.
section underlying the Surat basin (Brooke-Barnett Importantly, in reverse and strike-slip stress regimes
et al. 2015). coals exhibit generally lower overall stresses than the
surrounding country rock (Fig. 9a, b). This has been
established by numerous systematic DFIT tests in
Mechanical stratigraphy numerous wells in both the Bowen and Surat basins
(Fig. 9a, b; Flottmann et al. 2013). The same result
Rock properties (Young’s modulus, the uniaxial is achieved by establishing frictional limits theory
compressive strength and Poisson’s ratio) are based on image log analyses.
E. TAVENER ET AL.

Fig. 6. Logs and one-dimensional wellbore stress model for the Fairview F-V 03-15-1 well. Colour scheme as in
Figure 5. Note distinct change from a reverse to a strike-slip stress regime around 500 m. Black squares show DFIT
calibration points for stress model. Formation tops: PRECI, Precipice Sandstone; CLEM, Clematis Formation;
REW, Rewan Formation; BAND, Bandanna Formation; BLKA, Black Alley Shale.

The relationships given in equations (8) and (9) stress models, the essential elements impacting the
suggest that rocks with a high Poisson’s ratio (e.g. mechanical stratigraphy in the Bowen and Surat
coals) are more susceptible to accommodating basins can be reduced to three key components:
high stress in tectonic scenarios dominated by verti- (1) coals with a comparatively high Poisson’s ratio
cal ‘loading’ (overburden); conversely, horizontal and low Young’s modulus; (2) interburden rocks
‘loading’ (i.e. the tectonic component) is domi- with a comparatively low Poisson’s ratio and a
nantly accommodated in rocks with a high Young’s high Young’s modulus; and (3) a stress regime dom-
modulus. Based on the one-dimensional wellbore inated by horizontal (tectonic) components.

Fig. 7. Logs and one-dimensional wellbore stress model for the Durham Ranch DM 164 well. Colour scheme same
as Figure 5. Note distinct change from reverse stress regime above 400 m to strike-slip stress regime below 650 m.
Formation tops: EVER, Evergreen Formation; BOXS, Boxvale Sandstone; LEVE, Lower Evergreen; PRECI,
Precipice Sandstone; LPREC, Lower Precipice Sandstone; REW, Rewan Formation; BAND, Bandanna Formation;
KALM, Kaloola Member; BLKA, Black Alley Shale.
STRESS MECHANICAL STRATIGRAPHY

Fig. 8. Logs and one-dimensional wellbore stress model for the Hermitage 14 well (location Fig. 3). Colour scheme
same as Figure 5. Note numerous transitions from a reverse to a strike-slip regime. Black squares show DFIT
calibration points for stress model. Formation tops: WEAL, Weald Sandstone; SPBK, Springbok Formation; UPJU,
Upper Juandah coal measures; PROU, Proud Sandstone; LOJU, Lower Juandah coal measures; TANG, Tangalooma
Sandstone; TARO, Taroom coal measures.

Fig. 9. (a) Close up of DM 164 log over coal interval (grey shading); note that both SHmax and Shmin are smaller
than SV (straight brown line); coals are in normal stress regime. (b) Same observations as for well FV-03-15-1.
E. TAVENER ET AL.

In the following discussion we explore the oil- fact, the stress rotations may provide a guide for
field impacts of the key elements of the mechanical future remote deep data acquisition. Regardless,
stratigraphy in the Bowen and Surat basins, which is the observations presented here suggest that stresses
relevant to basins with similar conditions worldwide. in intracontinental basins can vary significantly
and without any visible first-order manifestation
in complementary datasets, such as faults or basin
Discussion structure.
However, from an oilfield perspective, it is
The data presented in this paper show variability in important to take local variations of SHmax into con-
the plan view stress azimuths and Andersonian sideration. For horizontal and/or deviated wells, for
stress geometry with depth. The data and interpreta- example, the SHmax azimuths can have significant
tions from the Permian Bowen basin presented here, implications with regard to optimizing wellbore
in combination with similar data from the Jurassic stability in deviated/horizontal wells in the context
Surat basin (Brooke-Barnett et al. 2015), show of rock strength and mud-weight optimization to
similar depths for the transition from a reverse to prevent wellbore collapse. Wellbore stability is
a strike-slip regime (400 –600 m). This suggests highest where the differential stresses around well-
that the first-order influence for the occurrence of bores are at a minimum. This is particularly true in
this transition in stress regime is depth (TVD) rather deviated/horizontal wells (regardless of the overall
than local conditions, such as the basin-specific stress magnitudes). In many strike-slip regimes,
stratigraphy. The transition from a reverse to a which are common around the world, this often
strike-slip stress state at c. 400 –600 m depth thus entails deviating the wellbore into SHmax as the
appears to be typical for eastern Queensland. This differential stresses between Shmin and SV are less
is in contrast with interior parts of Australia (e.g. than, for example, the differential stress between
the Cooper basin), where a transition from a strike- SV and SHmax.
slip stress regime at shallow depths to a reverse In the context of fracture stimulation comple-
stress regime is reported to occur at depths in excess tions, a well deviated into SHmax may not be
of c. 2.5 km (e.g. Reynolds et al. 2006). optimum because hydraulic fractures propagate
Previous datasets presented in the World Stress preferentially into the azimuth of SHmax, resulting
Map (Heidbach et al. 2008, 2010) suggest an overall in a longitudinal fracture stimulation. Generally,
NNE–NE orientation of SHmax in the Bowen basin. transverse stimulations (i.e. the fracture propagates
Our data support this observation over the entire perpendicular to the wellbore) are more advanta-
region, including both the Bowen and Surat basins geous, particularly in low-permeability formations,
(Table 2). However, our data show significant because they access greater reservoir rock volumes.
local deviations from the inferred regional trend. Consequently, the wellbore has to be deviated
In particular, the south Bowen domain of the into Shmin and, during drilling, the mud-weight
Bowen basin and the Taroom Trough domain of window has to be adjusted to deliver a stable well-
the Surat basin both display variable SHmax orienta- bore under high differential stresses acting perpen-
tions and the Denison Trough domain displays dicular to the wellbore axis. Regionally varying
an east– west SHmax which, although consistent, stress azimuths as documented here highlight the
deviates significantly from the regional SHmax need to acquire a complete stress dataset, allowing
orientation. full stress tensor and stress azimuth descriptions to
The reasons why the orientation of SHmax azi- optimize both drilling and completion options. The
muths is subject to significant local variations local variability of the SHmax azimuths in our dataset
remains, at this stage, a matter of speculation. demonstrates that it may be insufficient to rely
Brooke-Barnett et al. (2015) showed that SHmax azi- on regional datasets alone; local conditions have
muths in the Surat basin are significantly influenced to be established to allow for optimum drilling and
by the underlying basement structures. However, completion designs.
available SEEBASE data show no obvious major Stress state transitions with depth, and between
structural trends that could guide stress rotations lithologies, are of primary practical importance.
of the severity seen in the southern Bowen basin. Away from the wellbore-influenced hoop stresses,
The easterly stress orientations in the south-central hydraulic fractures open against the regional mini-
and southern domains may be guided by deep- mum principal stress. Consequently, hydraulic frac-
seated easterly trending fault systems and volcanism tures are expected to be vertical in both strike-slip
related to the opening of the Coral Sea during the and normal stress regimes; however, in a reverse
Tertiary (Cook & Jell 2013). The changing compo- stress regime fracture opening is expected to be hor-
sition of the deeper basement could result in changes izontal and leads to a horizontal hydraulic fracture.
in the bulk rock strength, which, in turn, could The occurrence of horizontal fracturing at shallow
contribute to the stress rotations observed – in depth is corroborated by tiltmeter data in the Surat
STRESS MECHANICAL STRATIGRAPHY

basin, which show a predominance of horizontal 8 & 9). Consequently, in reverse and strike-slip
components (up to 100%) at shallow depths. At regimes the interburden rocks (high Young’s modu-
depths .400– 500 m (Flottmann et al. 2013), lus and low Poisson’s ratio) will accommodate high
tiltmeter data show a predominance of vertical stresses and the coals will be comparatively
fracture components. less stressed.
Three-dimensional stress characterization is a The interaction of the Andersonian stress state
key requirement for the planning and implementa- and rocks of contrasting rheological properties
tion of drilling and completions such as hydraulic thus has significant implications for the propagation
fracture stimulations. Hydraulic stimulations propa- of hydraulic fractures. In a normal stress regime,
gate perpendicular to the lowest principal stress for example, where coals are comparatively highly
in the plane defined by the intermediate and maxi- stressed, hydraulic fractures will initiate in coals,
mum principal stresses; hydraulic fractures tend but will tend to grow into (interburden) formations
to grow in the azimuth of SHmax. The mechanical dominated by a lower mean stress (Fig. 10a). Our
stratigraphy and stress magnitudes are of particular interpretation of one-dimensional wellbore stress
importance for the vertical containment of fracture models shows that, in the Bowen basin, the actual
growth. Fractures initiate and grow preferentially Andersonian stress states are reverse and strike
in formations with low Shmin. Formations with low slip and the coals tend to be the least stressed mem-
Shmin have a low mean stress [(SV + SHmax + bers of the mechanical stratigraphy. Interburden
Shmin)/3]; conversely, formations with a high Shmin rocks with a high Young’s modulus accommodate
and a high mean stress tend to act as barriers to the the horizontal tectonic stress component (equations
vertical growth of hydraulic stimulations, thereby 8 & 9) and are the most highly stressed components
containing the stimulation to the target zone. of the mechanical stratigraphy in the Bowen basin
However, different Andersonian stress regimes (Fig. 10b). Consequently, hydraulic fracture com-
result in contrasting stress accommodation and pletions targeting coals are well contained in the
stress intensity in rocks with different rheological low stress coals (Fig. 10b). Similar observations
properties. Lithologies with a high Poisson’s ratio are documented in the Surat basin, where tracer
and a low Young’s modulus (such as coals) tend logs show containment in coal during hydraulic
to accommodate high stresses in a normal stress treatments (Kirk-Burnnand et al. 2015).
regime (SV . SHmax . Shmin), where the maximum Different Andersonian stress regimes can result
principal stress is vertical and the vertical component in potentially stark contrasts in rock-specific stress
of stress is governed by Poisson’s ratio (equations 8 states (Fig. 10a, b). This highlights the need for
& 9; see also Herwanger et al. 2015). Conversely, in the full three-dimensional characterization of stress
tectonic regimes where the maximum principal parameters to achieve the appropriate conditioning
stress is horizontal (reverse and strike-slip tectonic of one-dimensional geomechanical models, which
regimes), the stress distribution is governed by are key (software) inputs for planning fracture stim-
rocks with a high Young’s modulus and a low Pois- ulations or well planning. The three-dimensional
son’s ratio (i.e. the tectonic component in equations variation of stress states in the Bowen basin

Fig. 10. (a) Compression in vice represents normal stress regime (maximum stress vertical, i.e. perpendicular to
layering), resulting in high stress in coal (Poisson’s ratio (PR) c. 0.4) and low stress in clastic interburden rocks
(Poisson’s ratio c. 0.25); hypothetical fracture stimulation indicated by oval shapes; initiation in coal (solid outline),
but progressive growth into low stress interburden. (b) Compression in vice represents strike-slip/reverse stress
regime (maximum stress horizontal, i.e. parallel to layering), resulting in high stress in interburden and low stress in
coal. Hypothetical fracture stimulation (ovals) stays contained in coal.
E. TAVENER ET AL.

highlight the interdependency between the plate Bell, J.S. 1990. The stress regime of the Scotian Shelf
tectonics boundary conditions, the local structural offshore eastern Canada to 6 kilometres depth and
geological setting and rheological parameters, all implications for rock mechanics and hydrocarbon
of which ultimately contribute to the geomechanical migration. In: Maury, V. & Fourmaintraux, D.
(eds) Rock at Great Depth. Balkema, Rotterdam,
conditions at individual wellbores. 1243– 1265.
Bell, J.S. 1996a. Petro Geoscience 1. In situ stresses in
sedimentary rocks (part 1): measurement techniques.
Conclusions Geoscience Canada, 23, 85– 100.
Bell, J.S. 1996b. Petro Geoscience 2. In situ stresses in
This paper documents broad, but systematic, sedimentary rocks (part 2): applications of stress mea-
changes in the in situ stress orientations and in situ surements. Geoscience Canada, 23, 135 –153.
stress state with depth in the intracontinental Brooke-Barnett, S., Flottmann, T. et al. 2015. Influ-
Bowen and Surat basins of eastern Queensland, Aus- ence of basement structures on in-situ stresses over the
tralia. Both basins show a transition from a reverse Surat Basin, southeast Queensland. Journal of Geo-
stress regime at depths shallower than 400 –600 m; physical Research: Solid Earth, 120, 4946–4965.
Chang, C., Zoback, M.D. & Khaksar, A. 2006. Empiri-
at greater depths strike-slip stress geometries are cal relations between rock strength and physical prop-
dominant. The (Andersonian) stress regime transi- erties in sedimentary rocks. Journal of Petroleum
tion is primarily depth-controlled, independent of Science and Engineering, 51, 223–237.
the basin and/or lithology. The spatial variability of Coblentz, D.D., Sandiford, M., Richardson, R.M.,
the in situ stress distribution results from the interac- Zhou, S. & Hillis, R. 1995. The origins of the intra-
tion of regional intra-plate stresses with basin-scale plate stress field in continental Australia. Earth and
structures and basement rheology. Stress geometries Planetary Science Letters, 133, 299 –309.
and rock properties materially influence drilling and Coblentz, D.D., Zhou, S., Hillis, R.R., Richardson,
completion considerations. In particular, hydraulic R.M. & Sandiford, M. 1998. Topography, boundary
forces, and the Indo-Australian intraplate stress field.
fracture completions have to be designed to take Journal of Geophysical Research, 103, 919– 931.
into account the rock-specific geomechanical condi- Cook, A.G. & Jell, J.S. 2013. Paleogene and Neogene. In:
tions established from log-derived one-dimensional Jell, P.A. (ed.) Geology of Queensland. Geological
wellbore stress models. The data and analyses pre- Survey of Queensland, Brisbane, 577– 685.
sented suggest that first-order regional stresses Eaton, B.A. 1968. Fracture gradient prediction and its
increase in complexity at a local scale; ultimately, application in oilfield operations. Paper SPE-2163-
the resolved geomechanical state at the wellbore PA, https://doi.org/10.2118/2163-PA
level reflects a scale-dependent interplay of plate Eaton, B.A. 1972. The effect of overburden stress on geo-
tectonic forces that are geologically modulated by pressures prediction from well logs. Paper SPE-3719-
PA, https://doi.org/10.2118/3719-PA
the local structure and the mechanical properties of Eaton, B.A. 1975. The equation for geopressure predic-
individual rock packages. tion from well logs. Paper SPE-5544-MS, https://
doi.org/10.2118/5544-MS
We thank Santos Ltd and Origin Energy for permission to Flottmann, T., Brooke-Barnett, S. et al. 2013. Influ-
publish this paper. We acknowledge the efforts of count- ence of in-situ stresses on fracture stimulations in
less colleagues and field personnel who contributed to the Surat Basin, southeast Queensland. Paper SPE
data acquisition and discussions leading to the results pre- 167064-MS, presented at the SPE Unconventional
sented here. Particular thanks to the convenors of the GSL Resources Conference and Exhibition-Asia Pacific,
conference Geology of Geomechanics for the opportunity 11–13 November 2013, Brisbane, Australia, https://
to present and encouragement to publish. Tony Addis and doi.org/10.2118/167064-MS
Joe English prepared thorough and insightful reviews, Heidbach, O., Tingay, M., Barth, A., Reinecker, J.,
which greatly improved the manuscript. Kurfeß, D. & Müller, B. 2008. The World Stress
Map Database Release 2008, https://doi.org/10.
1594/GFZ.WSM.Rel2008
References Heidbach, O., Tingay, M., Barth, A., Reinecker, J.,
Kurfeß, D. & Müller, B. 2010. Global crustal stress
Anderson, M.E. 1951. The Dynamics of Faulting. Oliver pattern based on the World Stress Map database release
and Boyd, Edinburgh. 2008. Tectonophysics, 482, 3– 15.
Barree, R.D., Barree, V.L. & Craig, D.P. 2007. Holistic Herwanger, J.V., Bottrill, A.D. & Mildren, S.D.
fracture diagnostics. Paper SPE 107877, presented at 2015. Uses and abuses of the brittleness index with
the SPE Rocky Mountain Oil & Gas Technology Sym- application to hydraulic stimulations. Paper URTEC-
posium, 16– 18 April, Denver, CO, USA. 2172545-MS, presented at the URTeC Conference,
Barree, R.D., Gilbert, J.V. & Conway, M.W. 2009. 20–22 July 2015, San Antonio, TX, USA, https://
Stress and rock property profiling for unconventional doi.org/10.15530/URTEC-2015-2172545
reservoir stimulation. Paper SPE 118703, presented Hillis, R.R. & Reynolds, S.D. 2000. The Australian
at the SPE Hydraulic Fracturing Convention, 19– 21 stress map. Journal of the Geological Society, London,
January, The Woodlands, TX, USA. 157, 915– 921, https://doi.org/10.1144/jgs.157.5.915
STRESS MECHANICAL STRATIGRAPHY

Hillis, R.R. & Reynolds, S.D. 2003. In situ stress field Moos, D. & Zoback, M.D. 1990. Utilization of observa-
of Australia. In: Hillis, R.R. & Müller, R.D. (eds) tions of well bore failure to constrain the orienta-
Evolution and Dynamics of the Australian Plate. tion and magnitude of crustal stresses: application to
Geological Society of America, Boulder, 49– 58. continental, Deep Sea Drilling Project and Ocean
Hillis, R.R., Enever, J.R. & Reynolds, S.D. 1999. In Drilling Program boreholes. Journal of Geophysical
situ stress field of eastern Australia. Australian Journal Research, 95, 9305–9325, https://doi.org/10.1029/
of Earth Sciences, 46, 813–825. JB095iB06p09305
Holcombe, R.J., Stephens, C.J. et al. 1997a. Tectonic Nelson, E.J., Chipperfield, S.T., Hillis, R.R., Gilbert,
evolution of the northern New England fold belt: Car- J., McGowen, J. & Mildren, S.D. 2007. The relation-
boniferous to Early Permian transition from active ship between closure pressures from fluid injection
accretion to extension. In: Ashley, P.M. & Flood, tests and the minimum principal stress in strong rocks.
P.G. (eds) Tectonics and Metallogenesis of the New International Journal of Rock Mechanics & Mining
England Orogen: Alan H. Voisey Memorial Volume. Sciences, 44, 787–801.
Geological Society of Australia, Special Publications, QDEX 2016. www.business.qld.gov.au/industry/mining/
19, 66–79. mining-online-services/qdex-data
Holcombe, R.J., Stephens, C.J. et al. 1997b. Tectonic Reynolds, S.D., Coblentz, D.D. & Hillis, R.R. 2002.
evolution of the northern New England fold belt: the Tectonic forces controlling the regional intraplate
Permian–Triassic Hunter-Bowen event. In: Ashley, stress field in continental Australia: results from new
P.M. & Flood, P.G. (eds) Tectonics and Metallogene- finite element modeling. Journal of Geophysical
sis of the New England Orogen: Alan H. Voisey Memo- Research, 107(B7), https://doi.org/10.1029/2001JB
rial Volume. Geological Society of Australia, Special 000408
Publications, 19, 52–65. Reynolds, S.D., Coblentz, D.D. & Hillis, R.R. 2003.
Kirk-Burnnand, E., Pandey, V.J., Flottman, T. & Influences of plate-boundary forces on the regional
Trubshaw, R.L. 2015. Hydraulic fracture design intraplate stress field of continental Australia. In:
optimization in low permeability coals. Paper SPE Hillis, R.R. & Müller, R.D. (eds) Evolution and
176895-MS, presented at the SPE Unconventional Dynamics of the Australian Plate. Geological Society
Resources Conference and Exhibition-Asia Pacific, of America, Boulder, 59–70.
9–11 November 2015, Brisbane, Australia, https:// Reynolds, S.D., Mildren, S.D., Hillis, R.R., Meyer, J.J.
doi.org/10.2118/176895-MS & Flottmann, T. 2005. Maximum horizontal stress
Kirsch, E.G. 1898. Die Theorie der Elastizität und die orientations in the Cooper Basin, Australia: implications
Bedürfnisse der Festigkeitslehre. Zeitschrift des Ver- for plate-scale tectonics and local stress sources. Geo-
eines deutscher Ingenieure, 42, 797–807. physical Journal International, 160, 331–343.
Korsch, R.J. & Totterdell, J.M. 2009. Evolution of the Reynolds, S.D., Mildren, S.D., Hillis, R.R. & Meyer,
Bowen, Gunnedah and Surat Basins, eastern Australia. J.J. 2006. Constraining stress magnitudes using
Australian Journal of Earth Sciences, 56, 271–272. petroleum exploration data in the Cooper– Eromanga
Korsch, R.J., Totterdell, J.M., Cathro, D.L. & Nicoll, Basins, Australia. Tectonophysics, 415, 123– 140.
M.G. 2009a. Early Permian East Australian Rift System. Sandiford, M. 2016. http://jaeger.earthsci.unimelb.edu.
Australian Journal of Earth Sciences, 56, 381–400. au/Images/images.html
Korsch, R.J., Totterdell, J.M., Cathro, D.L. & Nic- Sandiford, M., Wallace, M. & Coblentz, D.C. 2004.
oll, M.G. 2009b. Contractional structures and defor- Origin of the in situ stress field in south-eastern Austra-
mational events in the Bowen, Gunnedah and Surat lia. Basin Research, 16, 325–338.
Basins, eastern Australia. Australian Journal of Earth SEEBASE 2005. Public domain report to Shell Develop-
Sciences, 56, 477– 499. ment Australia by FrOG Tech Pty Ltd, edited.
Lal, M. 1999. Shale stability: drilling fluid interaction Thiercelin, M.J. & Plumb, R.A. 1994. A core-based pre-
and shale strength. Paper presented at the SPE Latin diction of lithologic stress contrast in east texas forma-
American and Caribbean Petroleum Engineering tions. SPE Formation Evaluation, 9, 251–258.
Conference, 21–23 April 1999, Caracas, Venezuela. Tingay, M., Reinecker, J. & Müller, B. 2008. Borehole
Mardia, K.V. 1972. Statistics of Directional Data. Aca- Breakout and Drilling-induced Fracture Analysis from
demic Press, New York. Image Logs. World Stress Map Project Guidelines:
McNally, G.H.N. 1987. Estimation of coal measures rock Image Logs. World Stress Map Project.
strength using sonic and neutron logs. Geoexploration, Zoback, M.D. 2007. Reservoir Geomechanics. Cambridge
24, 381–395. University Press, Cambridge.

You might also like