The present-day stress field of Australia

The present-day stress field of Australia

Accepted Manuscript The present-day stress field of Australia Mojtaba Rajabi, Mark Tingay, Oliver Heidbach, Richard Hillis, Scott Reynolds PII: DOI: ...

4MB Sizes 1 Downloads 56 Views

Accepted Manuscript The present-day stress field of Australia

Mojtaba Rajabi, Mark Tingay, Oliver Heidbach, Richard Hillis, Scott Reynolds PII: DOI: Reference:

S0012-8252(16)30436-6 doi: 10.1016/j.earscirev.2017.04.003 EARTH 2402

To appear in:

Earth-Science Reviews

Received date: Revised date: Accepted date:

21 November 2016 9 February 2017 7 April 2017

Please cite this article as: Mojtaba Rajabi, Mark Tingay, Oliver Heidbach, Richard Hillis, Scott Reynolds , The present-day stress field of Australia. The address for the corresponding author was captured as affiliation for all authors. Please check if appropriate. Earth(2017), doi: 10.1016/j.earscirev.2017.04.003

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT The Present-Day Stress Field of Australia

Authors: Mojtaba Rajabi1*, Mark Tingay1*, Oliver Heidbach2, Richard Hillis3, Scott Reynolds4

Australian School of Petroleum, University of Adelaide, Adelaide, SA 5005, Australia

2

Helmholtz Centre Potsdam, GFZ German Research Centre for Geosciences, Telegrafenberg,

PT

1

Deep Exploration Technologies Cooperative Research Centre, 26 Butler Boulevard, Burbridge

Business Park, Adelaide, SA 5950, Australia Ikon Science, Asia Pacific, Kuala Lumpur, Malaysia

*

Corresponding authors:

MA

NU

4

SC

3

RI

14473 Potsdam, Germany

Mojtaba Rajabi: Australian School of Petroleum, University of Adelaide, Adelaide, SA 5005,

PT E

D

Australia

Email: [email protected]

Australia

CE

Mark Tingay: Australian School of Petroleum, University of Adelaide, Adelaide, SA 5005,

AC

Email: [email protected]

Tel.: +61 (0) 8 8313 4304 Fax: +61 (0) 8 8313 4345

1

ACCEPTED MANUSCRIPT Abstract The unusual present-day crustal stress pattern of the Australian continent has been the subject of scientific debate for over 25 years. The orientation of maximum horizontal present-day stress (SHmax) in continental Australia is unlike all other major tectonic plates in that the stress pattern shows regional variability and it is not oriented sub-parallel to the direction of absolute plate motion. Previous studies on the stress pattern of Australia revealed that the complex stress pattern

PT

of the continent is controlled, at a first-order, by the superposition of plate tectonic forces exerted at the plate boundaries. However, prior analysis of the contemporary Australian crustal stress pattern have been unable to model or explain the stress pattern observed in most of eastern

RI

Australia, and has not extensively addressed the numerous smaller scale variations in stress

SC

orientation. The recent development of unconventional reservoirs in Australia has resulted in a greatly increased amount of new data for stress analysis in previously unstudied or poorlyconstrained areas in eastern Australia. In addition, stress analysis in conventional hydrocarbon,

NU

mineral and geothermal exploration in all other parts of the continent provides the opportunity to review and update the Australian Stress Map (ASM). This study presents the new release of the

MA

ASM, with a total of 2150 stress data records in Australia (increased from 594 data records in 2003). The 2016 ASM contains 1359 data records determined from the interpretation of drillingrelated stress indicators, 650 from earthquake focal mechanism solutions, 139 from shallow

D

engineering measurements and two from geological indicators. The results reveal four distinct regional trends for the SHmax orientation in Australia including a NNE-SSW SHmax trend in northern

PT E

and northwestern Australia, which rotates to a prevailing E-W orientation in most Western and South Australia. The orientation of SHmax in eastern Australia is primarily ENE-WSW and swings to NW-SE in southeastern Australia. A comparison between the new ASM database and

CE

neotectonic features further confirms the role of contemporary stress in recent deformation of the Australian crust. The 2016 ASM reveals significant discrepancies between newly observed SHmax

AC

orientations and predictions by published geomechanical-numerical models for the Australian continent. Forces generated at the boundary of Indo-Australian Plate remain the major control on the regional crustal stress pattern in continental Australia, however, the increased data density, particularly in eastern Australia, reveals numerous local perturbations of the stress field that were not previously clearly captured. Hence, the key findings of the new release are that local (intraplate) stress sources are more significant than previously recognized, particularly in eastern Australian basins, and cause substantial local scale rotations of the present-day crustal stress pattern that have not been factored into existing Australian stress models. Keywords: Australian Stress Map, intra-plate stress, plate boundary forces, plate tectonics, stress rotation, neotectonics.

2

ACCEPTED MANUSCRIPT Contents 1.

Introduction ......................................................................................................................... 4

2.

Tectonic setting, neotectonics and seismicity of the Australian continent ..................... 7

3.

An overview of earlier studies on the crustal stress pattern of Australia ...................... 8

3.1.

Early studies before the initiation of the Australian Stress Map project (pre-1996) ........ 8

3.2.

Early phase of the Australian Stress Map project (1996-2003) ...................................... 10

3.3.

New phase of the ASM project (2012-present)................................................................ 11 Stress indicator used in the Australian Stress Map database ....................................... 12

PT

4.

Wellbore breakouts and drilling-induced tensile fractures ............................................. 13

4.2.

Focal mechanism solution of earthquakes ...................................................................... 14

4.3.

Hydraulic fracturing measurements and overcoring ...................................................... 14

SC

5.

RI

4.1.

The new Australian Stress Map database (release-2016) .............................................. 15 Analysis of stress provinces............................................................................................. 17

5.2.

Analysis of the SHmax stress pattern.................................................................................. 19

5.3.

Present-day stress regimes throughout the Australian continent.................................... 20 Pattern of present-day crustal stress in the Australian continent ................................ 21

MA

6.

NU

5.1.

6.1.

Regional orientation of SHmax across Australia ............................................................... 21

6.2.

Comparison of present-day crustal stress field and neotectonic features ....................... 22 Review on the potential sources of the crustal stress of the Australian continent....... 23

D

7.

An overview on the potential stress sources at different spatial scales .......................... 23

7.2.

Australian crustal stress pattern and absolute plate motion ........................................... 23

7.3.

First-order control on the present-day crustal stress pattern ......................................... 24

7.4.

Controls on the crustal stress pattern at second-, third- and fourth-order ..................... 27

7.5.

Comparison between SHmax orientation and azimuthal seismic anisotropy ..................... 28

CE

PT E

7.1.

Applications of the Australian Stress Map database ..................................................... 29

9.

Conclusions ........................................................................................................................ 33

AC

8.

Acknowledgements .................................................................................................................... 34 Supplementary Information ........................................................ Error! Bookmark not defined. References .................................................................................................................................. 42

3

ACCEPTED MANUSCRIPT 1. Introduction The contemporary pattern of tectonic stress in the upper part of the Earth’s lithosphere has numerous implications in various economical, scientific and societal aspects of human’s life (Bell, 1996b; Cornet, 2015; Engelder, 1993; Montone and Mariucci, 2016; Palano, 2015; Reiter et al., 2014; Zoback, 2007). The present-day stress field controls neotectonic deformation and seismicity and can

PT

be used to better understand earthquakes and their recurrence rates (Hillis et al., 2008; Palano, 2015; Sandiford et al., 2004; Seeber and Armbruster, 2000; Sibson et al., 2011; Sibson, 1992; Sibson et al.,

RI

2012; Stein, 1999). The crustal stress is a primary cause of collapse of subsurface structures, and is

SC

thus vital knowledge in order to create safe and stable tunnels, mines and boreholes (Bell, 1996b; Zang and Stephansson, 2010; Zoback, 2007). The in-situ stress is the key control on hydraulically

NU

induced fractures, and thus the knowledge of the stress field has been critical in the rapid increase of unconventional hydrocarbon production seen globally, and particularly in the USA, Canada and

MA

Australia in the past 10 years (Bell, 1996b; Bell, 2006; Fisher and Warpinski, 2012; King, 2010; Lund

D

Snee and Zoback, 2016; Olsen et al., 2007; Palmer, 2010; Pitcher and Davis, 2016).

PT E

Regional compilations of the present-day crustal stress data at continental and plate scales, during the first phase of the World Stress Map (WSM) project from 1986-1992, demonstrated that plate tectonics forces provide the major controls on the contemporary crustal stress field. These forces

CE

are the ‘ridge push’ at mid-ocean ridges, ‘slab pull’ at subduction zones, ‘suction’ at trench zones, ‘resistance forces’ at collision zones, ‘shear traction’ at base of the lithosphere and the ‘gravitational

AC

potential energy’ at mountain belts (Coblentz et al., 1995; Coblentz et al., 1998; D'Agostino et al., 2014; Ghosh et al., 2009; Ghosh et al., 2006; Hast, 1969; McGarr and Gay, 1978; Müller et al., 1992; Naliboff et al., 2012; Rajabi et al., 2017a; Ranalli and Chandler, 1975; Richardson, 1992; Richardson et al., 1979; Sbar and Sykes, 1973; Sykes and Sbar, 1973; Zoback, 1992; Zoback and Zoback, 1980; Zoback et al., 1989). Hence, Zoback et al., (1989) and Richardson (1992) showed that the regional orientation of maximum horizontal stress (SHmax) in major continental plates, including Western Europe, North America and South America is sub-parallel to the absolute plate motion (APM)

4

ACCEPTED MANUSCRIPT direction, resulting in the general hypothesis that the present-day crustal stress field on the plate interior is primarily controlled by the same forces that control and drive tectonic plate motion. Since the first release of the WSM, there has been a dramatic increase in analysis of the presentday stress pattern at smaller scales, such as the wellbore, field and basin scales, due the numerous applications of present-day stress data in petroleum exploration and production (Barton and Moos,

PT

2010; Bell, 1996b; Tingay et al., 2005a; Zoback, 2007). Several studies have already demonstrated

RI

that the orientation of SHmax in a sedimentary basin can be consistent with plate tectonic forces, such as the Alberta Basin in Canada (Bell, 1996b; Bell and Gough, 1979; Reiter et al., 2014), or can be

SC

complex and perturbed due to basin geometry (King et al., 2009; Rajabi et al., 2016b; Tingay et al., 2005b; Yassir and Zerwer, 1997), detachment zones (Ask, 1997; Bell, 1996b; Brereton et al., 1991;

NU

Heidbach et al., 2007; Müller et al., 2010; Roth and Fleckenstein, 2001; Tingay et al., 2011; Tingay et

MA

al., 2012; Tingay et al., 2005a), topography and geological structures such as fractures, faults and geological contrasts (Bell, 1996b; Heidbach et al., 2007; Mount and Suppe, 1987; Pierdominici and Heidbach, 2012; Rajabi et al., 2016b; Rajabi et al., 2017b; Sahara et al., 2014; Seithel et al., 2015;

D

Tingay et al., 2010b; Yale, 2003; Zoback, 2007; Zoback et al., 1987). The detailed studies of stress in

PT E

basins during the second phase of the WSM project demonstrated that “plate boundary forces are not enough” to explain the state of stress at smaller scales (Heidbach et al., 2007; Heidbach et al., 2010;

CE

Sperner et al., 2003; Tingay et al., 2006; Tingay et al., 2005a). The old and fast-moving Australian continent is located in the Indo-Australia Plate which is a

AC

‘plate of extremes’ (Keep and Schellart, 2012) in which the Himalaya (the highest mountain range on Earth) is located in the northwestern margin of the plate, the Sunda subduction zones (one of the largest subduction zones) is the northeastern boundary of the Indo-Australian Plate, the TongaKermadec-Hikurangi subduction, one of the fastest subduction zones on Earth (Bevis et al., 1995), is the eastern boundary of the plate (Figure 1). The present-day crustal stress pattern in Australia is unique amongst major tectonic plates as being the only major plate in which the present-day SHmax orientation is not broadly aligned to absolute plate motion and it shows significant rotations of the SHmax stress pattern on regional scale (Hillis and Reynolds, 2000; Hillis and Reynolds, 2003; 5

ACCEPTED MANUSCRIPT Richardson, 1992; Zoback et al., 1989). This significant paradox has motivated numerous researchers to examine the origin and state of the crustal stress pattern in stable continental Australia for over 20 years (Burbidge, 2004; Cloetingh and Wortel, 1986; Coblentz et al., 1995; Coblentz et al., 1998; Dyksterhuis et al., 2005a; Müller et al., 2012; Reynolds et al., 2002; Sandiford et al., 2004). The crustal stress pattern of Australia has been examined from a wide range of different perspectives,

PT

including stress data compilation (Denham and Windsor, 1991; Hillis et al., 1999; Hillis et al., 1998; Hillis and Reynolds, 2000; Hillis and Reynolds, 2003; Mildren and Hillis, 2000; Mildren et al., 1994;

RI

Rajabi and Tingay, 2016; Rajabi et al., 2016b; Rajabi et al., 2016c; Rajabi et al., 2015; Rajabi et al.,

SC

2017b; Worotnicki and Denham, 1976), geomechanical-numerical modelling (Burbidge, 2004; Cloetingh and Wortel, 1985; Coblentz et al., 1995; Coblentz et al., 1998; Dyksterhuis et al., 2005a;

NU

Hillis et al., 1997b; Müller et al., 2012; Reynolds et al., 2002; Sandiford et al., 2004) and detailed analysis of neotectonic features (Clark et al., 2012; Hillis et al., 2008; Quigley et al., 2010; Quigley et

MA

al., 2006; Sandiford, 2003; Sandiford et al., 2004). Furthermore, recent high resolution stress orientation datasets compiled for eastern Australian basins have revealed that these regions contain

D

numerous and striking stress perturbations from the basin to the wellbore scales in association with

PT E

different geological structures (Brooke-Barnett et al., 2015; Rajabi et al., 2016b; Rajabi et al., 2017b). As such, continental Australia provides a natural laboratory to examine the controls of stress fields at all scales, from the plate scale (100s-1000s of km) to the wellbore scale (1-100 m). These stress field

CE

variations are now clearly identified due to the significant increase of stress data in Australia

AC

presented in this work.

This paper is organised in nine sections. In the first section, we explain the tectonic setting of Australia. We then review 25 years of studies on the crustal stress field in the Australian continent from different perspectives and highlight what we have learnt and what questions still remained to be investigated. In the fourth and fifth sections, we explain the new release of the ASM, in terms of SHmax orientation and tectonic stress regime. This is the first update of the ASM in 13 years, and documents an increase in the stress data in the continent from 594 to 2150 data records, as well as an increase in the number of defined ‘stress provinces’ to 30 (from 16 provinces in 2003). We also compare and

6

ACCEPTED MANUSCRIPT show the consistency of the Australian present-day crustal stress pattern with neotectonic features across the continent. Our extensive analysis of stress with high resolution data documents the significant stress perturbations in the Indo-Australian Plate at different scales. Hence, we compare the regional crustal stress pattern of Australia with published Australian stress models and discuss the stress patterns/sources of Australia from continental to wellbore scales. We highlight that the

PT

observed stress at any point results from the superposition of all forces acting from plate-wide to local scale. Finally, in the last part of the paper we explain the application of our newly compiled data in

SC

RI

different earth-science disciplines.

NU

2. Tectonic setting, neotectonics and seismicity of the Australian continent

MA

The tectonically and geologically complex Indo-Australian Plate (IAP) represents an ideal natural laboratory to investigate the interaction of various plate boundary forces in the pattern of SHmax orientation (Coblentz et al., 1998; Reynolds et al., 2002). It contains many different types of plate

D

boundaries (Figure 1); from active spreading centres to the south and west (mid-ocean ridge systems

PT E

that separates the IAP from the Antarctic and African Plate, respectively), to different continental collisional boundaries and subduction zones in the east and north (Coblentz et al., 1998; Reynolds et

CE

al., 2002). Australia, which is entirely located within the IAP, is often considered to be a tectonically and seismically stable continent, dominated by ancient flat landscapes, located far away from plate

AC

boundaries and having comparatively low earthquake occurrence rates (Johnston et al., 1994; Schulte and Mooney, 2005). Yet, this common perception of a tectonically quiescent Australia is in stark contrast to the widespread observations of seismicity (Figure 2) and major neotectonic structures documented in the last 10 years (Clark et al., 2014; Clark et al., 2015; Clark et al., 2012; Hillis et al., 2008; Quigley et al., 2006; Sandiford, 2003; Sandiford et al., 2004). Large inversion related features and significant late Cenozoic uplift have been observed along Australia’s southern, northern and north-western margins (Hillis et al., 2008; Holford et al., 2011; Sandiford, 2003; Tassone et al., 2014; Tassone et al., 2012). For example, over ten inversion related anticlines are identified in the offshore 7

ACCEPTED MANUSCRIPT Otway Basin, some of which witness up to one kilometre of exhumation since the Miocene-Pliocene (Holford et al., 2011). The Mount Lofty and Flinders Ranges contain numerous outcrop examples of recently active faults, such as the Milendella Fault near Cambrai that displays Cambrian rocks thrust over Quaternary sediments (Hillis et al., 2008; Rajabi et al., 2016c; Sandiford, 2003). The earthquake epicentres of Australia (Figure 2) show that four seismic zones can be defined

PT

across the continent (Hillis et al., 2008). Australia, regarded as a stable continental region, contains

RI

numerous neotectonic (late Neogene to Quaternary) faulting record (Clark et al., 2014; Clark et al., 2012; Quigley et al., 2010). The analysis of recently active faults in southeastern Australia suggests

SC

that earthquakes with M > 7 are possible, however, with a very low occurrence probability on short time scales due to the low seismic strain rates of 10-16 to 10-17 s-1 (Sandiford et al., 2004). Leonard and

NU

Clark (2011) also show that most of fault scarps in Western Australia indicate that seismicity with

MA

magnitudes of >6.5 has occurred in the last 100 ka. Indeed, large earthquakes are commonly observed in regions, both in Australia and worldwide, that displayed little prior seismicity. For example, the 1988 magnitude 6.3-6.7 events in Tennant Creek (Northern Territory, Australia), the 1941 magnitude

D

7.1 Meeberrie event (Western Australia), the 1906 magnitude 7.2 in offshore Geraldton, Western

PT E

Australia (Leonard, 2008; McCue, 1990; McCue, 2014; McEwin et al., 1976) and the 1811-1812 magnitude 7.0-7.5 events that struck New Madrid, in the central Mississippi River Valley, in eastern

CE

United States (Hough et al., 2000). Hence, it is possible that regions of Australia near recently active faults, such as Adelaide, Newcastle and Perth, may be at greater seismic hazard than currently

AC

recognized.

3. An overview of earlier studies on the crustal stress pattern of Australia 3.1. Early studies before the initiation of the Australian Stress Map project (pre-1996)

The first measurement of in-situ stress in the Australian continent was carried out in an underground power project in eastern Australia by the flat jack method (Alexander and Worotnicki, 1957; Worotnicki and Denham, 1976). The flat jack method (since 1957) and overcoring method 8

ACCEPTED MANUSCRIPT (since 1963) were the main tools to determine the in-situ stress at engineering sites, and were the first data used to broadly examine stress in the Australian continent (Worotnicki and Denham, 1976). In one of the earliest studies on the global stress pattern by Ranalli and Chandler (1975), Australia was represented by only four stress-relief measurements, which was then increased to eight by Richardson et al. (1979).

PT

Along with these measurements at engineering sites, which provide stress information at

RI

shallower parts of the crust, many researchers analysed earthquake focal mechanism solutions at various locations to infer the relative magnitudes and orientation of tectonic stresses in the deeper

SC

parts of the Australian crust (Bock and Denham, 1983; Cleary, 1962; Cleary et al., 1964; Denham, 1980; Denham et al., 1979; Denham et al., 1980; Denham et al., 1982; Denham et al., 1985; Denham

NU

et al., 1981; Everingham and Smith, 1979; Fitch, 1976; Fitch et al., 1973; Fredrich et al., 1988;

MA

Greenhalgh et al., 1994; Greenhalgh and Singh, 1988; Gregson and Denham, 1987; Lambeck et al., 1984; McCaffrey, 1989; McCue et al., 1990a; McCue et al., 1990b; McCue, 1989; McCue and Sutton, 1979). The majority of in-situ stress information in the Australian continent prior to 1990 was either

D

from shallow engineering methods focal mechanism solutions of earthquakes at greater depth

PT E

(Denham et al., 1979; Lambeck et al., 1984; Worotnicki and Denham, 1976). The first extensive study on the orientation of SHmax in the Australian continent based on drilling-related stress indicators was

CE

carried out by Blümling (1986) using the interpretation of borehole breakouts in 22 wells across Australia. The petroleum industry of Australia then was introduced to the link between present-day

AC

stress orientations and the occurrence of ‘borehole breakouts’ (Denham and Windsor, 1991; Hillis, 1991; Kingsborough et al., 1991; Lowry, 1990) and, since then, drilling-related stress indicators have been the predominant source of crustal stress information in Australia (Denham and Windsor, 1991; Hillis et al., 1999; Hillis and Reynolds, 2000; Hillis and Reynolds, 2003; Mildren and Hillis, 2000; Rajabi et al., 2016b). In the first compilation of the World Stress Map (WSM) project, initiated in 1986, the contemporary state of stress in the Australian continent was poorly understood (Zoback, 1992; Zoback et al., 1989). Whilst the first release of the WSM contained large datasets for many major continental areas, particularly North America and Europe, this stress compilation only

9

ACCEPTED MANUSCRIPT contained 95 reliable stress data records in Australia, suggesting a highly variable stress pattern (Figure 3a), (Zoback, 1992; Zoback et al., 1989).

3.2. Early phase of the Australian Stress Map project (1996-2003)

PT

Even the small compilation of crustal stress data for Australia in the first release of the WSM

RI

revealed a complex stress pattern for the continent than mid-plate North America or Western Europe. The Australian Stress Map (ASM) project was initiated in 1996 to compile the SHmax orientation from

SC

different stress indicators and build a public available stress database to examine the origin and

NU

implications of the present-day tectonic stress pattern in Australia. Studies were primarily investigating the enigmatic crustal stress pattern of the Australian continent, with a particular

MA

emphasis on the petroleum basins in North West Shelf of Western Australia and the Otway Basin of South Australia (Hillis et al., 1997a; Hillis et al., 1995; Hillis and Williams, 1993a; Hillis and

D

Williams, 1993b; Mildren et al., 1994).

PT E

The first release of the ASM, published by Hillis et al. (1998), contained 357 SHmax data records with new data from Bonaparte, Carnarvon, Amadeus, Cooper-Eromanga and Otway basins (Figure 3b). Hillis et al. (1999) compiled extensive stress data from coal mines in eastern Australia, based on

CE

hydraulic fracturing tests and overcoring measurements. These data revealed a prevailing thrust

AC

faulting stress regime in the Bowen and Sydney basins and showed a N-S SHmax trend in the Bowen Basin and a NE-SW SHmax trend (with significant deviations) in the Sydney Basin. Further studies in the Perth Basin of Western Australia by Reynolds and Hillis (2000), resulted in the second release of the ASM project by Hillis and Reynolds (2000) and further documented the regional variability of the SHmax orientation across the Australian continent (Figure 3c). Extensive data compilation based on several studies (Clark and Leonard, 2003; Leonard et al., 2002; Mildren and Hillis, 2000; Mildren et al., 2002; Reynolds, 2001) resulted the most recent prior release of the ASM project, containing 594 data records (Figure 3d) (Hillis and Reynolds (2003). The 2003 version of ASM, which was the most comprehensive stress map of Australia prior to this study, revealed a heterogeneous stress pattern in 10

ACCEPTED MANUSCRIPT Australia where the SHmax orientation in north and northwestern Australia is NE-SW to N-S, and SHmax is E-W in the Perth and Carnarvon basins of Western Australia. A regional NW-SE SHmax is observed in southeastern Australia and a ‘horse-shoe shape’ in northeastern Australia, including E-W orientation of SHmax in the Cooper Basin and N-S SHmax trend in the Bowen and Amadeus basins (Figure 3d).

PT

During the second phase of the WSM project by Heidbach et al., (2008b; 2010) the Australian

RI

continent was not investigated significantly (Figure 3e), however, some additional data records for Australia were included from published studies in the Gippsland, Cooper and Perth basins (King et

SC

al., 2008; Nelson et al., 2006; Reynolds et al., 2005). Along with the WSM and ASM projects, numerous researchers have studied the enigmatic pattern of the present-day crustal stress in Australia

NU

by using geomechanical-numerical models to investigate the principal crustal stress sources, and to

MA

predict the stress pattern in those parts of Australia where there is no, or sparse, stress information (Cloetingh and Wortel, 1985; Cloetingh and Wortel, 1986; Coblentz et al., 1995; Coblentz et al., 1998; Dyksterhuis et al., 2005a; Dyksterhuis et al., 2005b; Hillis et al., 1997b; Müller et al., 2012;

D

Rajabi et al., 2017a; Reynolds et al., 2002; Reynolds et al., 2003; Zhao and Müller, 2003). Despite the

PT E

complex stress field observed throughout Australia, the results of these studies suggested that the complex plate boundary forces around the IAP control, at the first-order, the intraplate SHmax

AC

CE

orientation (see more discussion in section 7).

3.3. New phase of the ASM project (2012-present)

Despite the extensive and widespread stress analysis undertaken in the early phases of the ASM project, the stress field for many parts of Australia, and particularly eastern and South Australia, remained poorly resolved, with the dataset primarily comprised of shallow engineering test data and predictions based on numerous numerical geomechanical models (Hillis and Reynolds, 2003; Reynolds et al., 2002). Whilst many other major continental areas had an extensive set of petroleum, earthquake and engineering data, one of the most significant obstacles in understanding the state of 11

ACCEPTED MANUSCRIPT crustal stress in the Australian continent is that previously compiled stress data was largely confined to just ten mature petroleum provinces. Hence, the ASM was largely composed of high-density stress data in ten sedimentary basins, but with large gaps of almost no stress data throughout most of the continent (Figure 3d and 3e). The recent increase in unconventional petroleum reservoirs and geothermal exploration in Australia has resulted in a greatly increased amount of new data for stress

PT

analysis in previously unstudied or poorly-constrained areas in Northern Territory, Queensland, New South Wales and South Australia. Furthermore, the use of petroleum technologies in mineral wells in

RI

areas including South Australia and central Victoria, thus providing opportunities to investigate, re-

SC

assess and update the present-day stress pattern throughout far more of the Australian continent. In the past four years, we compiled and analysed an extensive new stress dataset from all these different

NU

methods across the Australian continent to document the stress pattern of Australia in vastly improved

MA

detail.

D

4. Stress indicator used in the Australian Stress Map database

PT E

There are a variety of methods that provide the information of crustal present-day stress from specific depth ranges (Bell, 1996a; Engelder, 1993; Schmitt et al., 2012; Zang and Stephansson, 2010;

CE

Zoback, 2007). We compiled and interpreted the in-situ stresses from five well-known methods suggested by the WSM project (Heidbach et al., 2010; Sperner et al., 2003; Zoback and Zoback, 1991;

AC

Zoback, 1992; Zoback et al., 1989) to investigate the orientation of SHmax from the near surface down to 40 km depth in the lithosphere in the Australian continent. In order to assess the data quality and to guarantee comparability of different stress indicators, each data record is ranked according to the WSM quality ranking system from A to E quality (Heidbach et al., 2010). In this worldwide accepted ranking system, the A-quality is the most accurate data records (within ±15°) while the E-quality is the least reliable data records. Actually, the E-quality is used when no stress information can be determined or the results have a standard deviation of >±40° (Heidbach et al., 2010).

12

ACCEPTED MANUSCRIPT In addition, wherever possible, the standard Andersonian stress regime classification is assigned to classify the stress state as being either normal (Sv>SHmax>Shmin), strike-slip (SHmax>Sv>Shmin) and thrust (SHmax>Shmin>Sv) faulting stress regime (Anderson, 1905; Zoback, 1992). The methods used for determination of the present-day crustal stress field in Australia are briefly summarised below, with an emphasis on the drilling-related stress indicators that compose the majority of new stress data

SC

4.1. Wellbore breakouts and drilling-induced tensile fractures

RI

PT

interpreted herein.

NU

Drilling-related stress indicators are considered as well-established methods that provide reliable in-situ stress information from the upper five km of the Earth’s lithosphere (Bell, 1996a; Schmitt et

MA

al., 2012; Zoback, 2007). Drilling-induced tensile fractures (DIFs) and borehole breakouts are common stress-induced failures of the wellbore wall and form due to the concentration of stress that develops around a borehole as it is drilled (Bell, 1996a; Bradley, 1979; Kirsch, 1898; Schmitt et al.,

D

2012). Borehole breakouts are ‘ovalisation’ of borehole and form due to compressive shear failure of

PT E

the borehole wall (Bell and Gough, 1979; Kirsch, 1898). In vertical wells, breakouts form parallel to the Shmin orientation and can be interpreted by oriented-caliper and borehole image logs (Bell, 1990;

CE

Bell and Gough, 1979; Plumb and Hickman, 1985; Rajabi et al., 2010; Schmitt et al., 2012; Zoback, 2007). DIFs form parallel to the SHmax orientation when hoop stress is less than tensile rock strength

AC

and can only be observed via borehole image logs (Aadnoy, 1990; Aadnoy and Bell, 1998; Bell, 1996a; McCallum and Bell, 1994; McCallum and Bell, 1995; Schmitt et al., 2012; Zoback, 2007). In image logs in sub-vertical wells (<10° deviation), breakouts are characterised as a pair of elongated zones on opposite side of the wellbore wall while DIFs appear as a pair of narrow vertical fractures (Figure 4). In this study, various types of wellbore image logs (resistivity, acoustic and density) are interpreted to directly interpret borehole breakouts and DIFs, based on the WSM analysis guidelines and quality ranking system (Heidbach et al., 2010).

13

ACCEPTED MANUSCRIPT 4.2. Focal mechanism solution of earthquakes

Earthquakes that are observed from a sufficiently high number of stations with a reasonable azimuthal distribution can be used to determine a focal mechanisms solution (FMS) and infer the orientation of SHmax and the stress regime (Heidbach et al., 2010; McKenzie, 1969; Sbar and Sykes,

stress data at depths > 5 km (Heidbach et al., 2010; Zoback, 1992).

PT

1973; Zoback, 1992). Indeed, FMS data records are usually the only stress indicator that provides

RI

According to the WSM project, three types of stress data records inferred from focal mechanism

SC

solutions of earthquakes are available in the ASM database. Single focal mechanism solutions (FMS) estimate the SHmax orientation using three principal strain axes of the focal mechanism as a proxy to

NU

estimate the SHmax orientation (Barth et al., 2016; McKenzie, 1969; Raleigh et al., 1972; Zoback, 1992). However, they do not receive a WSM quality better than C (±25o) due to ambiguity of the two

MA

perpendicular nodal planes of the focal mechanism (Barth et al., 2016; Célérier et al., 2012; McKenzie, 1969). Composite or average focal mechanism solutions (FMA) combine several FMSs

D

and generally do not receive the WSM quality of better than D (Heidbach et al., 2010). Formal stress

PT E

inversion of focal mechanism solutions (FMF) minimise the difference between the maximum shear stress and the slip direction of earthquakes (Bott, 1959; Heidbach et al., 2010), which can provide better quality data records (A and B) according to the WSM quality ranking criteria (Barth et al.,

AC

CE

2016; Heidbach et al., 2010).

4.3. Hydraulic fracturing measurements and overcoring

Shallow present-day stress information is often obtained in mines to assess pillar and slope stability (Amadei and Stephansson, 1997; Leeman, 1964; Zang and Stephansson, 2010). This information is typically collected through two well-established methods: hydraulic fracturing (HF) and overcoring (Enever et al., 1984; Engelder, 1993; Haimson and Cornet, 2003; Hubbert and Willis, 1957; Leeman, 1964; McGarr and Gay, 1978; Sjöberg et al., 2003). Overcoring is a technique to

14

ACCEPTED MANUSCRIPT examine the components of the reduced stress tensor based on the strain relief of the rock mass (Amadei and Stephansson, 1997; Engelder, 1993; Sjöberg et al., 2003). The overcoring method measures the expansion of cored rock by a strain gauge after it is removed from subsurface stresses, then the measured strain is converted to estimated stresses by using the mechanical properties of the rock sample (Amadei and Stephansson, 1997; Sjöberg et al., 2003). The overcoring technique needs

PT

to be implemented close to free surfaces, and hence provide (local rather than regional) stress

RI

information from the shallower part of the Earth’s crust (Zoback, 1992).

Hydraulic fracture tests provide in-situ stress information at greater depths in comparison with

SC

overcoring technique. This method involves isolating a short section of the wellbore between rubber packers and increasing the hydraulic pressure within the isolated interval until a small tensile fracture

NU

is initiated, typically parallel to the SHmax (Bell, 1996a; Hubbert and Willis, 1957). The orientation of

MA

the induced hydraulic fracture needs to be observed and interpreted by running oriented impression packer or wellbore image log over the fractured section (Amadei and Stephansson, 1997; Bell, 1996a; Bell and Babcock, 1986; Blomberg, 1971; Paillet, 1985). It should be noted that shallow stress data

D

determined from overcoring and HF techniques must be carefully analysed, as they can be strongly

CE

PT E

affected by topography effects (Bell, 1996a; Zoback, 1992).

AC

5. The new Australian Stress Map database (release-2016)

The last prior release of the ASM by Hillis and Reynolds (2003) compiled 549 (A-E quality) SHmax data records (Figure 3d) and successfully constrained the SHmax orientation in 16 stress provinces through Australia and Papua New Guinea. However, the pattern of stress in many areas of Australia, particularly eastern Australian basins and many parts of South Australia were not investigated extensively due to lack of data. In the past four years, we conducted significant stress analysis throughout the continent to study the present-day crustal stress pattern of Australia.

15

ACCEPTED MANUSCRIPT In this latest phase of the ASM project, we mainly focused on the interpretation of borehole image and oriented-caliper logs in major active petroleum basins of Australia, with a particular emphasis on poorly constrained regions such as recent unconventional and conventional petroleum, geothermal and mineral wells in eastern Australia, South Australia and the Northern Territory. We conducted extensive analysis of borehole image and oriented-caliper logs and provide 735 stress data

PT

records in various parts of New South Wales (Darling, Sydney, Clarence-Moreton, Gunnedah, Bowen-Surat and Gloucester basins), Queensland (Cooper-Eromanga Galilee, Surat and Bowen

RI

basins), South Australia (Officer, Cooper and Eromanga basins and mineral/geothermal wells),

Tasmania (Gippsland, Otway, Bass and Sorell basins).

SC

Western Australia (Browse Basin), Northern Territory (McArthur and Pedirka basins), Victoria and

NU

In addition, we compiled literature data from analysis of borehole image logs in the Perth and

MA

Carnarvon basins of Western Australia (Bailey et al., 2014; Bailey et al., 2016; Jepson, 2014), Gippsland Basin of Victoria (Swierczek et al., 2015), Surat Basin of Queensland (Brooke-Barnett et al., 2015), as well as recent analysis of stress in other regions of Australia (Klee et al., 2011;

D

MacGregor, 2002; MacGregor, 2003; MacGregor and Gale, 2000; Rasouli et al., 2013; Tassone et al.,

PT E

2014). Hence, our compilation provides 939 new present-day stress data records from wellbore information. In order to investigate the present-day stress in other regions of the continent, we

CE

extensively compiled all earthquake focal mechanism solutions in and around Australia from different sources including Global CMT (Dziewonski et al., 1981; Ekström et al., 2012), GEOFON (GEOFON

AC

Data Centre, 1993), SLU (SLU Earthquake Center, 2015) and USGS (USGS, 2015), as well as all related published papers (Balfour et al., 2015; Leonard et al., 2002; Revets et al., 2009; Sippl et al., 2015). In total, we added 521 new FMS data records and one new FMF data record. The total number of stress data in the ASM database release 2016 has been increased from 549 A-E quality data records from the 2003 release to 2150 A-E quality data records (Figures 3f and 5). The new ASM database contains 1359 (A-E quality) SHmax data records from borehole stress indicators (breakouts and DIFs), 650 (A-E quality) data records from earthquake data (FMS; FMF and FMA), 139 (A-E quality) records from engineering methods (HF and overcoring) and two E-quality 16

ACCEPTED MANUSCRIPT SHmax orientations inferred from geological information. The details of the ASM database can be found in Figures 5 and 6, Table 1 and supplementary online information. The majority of the newly compiled ASM database (except for 10 data records) is fully integrated into the 2016 release of the WSM project (Heidbach et al., 2016a; Heidbach et al., 2016b). The wellbore stress indicators (breakouts and DIFs) provide the majority of intraplate stress

PT

information while earthquake focal mechanism solutions provide stress indicators near plate

RI

boundaries (Figure 5). Our knowledge about the crustal stress in the continental Australia at shallower depths comes from overcoring, HF and drilling-related stress data records (in shallow coal seam gas

SC

wells), while breakouts and DIF provide stress information at intermediate depths < 5 km. Focal mechanism solutions of earthquakes are the only sources of information at any depths > 5 km in the

NU

continent (Figure 6).

MA

In addition to the SHmax orientation information, almost 37% of the data records in the 2016 ASM database were assigned a tectonic stress regime (Figure 6) according to the Andersonian classification

D

(Anderson, 1905) and the stress regime categorization of Zoback (1992). This classification enables

PT E

us to assess the variability of the relative stress magnitudes, i.e. the stress regime, with depth. In order to analyse the regional pattern of stress at the continental scale, two statistical

CE

approaches were used. The first one defines so-called stress provinces and the second one plots a smoothed stress pattern either using stress trajectories or mean SHmax orientations on a regular grid.

AC

The stress trajectories were used in the early phases of the ASM project by Hillis and Reynolds (2000) and thus facilitate comparison between the prior and current database.

5.1. Analysis of stress provinces

The concept of stress provinces in the ASM project (Hillis and Reynolds, 2000) is based on directional statistics and the Rayleigh test, which is a well-known technique to determine the dispersion and significance of directional vectors (Davis, 2002; Mardia, 1972). The use of Rayleigh 17

ACCEPTED MANUSCRIPT test to evaluate whether the stress data are distributed systematically or randomly was first used by Coblentz and Richardson (1995) to quantify the global pattern of the present-day stress based on the first release of the WSM project. A minimum of four A-C quality SHmax data records in a close geographic region (mainly a basin) is required to define a stress province. All SHmax data records within a province are quality-weighted. The most reliable A-quality SHmax data records receive a

PT

weight of ‘four’, down to D-quality data records that receive a weight of ‘one’. Finally, the mean SHmax orientation and the mean resultant vector length (R-value) are determined by a directional

RI

statistics method at each stress province (Davis, 2002; Mardia, 1972). The Rayleigh test was used to

SC

test the randomness of the mean SHmax azimuth at each stress province. We consider the null hypothesis that ‘SHmax orientations in a province are random’. Six types of stress province

NU

classifications are used (Hillis and Reynolds, 2003) where a type-1 shows that the ‘null hypothesis’ can be rejected at the 99.9%, type-2 of province at the 99% , type-3 of province at the 97.5% level,

MA

type-4 of province at the 95% level, and type-5 of province at the 90% level. When the ‘null hypothesis’ cannot be rejected at the 90% confidence level the province is represented by type-6 and

D

is considered as a high standard deviation for the stress data.

PT E

In this study, we update all previously defined 16 stress provinces and derive 14 new stress provinces in different parts of the Australian continent. Among the new provinces, nine provinces are

CE

located in eastern Australia (Queensland, New South Wales and Victoria), two provinces in South Australia, one in the Northern Territory and two in Western Australia. The details and locations of the

AC

stress provinces are shown in Table 2 and Figures 7 and 8. Analysis of stress provinces throughout the Australian continent reveals four major S Hmax orientations (Figure 8). The mean SHmax orientation in the northern and northeastern Australia, including the Papua New Guinea, MacArthur Basin region, Bonaparte, Canning and some parts of the Bowen-Surat basins, is approximately NE-SW. A prevailing E-W SHmax orientation exists in the Browse Basin in the North West Shelf, almost all of Western Australia (Carnarvon and Perth basins), and most parts of South Australia (Flinders Ranges, Officer and Cooper-Eromanga basins). The mean orientation of SHmax in most parts of eastern Australia is approximately ENE-WSW (Figure 8) in 18

ACCEPTED MANUSCRIPT different basins (including the Sydney, Clarence-Moreton, Gunnedah, Surat and Bowen basins), and SHmax rotates to approximately NW-SE in southeastern Australia (Figure 8) within the Gippsland, Otway and eastern Bight basins. The statistical analysis of the stress provinces shows that the mean SHmax orientation in 25 provinces is of type 1 or 2, and thus considered highly reliable (Table 2 and Figure 8). The majority of stress provinces in the western half of the continent, as well as in

PT

southeastern Australia show low standard deviations (Table 2), while most of stress provinces in eastern Australia show higher standard deviation indicating a higher variability of the S Hmax

RI

orientation in these regions (Brooke-Barnett et al., 2015; Hillis et al., 1999; Rajabi et al., 2016b;

SC

Rajabi et al., 2017b). In addition, the Southern Carnarvon (s.d.: 43o) and SE seismogenic zone (s.d.: 41o), which are based on focal mechanism solutions of earthquakes, all show significant deviations of

NU

the SHmax orientation that are probably due to the inherent uncertainty with SHmax determined from the FMS method (Figures 7 and 8) or due to neotectonic activity that represent localised stress

D

MA

perturbations.

PT E

5.2. Analysis of the SHmax stress pattern

The concept of stress provinces determines the mean SHmax orientation and its consistency in a

CE

particular geographic area. However, this method does not examine or predict the stress pattern throughout the entire continent. In order to estimate the long wavelength stress pattern on a regular

AC

grid and filter the effect of local stresses perturbations, we used the smoothing algorithm of Müller et al. (2003) that is based on the approach of Hansen and Mount (1990). The stress trajectories method aims to show the mean SHmax orientation at each point on the trajectories. Different weighting terms are used in this algorithm, including quality weight, distance weight and robustness weight (Hansen and Mount, 1990; Hillis and Reynolds, 2003; Müller et al., 2003). We incorporated the A-C quality data, with a weight of 1 for A, 0.75 for B and 0.5 for C-quality, to calculate stress trajectories. Herein, we applied the distance weighted algorithm following the method developed by Müller et al. (2003), including a fixed number of nearest neighbours (FNN) and fixed search radius (FSR). The FNN 19

ACCEPTED MANUSCRIPT method is based on data density in the area of interest and uses local weights, while the FSR algorithm uses a global weight that is dependent on the Euclidean distance. More details about the smoothing methods used in this study can be found in Hansen and Mount (1990) and Müller et al. (2003). Figure 9 shows the stress trajectories and smoothed stress pattern of the continent for a search radius of 750 km and a minimum number of three stress data records with A-C quality within the search radius.

PT

The new trajectory stress map of Australia is significantly different from that was calculated by Hillis and Reynolds (2003), particularly in eastern and Western Australia where we compiled a significantly

SC

RI

amount of new stress data records (compare Figure 3d with Figure 3f).

NU

5.3. Present-day stress regimes throughout the Australian continent

795 data records in the 2016 ASM dataset also provide the information on the tectonic stress

MA

regime (relative stress magnitudes) based on the standard Anderson (1905) classification into thrust, strike-slip and normal faulting regimes (Figures 5 and 6). The tectonic stress regimes in the ASM

D

database are from focal mechanism solutions of earthquakes, shallow engineering measurements and,

PT E

to a lesser extent, from geomechanical data in petroleum wells (Figures 5 and 6) that indicate a prevailing thrust and strike-slip stress regime (Figure 6).

CE

The majority of the data records with stress regime information come from FMS data records that are clustered around plate boundaries. In order to investigate the intra-plate variability of the

AC

stress regime with depth, we filtered the ASM database and used only the A-D quality data records that are >100 km away from the nearest IAP boundaries and with depth < 5 km in the Australian lithosphere, as this constitutes the majority of the stress orientation data and is considered the depth range of most interest in petroleum, geothermal and mining industries. The resulting data set (i.e. 185 data records) represents the intra-plate present-day stress regime in the upper five kilometres of the continent. This dataset of stress regime information is derived from a variety of stress indicators, including hydraulic fracturing tests and overcoring measurements (111 data records mainly compiled from Hillis et al. (1999), Hillis and Reynolds (2003), Klee et al. (2011) and Worotnicki and Denham 20

ACCEPTED MANUSCRIPT (1976)), earthquake focal mechanism solutions (35 data records) and petroleum wellbore data (39 data records compiled from Brooke-Barnett et al. (2015), Jepson (2014) Rajabi et al. (2016a), Rasouli et al. (2013) and Reynolds et al. (2006)). The resulting stress regime dataset in the upper five kilometres of the Australian crust reveals a prevailing thrust faulting stress regime in the shallower part (< 1.5 km) of the crust that changes to

PT

strike-slip in the deeper part (Figure 10). The changes of stress regimes with depth have been reported

RI

in different sedimentary basins of Australia (Brooke-Barnett et al., 2015; Gui et al., 2016; Rajabi et

NU

SC

al., 2016a; Rasouli et al., 2013) and support the results highlighted by the ASM database.

6. Pattern of present-day crustal stress in the Australian continent

MA

6.1. Regional orientation of SHmax across Australia

In this study, different techniques have been implemented to determine and clarify the

D

continental-wide SHmax orientation. Both stress provinces and smoothing algorithms methods provide

PT E

consistent results for the regional SHmax orientation in continental Australia (Figures 8 and 9). The north of Australia (Bonaparte and Canning basins, McArthur Basin region, New Guinea and Irian Jaya stress provinces) is characterised by a NNE-SSW SHmax orientation which is rotates to E-W in

CE

most parts of Western Australia, including Browse, Carnarvon and Perth basins (Figure 8 and 9). The

AC

SHmax orientation in central Australia (Amadeus Basin) is NNE-SSW, which seems to be the transition between N-S SHmax orientation (in the northern parts of the Australian continent) and the E-W SHmax orientation to the east and west of the basin. However, it should be noted that there is a slight inconsistency between the SHmax orientation inferred from wellbore and earthquake data around the Amadeus Basin. The Cooper-Eromanga, eastern Officer and Darling basins and the Flinders Ranges, have an E-W SHmax orientation that changes to ENE-WSW in most of eastern Australia (Figures 8 and 9). Southeastern Australia (Gippsland, Otway, Bass and eastern Bight basins) exhibits a prevailing NW-SE SHmax orientation (Figures 8 and 9), as has been observed in previous studies, particularly in

21

ACCEPTED MANUSCRIPT Sandiford et al. (2004) who concluded that the NW-SE SHmax pattern of this region is a result of New Zealand continental collision.

6.2. Comparison of present-day crustal stress field and neotectonic features

PT

The Australian continent has a rich neotectonic record, despite the common perception that it is a

RI

highly stable continent (Clark et al., 2012; Geoscience Australia, 2016b; Hillis et al., 2008; Quigley et al., 2006; Sandiford, 2003). Figure 11 shows and compares young geological structures with the mean

SC

SHmax orientation (smoothed SHmax pattern) across the Australian continent. The majority of

NU

neotectonic structures in the continent show a thrust faulting stress regime, which is in agreement with the mean SHmax orientation. Even the variability of the regional mean SHmax orientation between stress

MA

provinces is consistent with neotectonic features and further confirms that the present-day stress pattern is a primary control on intraplate deformation throughout the continent (Clark et al., 2012; Hillis et al., 2008). The new database suggests that the tectonic stress regime in the upper five

D

kilometres depth of the Australian crust often exhibits a thrust faulting tectonic stress regime at the

PT E

shallowest upper parts (upper one km), but that this changes to a strike-slip stress regime in the deeper parts (Figure 10). The orientation of SHmax and prevailing thrust faulting stress regime in the

CE

shallowest parts of much of the Australian crust (i.e. upper one km) is in agreement with the neotectonic database of Australia (Figures 10 and 11). However, as indicated in Figures 10 and 11 the

AC

stress regime is depth-dependent, and the thrust faulting stress regime inferred from surface neotectonic features may not correlate with deeper parts in which a strike-slip regime is prevailing (Bailey et al., 2016; Rajabi et al., 2016a). Furthermore, a large number of petroleum industry stress studies have often suggested that a normal faulting stress regime is present in many Australian sedimentary basins (Bailey et al., 2012; Bailey et al., 2016; Berard et al., 2008; Jepson, 2014; King et al., 2012). Hence, there are potentially significant pitfalls in using shallow neotectonic features to help constrain the stress state at the depths of petroleum, geothermal or CO2 reservoirs. Changes in stress regime with depth have been reported in several case studies in Australia (Brooke-Barnett et al., 2015; 22

ACCEPTED MANUSCRIPT Gui et al., 2016; Klee et al., 2011; Rajabi et al., 2016a; Rasouli et al., 2013) and worldwide (Chang et al., 2010; Heidbach and Reinecker, 2013; Hergert and Heidbach, 2011a; Reis et al., 2013; Zang et al., 2012).

PT

7. Review on the potential sources of the crustal stress of the Australian continent

RI

7.1. An overview on the potential stress sources at different spatial scales

Regional stress patterns have commonly been analysed in order to examine major sources of

SC

crustal stress (Heidbach et al., 2007; Montone and Mariucci, 2016; Müller et al., 2010; Palano, 2015;

NU

Pierdominici and Heidbach, 2012; Rajabi et al., 2016d; Reiter et al., 2014; Tingay et al., 2010a; Ziegler et al., 2016a; Zoback, 1992; Zoback et al., 1989). These studies concluded that plate tectonic

MA

forces control the crustal stress field at the first-order (Figure 12). These studies also demonstrated that large intra-plate stress sources, such as lateral density changes in the crust (e.g. topography, coastal shelf) and buoyancy forces can regionally control the crustal stress at a second-order (Figure

D

12). In addition, numerous smaller scale studies, particularly in sedimentary basins, highlighted the

PT E

role of third-order sources of stress, such as geological structures and elastic property contrasts, in the crustal stress pattern (Figure 12). In the following sections the potential sources of crustal stress in the

AC

CE

Australian continent, at different scales, are summarized.

7.2. Australian crustal stress pattern and absolute plate motion

The first release of the WSM project by Zoback (1992), and further studies by Richardson (1992), revealed a poor correlation between the SHmax orientation and the direction of absolute plate motion (APM) in the IAP in general and the Australian continent in particular. However, the first release of the WSM contained only 95 reliable SHmax orientations for Australia. Now that the Australian continent has over 10 times more reliable data records, we re-visited the major observation of whether there is a correlation between the SHmax orientation from individual data records with A-C 23

ACCEPTED MANUSCRIPT and A-D quality (Figure 13b) and the APM azimuth. Figure 13 compares the orientation of SHmax orientation and the APM azimuth derived from the hotspot referenced global plate model HS3 NUVEL-1A (Gripp and Gordon, 2002). The results support the original observation of a poor correlation between APM and SHmax orientation in the Australian continent. Although there is a poor correlation between the SHmax orientation and the APM azimuth in the

PT

continent (Figure 13), numerous global and plate scale geomechanical-numerical modelling studies

RI

suggest that the long-wavelength stress pattern of the continent can be largely explained by the largescale driving mechanisms of plate tectonics (Bird et al., 2008; Coblentz et al., 1995; Coblentz et al.,

SC

2015; Coblentz et al., 1998; Ghosh and Holt, 2012; Ghosh et al., 2013; Heidbach et al., 2007; Lithgow-Bertelloni and Guynn, 2004; Lithgow-Bertelloni and Silver, 1998; Naliboff et al., 2009;

NU

Naliboff et al., 2012; Richardson, 1992; Steinberger et al., 2001; Wang et al., 2015; Zoback et al.,

MA

1989). However, in contrast to other major tectonic plates, where one or two plate boundary force types, acting in approximately one direction, provide the key stress sources, the Australian crust is

PT E

D

subjected to a range of complex and interacting plate boundary forces that act in different directions.

7.3. First-order control on the present-day crustal stress pattern

CE

As defined by the WSM project, first-order stress sources control the long wavelength (> 500

AC

km) crustal stress pattern (Figure 12). The first-order stress sources of the Australian continent, based on the published literature, can be discussed from two modelling perspectives: The plate-scale model approaches that focus on the crustal stress in the brittle part of the lithosphere, and global-scale models that focus on the investigation of mantle convection on the lithospheric stress distribution. i.

Plate scale models: These studies specifically aimed to explain the observed crustal stress pattern of Australia by examining the interactions of driving and resisting forces acting on the IAP. Over 20 published studies have investigated the present-day stress pattern of Australia using plate-scale geomechanical-numerical models. The majority of these models applied

24

ACCEPTED MANUSCRIPT different boundary conditions on the IAP, whilst assuming a homogeneous elastic rheology, and fitted the model predictions with observational data (Cloetingh and Wortel, 1985; Cloetingh and Wortel, 1986; Coblentz et al., 1995; Coblentz et al., 1998; Dyksterhuis et al., 2005a; Dyksterhuis et al., 2005b; Hillis et al., 1997b; Reynolds et al., 2002; Reynolds et al., 2003; Zhang et al., 1996). A few plate-scale studies divided the Australian continent in

PT

different geomechanical provinces (fold-belts, basins and cratons) and assigned individual elastic properties to these regions (Dyksterhuis et al., 2005a; Dyksterhuis et al., 2005b; Müller

RI

et al., 2012). Burbidge (2004) investigated the plate-scale stress pattern with a model that used

SC

a visco-elastic-plastic rheology using a 2.5D finite-element code (SHELLS, which was originally developed by Bird, 1999). In addition, other researchers investigated the effect of

NU

local structures (Zhang et al., 1998; Zhang et al., 1996; Zhao and Müller, 2001; Zhao and Müller, 2003), temporal effect and New Zealand continental collision on the pattern of stress

MA

in southeastern Australia (Sandiford et al., 2004). In almost all published models, the general NNE-SSW SHmax orientation in north and northwest of Australia, E-W orientation of SHmax in

D

Western Australia, the Cooper, Darling and Officer basins and Flinders Ranges, and the NW-

PT E

SE SHmax orientation in southeastern Australia have been explained by the interaction of different plate boundary forces acting on the IAP (Figure 14). However, there are significant deviations between the model predictions and first-order stress pattern, particularly in eastern

CE

Australia (Figure 14). Rajabi et al. (2016b) and Rajabi et al. (2017b) suggested two possible

AC

reasons for the inconsistencies between model predictions and the observed regional stress pattern, namely overly simple assumptions in the published geomechanical-numerical models and/or lack of model-independent data for model calibration. Nevertheless, the conclusion from these models is that plate boundary forces remain the most likely control on the firstorder crustal stress pattern in the IAP, and that issues with model fit can be largely resolved by imparting updated boundary conditions and improved model geometries and properties, such as the inhomogeneous distribution of density and strength (Rajabi et al., 2017a). ii.

Global convection models: Numerous global models have been used to investigate the stress state of the Earth’s lithosphere in the context of topography, thickness of lithosphere, plate 25

ACCEPTED MANUSCRIPT driving forces, density structures and gravitational potential energy (GPE) (Artyushkov, 1973; Ghosh et al., 2009; Koptev and Ershov, 2010; Molnar and Lyon-Caen, 1988; Naliboff et al., 2012; Richardson et al., 1976; Richardson et al., 1979; Solomon et al., 1975), mantle convection and joint modelling of lithospheric-mantle dynamics (Bai et al., 1992; Bird, 1998; Bird et al., 2008; Coltice et al., 2017; Ghosh and Holt, 2012; Ghosh et al., 2013; Ghosh et al.,

PT

2008; Lithgow-Bertelloni and Guynn, 2004; Naliboff et al., 2009; Steinberger et al., 2001; Wang et al., 2015). Among the global models, we could not find a single model that

RI

satisfactorily predicted the SHmax and stress regime across the Australian continent. However,

SC

the best-fit models of Ghosh et al. (2013) and Wang et al. (2015), where they predicted deviatoric stresses due to GPE, mantle tractions and a combination of both GPE and mantle

NU

traction, appear to provide results that are reasonably consistent with stress observations for large parts of the Australian continent. In particular, the deviatoric stress calculated from GPE

MA

differences, mainly due to the ridge push force, by Ghosh et al. (2013) fits the first-order stress pattern of eastern Australia, while the combined model (GPE and mantle traction)

D

presented by Ghosh et al. (2013) and Wang et al. (2015) provide a better fit for the western

PT E

side of the continent. More recently, the importance of GPE due to upper mantle geoid has been recognized as an important contribution to the lithospheric stress field in many intraplate regions (Coblentz et al., 2015). Becker et al. (2015), based on mantle flow modelling,

CE

proposed that dynamic topography variations associated with mantle flow are an important

AC

source of intraplate deformation and seismicity in intraplate regions, particularly in the Western United States. This is particularly important for the Australian continent where Sandiford (2007) and Quigley et al. (2010) proposed the role of dynamic topography and geoid in the long-wavelength (103 km) neotectonic deformation and continental tilting of the Australian continent. Despite the potential promising results obtained by mantle flow and GPE models, there have not, to date, been any published models examining the potential role of mantle flow and crustal thickness variations that have focussed specifically on the complex first-order Australian stress pattern. The global models discussed above are severely limited in terms of 26

ACCEPTED MANUSCRIPT resolution. Whilst plate-scale models commonly have lateral resolutions of 10-25km, the global mantle flow and GPE models have lateral resolutions ranging from 2-5° (220-540 km north-south resolution or at the equator). Thus a direct comparison with individual stress data that are affected from various sources can be misleading, but the general trend expressed in the mean SHmax orientation within the presented 30 stress provinces should be reflected.

PT

Hence, we suggest that more advanced mantle flow and GPE models may greatly aid in understanding the sources of stress in Australia and that it is feasible that the variable stress

RI

pattern observed in Australia is the result of a combination of different lithospheric and

NU

SC

mantle forces.

MA

7.4. Controls on the crustal stress pattern at second-, third- and fourth-order

As outlined above, most of the first-order Australian stress pattern is generally interpreted as being the result of large-scale plate tectonic processes. However, the stress pattern in most of eastern

D

Australia and the numerous perturbations of SHmax at different scales in several sedimentary basins are

PT E

difficult to explain purely by plate boundary forces or mantle flow and are generally considered to be the result of intra-plate stress sources (Figure 12). Numerous studies have observed localised present-

CE

day stress variations near different geological structures in eastern Australian sedimentary basins, and suggested these stress rotations are the result of geological structures, density or elastic rock property

AC

contrasts (Brooke-Barnett et al., 2015; Enever et al., 1999; Flottmann et al., 2014; Gale et al., 1984; Hillis et al., 1999; Rajabi et al., 2016b; Rajabi et al., 2017b). Second- and third-order of stress sources are particularly used to explain the stress pattern between and within sedimentary basins respectively, and are likely the reason why the Bowen, Surat, Gunnedah, Clarence-Moreton and Sydney basins of eastern Australia have high standard deviations in stress province analysis (Figure 8; Table 2). In addition to second and third-order of stresses, numerous highly localised stress perturbations are observed at the wellbore (1-1000m) scale in association with geological features such as lithological contrasts, faults and fractures (Figures 12 and 15). Hence, we suggest that fourth-order stress sources, 27

ACCEPTED MANUSCRIPT associated with localised elastic rock property contrasts and small-scale geological features, are responsible for these small scale variations (Figures 12 and 15). These wellbore scale stress variations only act to deviate the stress orientation at small scales, but have significant applications, particularly with regards to hydraulic fracture stimulation and fluid flow in fractured reservoirs in unconventional and geothermal exploration, as well as CO2 geo-sequestration (Figures 12 and 15). More details about

PT

the role and significance of the localised SHmax perturbations at wellbore scales can be found in Rajabi

SC

RI

et al. (2016b) and Rajabi et al. (2017b).

NU

7.5. Comparison between SHmax orientation and azimuthal seismic anisotropy

Present-day stress orientation is often correlated with the fast shear wave direction in crustal

MA

rocks and the mantle (Boness and Zoback, 2004; Crampin, 1991; Gavin and Lumley, 2016; Lockner and Stanchits, 2002). The seismic azimuthal anisotropy (SAA) of the Australian continent has been

D

investigated comprehensively using earthquake data (Debayle, 1999; Debayle et al., 2005; Debayle

PT E

and Kennett, 2000; Debayle and Kennett, 2003). These studies demonstrated the presence of two layers of seismic anisotropy beneath the continental Australia. The upper layer (shallower than 100 km) reveals variable direction of SAA similar to the SHmax pattern documented herein, whilst the

CE

lower layer (175-300 km) shows significant consistency between SAA and APM. The variable pattern in the upper SAA layer has been explained as ‘frozen deformation’ due to previous tectonic events

AC

(Debayle et al., 2005; Debayle and Kennett, 2000; Debayle and Kennett, 2003). Simons et al. (2002) claimed a poor correlation between large-scale geological trends and azimuthal anisotropy. However, due to the similarity between the orientation of SAA in the upper layer and the SHmax in the Australian continent, Hillis and Reynolds (2003) highlighted the role of in-situ stress associated with contemporary plate dynamics in the complex upper SAA pattern. More recently, Gavin and Lumley (2016) analysed SAA at borehole scales in 34 wells in the North West Shelf of Australia and showed significant consistency between SHmax orientation and SAA. Hence, Gavin and Lumley (2016) suggested the role of differential horizontal stress as the causes of shallow (< 5 km) SAA in this 28

ACCEPTED MANUSCRIPT region. Therefore, as highlighted by Hillis and Reynolds (2003), we also suggest that in-situ stresses and plate dynamics are a key control on the observed SSA pattern in the Australian lithosphere and upper asthenosphere.

PT

8. Applications of the Australian Stress Map database

RI

Crustal stresses are important to understand natural processes and to manage the safe and sustainable usage of the underground during storage and exploitation (Figure 12). Constraining the

SC

contemporary state of stress is significant for a wide range of earth science disciplines, ranging from

i.

NU

the tectonic plate-scale to reservoir and wellbore scales as following.

Knowledge of the crustal stress is essential for understanding the forces that drive tectonic

Richardson, 1992; Zoback, 1992).

Stress and its link to deformation play a critical role in neotectonic evolution and intraplate

D

ii.

MA

plates (Heidbach et al., 2008a; Heidbach et al., 2010; Hergert and Heidbach, 2011b;

PT E

deformation (Heidbach et al., 2010; Hillis et al., 2008; Liu and Stein, 2016; Sandiford, 2003; Sandiford et al., 2004). iii.

Earthquakes and stress are interconnected, and earthquake hazard is a function of subsurface

CE

stresses and rock failure criteria (Heidbach and Ben-Avraham, 2007; Holdgate et al., 2003;

AC

Liu and Stein, 2016; Sandiford et al., 2004; Seeber and Armbruster, 2000; Sibson et al., 2012; Stein, 1999). This is particularly important in zones with low seismic recurrence rates, such as Australia. iv.

In-situ stresses can be used to produce fault slip tendency maps and these have implications for seismic hazard and for recognising whether paleo-hydrocarbon accumulations may have been breached due to fault reactivation (Ferrill et al., 1999; Tingay et al., 2009; Ziegler et al., 2016b).

29

ACCEPTED MANUSCRIPT v.

In-situ stress data is of key importance to calibrate 3D geomechanical-numerical models in order to predict the stress state before any underground usage (Fischer and Henk, 2013; Hergert et al., 2015; Rajabi et al., 2017a; Reiter and Heidbach, 2014; Ziegler et al., 2016b).

vi.

In-situ stress controls the stability of mines, tunnels and boreholes (Reinecker et al., 2006; Zang et al., 2012; Ziegler et al., 2015; Zoback, 2007). Fluid flow in subsurface aquifers, oil reservoirs, geothermal fields (Heffer et al., 1997; Heffer

PT

vii.

and Lean, 1991), as well as the formation of some mineral deposits, is controlled by in-situ

The orientation of hydraulically induced fractures is controlled by the in-situ stress (Bell,

SC

viii.

RI

stresses (Micklethwaite et al., 2010; Weatherley and Henley, 2013).

1996b; Davies et al., 2012; Maxwell et al., 2009; Zoback, 2007). The potential for induced seismicity associated with either subsurface fluid injection (e.g.

NU

ix.

CO2, enhanced oil recovery) or fluid withdrawal (e.g. from oil or geothermal reservoirs), is

MA

controlled by the in-situ stress (Altmann et al., 2014; Altmann et al., 2010; Davies et al., 2013; Grünthal, 2014; Hillis, 2001; Kraft and Deichmann, 2014; Segall, 1989; Zang et al.,

D

2014).

PT E

This study has greatly improved our understanding about the present-day crustal stress field in Australia and can be directly applied to all of the above issues in the Australian continent. At the

CE

continental scale, the pattern of stress in Australia, and its inconsistency with the APMs (Figure 13), can be used to investigate the relative contribution of different plate boundary forces and mantle

AC

processes on intraplate deformation and the stress field (Coblentz et al., 1998; Dyksterhuis et al., 2005a; Müller et al., 2012; Reynolds et al., 2003). Although various studies have already examined the role of plate boundary forces for the Australian continent, the new release of the ASM database reveals significant deviations between published models and the observed data in eastern Australia (Figure 14), likely due to either modelling assumptions or lack of observed stress data for model calibration in previous studies (Rajabi et al., 2017a; Rajabi et al., 2016b; Rajabi et al., 2017b). Hence, the new ASM database can be used as a valuable source of information for model calibration, which is of key importance for geomechanical-numerical modelling (Buchmann and Connolly, 2007; Rajabi

30

ACCEPTED MANUSCRIPT et al., 2017a; Reiter and Heidbach, 2014; Ziegler et al., 2016b). At regional scales, where there are perturbations of stress owing to second-order of stresses, the database can be used to improve our understanding and methodologies for new high-resolution and 3D geomechanical-numerical models that are being increasingly used to investigate the intraplate sources of stresses and for petroleum applications (Fischer and Henk, 2013; Henk, 2005; Hergert and Heidbach, 2010; Hergert et al., 2015;

PT

Rajabi et al., 2017a; Reiter and Heidbach, 2014) and for petroleum applications.

RI

The stress pattern analyses conducted in this paper revealed that large scale stress analyses do not necessarily represent the in-situ stress pattern at smaller scales. Similarly, the analysis of just a

SC

couple of borehole measurements in one area might not yield a good representation of the regional stress pattern. For example, the regional pattern of SHmax in the Gunnedah and Clarence-Moreton

NU

basins (eastern Australia) is ENE-WSW. However, stress analyses at smaller, basin and wellbore,

MA

scales revealed that SHmax orientations can be highly variable, and can even be perpendicular to the regional trend. However, it should be noted that all stress sources, acting from the plate to the wellbore scale, are inter-related to each other, so that the in-situ stress at any given point in the Earth's

PT E

D

crust is the superposition of all these different stress sources (Sonder, 1990; Tingay et al., 2006). The key finding of this study is the significant perturbation of stress in smaller scales due to third- and fourth-order of stress sources (Figures 12 and 15). The orientation and magnitude of stress

CE

have significant implications in petroleum exploration and production, mining, geothermal energy extraction and CO2 sequestration (Figures 12), (Bell, 1996b; Zang et al., 2014; Zoback, 2007). The

AC

present-day stress field is a primary control on the stability of all underground openings, including boreholes, mines and tunnels (Amadei and Stephansson, 1997; Reinecker et al., 2006; Zang et al., 2012). Hence, the perturbation of the SHmax orientation and changes of stress regime with depth needs a particular attention in any geomechanical studies at smaller scales. The neotectonic features suggest a prevailing reverse stress regime in the Australian continent and the shallow stress regime inferred from the ASM database (upper one km) is in agreement with these features (Figure 11). However, the ASM database indicates a change of stress regime with

31

ACCEPTED MANUSCRIPT depth, and thus potential errors when attempting to infer the present-day stress regime from young and shallow geological features. In particular, erroneous evaluation of in-situ stresses can lead to expensive wellbore instability, such as the recent experience of Geodynamics Limited’s with Habanero-2 geothermal well (Wyborn, 2012) and some other examples in Western Australia (Borthwick, 2010; Hayes, 2012; Kingsborough et al., 1991; Willson, 2012).

PT

The present-day stress also controls the creation of fractures, and associated permeability, in the

RI

subsurface. For example, hydraulic fractures commonly propagate parallel to the contemporary SHmax orientation and open against the least principle stress (Bell, 1996b; Engelder, 1993; Zoback, 2007).

SC

Thus, knowledge of the stress field is vital for hydraulic fracture stimulation procedures used in hydrocarbon and geothermal energy production (Bell, 1996b; Gaucher et al., 2015; Warpinski et al.,

NU

2008).

MA

The contemporary stress field is also a primary control on seismicity and the reactivation of faults in the subsurface. Fault reactivation occurs when stresses exceed the rock strength or frictional

D

limit of pre-existing faults. In particular, the injection of fluids into the subsurface, such as during

PT E

geothermal power production, CO2 sequestration or enhanced oil recovery, reduces the effective stresses and can result in fault reactivation with associated destructive seismicity or the surface release of hazardous material such as oil, CO2 or mud. For example, the Habanero fracture stimulation in the

CE

Cooper Basin of South Australia caused seismic events up to magnitude 3.7 (Asanuma et al., 2005; Hunt, 2006; Morelli, 2009). The Paralana-2 well stimulation caused several events up to magnitude

AC

2.5 in eastern South Australia (Albaric et al., 2014). Geothermal activity in Basel, Switzerland induced several seismic events of up to 3.4 in magnitude resulting in the cancellation of the project and a drawback of geothermal research in Europe in general (Grünthal, 2014). Similarly, the 2006 blowout of the Banjar Panji-1 well in Sidoarjo (Indonesia) has been suggested to have triggered reactivation of the nearby Watukosek fault that resulted in the ongoing Lusi mud flow (Tingay et al., 2015). The Lusi disaster has claimed 17 lives, flooded a 8 km2 area of the city and displaced 40,000 people with an estimated damage of over $767 million (Tingay et al., 2008). The influence of the

32

ACCEPTED MANUSCRIPT stress field on fault reactivation and associated seismicity highlights the critical importance of understanding controls on stress in risking natural hazards.

9. Conclusions

PT

The state of present-day stress in the Australian continent has been the subject of considerable

RI

debate since the first global compilation of crustal stress information. The early phases of the ASM project, published by Hillis and Reynolds (2003) contained 549 (A-E) data records, identified average

SC

stress orientations in 16 stress provinces and successfully constrained the first-order of the

NU

contemporary tectonic stress within the Australian continent. Recent unconventional and geothermal exploration in eastern and South Australia, and the running of borehole image tools in existing

MA

mineral wells in basement rocks of South Australia and Victoria provide the opportunity to update the pattern of contemporary tectonic stress in the Australian continent. The 2016 release of the ASM database contains 2150 (A-E) data records, including average SHmax orientations for 30 stress

D

provinces, and has made significant improvements in both quantity and quality of the Australian

PT E

crustal stress pattern. Of particular importance are the new, and sometimes the first, detailed contemporary stress maps for various sedimentary basins in eastern Australia, as well as the data

CE

collected from minerals wells in basement rocks of South Australia, and exploration wells in different basins including Darling, McArthur and Officer basins and offshore Tasmania. This new dataset

AC

combined with newly compiled data in almost all conventional petroleum basins of Australia in Western Australia, South Australia, Victoria and Queensland, provide the most up to date stress map of Australia. The analyses of stress data, along with directional statistics reveal four major approximate SHmax trends in the Australian continent.

i.

NNE-SSW in north, northeast and northwest of Australia.

ii.

E-W in most Western Australia as well as South Australia and western side of New South Wales.

iii.

ENE-WSW in most of eastern Australian sedimentary basins. 33

ACCEPTED MANUSCRIPT iv.

NW-SE in southeastern Australia.

The new ASM stress data further confirms the role of large-scale driving mechanisms of plate tectonics as the main intraplate sources of stress in the Australian continental crust. However, detailed analysis of borehole image logs in numerous wells revealed extensive localised perturbations of stress due to the existence of geological features such as faults, fractures, lithological density and strength

PT

contrasts and intraplate volcanism. Most of these localised stress perturbations are in eastern

RI

Australian basins (Surat, Bowen, Clarence-Moreton, Gunnedah and Sydney basins) that contain vast unconventional reserves and, hence, the state of stress is critical for exploration and production in

SC

these regions.

NU

Although the state of present-day stress regime is not determined in as much detail as the SHmax orientation, the compiled stress regime data from the upper five kilometres of the crust obtained from

MA

numerous sources (shallow engineering methods, intermediate-depth borehole information and earthquake focal mechanism solution) clearly show that the shallower part of the Australian crust is

CE

Acknowledgements

PT E

strike-slip stress regime at depth.

D

primarily characterized by a present-day thrust faulting stress regime that changes to a prevailing

AC

This work was funded by the ARC Discovery Project (DP120103849) and ASEG Research Foundation Project (RF13P02). We would like to thank the Ikon Science in Adelaide (formerly JRS Petroleum Research) for the use of their JRS Suite. Adam Baily, Gilby Jepson and Ernest Swierczek are thanked for providing new stress measurements for the Perth, Carnarvon and Gippsland basins. Jeremy Meyer, Scott Mildren, Oliver Bartdorff, Moritz Ziegler and Karsten Reiter are thanked for their support and their times to answer numerous technical questions of the first author. We also would like to thank four anonymous reviewers of the Earth-Science Reviews for their constructive

34

ACCEPTED MANUSCRIPT comments. Tim J. Horscroft (Elsevier) is thanked for inviting us to contribute this paper to Earth Science Reviews, as well as Carlo Doglioni for editorial handling. The following organisations are thanked for providing raw data for this study: New South Wales Department of Resources and Energy; Geological Survey of Queensland (Department of Natural Resources and Mines); Department of State Development, South Australia; Department of Mines and

PT

Energy of Northern Territory; Department of Mines and Petroleum of Western Australia; Victoria

RI

Department of Primary Industries; Mineral Resources Tasmania; and Australian Geophysical Observing System (AGOS) project. The first author wishes to thank the Formation Evaluation Society

SC

of Australia (for the Hugh Crocker Scholarship), Elsevier and International Association for Mathematical Geosciences (for the Computers & Geosciences Research Scholarship), which also

NU

supported this research. The maps are generated with CASMI (Heidbach and Höhne, 2008) and GMT

MA

(Wessel et al., 2013). Note that all stress data presented in this paper, with the exception of 10 data records, are included in the 2016 release of the World Stress Map Project (Heidbach et al., 2016a;

PT E

D

Heidbach et al., 2016b).

Table and Figures

CE

Table 1: Present-day stress data type and quality in the Australian stress map database and a comparison between the latest release of the project in 2003 and the new database.

AC

A-C and A-E: Quality assignment from the World Stress Map quality ranking scheme for the orientation of maximum horizontal stress (Heidbach et al., 2010). FMF: formal inversion of focal mechanism solutions, FMS: single focal mechanism solutions, FMA: composite or average focal mechanism solutions, DIF: drilling induced-tensile fractures, BO: borehole/wellbore breakouts, HF: hydraulic fracture measurements, OC: overcoring, GVA: geological volcanic alignments.

Table 2: Present-day stress provinces across continental Australia and Papua New Guinea.

35

ACCEPTED MANUSCRIPT

A-C: data quality based on the World Stress Map ranking scheme (Heidbach et al., 2010). DIF: drilling-induced tensile fractures, BO: borehole/wellbore breakouts, FMS: earthquake focal mechanism solutions, OC: overcoring, HF: hydraulic fracture measurements. Mean SHmax: mean orientation of maximum horizontal stress in each stress province using statistics for bipolar data, SD: standard deviation of the mean S Hmax in each province, R: the mean resultant vector length for the SHmax (Davis, 2002; Mardia, 1972), Conf. confidence level that show the type of stress provinces based on cut off values explained in the text. Note that the bold rows are

AC

CE

PT E

D

MA

NU

SC

RI

PT

newly defined stress provinces.

36

ACCEPTED MANUSCRIPT Figure 1: Indo-Australian Plate tectonic setting including the type of boundaries (different colours) around the plate and plate motion from Bird (2003). SUB: subduction zone, OCB: oceanic convergent boundary, OTF: oceanic transform fault, OSR: oceanic spreading ridge, CRB: continental rift boundary, CTF: continental transform fault and CCB: continental convergent boundary. The background image is land topography and ocean bathymetry based on ETOPO1

PT

model from Amante and Eakins (2009).

Figure 2: Epicentre location of the contemporary earthquakes with M > 2 across the Australian

RI

continent (Geoscience Australia, 2016a). The events plotted on this map are based on the Geoscience Australia until September 2016 (Geoscience Australia, 2016a). The seismicity in the

SC

Australian continent mainly distributed in four seismic zones shown by rectangles (Hillis et al.,

NU

2008). The background topography image is the ETOPO1 model from Amante and Eakins (2009).

MA

Figure 3: The state of crustal stress in the Australian continent by six different release of World Stress Map (WSM) and Australian Stress Map (ASM) projects. (a) Shows the first release of WSM by Zoback et al., (1992) with only 95 reliable orientation of SHmax. (b, c and d) Show three

D

different release of the ASM project by Hillis et al. (1998), Hillis and Reynolds (2000 and 2003).

PT E

The most recent prior (2003) release of the ASM and the calculated stress trajectories (grey lines) show the regional crustal stress pattern of the continent (Hillis and Reynolds, 2003). The database contained 549 in-situ stress data records based on different stress indicators (see legend). (e) Show the 2008 release of the WSM project by Heidbach et al. (2010) that show a lack of stress

CE

information in most of eastern Australia. (f) The ASM database release 2016 (this study) with 1147 A-C quality data records showing the SHmax orientations and the complementary stress

AC

trajectories (grey lines). Different colours in these maps show different stress regimes (SS: strikeslip, NF: normal, TF: thrust and U: undefined tectonic stress regime), the length of the lines indicate the quality and reliability of the SHmax orientation based on the WSM quality ranking scheme. The azimuth of each line shows the orientation of SHmax in each location. Note that only A-C quality SHmax data records are plotted on the maps.

Figure 4: Typical examples of drilling-related stress indicators including drilling-induced tensile fractures (DIFs) and wellbore breakouts (BOs) in an acoustic wellbore image logs. The image log provides a pseudo-picture of the wellbore wall based on the acoustic contrasts. In vertical wells, 37

ACCEPTED MANUSCRIPT BOs and DIFs form in the opposite sides of the borehole (parallel to wellbore axis). The BOs are interpreted as a wide pair of low-amplitude zones (that shows elongations) and display the orientation of Shmin (i.e. minimum horizontal stress), while the DIFs are identified as a narrow pair of low-amplitude zones and indicate the SHmax orientation (i.e. maximum horizontal stress). The grey sinusoids show natural fractures with different dips in this interval.

PT

Figure 5: The new (2016) release of the Australian Stress Map with all 2150 (A-E) data records. Different colours in the map show different stress regimes (SS: strike-slip, NF: normal, TF: thrust

RI

and U: undefined tectonic stress regimes). The azimuth of each line shows the SHmax orientation in each location; length of the lines indicates the quality and reliability of SHmax orientations

NU

SC

according to the World Stress Map quality ranking scheme.

Figure 6: Distribution of present-day stress data in the Australian Stress Map (ASM) database. (a)

MA

Data type that shows the drilling-related stress indicators comprise the majority of the database. (b) Data quality of ASM based on the WSM quality ranking scheme. (d) Shows the tectonic stress regime of the data records. (d, e, and f) Show the distribution of A-D quality data records versus

D

depth. (d) Shows the percentage of each data type and their distributions with depth. The majority

PT E

of the stress data are from borehole-related stress indicators that provide the stress information from near surface down to five km. (e) Indicate the frequency and distribution of data quality in the ASM database. (f) Presents the frequency and distribution of tectonic stress regime in the

CE

database.

AC

Figure 7: Location of the 30 stress provinces in the continental Australia and Papua New Guinea. Only the A-C quality data are used to define the stress provinces and at least four stress data records with A-C quality are required to define the province in a particular geographical region.

Figure 8: Mean SHmax orientation at the centre of 30 stress provinces in the Australian continent. A-C quality data that do not belong to each stress province are also shown. The background image is land topography and ocean bathymetry based on ETOPO1 model (Amante and Eakins, 2009).

38

ACCEPTED MANUSCRIPT Figure 9: Stress trajectories (a) and smoothed stress pattern (b) according to the 2016 Australian Stress Map database following the method explained by Hansen and Mount (1990) and Müller et al. (2003). Herein, at least three data records, within a 750 km search radius, are required to calculate the mean SHmax orientation.

Figure 10: Variation of tectonic stress regime versus depth in the upper five km of the Australian

PT

crust according to the Australian Stress Map database. Note that we used only the data records that are > 100 km away from any plate boundary in this diagram to show the variation of stress regime

RI

only for the intraplate setting. The resulting dataset contains 185 data records assigned with tectonic stress regime in the upper five kilometres 111 data records from hydraulic fracturing tests

SC

and overcoring measurements, 35 data records from earthquake focal mechanism solutions and 39 data records interpreted from petroleum wellbore data. A change of stress regime with depth is

NU

clear where the thrust faulting stress regime is prevailing in shallower intervals which change to

MA

strike-slip in the deeper intervals.

Figure 11: Compilation of neotectonic features (from Geoscience Australia (2016b)) and the

D

present-day stress regime in upper five km of the Australian continent. Different colours show

PT E

different stress regimes. TF: thrust, Sin./Dex. TF: sinistral and dextral thrust, NF: normal, U: undefined, TS: predominantly thrust with strike-slip component, NS: predominantly normal with the strike-slip component. White lines show the smoothed mean orientation of maximum horizontal stress orientation (SHmax) from Figure 9. The majority of neotectonic features show

CE

thrust faulting stress regime that is consistent with the SHmax orientation across the Australian continent. However, there are some inconsistencies between the stress regimes in the upper five

AC

km with neotectonic information. According to Figure 10, the ASM database shows a predominant thrust faulting tectonic stress regime in the upper one km, which is consistent with neotectonic structures. However, the tectonic regime is depth-dependant and changes to strike-slip or normal at deeper intervals.

Figure 12: Classification of various orders of crustal stress sources with spatial and temporal effects, as well as major implications and some Australian examples (modified and complied from (Heidbach et al., 2007; Rajabi et al., 2016b; Reiter et al., 2014; Tingay et al., 2006; Zoback et al., 1989). According to this classification, the orders of stress in the Earth’s lithosphere can be 39

ACCEPTED MANUSCRIPT explained by the interaction of different sources. (a) Plate boundary forces are the first-order control on the crustal stress field of continental Australia. (b) and (c) show extensive rotation of stress orientations due to the presence of second and third-order of stress sources (lithological contrasts, basement structures and faults) in the Surat and Gunnedah basins in eastern Australia. (d) Shows an example of stress rotation at wellbore scale owing to fourth-order of stress sources

PT

(faults and fractures).

Figure 13: Comparison of SHmax orientation and azimuth of absolute plate motion (APM). (a):

RI

Smoothed SHmax orientation (presented in Figure 9) versus APM azimuth from the hotspot references global plate model HS3 NUVEL-1A (Gripp and Gordon, 2002). (b): Histogram shows

SC

the distribution of misfit angle between APM and individual SHmax orientation for each stress data record in the Australian Stress Map database. SHmax orientations in the Australian continent show

NU

no correlation with the APM. The distribution curve shows the typical trend for stable North

MA

America (Richardson, 1992).

Figure 14: Comparison between regional SHmax orientations inferred from our study (a) and

D

pseudo-stress provinces inferred from published stress models (b-f). Arrows show the mean SHmax

PT E

orientation in each province and the colour codes indicates the tectonic regime including normal faulting (NF), strike-slip (SS), predominately normal faulting with strike-slip component (NS), thrust faulting (TF) and predominately thrust faulting with strike-slip component (TS). The tectonic stress regime in each province is defined if it has more than five data records assigned

CE

with stress regime. Hence, the black stress provinces are those that contain less than five data records assigned with stress regime. Among these models, the Burbidge (2004) predicted first-

AC

order SHmax pattern reasonably, however, this model is unable to predict tectonic stress regime across Australia. The model results presented in panel (b), (d), (e) and (f) fail to predict the new stress data in eastern Australia (modified from Rajabi et al. (2017a)).

Figure 15: A schematic diagram that shows the control on the stress pattern at each location based on different forces at different scales (modified from Tingay, 2009). In this case, first-order stress sources (e.g. ridge push) generate a significant east-west SHmax orientation for the region. However, second-, third- and fourth- orders of stress sources perturb the stress pattern at different

40

ACCEPTED MANUSCRIPT scales. Hence, stress orientation at each location is the result of superposition and summation of all

AC

CE

PT E

D

MA

NU

SC

RI

PT

forces acting at hundreds of kilometres to couple of meters.

41

ACCEPTED MANUSCRIPT References

AC

CE

PT E

D

MA

NU

SC

RI

PT

Aadnoy, B.S., 1990. In-situ stress directions from borehole fracture traces. Journal of Petroleum Science and Engineering, 4(2): 143-153. Aadnoy, B.S. and Bell, J.S., 1998. Classification of drilling-induced fractures and their relationship to in-situ stress directions. The Log Analyst, 39(6): 27-42. Albaric, J., Oye, V., Langet, N., Hasting, M., Lecomte, I., Iranpour, K., Messeiller, M. and Reid, P., 2014. Monitoring of induced seismicity during the first geothermal reservoir stimulation at Paralana, Australia. Geothermics, 52(0): 120-131. Alexander, L.G. and Worotnicki, G., 1957. Rock stress tests in Tumut 2 exploratory tunnel, Snowy Mountains Hydro-Electric Authority. Altmann, J.B., Müller, B.I.R., Müller, T.M., Heidbach, O., Tingay, M.R.P. and Weißhardt, A., 2014. Pore pressure stress coupling in 3D and consequences for reservoir stress states and fault reactivation. Geothermics, 52(0): 195-205. Altmann, J.B., Müller, T.M., Müller, B.I.R., Tingay, M.R.P. and Heidbach, O., 2010. Poroelastic contribution to the reservoir stress path. International Journal of Rock Mechanics and Mining Sciences, 47(7): 1104-1113. Amadei, B. and Stephansson, O., 1997. Rock stress and its measurement. Chapman & Hall, London. Amante, C. and Eakins, B.W., 2009. ETOPO1 1 arc-minute global relief model procedures, data sources and analysis. NOAA Technical Memorandum NESDIS NGDC-24. National Geophysical Data Center, NOAA. . Anderson, E.M., 1905. The dynamics of faulting. Transactions of the Edinburgh Geological Society, 8: 387–402. Artyushkov, E.V., 1973. Stresses in the lithosphere caused by crustal thickness inhomogeneities. Journal of Geophysical Research, 78(32): 7675-7708. Asanuma, H., Soma, N., Kaieda, H., Kumano, Y., Izumi, T., Tezuka , K., Niitsuma, H. and Wyborn, D., 2005. Microseismic Monitoring of Hydraulic Stimulation at the Australian HDR Project in Cooper Basin World Geothermal Congress 2005, Antalya, Turkey, pp. 5. Ask, M.V.S., 1997. In situ stress from breakouts in the Danish sector of the North Sea. Marine and Petroleum Geology, 14(3): 231-243. Bai, W., Vigny, C., Ricard, Y. and Froidevaux, C., 1992. On the origin of deviatoric stresses in the lithosphere. Journal of Geophysical Research: Solid Earth, 97(B8): 11729-11737. Bailey, A., King, R. and Backe, G., 2012. Integration of structural, stress, and seismic data to define secondary permeability networks through deep-cemented sediments in the Northern Perth Basin. Australian Petroleum Production and Exploration Association Journal, 52: 455-474. Bailey, A., King, R., Holford, S., Sage, J., Hand, M. and Backe, G., 2014. Defining structural permeability in Australian sedimentary basins. Australian Petroleum Production and Exploration Association Journal, 2015: 119-147. Bailey, A.H.E., King, R.C., Holford, S.P. and Hand, M., 2016. Incompatible stress regimes from geological and geomechanical datasets: Can they be reconciled? An example from the Carnarvon Basin, Western Australia. Tectonophysics, 683: 405-416. Balfour, N.J., Cummins, P.R., Pilia, S. and Love, D., 2015. Localization of intraplate deformation through fluid-assisted faulting in the lower-crust: The Flinders Ranges, South Australia. Tectonophysics, 655: 97-106. Barth, A., Reinecker, J. and Heidbach, O., 2016. Stress derivation from earthquake focal mechanisms In: O. Heidbach (Editor), Scientific Technical Report - WSM quality ranking scheme, database description and analysis guidelines for stress indicator (http://www.world-stress-map.org). GFZ German Research Centre for Geosciences, Potsdam, Germany, pp. 15-26. Barton, C.A. and Moos, D., 2010. Geomechanical wellbore imaging: key to managing the asset life cycle. In: M. Pöppelreiter, C. Garcia-Carballido and M. Kraaijveld (Editors),

42

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Dipmeter and Borehole Image Log Technology. American Association of Petroleum Geologists Memoir pp. 81-112. Becker, T.W., Lowry, A.R., Faccenna, C., Schmandt, B., Borsa, A. and Yu, C., 2015. Western US intermountain seismicity caused by changes in upper mantle flow. Nature, 524(7566): 458-461. Bell, J.S., 1990. Investigating stress regimes in sedimentary basins using information from oil industry wireline logs and drilling records. Geological Society London, Special Publications, 48(1): 305-325. Bell, J.S., 1996a. Petro Geoscience 1. In situ stresses in sedimentary rocks (part 1): measurement techniques. Geoscience Canada, 23(2): 85-100. Bell, J.S., 1996b. Petro Geoscience 2. In situ stresses in sedimentary rocks (part 2): applications of stress measurements. Geoscience Canada, 23(3): 135-153. Bell, J.S., 2006. In-situ stress and coal bed methane potential in Western Canada. Bulletin of Canadian Petroleum Geology, 54(3): 197-220. Bell, J.S. and Babcock, E.A., 1986. The stress regime of the Western Canadian Basin and implications for hydrocarbon production. Bulletin of Canadian Petroleum Geology, 34(3): 364-378. Bell, J.S. and Gough, D.I., 1979. Northeast-southwest compressive stress in Alberta evidence from oil wells. Earth and Planetary Science Letters, 45(2): 475-482. Berard, T., Sinha, B.K., van Ruth, P., Dance, T., John, Z. and Tan, C.P., 2008. Stress Estimation at the Otway CO2 Storage Site, Australia, SPE Asia Pacific Oil and Gas Conference and Exhibition. Society of Petroleum Engineers, Perth, Australia, pp. 26. Bevis, M., Taylor, F.W., Schutz, B.E., Recy, J., Isacks, B.L., Helu, S., Singh, R., Kendrick, E., Stowell, J., Taylor, B. and Calmantli, S., 1995. Geodetic observations of very rapid convergence and back-arc extension at the Tonga arc. Nature, 374(6519): 249-251. Bird, P., 1998. Testing hypotheses on plate-driving mechanisms with global lithosphere models including topography, thermal structure, and faults. Journal of Geophysical Research: Solid Earth, 103(B5): 10115-10129. Bird, P., 1999. Thin-plate and thin-shell finite-element programs for forward dynamic modeling of plate deformation and faulting1. Computers & Geosciences, 25(4): 383-394. Bird, P., 2003. An updated digital model of plate boundaries. Geochemistry, Geophysics, Geosystems, 4(3): 1027. Bird, P., Liu, Z. and Rucker, W.K., 2008. Stresses that drive the plates from below: Definitions, computational path, model optimization, and error analysis. Journal of Geophysical Research: Solid Earth, 113(B11): n/a-n/a. Blomberg, J.R., 1971. The Determination and Considerations of Induced Fracture Strike, SPE Eastern Regional Meeting Society of Petroleum Engineers, Charleston, West Virginia, pp. 8. Blümling, P., 1986. In situ Spannungsmessungen in Tiefbohrungen mit Hilfe von Bohrlochrandausbrnchen und die Spannungsverteilung in der Kruste Mitteleuropas und Australiens, University of Karlsruhe, Karlsruhe, Germany, 135 pp. Bock, G. and Denham, D., 1983. Recent earthquake activity in the Snowy Moutains region and its relationship to major faults. Journal of the Geological Society of Australia, 30(3-4): 423429. Boness, N.L. and Zoback, M.D., 2004. Stress-induced seismic velocity anisotropy and physical properties in the SAFOD Pilot Hole in Parkfield, CA. Geophysical Research Letters, 31(15): L15S17 1-4. Borthwick, D., 2010. The Report of the Montara Commission of Inquiry. Montara Commission of Inquiry, Canberra (2010) http://www.industry.gov.au/AboutUs/CorporatePublications/MontaraInquiryResponse/Do cuments/Montara-Report.pdf (accessed 22.03.16). Bott, M.H.P., 1959. The mechanics of oblique slip faulting. Geological Magazine, 96(2): 109-117. Bradley, W.B., 1979. Failure of Inclined Boreholes. Journal of Energy Resources Technology, 101(4): 232-239. 43

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Brereton, R., Muller, B., Hancock, P., Harper, T., Bott, M.H.P., Sanderson, D. and Kusznir, N., 1991. European Stress: Contributions from Borehole Breakouts [and Discussion]. Philosophical Transactions: Physical Sciences and Engineering, 337(1645): 165-179. Brooke-Barnett, S., Flottmann, T., Paul, P.K., Busetti, S., Hennings, P., Reid, R. and Rosenbaum, G., 2015. Influence of basement structures on in situ stresses over the Surat Basin, southeast Queensland. Journal of Geophysical Research: Solid Earth, 120(7): 4946-4965. Buchmann, T.J. and Connolly, P.T., 2007. Contemporary kinematics of the Upper Rhine Graben: A 3D finite element approach. Global and Planetary Change, 58(1–4): 287-309. Burbidge, D.R., 2004. Thin plate neotectonic models of the Australian plate. Journal of Geophysical Research: Solid Earth, 109(B10): B10405. Célérier, B., Etchecopar, A., Bergerat, F., Vergely, P., Arthaud, F. and Laurent, P., 2012. Inferring stress from faulting: From early concepts to inverse methods. Tectonophysics, 581: 206219. Chang, C., McNeill, L.C., Moore, J.C., Lin, W., Conin, M. and Yamada, Y., 2010. In situ stress state in the Nankai accretionary wedge estimated from borehole wall failures. Geochemistry, Geophysics, Geosystems, 11(12): n/a-n/a. Clark, D. and Leonard, M., 2003. Principal stress orientations from multiple focal-plane solutions: new insight into the Australian intraplate stress field. Geological Society of America Special Papers, 372: 91-105. Clark, D., McPherson, A. and Allen, T., 2014. Intraplate Earthquakes in Australia. In: P. Talwani (Editor), Intraplate Earthquakes. Cambridge University Press, pp. 398. Clark, D., McPherson, A., Cupper, M., Collins, C.D.N. and Nelson, G., 2015. The Cadell Fault, southeastern Australia: a record of temporally clustered morphogenic seismicity in a lowstrain intraplate region. Geological Society, London, Special Publications, 432. Clark, D., McPherson, A. and Van Dissen, R., 2012. Long-term behaviour of Australian stable continental region (SCR) faults. Tectonophysics, 566–567(0): 1-30. Cleary, J.R., 1962. Near earthquake studies in south-eastern Australia, Australian National University. Cleary, J.R., Doyle, H.A. and Moye, D.G., 1964. Seismic activity in the snowy mountains region and its relationship to geological structures. Journal of the Geological Society of Australia, 11(1): 89-106. Cloetingh, S. and Wortel, R., 1985. Regional stress field of the Indian Plate. Geophysical Research Letters, 12(2): 77-80. Cloetingh, S. and Wortel, R., 1986. Stress in the Indo-Australian plate. Tectonophysics, 132(1–3): 49-67. Coblentz, D. and Richardson, R.M., 1995. Statistical trends in the intraplate stress field. Journal of Geophysical Research: Solid Earth, 100(B10): 20245-20255. Coblentz, D., Sandiford, M., Richardson, R.M., Zhou, S. and Hillis, R., 1995. The origins of the intraplate stress field in continental Australia. Earth and Planetary Science Letters, 133(3– 4): 299-309. Coblentz, D., van Wijk, J., Richardson, R.M. and Sandiford, M., 2015. The upper mantle geoid: Implications for continental structure and the intraplate stress field. Geological Society of America Special Papers, 514. Coblentz, D., Zhou, S., Hillis, R., Richardson, R.M. and Sandiford, M., 1998. Topography, boundary forces, and the Indo-Australian intraplate stress field. Journal of Geophysical Research: Solid Earth, 103(B1): 919-931. Coltice, N., Gérault, M. and Ulvrová, M., 2017. A mantle convection perspective on global tectonics. Earth-Science Reviews, 165: 120-150. Cornet, F.H., 2015. Elements of crustal geomechanics. Cambridge University Press, 490 pp. Crampin, S., 1991. Wave propagation through fluid-filled inclusions of various shapes: interpretation of extensive-dilatancy anisotropy. Geophysical Journal International, 104(3): 611-623. D'Agostino, N., England, P., Hunstad, I. and Selvaggi, G., 2014. Gravitational potential energy and active deformation in the Apennines. Earth and Planetary Science Letters, 397: 121132. 44

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Davies, R., Foulger, G., Bindley, A. and Styles, P., 2013. Induced seismicity and hydraulic fracturing for the recovery of hydrocarbons. Marine and Petroleum Geology, 45(0): 171185. Davies, R.J., Mathias, S.A., Moss, J., Hustoft, S. and Newport, L., 2012. Hydraulic fractures: How far can they go? Marine and Petroleum Geology, 37(1): 1-6. Davis, J.C., 2002. Statistics and data analysis in geology. Wiley, New York, 656 pp. Debayle, E., 1999. SV-wave azimuthal anisotropy in the Australian upper mantle: preliminary results from automated Rayleigh waveform inversion. Geophysical Journal International, 137(3): 747-754. Debayle, E., Kennett, B. and Priestley, K., 2005. Global azimuthal seismic anisotropy and the unique plate-motion deformation of Australia. Nature, 433(7025): 509-512. Debayle, E. and Kennett, B.L.N., 2000. Anisotropy in the Australasian upper mantle from Love and Rayleigh waveform inversion. Earth and Planetary Science Letters, 184(1): 339-351. Debayle, E. and Kennett, B.L.N., 2003. Surface-wave studies of the Australian region. Geological Society of America Special Papers, 372: 25-40. Denham, D., 1980. The 1961 Robertson earthquake - more evidence for compressive stress in Southeast Australia. BMR Journal of Australian Geology & Geophysics(5): 153-156. Denham, D., Alexander, L.G. and Worotnicki, G., 1979. Stresses in the Australian crust: evidence from earthquakes and in-situ stress measurements. BMR Journal of Australian Geology and Geophysics, 67(4): 289-295. Denham, D., Alexander, L.G. and Worotnicki, G., 1980. The stress field near the sites of the Meckering (1968) and Calingiri (1970) earthquakes, Western Australia. Tectonophysics, 67(3–4): 283-317. Denham, D., Bock, G. and Smith, R.S., 1982. The Appin (New SouthWales) earthquake of 15 November 1981. BMR Journal of Geology & Geophysics, 28: 323-332. Denham, D., Jones, T. and Weekes, J., 1985. The 1982 Wyalong earthquakes (NSW) and recent crustal deformation. BMR Journal of Australian Geology & Geophysics, 9: 255-260. Denham, D., Weekes, J. and Krayshek, C., 1981. Earthquake evidence for compressive stress in the southeast Australian crust. Journal of the Geological Society of Australia, 28(3-4): 323-332. Denham, D. and Windsor, C.R., 1991. The crustal stress pattern in the Australian continent. Exploration Geophysics, 22(1): 101-106. Dyksterhuis, S., Albert, R.A. and Müller, R.D., 2005a. Finite-element modelling of contemporary and palaeo-intraplate stress using ABAQUS™. Computers & geosciences, 31(3): 297307. Dyksterhuis, S., Müller, R.D. and Albert, R.A., 2005b. Paleostress field evolution of the Australian continent since the Eocene. Journal of Geophysical Research: Solid Earth, 110(B5): B05102. Dziewonski, A.M., Chou, T.A. and Woodhouse, J.H., 1981. Determination of earthquake source parameters from waveform data for studies of global and regional seismicity. Journal of Geophysical Research: Solid Earth, 86(B4): 2825-2852. Ekström, G., Nettles, M. and Dziewoński, A.M., 2012. The global CMT project 2004–2010: Centroid-moment tensors for 13,017 earthquakes. Physics of the Earth and Planetary Interiors, 200–201: 1-9. Enever, J.R., Gale, W. and Fabjanczvk, M., 1999. A Study of the Variability of the Horizontal Stress Field In a Sedimentary Basin, 9th ISRM Congress. International Society for Rock Mechanics, Paris, France, pp. 1257-1260. Enever, J.R., Hattersley, P. and Wooltorton, B., 1984. In situ rock stress measurements using the hydraulic fracturing technique for the proposed Sydney Ocean outfalls project, Fifth Australian Tunnelling Conference: State of the Art in Underground Development and Construction. Institution of Engineers, Australia, Barton, ACT, pp. 114-121. Engelder, T., 1993. Stress regimes in the lithosphere. Princeton University Press, New Jersey. Everingham, I.B. and Smith, R.S., 1979. Implications of fault-plane solutions for Australian earthquakes of 4 July 1977, 6 May 1978 and 25 November 1978. BMR Journal of Australian Geology & Geophysics, 4: 297-301. 45

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Ferrill, D.A., Stamatakos, J.A. and Sims, D., 1999. Normal fault corrugation: implications for growth and seismicity of active normal faults. Journal of Structural Geology, 21(8–9): 1027-1038. Fischer, K. and Henk, A., 2013. A workflow for building and calibrating 3-D geomechanical models &ndash a case study for a gas reservoir in the North German Basin. Solid Earth, 4(2): 347-355. Fisher, M.K. and Warpinski, N.R., 2012. Hydraulic-Fracture-Height Growth: Real Data. SPE Production & Operations, 27(01): 8-19. Fitch, T.J., 1976. The Picton earthquake of 9 March 1973: A seismic view of the source. In: D. Denham (Editor), Seismicity and earthquake risk in eastern Australia. Department of National Resources, Burea of mineral resources, geology and geophysics, Canberra. Fitch, T.J., Worthington, M.H. and Everingham, I.B., 1973. Mechanisms of Australian earthquakes and contemporary stress in the Indian ocean plate. Earth and Planetary Science Letters, 18(2): 345-356. Flottmann, T., Brooke-Barnett, S., Trubshaw, R., Naidu, S.-K., Kirk-Burnnand, E., Paul, P., Busetti, S. and Hennings, P., 2014. Influence of in-situ stresses on fracture stimulations in the Surat Basin, southeast Queensland, SPE Unconventional Resources Conference and Exhibition-Asia Pacific. SPE-167064-MS, Brisbane, Australia pp. 14 pages. Fredrich, J., McCaffrey, R. and Denham, D., 1988. Source Parameters of Seven Large Australian Earthquakes Determined By Body Waveform Inversion. Geophysical Journal International, 95(1): 1-13. Gale, W.J., Enever, J.R., Blackwood, R.L. and McKay, J., 1984. An investigation of the effect of a fault /monocline structure on the in-situ stress field and mining conditions at Nattai Bulli Colliery NSW, Australia, . CSIRO Australia, Division of Geomechanics, Geomechanics of Coal Mining Report No. 48. Gaucher, E., Schoenball, M., Heidbach, O., Zang, A., Fokker, P.A., van Wees, J.-D. and Kohl, T., 2015. Induced seismicity in geothermal reservoirs: A review of forecasting approaches. Renewable and Sustainable Energy Reviews, 52: 1473-1490. Gavin, L.J. and Lumley, D., 2016. Stress-induced seismic azimuthal anisotropy in the upper crust across the North West Shelf, Australia. Journal of Geophysical Research: Solid Earth, 121(2): 1023-1039. GEOFON Data Centre, 1993. GEOFON Seismic Network. doi:10.14470/TR560404. Geoscience Australia, 2016a. Earthquake database: http://www.ga.gov.au/earthquakes/#/ retrieved, 5 May 2016. Geoscience Australia, 2016b. Neotectonic features Geoscience Australia: http://www.ga.gov.au/earthquakes/staticPageController.do?page=neotectonics retrieved, 25 January 2016. Ghosh, A. and Holt, W.E., 2012. Plate Motions and Stresses from Global Dynamic Models. Science, 335(6070): 838-843. Ghosh, A., Holt, W.E. and Flesch, L.M., 2009. Contribution of gravitational potential energy differences to the global stress field. Geophysical Journal International, 179(2): 787-812. Ghosh, A., Holt, W.E., Flesch, L.M. and Haines, A.J., 2006. Gravitational potential energy of the Tibetan Plateau and the forces driving the Indian plate. Geology, 34(5): 321-324. Ghosh, A., Holt, W.E. and Wen, L., 2013. Predicting the lithospheric stress field and plate motions by joint modeling of lithosphere and mantle dynamics. Journal of Geophysical Research: Solid Earth, 118(1): 346-368. Ghosh, A., Holt, W.E., Wen, L., Haines, A.J. and Flesch, L.M., 2008. Joint modeling of lithosphere and mantle dynamics elucidating lithosphere-mantle coupling. Geophysical Research Letters, 35(16). Greenhalgh, S.A., Love, D., Malpas, K. and McDougall, R., 1994. South Australian earthquakes, 1980–92. Australian Journal of Earth Sciences, 41(5): 483-495. Greenhalgh, S.A. and Singh, R., 1988. The seismicity of the Adelaide Geosyncline, South Australia. Bulletin of the Seismological Society of America, 78(1): 243-263. Gregson, P.J. and Denham, D., 1987. Australian Seismological Report 1983. Dept. of Primary Ind. and Energy, Canberra, ACT, Australia - Report 280. 46

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Gripp, A.E. and Gordon, R.G., 2002. Young tracks of hotspots and current plate velocities. Geophysical Journal International, 150(2): 321-361. Grünthal, G., 2014. Induced seismicity related to geothermal projects versus natural tectonic earthquakes and other types of induced seismic events in Central Europe. Geothermics, 52: 22-35. Gui, F., Rahman, K., Lockhart, D., Cunha, D. and Meadows, D., 2016. High Stress Contrast between Different Lithologies and its Implications for Drilling and Stimulation in the Northern Cooper Basin, SPE Asia Pacific Oil & Gas Conference and Exhibition. Society of Petroleum Engineers, Perth, Australia. Haimson, B.C. and Cornet, F.H., 2003. ISRM Suggested Methods for rock stress estimation—Part 3: hydraulic fracturing (HF) and/or hydraulic testing of pre-existing fractures (HTPF). International Journal of Rock Mechanics and Mining Sciences, 40(7–8): 1011-1020. Hansen, K.M. and Mount, V.S., 1990. Smoothing and extrapolation of crustal stress orientation measurements. Journal of Geophysical Research: Solid Earth, 95(B2): 1155-1165. Hast, N., 1969. The state of stress in the upper part of the earth's crust. Tectonophysics, 8(3): 169211. Hayes, J., 2012. Operator competence and capacity – Lessons from the Montara blowout. Safety Science, 50(3): 563-574. Heffer, K.J., Fox, R.J., McGill, C.A. and Koutsabeloulis, N.C., 1997. Novel Techniques Show Links between Reservoir Flow Directionality, Earth Stress, Fault Structure and Geomechanical Changes in Mature Waterfloods. SPE Journal, 2(02): 91-98. Heffer, K.J. and Lean, J.C., 1991. Earth stress orientation - a control on, and guide to, flooding directionality in a majority of reservoirs. In: B. Linville (Editor), Reservoir Characterization III. PennWell Books, Tulsa, pp. 799-822. Heidbach, O. and Ben-Avraham, Z., 2007. Stress evolution and seismic hazard of the Dead Sea Fault System. Earth and Planetary Science Letters, 257(1–2): 299-312. Heidbach, O. and Höhne, J., 2008. CASMI—A visualization tool for the World Stress Map database. Computers & Geosciences, 34(7): 783-791. Heidbach, O., Iaffaldano, G. and Bunge, H.-P., 2008a. Topography growth drives stress rotations in the central Andes: Observations and models. Geophysical Research Letters, 35(8): L08301. Heidbach, O., Rajabi, M., Reiter, K. and Ziegler, M., 2016a. World Stress Map 2016. GFZ Data Services. http://doi.org/10.5880/WSM.2016.002. Heidbach, O., Rajabi, M., Reiter, K., Ziegler, M. and WSM-Team, 2016b. World Stress Map Database Release 2016. GFZ Data Services. http://doi.org/10.5880/WSM.2016.001. Heidbach, O. and Reinecker, J., 2013. Analyse des rezenten Spannungsfeldes der Nordschweiz, (Arbeitsbericht NAB 12-05 / nagra), Wettingen : Nationale Genossenschaft für die Lagerung radioaktiver Abfälle (nagra). http://www.nagra.ch/data/documents/database/dokumente/$default/Default%20Folder/Pu blikationen/NABs%202004-2012/d_nab12-005.pdf. Heidbach, O., Reinecker, J., Tingay, M., Müller, B., Sperner, B., Fuchs, K. and Wenzel, F., 2007. Plate boundary forces are not enough: Second- and third-order stress patterns highlighted in the World Stress Map database. Tectonics, 26(6): TC6014. Heidbach, O., Tingay, M., Barth, A., Reinecker, J., Kurfeß, D. and Müller, B., 2008b. The World Stress Map database release 2008. URL: www.world-stress-map.org. Heidbach, O., Tingay, M., Barth, A., Reinecker, J., Kurfeß, D. and Müller, B., 2010. Global crustal stress pattern based on the World Stress Map database release 2008. Tectonophysics, 482(1–4): 3-15. Henk, A., 2005. Pre-drilling prediction of the tectonic stress field with geomechanical models. First break, 23(11): 53-57. Hergert, T. and Heidbach, O., 2010. Slip-rate variability and distributed deformation in the Marmara Sea fault system. Nature Geoscience, 3(2): 132-135. Hergert, T. and Heidbach, O., 2011a. Geomechanical model of the Marmara Sea region—II. 3-D contemporary background stress field. Geophysical Journal International, 185(3): 10901102. 47

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Hergert, T. and Heidbach, O., 2011b. Geomechanical model of the Marmara Sea region - II. 3-D contemporary background stress field. Geophysical Journal International: doi:10.1111/j.1365-246X.2011.04992.x, 1090-1102. Hergert, T., Heidbach, O., Reiter, K., Giger, S.B. and Marschall, P., 2015. Stress field sensitivity analysis in a sedimentary sequence of the Alpine foreland, Northern Switzerland. Solid Earth Discuss., 7(1): 711-756. Hillis, R., 1991. Australia-Banda Arc collision and in situ stress in the Vulcan Sub-basin (Timor Sea) as revealed by borehole breakout data. Exploration Geophysics, 22(1): 189-194. Hillis, R., 2001. Coupled changes in pore pressure and stress in oil fields and sedimentary basins. Petroleum Geoscience, 7(4): 419-425. Hillis, R., Enever, J.R. and Reynolds, S.D., 1999. In situ stress field of eastern Australia. Australian Journal of Earth Sciences, 46(5): 813-825. Hillis, R., Meyer, J.J. and Reynolds, S.D., 1998. The Australian stress map. Exploration Geophysics, 29(4): 420-427. Hillis, R., Mildren, S.D., Pigram, C.J. and Willoughby, D.R., 1997a. Rotation of horizontal stresses in the Australian North West continental shelf due to the collision of the IndoAustralian and Eurasian Plates. Tectonics, 16(2): 323-335. Hillis, R., Monte, S.A., Tan, C.P. and Willoughby, D.R., 1995. The contemporary stress field of the Otway Basin, South Australia: implications for hydrocarbon exploration and production. Australian Petroleum Production and Exploration Association Journal, 35: 494-506. Hillis, R. and Reynolds, S.D., 2000. The Australian Stress Map. Journal of the Geological Society, 157(5): 915-921. Hillis, R. and Reynolds, S.D., 2003. In situ stress field of Australia. Geological Society of America Special Papers, 372: 49-58. Hillis, R., Sandiford, M., Coblentz, D.D. and Zhou, S., 1997b. Modelling the contemporary stress field and its implications for hydrocarbon exploration. Exploration Geophysics, 28(2): 8893. Hillis, R., Sandiford, M., Reynolds, S.D. and Quigley, M.C., 2008. Present-day stresses, seismicity and Neogene-to-Recent tectonics of Australia's ‘passive’ margins: intraplate deformation controlled by plate boundary forces. Geological Society, London, Special Publications, 306(1): 71-90. Hillis, R. and Williams, A.F., 1993a. The contemporary stress field of the Barrow-Dampier SubBasin and its implications for horizontal drilling. Exploration Geophysics, 24(4): 567-576. Hillis, R.R. and Williams, A.F., 1993b. The stress field of the North West Shelf and wellbore stability. APPEA Journal, 33: 373-385. Holdgate, G.R., Wallace, M.W., Gallagher, S.J., Smith, A.J., J.B., K., Moore, D. and Shafil, S., 2003. Plio-Pleistocene tectonics and eustasy in the Gippsland Basin, southeast Australia: evidence from magnetic imagery and marine geological data. Australian Journal of Earth Sciences, 50: 403-426. Holford, S.P., Hillis, R., Duddy, I.R., Green, P.F., Stoker, M.S., Tuitt, A.K., Backé, G., Tassone, D.R. and MacDonald, J., 2011. Cenozozic post-breakup compressional deformation and exhumation of the Southern Australian Margin. Australian Petroleum Production and Exploration Association Journal, 51: 613-638. Hough, S.E., Armbruster, J.G., Seeber, L. and Hough, J.F., 2000. On the Modified Mercalli intensities and magnitudes of the 1811–1812 New Madrid earthquakes. Journal of Geophysical Research: Solid Earth, 105(B10): 23839-23864. Hubbert, M.K. and Willis, D.G., 1957. Mechanics of hydraulic fracturing. Transactions of the American Institute of Mining, Metallurgical, and Petroleum Engineers, 210: 153-166. Hunt, S.P.M., C.P., 2006. Cooper Basin HDR hazard evaluation: Predictive modelling of local stress changes due to HFR geothermal energy operations in South Australia. South Australia. Department of Primary Industries and Resources. Report Book 2006/16. Jepson, G., 2014. In-situ stress and natural fracture networks in the Carnarvon Basin, North West Shelf, Australia, University of Adelaide, Adelaide. 48

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Johnston, A.C., Coppersmith, K.J., Kanter, L.R. and Cornell, C.A., 1994. The earthquakes of stable continental regions. Electric Power Research Institute Report TR102261V1. Keep, M. and Schellart, W.P., 2012. Introduction to the thematic issue on the evolution and dynamics of the Indo-Australian plate. Australian Journal of Earth Sciences, 59(6): 807808. King, G.E., 2010. Thirty Years of Gas Shale Fracturing: What Have We Learned?, SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers (SPE-133456-MS), Florence, Italy, pp. 50. King, R.C., Hillis, R. and Reynolds, S.D., 2008. In situ stresses and natural fractures in the Northern Perth Basin, Australia. Australian Journal of Earth Sciences, 55(5): 685-701. King, R.C., Hillis, R., Tingay, M.R.P. and Morley, C.K., 2009. Present-day stress and neotectonic provinces of the Baram Delta and deep-water fold–thrust belt. Journal of the Geological Society, 166(2): 197-200. King, R.C., Holford, S.P., Hillis, R., Tuitt, A., Swierczek, E., Backé, G., Tassone, D.R. and Tingay, M., 2012. Reassessing the in-situ stress regimes of Australia's petroleum basins. Australian Petroleum Production and Exploration Association Journal, 52: 415-425. Kingsborough, R.H., Williams, A.F. and Hillis, R., 1991. Borehole Instability on the Northwest Shelf of Australia, SPE Asia-Pacific Conference Society of Petroleum Engineers, Perth. Australia, pp. 653-662. Kirsch, E.G., 1898. Die Theorie der Elastizität und die Bedürfnisse der Festigkeitslehre. Zeitschrift des Vereines Deutscher Ingenieure, 42: 797-807. Klee, G., Bunger, A., Meyer, G., Rummel, F. and Shen, B., 2011. In Situ Stresses in Borehole Blanche-1/South Australia Derived from Breakouts, Core Discing and Hydraulic Fracturing to 2 km Depth. Rock Mechanics and Rock Engineering, 44(5): 531-540. Koptev, A.I. and Ershov, A.V., 2010. The role of the gravitational potential of the lithosphere in the formation of a global stress field. Izvestiya, Physics of the Solid Earth, 46(12): 10801094. Kraft, T. and Deichmann, N., 2014. High-precision relocation and focal mechanism of the injection-induced seismicity at the Basel EGS. Geothermics, 52: 59-73. Lambeck, K., McQueen, H., Stephenson, R. and Denham, D., 1984. The state of stress within the Australian continent, Annales Geophysicae. Gauthier-Villars, pp. 723-741. Leeman, E.R., 1964. The measurement of stress in rock - Parts I, II and III. Journal of the Southern African Institute of Mining and Metallurgy, 65: 45–114, 254–284. Leonard, M., 2008. One hundred years of earthquake recording in Australia. Bulletin of the Seismological Society of America, 98(3): 1458-1470. Leonard, M. and Clark, D., 2011. A record of stable continental region earthquakes from Western Australia spanning the late Pleistocene: Insights for contemporary seismicity. Earth and Planetary Science Letters, 309(3–4): 207-212. Leonard, M., Ripper, I.D. and Yue, L., 2002. Australian earthquake fault plane solutions. Geoscience Australia, Record 2002/19. Lithgow-Bertelloni, C. and Guynn, J.H., 2004. Origin of the lithospheric stress field. Journal of Geophysical Research: Solid Earth, 109(B1): B01408. Lithgow-Bertelloni, C. and Silver, P.G., 1998. Dynamic topography, plate driving forces and the African superswell. Nature, 395(6699): 269-272. Liu, M. and Stein, S., 2016. Mid-continental earthquakes: Spatiotemporal occurrences, causes, and hazards. Earth-Science Reviews, 162: 364-386. Lockner, D.A. and Stanchits, S.A., 2002. Undrained poroelastic response of sandstones to deviatoric stress change. Journal of Geophysical Research: Solid Earth, 107(B12): ETG 13-1-ETG 13-14. Lowry, D., 1990. Breakouts in oil wells. PESA Journal, 17: 43-44. Lund Snee, J.-E. and Zoback, M.D., 2016. State of stress in Texas: Implications for induced seismicity. Geophysical Research Letters, 43: 10,208-10,214. MacGregor, S., 2002. Acoustic scanner analysis of borehole breakout to define the stress field across mine sites in the Sydney and Bowen Basins, Australia 21st International Conference on Ground Control in Mining Morgantown, WV, USA, pp. 278-287. 49

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

MacGregor, S., 2003. Maximising in situ stress measurement data from borehole breakout using acoustic scanner and wireline tools, Final report ACARP Project C10009. MacGregor, S. and Gale, W., 2000. In situ stress measurement using acoustic scanner analysis of borehole breakout, Bowen Basin Symposium, pp. 97-106. Mardia, K.V., 1972. Statistics of directional data / K.V. Mardia. Probability and mathematical statistics. Academic Press, London ; New York. Maxwell, S.C., Zimmer, U., Gusek, R.W. and Quirk, D.J., 2009. Evidence of a Horizontal Hydraulic Fracture From Stress Rotations Across a Thrust Fault. SPE Production & Operations, 24(2): 312-319. McCaffrey, R., 1989. Teleseismic investigation of the January 22, 1988 Tennant Creek, Australia, earthquakes. Geophysical Research Letters, 16(5): 413-416. McCallum, R.E. and Bell, J.S., 1994. Western Canada Sedimentary Basin borehole imagery analysis project: a summary of TOTAL CONWEST Cypress C-50-B/94-B-15, Geological Survey of Canada, Open File 2945. McCallum, R.E. and Bell, J.S., 1995. Diagnosing natural and drilling-induced fractures from borehole images of western Canadian wells. In: J.S. Bell, T.D. Bird, T.L. Hillier and P.L. Greener (Editors), Oil and Gas Forum '95 - Energy from Sediments. Geological Survey of Canada, Open File Report 3058, pp. 79-82. McCue, K., 1990. Australia's large earthquakes and Recent fault scarps. Journal of Structural Geology, 12(5–6): 761-766. McCue, K., 2014. Historical earthquakes in Western Australia. Australian Earthquake Engineering Society, Canberra: 46. McCue, K., Gibson, G. and Wesson, V., 1990a. The earthquake near Nhill, western Victoria, on 22 December 1987 and the seismicity of eastern Australia. BMR Journal of Australian Geology and Geophysics, 11(4): 415-420. McCue, K., Wesson, V. and Gibson, G., 1990b. The Newcastle, New South Wales, earthquake of 28 December 1989. BMR Journal of Australian Geology and Geophysics, 11(4): 559-567. McCue, K.F., 1989. Australian Seismological Report 1985. Dept. of Primary Ind. and Energy, Canberra, ACT, Australia, Canberra, Australia - Report 285: 50. McCue, K.F. and Sutton, D.J., 1979. South Australian earthquakes during 1976 and 1977. Journal of the Geological Society of Australia, 26(5-6): 231-236. McEwin, A., Underwood, R. and Denham, D., 1976. Earthquake risk in Australia. BMR Journal of Australian Geology and Geophysics, 1(1): 15-21. McGarr, A. and Gay, N.C., 1978. State of stress in the Earth's crust. Annual Review of Earth and Planetary Sciences, 6: 405-436. McKenzie, D.P., 1969. The relation between fault plane solutions for earthquakes and the directions of the principal stresses. Bulletin of the Seismological Society of America, 59(2): 591-601. Micklethwaite, S., Sheldon, H.A. and Baker, T., 2010. Active fault and shear processes and their implications for mineral deposit formation and discovery. Journal of Structural Geology, 32(2): 151-165. Mildren, S.D. and Hillis, R., 2000. In situ stresses in the Southern Bonaparte Basin, Australia: Implications for first- and second-order controls on stress orientation. Geophysical Research Letters, 27(20): 3413-3416. Mildren, S.D., Hillis, R., Fett, T. and Robinson, P.H., 1994. Contemporary stresses in the Timor Sea: implications for fault trap integrity. In: P.G. Purcell and R.R. Purcell (Editors), Petroleum Exploration Society of Australia Symposium: The Sedimentary Basins of Western Australia, Perth, Western Australia, pp. 291-300. Mildren, S.D., Hillis, R. and Kaldi, J., 2002. Calibrating predictions of fault seal rectivation in the Timor Sea. Australian Petroleum Production and Exploration Association Journal, 42: 187-202. Molnar, P. and Lyon-Caen, H., 1988. Some simple physical aspects of the support, structure, and evolution of mountain belts. Geological Society of America Special Papers, 218: 179-208. Montone, P. and Mariucci, M.T., 2016. The new release of the Italian contemporary stress map. Geophysical Journal International, 205(3): 1525-1531. 50

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Morelli, C.P., 2009. Analysis and management of seismic risks associated with engineered geothermal system operations in South Australia, Report Book 2009/11, Petroleum and Geothermal Group, Department of Primary Industries and Resources South Australia, Adelaide. Mount, V.S. and Suppe, J., 1987. State of stress near the San Andreas fault: Implications for wrench tectonics. Geology, 15(12): 1143-1146. Müller, B., Heidbach, O., Negut, M., Sperner, B. and Buchmann, T., 2010. Attached or not attached—evidence from crustal stress observations for a weak coupling of the Vrancea slab in Romania. Tectonophysics, 482(1–4): 139-149. Müller, B., Wehrle, V., Hettel, S., Sperner, B. and Fuchs, K., 2003. A new method for smoothing orientated data and its application to stress data. Geological Society, London, Special Publications, 209(1): 107-126. Müller, B., Zoback, M.L., Fuchs, K., Mastin, L., Gregersen, S., Pavoni, N., Stephansson, O. and Ljunggren, C., 1992. Regional patterns of tectonic stress in Europe. Journal of Geophysical Research: Solid Earth, 97(B8): 11783-11803. Müller, R.D., Dyksterhuis, S. and Rey, P., 2012. Australian paleo-stress fields and tectonic reactivation over the past 100 Ma. Australian Journal of Earth Sciences, 59(1): 13-28. Naliboff, J.B., Conrad, C.P. and Lithgow-Bertelloni, C., 2009. Modification of the lithospheric stress field by lateral variations in plate-mantle coupling. Geophysical Research Letters, 36(22): n/a-n/a. Naliboff, J.B., Lithgow-Bertelloni, C., Ruff, L.J. and de Koker, N., 2012. The effects of lithospheric thickness and density structure on Earth's stress field. Geophysical Journal International, 188(1): 1-17. Nelson, E.J., Hillis, R., Sandiford, M., Reynolds, S. and Mildren, S.D., 2006. Present-day state-ofstress of southeast Australia. Australian Petroleum Production and Exploration Association Journal, 2006: 283-305. Olsen, T.N., Bratton, T.R., Tanner, K.V., donald, A. and Koepsell, R., 2007. Application of Indirect Fracturing for efficient Stimulation of Coalbed Methane, Rocky Mountain Oil & Gas Technology Symposium. Society of Petroleum Engineers, Denver, Colorado, . Paillet, F.L., 1985. Applications Of Borehole-Acoustic Methods In Rock Mechanics, The 26th U.S. Symposium on Rock Mechanics (USRMS). American Rock Mechanics Association, Rapid City, South Dakota, pp. 207-220. Palano, M., 2015. On the present-day crustal stress, strain-rate fields and mantle anisotropy pattern of Italy. Geophysical Journal International, 200(2): 967-983. Palmer, I., 2010. Coalbed methane completions: A world view. International Journal of Coal Geology, 82(3–4): 184-195. Pierdominici, S. and Heidbach, O., 2012. Stress field of Italy — Mean stress orientation at different depths and wave-length of the stress pattern. Tectonophysics, 532–535(0): 301311. Pitcher, T. and Davis, T.L., 2016. Geomechanical analysis of in-situ stress and its influence on hydraulic fracturing at the Wattenberg Field, Colorado. First Break, 34(3): 45-50. Plumb, R.A. and Hickman, S.H., 1985. Stress-induced borehole elongation: A comparison between the four-arm dipmeter and the borehole televiewer in the Auburn Geothermal Well. Journal of Geophysical Research: Solid Earth, 90(B7): 5513-5521. Quigley, M., Clark, D. and Sandiford, M., 2010. Tectonic geomorphology of Australia. Geological Society of London Special Publications, 346: 243-265. Quigley, M., Cupper, M. and Sandiford, M., 2006. Quaternary faults of south-central Australia: palaeoseismicity, slip rates and origin. Australian Journal of Earth Sciences, 53(2): 285301. Rajabi, M., Heidbach, O., Tingay, M. and Reiter, K., 2017a. Prediction of the present-day stress field in the Australian continental crust using 3D geomechanical-numerical models. Australian Journal of Earth Sciences http://dx.doi.org/10.1080/08120099.2017.1294109 Rajabi, M., Sherkati, S., Bohloli, B. and Tingay, M., 2010. Subsurface fracture analysis and determination of in-situ stress direction using FMI logs: An example from the Santonian 51

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

carbonates (Ilam Formation) in the Abadan Plain, Iran. Tectonophysics, 492(1–4): 192200. Rajabi, M. and Tingay, M., 2016. Pattern and Origin of the Present-Day Tectonic Stress in the Australian Sedimentary Basins. ASEG Extended Abstracts, 2016: 1-5. http://dx.doi.org/10.1071/ASEG2016ab275 Rajabi, M., Tingay, M. and Heidbach, O., 2016a. The present-day state of tectonic stress in the Darling Basin, Australia: Implications for exploration and production. Marine and Petroleum Geology, 77: 776-790. Rajabi, M., Tingay, M. and Heidbach, O., 2016b. The present-day stress field of New South Wales, Australia. Australian Journal of Earth Sciences, 63(1): 1-21. Rajabi, M., Tingay, M., Heidbach, O., Belton, D., Balfour, N. and Bendall, B., 2016c. New constraints on the neotectonic stress pattern of the Flinders and Mount Lofty Ranges, South Australia. Exploration Geophysics, doi: 10.1071/EG16076. Rajabi, M., Tingay, M., King, R. and Cooke, D., 2015. The present-day stress field of Australia: New release of the Australian Stress Map. ASEG Extended Abstracts, 2015(1): 1-3. http://dx.doi.org/10.1071/ASEG2015ab303. Rajabi, M., Tingay, M., King, R. and Heidbach, O., 2017b. Present-day stress orientation in the Clarence-Moreton Basin of New South Wales, Australia: a new high density dataset reveals local stress rotations. Basin Research, 29(S1): 622-640. Rajabi, M., Ziegler, M., Tingay, M., Heidbach, O. and Reynolds, S., 2016d. Contemporary tectonic stress pattern of the Taranaki Basin, New Zealand. Journal of Geophysical Research: Solid Earth, 121(8): 6053–6070. Raleigh, C.B., Healy, J.H. and Bredehoeft, J.D., 1972. Faulting and Crustal Stress at Rangely, Colorado. In: H.C. Heard, I.Y. Borg, N.L. Carter and C.B. Raleigh (Editors), Flow and Fracture of Rocks. U.S. Geological Survey, Washington, D. C., pp. 275-284. Ranalli, G. and Chandler, T.E., 1975. The stress field in the upper crust as determined from in situ measurements. Geologische Rundschau, 64(1): 653-674. Rasouli, V., Pervukhina, M., Müller, T.M. and Pevzner, R., 2013. In-situ stresses in the Southern Perth Basin at the GSWA Harvey-1 well site. Exploration Geophysics, 44(4): 289-298. Reinecker, J., Tingay, M.R.P. and Müller, B., 2006. The use of the WSM database for rock engineers. In: M. Lu, C.C. Li, H. Kjorholt and H. Dahle (Editors), In-Situ Rock Stress: Measurement, interpretation and application. Taylor & Francis, London, pp. 505-510. Reis, Á.F.C., Bezerra, F.H.R., Ferreira, J.M., do Nascimento, A.F. and Lima, C.C., 2013. Stress magnitude and orientation in the Potiguar Basin, Brazil: Implications on faulting style and reactivation. Journal of Geophysical Research: Solid Earth, 118(10): 5550-5563. Reiter, K. and Heidbach, O., 2014. 3-D geomechanical–numerical model of the contemporary crustal stress state in the Alberta Basin (Canada). Solid Earth, 5(2): 1123-1149. Reiter, K., Heidbach, O., Schmitt, D., Haug, K., Ziegler, M. and Moeck, I., 2014. A revised crustal stress orientation database for Canada. Tectonophysics, 636: 111-124. Revets, S.A., Keep, M. and Kennett, B.L.N., 2009. NW Australian intraplate seismicity and stress regime. Journal of Geophysical Research: Solid Earth, 114(B10): B10305. Reynolds, S.D., 2001. Characterization and modelling of the regional in situ stress field of continental Australia, The University of Adelaide, Adelaide, 216 pp. Reynolds, S.D., Coblentz, D.D. and Hillis, R., 2002. Tectonic forces controlling the regional intraplate stress field in continental Australia: Results from new finite element modeling. Journal of Geophysical Research: Solid Earth, 107(B7): ETG 1-1-ETG 1-15. Reynolds, S.D., Coblentz, D.D. and Hillis, R., 2003. Influences of plate-boundary forces on the regional intraplate stress field of continental Australia. Geological Society of America Special Papers, 372: 59-70. Reynolds, S.D. and Hillis, R., 2000. The in situ stress field of the Perth Basin, Australia. Geophysical Research Letters, 27(20): 3421-3424. Reynolds, S.D., Mildren, S.D., Hillis, R. and Meyer, J.J., 2006. Constraining stress magnitudes using petroleum exploration data in the Cooper–Eromanga Basins, Australia. Tectonophysics, 415(1–4): 123-140. 52

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Reynolds, S.D., Mildren, S.D., Hillis, R., Meyer, J.J. and Flottmann, T., 2005. Maximum horizontal stress orientations in the Cooper Basin, Australia: implications for plate-scale tectonics and local stress sources. Geophysical Journal International, 160(1): 331-343. Richardson, R.M., 1992. Ridge forces, absolute plate motions, and the intraplate stress field. Journal of Geophysical Research: Solid Earth, 97(B8): 11739-11748. Richardson, R.M., Solomon, S.C. and Sleep, N.H., 1976. Intraplate stress as an indicator of plate tectonic driving forces. Journal of Geophysical Research, 81(11): 1847-1856. Richardson, R.M., Solomon, S.C. and Sleep, N.H., 1979. Tectonic stress in the plates. Reviews of Geophysics, 17(5): 981-1019. Roth, F. and Fleckenstein, P., 2001. Stress orientations found in north-east Germany differ from the West European trend. Terra Nova, 13(4): 289-296. Sahara, D.P., Schoenball, M., Kohl, T. and Müller, B.I.R., 2014. Impact of fracture networks on borehole breakout heterogeneities in crystalline rock. International Journal of Rock Mechanics and Mining Sciences, 71(0): 301-309. Sandiford, M., 2003. Neotectonics of southeastern Australia: linking the Quaternary faulting record with seismicity and in situ stress. Geological Society of America Special Papers, 372: 107-119. Sandiford, M., 2007. The tilting continent: A new constraint on the dynamic topographic field from Australia. Earth and Planetary Science Letters, 261(1-2): 152-163. Sandiford, M., Wallace, M. and Coblentz, D.D., 2004. Origin of the in situ stress field in southeastern Australia. Basin Research, 16(3): 325-338. Sbar, M.L. and Sykes, L.R., 1973. Contemporary Compressive Stress and Seismicity in Eastern North America: An Example of Intra-Plate Tectonics. Geological Society of America Bulletin, 84(6): 1861-1882. Schmitt, D.R., Currie, C.A. and Zhang, L., 2012. Crustal stress determination from boreholes and rock cores: Fundamental principles. Tectonophysics, 580(0): 1-26. Schulte, S.M. and Mooney, W.D., 2005. An updated global earthquake catalogue for stable continental regions: reassessing the correlation with ancient rifts. Geophysical Journal International, 161(3): 707-721. Seeber, L. and Armbruster, J.G., 2000. Earthquakes as beacons of stress change. Nature, 407(6800): 69-72. Segall, P., 1989. Earthquakes triggered by fluid extraction. Geology, 17(10): 942-946. Seithel, R., Steiner, U., Müller, B., Hecht, C. and Kohl, T., 2015. Local stress anomaly in the Bavarian Molasse Basin. Geothermal Energy, 3(1): 1-22. Sibson, R., Ghisetti, F. and Ristau, J., 2011. Stress Control of an Evolving Strike-Slip Fault System during the 2010–2011 Canterbury, New Zealand, Earthquake Sequence. Seismological Research Letters, 82(6): 824-832. Sibson, R.H., 1992. Implications of Fault-Valve Behavior for Rupture Nucleation and Recurrence. Tectonophysics, 211(1-4): 283-293. Sibson, R.H., Ghisetti, F.C. and Crookbain, R.A., 2012. Andersonian wrench faulting in a regional stress field during the 2010–2011 Canterbury, New Zealand, earthquake sequence. Geological Society, London, Special Publications, 367(1): 7-18. Simons, F.J., van der Hilst, R.D., Montagner, J.-P. and Zielhuis, A., 2002. Multimode Rayleigh wave inversion for heterogeneity and azimuthal anisotropy of the Australian upper mantle. Geophysical Journal International, 151(3): 738-754. Sippl, C., Kennett, B.L.N., Tkalčić, H., Spaggiari, C.V. and Gessner, K., 2015. New constraints on the current stress field and seismic velocity structure of the eastern Yilgarn Craton from mechanisms of local earthquakes. Australian Journal of Earth Sciences, 62(8): 921-931. Sjöberg, J., Christiansson, R. and Hudson, J.A., 2003. ISRM Suggested Methods for rock stress estimation—Part 2: overcoring methods. International Journal of Rock Mechanics and Mining Sciences, 40(7–8): 999-1010. SLU Earthquake Center, 2015. Saint Louis University Earthquake Center. URL:http://www.eas.slu.edu/eqc/.

53

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Solomon, S.C., Sleep, N.H. and Richardson, R.M., 1975. On the Forces Driving Plate Tectonics: Inferences from Absolute Plate Velocities and Intraplate Stress. Geophysical Journal of the Royal Astronomical Society, 42(2): 769-801. Sonder, L.J., 1990. Effects of density contrasts on the orientation of stresses in the lithosphere: Relation to principal stress directions in the Transverse Ranges, California. Tectonics, 9(4): 761-771. Sperner, B., Müller, B., Heidbach, O., Delvaux, D., Reinecker, J. and Fuchs, K., 2003. Tectonic stress in the Earth’s crust: advances in the World Stress Map project. Geological Society, London, Special Publications, 212(1): 101-116. Stein, R.S., 1999. The role of stress transfer in earthquake occurrence. Nature, 402(6762): 605609. Steinberger, B., Schmeling, H. and Marquart, G., 2001. Large-scale lithospheric stress field and topography induced by global mantle circulation. Earth and Planetary Science Letters, 186(1): 75-91. Swierczek, E., Holford, S. and Backè, G., 2015. Stress orientation and subsurface natural fractures model along the Rosedale Fault System, Gippsland Basin, Australia—Implications for fault related fluid migration along inverted normal fault planes, International Conference and Exhibition. Society of Exploration Geophysicists and American Association of Petroleum Geologists, Melbourne, Australia, pp. 564. doi: 10.1190/ice2015-2211387. Sykes, L.R. and Sbar, M.L., 1973. Intraplate Earthquakes, Lithospheric Stresses and the Driving Mechanism of Plate Tectonics. Nature, 245(5424): 298-302. Tassone, D.R., Holford, S.P., Duddy, I.R., Green, P.F. and Hillis, R., 2014. Quantifying Cretaceous–Cenozoic exhumation in the Otway Basin, southeastern Australia, using sonic transit time data: Implications for conventional and unconventional hydrocarbon prospectivity. AAPG Bulletin, 98(1): 67-117. Tassone, D.R., Holford, S.P., Hillis, R. and Tuitt, A.K., 2012. Quantifying Neogene plateboundary controlled uplift and deformation of the southern Australian margin. Geological Society, London, Special Publications, 367(1): 91-110. Tingay, M., 2009. State and origin of present-day stress fields in sedimentary basins. ASEG Extended Abstracts, 2009(1): 1-10. Tingay, M., Bentham, P., De Feyter, A. and Kellner, A., 2011. Present-day stress-field rotations associated with evaporites in the offshore Nile Delta. Geological Society of America Bulletin, 123(5-6): 1171-1180. Tingay, M., Bentham, P., De Feyter, A. and Kellner, A., 2012. Evidence for non-Andersonian faulting above evaporites in the Nile Delta. Geological Society, London, Special Publications, 367(1): 155-170. Tingay, M., Heidbach, O., Davies, R. and Swarbrick, R., 2008. Triggering of the Lusi mud eruption: Earthquake versus drilling initiation. Geology, 36(8): 639-642. Tingay, M., Morley, C., King, R., Hillis, R., Coblentz, D.D. and Hall, R., 2010a. Present-day stress field of Southeast Asia. Tectonophysics, 482(1–4): 92-104. Tingay, M., Muller, B., Reinecker, J. and Heidbach, O., 2006. State and Origin of the Present-Day Stress Field in Sedimentary Basins: New Results from the World Stress Map Project, The 41st U.S. Symposium on Rock Mechanics (USRMS). American Rock Mechanics Association, Golden, Colorado Tingay, M., Müller, B., Reinecker, J., Heidbach, O., Wenzel, F. and Fleckenstein, P., 2005a. Understanding tectonic stress in the oil patch: The World Stress Map Project. The Leading Edge, 24(12): 1276-1282. Tingay, M.R.P., Hillis, R., Morley, C.K., King, R.C., Swarbrick, R.E. and Damit, A.R., 2009. Present-day stress and neotectonics of Brunei: Implications for petroleum exploration and production. AAPG Bulletin, 93(1): 75-100. Tingay, M.R.P., Hillis, R., Morley, C.K., Swarbrick, R.E. and Drake, S.J., 2005b. Present-day stress orientation in Brunei: a snapshot of ‘prograding tectonics’ in a Tertiary delta. Journal of the Geological Society, 162(1): 39-49. Tingay, M.R.P., Morley, C.K., Hillis, R. and Meyer, J., 2010b. Present-day stress orientation in Thailand's basins. Journal of Structural Geology, 32(2): 235-248. 54

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Tingay, M.R.P., Rudolph, M.L., Manga, M., Davies, R.J. and Wang, C.-Y., 2015. Initiation of the Lusi mudflow disaster. Nature Geosci, 8(7): 493-494. USGS, 2015. United States Geological Survey’s (USGS) Earthquake Hazards Program. URL:http://earthquake.usgs.gov/. Wang, X., Holt, W.E. and Ghosh, A., 2015. Joint modeling of lithosphere and mantle dynamics: Evaluation of constraints from global tomography models. Journal of Geophysical Research: Solid Earth, 120(12): 8633-8655. Warpinski, N.R., Mayerhofer, M.J., Vincent, M.C., Cipolla, C.L. and Lolon, E., 2008. Stimulating Unconventional Reservoirs: Maximizing Network Growth While Optimizing Fracture Conductivity, SPE Unconventional Reservoirs Conference. Society of Petroleum Engineers, Keystone, Colorado, USA. Weatherley, D.K. and Henley, R.W., 2013. Flash vaporization during earthquakes evidenced by gold deposits. Nature Geosci, 6(4): 294-298. Wessel, P., Smith, W.H.F., Scharroo, R., Luis, J. and Wobbe, F., 2013. Generic Mapping Tools: Improved Version Released. Eos, Transactions American Geophysical Union, 94(45): 409-410. Willson, S.M., 2012. A Wellbore Stability Approach For Self-Killing Blowout Assessment, SPE Deepwater Drilling and Completions Conference. Society of Petroleum Engineers, Galveston, Texas, USA, pp. 16. Worotnicki, G. and Denham, D., 1976. The state of stress in the upper part of the earth's crust in Australia according to measurements in mines and tunnels and from seismic observations, Symposium on Investigation of Stress in Rock: Advances in Rock Measurment Institution of Engineers, Australia., pp. 71-82. Wyborn, D., 2012. The Innamincka Enhanced Geothermal System (EGS) – dealing with the overpressures. In: H.-H. C. and G. E. (Editors), The 2012 Australian Geothermal Energy Conference. Geoscience Australia (Record 2012/73), Coogee Beach, Sydney, pp. 251258. Yale, D.P., 2003. Fault and stress magnitude controls on variations in the orientation of in situ stress. Geological Society, London, Special Publications, 209(1): 55-64. Yassir, N.A. and Zerwer, A., 1997. Stress regimes in the Gulf Coast, offshore Louisiana; data from well-bore breakout analysis. AAPG Bulletin, 81(2): 293-307. Zang, A., Oye, V., Jousset, P., Deichmann, N., Gritto, R., McGarr, A., Majer, E. and Bruhn, D., 2014. Analysis of induced seismicity in geothermal reservoirs – An overview. Geothermics, 52: 6-21. Zang, A. and Stephansson, O., 2010. Stress Field of the Earth's Crust. Springer Dordrecht Heidelberg London New York, Springer, 322 pp. Zang, A., Stephansson, O., Heidbach, O. and Janouschkowetz, S., 2012. World Stress Map Database as a Resource for Rock Mechanics and Rock Engineering. Geotechnical and Geological Engineering, 30(3): 625-646. Zhang, Y., Scheibner, E., Hobbs, B.E., Ord, A., Drummond, B.J. and Cox, S.J.D., 1998. Lithospheric Structure in Southeast Australia: a Model Based on Gravity, Geoid and Mechanical Analyses, Structure and Evolution of the Australian Continent. American Geophysical Union, pp. 89-108. Zhang, Y., Scheibner, E., Ord, A. and Hobbs, B., 1996. Numerical modelling of crustal stresses in the eastern Australian passive margin. Australian Journal of Earth Sciences, 43(2): 161175. Zhao, S. and Müller, R.D., 2001. The tectonic stress field in eastern Australia, Eastern Australasian Basins Symposium, A Refocussed Energy Perspective for the Future. Petroleum Exploration Society of Australia Special Publication, pp. 61-70. Zhao, S. and Müller, R.D., 2003. Three-dimensional finite-element modelling of the tectonic stress field in continental Australia. Geological Society of America Special Papers, 372: 71-89. Ziegler, M., Rajabi, M., Heidbach, O., Hersir, G.P., Ágústsson, K., Árnadóttir, S. and Zang, A., 2016a. The stress pattern of Iceland. Tectonophysics, 674: 101-113. Ziegler, M., Reiter, K., Heidbach, O., Zang, A., Kwiatek, G., Stromeyer, D., Dahm, T., Dresen, G. and Hofmann, G., 2015. Mining-Induced Stress Transfer and Its Relation to a Mw 1.9 55

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

NU

SC

RI

PT

Seismic Event in an Ultra-deep South African Gold Mine. Pure and Applied Geophysics, 172(10): 2557-2570. Ziegler, M.O., Heidbach, O., Reinecker, J., Przybycin, A.M. and Scheck-Wenderoth, M., 2016b. A multi-stage 3-D stress field modelling approach exemplified in the Bavarian Molasse Basin. Solid Earth, 7(5): 1365-1382. Zoback, M. and Zoback, M.L., 1991. Tectonic stress field of North America and relative plate motions. In: D.B. Slemmons, E.R. Engdahl, M. Zoback and D. Blackwell, D. (Editors), Neotectonics of North America. DNAG, Continent-Scale Map Series. Geological Society of America, Boulder, Colorado, pp. 339-366. Zoback, M.D., 2007. Reservoir Geomechanics. Cambridge University Press, New York, 449 pp. Zoback, M.D., Zoback, M.L., Mount, V.S., Suppe, J., Eaton, J.P., Healey, J.H., Oppenheimer, D., Reasenberg, P., Jones, L., Raleigh, C.B., Wong, I.G., Scotti, O. and Wentworth, C., 1987. New Evidence on the State of Stress of the San Andreas Fault System. Science, 238(4830): 1105-1111. Zoback, M.L., 1992. First- and second-order patterns of stress in the lithosphere: The World Stress Map Project. Journal of Geophysical Research: Solid Earth, 97(B8): 11703-11728. Zoback, M.L. and Zoback, M., 1980. State of stress in the conterminous United States. Journal of Geophysical Research: Solid Earth, 85(B11): 6113-6156. Zoback, M.L., Zoback, M.D., Adams, J., Assumpcao, M., Bell, S., Bergman, E.A., Blumling, P., Brereton, N.R., Denham, D., Ding, J., Fuchs, K., Gay, N., Gregersen, S., Gupta, H.K., Gvishiani, A., Jacob, K., Klein, R., Knoll, P., Magee, M., Mercier, J.L., Muller, B.C., Paquin, C., Rajendran, K., Stephansson, O., Suarez, G., Suter, M., Udias, A., Xu, Z.H. and Zhizhin, M., 1989. Global patterns of tectonic stress. Nature, 341(6240): 291-298.

56

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

Figure 1

57

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

Figure 2

58

AC

CE

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 3

59

Figure 4

AC

CE

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

60

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

Figure 5

61

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

PT E

Figure 6

62

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

Figure 7

63

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

Figure 8

64

AC

CE

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 9

65

AC

CE

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 10 66

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

Figure 11

67

AC

CE

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 12

68

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

Figure 13

69

AC

CE

PT E

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 14

70

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC

CE

PT E

D

MA

Figure 15

71

ACCEPTED MANUSCRIPT

Table 1: Data type and quality of the present-day stress data records in the Australian stress map database and a comparison between the latest release of the project in 2003 and the new database.

SC

RI

PT

New (2016) release of ASM A-E A-C 2 2 643 551 5 0 1027 426 332 101 81 62 58 5 2 0 2150 1147

NU

Data type FMF FMS FMA BO DIF HF OC GVA Total

2003 release of ASM A-E A-C 1 1 122 95 5 0 258 146 27 24 78 59 56 6 2 0 549 331

A-C and A-E: Quality assignment from the World Stress Map quality ranking scheme for the orientation of

MA

maximum horizontal stress (Heidbach et al., 2010). FMF: formal inversion of focal mechanism solutions, FMS: single focal mechanism solutions, FMA: composite or average focal mechanism solutions, BO: borehole breakouts, DIF: drilling induced-tensile fractures, HF: hydraulic fracture measurements, OC:

AC

CE

PT E

D

overcoring, GVA: geological volcanic alignments.

72

ACCEPTED MANUSCRIPT Table 2: Present-day stress provinces across continental Australia and Papua New Guinea.

Bight Basin

Canning Basin

D

Mean SHmax

s. d.

1 1

0

6

0

0

1

7

9

9

27

31

4

4

0

0

0

0

0

1

3

1

134

19

5 5

3 4

1 9

2

0

0

1 1

2 5

1 9

3 5

4

4

0

0

0

0

3

1

0

1

6 4

1 7

1 3

3

3 1

0

7

2 2

3 5

5 2

4 9

3

0

0

0

5

1 6

3 1

2 8 1 0

1 7

1 0

1

0

0

8

1 3

7

7

0

2

1

0

2

2

6

6 0

4 8

6

6

1 6

0

0

16

2 1

2 1

7 8

6 2

Irian Jaya 1 Irian Jaya 2 McArthur Region New Guinea 1

PT

A B C

R

Conf .

0. 57 0. 81

99.0 0% 90.0 0%

Provi nce's Type 2 5

16

0. 85

99.9 0%

1

57

4

0. 99

99.0 0%

2

32

30

0. 58

99.9 0%

1

65

33

0. 51

99.9 0%

1

92

24

1

53

11

0. 70 0. 93

99.9 0% 99.9 0%

RI

47

SC

8 9 1 2 4 1 1

1 1

0

0

1 2

2 3

2 5

6 8

108

24

0. 71

99.9 0%

1

0

0

0

0

1 6

0

109

43

0. 32

<90 %

6

0

0

0

0

1

1 6

4

1 2

69

23

0. 71

99.9 0%

1

1 6

0

0

0

2 8

3 0

2 0

5 8

99

14

0. 89

99.9 0%

1

0. 93 0. 63 0. 96 0. 72 0. 19 0. 51

99.9 0% 99.0 0% 99.0 0% 99.9 0% <90 % 99.9 0%

0. 60

99.9 0%

0. 76 0. 95 0.

99.9 0% 99.9 0% 99.9

8

2

1

1

1

3

2

1

5

2

91

11

1 9

5

0

13

1

0

2

3

1 4

6 2

91

27

5

2

2

0

0

1

1

1

3

1

87

8

1 7

1 5

2

0

0

0

4

3

1 0

2 1

133

23

5

5

0

0

0

0

0

2

3

6

34

52

3 0 1 3 5 9 9

2 5

3

2

0

0

1

6

6 1

57

33

0

0

135

0

0

0

0

0

18

29

0

0

99

0

0

0

0

2 3 1 3 5 9 9

0

40

21

6

3

0

0

2

1

2

3

1

5

26

9

1

0

0

151

0

0

0

0

1

0

40

23

AC

Northern Carnarvon Basin Southern Carnarvon Basin ClarenceMoreton Basin CooperEromanga Basins Darling Region Flinders Ranges GalileeEromanga Gippsland Basin Gloucester Basin Gunnedah Basin

O C

NU

Browse Basin

H F

D

Bowen-Surat 2

B O

FMS/F MF/FM A

PT E

Northern Bonaparte Basin Southern Bonaparte Basin Bowen-Surat 1

Statistics using A-C quality data records

Quality

D I F

CE

Province

Type (A-C)

MA

Amadeus Basin Region

N o. A C 1 7

73

1 1 2 1 6 1 1 1 1 1

ACCEPTED MANUSCRIPT 5 1 6 3 1 9

0

0

63

0

0

0

0

1 4

5

0

0

0

7

4

8

5

3

0

0

0

4

Officer Basin

6

4

2

0

0

0

Perth Region

6 4

4 0

1 5

8

1

1 4

0

0

12

1 2

0

3

0

8

1 4

131

13

3

1

6

137

5

1

2

3

5

95

17

0

8

2 3

3 3

2 5

88

18

2

0

0

1

1 3

1 5

4

1 6

0

2

1 1

1 9

2

6

0

1

2

8

7 7 2 0

MA D PT E CE AC

74

PT

3 2 1 1

27

NU

SE Seismogenic Zone North Sydney Basin South Sydney Basin

25

122

41

65

40

41

35

RI

SA Otway Basin Vic Otway Basin

0

SC

New Guinea 2

5 1 6 3

73

0%

0. 64 0. 91 0. 98 0. 84 0. 82

99.9 0% 99.9 0% 99.9 0% 99.0 0% 99.9 0%

0. 37

<90 %

0. 39 0. 48

99.0 0% 90.0 0%

1 1 1 2 1 6 2 5

ACCEPTED MANUSCRIPT HIGHLIGHTS 

An extensive review of 25 years of research on the pattern of stress in the Australian continent.



New release of the Australian Stress Map project with 2150 data records in 30 geological provinces across Australia.

PT

Characterisation of the controls on the Australian stress pattern from plate to wellbore scale.

RI

 Globally relevant classification scheme of the orders of stress patterns and associated

CE

PT E

D

MA

NU

SC

stress sources in the lithosphere.

AC



75