fault seal analysis: constraining fault seal risk using ...fault seal capacity is an important...

107
i Fault Seal Analysis: Constraining fault seal risk using seismic velocities Tobias Colson Supervisor: Prof. Myra Keep Co-Supervisor: Adj. Prof. David Castillo This thesis is submitted to fulfil the requirements for Master of Science (Geology) by way of Thesis and Coursework Faculty of Science October 2014

Upload: others

Post on 17-Apr-2021

24 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

i

Fault Seal Analysis:

Constraining fault seal risk using seismic velocities

Tobias Colson

Supervisor: Prof. Myra Keep

Co-Supervisor: Adj. Prof. David Castillo

This thesis is submitted to fulfil the requirements for Master of Science (Geology) by way of

Thesis and Coursework

Faculty of Science

October 2014

Page 2: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

ii

Page 3: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

iii

TABLE OF CONTENTS

Acknowledgments ......................................................................................................................... v

Glossary ........................................................................................................................................ vii

Abstract.......................................................................................................................................... ix

1. Introduction ............................................................................................................................... 1

2. Background ................................................................................................................................ 2

2.1 The fault and capillary seal ............................................................................................... 2

2.2 The application of seismic data ......................................................................................... 7

2.3 Geological structure of the study area - Dampier Sub-basin and Rankin Trend, North West Shelf ..................................................................................................................... 10

3. Aims (and objectives): ............................................................................................................ 16

4. Methodology: ........................................................................................................................... 17

5. Results ....................................................................................................................................... 24

5.1 Structural Characteristics of the Rankin Dockrell Keast Blocks - Rankin Trend . 24

5.2 Velocity Analysis ............................................................................................................... 44

6. Discussion/Analysis ................................................................................................................. 66

7. Conclusions .............................................................................................................................. 78

8. Further work ........................................................................................................................... 79

9. Appendix A -Tectonostratigraphic diagram ...................................................................... 80

10. Appendix B – Stratigraphic column for the West Dampier Sub-basin ........................ 81

11. Appendix C – Porosity and permeability in the NCB ..................................................... 82

12. Appendix D - Pore pressure ................................................................................................ 86

13. Appendix E - Empirical calculation of shear wave velocity ........................................... 89

14. Appendix F - Shear and normal stress .............................................................................. 91

15. Appendix G – Thermal gradient for NCB ........................................................................ 92

16. References .............................................................................................................................. 93

Page 4: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

iv

Page 5: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

v

ACKNOWLEDGMENTS

This project was supported by BP Developments Australia Pty Ltd. Many thanks to Ben

Boterhoven, Vincent Dykman and Gerchard Pfau in the North West Shelf Joint Venture

Exploration team of BP for their assistance during this research.

Thanks to Prof. Myra Keep for supervising the project and co-supervisor Adj. Prof. David

Castillo for advice.

Finally to my MSc research peers in petroleum exploration geology, M.S. and C.vR.,

Prodeo Quo Victum !

T.Colson 2014

Page 6: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

vi

Page 7: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

vii

GLOSSARY

Checkshot - Seismic Velocity Checkshot Survey - Used to calibrate time-depth curves by conducting a series of sound shots with a receiver at various depths in the well. CFF - Coulomb Failure Function. FIT - Formation Integrity Test - used to test the strength of a formation by increasing the bottom hole pressure in a well. Where LOT tests are conducted and 'Leak off' not achieved, a FIT test has been conducted, often quoted as 'Max Pressure'. LOT - Leak Off Test - A test to determine the strength or fracture pressure of the open formation, usually conducted immediately after drilling below a new casing shoe. During the test, the well is shut in and fluid is pumped into the wellbore to gradually increase the pressure until the fluid enters the formation, or leaks off, either moving through permeable paths in the rock or by creating a space by fracturing the rock. The results of the leak-off test dictate the maximum pressure or mud weight that may be applied to the well during drilling operations. NCB - North Carnarvon Basin. NWS - North West Shelf, Western Australia. RFT/FIT - Repeat Formation Tester/Formation Interval Test - Tool/measurement used to assess formation fluid pressure, density and other fluid properties (FIT is used in older Wells). Sonic Log - Well Log used to measure interval transit time by sending out a high-frequency sound signal and receiving a return signal at a receiver in the same tool. Resolution and frequency are far higher than with a checkshot, so it has important scale dependent differences compared with the checkshot. Stacks - Stacking Velocities. Once seismic traces are collected according to a common mid-point (CMP), their velocities are corrected for the normal hyperbolic moveout, and combined/stacked to improve the signal-to-noise ratio. Stacks can be interval velocities, RMS or average velocities. Shmax - Maximum horizontal stress Shmin - Minimum horizontal stress

Page 8: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

viii

Page 9: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

ix

ABSTRACT

Fault seal capacity is an important component in the conventional petroleum system.

Assessing the capacity for a fault to seal or leak can be difficult, particularly where well

constraint is lacking. In the frontier basin, in a marine setting, seismic velocities may be the

only data available. However, useful constraints on a faults sealing capacity can be extracted

from this data alone. This study investigates the robustness of a number of empirical relations

that can assist in extracting useful constraints from seismic velocities and amplitudes.

Reliable estimates on maximum and minimum stress tensors and pore pressures can be

calculated and combined with basic fault architecture analysis, to place practical constraints

on fault risk. In this study of an area on the Rankin Trend found good correlation between

well-based and seismic velocity-based pore pressures and stress magnitudes allowing a

coulomb failure function based only on stacking velocities to be calculated. Faults separating

the Rankin 1 well block from the Dockrell/Keast Field, were shown to be within a stable

stress regime. Calculated pore pressures match known RFT measurements and show that

overpressure can be assessed using basic time-migrated velocity stacks. Furthermore,

theoretical capillary pressure and hydrocarbon column calculations correlate with known

values and highlight the capillary seal potential of sand-sand juxtaposition seals which can

support 100 to 200 m column heights within the Triassic play. They suggest that many traps

on the Rankin Trend within the Triassic play are limited by a combination of pore pressure

and stress orientation. However, deeper intra-formational seals are likely to have increased

seal capacity by virtue of overburden stress and reduced porosity exceeding any capacity for

shear stress failure. Well data for this area confirms a Shmax orientation of approximately 110

+/- 10o and calculations show that faults striking within 20° of this direction may be at high

risk of failure within the neotectonic setting, where pore pressures and dip predicate fault slip.

Page 10: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

x

Page 11: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

1

1. INTRODUCTION

The primary components of a petroleum system include source, seal, trap, reservoir and a

mechanism for migration of petroleum. These factors must all be assessed, not just for their

presence, but for the degree to which they are associated and the sequence in which they were

formed. Key to the effectiveness of a petroleum system is the extent to which secondary

migration and charge access can facilitate the linking of these primary components. An

element of assessing the risk factors associated with secondary migration and charge access is

the capacity for faults to inhibit or facilitate secondary migration, as well as their capacity to

contribute to trap formation.

Many petroleum exploration targets that are drilled on faulted traps and fail to return positive

results have a lack of petroleum accumulation attributed to charge access. Charge access

relates to the capacity for secondary migration to facilitate the 'filling' of a defined trap.

However, the contribution that fault seal has played in this is poorly understood.

Advancement in reservoir petroleum management, particularly with 4D time-dependent

seismic monitoring, has advanced our understanding of why faults fail when they do.

However, the ability to make predictions of fault seal capacity from the geometry and seismic

velocities alone allows more accurate risk factors to be applied in the 'green fields' stage of oil

and gas field exploration.

This study addresses some key constraints used in assessing whether a fault is likely to seal or

leak. Principal to these constraints is the typical lack of well control in frontier exploration. A

first-pass methodology based on seismic data alone enables better defined risking parameters

to be placed on the fault component of the petroleum system analysis where well control is

lacking.

Page 12: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

2

2. BACKGROUND

2.1 THE FAULT AND CAPILLARY SEAL

Secondary migration refers to the movement of petroleum from its source rock (in the

'kitchen') to a site where it accumulates (the trap). Secondary migration (movement through a

carrier bed) differs from primary migration (out of the source rock) via the porosity,

permeability and pore size distribution in which the petroleum migrates.

The process of secondary migration is essentially one of fluid movement. Density and

viscosity differences notwithstanding, the fundamental principles of fluid mechanics are

applied equally, whether the migration of gas, oil or water is being assessed. These principles

have being applied in studies on migration and seal capacity by Hubbert (1953), Smith

(1966), Berg (1975) and Schowalter (1979). Watts (1987) further developed cap (top seal) and

fault rock seal capacity based on capillary leakage in which cap rock seals are divided into

those that fail by capillary leakage and those that fail by hydraulic fracturing due to capillary

entry pressures being higher than the tensile strength of the sealing lithology.

Capillary leakage refers to interstitial pore pressure exceeding the capillary pressure of the

formation and subsequently leaking. Capillary pressure exists whenever two immiscible

phases are present in a fine bore tube or analogous void or pore space in a rock. Capillary

pressure is then defined as the pressure drop or differential across a curved liquid interface.

Watts (1987) shows how capillary pressure Pc in a pore space can be calculated if the fluid

interfacial tension γ, rock-fluid contact angle θ and the pore radius, r, are known;

𝑷𝑪 =𝟒𝜸 𝐜𝐨𝐬(𝜽)

𝒓 ... Eqn. 1

Alternatively, capillary pressure can be expressed as a hydrostatic head. It is equal to the

product of the height of the liquid (column height) the density difference between the two

liquids and the gravitational constant g ;

𝒉𝒄 = 𝑯𝒈(𝝆𝒘 − 𝝆𝒄)(Watts, 1987) ...Eqn. 2

For capillary seals, Watts (1987) noted that cap-rock thickness becomes unimportant once the

capillary slug pinch-off length is exceeded. The seal capacity of a hydraulic seal on the other

Page 13: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

3

hand is highly dependent on thickness, together with the in-situ stress state of the sealing layer

and the degree of overpressure development.

The concept of cap rock failure by capillary leakage or hydraulic fracturing is directly

transferable to our understanding of whether a fault will seal and inhibit, or leak and/or

facilitate secondary migration. Sealing faults can thus be thought of as a capillary cap rock at

an angle.

Fault related seals may be further divided into those where the fault plane itself acts as a seal

and those where the sealing unit is laterally juxtaposed against reservoir sands (‘juxtaposition

seals’). Most if not all fault seals will preferentially leak via capillary leakage rather than

hydraulic fracture (not including leakage along the fault) (Aydin, 2000; Bailey et al., 2006;

Dodds et al., 2007; Shipton et al., 2005; Tueckmantel et al., 2010; Watts, 1987; Shepherd,

2009)

The concept of capillary pressure or membrane seal potential and associated pressure

differential is a direct reflection of the permeability of the fault in question. The permeability

is known to be highly dependent on the clay content of the fault (Aydin, 2000; Berg, 1975;

Bretan et al., 2003; Busch and Amann-Hildenbrand, 2013; Dehandschutter et al., 2005;

Dewhurst et al., 2002; Fisher et al., 2001; Fisher and Knipe, 2001; Kachi et al., 2005;

Ngwenya et al., 2000; Schowalter, 1979; Shipton et al., 2005; Watts, 1987) and clay content

in particular is a primary control on seal behaviour of faults in mixed clastic sequences

(Ngwenya et al., 2000; Torabi et al., 2013; Tueckmantel et al., 2010).

There are a number of empirical based algorithms that are widely used to attempt to predict

seal capacity based on the clay component of a fault. These include, the Shale Gouge Ratio

(SGR), the Smear Gouge Ratio (SMGR), the Shale Smear Factor (SSF) and the Clay Smear

Potential (CSP) (Allen and Allen, 2005; Childs et al., 2009; Kachi et al., 2005; Yielding et al.,

1997). The shale gouge ratio (SGR) is one of the more widely used algorithms and is based on

the assertion that the sealing capacity of a fault is directly related to the number of shale beds

and therefore amount of clay material within the slipped interval. If clay is incorporated in the

fault gouge, the clay will increase the membrane seal potential of the fault and increase the

chances of an effective seal.

Page 14: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

4

The studies by Fisher and Knipe (2001) directly attributed increased clay percentage in the

fault gouge to decrease in permeability of the fault. However, the concept assumes a

homogeneous distribution of the percentage of clay within the fault, an assumption that

develops an inherent weakness in this approach for predicting seal potential. Studies have

shown consistently that there will generally be a high degree of variability in the amount of

clay within a fault (Bailey et al., 2006; Bretan et al., 2003; Fisher et al., 2001; Fisher and

Knipe, 2001; Torabi et al., 2013). Furthermore, this method inherently requires an

understanding of the amount of clay that has been incorporated within the fault gouge. Whilst

mature basins, where core analysis is available, benefit from a deterministic petro-physical

constraint placed on SGR estimates, new exploration fields generally lack this information.

Without robust clay component estimates, SGR parameters can become misleading. SGR

estimates suggest that faults with shale gouge of <15% clay component will tend to leak

(Yielding et al., 1997). However, quartz dissolution-related diagenetic processes can

contribute up to 2 orders of magnitude in permeability reduction in faults (Fisher and Knipe,

2001; Ngwenya et al., 2000; Tueckmantel et al., 2010). SGR estimates do not take this into

account, nor do they take into account increasing confining pressures with depth. The greater

a cap rock of any kind is buried, the greater the compaction and cementation contribution to

permeability reduction will be.

Empirically derived equations that attempt to incorporate factors such as cementation and

increased confining pressures are highly dependent on the data set from which they are based

and have limited applicability to different tectonic settings, fault types and lithologies.

Nonetheless the SGR and SMGR, where reliable estimates of clay content are available, can

provide a first pass assessment, and studies that show a reduced permeability in fault gouge

with increasing clay content suggest that higher clay content should indicate higher Pc. A

higher Pc alone, on the other hand, will not necessarily be indicative of a sealing fault,

because if the pressure differential, ΔP, across the fault exceeds Pc, the fault will still leak

(Fig. 1) (Fisher and Knipe 2001).

Thus, if the capacity for cross fault migration depends on the pressure differential, it then

depends on how well a trap on one side of the fault is charged as to whether a trap will

equilibrate on both sides of the fault (Fisher et al., 2001). Studies also show the sealing

capacity of a fault should be calculated from the difference in pressure between the

Page 15: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

5

hydrocarbon and pore-water at the position along the fault where leakage is most likely to

occur (Fisher et al., 2001).

Figure 1 Representation of a typical fault zone with fault gouge in the damage zone depicted

as varying from siliceous cataclastic gouge to shale smear gouge with a higher clay content

(from Shipton et al., 2005). Pc denotes the capillary pressure of the gouge, and p1 and p2

denote the host rock or reservoir pressure, where Δp is the difference in pressure from one

side of the fault to the other.

The pressure differential can therefore provide an indication as to whether there is fluid

communication on one side of the fault to the other. However, as (Fisher and Knipe, 2001)

and Fisher (2001) note, pressure differences and heights of hydrocarbon columns are often

more likely to be controlled by factors such as the capillary entry pressure differences of the

undeformed reservoir lithologies and the amount of hydrocarbons entering the reservoir than

Pc

P1

P2

ΔP = P1-P2

Page 16: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

6

by the actual capillary entry pressure of the fault rock itself. So whilst the pressure difference

will not necessarily reflect the capillary entry pressure of the fault gouge, it may provide an

indication of a potential effective fault seal. This can be complicated since a lack of pressure

difference does not necessarily mean that the fault is leaking. Without knowledge of absolute

hydrocarbon column height potential of the cap rock, a ΔP of zero may mean that that there

has been sufficient charge to fill the trap up to the maximum capillary pressure of the fault

gouge, and leakage has occurred whenever the pressure exceeds Pc max. This would present a

repetitive leak-seal sequence analogous to a ‘fill to spill’ scenario. Given such a scenario, an

assessment would ideally be made on pressure differences between hydrocarbon and pore

water pressures in areas with minimum shale/clay component as this would have minimal

capillary pressure potential and be where the fault was most likely to leak.

There are numerous studies (Aydin, 2000; Bense et al., 2013; Losh, 1998) that ascribe fault

gouge acting as a fluid conduit for hydrocarbon migration without providing significant direct

evidence. Rather, they conclude that fault formation and trap formation are related and

typically infer that fluid conduit migration ‘probably’ facilitated the migration pathway.

Studies do show that fault-related dilatant fractures can increase permeability (Aydin, 2000;

Dewhurst and Hennig, 2003; Dewhurst et al., 2002; Færseth et al., 2007; Ngwenya et al.,

2000; Shepherd, 2009), which, along with mode II failure, places a dependency of fluid

migration on the seismicity of the fault and a dependency on conditions necessary for

reactivation.

The magnitudes of the maximum principal stress and the minimum cohesive shear strength

necessary for reactivation sensitively depend on the orientation of the horizontal intermediate

principal stress. The reactivation becomes easiest when the intermediate principal stress axis

is rotated out of strike by 45°. Here it requires only one third of the minimum cohesion and

little more than half of the maximum compressive stress necessary for pure dip-slip reversal

(Mandl, 2000). This implies that the axis of maximum horizontal stress, σH, during

reactivation is unlikely to coincide with that of the earlier axis of minimum horizontal stress,

σh. Hence, most inversion episodes for example, are non-coaxial, with oblique-slip or strike-

slip kinematics being the important modes of fault reactivation. In such a setting, shortening

strains will often be partitioned between faults displaying dip-slip, oblique-slip and strike-slip

kinematics (Turner and Williams, 2003).

Page 17: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

7

The Coulomb Failure Function provides an assessment on the likelihood of mode I and II

failure where shear stress on a fault plane exceeds the effective normal stress. This function

takes into account the effect pore pressure has on the effective stresses on a fault plane. If

such a failure mode is prevalent, the in-situ stress field needs to be assessed and fault seal

reactivation risked (Castillo et al., 2000; Jones and Hillis, 2003; Reynolds et al., 2005;

Zoback, 2010; Zoback and Healy, 1984).

2.2 THE APPLICATION OF SEISMIC DATA

One of the principal indicators of fault seal potential is the pressure differential across the

fault. How then can we attempt to measure petro-physical properties such as pore fluid

pressure, rock strength and porosity? We know that when external forces are applied to a

body, balanced internal forces are set up within it. Stress is a measure of the intensity of these

balanced internal forces. The stress acting on an area of any surface within the body may be

resolved into a component of normal stress perpendicular to the surface, and a component of

shear stress in the plane of the surface. A fluid does not shear stress however, so no shear

stress is developed in a fluid. A body subjected to stress undergoes a change of shape and/or

size known as strain (Brown, 2004; Kearey et al., 2002; Yilmaz, 2008).

Up to a certain limiting value of stress, known as the yield strength of a material, the strain is

elastic and directly proportional to the applied stress. The linear relationship between stress

and strain in the elastic field is specified for any material by its various elastic moduli, each of

which expresses the ratio of a particular type of stress to the resultant strain.

It is the bulk elastic properties of a material that determine how much it will compress under

an external pressure. The ratio of the change in pressure to fractional volume compression is

referred to as the bulk elastic modulus (Bulk Modulus) and it is the bulk modulus that

influences the speed of sound such as seismic waves in a material, as well as being a factor in

the amount of elastic energy or stress that is built up in a solid. This bulk modulus can be

thought of as the state of incompressibility of a solid and it is this incompressibility that

allows the propagation of sound.

Page 18: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

8

Seismic waves can then be thought of as parcels of elastic strain energy that propagate

outwards from the seismic source. Since rocks differ in their elastic moduli and densities, they

differ in their response to seismic waves measured through differing seismic velocities and

energy propagation. Information on the compressional P (Primary) and S (shear) wave

velocities, Vp and Vs, of rock layers measured through seismic wave response can therefore

provide an indication of the lithology of the rock through which they propagate and the nature

of the pore fluids contained within it. Because shear strain will not develop either in the pore

space or in the pore filling fluid, it is only the P-wave velocity that is influenced by pore

fluids. Thus, in principle, a variation in pore fluid type as well as pore fluid pressure should be

detectable through P-wave velocity variation (Brown, 2004; Kearey et al., 2002; Yilmaz,

2008).

Because the velocity of seismic waves is highly dependent on the density of the medium they

propagate through, subtle velocity changes can also give an indication of the densities and

pressure through which they propagate. This enables P wave velocity changes to give an

indication of not only the relative pressures and densities but also the bulk modulus and

subsequently the stress states of a rock (Hatchell and Bourne, 2005). The change in reservoir

pore-fluid properties due to pressure depletion such as gas coming out of solution can also

have a large impact on the seismic velocity, and studies show subtle variation in P-wave

velocities depending on whether the pore fluids are brine (water) or gas (Dvorkin, 2008;

Hatchell and Bourne, 2005)

The bulk modulus can be calculated from shear wave velocities based on empirically derived

relationships between Vp and Vs (Castagna, 1993). However, Vs can also be derived from

porosity velocity relationships (Castagna, 1993; Castagna et al., 1985; Dvorkin, 2008). Where

known porosity trends are available, this allows further deterministic assessment of the elastic

moduli response of a stress regime, as well providing comparison with bulk moduli

determined from the empirical relationship between Vs and Vp. Furthermore, we see that since

the pore fluid type should not affect the shear wave response, this would also allow a further

assessment on the impact on Vp by the pore fluid, independent of any anisotropic effect on the

Vs. The inter-relationship between Vp and capillary pressure becomes apparent when we

consider the empirical equation relation of porosity to pore throat radius (Pittman, 1992;

Torabi et al., 2013). Torabi et al. (2013) also show a porosity permeability power law

relationship that relates a lower porosity and permeability to fault core in relation to host rock.

Page 19: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

9

However, apart from the exponent, the empirical relationship is the same, such that the

hydrocarbon column height can then be related to pore throat radius via;

𝒉𝒄 =𝟐𝜸

𝑹𝒈(𝝆𝒘−𝝆𝒄) ... Eqn. 3

and where;

𝒉𝒄 =∆𝑷

𝒈(𝝆𝒘−𝝆𝒄) => ∆𝑷 =

𝟐𝜸

𝑹= 𝑷𝑪 ...Eqn. 4 (Hasegawa et al., 2005; Shipton et al.,

2005)

where Pc is the capillary pressure, ΔP the pore fluid pressure differential across the fault, g

gravitational constant, R the pore throat radius, γ fluid viscosity, ρ the density of w (water)

and c (hydrocarbon). Pore throat radius can be estimated from porosity through Pittman's

(1992) empirical relationship. Thus, through equations 3 and 4, an empirical determination of

capillary pressure for a given lithology can be assessed (Hasegawa et al., 2005; Pittman, 1992;

Shipton et al., 2005; Torabi et al., 2013; Watts, 1987).

Stacking velocities in a seismic survey is related to the normal-moveout (NMO) velocity

which in turn, is related to the root-mean-squared velocity from which the average and

interval velocities are derived. Interval velocity is the average velocity in an interval between

two reflectors. Pore shape, pore pressure, pore fluid saturation, confining pressure and

temperature all influence the interval velocity within rock unit (Yilmaz, 2008).

One of the most prominent factors influencing velocity in a rock of given lithology and

porosity is confining pressure. This type of pressure arises from the overburden and increases

with depth. It is generally true that velocity increases with depth, however, because of factors

such as pore pressure, there may be inversion in the velocity within a layer (Yilmaz, 2008).

There are a number of empirically derived equations relating the effect of overburden stress

and pore pressure to the Vp. Amongst the more robust relationships are the relationship of

pore pressure to vertical stress, first proposed by Terzaghi (1923) in 1923;

𝝈𝒗 = 𝝈𝒗 − 𝜶𝑷𝒑 ... Eqn. 5

Page 20: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

10

and

𝝈𝒗 = 𝝆𝒘𝒈𝒛𝒘 + 𝝆 𝒛 𝒈𝒅𝒛𝒛

𝒔𝒃 ...Eqn. 6 (Gardner et al., 1974; King et al., 2010;

Tingay et al., 2003; Castagna et al., 1985)

Where σv is the vertical stress field, z is the depth or amount of overburden, g gravitational

constant and ρ the density of w; water, for the water column, or a density as a function of

depth below mudline. Density of the lithology is the dominant petrophysical property

affecting the velocity and Gardner et al's., (1974) equation 𝜌 = 1.741𝑉𝑝0.25 is typically used.

However, Castagna's (1993) separate empirical relations for sandstone, limestone and shale

(𝜌 = 𝑎𝑉𝑝2 + 𝑏𝑉𝑝 + 𝑐) show a closer relation to typical trends. The coefficients provided by

Castagna (1993) provide good results, notwithstanding the fact that if sonic log data is

available, these should always be corrected (Castagna et al., 1985; Gardner et al., 1974;

Tingay et al., 2003).

2.3 GEOLOGICAL STRUCTURE OF THE STUDY AREA - DAMPIER SUB-BASIN AND

RANKIN TREND, NORTH WEST SHELF

The Dampier Sub-basin forms part of the North Carnarvon Basin (NCB) which itself

constitutes one of the four main basins of the North West Shelf (NWS) of Australia. The

NWS is a collection of marginal and offshore rift basins along the northern Australian margin

(Fig. 2). It is a marginal rift with pre-rift Permo-Triassic intracratonic sediments overlain by

Jurassic to Cenozoic syn- and post-rift successions deposited in response to rifting and

seafloor spreading throughout the Paleozoic and Mesozoic (Keep et al., 2000; Keep et al.,

2007; Longley et al., 2002). The NWS has pervasive NNE-SSW trending faults, some of

which show reactivation (Etheridge et al., 1991). The primary faults that influence the

structural trends are inherited from the final stages of rifting and breakup of Gondwana

throughout the Jurassic and early Cretaceous (Keep et al., 2007).

Page 21: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

11

Figure 2. Regional geology of the North West Shelf showing the four main constituent basins (Bonaparte, Browse, Canning and Carnarvon). Modified from (Keep and Moss, 2000).

Tectonic collision during the Neogene modified the pre-existing rift architecture, causing

localised fault reactivation (Longley et al., 2002). The change in tectonic regime from one of

overall extension to one of overall shortening, manifested as pulses of deformation during the

Miocene and Recent (Harrowfield et al., 2003). Bulk strain, estimated at less than 1%

shortening, was strongly partitioned on to a small number of favourably orientated structures

(Harrowfield et al., 2003; Keep et al., 2007). Areas of discrete reactivation/inversion occur in

some parts of the Bonaparte, Browse and Carnarvon basins (Keep et al., 2007). These plate

boundary forces continue to have an impact on the present day stress field and neotectonic

intraplate deformation of Australia (Fig. 3) (Hillis et al., 2008; Longley et al., 2002; Müller et

al., 2012).

Page 22: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

12

Figure 3 Australian stress map showing the orientations (small black lines) of modelled contemporary (6-0 Ma) maximum horizontal surface stress with small white lines showing maximum horizontal stress measured from bore hole breakouts (colour bar is in MPa). In the Carnarvon basin (Inset), the stress is shown to be in a NW-SE orientation (Müller et al., 2012).

Differences in neotectonic activity along the NWS reflect rift-related compartmentalisation

which resulted in the division of the major basins. The compartmentalisation takes place

along NW-trending features that have been shown to be Proteozoic fracture systems. The

basement involvement in this compartmentalisation has resulted in the compartments

responding to regional stress fields in different ways (O'Brien et al., 1997). As a result, there

is a marked change in structural orientation on the NWS in the Carnarvon Basin from a

northeast orientation in the north (Dampier Sub-basin) to a north-northeast orientation in the

south (Exmouth Sub-Basin) (Baillie and Jacobson, 1995). Muller et al. (2012) modelled the

σH (horizontal stress) regime which changes considerably in the late Miocene with the onset

of collision at the Papua New Guinea margin. Late Miocene σH orientations rotate in a

Page 23: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

13

clockwise direction from east–west in the Carnarvon Basin to northeast in the Timor Sea,

correlating with structural observations.

The NCB which forms the westernmost basin of the NWS has undergone multiple

deformation episodes including Late Permian extension (Cathro and Karner, 2006), Late

Triassic inversion and further localised extension in the Early Jurassic (King et al., 2010) (see

Appendix A for tectonostratigraphic column of the NCB).

The two final phases of extension in the Mesozoic resulted in continental breakup and

formation of oceanic crust in the Argo (Callovian) and Gascoyne-Cuvier abyssal plains

(Cathro and Karner, 2006). Minor inversion at this time is interpreted to be the result of a

major plate re-organisation as India moved northwest and break up commenced between

Australia and Antarctica (Cathro and Karner, 2006; King et al., 2010).

Extensive faulting in the Jurassic produced the five main structural elements of the Dampier

Sub-basin, including the Rankin Platform, the Madeline-Dampier Trend, the Lewis Trough,

the Legendre-Rosemary Trend and the Enderby fault zone (Fig. 4). The Rankin Platform is a

relatively high structural platform of Triassic rocks between the Rankin Fault and the shelf

edge.

Since the Neogene, collision of the Australian Plate with the Java-Banda arc system to the

north has dominated tectonism along the NWS (Keep and Harrowfield, 2008; Keep et al.,

2007). As a result, present day stress fields (Fig. 3), determined from modelling, borehole

breakouts and seismicity, indicate that maximum horizontal stress trends NE-SW across the

Bonaparte, Browse and Canning basins, reflect the dominant control by plate collision in the

PNG region (Hillis and Reynolds, 2003; Hillis et al., 2008; Hillis and Williams, 1993a; Hillis

and Williams, 1993b; Müller et al., 2012). However, in the Carnarvon Basin the modelled

stress field changes orientation, with maximum horizontal stress oriented ESE, suggesting

dominant control by far field stress generated from the plate collision in the Himalayas (Hillis

et al., 2008; Revets et al., 2009).

Page 24: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

14

Figure 4 Dampier Sub-basin and Rankin Platform (modified from (Baillie and Jacobson, 1995)).

Cretaceous inversion concentrated along major boundary faults of the northeast–southwest

orientated Rankin, Madeleine, and Rosemary–Legendre trends, with the locus of Miocene

inversion located adjacent to the northwest limits of the Rosemary–Legendre Trend (Fig. 5).

Inversion has critically controlled the major structures for trapping hydrocarbons within the

Dampier Sub-basin (Cathro and Karner, 2006).

The Dampier Sub-basin can be divided broadly into two structures, flanking either side of the

main Oxfordian depocenter. Fields along the western flank include the large rift-related horst

block traps such as the Goodwyn, North Rankin, and Echo-Yodel wet gas fields, the single

unique Perseus lowside saddle trap and numerous other smaller horst and tilted fault block

traps such as the Searipple field beneath Perseus. Subcrop of different Early Jurassic and Late

Triassic sands within many traps at the level of the base regional seal produce complex fields

containing oil and dry gas and wet gas phase combinations (Longley et al., 2002).

Rosemary trend

Madeleine trend

Page 25: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

15

The accumulation of significant volumes of hydrocarbons at multiple stratigraphic levels

along both the flanks and within the main basinal area demonstrates the laterally drained high-

impedance nature of the petroleum system within the central Dampier Sub-basin (Longley et

al., 2002). In the Dampier Sub-basin and Rankin platform in particular, the complex charge

history is reflected in at least 5 distinct fluid families, ranging from oil accumulations in the

Jurassic to dry gas accumulations suspected to have undergone a component of vertical

migration from the Triassic (Longley et al., 2002).

Figure 5 schematic cross-section of the West Dampier Sub-basin (Bennett, 2005) (see Appendix B for typical stratigraphic column of the Dampier Sub-basin).

So we see the structural history of the Rankin platform, along with the numerous well-

defined, yet complex hydrocarbon fields, provide the ideal setting to better understand the

implications that the regional stress field and multiple episodes of fault reactivation have had

on sealing faults. The deep seated faults penetrating the Triassic are well suited as they have

typically been formed during the early stages of the Gondwana breakup and have undergone

multiple phases of reactivation. This is particularly relevant as industry now more actively

pursues the Triassic Play of the NWS.

Page 26: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

16

3. AIMS (AND OBJECTIVES):

This study assesses parameters that can be used to define risk factors on fault seal when well

data is scarce. Fault seal risk analysis is not only important in the assessment of petroleum

reservoirs, but also in the assessment on the impact of any type of fluid flow within the

ground. The cross-discipline implications none more apparent than in the assessment of fluid

flow in ground water. Ground water management continues to of immense importance to the

broader economy and livelihood of those that depend on it. The developing science of CO2

sequestration is yet another discipline where fault seal capacity can have significant

implications. The assessment of CO2 reservoir storage potential, relies on the accurate

assessment of the impact that pre-existing faults and future potential seismicity may have on

its integrity.

The principle objectives focus on:

Identifying the degree to which fault architecture and stress regime can contribute to

fault seal capacity assessment.

Endeavouring to link fault seal capacity to formation pressure and seismic velocity.

Evaluation of capacity to make passive evaluation (no wells) of fault seal risk factor.

Fundamental to these objectives is the assertion (hypothesis) that the risk factor on fault seal

can be better defined using seismic velocities alone, to a practical level, where well log data is

unavailable. This study aims to test this hypothesis.

Page 27: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

17

4. METHODOLOGY:

The study involved the analysis of the Demeter 3D survey. Acquired in 2004 by Woodside

Pty Ltd on behalf of the North West Shelf Joint Venture, it is a SEG normal polarity pre-stack

time migrated volume. The survey covers approximately 3590 square km over the Echo-

Yodel, Rankin, Dockrell/Keast, Goodwyn and Perseus fields, to name a few, which sit on the

northern end of the Rankin Trend (Fig. 6). The survey has an inline bin spacing of 6.25m and

crossline bin spacing of 25m with a fold of coverage of 133. The survey was reportedly

optimised for maximum bandwidth to improve imaging of shallow targets (Bennett, 2005).

Figure 6 Map showing the blocks comprising the NWSJV and outline (red line) of the Demeter 3D survey area, on the western edge of the Dampier Sub-basin, NCB (modified from (Bennett, 2005).

A small subset of the area was selected, covering the Rankin and Dockrell/Keast fields and

associated fault blocks (Fig. 7). The subset area (summarised in Table 1) was selected as it

covers some well known fault blocks which could be used for assessment. Furthermore, the

associated stacking velocity file had to be small enough in size to be loaded on the software in

use, which has a 4GB restriction. Seismic analysis was conducted using the Landmark

software suite V5000.0.2 and final maps were generated using the Petrosys mapping software

Page 28: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

18

suite. The stacking velocities were High Density Velocity Analysis (HDVA) ‘unsmoothed’

pre stack time migrated CMP stacks of a 12.5 m x 25 m grid. HDVA stacks were generated to

create a velocity field with higher spatial resolution for final NMO correction, rather than the

500m x 500m sparse array that may typically be generated. The stacking velocities are

sampled on every shotpoint at a spacing of 12.5 x 25 m with a vertical sampling of 32ms.

Table 1 Survey Subset details

Survey Inlines Crosslines Depth Range (ms) Bin Size Area (km2)

Demeter 3D -

subset 1800 1400 0 - 6000 6.25 x 25 394

HDVA Vels 900 1400 32..sampling

interval 12.5 x 25 394

Checkshot survey data was available for 32 wells in the Rankin Trend and Exmouth Plateau.

Exmouth Plateau wells were analysed for comparative purposes in a tectonic area with a

'similar' history. The checkshot data, sometimes referred to as the 'poor man’s’ sonic log,

provide accurate time depth conversion, as well as interval velocities that are based on Dix's

Equation (Dix, 1955). This allowed calculations to be compared to well-based data at

common depths. A quality check of the data was also performed, comparing HDVA,

checkshot and sonic velocities. Sonic logs have a vertical resolution greatly exceeding that of

the seismic velocities, thus, these were not used in calculation checks, rather the checkshots

represent a far more realistic approach to qualifying the validity of using the various empirical

relationships that may be employed on the stacking velocities.

Page 29: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

19

Figure 7 Base Oxfordian regional unconformity that intersects the Horst Graben structure of the Rankin Trend. The escarpment fault which bounds the eastern edge of the Rankin Trend is also indicated. The study area is bounded in the green box (modified from (Bennett, 2005)).

The empirical relationships used rely on established properties in standard rock physics

models. In basin analysis, the overburden stress Sv can be calculated as a function of thickness

Z and density ρ of the overlying rocks multiplied by gravity g.

𝑺𝑽 = 𝒁𝝆𝒈 ... Eqn. 7

The bulk density was calculated using the empirical relationship of (Castagna, 1993). This

polynomial fit allows density to be determined in the absence of sonic log, core sampling, or

locally calibrated trends:

𝝆 = 𝒂𝑽𝒑𝟐 + 𝒃𝑽 + 𝒄; ...Eqn. 8

Here, a = - 0.0261, b = 0.3730 and c = 1.4580 . These constants match the fit for shales and

feldspathic sandstones, and best approximate the marls, claystones and clay and shale inter-

Rankin High

Kendrew

Trough

Rankin Trend

10 km

Rankin Escarpment

Page 30: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

20

bedded sandstones that dominate the lithology in the study area. The appropriateness of

Catagna's (1993) relationship was compared to (Gardner et al., 1974) relationship;

𝝆 = 𝟎.𝟐𝟑 ∗ 𝟑. 𝟐𝟖𝟎𝟖𝑽𝒑𝟎.𝟐𝟓 ...Eqn. 9

as well as a porosity based empirical relationship (Swarbrick and Osborne, 1998) relating

porosity to density;

𝝆 = 𝝆𝒎𝒂 𝟏 − ∅ + 𝝆𝒇𝒍∅ ...Eqn. 10

where ma and fl refer to the matrix and fluid density respectively. Equation 10, based on

(Wyllie et al., 1956) relationship, had the advantage of comparing density trends from two

different sources, thus comparing not just the closeness of 'fits'. Here, porosity is related to

depth via a porosity depth trend for the NCB (see Appendix C for details). The resulting

density allows Terzaghi's (Terzaghi, 1923) equation to calculate overburden stress;

𝝈 = 𝑺𝑽 − 𝑷𝒑; ...Eqn. 11 where Pp is the pore pressure.

Pore pressure can be measured in the well via a Repeat Formation Tester (RFT) tool and RFT

pressure data was collected from publically available Well Completion Reports (WCR) for 28

Wells along the Rankin Trend. This allows the effective stress to be calculated for depths

corresponding to the RFT measurement points. However, this also allows a QC check to be

performed on an empirical relation between seismic velocity and pore pressure. The common

pore pressure relationship given by Eaton (1969) relates stress measured to pore pressure from

the ratio of measured seismic velocities to the normal velocity trend for the basin. Although, it

is noted that a ‘normal’ trend can be difficult to confirm in many instances and possibly

erroneous where for example, as is the case for the Rankin trend, overpressure is so consistent

that the overpressure trend is 'normal'.

There are a variety of additional empirical relationships that reportedly provide good results

where good well calibration is available (Sayers, 2006; Soleymani and Riahi, 2012; Zhang,

2011). This study uses Bower's (1995) method, which required less calibration compared to

some others and was therefore considered more readily applied to the frontier basin. Bower's

(1995) modification relates velocity to pore pressure based on a virgin pressure curve (see

Appendix D) taken from a reference well, Rankin 1 in this case, and subtracts this from the

lithostatic pressure gradient;

Page 31: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

21

𝑷 = 𝟑.𝟗𝟓𝟖 ∗ 𝒛 − 𝑽𝒊𝒏𝒕−𝟏𝟐𝟓𝟎

𝟏𝟔.𝟓 𝟏/𝟎.𝟔

...Eqn. 12

where z is depth in meters, Vint the interval velocity in m/s.

As the stacking velocities only contained the P-wave velocities. Shear wave velocities are

calculated using the empirical relationship of Castagna (1993);

𝑽𝒔 = 𝒃𝑽 + 𝒄 ...Eqn. 13

where b = 0.77 and c = -0.867. Again this is for a lithology that best fits the dominant rock

type in the formations being studied (see Appendix E for the 'goodness of fit' for this

empirical relation). This allows Poisson's ratio to be calculated through the relationship;

𝝂 = 𝑽𝒑𝟐−𝟐𝑽𝒔

𝟐

𝟐(𝑽𝒑𝟐−𝑽𝒔

𝟐) ...Eqn. 14

And Poisson’s ratio can be used to equate the horizontal stress Shmin extent via;

𝑺𝑯𝒎𝒊𝒏 = 𝑺𝒗

𝟏 𝝂 −𝟏 ; ...Eqn. 15

A range of different empirical relationships to calculate Shmin are available (Zoback, 2010).

Those employing an effective stress ratio consistent with Byerlee’s (Byerlee, 1978) range of

values for constant of internal friction are compared (Eaton, 1969; Hubbert and Willis, 1957;

Zoback and Healy, 1984). Figure 8 summarises the work flow utilising the relationships

described above.

This allows a range or properties to be assessed; importantly the Coulomb Failure Function

(CFF):

𝑪𝑭𝑭 = 𝝈𝒔 − 𝝁𝒔 𝝈𝒏 − 𝑷𝒑 ...Eqn. 16

where, σn is the normal stress, σs the shear stress and Pp pore pressure. Appendix F describes

the three dimensional calculation of the normal and shear stress from the principal stress

vectors.

The CFF measures the difference between shear stress and effective normal stress. The effect

of increased pore pressure is to decrease the effect of overburden stress such that shear stress

is greater and Mode I and II failure is of higher risk.

Page 32: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

22

If we know the normal and shear stress, we can calculate the angle of internal friction, or

angle at which the ratio of shear and normal stress puts the rock in a critical state of stress

through

𝝈𝟑

𝝈𝟏= 𝛍𝐜 = 𝐭𝐚𝐧 𝜽 ...Eqn. 17

μson the other hand refers to the coefficient of sliding friction, equal to;

𝑺𝑺

𝑺𝒏= 𝛍𝐬 ...Eqn. 18

This is essentially Amontons’ law (Amontons, 1699), who had expanded on experiments on

friction that had first been conducted by Leonardo da Vinci (Gao et al, 2004). The coefficient

of sliding friction describes the ratio of shear to normal stress in a fault that will initiate slip or

movement. Byerlee (1978) found that for a range of rock types, the ratio of shear stress to

effective normal stress required to initiate slip was between 0.6 and 1, a phenomenon known

as Byerlee’s law. Further studies (Jaeger et al., 2007), also showed that the ratio, or μs was

remarkably consistent and approximate to 0.6. Some mudstones and clays have been shown to

have a μs of less than 0.6 and as low as 0.2 at low effective pressures (Lockner et al., 2011;

Moore and Lockner, 2004; Moore et al., 2009; Zoback, 2010). Low μs in the fault gouge of

some rock types is attributed to a high concentration of certain phyllosilicates, of which the

more common minerals have been shown to range between 0.2-0.5 (Moore and Lockner,

2004). However, aside from these minerals and the fault gouge that is comprised of them,

most rocks typically adhere to Byerlee's law thus, where μs is applied in this study, it is

assumed to be 0.6. This carries forward into the various empirical relations for Shmin

employed, where an Effective Stress Ratio (ESR) of 0.32 is equivalent to a value of μsof 0.6.

Page 33: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

23

Figure 8 Workflow diagram summarising the process of assessing the validity of select empirical relationships used to relate velocity to fault seal properties.

Page 34: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

24

5. RESULTS

5.1 STRUCTURAL CHARACTERISTICS OF THE RANKIN DOCKRELL KEAST BLOCKS -

RANKIN TREND

Spudded in July 1971, Rankin 1 (Fig. 9) tested a structural culmination on the Rankin Trend

structural high. The Rankin 1 well drilled through Pleistocene to Upper Triassic formations

with Cretaceous rocks lying unconformably on the Triassic with the entire Jurassic and much

of the Neocomian absent. The Triassic sandstones contained significant hydrocarbon

accumulations with 63m of gas/condensate pay underlain by a 3.4m oil column and an

oil/water contact at approximately 2955m, 274m higher than at North Rankin.

Figure 9 The study area with main oil and gas fields and well locations. Inset: tectonic map highlighting the Sub-basins and Rankin Trend in the NCB. The blue box over the inset relates to the area covered by the larger picture (inset: (Geoscience_Australia, 2010)).

Rankin-Dockrell/Keast Wells/Fields

Io/Jansz Field

Gorgon Field

Barrow Island

Exmouth Plateau

Page 35: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

25

The Dockrell Field was discovered in 1973 with the Dockrell 1 well intersecting a stacked

reservoir of three hydrocarbon accumulations of gas and oil, with approximately 88m of oil

and gas in the top Brigadier reservoir, and 15m of gas/condensate in the lower two Mungaroo

Triassic intersections.

The Keast Field was discovered in 1996 and has two penetrations. Keast-1, drilled in 1996,

encountered a stacked reservoir with five hydrocarbon accumulations with the fluid

intersected being gas. Keast-2, drilled in 1997, was a dry hole. Whilst both Keast and

Dockrell penetrated the Upper Triassic, core data is limited to the Brigadier formation.

Together, the three wells, which are not under production, provide pressure information

across two faults separating the Rankin High, from the Keast Graben and Dockrell high (Fig.

10). The fault blocks of the Rankin High, Keast Graben and Dockrell High are comprised of

extensional graben, half graben blocks separated by moderately dipping (40-50°) planar

normal faults. These normal faults show differential offset in the strata indicative of

continuing reactivation throughout the Jurassic age (Figs. 10-12). The main southern

bounding Rankin Escarpment fault has been reactivated well into the middle Neogene with

hard and soft linkage between the Triassic and Cretaceous/Cenozoic overburden, alternating

along the NE-SW strike of the fault. The faults show significant throw in the Triassic strata

with the Rankin Keast fault and Keast Dockrell faults showing more than 300 ms TWT

(approximately more than 400 m). Pre-kinematic strata fill the Keast Graben within the

Triassic and along the Rankin Escarpment (Figs. 12-13).

Figures 10 and 11 both show two-way time slice through 2920 ms in the Triassic. Figure 11

displays amplitude variation across the time slice with red (negative) amplitudes principally

reflecting a shale-dominant lithology with a resulting high impedance contrast. The amplitude

time slice shows horizontal separation in the NW-SE faults, which, when considered with the

relative dip of reflectors suggests that a dextral strike-slip or oblique slip movement has

occurred.

Most of the Jurassic is missing with the reported Oxfordian sub-aerial exposure resulting in

the erosion of these sequences. The upper part of the Oxfordian is dominated by claystones

which form a seal over the underlying lowstand sands. This in turn is overlain by regional

transgressive and highstand marine claystones of the Early Cretaceous period.

Page 36: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

26

Figure 10 TWT structure map of the base Norian Late Triassic surface showing the main structural elements in the study area. Triassic fault blocks have undergone extensive faulting and reactivation with the southern bounding Rankin escarpment fault representing the southern boundary of the Rankin Trend high. The Rankin, Keast and Dockrell wells have all been drilled in fault bounded structural highs with recorded hydrocarbon accumulations.

These Oxfordian and Neocomian claystones dominate the seismic cross sections with high

impendence contrast. Only extensive and continual reactivation of the Rankin Escarpment and

Goodwyn fault block offset these regional claystone seals. However, whilst Neogene

reactivation is hardlinked to the escarpment fault, many Triassic fault blocks show some

minor evidence of reactivation, with soft-linking through the Jurassic and Cretaceous

Rankin-Echo/Yodel Block

Keast Graben

Dockrell Block

Rankin

Escarpment/Fault

Rankin/Keast N-S Fault

Keast/Dockrell N-S Fault

ms TWT

E-W faults within

the Rankin Block

Page 37: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

27

claystones (Figs. 12-13). The claystone horizons themselves are dominated by classic

polygonal faulting that is typically associated with compaction and dewatering, but without

clear evidence of hardlinks through the more ductile shale dominant lithology (Fig. 13).

The semblance attribute time slice at 2910 ms TWT in figure 14 shows faulting within the

main fault blocks. The Keast Graben and Dockrell high have small faults cross-cutting the

blocks at an angle of approximately 60o to the bounding faults and with small vertical offsets,

typically less than 10 ms TWT. These cross-cutting faults terminate at the bounding faults.

The E-W faults within the Rankin/Echo-Yodel fault block are somewhat different, with very

little vertical offset and a E-W striking trend. The E-W faults also indicate horizontal

separation (Fig. 15). These E-W faults comprise a small horst and graben set cross-cutting the

Rankin/Echo-Yodel fault block and offsetting some N-S faults at an angle of approximately

60°.

Figure 11 2910 ms TWT amplitude time slice of the study area. The principal faults in the study can be seen, along with horizontal separation in the east-west faults cross-cutting the Rankin block.

Rankin escarpment

Keast/Dockrell fault

E-W faults within Rankin Block

Rankin/Keast fault

Page 38: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

28

Figure 12 a) NE-SW cross section (inset shows location (inset: red lines show correct spatial location of inline and crossline, but are not to scale in length) showing the Rankin High and well bounded by the escarpment fault to the SW and Keast Graben to the NE. Both faults are reactivated normal faults with fault bend folds in the upper Triassic and reactivation up into the middle Neogene horizons b) showing principal stratigraphic boundaries (TWT, v/h x 2 and x 4 respectively).

Fault bend fold

Reactivation of Triassic

Rankin fault into Neogene

overburden

Rankin high

Keast Graben

Rankin escarpment

fault

Eocene/Paleocene SB

KT boundary

Upper Norian

Middle Norian

400m

2500

m

2000

m

1500

500

-500

Am

plit

ud

e

400m

100 msTWT

a)

b)

SE SW SE NE

SW

Page 39: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

29

Figure 13 Semblance images show time slices in the Eocene, Middle Cretaceous and Base Cretaceous respectively. They highlight the reactivation of the Rankin Escarpment fault throughout, however, no reactivation of the underlying horst-graben architecture is evident within the geomorphology above the regional unconformity which is controlled by the depositional setting of the area.

Middle Cretaceous: Rankin Escarpment still

apparent with the geomorphology revealing little

about underlying structure in this calcilutite and

carbonate-dominated depositional sequence.

Eocene: Rankin Escarpment still present. Shore

line depositional features dominate with no

further evidence of reactivation of underlying

structure.

Shoreline channels in

siliciclastic-dominated

depositional sequence

Rankin Escarpment shows Neogene

reactivation

Goodwyn/Dockrell fault

Rankin Escarpment fault

Base Cretaceous: Classic polygonal faulting in the

shale sequence with the Rankin escarpment

evident, but underlying architecture hidden under

the unconformity.

Page 40: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

30

Figure 14 Semblance attribute of the 2910 ms TWT slice (Late Triassic Carnian) through the study area highlighting the fault array. The main Rankin Escarpment fault bounds the Rankin, Keast and Dockrell blocks to the south and the smaller East -West faults cross-cutting the Rankin Block can be seen.

Rankin escarpment

E-W faults restricted

to the Triassic -

vertical offset < 5ms

TWT

Keast Graben

Complex faulting -

process zone

crossover

Rhombohedral geometries

cross-cut the horst-graben

architecture throughout

the area.

3 km

Faults cross-cut the

graben and horst at

approximately 60o to

the bounding faults

Page 41: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

31

Figure 15 Cross section through E-W striking faults that manifest within the upper Triassic of the Rankin block. These faults show small vertical offset (approx. 10 ms TWT) but 100s of metres of horizontal separation (v/h x8).

The Keast/Graben and Rankin and Dockrell Highs are underlain by a large listric fault that

terminates at the escarpment and strikes along the uplifted basement structure that constitutes

the greater Rankin Trend (Fig. 16). Figure 16 shows that any tilting of the horst graben fault

blocks will not occur without influence from the underlying Rankin Trend structure.

Figure 17 also shows the Keast Graben displaying subtle strike slip features indicative of fault

reactivation with an oblique slip component, with strata peaking across offset (pseudo teepee).

Horizons with varying angle of dip and some reverse offset are also possible indications of

oblique slip. The regional base Cretaceous unconformity (Figs. 19 and 20) shows that whilst

the horst-graben bounding faults do not offset the shale, the geomorphology is still controlled

by the oblique slip reactivated horst-graben architecture. This architecture is also shown in

figure 18, which again exhibits oblique slip features on what would otherwise be considered a

classic extensional horst graben array.

2500

m

2250

m

2000

2 km

E-W faults within the Rankin block

NE SW

Page 42: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

32

Figure 16 Inline cross section of the Keast Graben highlighting listric faulting underlying the horst graben architecture (v/h x 2).

Listric fault underlying horst graben architecture

SE NW

Page 43: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

33

Figure 17 The horst graben architecture can be seen with transpressive/transtensive reactivation generating teepee and inversion rollover along the main horst-graben bounding faults. Varying angle of dip also indicate block rotation (oblique slip) rather than a simple strike-slip sideways movement (v/h x8).

Rankin Block Keast Graben

2500

2250

2000

3 km

500

Am

plit

ud

e

Rankin Block Keast Graben Dockrell Block

teepee transtensive / transpressive fault drag Rankin Escarpment

2500

2250

2000

500

-500

Am

plit

ud

e

NE SW

Page 44: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

34

Figure 18 An inline cross section across the Keast Graben showing features indicative of fault reactivation with an oblique slip component, with strata peaking across offset (pseudo teepee). Horizons with varying angle of dip and some reverse offset suggest that horizontal movement may be associated with the episodes of fault reactivation that have occurred (v/h x8).

Figure 19 Map of the regional Cretaceous unconformity: a deep marine shale deposited on top of the erosional Jurassic unconformity. Whilst only the main reactivated escarpment faults offset the horizon, the horizontal movement and rotation of the Triassic fault blocks are reflected in the geomorphology of this surface.

4 km

Fault drag across fault with normal offset.

Varying dip in strata, plus reverse polarity in offset and teepee-like conversion suggest inversion with an oblique slip component has reactivated the fault.

Main escarpment fault maintains net normal displacement - any oblique slip reactivation is transtensional

3000

2500

2000

SE NW SE

Page 45: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

35

Figure 20 Dockrell a) shows the Dockrell horst fault block along strike NE-SW (inset: red lines show correct position of inline and crossline, not to scale) which has NE dipping strata from the northerly tilt of the fault block. b) NW-SE cross-section on the edge of the escarpment again shows the reactivation into the Neogene (v/hx2).

Fault bend fold

Reactivation of Triassic Rankin fault into Cretaceous overburden

Dockrell high

Rankin escarpment fault

Dockrell high

Keast Graben

B A

3000

2500

2000

3000

2500

2000

500m 500m

500

Am

plit

ud

e

Rankin escarpment fault

Dockrell 1 Dockrell 1

B ́A´

B ́

A

B

Page 46: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

36

The implications of this fault reactivation on lithology and subsequently, pressure

communication across faults, may be further assessed from consideration of the velocity

profile across this fault array. Figures 21 to 25 show the stacking velocities as colour

overlaying the amplitude cross-sections for the fault blocks being studied. The cross sections

through the Rankin closure show significant velocity variation though the Triassic fault

blocks. In the Upper Triassic, velocity differential can be seen across the Rankin escarpment

fault, and interestingly within the Rankin High itself, where a small fault appears to be

delineating the southern and northern parts of the Rankin Closure (Fig. 21). Beyond 3000 -

3200 ms TWT, the velocity differential is too poor to make any significant observations

regarding relevant variation.

The Keast Graben and Dockrell High show distinct velocity shifts from the Cretaceous to

Triassic lithology (Fig. 22-23). Low velocities persist at greater depths around the bounding

faults. The velocity also shows a significant increase at the base of the Norian. In all the

velocity overlays, a low velocity zone can be seen dominating the Cretaceous formation. This

is a region consisting of a mixture of marls, claystones and calcilutite. The velocity increases

again over the claystone shale layers that make up the regional unconformity above the

Triassic. Lateral variation in velocity is minimal above the regional claystone seals and

relatively high below the seal within the Triassic fault blocks. Low velocities focus around

bounding faults and at the 'up dip' section of dipping strata (Figs. 22- 25). A distinct

differential in relative velocity across the faults is not evident. However, as figure 25

illuminates, the velocity gradient is highly sensitive to formation boundaries. The shift from

one major formation to another can almost be picked from the velocity gradient shift alone.

The velocity distribution map, for the base Norian surface (Fig. 26), shows how the velocity is

principally dependent on density and therefore depth, with the deeper parts of the surface

correlating with higher velocities. The surface is within the Mungaroo formation, and shows

lateral variation in lithology that is not seen in the TWT surface.

Page 47: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

37

Figure 21 Cross section (location inset) showing velocity variation across the Triassic fault blocks of the study area. A velocity differential is observed between the minor fault transecting the Rankin structural high. This variation may suggest a pressure differential associated with fault seal. This delineates the Rankin accumulation from the south-bounding escarpment, however, little differential is evident between the Rankin block and Keast Graben (v/h x2).

3000

2500

2000

5000

3500

Vin

t m

/s

Rankin 1

Rankin escarpment fault

Keast Graben

Velocity variation noted across minor fault transecting the Rankin high in which the Rankin 1 Well has been drilled.

Little velocity variation across the

Rankin/Keast fault

SW NE

Page 48: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

38

Figure 22 Crossline through the Dockrell and Keast Graben. The velocity profile shows relatively low velocities around the main horst-graben bounding faults. The velocities also appear to delineate the main formations with high consistency (v/h x1.5).

3000

2500

2000

Dockrell 1

Dockrell high Rankin block

Keast Graben

Relatively slower

velocities appear

around the faults

5000

3500

Vin

t m

/s

NW SE

Page 49: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

39

Figure 23 Inline and crossline across the Dockrell accumulation showing significant velocity variation across the Triassic fault blocks of the study area. The Rankin escarpment fault displays a velocity differential across it, however, again there is little variation between the Keast Graben and Dockrell high. Both the inline, B, across the Dockrell Horst and the crossline A, aligned with the strike of the horst show evidence of slow velocity patches within the Carnian which may suggest fluids or gas within the Middle Triassic (v/h x4).

5000

3500

Vin

t m

/s

3000

2500

2000

3000

2500

2000

B A

Dockrell 1

Dockrell high Dockrell high

Dockrell 1

A

B

500m 500m B ́

B ́A´

Rankin escarpment fault

Rankin escarpment fault

Page 50: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

40

Figure 24 This crossline across the Rankin, Keast and Dockrell fault blocks shows evidence of inversion, principally through the structural culmination within the greater Keast, Dockrell Graben where sediment has been deposited prior to extension, extensionally listrically faulted and inverted. The varying dip within each fault block is itself a possible result of fault block rotation associated with oblique slip. Unlike the inlines, this cross line at a smaller scale does show some velocity variation across the Rankin/Keast fault, however in this instance, it is difficult to separate this from fault shadow effectsf (v/h x4).

Structural culmination over otherwise extensional architecture is highly suggestive of inversion

5000

3500

Vin

t m

/s

SW NE

Rankin escarpment fault

Rankin/Keast fault

Page 51: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

41

Figure 25 Whilst velocity differential across faults is ambiguous, the velocity profile shows remarkable consistency with the formation boundaries. ‘Formation’, as being distinct from ‘lithology’ (such as shale or sandstone), means the overall gross lithological properties that define a formation such as the Haycock Marl or Mungaroo Formation (v/hx2).

marl an

d

calcareo

us

clayston

e

Calcilu

tite

clayston

e

and

marl

sandstone& claystone

claystone

Co

arse Sst with

inte

rbed

ded

clayston

e

5000

3500

Vin

t m

/s

3000

2500

2000

600m

0 1000 2000 3000 4000 5000 6000 7000 8000

Interval Velocity (m/s)

Rankin 1 HDVA SPT Stack

SW NE

Page 52: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

42

Figure 26 a) Base Norian Horizon in TWT showing the principal fault blocks of the study, including Goodwyn 6 and the eastern edge of the Goodwyn fault block. b) showing the P-wave velocity profile of the Base Norian horizon. The velocity profile is a good indicator of lithology and density, and therefore the fault block architecture, however, lateral resolution is too poor to infer any pore pressure differential.

TWT ms

Velocity m/s

2 km

Page 53: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

43

Figure 27 shows the Depth Structure map of the Base Norian Horizon. It shows that the

general structural geomorphology is consistent with the TWT maps with no culminations or

depressions removed during depth conversion. The depth map is used in the pore pressure

calculation for the base Norian horizon.

Figure 27 Depth Structure map of the base Norian (Triassic) horizon. The structure remains consistent with the time structure maps.

Depth mTVDSS

5 km

Page 54: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

44

5.2 VELOCITY ANALYSIS

The velocity profile of the Rankin Trend is broadly consistent along its entirety (Fig. 28).

However, when compared to a similar region within the same basin; the Exmouth Plateau,

each displays a different rate of increase with depth. The exception is Gorgon 1, which is

known to sit over an anomalous 'hot spot' within the basin (see Appendix G). In addition, the

interval velocities (Fig. 29) as opposed to the instantaneous velocity, show a velocity reversal

at approximately 2000 mBML which is also different when compared to the Exmouth

Plateau, which has no reversal. The velocity reversal, present over the entire Rankin Trend, is

clearly shown in figure 30.

Figure 28 The velocity profile of 33 wells in the NCB comparing the Rankin Trend to the Exmouth Plateau, two sub-regions of the Basin that are relatively uplifted. Each has a distinct velocity trend of its own.

0.00

1000.00

2000.00

3000.00

4000.00

5000.00

6000.00

0.00 500.00 1000.00 1500.00 2000.00 2500.00

De

pth

(m)

OWT (ms)

Velocity Profile of the Rankin Trend and Exmouth Plateau Jansz 3

Nickols Bay

Jupiter

Vinck 1

Eendracht 1

Goodwyn 6

Goodwyn 9

Goodwyn 1

Zeewulf 1

Zagreus 1

Athena 1

Dockrell 1

Dockrell 2

Rankin 1

North Rankin 1

North Rankin 2

Lady Nora 1

Dixon 1

Dixon 2

West Dixon 1

Sculptor 1

Miller 1

Haycock 1

Yodel 1

Echo 1

Keast 1

Perseus 1

Dampier 1

Malmsey 1

Eastbrook 1

Gorgon 1

Mecury 1

Saturn 1

Exmouth Plateau Wells

Rankin Trend Wells

Page 55: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

45

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

5000.00

0 1000 2000 3000 4000 5000 6000 7000 8000

De

pth

(m)

Interval Velocity (m/s)Velocity Profile of the Rankin Trend

Dockrell 1

Dockrell 2

Rankin 1

North Rankin 1

North Rankin 2

West Dixon 1

Sculptor 1

Keast 1

Figure 29 Profile of the interval velocity in the Rankin trend and Exmouth Plateau for comparison. Unlike the Exmouth Plateau, the Rankin trend has a distinct change in slope at approximately 1200m and 2500m TVDSS.

Figure 30 The interval velocity profile of the wells within and adjacent to the blocks of the study area on the Rankin trend. A decrease in interval velocity with increasing depth is seen between approximately 2200m and 2700m.

Exmouth Plateau Wells

Rankin Trend Wells

Page 56: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

46

The Vp/Vs ratio is approximately 2, albeit with a slow decrease (increasing Vs) with depth.

The first 1000m below mudline has a slow decrease in the ratio in both the Rankin trend and

Exmouth plateau. The ratio remains remarkably constant beyond 1000m BML, with the

exception of significant lithology variation or overpressure/fluid contact zones.

Figure 31 The Vp/Vs ratio with depth is rather consistent, at approximately 2, and 2.3 in depths correlating with overpressure.

When comparing the stress and pore pressure differential across a fault block it is important to

ensure that the stacking velocities have been 'picked' to sufficient accuracy. Figure 32 shows a

comparison between the sonic log and checkshot velocities from four wells within the study

area. These well velocities are compared to the stacking velocity over the seismic shotpoint

that is nearest to the respective wells (within 75m of the well by horizontal distance). The

velocities show a good correlation: however, the sonic logs display vertical resolution of less

than 1 m whereas the vertical resolution of the checkshot and stacking velocities are on the

order of tens of metres. This is important to consider within the context of differentiating

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

5000.00

1.00 1.50 2.00 2.50 3.00 3.50 4.00 4.50 5.00

De

pth

(mTV

DSS

)

Vp/Vs

Vp/Vs Trends for the Rankin Trend and Exmouth Plateau Jansz 3Nickols BayJupiterVinck 1Goodwyn 6Goodwyn 9Goodwyn 1Zeewulf 1Zagreus 1Athena 1Dockrell 1Dockrell 2Rankin 1North Rankin 1North Rankin 2Lady Nora 1Dixon 1Dixon 2West Dixon 1Sculptor 1Miller 1Haycock 1Yodel 1Echo 1Keast 1Perseus 1Dampier 1Malmsey 1Eastbrook 1Gorgon 1Mercury 1Saturn 1

Exmouth Plateau Wells

Rankin Trend Wells

Exmouth water depth approx 1400m

Rankin Trend water depth approx. <200m

Page 57: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

47

individual channel sands or intra-formational seals that may themselves be less than 1m thick.

However, within the context of determining gross formation stress and pore pressure

variation, the stack and checkshot interval velocities provide a workable resolution.

Hence, we contend that the stacking and checkshot interval velocities are valid data sources

for estimating empirically related properties such as pore pressure and stress in a geological

formation. Figure 33 shows the relationship between velocity and density, as described in

Section 4: (Methodology). The densities were derived from the velocity profiles over the

Rankin Trend and Exmouth Plateau. These velocity-based densities broadly match the density

trend derived from the NCB porosity trend.

RFT/FIT pressure tests (Fig. 34-35) in the Rankin and Keast/Dockrell field suggests that these

accumulations have a pressure differential of up to 750 psi. The Rankin 1 well RFT data is old

(1973) and perhaps not as precise as newer tests, however, it suggests that there is capillary

seal of at least 750 psi across the Rankin/Keast fault. The Keast 1 and Dockrell 1 Wells

indicate a pressure differential of 280 psi, suggesting that the N-S fault separating Dockrell

and Keast has a capillary potential of at least 280psi. Most wells follow the hydrostatic

gradient with notable exceptions including Zagreus, Orthrus and Gorgon (Fig. 34). With

respect to the Rankin Trend wells, Lady Nora 1 and Rankin 1 stand out, both overpressured

relative to the hydrostatic gradient at shallower depths. Most wells show a deviation from the

hydrostatic gradient around 3500 mTVDSS, with subtle variations, for example, Goodwyn 6

is slightly deeper, around 3600 m, whereas Rankin 1, located structurally on one of the

shallowest locations within the Triassic, has elevated pore pressure from 2900m.

Page 58: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

48

Figure 32 Wells from the main fault blocks being studied show a comparison between the interval velocities extracted from the HDVA stacks being examined and the checkshot and sonic log velocities from each respective well. Good correlation supports the validity of using either velocity source to derive stress and pore pressure relationships.

0

500

1000

1500

2000

2500

0

500

1000

1500

2000

2500

3000

3500

4000

4500

0 2000 4000 6000 8000

TWT

(ms)

De

pth

(mTV

DSS

)

Interval Velocity (m/s)

Comparison of velocity data for Dockrell 1

Sonic

Checkshot

HDVA SPT Stack

0

500

1000

1500

2000

2500

0

500

1000

1500

2000

2500

3000

3500

4000

4500

0 2000 4000 6000 8000

TWT

(ms)

De

pth

(mTV

DSS

)

Velocity (m/s)

Comparision of velocity data for Rankin 1

Sonic Log

Checkshot Vint

HDVA SPT Stack

0

500

1000

1500

2000

2500

0

500

1000

1500

2000

2500

3000

3500

4000

0 2000 4000 6000 8000

TWT

(ms)

Dep

th (m

TVD

SS)

Interval Velocity (m/s)

Comparison of velocity data for Keast 1

Sonic

Checkshot

HDVA SPT Stack

0

500

1000

1500

2000

2500

3000

0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

0 2000 4000 6000 8000

TWT

(ms)

De

pth

(mTV

DSS

)

Interval velocity (m/s)

Comparison of velocity data for

Goodwyn 6

Sonic

Checkshot

HDVA SPT Stack

Page 59: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

49

Pore fluid retention leading to overpressure is restricted to the Triassic, below the regional

claystone seals. The over pressure can be seen (Figs. 34-36) with the pressure deviation from

the hydrostatic trend occurring at approximately 3600 mTVDSS, which is the typical depth of

the regional claystone seal within the study area.

The effect of this overpressure can be seen in figure 36 where the effective pressure is not a

simple linear trend with depth. Rather, the effective pressure trend is concave downwards

with further reduction in effective pressure around 3500 mTVDSS. Further evidence for

effective stress reduction being due to overpressure is assessed through the LOT data from the

NCB. Figure 37 shows this data including FIT tests that recorded maximum pressures. A

linear trend throughout the NCB is evident and the maximum pressure data points scatter with

the normal distribution around the actual LOT data. Figure 37b shows just the LOT data and

the consistency of the trend. Figure 38 shows the methods based on an empirical Poisson’s

ratio (Eqn. 15) and porosity trend underestimate the Shmin at depth, when compared to values

derived from LOT data. The methods of (Hubbert and Willis, 1957; Zoback and Healy, 1984)

and (Eaton, 1969) provide a closer fit, with the Zoback and Healy (1984) and Hubbert and

Willis (1957) relations underestimating Shmin at depth and Eaton overestimating the stress. In

these relations, Eaton’s (1969) and Zoback and Healy’s (1984) relations use an effective

stress ratio of 0.32 (coefficient of internal friction (CIF) of 0.6), whilst that of Hubbert and

Willis’ is 0.38 (CIF 0.5). Eaton’s (1969) relationship provides a poorer fit for the Rankin

Trend alone (Fig. 39), although may fit better at greater depths.

The difference in Shmin derived from the velocity and from LOT is plotted in Figure 40. The

power law relationship resulting from the underestimation of Shmin from the velocity allows a

'correction factor' to be calculated. This correction factor assumes a calibration well is present,

nonetheless, a clear trend is discernable.

Figure 43a-d show the coulomb failure function described for the principal fault blocks in the

study. Described in Section 4, these functions are graphed for a fixed orientation

representative of the main graben bounding NE-SW faults that separated Rankin from Keast

and Dockrell. A constant coefficient of sliding friction of 0.6 is assumed. The CFF suggests,

that under the current stress regime, these fault blocks should not consist of a stress regime

that would lead to fault slip for a plane representing the Rankin/Keast/Dockrell N-S faults

(Fig. 10). However, when the Shmin derived from the P-wave interval velocity is employed, an

Page 60: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

50

erroneous data set is found. When the velocity-based CFF with a correction factor applied

shows a broad consistency with the LOT-based CFF.

Pore pressures are calculated from Bowers’ (1995) relation with a Biot factor of 0.8 (see

Appendix D for details) and pore throat radius calculated from (Pittman, 1992) relation. With

empirically derived pore pressure, Eaton’s (1969) or Zoback and Healy’s (1984) equation can

be used to calculate the CFF in place of a ‘corrected’, well-dependent CFF. The calculated

pore pressure correlate well with the RFT measured pore pressures (Figs. 44-45). The pore

pressure variation across the area is broadly consistent, with some scatter coming from the use

of checkshot velocities, particularly in the case of Keast (Fig. 46). However, the calculated

pore pressures from stacks, mapped on to the base Norian surface in the case of figure 47,

suggest there is a lateral variability.

The degree of this lateral variability further is assessed by looking at pore pressure

distribution for a constant depth(time) over a time slice of 2300 ms TWT. Eliminating the

depth-dependent pressure effects that occur with a depth-varying surface, it shows that a

lateral variability in pore pressure remains. However, anomalous velocities (and hence

pressures) are observed.

The CFF was subsequently calculated using Bowers (1995) pore pressure and LOT based

Shmin and Zoback and Healy (1984) based Shmin (Fig. 49). Both methods suggest that even

within the overpressure window there is little risk of reactivation on the modelled N-S fault.

Final CFF’s used stacking velocities rather than checkshot velocities (Figs. 50-52). The three

primary fault planes representing the Rankin escarpment, Rankin/Keast fault and small East-

West Rankin Block faults are modelled. The effects of pore pressure in the overpressure

window can be seen. Whereas the Rankin/Keast fault is stable, the escarpment fault has a

positive CFF at depths equivalent to the top of the Triassic. Within the current stress regime,

the escarpment fault may still be subject to failure in places.

Page 61: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

51

Figure 33 The density relationship with depth for the Rankin Trend, NCB. Scatter in the density derived from interval velocities reflects differing lithologies and pore pressures, whereas Equation 10 (Swarbrick and Osborne, 1998) relates an empirically derived porosity with depth trend to density, hence, the smooth exponential type curve.

Figure 34 The formation pressures of selected wells from the Rankin trend and Exmouth plateau. The lithostatic gradient is based on core measured Rankin trend porosities resulting in the Exmouth plateau well, Jansz 3 exceeding it.

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

5000.00

1.5 1.7 1.9 2.1 2.3 2.5 2.7 2.9D

ep

th (m

BM

L)Density (g/mm3)

Density Relationships

Gardner 1974

Castagna 1993

Swarbick and Osborne 1998

Page 62: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

52

Figure 35 With the exception of the *wells from the Exmouth Plateau, the Rankin Trend wells all show changes in formation pressure gradient associated with the overpressure from disequilibrium compaction. Goodwyn 6 is pointed to as its numerous data points make the gradient change easier to see.

Figure 36 Effective pressures from wells in the study area. They show a concave downwards trend as overpressure at depth of 3400m and below reduces the effective stress and associated stress to depth pressure gradient.

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

4000.0

4500.0

5000.0

2000.00 3000.00 4000.00 5000.00 6000.00 7000.00 8000.00 9000.00 10000.00

De

pth

(mB

ML)

Pressure (PSI)Rankin Trend Formation Pressures

Dixon 2

Dixon 1

Rankin 1

Dockrell 1

Zagreus 1*

Orthrus 1*

Maenad 1*

Goodwyn 1

Goodwyn 6

Keast 1

Sculptor 1

Keast 2

Peublo 1

2000

2500

3000

3500

4000

4500

5000

2000 3000 4000 5000 6000 7000 8000 9000 10000

De

pth

(mB

ML)

Pressure (PSI)

Rankin Trend Effective Pressures

Dixon 2

Dixon 1

Rankin 1

Dockrell 1

Goodwyn 6

Keast 1

Keast 2

Lithostatic Pressure Gradient

Hyrdrostatic Pressure Gradient

Linear (Lithostatic Pressure Gradient)

Exmouth Plateau wells

example of pressure gradient

variation due to overpressure.

Page 63: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

53

Figure 37 XLOT and ISIP pressures from 47 wells from the NCB show a linear trend between depth and inferred Shmin at 0 to 4 km depth below sea floor. It suggests a broad consistency in regional Shmax throughout the NCB and well as a consistent vertical to horizontal stress ratio (data from the CSIRO pressure plotTM database).

a)

b)

Dep

th TV

DSS

Dep

th TV

DSS

Dep

th T

VD

SS

Dep

th T

VD

SS

Page 64: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

54

Figure 38 Comparison of the main empirical relationships used to determine Shmin from interval velocities. Bilateral constraint based relations such empirical Poisson and Porosity based trends underestimate the Shmin at depth.

Figure 39 Based on the assumed effective stress ratio of 0.32, Zoback and Healy's and Hubbert and Willis' relations provides a good fit to LOT trends, Eaton's tends to overestimate at depth.

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

5000.00

0 2000 4000 6000 8000 10000 12000

De

pth

mTV

DSS

Shmin (psi)

Comparison of Shmin empirical methods to Shmin from LOT

Shmin from empiricalPoissonShmin from PorosityrelationShmin from Hubbert &Willis (1957)Shmin from Zoback & Healy(1984)Shmin from LOT trend

Shmin Eaton (1969)

Linear (Shmin from LOTtrend)

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

5000.00

0 2000 4000 6000 8000 10000 12000 14000

De

pth

mTV

DSS

Shmin (psi)

Comparison of Shmin empirical methods accounting for bilaterial constraint on the Rankin trend

Eaton (1969)

Shmin from Hubbert & Willis(1957)Shmin from Zoback & Healy(1984)Shmin from LOT

Overburden Stress psi

Increase in uncertainty due to increase in

lithological heterogeneity (see Appendix 11) Underestimation in Shmin by

bilateral constrain based relation

Page 65: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

55

0.00

1000.00

2000.00

3000.00

4000.00

5000.00

6000.00

0.00 500.00 1000.00 1500.00 2000.00 2500.00 3000.00 3500.00 4000.00 4500.00 5000.00

dif

fere

nce

Sh

min

(p

si)

Depth mTVDSS

Difference plot between velocity-based and LOT-based Shmin

Figure 40 The ratio of Sv and Shmin is plotted for the Rankin trend wells in the study area and compares Shmin estimated from LOT data to Shmin calculated from an empirical relation. The green calculated values have an average value of 0.62, however, tend to underestimate the ratio compared to the LOT based data which has an average of 0.64. Apart from a deviation around 3000m, the ratio has a small decrease with depth due to the gradual increase in overburden stress.

Figure 41 plots the difference in Shmin where σ3 is derived from shear wave velocity using Castagna's (1993) relation and σ3 is inferred from the LOT measurements. This presents a 'correction factor' that can account for the error in the assumption of bilateral constraint. The validity of using shear wave based Shmin may be appropriate where this difference is known, e.g. from a reference well (see Appendix E for validity check).

𝑦 = 0.0002𝑥2 + 0.1355𝑥 R2=0.78

Page 66: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

56

2000.0

2500.0

3000.0

3500.0

4000.0

4500.0

5000.0

1000.00 6000.00 11000.00 16000.00 21000.00

De

pth

(mB

ML)

Pressure (PSI)

Comparison between Pore Pressure, σv and Shmax for selected Wells on the Rankin Trend

Dixon 1

Rankin 1

Dockrell 1

Goodwyn 6

Keast 1

Keast 2

Lithostatic Pressure Gradient

Hydrostatic Pressure Gradient

sHmin

Dixon 1 sHmax

Dockrell 1 sHmax

Rankin 1 sHmax

Goodwyn 6 sHmax

Echo 1 sHmax

Keast 1 sHmax

Dixon 1 σv

Rankin 1 σv

Dockrell 1 σv

Goodwyn 6 σv

Echo 1 σv

Keast 1 σv

Figure 42 Comparison between pore pressure and vertical and minimum horizontal stress components. The graph shows effective stress reduction where pore pressure increases. Furthermore, σ1 remains in the vertical direction, suggesting a normal/strike slip tectonic regime.

Page 67: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

57

``

Figure 43 Coulomb Failure Function for a fault plane that is representative of the N-S steeply dipping faults that separate the Rankin/Keast and Dockrell blocks (Schematic of the theoretical plane used can be seen in figure 56). The CFF is calculated using Pp and LOT measurements at a depth which has Norian (Triassic) Mungaroo formation geology in the Rankin Trend. a) The shear stress is measured from LOT data, are all below zero suggesting no prevalence for failure. b) The shear stresses are derived from the seismic velocities using Vp to Vs conversion for an estimation of Poisons ration. C) a correction factor is applied accounting differences between LOT-based shear stress and Vp based shear stress d) shear stress calculated assuming a fixed principal stress ratio i.e. fixed ESR of 0.32, as per Zoback and Healy (1984) or Eaton's (1969) equations. The figures show the degree to which the 'corrected' c, or calculated d, CFF broadly correlate with a, the LOT based CFF.

0.0

500.0

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

4000.0

-1400.00 -1200.00 -1000.00 -800.00 -600.00 -400.00 -200.00 0.00

De

pth

(mTV

DSS

)

CFF (psi)

Coulomb Failure Function using Shmin derived from Zoback and Healy's (1985) relation

Rankin 1

Dockrell 1

Keast 1

0.0

500.0

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

4000.0

-1600.00 -1400.00 -1200.00 -1000.00 -800.00 -600.00 -400.00 -200.00 0.00

De

pth

(mTV

DSS

)

CFF (psi)

Coulomb Failure Function using Vs with empirical correction factor applied

Rankin 1

Dockrell 1

Keast 1

0.0

500.0

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

4000.0

-600.00 -400.00 -200.00 0.00 200.00 400.00 600.00 800.00

De

pth

(mTV

DSS

)

CFF (psi)

Coulomb Failure Function using Vs from Castagna's (1985) relation

Rankin 1

Dockrell 1

Keast 1

0.0

500.0

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

4000.0

-1400.00 -1200.00 -1000.00 -800.00 -600.00 -400.00 -200.00 0.00

De

pth

(m

TV

DSS

)

CFF (psi)

Coulomb Failure Function using LOT stress

Rankin 1

Dockrell 1

Keast 1

b a

c

d c

Page 68: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

58

Figure 44 Comparison of pore pressure from checkshot interval velocity, RFT and HDVA velocity stack. The velocity picks from the stack closely match the least squares fit of the checkshot.

Figure 45 Comparison between Rankin 1 RFT measured pore pressure and calculated pore pressure using the Bowers’ (1995) relationship with velocity. RFT measurements are biased towards the target reservoir (only measured at the depth corresponding to reservoir), but show a reasonable correlation.

0

500

1000

1500

2000

2500

3000

3500

4000

0.00 1000.00 2000.00 3000.00 4000.00 5000.00 6000.00 7000.00 8000.00

TVD

SS

(ms)

Pore Pressure (psi)

Pore pressure comparison between RFT and velocity measurements on Keast 1

Velocity based pore pressure

RFT

HHDVA stack SPT

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

0.00 1000.00 2000.00 3000.00 4000.00 5000.00 6000.00 7000.00 8000.00

TVD

SS

(ms)

Pore Pressure (psi)

Pore pressure comparison between RFT and velocity measurements on Rankin 1

Velocity based pore pressure

RFT measured pore pressure

De

pth

(m

TVD

SS)

De

pth

(m

TVD

SS)

Page 69: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

59

Figure 46 The estimated pore pressure trend based on checkshot velocities from the respective wells in the list is broadly consistent across the different fault blocks within which the wells are located.

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

0.00 2000.00 4000.00 6000.00 8000.00 10000.00

De

pth

mTV

DSS

Pore Pressure (psi)

Pore Pressure from Velocity on the Rankin Trend

Goodwyn 6

Rankin 1

North Rankin 1Yodel 1

Keast 1

Sv

Page 70: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

60

Figure 47 Pore pressure variation across the base Norian horizon. Higher pore pressures are focussed around fault boundaries, as well as broadly within the Rankin High and Dockrell Highs and near the Goodwyn block. Pore pressures on the scale bar are in psi. This map does not indicate pressure differential across faults as the horst and graben are at different depths.

Higher pore pressure in the shallower section of the base

Norian horizon- localised around the Rankin, Dockrell and

Goodwyn 6 blocks

Pore pressure (psi)

Page 71: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

61

Figure 48 Observation of lateral variation in a) velocity and b) pore pressure from a time slice at 2300 ms TWT shows good consistency with values for most of the study area. However, anomalous results are evident in the lower corner (see schematic top left corner) of the study area that is south of the Rankin escarpment. This is shown to be due to poorly picked stacking velocities in this part of the study area (inset). Whilst relative variation is useful, absolute values are only as good as the velocity picks.

0

500

1000

1500

2000

2500

0

500

1000

1500

2000

2500

3000

3500

4000

4500

0 2000 4000 6000 8000

TWT

(ms)

De

pth

(mTV

DSS

)

Interval Velocity (m/s)

Comparison of velocity data for Dockrell 1

Sonic

Checkshot

HDVA SPT Stack

As Dockrell 1 velocity comparison shows,

the stacking velocities in the SE corner of

the study area poorly match calibration

below 3100 m(2000 ms TWT).

Velocities anomalously high in this area

produce anomalously low pore pressures

Incorrect velocities result in anomalously low

pore pressure in the Se corner of the study

area

Velocity m/s

Pore Pressure (psi)

5 km 5 km

5 km

Poor velocity picks in the lower

(red) corner of the study area

or everywhere south of the

Rankin escarpment.

Page 72: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

62

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

-3500.00 -3000.00 -2500.00 -2000.00 -1500.00 -1000.00 -500.00 0.00

Dep

th (

mTV

DSS

)

CFF (psi)

Coulomb Failure Function using Shmin from LOT and Pore Pressure from Bowers relation

Rankin 1

Dockrell 1

Yodel 1

Keast 1

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

-3000.00 -2500.00 -2000.00 -1500.00 -1000.00 -500.00 0.00

Dep

th (

mTV

DSS

)

CFF (psi)

Coulomb Failure Function using Shmin from Zoback & Healy (1984) and Pore Pressure from Bowers relation

Rankin 1

Dockrell 1

Yodel 1

Keast 1

a)

b)

Figure 49 a) Coulomb Failure Function using LOT-derived Shmin shows the fault plane representing the Rankin/Keast fault should be stable, albeit with relatively reduced stability between 2000 and 3000m BML due to increased pore pressure. The Zoback and Healy (1984) relation is similar, but with reduced variation due to pore pressure.

Page 73: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

63

0

1000

2000

3000

4000

5000

6000

-3000.00 -2500.00 -2000.00 -1500.00 -1000.00 -500.00 0.00 500.00 1000.00 1500.00 2000.00

De

pth

(m

TVD

SS)

CFF (psi)

CFF comparison for three fault planes representing the main fault within the Rankin Fault Block with Rankin Block Pore Pressures

Rankin BlocK: Fault plane 010/50/E

Rankin Block: Fault Plane 070/40/E

Rankin Bloock: Fault Plane 080/60/W

Figure 50 Coulomb failure function profile using stacking velocities and the empirical relations from Zoback and Healy (1985) and Bowers (1983).The stacks have reduced variability, but are broadly consistent with the checkshot CFF.

Figure 51 The CFF for three fault planes in the study area. The escarpment fault in the only one within the current stress regime that high risk of mode I or II failure. In this calculation, the intermediate stress is approximated as 0.9σ1.

0

1000

2000

3000

4000

5000

6000

-3000.00 -2500.00 -2000.00 -1500.00 -1000.00 -500.00 0.00

Dep

th (

mTV

DSS

)

CFF (psi)

CFF for fault plane representing the Rankin/Keast Fault with Rankin Block Pore Pressures

Page 74: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

64

Further analysis of the possible range of hydrocarbon column heights was made. Based on

mercury injection permeability and porosity tests on core sample at STP conditions.

Combined with NCB porosity trends for the Norian Mungaroo (Appendix C), it presents a

range of possible capillary pressures and HC heights for shale and sandstone. At depths

equivalent to typical oil/gas-water contacts, sandstone capillary pressures measure

approximately 500 psi, with HC heights up to 126m. Shale on the other hand results in

calculated capillary pressure of up to 3700 psi and HC heights of over 1000m. Mixed

lithology trend lay somewhere in between and are representative of low Vshale sands due to

the statistical bias in the porosity trends (see Appendix C).

Figure 52 Theoretical capillary pressures for sandstone, shale and sandstone dominant mixed lithologies calculated using mercury injection STP trends and Mungaroo core based porosity trends. Maximum capillary pressures increase with depth and are significantly higher for shale vs. sandstone.

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

0 500 1000 1500 2000 2500 3000 3500 4000

De

pth

(TV

DSS

)

Capillary Pressure (psi)

Capillary Pressure from Mungaroo Porosity Trends on the NCB

Sandstone

Shale

Sst dominant mixed lith

Page 75: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

65

Figure 53 Resulting permissible HC column heights based on theoretical capillary pressures calculated from Mungaroo porosity trends. This shows that even for moderately porous sandstone, significant HC heights can be sustained at depth.

0.00

500.00

1000.00

1500.00

2000.00

2500.00

3000.00

3500.00

4000.00

0 200 400 600 800 1000 1200 1400D

ep

th m

TVD

SSHC column height (m)

HC column heights from NCB porosity trends in the Norian Mungaroo

Sst

Shale

Sstdominantmixed lith

Page 76: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

66

6. DISCUSSION/ANALYSIS

Structurally, the Rankin and Dockrell/Keast fault blocks show evidence for fault reactivation

through to the Neogene (Fig. 12-18). However, whilst multiple episodes of reactivation are

evident, this reactivation is highly partitioned, with main horst-graben bounding faults

showing differential offset within the Triassic, up to the Jurassic/Cretaceous unconformity.

Within the Cretaceous and Cenozoic, reactivation is focused on NE-SW bounding faults of

the Rankin escarpment and Goodwyn/Dockrell blocks (Fig. 12- 18).

The faults show characteristic features associated with oblique slip which, along with varying

dip of strata indicates that oblique slip is associated with fault block rotation during extension.

The orientation of fault arrays and classic rhombohedral geometries within the fault blocks

also suggests an element of oblique slip associated with seismicity during the formation of the

horst-graben architecture (Fig. 14). A rose diagram of the fault orientation also has an array

consistent with oblique extension (Fig. 54).

Figure 54 The fault orientation within the Upper Triassic (Rose diagram) is consistent with a typical array that results from oblique extension. As the range of angles of obliquity shown modelled by (Withjack and Jamison, 1985). The angle of maximum extension shows inversion oblique to extension. The pink line is in the direction of maximum extension, red arrows indicate Shmax orientation: approx. 110o.

Escarpment fault orientation

Horst Graben orientation

Page 77: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

67

Whilst fault offset maintains net normal displacement, subtle inversion features are evident

around the Rankin escarpment which shows evidence of Neogene reactivation. Fault

reactivation along the horst-graben bounding faults that separate the fault blocks show signs

of late reactivation to be predominately via oblique slip. Stereonets in figure 57 using

Angelier's (Angelier, 1979) method, show the likely direction of slip for these reactivated

faults is consistent with the right lateral horizontal movement that is evident (Fig. 14, 17, 19,

20). However, this remains comparatively minor and the faults have broadly accommodated

very little strain throughout the Cenozoic. This evidence suggests little seismicity along these

faults, which are likely to be relatively stable in the present day. Some indication of the

validity of this assertion comes from passive seismic data recorded on the NWS. Figure 55

shows all earthquakes recorded at a depth of less than 10km since 1954 (when recording

began). The only earthquake measured near the study area was in proximity and likely

associated with reactivation of the Rankin escarpment fault.

Figure 55 Yellow circles are earthquakes recorded since 1954 which were located at depths of 10km or less. Magnitudes 2 or greater (non labelled <2) (data from (Geoscience_Australia, 2014)).

Local borehole breakout data from around the study area show present day Shmax orientations

to be approximately NW-SE. This is relatively consistent with the far field Shmax for the NCB,

reported to be approximately 110o (Hillis et al., 2008; Hillis and Williams, 1993b). The fault

array shows a geometry that is suggested to be consistent with far field stress dominance,

which is typical for the NCB (Dyksterhuis and Muller, 2008; Hillis et al., 2008; Keep et al.,

2002).

Rankin escarpment bounds the

southern edge of NWSJV fields

2.7

4.0

3.4

3.0

1997

1990 1991 3.7

3.2

4.8

1987

1986 3.7

5.6

1976

3.3 2002

3.1 3.2

1997 1991

3.2

1979

1979 1984

1986

Page 78: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

68

Whilst there is lateral variation in velocity within the Triassic, the variation across the faults

themselves is subtle, if not typically indistinguishable. A notable exception is within the

Rankin Block near the Rankin 1 well, across a small synthetic fault (Fig. 21). Whilst there is

little differential across the faults, this is a relative velocity reduction around the fault damage

zone. This suggests that there is some change in the petrophysical properties in the lithology

immediately adjacent to the fault surface. As that change is likely to be a combination of

density and elastic response, it may be indicative of the greater damage zone envelope of the

fault. The fact that there is little velocity variation across the faults separating the Rankin

High and the Keast Graben is interesting in so far as the Rankin/Keast fault juxtaposes fluvial

dominated Mungaroo units with dominantly deltaic Brigadier lithology. This suggests that the

two distinct formations must have similar gross petrophysical properties as well as reducing

the prospect of the velocities being able to show subtle pressure differences, which the RFT

data (Fig. 35) indicates should exist.

However, as figure 25 illuminates, the velocity gradient is highly sensitive to formation

boundaries. The shift from one major formation to another can be picked from the velocity

gradient shift alone. Particularly prominent is the velocity reversal on the Rankin Trend due to

the 'siliciclastic' formation in the Eocene (see Fig 5), which is thin to absent in the offshore

Exmouth Plateau, hence no velocity reversal manifests over it. This suggests that whilst the

particular stacking velocities used in this study may not have sufficient resolution to detect

minor fluid pressure changes, the velocities are sensitive to the contrast between different

bulk petrophysical properties of different formations. The Vp/Vs ratio (Fig. 31) is a typical

indication of lithology, particularly when correlated with acoustic impedance. Whilst the

Vp/Vs ratio approximates 2, it also shows subtle variation, particularly at depths that

correspond to overpressure zones and significant lithology shifts such as we see in a

formation change.

The 'siliciclastic' formation referred to is part of the Winning Group where claystones at the

base of the Wilcox formation grade into limestones of the Walcott formation with a mix of

claystones and minor interbedded standstones (Apthorpe, 1988; Hocking et al., 1987). Whilst

this lithology is not likely to be particularly slow, it is in relation to the underlying and

overlying lithology, hence the velocity reversal that is evident. This also highlights caution in

making simplistic interpretations of lithology based on velocity, because any velocity reversal

or increase is going to be relative to the overlying/underlying geology.

Page 79: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

69

Rather than absolute velocities, an understanding of the Poisson ratio extracted from these

velocities provides a more robust insight into the type of lithology, be it carbonate, shaly or

clean sands. Furthermore, studies have reliably correlated Vp/Vs vs. Acoustic Impedance (AI)

correlations to infer general lithology type (Pendrel et al.; Saussus and Jarvis; Vernik et al.).

This is also the premise to studies that use impedance inversion to either deterministically

match elastic properties of lithology to seismic data, or model seismic data based on lithology

type with a stochastic type feedback to acquire a match to a set of calibration wells (Pendrel et

al.; Rowbotham et al., 2003; Spikes and Dvorkin, 2003). Lithology determination from a

stochastic model-based method is independent of velocity error and far more accurate,

although, requires a calibration well in the first place and is computationally demanding. So a

simple Vp/Vs vs. AI relation can still be more useful in the frontier basin.

Figure 26 shows the base Norian TWT structure map and associated velocity map of the

varying interval velocity for a constant lithology. It shows that within a single formation,

there remains significant velocity variation which is highly dependent on depth. The deeper

the horizon, such as in the Keast Graben, the faster the velocity. The variation is great enough

to exceed any possible distinct observation of a pore pressure differential across a fault, yet

again does provide a distinct variation with depth and pressure.

The density can be seen in figure 33 where the variation in lithology creates a significant

variation in bulk rock density through the formations. This density is empirically derived from

the velocity and broadly correlates with the porosity-based trend of equation 10. The porosity-

based trend is derived from core log porosities over the entire NCB. The velocity based

densities from the Rankin Trend and Exmouth Plateau reflect the general NCB trend, and

confirm a high degree of validity for the density relationships investigated.

It is through these velocity-derived densities that we are able to calculate Sv and Shmin to place

some preliminary geomechanical constraints on the lithology. This is compared to the LOT

measured Shmin which, as figure 37 shows, is not only consistent across the entire NCB, but

increases linearly. The implications of the linear relation between stress and strain, i.e. the

elastic moduli of the material, can be seen in the linear increase in Shmin with depth

(overburden stress).

Nonetheless, as figure 38 shows, calculating Shmin from a porosity-based or velocity-

determined Poisson ratio, underestimates the Shmin at depth. This is because they rely on

bilateral constraint to determine the horizontal stress. Bilateral constraint refers to the

Page 80: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

70

expansion of a material and hence increase in horizontal stress, when it is squeezed. This is

the property of elasticity that the Poisson ratio measures, hence the reliance on the empirically

determined Poisson ratio to assess the Shmin. Calculating the stress strain ratio from bilateral

constraint relies on a linear relationship, however, as Appendix C highlights, porosity, and

therefore density and the associated bulk modulus of a material will depend on a negative

exponential relation with increasing pressure (depth). Thus the stress strain ratio appears non-

linear where it is determined from P-wave velocities. Zoback and Healy’s (1984) relation,

Hubbert and Willis’s original (1957) relation and Eaton’s (Eaton, 1969) relation result in

more accurate values for horizontal stress and generally confirm the coefficient of friction’s

nominal value of 0.6, for a vast array of circumstances, hence its validity in use for a frontier

region for example. This is because these empirical relations fix the effective stress ratio. This

fixed ESR probably begins to break down at depth as the crust transitions from the brittle to

plastic elastic moduli regime or where nonlinear viscoelastic solids, such as certain clays and

salt, dominate the lithology. However, in sedimentary basin lithology, a constant ESR is

shown to remain particularly valid. This is confirmed in figure 40, where the ratio of Shmin to

Sv, is a little over 0.6. The ratio of vertical to horizontal stress is assessed as approximately

0.62 for the Rankin, Dockrell and Keast blocks.

The determination of Shmin by bilateral constraint underestimates true values and there is an

increased uncertainty in the bilateral-based empirical relation where there is a significant

heterogeneity in the lithology or pore fluid (see Appendix E). However, the relation is

remarkably consistent. The difference between the LOT-based and bilateral constraint

velocity-based Shmin shows a power-law relation with depth (Fig. 41). This consistency

facilitates the determination of a 'correction' factor to be applied to the velocity- based ratio.

Shmin can then be calculated using this correction factor, as an alternative method to the

‘assumed’ effective stress ratio of 0.32. However, as figure 30 shows, a distinct difference in

the trend between Rankin and the Exmouth Plateau is evident, which suggests that the

underestimation will vary according to broad differences in the petrophysical properties area

to area.

Furthermore, variation from area to area is likely to be enhanced by the fact that Castagna's

(1993) empirical relation used to determine Vs has not had its constants calibrated for the

formations being studied. Whilst this correction factor can account for underestimation in

certain circumstances, figure 30 shows that care must be taken not to assume a trend over an

entire basin; the sub-basins are likely to vary in their shear wave velocity response. If a well

Page 81: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

71

calibration cannot be carried out for a study area, such a method should be avoided, and one

of the empirical relations using a fixed effective stress ratio used instead.

The distinct effect of overpressure on the effective stress (Fig. 42) manifests as a sharp

effective stress reduction on the Rankin trend at approximately 3000m and 3500 mTVDSS.

Whilst the CFF for the Rankin/Keast fault does not suggest any fault instability, it shows a

distinct deviation from the trend with depth where overpressures are reported. This is an

indication of the effect that overpressure and pore fluid retention can have on the stability of a

fault. The capacity to assess this for a range of fault blocks using seismic velocities alone, can

provide a useful initial constraint on fault risk and reservoir integrity. As figure 43 shows, for

a number of fault blocks, a similar pattern is observed, all being in the stable regime, but with

less stability around the 3000m mark where overpressure is reported.

Figure 43b shows the effect of using the incorrect Shmin from equation 15. The under

estimation of Shmin result in an incorrect shear stress and an unrealistic conclusion of fault

instability. We can apply a correction factor and in so doing acquire a reasonable

approximation on the CFF from velocities alone (Fig. 43c). The ‘corrected’ CFF shows

similar relative pressures and trends with a deviation around 3000m due to overpressure. As

previously noted, however, the problem with the corrected CFF is that the correction factor is

highly variable from area to area and cannot be applied on a general basis. Hence, where well

calibration for the study area is not available, the constant ESR methods should be applied.

Thus a confidence can be established in the calculated CFF’s used in velocity stacks. Figure

51 shows this for the theoretical planes which are illustrated in the schematic of figure 56 . It

is important to remember that the CFF is dependent on the relative strike and dip of a fault

with respect to the principle stress orientations. Faults will vary in dip and strike and a

thorough CFF analysis needs to ‘segment’ a fault to account for this variation. This study only

employed a single plane for each fault to establish the validity of the stack-based calculations.

Page 82: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

72

Figure 56 Schematic showing the theoretical planes representing the Rankin Escarpment, Rankin/Keast/Dockrell faults and E-W faults modelled in the CFF plot using the stacking velocities. The CFF plot suggests that the green planes are stable and orange planes have a high mode I/II failure risk in the current Shmax regime. The escarpment fault, represented by a orange plane, is the only fault to show evidence of reactivation in the Neogene. Maximum pressure differentials are both in the Triassic. b) schematic showing base depths of Rankin, Keast and Dockrell gas columns. Oil-water and gas-water contacts are in blue, depths are mTVDSS. Heights of columns are relative only and not actual heights. Adjacent fault block water contacts at different heights also suggest a degree of sealing capacity to the faults.

Nonetheless, the implications of the stack-based CFF’s are twofold: the capacity to acquire a

CFF assessment from seismic velocities alone, and the necessity to have either a correction

factor based on a reference well or the assumption of a fixed ESR. It is rare to not have at

least one exploration well in a basin, and whilst the comparison between the Rankin Trend

and Exmouth Plateau shows that petrophysical properties and seismic anisotropy should not

be over-extrapolated, a correction factor from a similar region could be adequate to provide an

initial approximate CFF. It is noted that a CFF based on LOT and sonic log data will always

provide a superior assessment of the CFF and stress conditions in general, and should not be

superseded by an approximation from seismic velocities. However, the contention remains

that where well based data is lacking, this framework provides far greater constraints on stress

and fault failure assessment than would otherwise be available.

Rankin Escarpment

East-West faults

Rankin/Keast fault

Max ΔP 280psi

Max ΔP 750psi

Dockrell Keast Rankin

a) b)

water contact

relatively higher

Page 83: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

73

Figure 57 Cross-section across the study area and TWT map of the regional Cretaceous unconformity. The figure shows that with faults orientated NNW-SSE and ENE-WSW at depths between 2500 and 4000 mTVDSS (approximately 1600-3000 ms TWT) there is a higher risk of fault reactivation. Pore pressure plays a large part in making this range unstable. The only faults that show evidence of reactivation in the Cenozoic are the Rankin escarpment and Goodwyn/Dockrell fault: both are in higher risk orientations. Most fault-seal field limitations therefore appear to be due to capillary seal potential which ranges between 500 and 3500 psi depending on the clay content of the fault gouge.

Brigadier Fm

Mungaroo Fm

Base regional Cretaceous Uc

Goodwyn/Dockrell fault

Between 2500 and 4500 m

TVDSS

Above 2500m and Below

4500m TVDSS Whilst pore pressure reduces stability

on the Rankin/Keast fault, it does not

touch the failure line (for CIF =0.6)

4 km

Page 84: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

74

The availability of pore pressure trends is a further constraint that can be made using the

seismic velocities. Bowers (Bowers, 1995) formula shows overlap between the RFT pressures

and velocity-based pore pressure. This indicates that the use of velocities to infer a pore

pressure range or shift may be useful where further constraints are absent. There remains a

considerable amount of uncertainty due to the scatter in the velocity data in some locations, be

it well location or shotpoint from the stack. This scatter is a product of the lithology varying

the velocity, rather than pore pressure, and leads to a standard deviation of up to 1800 psi in

the case of Keast 1 (Fig. 44). However, this deviation is lost in the stacking velocities, which

are representative of the least square mean value of the interval velocity at any given depth.

The checkshot velocities show the standard deviation to vary with depth, a reflection of the

level of heterogeneity in different lithological formations.

A pore pressure deviation from the normal trend could be due to overpressure or fluid

accumulation or may be a reflection of significant heterogeneity in the lithology. Furthermore,

whilst lateral variability is significant, figure 48 shows that poor velocity picks have a

significant effect on calculated pore pressure, giving rise to anomalous results. Hence caution

is required before taking velocity-based pore pressures as an absolute direct indicator without

well constraint. It is noted that good correlation between velocity-based pore pressure and

RFT pore pressure was only achieved where the Biot factor was taken into account (Appendix

D). This suggests that always assuming the Biot factor to be 1 may be erroneous in many

circumstances.

The degree of correlation between RFT and velocity-based pore pressures, LOT-based and

velocity-based CFF's, and log-based and velocity-based formation heterogeneity and density

variation is high. Using stacking velocities to calculate CCF and pore pressure presents an

inherent uncertainty that comes with the method by which they are picked. However, error

remains from the empirical relationship utilised to extract density and shear wave velocity.

This error could be reduced by calibrating the empirical relationships although this is

somewhat difficult in frontier basins. Error could also be reduced by using measured Vs rather

than using a calculated Vp:Vs relation, although this is often not available.

These same sources of error place uncertainty in pore pressure prediction, however, resolution

is also a key issue. Where vertical sampling of the stacking velocities is, in this case 32ms,

this can amount to more than 60m, combined with the half wavelength of the seismic wave;

on the order of 30m, single pay pressure differences can be quickly lost. Absolute pore

Page 85: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

75

1

10

100

1000

10000

Rankin 1 Dockrell 1 Keast 1 Goodwyn 6

HC

co

lum

n h

eig

ht

(lo

g m

)

Comparison of recorded HC column heights to Hg capillary based theoretical min and max heights (based on Mungaroo porosity

trends in the NCB)

Sst/Sst HC Column potential Shale/Shale HC Column potential Recorded HC column

pressure determination remains important for effective stress calculations and providing

useful information for drilling in frontier regions where mud weights and pressure kick

tolerance may be poorly constrained. However, capillary seal potential remains dependent on

the pressure differential across the seal i.e. relative pressure, rather than absolute pressure.

Thus, in the instance that a calibration well was available, theoretical capillary pressures and

hydrocarbon column heads can be constrained. Pittman's (Pittman, 1992) equation was used

as it enables a constraint to be placed on a possible range in capillary pressures, where fluid

type and clay content in fault gouge is unknown. Sand based column heights give an

indication of possible sand-sand juxtaposition, arguably the weakest link in the fault seal

capacity. Sand capillary pressures based column heights closely match measured values (Fig.

58).

Figure 58A comparison of theoretical HC column heights calculated from mercury-based capillary pressures and NCB Mungaroo (Norian) porosity trends. It suggests that where sand-sand juxtaposition is taken as the ‘weakest’ link in fault seal, significant column heights can still be achieved, assuming geomechanically stable lithology.

Page 86: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

76

One must expect that if the fault is stable, a potential HC column height will sit somewhere

between the sand and shale values, depending on the clay content of the fault gouge. Where

column heights are less than the sand value, if the fault is stable, charge access may well be an

issue. These values are based on mercury capillary threshold pressures, thus whilst they can

provide a useful range, once the source rock was defined and fluid type assessed, they could

be constrained using relationships that take this into account (Berg, 1975; Busch and Amann-

Hildenbrand, 2013; Watts, 1987) Nonetheless they proved a useful first pass assessment,

particularly when considered in conjunction with the stress regime and CFF of the fault.

Thus we can see that seismic velocities and amplitudes provide useful constraints on fault seal

assessment where good well constraint is lacking. A correlation matrix (Fig. 59) summarises

that match between measured values and seismic velocity-based empirically calculated values

in this case study. A summary of an effective workflow is presented in figure 60. This shows

that depending on the well constraint available, constants can be refined over time to make

better predictions of the fault seal capacity.

Figure 59 Correlation matrix highlights the good correlation between directly measured values and seismic-based empirically calculated values. It supports the applicability of calculating these values where well constraint is lacking.

Correlation Matrix

Good correlation between Well -

based and Velocity-based risk

assessment

Correlation Key

Page 87: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

77

Figure 60 Recommended work flow for using stacking velocities to calculate CFF, pore pressure and capillary pressure for fault seal constraint.

Page 88: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

78

7. CONCLUSIONS

This study has shown that a number of empirical relationships that are available enable a

robust approximation of pore pressures and stress magnitudes to be calculated from seismic

velocities. When combined with a basic fault architecture assessment (dip, strike, sense of

offset, and evidence of reactivation), an assessment of fault seismicity can be made. This

study successfully calculated a coulomb failure function using P-wave seismic velocities as

the only source of input. A subsequent porosity evaluation also enabled a theoretical

calculation to be made on the capillary pressures which further constrain fault seal risk. A

case study of the Rankin and Dockrell/Keast fault blocks shows that there is robust correlation

between measured data and empirically derived values. They show that the main faults

separating the Rankin and Dockrell and Keast fault blocks are in a stable stress regime and

not subject to further reactivation. They have an increased risk of failure around the base of

the top-sealing regional Jurassic and Cretaceous unconformity, where overpressure is

prevalent and principal stress ratios are large. This highlights the importance of taking all of

the following factors into consideration in fault seal risk: pore pressure, stress ratio, and dip

and strike of fault. All of these will affect whether a fault is high or low risk for either slip or

capillary-based leakage.

Rankin and Dockrell/Keast faults orientated approximately NNE-SSW in strike, with dips of

approximately 50o show no risk of slip, and have a minimum hydrocarbon column head

sealing potential of 120m .

The final uncertainty remains with the clay content of the fault gouge and fluid type, both of

which affect capillary seal potential. Whilst these two factors cannot be assessed from seismic

data alone, theoretical boundaries can be accommodated through assessment of potential

sand-sand and shale-shale potential. In this case, sand-sand capillary pressures were found to

be approximately 500 psi in the Late Triassic, corresponding to some capillary pressures

measured in the study area. Where source rock presence, maturity and type were all within

ideal windows, if the hydrocarbon column head was less than this, charge access should be

considered the primary accumulation risk.

Page 89: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

79

8. FURTHER WORK

In using the velocity stacks for CFF and pore pressure range determination, a systematic risk

profile could be implemented for an entire region. This is in the spirit of studies such as those

conducted by (Castillo et al., 2000; Reynolds et al., 2005). Whilst general strikes and dips

may put groups of faults in a high risk or low risk category, this process would need to be

refined for prospect-level evaluation whereby trap-relevant faults would need to be

segmented. This is because a fault will typically vary in dip and strike and this variation

places each respective dip and strike segment under a slightly different stress. In addition pore

pressure plays an important role in creating fault instability and, as we have seen, pore

pressure also varies both laterally and vertically. Hence, the level of resolution considered in a

fault study should be selected based on whether a risk profile is needed for a regional or

prospect level.

This should be compared to a mature basin province for a statistical Vs empirical faulted trap

comparison. This would allow a statistically robust deterministic Vs empirical comparison to

be conducted between faults that have been recorded to have sealed with their risk profile, as

calculated from the seismic velocity empirical relations.

Vp/Vs ratio's also lend use for the purpose of lithology determination. Regional velocity cubes

can be inverted and Vp/Vs vs. acoustic impedance plots compared with velocity-based

Poisson ratio's to allow a relatively robust constraint on lithology in the frontier basin.

Finally, this study presents an example of the potential application of model-based full

waveform inversion (FWI) velocity data. The accuracy of FWI velocities can reduce the

uncertainty associated with standard velocity picks and present a broader scope for the

utilisation of the FWI process in future surveys.

Page 90: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

80

9. APPENDIX A -TECTONOSTRATIGRAPHIC DIAGRAM

Overview shows the general stress regime and the main tectonic event association in the NWS(Baillie and Jacobson, 1995; Cathro and Karner, 2006; Keep et al., 2007; Longley et al., 2002). Stress Regime Epoch (Age) Tectonic Event Deposition Event in

the Carnarvon Basin Barrow Sub-basin

Compressional

Pliocene Generation of Timor Island and Trough

7 Ma. Miocene (Messinian)

Inversion

10 Ma. Miocene (Tortonian)

Inversion Delambre Fm

13 Ma. Miocene (Serravallian)

Inversion Yardie Gp

18 Ma. Miocene (Burdigalian)

Inversion Cape Range Gp

Miocene Minor siliciclastic deposition

Oligocene Australian and Eurasian (Banda Arc) Plates Collide

Carbonate sedimentation

Extensional

Late Paleogene (Oligocene)

Final Separation of Australia and Antarctica

Early Paleogene (Paleocene)

Onset of Coral Sea Rifting

Late Cretaceous

Minor Inversion due to intraplate deformation resulting from AA breakup and India migration. Creation of Barrow Anticline.

Toolonga Calcilutite

Early Cretaceous

Australia/Antarctica Breakup

Regional Transgression deposits marine (Muderong) Shale.

Australia/India Rift Separation

Crust formation of Argo & Gascoyne-Cuvier Abyssal plains

Australia/India Rift Onset Transgressive highstand deltaic complexes

Late Jurassic

West Burma Block III Separation

East and West Gondwana Break up, submarine fans accumulate large volumes of sediment in syn-rift and eustatic sea level rise and fall.

West Burma Block III Rift Onset

West Burma Block II Separation

Middle Jurassic West Burma Block II Rift Onset

Early Jurassic West Burma Block I Separation

Localised extension initiates NCB

Late Triassic

West Burma Block I Rift Onset

Interbedded sand and muds deposited in a fluvial deltaic coastal (Mungaroo) complex

Inversion

Lhasa Block Separation Pangea Breakup - Locker Shale deposited in marine settings

Permo-Triassic Boundary

Regional Transpression Tectonism

Permian sediments comprise marine and marginal marine successions

Compressional

Mid-Carboniferous Gondwana and Laurussia collide to form Pangea

Late Carboniferous glacial conditions

Devonian Uplift and warping, marine conditions.

1 Km of Devonian Sediments with Frasnian reef complex.

Ordovician-Silurian Carnarvon intracratonic basin formation in Gondwana

Page 91: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

81

10. APPENDIX B – STRATIGRAPHIC COLUMN FOR THE WEST DAMPIER

SUB-BASIN

(Bennett, 2005)

Page 92: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

82

11. APPENDIX C – POROSITY AND PERMEABILITY IN THE NCB

The porosity and permeability trends for the NCB were acquired from 30 wells within the

NCB, biased towards wells along or near the Rankin Trend (Fig. 61). All porosity and

permeability data are from core analysis on samples from the Mungaroo formation of the

Triassic Period, NCB, NWS. This selection method was chosen so as not to 'skew' trends by

mixing formations. The Triassic was most pertinent to this study because fault seal risk

uncertainty is targeted in the Triassic. In addition, both mixed and single lithology trends have

been acquired, however, due to the core sampling bias towards the 'reservoir', sandstone has

the most reliable statistics.

Figure 61 The arithmetic average for 32 Wells in the NCB. 4 Wells were excluded as their inboard and shallow setting mark a statistically significant separation from the rest.

Broad trends in porosity with depth relations are very similar, irrespective of the lithology

(Fig. 62). The trends confirm to the typical (Mondol et al., 2007; Weaver, 1989; Wood, 2010)

exponential-type relationship between porosity and depth. Shales will tend to have a much

faster reduction in porosity with depth as the clay mineralogy dewaters and compacts. This

Dailey 1

Griffin 1

Griffin 1

North Gorgon 3

North Gorgon 4

Long Island 1

Observation 1

West Tryal Rocks 1

Bowers 1

Nimrod 1ST3Hilada 1A

Gorgon 1

Gorgon 3

Sirius 1 Briseis 1Maenad 1

Bluebell 1

Orthrus 1

Vinck 1

Glencoe 1

Rankin 1Jupiter 1

Goodwyn 6Malmsey 1

Noblige 1Echo 1Yodel 1 Keast 1

Dockrell 1Lady Nora 2

Athena 1

North Rankin 1

North Rankin 2

2000.00

2500.00

3000.00

3500.00

4000.00

4500.00

5.00 10.00 15.00 20.00 25.00

De

pth

(m)

Porosity %

Averaged Porosity vs Average Depth for the Triassic (Norian) Mungaroo Formation in the North Carnarvon Basin

Page 93: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

83

then slows where further porosity reduction is due to diagenetic reactions. Sandstone

compacts much less in the first instance, then slows where further porosity reduction is due to

diagenesis and cementation. The Mungaroo Formation is an intra-formational claystone

sandstone formation and it is the depositional environment that is by far the greatest

determining factor of the ultimate porosity trend and porosity permeability relationship in a

formation. Sorting and maturity, along with clay content, have a large impact on the ultimate

porosity permeability trend. We see this in figures 63 and 64 for example where the 'slope' of

the porosity permeability plot is different depending on well location and its depositional

setting, with wells in steep slope setting having coarser and typically cleaner sandstones.

Results from this study suggest that good correlation and control should allow depositional

setting to be inferred from the porosity permeability trend alone. However, in this instance,

the porosity permeability trend used in Pittman’s (Pittman, 1992) relationship was taken from

the Rankin Trend only wells, for sandstone lithology, as this lithology type is biased in the

RFT data and hence allows for more robust comparison. Capillary pressure determination and

subsequent hydrocarbon column height calculation were assessed using sandstone, shale and

mixed lithologies for comparison.

Figure 62 Porosity with depth trends show typical exponential curves. Shale porosity reduction occurs quicker than with sandstone, however, broad trends are similar. These relationships are used for the Shmin and capillary pressure relations in the study.

y = 8389.8e-0.078x

R² = 0.5405

y = 3262.8e-0.084x

R² = 0.9229

y = 3313.7e-0.047x

R² = 0.4176

0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

1.00 6.00 11.00 16.00 21.00 26.00 31.00

De

pth

(m

)

Porosity %

100 m average porosity for different lith types in the Triassic (Norian) Mungaroo Fm

SstShSlstClst

Porosity Permeability Trends for sandstone in the Triassic (Norian) Mungaroo Fm. - North Carnarvon Basin

Dep

th m

BM

L

Page 94: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

84

Figure 63 Porosity vs. Permeability plot showing that lithology is not a primary determining factor in the resulting porosity-permeability relationship.

0.001

0.01

0.1

1

10

100

1000

10000

100000

0 5 10 15 20 25 30 35 40

Pe

rme

abili

ty (

mD

)

Porosity %

Poro Perm Trends for sandstone in the Triassic (Norian) Mungaroo Fm - North Carnarvon Basin

Sst Sandstone

Calc Sltst Calcic Siltstone

Sst & Sh Sandy Siltstone

Ls Limestone

Sltst Siltstone

Clst Claystone

Sh Shale

Slty Sh Slty Clst

Slty Sst Silty Sandstone

Marl limy mudstone

Page 95: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

85

0.00

0.01

0.10

1.00

10.00

100.00

1000.00

10000.00

100000.00

0.00 5.00 10.00 15.00 20.00 25.00 30.00 35.00 40.00

Pe

rme

abili

ty (

mD

)

Porosity %

Poro Perm Trends for sandstone in the Triassic (Norian) Mungaroo Fm - NWS Bluebell 1

Bowers 1

Briseis 1

Clio 2

Glencoe 1

Gorgon 3

Gorgon 1

Long Island 1

Nimblefoot 1

North Gorgon 3

North Gorgon 4

North Gorgon 1

Observation 1

Onslow 1

Sirius 1

West Tryal Rocks 1

Zeepaard 1

Goodwyn 6

Malmsey 1

Nickols Bay 1

Noblige 1

Vinck 1

Athena 1

Dockrell 1

Echo 1

Gaea 1

Keast 1

North Rankin 1

North Rankin 2

Yodel 1

Figure 64 Porosity/log permeability plot highlighting how well location and associated depositional environment area greater determining factor in the porosity-permeability trend. Inset: Gamma ray and resistivity logs indicating (shale content) and GR trend matching the depositional environment.

Clio 1

Glencoe 1

Nimblefoot 1

Gorgon 3

Briseis 1

Coarse and/or Clean

Fine and/or Shaly

Page 96: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

86

12. APPENDIX D - PORE PRESSURE

This study uses Bowers (1995) relation for pore pressure calculations using seismic velocities.

Bowers (1995) relation employs constants derived from a so-called ‘virgin’ curve. This

describes the increase in velocity effective stress relation with depth where fluids do not affect

the trend by reducing the effective stress. The virgin curve constants in this were derived from

a corrected sonic log from the Rankin 1 Well (Fig. 65). This sonic log curve was ‘corrected’

to account for hydrocarbon accumulations, hence giving the background or ‘virgin’ relation.

Figure 65 Pore pressure Vs, effective stress from sonic log values, corrected for effects of background fluid pressure. The power regression fit provides constants applied to Bowers pore pressure empirical relationship.

In Bowers’ (1995) relation the mudline velocity is subtracted from the interval velocity and is

typically approximated as 1500 m/s; seawater velocity. In this study, the value of 1250 m/s

was used, as this relates to the velocity extrapolated to zero pore pressure, because the sea

column at the mudline exerts 100's of psi in pressure, depending on depth. The typical water

depth in this study was 120 m, relating to 190 psi, so such an extrapolation is considered to

better account for the effect of seawater column height on hydrostatic pressure.

y = 16.5x0.6

R² = 0.6

0

1000

2000

3000

4000

5000

6000

2000 4000 6000 8000 10000 12000 14000 16000 18000

Po

re p

ress

ure

(p

si)

Stress (psi)

Virgin Pore Pressure trend based on a hydrocarbon corrected Vp

Sonic Log from Rankin 1

Page 97: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

87

Poroelasticity refers to the time-dependant deformation of a porous elastic solid.

Poroelasticity results in pore pressure within a porous solid responding to an external pressure

applied and affecting the response to that external pressure. Because of this response, the pore

pressure can affect the measured bulk modulus of the material. The Biot coefficient attempts

to account to the poroelastic effect on the bulk modulus (seismic response) of a material. This

study infers this poroelastic response effect on the seismic-based pore pressure calculations by

multiplying the seismic-based pore pressure by a poroelastic (Biot) correction factor. This

correction factor is estimated by taking the ratio of RFT pore pressure to seismic-based pore

pressure. The average correction factor is approximately 0.8. Whilst there is up to a 50%

variation in this factor, the results of applying 0.8 across all values provide a very good

correlation for wells and fault blocks, whereas assuming a correction factor of 1 leads to

erroneous results.

Figure 66 A value for poroelasticity is taken from the ratio of empirically based pore pressure and RFT based pore pressure. An average is taken as 0.8, seen to be the approximate median in this scatter plot and which results in consistently good agreement between calculated and measured pore pressure.

0.0

500.0

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

4000.0

4500.0

5000.0

0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20

De

pth

mTV

DSS

Poroelasticity correction factor

Rankin Trend and Exmouth Plateau Wells

Rankin

Dockrell

Keast

Page 98: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

88

y = 96.684x0.3921

R² = 0.7658

0

500

1000

1500

2000

2500

3000

3500

4000

0.00 1000.00 2000.00 3000.00 4000.00 5000.00 6000.00 7000.00 8000.00 9000.00

TVD

SS (

ms)

Pore Pressure (psi)

Pore pressure comparison between RFT and velocity measurements on Keast 1

Velocity based porepressure

RFT

HDVA stack SPT

Power (Velocity basedpore pressure)

Figure 67 This pore pressure plot using a Biot factor of 1 does not fit the RFT data, rather it is anomalously high. This is consistent with all the blocks, and the effect of applying the correction factor of 0.8 can be seen in the results (section 5.2).

Page 99: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

89

13. APPENDIX E - EMPIRICAL CALCULATION OF SHEAR WAVE VELOCITY

Figure 68 compares Vs measured from the sonic log of the Rankin 1 well to Vs calculated

using Castagna's (1993) Vp relation. The relation is based on an empirical 'fit' from laboratory

tests on feldspathic sandstone. The comparison plot shows good correlation with the linear

regression fit (pseudo mudrock line) shows a Vp of approximately 1500 m/s where Vs is zero.

The correlation plot (Fig. 69) is another qualifier of the appropriateness of using Castagna's

(1993) relation. Some scatter persists in the correlation plot, as is expected with heterogeneity

in the lithology. More scatter (a poorer fit in the empirical relation) occurs at depths where the

Mungaroo formation is penetrated, as well as at depths with significant gas flow is recorded.

Considering the lack of calibration of this empirical relation, the correlation was remarkably

good, considering the context of having a relation that could be applied in a frontier basin

where no calibration well is available.

Figure 68 Comparison of measured to empirically calculated S-wave velocity plotted vs. the P-wave velocity. The plot show good correlation with the linear regression fit representing a pseudo mudrock line. The mudrock line confirms a P-wave velocity of approximately 1500 m/s where shear wave velocity is zero.

0

1000

2000

3000

4000

5000

6000

0 500 1000 1500 2000 2500 3000 3500

Vp

m/s

Vs m/s

Vp Vs plot comparing measured and calculated Vs

Sonic log measurement

Vs calculated from Castagna's

(1993) Vp relation

Linear (Sonic log measurement)

Page 100: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

90

Figure 69 The correlation plot of calculated vs. measured shear wave velocity. Increased scatter indicates poorer correlation and therefore poorer validity of the empirical trend used. There is a poorer fit at depths correlating with greater lithological heterogeneity and overpressured zones.

0

500

1000

1500

2000

2500

3000

3500

4000

0 500 1000 1500 2000 2500 3000 3500 4000

Vs

(cal

cula

ted

fro

m C

asta

gna

(19

93

) Vp

re

lati

on

)

Vs (measured from sonic log)

Correlation plot of measured vs. calculated Vs

2500-2800

2800-3100

3100-3400

3400-3700

3700-4000 ●

● ●

Depth range mTVDSS

Page 101: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

91

14. APPENDIX F - SHEAR AND NORMAL STRESS

σn and σs are the normal and shear stress components of the principal stress tensors σ1 σ2 and

σ3 . As described in the method, the principal stresses can be determined from a combination

of overburden stress, LOT measurements and either mud weight recordings, where well data

is available, or constraining the intermediate stress to one that fits the Andersonian tectonic

regime, as defined by evidence in the fault architecture. The three dimensional vector

components of these principal stress vectors acting on a fault plane, i.e. the normal and shear

stress, is calculated via a third order tensor, basis vector transformation, whereby the vectors

are rotated to the coordinate system of the fault plane from that of the orthogonal Cartesian

system of the principal stress vectors. The equations used in this study for the calculation of

σn and σs, as described by Zoback (2010) and Jaeger et al. (2007) are as follows:

𝜎𝑆 = 𝑎11𝑎12𝜎1 + 𝑎12𝑎22𝜎2 + 𝑎13𝑎23𝜎3 ...Eqn. 17

𝜎𝑛 = 𝑎112 𝜎1 + 𝑎12

2 𝜎2 + 𝑎132 𝜎3 ...Eqn. 18

where aij are the matrix components as follows,

𝐴 = 𝑐𝑜𝑠𝛾𝑐𝑜𝑠𝜆 𝑐𝑜𝑠𝛾𝑠𝑖𝑛𝜆 −𝑠𝑖𝑛𝛾−𝑠𝑖𝑛𝜆 𝑐𝑜𝑠𝜆 0

𝑠𝑖𝑛𝛾𝑐𝑜𝑠𝜆 𝑠𝑖𝑛𝛾𝑠𝑖𝑛𝜆 𝑐𝑜𝑠𝛾 ...Eqn, 19

where γ is the angle between the fault normal and minimum principal stress vector, and λ is

the angle between the strike of the fault and the plane perpendicular to the minimum

principal stress vector.

Page 102: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

92

15. APPENDIX G – THERMAL GRADIENT FOR NCB

Figure 70 A thermal gradient of approximately 29.3 + 3.0 oC/km is measured from 28 wells in the NCB. This thermal gradient is measured from a least squares linear regression fit excluding values measured within the Gorgon field as it is consistently assessed as being anomalously ‘hot’ compared with the rest of the basin. Larsen Deep 1 off the Exmouth Plateau is another that appears relatively hot in comparison.

0.00

1000.00

2000.00

3000.00

4000.00

5000.00

6000.00

0.00 20.00 40.00 60.00 80.00 100.00 120.00 140.00 160.00 180.00

De

pth

(mB

ML)

Temperature ( C)Temperature Gradient - North Carnarvon Basin

Gorgon and Larsen Deep 1

Page 103: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

93

16. REFERENCES

Allen, P.A. and Allen, J.R., 2005. Basin Analysis: Principles and Applications. Blackwell Publishing, Malden, MA.

Angelier, J., 1979. Determination of the mean principal directions of stresses for a given fault population. Tectonophysics, 56(3–4): T17-T26.

Apthorpe, M., 1988. Cainozoic Depositional History of the North West Shelf. In: P.G. Purcell and R.R. Purcell (Editors), Petroleum Exploration Society of Australia Symposium. PESA, Perth, pp. 55-84.

Aydin, A., 2000. Fractures, faults, and hydrocarbon entrapment, migration and flow. Marine and Petroleum Geology, 17(7): 797-814.

Bailey, W.R., Underschultz, J., Dewhurst, D.N., Kovack, G., Mildren, S. and Raven, M., 2006. Multi-disciplinary approach to fault and top seal appraisal; Pyrenees–Macedon oil and gas fields, Exmouth Sub-basin, Australian Northwest Shelf. Marine and Petroleum Geology, 23(2): 241-259.

Baillie, P.W. and Jacobson, E., 1995. Structural Evolution of the Carnarvon Terrace, Western Australia. Australian Petroleum and Exploration Association Journal, 35: 321-332.

Bennett, K., 2005. Demeter 3D Seismic Interpretation Report, Northwest Shelf Venture Area, Carnarvon Basin, Australia., Woodside Petroleum Ltd.

Bense, V.F., Gleeson, T., Loveless, S.E., Bour, O. and Scibek, J., 2013. Fault zone hydrogeology. Earth-Science Reviews, 127(0): 171-192.

Berg, R.R., 1975. Capillary pressure in stratigraphic traps. AAPG Bulletin, 59: 939-956. Bowers, G.L., 1995. Pore Pressure Estimation From Velocity Data: Accounting for Overpressure

Mechanism Besides Undercompaction. SPE Drilling and Completion, 27488: 89-95. Bretan, P., Yielding, G. and Jones, H., 2003. Using calibrated shale gouge ratio to estimate

hydrocarbon column heights. AAPG Bulletin, 87(3): 397-413. Brown, A.R., 2004. Interpretation of Three-Dimensional Seismic Data. AAPG Memoir 42

SEG Investigations in Geophysics, No. 9. AAPG and the Society of Exploration Geophysicists, Tulsa, Oklahoma.

Busch, A. and Amann-Hildenbrand, A., 2013. Predicting capillarity of mudrocks. Marine and Petroleum Geology, 45(0): 208-223.

Byerlee, J., 1978. Friction of Rocks. Pure and applied geophysics, 116(4-5): 615-626. Castagna, J.P., 1993. Petrophysical imaging using AVO, The Leading Edge. Society of Exploration

Geophysics, Tulsa, Oklahoma, pp. 172-178. Castagna, J.P., Batzle, M.L. and Eastwood, R.L., 1985. Relationships between compressional-wave and

shear-wave velocities in clastic silicate rocks. Geophysics, 50(4): 571-581. Castillo, D.A., Bishop, D.J., Donaldson, I., Kuek, D., deRuig, M., Trupp, M. and Shuster, M.W., 2000.

Trap Integrity in the Laminaria High-Nancar Trough Region, Timor Sea: Prediction of Fault Seal Failure Using Well-Constrained Stress Tensors and Fault Surfaces Interpreted from 3D Seismic, APPEA. APPEA, Brisbane, pp. 151-173.

Cathro, D.L. and Karner, G.D., 2006. Cretaceous–Tertiary inversion history of the Dampier Sub-basin, northwest Australia: Insights from quantitative basin modelling. Marine and Petroleum Geology, 23(4): 503-526.

Childs, C., Sylta, Ø., Moriya, S., Morewood, N., Manzocchi, T., Walsh, J.J. and Hermanssen, D., 2009. Calibrating fault seal using a hydrocarbon migration model of the Oseberg Syd area, Viking Graben. Marine and Petroleum Geology, 26(6): 764-774.

Dehandschutter, B., Gaviglio, P., Sizun, J.P., Sintubin, M., Vandycke, S., Vandenberghe, N. and Wouters, L., 2005. Volumetric matrix strain related to intraformational faulting in argillaceous sediments. Journal of the Geological Society, 162(5): 801-813.

Page 104: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

94

Dewhurst, D.N. and Hennig, A.L., 2003. Geomechanical properties related to top seal leakage in the Carnarvon Basin, Northwest Shelf, Australia. Petroleum Geoscience, 9(3): 255-263.

Dewhurst, D.N., Jones, R.M. and Raven, M.D., 2002. Microstructural and petrophysical characterization of Muderong Shale: application to top seal risking. Petroleum Geoscience, 8(4): 371-383.

Dix, C.H., 1955. Seismic Velocities from Surface Measurements. Geophysics, 20(1): 68. Dodds, K.J., Dewhurst, D.N., Siggins, A.F., Ciz, R., Urosevic, M., Gurevich, B. and Sherlock, D.H., 2007.

Experimental and theoretical rock physics research with application to reservoirs, seals and fluid processes. Journal of Petroleum Science and Engineering, 57(1–2): 16-36.

Dvorkin, J.P., 2008. Yet another Vs equation. Geophysics, 73(2): 35-39. Dyksterhuis, S. and Muller, R.D., 2008. Cause and evolution of intraplate orogeny in Australia.

Geology, 36: 495-498. Eaton, B.A., 1969. Fracture gradient prediction and its application in oilfield operations. Journal of

Petroleum Technology, 246: 1353-1360. Etheridge, M., McQueen, H. and Lambeck, K., 1991. The Role of Intraplate Stress in Tertiary (and

Mesozoic) Deformation of the Australian Continent and Its Margins: A Key Factor in Petroleum Trap Formation. Exploration Geophysics, 22: 123-128.

Færseth, R.B., Johnsen, E. and Sperrevik, S., 2007. Methodology for risking fault seal capacity: Implications of fault zone architecture. AAPG Bulletin, 91(9): 1231-1246.

Fisher, Q.J., Harris, S.D., McAllister, E., Knipe, R.J. and Bolton, A.J., 2001. Hydrocarbon flow across faults by capillary leakage revisited. Marine and Petroleum Geology, 18(2): 251-257.

Fisher, Q.J. and Knipe, R.J., 2001. The permeability of faults within siliciclastic petroleum reservoirs of the North Sea and Norwegian Continental Shelf. Marine and Petroleum Geology, 18(10): 1063-1081.

Gardner, G., Gardner, L. and Gregory, A., 1974. Formation velocity and density - The diagnostics basics for stratigraphic traps. Geophysics, 39(6): 770-780.

Geoscience_Australia, 2010. Barrow Sub-basin, Northern carnarvon basin release areas w10-15, w10-16, w10-17, w10-18 and w10-19. In: Department, of, Resources, Energy and Tourism (Editors). Australian Government, pp. 1-42.

Geoscience_Australia, 2014. Earthquakes@GA, Australia. Harrowfield, M., Cunneen, J., Keep, M. and Crowe, W., 2003. Early-stage orogenesis in the Timor Sea

region, NWAustralia. Geological Society of London, 160: 991-1001. Hasegawa, S., Sorkhabi, R., Iwanaga, S. and Sakuyama, N., 2005. Fault-seal Analysis in the Temana

Field, Offshore Sarawak, Malaysia. In: R. Sorhabi and Y. Tsuji (Editors), Faults, Fluid Flow and Petroleum Traps. AAPG Memoir 85, pp. 43-58.

Hatchell, P. and Bourne, S., 2005. Rocks under strain: Strain-induced time-lapse time shifts are oserved for depleting reservoirs, The Leading Edge. Society for Exploration Geophysics.

Hillis, R.R. and Reynolds, S.D., 2003. In situ stress field of Australia. In: R.R. Hillis and M. R.D. (Editors). Geological Society of Australia, pp. 49-58.

Hillis, R.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. In: H. JOHNSON, A.G. DORE´ , R.W. GATLIFF, R. HOLDSWORTH, E.R. LUNDIN and J.D. RITCHIE (Editors), Geological Society London, Special Publications. The Geological Society of London, pp. 71-90.

Hillis, R.R. and Williams, A.F., 1993a. The contemporary stress field of the Barrow-Dampier Sub-Basin 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. Australian Petroleum and Exploration Association Journal, 33: 373-385.

Hocking, R.M., Moors, H.T. and Van de Graaff, W.J.E., 1987. Geology of the Carnarvon Basin Western Australia, Geological Survey of Western Australia

Hubbert, M.K., 1953. Entrapment of Petroleum under hydrodynamic conditions. AAPG Bulletin, 37: 1954-2026.

Page 105: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

95

Hubbert, M.K. and Willis, D.G., 1957. Mechanics of Hydraulic fracturing. Petroleum Transactions of the American Institue of Mechanical Engineers, 210: 153-169.

Jaeger, J.C., Cook, N.G.W. and Zimmerman, R.W., 2007. Fundamentals of Rock Mechanics. Blackwell Publishing, USA.

Jones, R.M. and Hillis, R.R., 2003. An integrated, quantitative approach to assessing fault-seal risk. AAPG Bulletin, 87(3): 507-524.

Kachi, T., Yamada, H., Yasuhara, K., Fujimoto, M., Hasegawa, S., Iwanaga, S. and Sorhabi, R., 2005. Fault Seal Analysis Applied to the Erawan Gas-condensate Field in the Gulf of Thailand. In: R. Sorhabi and Y. Tsuji (Editors), Faults, Fluid flow and petroleum traps. AAPG Memoir 85, pp. 59-78.

Kearey, P., Brooks, M. and Hill, I., 2002. An Introduction to Geophysical Exploration. Blackwell Publishing, London.

Keep, M., Bishop, A. and Longley, I., 2000. Neogene wrench reactivation of the Barcoo Sub-basin, northwest Australia: implications for Neogene tectonics of the northern Australian margin. Petroleum Geoscience, 6: 211-220.

Keep, M. and Harrowfield, M., 2008. Elastic flexure and distributed deformation along Australia’s North West Shelf: Neogene tectonics of the Bonaparte and Browse basins. In: H. JOHNSON, A.G. DORE ´ , R.W. GATLIFF, R. HOLDSWORTH, E. LUNDIN and J.D. RITCHIE (Editors), The Nature and Origins of Compressive Margins. Geological Society, London, pp. 185-200.

Keep, M., Harrowfield, M. and Crowe, W., 2007. The Neogene tectonic history of the North West Shelf, Australia. Exploration Geophysics, 38: 151-174.

Keep, M., Longley, I. and Jones, R., 2002. Sumba and its effect on Australia's north-western margin. Geological Society of Australia Special Publication, 22: 303-312.

Keep, M. and Moss, S.J., 2000. Basement reactivation and control of Neogene structures in the Outer Browse Basin, North West Shelf. Exploration Geophysics, 31: 424-432.

King, R.C., Neubauer, M., Hillis, R.R. and Reynolds, S.D., 2010. Variation of vertical stress in the Carnarvon Basin, NW Shelf, Australia. Tectonophysics, 482(1–4): 73-81.

Lockner, D.A., Morrow, C., Moore, D. and Hickman, S., 2011. Low strength of deep San Andreas fault gouge from SAFOD core. Nature, 472(7341): 82-85.

Longley, I., Buessenschuett, C., Clydsdale, L., Cubitt, C.L., Davis, R.C. and 2002. The North West Shelf of Australia - a Woodside perspective. In: M. Keep and S. Moss (Editors), Western Australian Basins Symposium. Petroleum Exploration Society of Australia, Perth, pp. 27-86.

Losh, S., 1998. Oil Migration in a Major Growth Fault: Structural analysis of the Pathfinder core, South Eugene Island Block 330, Ofshore Louisiana. AAPG Bulletin, 82(9): 1694-1710.

Mandl, G., 2000. Faulting in Brittle Rocks. Springer-Verlag, Germany. Mondol, N.H., Biolykke, K., Jahren, J. and Hoeg, K., 2007. Experimental mechanical compaction of clay

mineral aggregates—Changes in physical properties of mudstones during burial. Marine and Petroleum Geology, 24: 289-311.

Moore, D.E. and Lockner, D.A., 2004. Crystallographic controls on the frictional behavior of dry and water-saturated sheet structure minerals. Journal of Geophysical Research: Solid Earth, 109(B3): B03401.

Moore, D.E., Lockner, D.A., Ito, H., Ikeda, R., Tanaka, H. and Omura, K., 2009. Geometry of the Nojima Fault at Nojima-Hirabayashi, Japan - II. Microstructures and their Implications for Permeability and Strength. Pure & Applied Geophysics, 166(10-11): 1669-1691.

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.

Ngwenya, B.T., Elphick, S.C., Main, I.G. and Shimmield, G.B., 2000. Experimental constraints on the diagenetic self-sealing capacity of faults in high porosity rocks. Earth and Planetary Science Letters, 183(1–2): 187-199.

O'Brien, G.W., Morse, M., Wilson, D., Quaife, P., Colwell, J., Higgins, R. and Foster, C.B., 1997. Margin-Scale, Basin-Involved Compartmentalisation of Australia's North West Shelf: A Primary

Page 106: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

96

Control on Basin Scale Rift, Depositional and Reactivation Histories. Australian Petroleum Production and Exploration Association Journal, 39: 40-61.

Pendrel, J., Debeye, H., Pedersen‐Tatalovic, R., Goodway, B., Dufour, J., Bogaards, M. and Stewart, R.R., Estimation and interpretation of P and S impedance volumes from simultaneous inversion of P‐wave offset seismic data, SEG Technical Program Expanded Abstracts 2000, pp. 146-149.

Pittman, E.D., 1992. Relationship of Porosity and Permeability to Various Parameters Derived from Mercury Injection-Capillary Pressure Curves for Sandstone. AAPG Bulletin, 76(2): 191-198.

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., Paraschivoiu, E., Hillis, R.R. and Obrien, G.W., 2005. A Regional Analysis of Fault Reactivation and Seal Integrity Based on Geomechanical Modeling: An Example from the Bight Basin, Australia. In: P. Boult and J. Kaldi (Editors), AAPG Hedberg Series. Evaluating fault and cap rock seals, pp. 57-71.

Rowbotham, P.S., Marion, D., Lamy, P., Insalaco, E., Swaby, P.A. and Boisseau, Y., 2003. Multidisciplinary stochastic impedance inversion: integrating geological understanding and capturing reservoir uncertainty. Petroleum Geoscience, 9(4): 287-294.

Saussus, D. and Jarvis, K., Extracting detailed lithology from seismic data, ASEG Extended Abstracts 2009: 20th Geophysical Conference, pp. 1-5.

Sayers, C.M., 2006. An introduction to velocity-based pore-pressure estimation. The Leading Edge, 25(12): 1496-1500.

Schowalter, T.T., 1979. Mechanisms of secondary hydrocarbon migration and entrapment. AAPG Bulletin, 63: 723-760.

Shepherd, M., 2009. Structural Geology: Faults. In: M. Shepherd (Editor), Oil field production geology. AAPG Memoir 91, pp. 107-122.

Shipton, Z.K., Evans, J.P. and Thompson, L.B., 2005. The Geometry and Thickness of Deformation-band Fault Core and its Influence on Sealing Characteristics of Deformation Fault Zones. In: R. Sorhabi and Y. Tsuji (Editors), Faults, fluid flow and petroleum traps. AAPG Memoir 85, pp. 181-195.

Smith, D.A., 1966. Theoretical considerations of sealing and non-sealing faults. AAPG Bulletin, 50: 363-374.

Soleymani, H. and Riahi, M.-A., 2012. Velocity based pore pressure prediction—A case study at one of the Iranian southwest oil fields. Journal of Petroleum Science and Engineering, 94–95(0): 40-46.

Spikes, K.T. and Dvorkin, J.P., 2003. Model-based prediction of porosity and reservoir quality from P- and S-wave data. Geophysical Research Letters, 30(20): 2029.

Swarbrick, R.E. and Osborne, M.J., 1998. Mechanisms that Generate Abnormal Pressures: An Overview. In: B.E. Law, G.F. Ulmishek and V.I. Slavin (Editors), AAPG Memoir, pp. 13-34.

Terzaghi, K., 1923. Die Berechnung der Durchlassigkeitsziffer des Tones aws dem Verlanf der Hydrodynamischen Spannungserscheinungen. Sb. Akad. Wiss, Wien: 132-135.

Tingay, M.R.P., Hillis, R.R., Morley, C.K., Swarbrick, R.E. and Okpere, E.C., 2003. Variation in vertical stress in the Baram Basin, Brunei: tectonic and geomechanical implications. Marine and Petroleum Geology, 20(10): 1201-1212.

Torabi, A., Fossen, H. and Braathen, A., 2013. Insight into petrophysical properties of deformed sandstone reservoirs. AAPG Bulletin, 97(4): 619-637.

Tueckmantel, C., Fisher, Q.J., Knipe, R.J., Lickorish, H. and Khalil, S.M., 2010. Fault seal prediction of seismic-scale normal faults in porous sandstone: A case study from the eastern Gulf of Suez rift, Egypt. Marine and Petroleum Geology, 27(2): 334-350.

Turner, J.P. and Williams, G.A., 2003. Sedimentary Basin Inversion and Intra-Plate Shortening. Earth Science Reviews, 65: 277-304.

Vernik, L., Fisher, D. and Bahret, S., Estimation of net‐to‐gross from P and S impedance: Part II — 3D seismic inversion, SEG Technical Program Expanded Abstracts 2001, pp. 211-214.

Page 107: Fault Seal Analysis: Constraining fault seal risk using ...Fault seal capacity is an important component in the conventional petroleum system. Assessing the capacity for a fault to

97

Watts, N.L., 1987. Theoretical aspects of cap-rock and fault seals for single- and two-phase hydrocarbon columns. Marine and Petroleum Geology, 4(4): 274-307.

Weaver, C.E., 1989. Clays, Muds and Shales. Elsevier, Amsterdam. Withjack, M.O. and Jamison, W.R., 1985. Deformation produced by oblique rifting. Tectonophysics,

126: 99-124. Wood, L.J., 2010. Shale Tectonics. AAPG Memoir, 93. AAPG, Tulsa, OK. Wyllie, M., Gregory, A. and Gardner, L., 1956. ELASTIC WAVE VELOCITIES IN HETEROGENEOUS AND

POROUS MEDIA. GEOPHYSICS, 21(1): 41-70. Yielding, G., Freeman, B. and Needham, D.T., 1997. Quantitative Fault Seal Prediction. AAPG Bulletin,

81: 897-917. Yilmaz, O., 2008. Seismic Data Analysis - Processing, Inversion and Interpretation of Seismic Data, 1.

Society of Exploration Geophysicists, USA. Zhang, J., 2011. Pore pressure prediction from well logs: Methods, modifications, and new

approaches. Earth-Science Reviews, 108(1–2): 50-63. Zoback, M.D., 2010. Reservoir Geomechanics. Cambridge University Press, UK. Zoback, M.D. and Healy, J.H., 1984. Friction, Faulting and in-situ stress. Annales Geophysicae, 2: 689-

698.