Sei sulla pagina 1di 11

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

net/publication/317061233

A new approach for estimating the amount of eroded sediments, a case study
from the Canning Basin, Western Australia

Article  in  Journal of Petroleum Science and Engineering · May 2017


DOI: 10.1016/j.petrol.2017.05.008

CITATION READS

1 251

5 authors, including:

Lukman Mobolaji Johnson Reza Rezaee


Curtin University Curtin University
12 PUBLICATIONS   10 CITATIONS    196 PUBLICATIONS   1,626 CITATIONS   

SEE PROFILE SEE PROFILE

Ali Kadkhodaie Gregory Smith


University of Tabriz Curtin University
130 PUBLICATIONS   875 CITATIONS    7 PUBLICATIONS   2 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Rock Typing and Total Gas content of Shale Gas Reservoirs. View project

Gas sorption hysteresis in coals & shales View project

All content following this page was uploaded by Lukman Mobolaji Johnson on 29 January 2018.

The user has requested enhancement of the downloaded file.


Journal of Petroleum Science and Engineering 156 (2017) 19–28

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

A new approach for estimating the amount of eroded sediments, a case


study from the Canning Basin, Western Australia
Lukman Mobolaji Johnson a, *, Reza Rezaee a, Ali Kadkhodaie a, d, Gregory Smith b,
Hongyan Yu a, c
a
Department of Petroleum Engineering, Curtin University, Perth, Australia
b
Department Applied Geology, West Australian School of Mines, Curtin University, Perth, Australia
c
State Key Laboratory of Continental Dynamics, Xi'an, PR China
d
Department of Geology, Faculty of Natural Sciences, University of Tabriz, Tabriz, Iran

A R T I C L E I N F O A B S T R A C T

Keywords: In order to accurately reconstruct the burial and thermal history of sedimentary basins, one of the key uncertainty
Exhumation thickness that should be addressed is the accurate delineation of the maximum burial depths and estimation of the thickness
Burial history of eroded sections in the basin. Several methods have been routinely utilized to account for this, but each with its
Sonic transit travel time
limitations. Here, we have utilized the sonic transit time data to account for the amounts of removed sections in
Compaction curve
the Broome Platform of the partly exhumed Canning Basin, Western Australia. The observed exhumed values are
Canning Basin
further compared with the values from AFTA (Apaptite Fission Track Analysis) data as well as results of exhu-
mation observed from the projection of vitrinite reflectance data. Furthermore, a mathematical relationship is
established between the amounts of exhumation and the sonic transit time (DT) and depth. The results from the
calculated exhumed sections show a good correlation with the estimated removed sections reported from AFTA
but poor correlations with the projected vitrinite reflectance profiles. While AFTA and vitrinite reflectance
methods are widely used in the industry, the exhumation estimates derived from sonic-porosity logs are usually
independent of the thermal history of the basin and thereby would provide more accurate inputs or further
constraints into the burial history model. Also, in this study, the results from the AFTA usually allows for a wide
range of uncertainty, while the calculation of exhumation values from the new approach gives an absolute
exhumation value for the well locations. Furthermore, the sonic log approach has been able to identify and es-
timate the magnitude of an older erosional event that was not previously reported in the Geotrack report based on
thermal history methods.

1. Introduction unconventional shale gas resource, with “limited” information from the
Ordovician shales (Triche and Bahar, 2013). Carlsen and Ghori (2005)
Understanding the amount of exhumed sections in sedimentary ba- suggested that the Canning Basin is one of the least explored Palaeozoic
sins is a key step in the determination of the basins prospectivity. This is systems in the world, and detailed understanding of some of the systems
important in terms of defining the source rock burial history and subse- has been a long-standing issue due to limited outcrop exposure (Eyles
quently, its maturation and hydrocarbon generation potential. In the et al., 2001). Like the Canning Basin, one major exploration risk associ-
conventional realm, this is important in constraining the time of trap ated with global Palaeozoic systems (eg the North Caspian Basin,
formation relative to the time of hydrocarbon expulsion from source Kazakhstan; the Alberta Basin, Canada; and the Shetland Basin, UK) is the
rocks. In the unconventional reservoirs, on the other hand, deep burial timing of hydrocarbon charge, and subsequent preservation within the
and subsequent exhumation “excites” the kerogen, and brings the ther- complex stratigraphy and tectonic history (Ghori and Haines, 2006).
mally mature source rocks closer to the surface. Despite the significance of exhumation estimation to the conventional
The Ordovician Canning Basin, Western Australia has undergone at and unconventional hydrocarbon recovery, there is very limited litera-
least three major events of uplift and erosion during its geologic history. ture that discusses this in details for the Canning Basin. However, Duddy
This basin has been described to hold the largest amounts of et al. (2006) presented a report using Apatite Fission Track Analysis

* Corresponding author.
E-mail address: lukman.johnson@postgrad.curtin.edu.au (L.M. Johnson).

http://dx.doi.org/10.1016/j.petrol.2017.05.008
Received 18 August 2016; Received in revised form 1 May 2017; Accepted 11 May 2017
Available online 16 May 2017
0920-4105/© 2017 Elsevier B.V. All rights reserved.
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

(AFTA) to quantify the amounts of the removed section from the Triassic- The exhumation values obtained from this study is compared with
Jurassic and Eocene-Present day erosions respectively. This method exhumation values obtained from the thermal history dependent
works by providing information on the maximum paleotemperatures and approach (VR and AFTA) in order to validate or accurately constrain the
subsequent cooling that sediments have been exposed to over geologic obtained values.
time. This information is stored in the Apatite grains or Vitrinites and is
used to infer the possible amounts of exhumation. While this method is 2. Geologic setting
valid and in general practice in the industry, it is often desirable to obtain
exhumation magnitudes irrespective of burial temperatures, given the The Canning Basin, Western Australia occupies 506,000 km2 of which
susceptibility of thermal methods of estimation to occasional transient over 430,000 km2 is onshore (Fig. 1). The development of this basin
heating and hydrothermal flows in the basin. Although, hydrothermal initially began in the Early Palaeozoic as an intracratonic sag between the
flows haven't been reported in the Canning Basin, It is always a good Precambrian Pilbara and Kimberley Basins. This basin contains two
approach to utilize a number of independent, yet complimentary ap- major north-westerly trending troughs separated by a mid-basin arch,
proaches to constrain the thickness of exhumed sections in a sedimentary and marginal shelves (GWA, 2014). The several subdivisions that have
basins (Corcoran and Dore, 2005). Furthermore, the exposure of older been documented for this basin are based on the structural elements
Formations to higher temperatures as a result of deeper burial, leads to (mostly extension, subsidence, compression, and erosion) which have
annealing of the Apatite Fission Tracks, thereby, inaccurate paleo- been active at different times in the evolution of the basin. The defor-
temperature estimation, and by extrapolation, the magnitude of mation recorded in the southern Canning Basin is significantly lower
exhumation. than in the counterpart northern region. This is attributed to the absence
This study has attempted to document exhumation magnitudes from a of major fault block movements in the south (GWA, 2014).
portion of the Canning Basin based on the irreversible effect of burial on For this study, 4 wells in the Broome platform have been selected in
the physical properties of shales. This is well-documented in the sonic the first instance, along a northeast-southwest transect. The results is
logs, thereby, the sonic transit time of compacted shales is plotted against then further applied to other wells in the sub-basin.
depth. The sonic tool typically measures the present day transit time, Western Australia has been affected by a series of geodynamic events
which is an outcome of all stages of tectonic activities; and subsequently, since the Ordovician Period. The Canning Basin, in particular, was
the total compaction the sedimentary succession has been exposed to at affected by the early Ordovician extensional events that came about as a
the time of log acquisition. This approach provides a complimentary result of the breakup of micro-continents from the northeastern Gond-
exhumation estimation in the Canning Basin, as well as an improved wana. The dominant structure in the Canning Basin is the northwest
spatial resolution across the sub-basin relative to AFTA data availability. trending elevated Precambrian mid-basin platform. This separates the
In this study, four wells in the Broome Platform are studied, and subse- northern terraces from the southern Kidson and Willara sub-basins. In the
quently, the results from these wells are further applied to other wells northern Canning, it is referred to as the Broome platform, representing a
within the sub-basin. A method, first proposed by Athy (1930) has been horst between the Willara sub-basin and the Fitzroy Trough.
applied to several global basins and has proved to be effective in the Deposition in the Lower Paleozoic times took place across the entire
study areas. Magara (1976, 1986) also proposed an additional technique, basin until the Devonian, when this dominant broome swell developed.
still utilizing the shale compaction trends. This led to confinement of sediments into localized depressions in the

Fig. 1. Location of the study area. The green circles roughly indicate the Hilltop 1, Aquila 1, McLarty 1, and Kunzea 1 well locations (not to scale) (Cadman et al., 1993). (For interpretation
of the references to colour in this figure legend, the reader is referred to the web version of this article.)

20
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

Fig. 2. The general stratigraphy of the Canning Basin (GSWA, 2014).

21
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

south, and clastic sediments were deposited in the rapidly subsiding tectonically stable basin, with minor growth faulting at the Admiral Bay
graben (Seymour, 1972). As a consequence, structures in the southern Fault. Widespread deposition continued until the early Silurian, with salt
Canning Basin are less complex, with fewer periods of tectonism. Here, deposition in parts of the Canning. The Fitzroy Trough opened up with
the mid-basin platform is locally known as the Crossland Platform, with rifting and uplift of the platform in the Devonian (Giventian), with little
adjacent Kidson and Gregory sub-basins. From the eastern edge of the stabilization until the post-Triassic. Events included faulting and
Broome Platform, there are a number of downstepping fault-bounded compression in the upper Carboniferous time – before the phase of oil
sub-basins, including the Jugura Terrace, the Fitzroy Trough which generation in the Permian - late Triassic depositional phase (Horst-
then thins out towards the Lennard Shelf (with no Ordovician sedimen- man, 1984).
tation recorded). The down-thrown Willara sub-basin is separated from Veevers (1980) posited that the Canning Basin is an aulacogen, with
the Broome Platform at the Bay Fault System, and separated from the series of northwest trending half grabens related to south-dipping faults.
Fitzroy Trough by the Dampeir Fenton Fault (Taylor, 1992). The extensional event was followed by rapid subsidence, followed by a
During the Ordovician, marine sedimentation took place on the then sag stage, characterized by tilting, uplift, compression, and erosion, as
well as widespread evaporitic and playa conditions till the Silu-
rian Period.
Table 1
The second phase of evolution is characterized by minor folding and
Removed section estimates from the Acacia 1 and 2 wells, Canning Basin (Duddy et al.,
2006). regional uplift in the Early Devonian. By the Middle – Late Devonian,
extensional tectonics dominated, which led to the introduction of con-
Estimates of removed section (m)
tinental sediments in the northeastern part of the basin and in the
Triassic-Jurassic (230- L.Eocene - Present Day (40- southwest, restricted marine conditions prevailed. Following this phase,
170 Ma) 0) Ma
a major phase of compression and uplift took place across the basin until
Maximum Likelihood Estimate 1 300 1950 the mid-Carboniferous period, marked by inversion of the Devonian
Lower and upper 95% 800–2 800 400->10 000
faults, and deposition of syntectonic fluvial sediments.
confidence limits
Fixed paleo-geothermal gradients From Late Carboniferous, renewed extension and rapid subsidence
5  C/km not allowed >10 000 dominated, with widespread transgression in the Early Permian. This was
10  C/km not allowed 3 050–4 950 followed by a period of no deposition in most parts of the basin. Regional
15  C/km not allowed 1800–3 300 and dextral wrench movements in the Late Triassic – Early Jurassic led to
20  C/km not allowed 1 200–2 300
uplift and erosion of up to 3 km sediments. This was accompanied by the
25  C/km 2 650–2 950 900–1800

30 C/km 2000–2 300 700–1 300 deposition of fluvio-deltaic and marine sediments (Laurie and Stewart,
 2012). Some folding and faulting took place after the sea slowly retrea-
33.9 C/km -Present Day 1900 ± 200 800 ± 200
40  C/km 1 300–1700 450–650 ted, till the late Cretaceous times. From Cretaceous to recent, the basin
50  C/km 950–1 250 not allowed
has been land, except for the intermittent submerges of the present coast.
60  C/km 750–85 not allowed
65  C/km not allowed not allowed The general stratigraphy of the Canning Basin is summarised in Fig. 2.

Fig. 3. Standard compaction curve (Jankowsky, 1962).

22
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

Fig. 4. Gamma-ray and sonic logs (A) and the extrapolated ΔT shale values to the surface (B) (Poelchau, 2001).

3. Previous work from Apatite Fission Track Analysis 1 Present day geothermal gradients supplied from the Canning Basin
wells Bottom Hole Temperatures (BHT) are too high. Generally, these
Duddy et al. (2006) analysed some Canning Basin wells for qualitative BHT's were obtained from log headers, where they have been cor-
thermal histories, including the timing of thermal episodes from Apatite rected using a simplified correction procedure. These were further
Fission Track Analysis (AFTA) and Vitrinite Reflectance (VR) study. The corrected using AFTA steady state geothermal gradient.
study concluded with the following assertions. 2 AFTA and VR data showed that pre-Jurassic sequences cooled from
maximum paleo-temperatures in the Triassic, thereby responsible for

23
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

Fig. 5. Interpolation of the sonic data from Hilltop 1 well on the standard compaction
curve.

the cessation of active source rock maturation in the pre-Jurassic


sequences.
3 AFTA and VR data also reveal two subsequent regional heating and
cooling episodes in the Cretaceous and Tertiary.
4 The geothermal gradient in the basin has been consistent with the
present-day heat flow with no sign of elevated basal heat flow during
the exhumation period. This suggests that heating in each event was
as a result of sediment loading, with associated cooling during uplift
and erosion.

From Duddy et al. (2006), quantitative estimates of removed sections


in the Acacia#1 and Acacia #2 wells show approximately 1 300 and
2000 m of eroded sections in the Triassic and Eocene sediments respec-
tively. This is summarised in Table 1. It should be noted that this study
indicates significant spatial variability in confidence limits of the
magnitude of exhumation.

4. Estimation of the amount of erosion

4.1. Compaction trends from sonic logs

The reduction in sedimentary thicknesses as a result of successive


burial is known as compaction. Here, we assume a vertical principal
stress for successive burial. Sonic logs are used as a proxy for compaction
because sonic interval travel time depends on porosity (Hillis et al.,
1994). The sonic curve represents the imprint of different stages of uplift
and erosion that is stored in the basins stratigraphic record. One Fig. 6. The plot of sonic transit time versus depth showing the trend for uncompacted
shales of the Hilltop#1 well. A Gamma ray cut off value of 130API was used to filter for
advantage of this method is the wide availability of sonic logs, as they
shale values.
have become routinely acquired for formation evaluation. This allows for
a greater spatial resolution across the basin compared to the limited data
from AFT studies (Burns et al., 2005). extrapolation of the filtered sonic transit data to this curve. The mid-
Quantifying the amount of exhumation with sonic logs was carried point (mean) value represents difference in depth or the observed
out in two ways. vertical displacement of the present day sonic transit time values from
the standard compaction curve is measured and represents an esti-
1 Obtaining a representative standard compaction curve: The mate the average exhumation thickness (Ware and Turner, 2002) for
compaction curve commonly defines the sedimentary succession in a a particular horizon. This is represented by equation (1) below
basin at its maximum burial depth (Tassone et al., 2014a,b). In this (Tassone et al., 2014a,b)
study, the scheme of Heasler and Kharitonova (1996) and Magara
(1976) was modified to obtain a best fit transit time for the Canning ET ¼ BmaxBpresent-day (1)
Basin, using a polynomial function of the sonic transit time with whereET is thickness of eroded section, Bmax is maximum burial depth
depth. Here, we consider the boundary condition of shales at zero from the compaction curve, and Bpresent-day is the mean present day
depth as 200 μs/ft, which also shows a decreased porosity as well as burial depth.
sonic transit time with depth (Fig. 3). With the gamma ray log, sand and shale units are differentiated. The
2 The reduction in porosity is as a result of normal mechanical and sand intervals are eliminated while the shale values of the corresponding
thermochemical compaction (Japsen et al., 2007). Data from studied sonic log interval is used for extrapolation. The resultant shale interval is
wells is then compared with the standard compaction curve by

24
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

the representative shale at the maximum burial depth at a specific 5.2. Extrapolation from Vitrinite reflectance trends
stratigraphic unit. As shown in Fig. 3, the shales are projected on the
compaction curve, and the difference in depth with the curve represents The plot of Vitrinite Reflectance (and its equivalent) shows a
the uplifted values. In this study, the present day depths of the resultant consistent increase with depth from Recent to Permian deposits. From
extrapolated shales is compared with the Formation tops and biostrati- the formation tops, there is a sharp increase in VR that suggests a major
graphic data of each wells to obtain the relative timing of exhumation. hiatus between the Permian Grant Formation and the Ordovician Gold-
This data was obtained from the Department of Mines biostratigraphic wyer Formation (Fig. 7).
and well completion reports. The original profile is projected downwards, showing a steady in-
crease in the maturation profile of the subsiding sediments with depth.
3 Cross-plotting the sonic transit travel time (ΔT) of shales nearest to The major unconformity is observed as a result of a discontinuity in the
the surface against depth (Fig. 4). The resultant trend line is subse- original profile, with an offset range between 700 m and 1600 m, with a
quently extrapolated to the surface value of uncompacted shale more probable estimate of ~1000 m. The discontinuity shows that the
(200 μs/ft) sediments have been exposed to different thermal regimes, and suggests
that there was an uplift between the Permian Grant formation and the
In extracting the representative shaly sections, the gamma ray log was Ordovician Goldwyer Formation. The profile also suggests that the older
used as a filter for the shales and non-shale sections. A cut-off value of vitrinites must have been exposed to extreme temperatures over time,
130API was applied for the shales. prior to the uplift, thereby, in the kinetics of the system, most of the
vitrinites have been annealed (Fig. 7). As described by Dembicki Jr.
(1984), variation in organic provenance or maceral misidentification is
4.2. Extrapolation from Vitrinite reflectance trends one of the limitations to the Vitrininite Reflectance methods of erosion
estimation. This could in part be responsible for the anomalous increase
Vitrinite reflectance (VR) data were obtained from the Department of in Ro(%) over sediment thickness <100 m (Fig. 7). However, compared
Mines and Petroleum reports as well as from converted Tmax data. The to the AFT approach and the sonic log approach, exhumation estimates
VR values were plotted against depth, and the offset from the normal from vitrinite reflectance yields results that are markedly inconsistent
trend is measured as the amount of exhumation (Dow, 1977). with the results from the aforementioned. Here, some part of the basins
evolution have been not accounted for (Fig. 7). This may in part be
attributed to limited sample size, sample preparation methods and vit-
5. Results
rinite maceral misidentification.
5.1. Compaction trends from sonic logs
6. The new approach
In order to compute the magnitude of erosion from sonic logs, the
polynomial function of the sonic transit time against depth is obtained In an attempt to establish an empirical relationship between the
(Fig. 5). This represents the normal compaction trend in the basin, similar magnitudes of erosion and ΔT with respect to present day depths for
to that proposed by Jankowsky (1962); Magara (1976). Utilizing equa- compacted shales within the Canning Basin, several ΔT and observed
tion (1), the subsequent superposition of sonic data on the normal displacement values was computed and utilized to derive a regression
compaction trend (NCT) curve reveals a mean offset of 1600 m and equation, which is as follows:
1400 m of exhumation in both the Triassic-Jurassic unconformity and
Thickness of eroded section (TES) ¼ 5 417–32.86ΔT 0.844depth; (2)
Silurian-Carboniferous unconformity for the Hilltop 1 well. This is the
offset of the shales from the NCT (Fig. 5). This value, when compared to with a correlation coefficient of R ¼ 0.86.where
other values within the basin is close to the values reported from the TES is Thickness of eroded section/displacement, ΔT is sonic transit
AFTA analysis. However, this haven't accounted for the Eocene uncon- time for shale, and depth is the corresponding depth for the
formity. Also, the Silurian-Carboniferous erosional event was not shale interval.
captured in the Geotrack report by Duddy et al. (2006), probably as a The equation is synonymous to the general regression equation:
result of the exposure of the older Formations to higher temperatures and
subsequently, annealed fission tracks. Y ¼ β0 þ β1 X1 þ β2 X2 þ …βn Xn þ ε (3)
Furthermore, in the relatively uncompacted Eocene sediments, a
Therefore, TES ¼ y ¼ the dependent variable.
method proposed by Jankowsky (1962); Magara (1976) was employed
to account for the exhumation magnitudes. The extrapolation of the
normal compaction trend of these near-surface sediments to the surface
transit time value (ΔTo) of 200 μs/ft shows 800 m of exhumed sec-
tion (Fig. 6).
From these, it is estimated that a net of at least 3000 m of sediments
has been removed from the succession in this part of the Canning Basin.
The method was repeated for three (3) other wells along a rough
northeast-southwest transect as shown in Fig. 1, and the results are
shown in Table 2.

Table 2
Erosion estimation from 4 (four) Broome Platform wells, Canning Basin.

Well Name Sections

Eocene Triassic-Jurassic Permian-Carboniferous

Hilltop 1 400–800 (600) 400–2 100 (1 600) 500–2000 (1 400)


Aquila 1 600–1 200 (800) 800–1900 (1 400) 500–2 200 (1 300)
McLarty 1 500–800 (600) 200–1 200 (800) 1 000–2000 (1700)
Fig. 7. Vitrinite reflectance vs depth plot for Hilltop#1, showing a discontinuity in the
Kunzea 1 150–300 (300) 400–800 (500) 200–1800 (900)
maturation profile of the sediment, suggestive of an uplift.

25
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

Table 3 Table 4
Data from different intervals showing the sonic transit time (ΔT) and the corresponding Erosion estimation from 8 Broome Platform wells.
depth, the visual observed displacement (Jankowsky, 1962), and the calculated displace-
Well Name Sections
ment using equation (1).
Eocene Triassic-Jurassic Permian-Carboniferous
Depth (m) ΔT (μs/ft) Observed TES (m) Calculated TES (m)
Hilltop 1 400-800 (600) 400-2 100 (1 600) 500-2000 (1 400)
230 120 1 250 1 279.7
Aquila 1 600-1 200 (800) 800-1900 (1 400) 500-2 200 (1 300)
529 114 1 350 1 224.5
McLarty 1 500-800 (600) 200-1 200 (800) 1 000-2000 (1700)
268 134 930 787.6
Kunzea 1 150-300 (300) 400-800 (500) 200-1800 (900)
1 150 90 1 650 1 489.0
Musca 1 300-700 (400) 600-2000 (1 200) 300-1 500 (1 200)
268 120 1 250 1 247.6
Matches Springs 1 400-800 (600) 400-1 300 (1 000) 300-900 (600)
321 112 1 400 1 465.8
Santalum 1 150-400 (300) 900-2000 (1800) No Records
774 107 1 226 1 247.7
Edgar Range 1 400-800 (600) 300-2 100 (1 000) 200-1700 (1 400)
250 125 1 150 1 098.5
754 107 1 050 1 264.6
800 95 1 600 1 620.1
577 130 550 658.2 The contour maps indicate that, Tertiary Period, and the Silurian-
Carboniferous Period, the magnitude of erosion in the northwest quad-
rant of the Broome Platform was relatively higher than it was in the
Southwest quadrant towards the Kidson sub-basin. However, in the
Triassic-Jurassic Period, erosional activities was relatively uniform
across the sub-basin. This period is associated with the Fitzroy trans-
pression that took place across the Canning Basin.

7. Discussion and conclusion

In sedimentary basins, accurate delineation of the thickness of


exhumed sections is important in the establishment of the basins subsi-
dence history. This directly imparts the maximum temperatures that a
Formation has been exposed to over geologic time, and this could have
both positive and negative impacts on organic matter maturation, trap
formation and ultimately, hydrocarbon prospectivity in the basin. In this
study, gamma ray logs, sonic transit time logs and formation tops and
biostratigraphic data have been employed for this purpose. After
extraction of the representative shale layers for different formations, the
Fig. 8. A plot showing TES for the Acacia#1. This further validates the equation; at 250 m
depth for the Triassic-Jurassic sediments, and ΔT of 120 μs/ft. Calculated TES in this case
deviation or mean vertical displacement of the present day sonic transit
is 1262 m, while the observed displacement value is 1250 m. AFTA reports also indicate a time values from the standard compaction curve is measured and rep-
value of 1300 m. resents an estimate the average amount of exhumed sections in the well
locations. This was repeated for several wells in the Broome Platform,
5 417 ¼ β0 ¼ intercept Canning Basin. Of the several porosity logs, sonic logs are preferred
() 32.86 ¼ β1 ¼ regression model coefficient because they display relatively simple normal compaction trends (Poel-
() 0.844 ¼ β2 ¼ regression model coefficient chau, 2001; Tassone et al., 2014a,b). Furthermore, in this study, a new
ΔT ¼ X1 ¼ an independent variable equation have been derived, which forms an empirical relationship be-
Depth ¼ X2 ¼ an independent variable tween shale DT values, depth and the magnitude of erosion. This equa-
Е ¼ Error Term tion appears to be in good agreement with both the measured AFTA data
and the observed exhumation from several wells in the Canning Basin.
This equation is basin specific, and has been tested on several in- The study reveals an estimate of at least 3 000 m of exhumed sections
tervals across the basin (Table 3) as well as the Triassic – Jurassic in the basin.
exhumation values as reported by Duddy et al. (2006) on the Acacia 1 In the studied basin, exhumation data from 8 wells are computed and
well in the Barbwire Terrace (Fig. 8). The observed TES is in agreement contoured. These contours show that exhumation has been relatively
with the measured data from AFTA studies (Table 1), and also have a uneven across the basin. The results suggest that for each tectonic
good correlation with the calculated exhumation values, using equation episode, deformation and erosion is less intense onshore – towards the
(3) above. Also, further validation is carried out on the studied data as Kidson sub-basin, and progressively increases in the offshore direction.
well as data across the basin (Table 3). However, in the deeper parts of Several methods have been proposed to account for the estimation of
the basin, where the compaction curve/porosity approaches zero, esti- thickness of eroded sections in a sedimentary basin, and each have their
mation of exhumed sections in that section would have a shallow slope own limitations. While AFTA and vitrinite reflectance methods are
and projection of the sonic data on the curve would yield a high, inac- routinely a standard practice in exploration, the yielded erosion esti-
curate erosion estimate. Here, the equation (TES) ¼ 5 417–32.86 (ΔT) - mates may be inaccurate (Burns et al., 2005) as these methods often
0.844depth is used to calculate the thickness of exhumed sections. require assumed values for the basins geothermal gradient and thermal
With a p-value of less than 0.05 and calculated erosion values (Y) that conductivities. This is more so significant in basins that have been
does not differ significantly (in exhumation magnitudes, ±150 m) of the affected by heating related to fluid flow or volcanic activities. Therefore,
observed TES, we consider that depth and ΔT are the two main a major advantage the compaction based method for exhumation esti-
contributing factors that influence (Y). Therefore, the error term is mation is that the estimates are independent of the basins thermal his-
assumed to be zero (0) in this study. tory. The decrease in sonic transit travel and porosity loss is with depth is
The introduced equation has further been used to calculate erosion usually associated burial depths. Also, sonic logs are routinely acquired
estimates from 4 additional Broome Platform wells (Table 4), and the during drilling therefore, the wide availability of sonic logs relative to
magnitudes of erosion across the sub-basin is represented in the contour AFTA and VR data makes the process relatively inexpensive. Thus,
map shown in Fig. 9a–c. providing cheap and more accurate input into the basins thermal and
burial history model. Furthermore, from this study, the exhumation

26
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

Fig. 9. a–c. Contour maps showing the magnitudes of erosion from the northwest to the south eastern parts of the Broome Platform. Data from 8 wells.

27
L.M. Johnson et al. Journal of Petroleum Science and Engineering 156 (2017) 19–28

values reported from AFTA usually allows for a wide range of uncer- Ghori, K.A.R., Haines, P.W., 2006. Paleozoic Petroleum Systems of the Canning Basin,
Western Australia: a Review. AAPG (Perth, Australia, Search and Discovery Article).
tainty, while calculation of exhumation values from the new approach
GSWA, 2014. In: Geological Survey of Western Australia (Ed.), Petroleum Prospectivity of
gives an absolute value. However, a limitation to the compaction based State Acreage Release Area L14-2, Canning Basin, Western Australia, p. 17.
method of exhumation estimation is that the absolute timing of the GWA, 2014. In: DMP (Ed.), Western Australia's Petroleum and Geothermal Explorer's
erosional or cooling event cannot be independently constrained (Tassone Guide. Government of Western Australia, Perth.
Heasler, H.P., Kharitonova, N.A., 1996. Analysis of Sonic Well Logs Applied to Erosion
et al., 2014a,b). Also, change in rock composition is often overlooked as it Estimates in the Bighorn Basin, Wyoming: AAPG Bulletin, vol. 80, pp. 630–646.
may be difficult to differentiate radiogenic sands from shales. Hillis, R.R., Thomson, K., Underhill, J.R., 1994. Quantification of tertiary erosion in the
However, since these studies are usually prone to high levels of un- Inner Moray Firth using sonic velocity data from the Chalk and the Kimmeridge Clay.
Mar. Petrol. Geol. 11, 283–293.
certainties, to accurately constrain the thickness of exhumed sections, a Horstman, E.L., 1984. Source rocks in the Canning Basin: a review. In: Purcell, P.G. (Ed.),
combination of different approaches is usually recommended (Corcoran Proceedings of the Geological Society of Australia. Petroleum Exploration Society of
and Dore, 2005; Tassone et al., 2014a,b). In this study, both the Australia, Perth.

Jankowsky, W., 1962. Diagenese und Olinhalt als Hilfsmittel für die
compaction based estimate and thermal based estimation of exhumed strukturgeschichtliche Analyse des Nordwestdeutschen Beckens. Z. Dtsch. Geol. Ges.
sections have been considered. At this time, the access to seismic data has 452–460.
proved abortive, thereby, further validation by obtaining continuous Japsen, P., Mukerji, T., Mavko, G., 2007. Constraints on Velocity-depth Trends from Rock
Physics Models: Geophysical Prospecting, vol. 55, pp. 135–154.
information using seismic cross section is not possible. Laurie, J.R., Stewart, A.J., 2012. Canning Basin, Australian Capital Territory, Geoscience
Australia.
References Magara, K., 1976. Thickness of removed sedimentary rocks, paleopore pressure, and
paleotemperature, southwestern part of Western Canada Basin. AAPG Bull. 60,
554–565.
Athy, L.F., 1930. AAPG Bulletin. Density, Porosity, and Compaction of Sedimentary
Magara, K., 1986. Geological Models of Petroleum Entrapment. Elsevier Applied Science
Rocks, vol. 14, pp. 1–24.
Publishers, London.
Burns, W.M., Hayba, D.O., Rowan, E.L., Houseknecht, D.W., 2005. Estimating the Amount
Poelchau, H.S., 2001. Modeling an exhumed Basin: a method for estimating eroded
of Eroded Section in a Partially Exhumed Basin from Geophysical Well Logs: an
overburden. Nat. Resour. Res. 10, 73–84.
Example from the North Slope: Studies by the US Geological Survey in Alaska: US
Seymour, M.D., 1972. In: DMP (Ed.), Canning Basin Regional Report. B.O.C. of Australia
Geological Survey, Special Paper, pp. 1–18.
Limited, Perth, Australia, p. 71.
Cadman, S.J., Pain, L., Vuckovic, V., Le Poidevin, S.R., 1993. Canning Basin. Australian
Tassone, D.R., Holford, S.P., Stoker, M.S., Green, P., Johnson, H., Underhill, J.R.,
Petroleum Accumulations, Canberra, ACT. W.A., in B. o. R. Sciences.
Hillis, R.R., 2014a. Constraining Cenozoic exhumation in the Faroe-Shetland region
Carlsen, G., Ghori, K., 2005. Canning Basin and global Palaeozoic petroleum systems — a
using sonic transit time data. Basin Res. 26, 38–72.
review. APPEA J. 45, 349–364.
Tassone, D.R., Holford, S.P., Duddy, I.R., Green, P.F., Hillis, R.R., 2014b. Quantifying
Corcoran, D., Dore, A., 2005. A review of techniques for the estimation of magnitude and
Cretaceous–Cenozoic exhumation in the Otway Basin, southeastern Australia, using
timing of exhumation in offshore basins. Earth Sci. Rev. 72, 129–168.
sonic transit time data: Implications for conventional and unconventional
Dembicki Jr., H., 1984. 1984, an interlaboratory comparison of source rock data.
hydrocarbon prospectivity. AAPG Bull. 98, 67–117.
Geochimica Cosmochimica Acta 48, 2641–2649.
Taylor, D., 1992. In: DMP (Ed.), A Review of Ordovician Source Rocks, Canning Basin,
Dow, W.G., 1977. Kerogen studies and geological interpretations. J. Geochem. Explor. 7,
Western Australia. Bureau Mineral Resources, Geology and Geophysics, Australia.
79–99.
Triche, N.E., Bahar, M., 2013. Shale Gas Volumetrics of Unconventional Resource Plays in
Duddy, I.R., Moore, M.E., O'Brien, C., 2006. Thermal History Reconstruction in Five
the Canning Basin, Western Australia, SPE, Brisbane (Australia, SPE).
Canning Basin Wells: Acacia-1 &-2, Kidson-1, Willara-1 & Yulleroo-1 Based on
Veevers, J.J., 1980. Basins of the Australian craton and margin. In: Bally, A.W. (Ed.),
Apatite Fission Track Analysis (AFTA®) and Vitrinite Reflectance Data, Geotrack
Dynamics Ofplate Interiors: Geodynamics, American Geophyscial Union, pp. 73–80.
Report. Victoria, Australia p. 28.
Ware, P.D., Turner, J.P., 2002. Sonic Velocity Analysis of the Tertiary Denudation of the
Eyles, N., Eyles, C.H., Apak, S.N., Carlsen, G.M., 2001. Permian-Carboniferous tectono-
Irish Sea Basin, vol. 196, pp. 355–370. Geological Society, London, Special
stratigraphic evolution and petroleum potential of the northern Canning Basin,
Publications.
Western Australia. AAPG Bull. 85, 989–1006.

28

View publication stats

Potrebbero piacerti anche