You are on page 1of 17

Importance of pre-existing fault size for the evolution of

an inverted fault system


CATHAL REILLY1,2,3*, ANDREW NICOL1,2,4 & JOHN WALSH2
1
GNS Science, PO Box 30368, Lower Hutt 5040, New Zealand
2
Fault Analysis Group, School of Geological Sciences, University College Dublin,
Belfield, Dublin 4, Ireland
3
Present address: Midland Valley, Floor 9, 2 West Regent Street, Glasgow G2 1RW, UK
4
Present address: Department of Geological Sciences, University of Canterbury,
Private Bag 4800, Christchurch, New Zealand
*Corresponding author (e-mail: cathal@mve.com)

Abstract: Fault inversion has been documented in many basins worldwide, yet the details of how
the initial extensional faults impact on the geometry and growth of the reactivated contractional
system is often poorly resolved by the available data. Two-dimensional (2D) and 3D seismic reflec-
tion, and well data have been used to chart the evolution of inverted faults from the Taranaki Basin,
offshore New Zealand. Sedimentary rocks up to 8 km thick record Late Cretaceous–Paleocene
normal faults inverted during Miocene and younger shortening. The displacement and length of
early normal faults is a key determinant for the reactivation and size of the subsequent reverse
faults. All normal faults with maximum vertical displacements ≥600 m and lengths ≥9 km
were inverted along their entire length, while smaller faults were not inverted. The proportion
of the total basin-wide strain accommodated on each fault is comparable between deformational
episodes. The hierarchy of reverse fault lengths was established rapidly, with longer faults accruing
a greater proportion of the total strain from an early stage of shortening. The reverse fault system
is dominated by inverted normal faults, which accrue displacement at the expense of smaller faults,
and utilize the largest crustal-scale elements of the pre-existing system. The size of pre-existing
heterogeneities is an important control for the magnitude and spatial extent of elevated stresses
during contraction, which, in turn, control the dimensions, locations and displacements of
subsequent fault growth.

Fault inversion (Fig. 1) occurs globally in a variety constrain the geometrical and kinematic relation-
of sedimentary basin settings, including subduction ships between extensional and contractional fault
margins, orogenic foredeeps, intracratonic basins systems (e.g. Davis 1983; Badley et al. 1989; Bishop
and ‘peri-orogenic’ regions of continental escape & Buchanan 1995; Wang et al. 1995; Bulnes &
tectonics (Turner & Williams 2004 and references McClay 1998; Withjack et al. 2010; Grimaldi &
therein). Basin inversion has been recognized for Dorobek 2011; Jackson et al. 2013). Such seismic
almost a century (Lamplugh 1920; Stille 1924) data are utilized in this paper to investigate the influ-
and is typically achieved by fault inversion with ence of early normal faults on the evolution of
early normal faults reactivated during later contrac- an inverted fault system in the offshore Taranaki
tion (Glennie & Boegner 1981; Hayward & Graham Basin, New Zealand.
1989; Williams et al. 1989; Bishop & Buchanan Fault inversion is selective and often only a small
1995; Brun & Nalpas 1996; Turner & Williams percentage (,10%) of the initial normal fault sys-
2004). Hanging-wall anticlines formed in associa- tem are reactivated (e.g. Davis 1983; Badley et al.
tion with inversion produce structural culminations 1989; Williams et al. 1989; Sibson 1995; Kelly
that, in many basins, contain important hydro- et al. 1999; Panien et al. 2005). A number of fac-
carbon accumulations (Fig. 1) (e.g. Fraser & Gaw- tors may control which faults are reactivated dur-
thorpe 1990; Uliana et al. 1995; Gluyas et al. 2003; ing inversion, including: the relative orientation of
Madon et al. 2004; Warren 2009). Hydrocarbon the pre-existing faults and the principal horizon-
exploration of inversion structures has produced tal shortening direction during inversion (Brun &
a large body of seismic reflection data that help Nalpas 1996; Panien et al. 2005; Del Ventisette

From: Childs, C., Holdsworth, R. E., Jackson, C. A.-L., Manzocchi, T., Walsh, J. J. & Yielding, G. (eds)
The Geometry and Growth of Normal Faults. Geological Society, London, Special Publications, 439,
http://doi.org/10.1144/SP439.2
# 2016 The Author(s). Published by The Geological Society of London. All rights reserved.
For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics
C. REILLY ET AL.

(a) Normal Fault


Post-Inversion
sedimentation
Syn-Inversion
sedimentation
Vertical
displacement Post-Rift
(Vd) sedimentation
Syn-Rift
sedimentation

(b) Fault and basin inversion Basement

Vertical
separation
(Vs)

Inverted
displacement
(Id)

Id=Vd-Vs

Fig. 1. Schematic cross-section of an inverted fault and basin. Growth strata deposited during extension (green) and
contraction (white) are separated by intervals of tectonic quiescence (grey, blue and yellow units) (modified from
Williams et al. 1989). Displacements measured in this paper are also shown.

et al. 2006; Withjack et al. 2010); fault dip, with Taranaki Basin in offshore New Zealand. Within
shallower dips more conducive to inversion (Sibson this portion of the basin, normal faults of Late
1995; Kelly et al. 1999); fluid overpressure, which Cretaceous –Paleocene age (c. 80 –55 Ma) were in-
promotes reduction of the effective normal stress verted during Miocene and younger shortening
on the fault surface (Sibson 1995; Turner & Wil- (Figs 2b, c & 3). The strike and dip of the early, nor-
liams 2004); and the size of original normal faults, mal faults are approximately uniform (NE–SW
with larger faults more likely to invert (e.g. Eisen- and 55 –708, respectively), while the extension and
stadt & Withjack 1995; Kelly et al. 1999; Panien shortening directions during respective deformation
et al. 2005). Whatever the precise mechanism con- episodes were approximately coaxial and trend per-
trolling which faults reactivate, it is clear that inher- pendicular to fault strike (King & Thrasher 1996;
itance from pre-existing structures can profoundly Reilly et al. 2015). Therefore, neither changes in
influence the evolution of later fault systems (Chad- the obliquity of slip nor variations in fault dip are
wick & Evans 1995; Morley 1995; Lezzar et al. likely to control which structures were inverted.
2002; Walsh et al. 2003; Morley et al. 2004; Wor- However, the Late Cretaceous – Paleocene normal
thington & Walsh 2011). In addition to controlling faults have a range of sizes, with vertical displace-
the orientation, location and size of inverted faults, ments of 10 m –3.5 km resolved on 2D and 3D seis-
early normal faults may promote rapid establish- mic reflection lines (Fig. 3). The wide range of fault
ment of final fault lengths and strain localization sizes (i.e. three orders of magnitude) enables the
onto a small number of primary structures (Morley dependence of inversion on fault size to be exam-
1999; Meyer et al. 2002; Childs et al. 2003; Walsh ined (Figs 4 & 5). The seismic data also provide a
et al. 2002; Giba et al. 2012). rare 3D perspective of the geometrical and kine-
The role of fault size (i.e. length and maximum matic evolution of inverted normal faults. These
displacement) in selective inversion, and the impact data show a clear control of fault size on inver-
that reactivation has on fault growth, is considered sion, and suggest that size, spatial distribution and
here using seismic reflection data from the southern relative fault sizes in the early extension system
SIZE CONTROLS ON INVERSION

pre-condition the geometry and kinematics of the sedimentary succession, with a total thickness of
inverted fault array. approximately 800 m –8 km, unconformably over-
lying Mesozoic basement (Fig. 2). Within the sedi-
mentary sequence, up to 16 horizons were mapped
Geological setting on 916 2D seismic reflection lines (with a total
length of almost 25 000 km) and within five 3D seis-
The Taranaki Basin has experienced both extension
mic reflection surveys (with a total area of almost
and shortening since about 80 Ma (e.g. Figs 2 & 3)
3500 km2) (Reilly et al. 2015). The interpreted
(King & Thrasher 1996). Normal faults develo-
seismic horizons are approximately isochronous,
ped throughout the Taranaki Basin during the Late
with ages estimated using biostratigraphy from
Cretaceous–Paleocene (c. 80–55 Ma) and reflect
35 wells (Roncaglia et al. 2008, 2010, 2013; Mor-
crustal extension during Gondwana break-up
gans 2013). The interpreted horizons have a tempo-
(Thrasher 1990; King & Thrasher 1996; Hemmings-
ral resolution of c. 1–4 myr since 24 Ma and c. 5–
Sykes 2011). These normal faults are commonly
10 myr prior to the Miocene. Approximately 1000
the oldest structures that displace sedimentary strata
faults with vertical displacements of around 15 –
(Figs 2 & 3). They produced a series of half-graben
2200 ms (10 m –3.5 km) have been mapped in the
sedimentary basins that contain strata up to 4 km
southern Taranaki Basin. The basin hosts normal
thick. The normal faults mainly strike north–south
faults that formed during Late Cretaceous and
and NE–SW with dips of 55 –708, suggesting a
Paleocene rifting, reverse faults that formed during
principal horizontal extension direction of approxi-
the Late Eocene –Miocene, and Plio-Pleistocene
mately NW–SE. Their displacements range from
(in the south of the southern Taranaki Basin) con-
,10 m to large, basin-bounding faults (.3.5 km)
traction. In addition, the north part of the southern
and have maximum lengths of up to around 90 km
Taranaki Basin hosts normal faults that formed dur-
(Figs 2–5).
ing Plio-Pleistocene extension (Fig. 6) (Pilaar &
Contraction and associated inversion in the
Wakefield 1978; Knox 1982; Schmidt & Robinson
basin commenced no later than Early Oligocene
1989; King 1990; Hoolihan & Yang 1991; Thrasher
and migrated westwards across the basin during
1992; Palmer & Andrews 1993; Holt & Stern 1994;
the Miocene (Pilaar & Wakefield 1978; Knox
King & Thrasher 1996; Nicol et al. 2005; Stagpoole
1982; Schmidt & Robinson 1989; King & Thrasher
& Nicol 2008; Giba et al. 2010; Reilly et al. 2015).
1996; Reilly et al. 2015). In the SW of the basin,
A small proportion (,10%) of the earliest formed
some reverse faults have been active during the
normal faults was inverted during later contraction,
Quaternary (Fig. 6). Shortening in the basin is
which also resulted in the formation of new reverse
inferred to have been driven by convergence along
faults. Examination of individual seismic lines sug-
the Australian–Pacific plate boundary through New
gests that inversion is mainly limited to the largest
Zealand (Ballance 1976; Walcott 1978; Holt &
displacement faults in the basin (Figs 2 & 3).
Stern 1994; Stagpoole & Nicol 2008; Reilly et al.
Displacement measurements and displacement
2015). The resulting reverse faults are approxima-
back-stripping were used to determine the kinema-
tely dip-slip, with an inferred principal horizon-
tic history of 21 faults, with maximum vertical
tal shortening direction of approximately NW–SE,
displacements of ≥500 m. To measure both fault
and are commonly associated with asymmetric
displacement (fault-surface slip) and fault-related
footwall-verging folds (i.e. hanging-wall anticlines
folding, vertical separations were recorded between
and footwall synclines) (Figs 2 & 3).
footwall and hanging-wall fold hinges. Vertical sep-
aration of the hinges is considered to be a proxy for
Dataset and methods fault vertical displacement, and here the term dis-
placement refers to vertical displacement and/or
2D and 3D seismic reflection surveys tied to petro- vertical separation. In cases where the anticlinal
leum exploration wells have been utilized to esti- hinge had been eroded, the eroded surfaces have
mate the displacement histories of faults and the been projected to the hanging-wall anticline hinge
influence of fault size on subsequent inversion (e.g. to enable an estimation of vertical separation. Dis-
Fig. 3). These seismic data provide information on placement back-stripping was performed on indi-
the style of deformation, while displacement back- vidual seismic sections by sequentially subtracting
stripping using seismic reflection lines enables the displacements on progressively older horizons from
timing of fault growth to be estimated. all underlying horizons (for further discussion of
The evolution of the faults is recorded by syn- the methodology, refer to: Petersen et al. 1992;
faulting growth strata and by up-sequence changes Childs et al. 1993, 2003; Clausen et al. 1994). The
in the magnitude and style (normal and reverse back-stripping method has been modified slightly
dip slip) of displacements (e.g. Fig. 1). Faulted to account for inverted faults with early normal
strata comprise a Late Cretaceous and younger and late reverse displacements (and folding). Taking
C. REILLY ET AL.
Fig. 2.
SIZE CONTROLS ON INVERSION
Fig. 2. (Continued) Faults and structural setting of the southern Taranaki Basin. (a) Plate boundary setting and study area location. New Zealand Pacific Plate relative motion
vectors relative to the Australian Plate are from Beavan & Haines (2001). (b) Map of main faults in the southern Taranaki Basin active since approximately 80 Ma. Triangles
and ticks on faults indicate dip directions of present reverse and normal faults, respectively. Thick red lines across northern and southern parts of the basin, a– a′ and b– b′ ,
show the locations of the regional cross-sections in (c) and (d). (c) & (d) Cross-sections a–a′ and b –b′ across the NE and SW of the southern Taranaki Basin, respectively.
Individual seismic line names that make up the cross-sections and extents on the sections are shown in black above each section. Sediment thickness variations indicate the
timing and sense of movement on large faults in the basin (e.g. Fig. 1). Note also the death of small normal faults, terminating in the Late Cretaceous–Paleocene sequence
and not undergoing later reactivation.
C. REILLY ET AL.
Fig. 3. Example of (a) an uninterpreted and (b) an interpreted seismic reflection line showing inverted and non-inverted normal faults, and new reverse faults (labelled in b).
The location of the seismic line is shown in Figure 2b. Schematic tectonic history outlines Cretaceous– Paleocene rifting, Eocene quiescence, Oligocene –Recent contraction
and Plio-Pleistocene rifting in the north of the study area. TTWT, two-way transit time.
SIZE CONTROLS ON INVERSION

Fig. 4. Maximum displacement v. length plots for (a) normal and (b) reverse faults in the southern Taranaki Basin.
Inverted (filled circles) and non-inverted normal faults (a) and newly formed reverse fault (b) (open circles) are
differentiated in each plot. Inverted faults represent the largest approximately 10% of the original normal faults and
have maximum vertical displacements and trace lengths of approximately .600 m and .9 km, respectively.
Displacements were measured in milliseconds and converted to metres using the depth-conversion method of
Thrasher & Cahill (1990), with more recently calculated velocity values for southern Taranaki Basin intervals (Hill
& Milner 2012). Global fault data (filled small grey circles) are compiled from the literature (e.g. Schlishe et al.
1996; Bailey et al. 2005) augmented by recent unpublished compilations.

measured reverse vertical displacements as negative compaction is usually ,20% (Taylor et al. 2008)
values and subtracting them from the apparent dis- and does not impact our first-order conclusions.
placements on the oldest syn-rift strata in the same In addition to back-stripping, the maximum dis-
cross-section allows the original normal displace- placement and lengths of 165 faults were measured
ment associated with the extensional event to be for one or both of the deformation episodes, and are
quantified (Fig. 1). For 14 of the 21 faults, multiple shown in Figure 4. Collectively, these data provide
seismic sections, spaced at 0.5–2 km, were back- information on the evolution of the fault systems
stripped to produce along-strike displacement pro- and the role that fault size plays in inversion.
files that record displacements for extensional and
contractional phases of faulting (e.g. Fig. 7) (for Fault size and inversion
further discussion of the back-stripping technique
as applied to inverted faults refer to Reilly et al. In the southern Taranaki Basin, fault size (displace-
2015). Strata were not decompacted during back- ment and length) provides an important control
stripping as displacement loss due to sediment on subsequent reactivation, with only the larger
C. REILLY ET AL.

Fig. 5. Plots comparing (a) normal and inverted vertical displacements and (b) trace lengths for large inverted faults
in the southern Taranaki Basin. The data show a positive relationship clustered around the 1:1 lines, suggesting that
displacements and lengths during reverse and normal phases of faulting are similar.

normal faults inverted during shortening. The size Stable fault lengths are often accompanied by
dependence of inversion is illustrated in Figure 4, comparable maximum displacement magnitudes
which shows the maximum vertical displacement for normal and reverse faulting. The correspondence
v. length for normal and reverse faults. On each of maximum displacements is considered to be
graph, the Taranaki Basin data are consistent with fortuitous, and arises because regional finite strains
the global dataset, suggesting that in displacement– during extension and shortening were approxima-
length space the faults examined in this study are tely equal across the southern Taranaki Basin.
typical of many faults worldwide. The most striking In circumstances where regional extensional and
aspect of these graphs is the segregation of inver- contractional strains differed significantly, it is
ted and non-inverted faults. In Figure 4a, only the expected that the maximum displacements would
largest 14 normal faults in the initial extensional also be different between faulting phases. Even
system were inverted. At the end of the Paleocene, though the maximum displacements and fault
these 14 extensional faults had maximum displace- lengths are often comparable, the shapes of dis-
ments and lengths of ≥600 m and ≥9 km, respecti- placement profiles between deformation episodes
vely. Displacement–length relationships for reverse can be similar or vary significantly (Fig. 7). In
faults (Fig. 4b) are similar to those observed for Figure 7a, for example, the location of the maxi-
normal faults (Fig. 4a), with the majority of the larg- mum displacement and the associated displacement
est structures in the reverse system (i.e. ≥600 m gradients are similar between faulting phases,
maximum separation and ≥9 km length) being in- suggesting that the fault interactions that gave rise
verted normal faults. to these displacement variations (cf. Nicol et al.
Along-strike displacement profiles support the 1996) did not vary markedly between episodes.
view that fault lengths and maximum displacements The strong positive correlation of lengths and
were generally comparable during each faulting displacements between deformation phases is indi-
episode (Fig. 7). The likelihood of fault propa- cated in Figure 5 and arises because the largest
gation during inversion appears to increase with faults during the initial normal faulting were also
decreasing size of the initial normal fault (Fig. 7). the largest inverted faults. The similarity of fault
For the largest faults in the system (displacement lengths during each deformation phase suggests
≥c. 800 m and trace length ≥c. 25 km), the tips that the crustal heterogeneities produced by the orig-
remained stationary with no resolvable propa- inal normal faults provided a pre-eminent control
gation during inversion beyond the pre-existing on the location, orientation and length of the subse-
normal fault (Fig. 7a–d, left). In the few cases, quent reverse faults. As a consequence, the hierar-
where fault lengths differed pre- and post-inversion, chy of the largest reverse faults (i.e. their relative
contraction resulted in the propagation and crea- sizes and locations) appears to have been strongly
tion of new fault lengths (Fig. 7e, f). Examples influenced by the sizes of the pre-existing normal
where parts of the initial normal faults were only faults. As the inverted faults accommodated most
inverted along part of their length have not been (c. 80%) of the contraction in the basin and their
observed. relative sizes were inherited from the pre-existing
SIZE CONTROLS ON INVERSION

weakness (e.g. McKenzie 1969; Raleigh et al.


1972; White et al. 1986; Imber et al. 1997; Holds-
worth 2004; Worthington & Walsh 2011). At out-
crop scale, weak fault rock is commonly observed:
within fault zones, however, the width, structure
and thickness of weak fault rock are highly hetero-
geneous on metre to decametre scales (e.g. Childs
et al. 2009). Walsh et al. (2001) contended that
outcrop-scale heterogeneities are unlikely to control
regional-scale reactivation and, instead, suggested
that the geometrical properties of a fault system
(e.g. fault dimensions and locations relative to
other elements in fault array) exert a first-order
control on fault reactivation and system evolution.
Their conclusions are supported by numerical
Distinct Element Models, which suggest that fault
dimensions and locations have a greater impact
on subsequent growth than fault strength (Walsh
et al. 2001). The geometrical properties of pre-
existing faults are important for subsequent reacti-
vation because they produce crustal heterogeneity
that locally concentrates stress. Such heterogeneity
may reflect the presence of a shear fabric and/or
the juxtaposition of rocks with different strengths,
and need not be characterized by an inherent
fault-zone weakness. The role of heterogeneities in
concentrating stress is widely recognized in material
science literature, with the load required to rupture
an existing crack inversely proportional to the
length of the heterogeneity (Griffith 1921; Gordon
1978; Murakami 2002; Pilkey & Pilkey 2008).
Therefore, the size of the fault is likely to be an
important determinant for the magnitude and spatial
extent of elevated stresses, which, in turn, con-
trol the dimensions and locations of future failure
(Walsh et al. 2001).
Newly formed reverse faults are typically smaller
than inverted faults, with displacements of ≤600 m.
Furthermore, the locations of the newly formed
reverse faults differ from the smaller original nor-
mal faults. New reverse faults are predominantly
located distal to inverted faults and appear to form
in areas of regional shortening deficit rather
than being clustered close the inverted faults
(Fig. 8e). Lines A and B, drawn parallel to the
Fig. 6. Horizon surfaces and fault polygons from principal horizontal shortening direction (Fig. 8e),
interpretations carried out in this study for for example, show that the number of new faults
(a) Pleistocene, (b) Late Miocene and (c) Late are at a maximum across the centre of the basin
Cretaceous time periods.
on line B where the number and size of inverted
faults is at a minimum. The resulting shortening def-
normal fault system, individual inverted faults icit in the central portion of line B is inferred to
accommodated a similar proportion of the total have been partly accommodated by the initiation
strain across the basin during each deformation of new small reverse faults. The dearth of newly for-
episode. med small reverse faults close to the inverted faults
Fault reactivation (both extensional and con- may partly arise because, in these areas, the majo-
tractional) and associated strain localization onto rity of the stress was concentrated on the inverted
pre-existing structures is widely attributed to structures, and the areas enclosing these faults
the original faults representing zones of crustal were effectively destressed (e.g. Ackermann &
C. REILLY ET AL.

(a) Cape Egmont Fault 1600


(b) Kahurangi Fault
Pre-existing 1400
1000
fault 1200

Inverted fault 1000

800

500
600

400

200

0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60

2500

(c) Wakamarama Fault (d) Maari Fault


1200
Displacement (ms)

2000
1000

1500
800

600
1000

400

500
200

0 0
0 5 10 15 20 25 30 35 40 45 0 5 10 15 20 25

400
(e) Surville Fault (f) Whitiki Fault
1200

1000 300

800
Inverted fault tips
200
600
Inverted fault tips
400 Original
fault tips 100

200 Original
fault tips
0 0
0 10 20 30 40 50 60 70 0 5 10 15 20 25

Distance along fault trace (km)

Fig. 7. Along-strike displacement profiles for the (a) Cape Egmont, (b) Kahurangi, (c) Wakamarama, (d) Maari,
(e) Surville and (f) Whitiki faults in the southern Taranaki Basin (see Fig. 2 for locations). Total normal displacement
on the top basement horizon is shown by the grey curve, and reverse displacement on the base Oligocene horizon,
accommodated during inversion by the black curve. In (a–d), the normal and inverted profiles have comparable
lengths, while (b –d) also have comparable displacement distributions. By contrast, the Surville (e) and Whitiki
(f) faults propagated laterally during inversion, and presently exceed the original normal fault length.

Schlische 1997). Our observations support the view to their final configuration. Displacement back-
that large faults can increase or decrease stresses to stripping of individual faults constrains the evolu-
nearby faults in the system, which has the potential tion of the reverse fault system in the southern
to locally promote or impede the development of Taranaki Basin since about 20 Ma (Fig. 8). Perform-
inverted faults (Cowie 1998; Walsh et al. 2001). ing displacement back-stripping on multiple seis-
mic sections along the strike of each fault enables
changes in their length to be mapped through time
Fault-system evolution and strain (for a full description of method refer to Reilly
localization et al. 2015, fig. 8). During the Miocene, contrac-
tion and fault inversion episodically stepped west-
Analysis of finite displacements and lengths at the wards across the southern Taranaki Basin (Reilly
cessation of each faulting episode (Figs 4– 7) pro- et al. 2015) (Fig. 8). By the end of the Miocene
vides limited information on how the faults grew (c. 6 Ma), the areal extent of contraction was similar
SIZE CONTROLS ON INVERSION
Fig. 8. Maps showing temporal changes in the spatial distribution and types of faulting in the southern Taranaki Basin: (a) (60 Ma) Late Cretaceous– Paleocene extension;
(b) (30 Ma) Oligocene onset of eastern basin contraction; (c) (20 Ma) Early Miocene contraction; (d) (10 Ma) Mid-Miocene westwards migration of contraction; (e) (6 Ma) the
Latest Miocene greatest extent of contraction; and (f) (3 Ma) Plio-Pleistocene southwards migration of extension. The maps were produced using displacement back-stripping
of strata imaged in 2D and 3D seismic reflection lines (e.g. Fig. 3). Active normal faults (black), possible active normal faults (dashed black), active inverted faults (red),
possible active inverted faults (dashed red) and active, newly formed reverse faults (green). Lines A and B show the locations of regional strain estimates (see the text for
further discussion).
C. REILLY ET AL.

Fig. 9. Schematic block diagram showing the evolution of the inverted fault system in the southern Taranaki Basin.
(a) Extensional normal faults with a range of sizes form and create half-graben that fill with sediment (b).
(c) Following extension, uniform thickness strata are deposited across the basin during a period of thermal
subsidence and tectonic quiescence. (d) Basin contraction coaxial with the original extension direction. The largest
of the original normal faults are inverted with associated anticlines in the fault hanging walls and low-amplitude
synclines in their footwalls. The uplifted hanging-wall anticlines are partially eroded with the sediment deposited in
the adjacent synclines.
SIZE CONTROLS ON INVERSION

to that of the earlier extension. The positive relation- Fault death in the central basin was probably due,
ships in Figures 4 and 5 arise, in part, because of the in part, to a change in plate boundary conditions
spatial coincidence of the regions of normal and and a decrease in the rates of shortening associated
reverse faulting, and would not be expected to be with the southwards migration of intra-arc exten-
so pronounced if the region of overlap was smaller. sion (King & Thrasher 1996; Cande & Stock
If, for example, we had sampled the normal and 2004; Giba et al. 2010; Seebeck et al. 2014; Reilly
reverse fault systems at 20 Ma, when contraction et al. 2015). Further south, where the basin extends
was concentrated along the eastern margin of the onshore in the northern South Island, the rates of
basin (Fig. 8), all but three of the faults in Figure 5 shortening have not decreased significantly since
would be lying on the x-axis with no clear relation- 5 Ma, yet few of the smaller reverse faults remained
ship between the lengths and displacements during active at 3 Ma (Fig. 8). Near the coastline, the death
each deformation phase. Therefore, for the relation- of these small faults may be partly due to strain
ships in Figure 5 to apply widely to other basins, localization onto the larger faults.
each deformation phase should reoccupy appro- The evolution of the inverted fault system in the
ximately the same regions of the crust: this is not southern Taranaki Basin, where the principal hori-
always the case. It has, for example, been variously zontal extension and shortening axes were coaxial,
shown elsewhere that inversion episodes can affect is illustrated schematically in Figure 9. The key ele-
a part of a previously extensional basin (e.g. Wang ments of the model are that: (i) only the largest of
et al. 1995), the entire basin (e.g. Uliana et al. the early normal faults are reactivated during inver-
1995; Grimaldi & Dorobek 2011) or different parts sion; (ii) the locations and lengths of inverted faults
of a basin at different times (e.g. Jackson et al. are generally inherited from earlier normal faults;
2013). In addition to inversion at 20 Ma being lim- and (iii) for those faults that were inverted, the hier-
ited to the large basin-bounding faults, it seems that archy of fault sizes is maintained between defor-
few new smaller faults were generated at this time. mation phases. As the locations and displacements
The absence of smaller faults may arise because of newly formed reverse faults appear to be con-
the basin-bounding faults collectively extend along trolled by the larger inverted faults, it is suggested
strike for at least 400 km, producing crustal-scale that the inverted faults establish their final length
heterogeneities that are large enough to accommo- early. Displacement profiles for inverted faults in
date the entire stress (and strain) associated with the Taranaki Basin typically show little evidence
the regional shortening at this time. of propagation, which is consistent with the idea
Normal fault systems typically develop in asso- that the lengths of contractional faults were estab-
ciation with strain localization and progressive con- lished early. Such rapid early growth in fault
centration of displacements onto the largest fault(s) lengths has been proposed for natural fault systems
in a system, with the remainder of faults dying (Walsh et al. 2002) and from analogue models
(Nicol et al. 1997; Cowie 1998; Walsh et al. 2001, (Schlagenhauf et al. 2008), and may be a feature
2002; Meyer et al. 2002; Gawthorpe et al. 2003; of many fault systems whether they are reactivated
Whipp et al. 2014). Superficially, the fault maps in or new, extensional or contractional. It is possible
Figure 8 appear to indicate that, rather than localiz- that, because of the crustal-scale heterogeneities
ing with increasing finite strain, contractional fault- produced by early normal faults, the final lengths
ing became more spatially distributed with time. of inverted faults formed instantaneously on geo-
The increase in geographical spread of the active logical timescales (e.g. ,100 kyr): that is, within
faulting is, in part, due to a change in boundary con- several seismic cycles. However, in common with
ditions during contraction that may reflect increases normal fault systems (e.g. Walsh et al. 2002; Meyer
in the rates of convergence across the plate boun- et al. 2002), the seismic reflection and well data
dary during the Miocene, as proposed by Reilly are not sufficiently detailed to quantify the timing
et al. (2015) from regional analysis of fault evo- or the rates of this early propagation.
lution. In addition to the increase in regional spread
of contraction, many new smaller faults (displa-
cements ,400 m and lengths ,10 km) formed Conclusions
between 5 and 7 Ma, when the rates of shortening
across the basin were at a maximum of up to about The role of fault size in the inversion of pre-existing
1 mm a21. These smaller faults accommodate a normal faults has been examined using 2D and 3D
small, but important, component of the total strain seismic reflection lines tied to wells. The size (dis-
(c. 12%) and appear to have been initiated in asso- placement and length) of early normal faults is
ciation with an order of magnitude increase in the shown to be a key determinant for the reactivation
regional rates of shortening (Reilly et al. 2015). and size of the subsequent reverse faults when short-
Smaller faults initiated in the Late Miocene cea- ening and extension directions are parallel. In the
sed to accrue displacement by 3 Ma (Fig. 7f). southern Taranaki Basin, all normal faults with
C. REILLY ET AL.

maximum vertical displacements ≥c. 600 m and Bulnes, M. & McClay, K.R. 1998. Structural analysis and
trace lengths ≥c. 9 km were inverted, while few kinematic evolution of the inverted South Celtic Sea
faults with maximum vertical displacements and Basin. Marine and Petroleum Geology, 15, 667–687.
trace lengths smaller than this were inverted. In Cande, S.C. & Stock, J.M. 2004. Pacific–Antarctic–
Australia motion and the formation of the Mac-
the majority of cases, the entire length of the initial quarie Plate. Geophysical Journal International, 157,
normal fault was inverted, with the proportion of 399–414.
the total basin-wide strain accommodated on the Chadwick, R.A. & Evans, D.J. 1995. The timing and
fault comparable between both extensional and direction of Permo-Triassic extension in southern
contractional deformational episodes. The hierarchy Britain. In: Boldy, S.A.R. (ed.) Permian and Triassic
of reverse fault lengths was established rapidly, Rifting in Northwest Europe. Geological Society,
with longer faults accruing a greater proportion of London, Special Publications, 91, 161– 192, http://
the total strain from an early stage of deformation. doi.org/10.1144/GSL.SP.1995.091.01.09
The reverse fault system is dominated by inverted Childs, C., Easton, C.J., Vendeville, B.C., Jackson,
M.P.A., Lin, S.T., Walsh, J.J. & Watterson, J.
faults, which accrue displacement at the expense 1993. Kinematic Analysis of faults in a physical model
of smaller faults, and utilize the largest crustal-scale of growth faulting above a viscous salt analogue.
elements of the pre-existing system. The size of Tectonophysics, 228, 313– 329.
pre-existing crustal heterogeneities is an important Childs, C., Nicol, A., Walsh, J.J. & Watterson, J.
determinant for the magnitude and spatial extent 2003. The growth and propagation of synsedimentary
of elevated stresses, which, in turn, control the dimen- faults. Journal of Structural Geology, 25, 633–648.
sions, locations and displacements of subsequent Childs, C., Manzocchi, T., Walsh, J.J., Bonson, C.G.,
fault growth. Nicol, A. & Schöpfer, M.P.J. 2009. A geometric
model of fault zone and fault rock thickness variations.
Funding for this work was provided by a GNS Science Journal of Structural Geology, 31, 117 –127.
Scholarship, with support from the Fault Analysis Group Clausen, O.R., Howard, C.B. et al. 1994. Systematics
at University College Dublin. Constructive comments of faults and fault arrays. In: Helbig, K. (ed.) Model-
from Jonathan Imber, Chris Jackson, David McNamara ling the Earth for Oil Exploration. Final Report of
and Hannu Seebeck helped to improve the manuscript. the CEC’s Geoscience Program 1990–1993. Elsevier,
Amsterdam, 205– 316.
Cowie, P.A. 1998. A healing–reloading feedback control
References on the growth rate of seismogenic faults. Journal of
Structural Geology, 20, 1075–1087.
Ackermann, R.V. & Schlische, R.W. 1997. Anticluster- Davis, P.N. 1983. Gippsland Basin, SE Australia. In:
ing of small normal faults around larger faults. Geo- Bally, A.W. (ed.) Seismic Expression of Structural
logy, 25, 1127–1130. Styles. American Association of Petroleum Geologists,
Badley, M.E., Price, J.D. & Backshall, L.C. 1989. Special Publications, 3, 3.3.19– 3.3.24.
Inversion, reactivated faults and related structures: Del Ventisette, C., Montanari, D., Sani, F. & Bonini,
seismic examples from the southern North Sea. In: M. 2006. Basin inversion and fault reactivation in lab-
Cooper, M.A. & Williams, G.D. (eds) Inversion oratory experiments. Journal of Structural Geology,
Tectonics. Geological Society, London, Special Publi- 28, 2067–2083.
cations, 44, 201–219, http://doi.org/10.1144/GSL. Eisenstadt, G. & Withjack, M.O. 1995. Estimating
SP.1989.044.01.12 inversion: results from clay models. In: Buchanan,
Bailey, W.R., Walsh, J.J. & Manzocchi, T. 2005. Fault J.G. & Buchanan, P.G. (eds) Basin Inversion. Geo-
populations, strain distribution and basement reacti- logical Society, London, Special Publications, 88,
vation in the East Pennines Coalfield, U.K. Journal 119–136, http://doi.org/10.1144/GSL.SP.1995.088.
of Structural Geology, 27, 913– 928. 01.08
Ballance, P.F. 1976. Evolution of the upper Cenozoic Fraser, A.J. & Gawthorpe, R.L. 1990. Tectono-
magmatic arc and plate boundary in northern New stratigraphic development and hydrocarbon habitat of
Zealand. Earth and Planetary Science Letters, 28, the Carboniferous in northern England. In: Hardman,
356– 370. R.F.P. & Brooks, J. (eds) Tectonic Events Responsible
Beavan, J. & Haines, J. 2001. Contemporary horizontal for Britain’s Oil and Gas Reserves. Geological Soci-
velocity and strain rate fields of the Pacific– Australian ety, London, Special Publications, 55, 49–86, http://
plate boundary zone through New Zealand. Journal of doi.org/10.1144/GSL.SP.1990.055.01.03
Geophysical Research, 106, 741 –770. Gawthorpe, R.L., Jackson, C.A.L., Young, M.J., Sharp,
Bishop, D.J. & Buchanan, P.G. 1995. Development of I.R., Moustafa, A.R. & Leppard, C.W. 2003. Normal
structurally inverted basins: a case study from the fault growth, displacement localisation and the evolu-
West Coast, South Island, New Zealand. In: Bucha- tion of normal fault populations: the Hammam Faraun
nan, J.G. & Buchanan, P.G. (eds) Basin Inversion. fault block, Suez rift, Egypt. Journal of Structural Geo-
Geological Society, London, Special Publications, logy, 26, 883–895.
88, 549–585, http://doi.org/10.1144/GSL.SP.1995. Giba, M., Walsh, J.J. & Nicol, A. 2010. Evolution
088.01.28 of faulting and volcanism in a backarc basin and its
Brun, J.P. & Nalpas, T. 1996. Graben inversion in nature implications for subduction processes. Tectonics, 29,
and experiments. Tectonics, 15, 677– 687. TC4020, http://doi.org/10.1029/2009TC002634
SIZE CONTROLS ON INVERSION

Giba, M., Walsh, J.J. & Nicol, A. 2012. Segmentation King, P.R. 1990. Polyphase evolution of the Taranaki
and growth of an obliquely reactivated normal fault. Basin, New Zealand: changes in sedimentary and
Journal of Structural Geology, 39, 253– 267. structural style. In: 1990 New Zealand Oil Exploration
Glennie, K.W. & Boegner, P.L.E. 1981. Sole pit Conference Proceedings. Ministry of Commerce,
inversion tectonics. In: Illing, L.V. & Hobson, G.D. Wellington, 134–150.
(eds) Petroleum Geology of the Continental Shelf of King, P.R. & Thrasher, G.P. 1996. Cretaceous –
Northwest Europe. Institute of Petroleum, London, Cenozoic Geology and Petroleum Systems of the Tara-
110–120. naki Basin, New Zealand. Institute of Geological &
Gluyas, J.G., Evans, I.J. & Richards, D. 2003. The Nuclear Sciences, Lower Hutt, New Zealand, Mono-
Kimmeridge Bay Oilfield, UK Onshore. In: Gluyas, graphs, 13.
J.G. & Hichens, H.M. (eds) United Kingdom Oil Knox, G.J. 1982. Taranaki Basin, structural style and
and Gas Fields, Commemorative Millennium Volume. tectonic setting. New Zealand Journal of Geology
Geological Society, London, Memoirs, 20, 943–948, and Geophysics, 25, 125–140.
http://doi.org/10.1144/GSL.MEM.2003.021.01.80 Lamplugh, G.W. 1920. Structure of the Weald and
Gordon, J.E. 1978. Structures, or Why Things Don’t Fall analogues tracts. Quarterly Journal of the Geological
Down. Plenum, New York. Society, London, 75, LXXIII– XCV (Anniversary
Griffith, A.A. 1921. The phenomena of rupture and flow address of the President).
in solids. Philosophical Transactions of the Royal Lezzar, K.E., Tiercelin, J.-J., Le Turdu, C., Cohen,
Society of London A, 221, 163–198. A.S., Reynolds, D.J., Le Gall, B. & Scholz, C.A.
Grimaldi, G.O. & Dorobek, S.L. 2011. Fault framework 2002. Control of normal fault interaction on the dis-
and kinematic evolution of inversion structures: natu- tribution of major Neogene sedimentary depocen-
ral examples from the Neuquen Basin, Argentina. tres, Lake Tanganyika, East African rift. American
American Association of Petroleum Geologists Bulle- Association of Petroleum Geologists Bulletin, 86,
tin, 95, 27–60. 1027– 1059.
Hayward, A.B. & Graham, R.H. 1989. Some geometrical Madon, M., Yang, J-S., Abolins, P., Hassan, R.A.,
characteristics of inversion. In: Cooper, M.A. & Yakzan, A.M. & Zainal, S.B. 2004. Petroleum sys-
Williams, G.D. (eds) Inversion Tectonics. Geological tems of the Northern Malay Basin. Paper presented
Society, London, Special Publications, 44, 17– 40, at the Petroleum Geology Conference & Exhibition,
http://doi.org/10.1144/GSL.SP.1989.044.01.03 15–16 December 2004, Kuala Lumpur, Malaysia.
Hemmings-Sykes, S. 2011. The influence of faulting McKenzie, D.P. 1969. The relation between fault plane
on hydrocarbon migration in the Kupe area, south solutions for earthquakes and the directions of the prin-
Taranaki Basin, New Zealand. MSc thesis, Victo- cipal stresses. Bulletin of the Seismological Society of
ria University of Wellington, Wellington, New America, 59, 591–601.
Zealand. Meyer, V., Nicol, A., Childs, C., Walsh, J.J. &
Hill, M.G. & Milner, M. 2012. The Kupe Velocity Watterson, J. 2002. Progressive localisation of strain
Model – 4D Taranaki Project. GNS Science Report during the evolution of a normal fault population. Jour-
2012/3. GNS Science, Lower Hutt, New Zealand. nal of Structural Geology, 24, 1215– 1231.
Holdsworth, R.E. 2004. Weak faults, rotten cores. Sci- Morgans, H.E.G. 2013. Oligocene to Pliocene Biostratig-
ence, 303, 181–182, http://doi.org/10.1126/science. raphy of the Marine Sequences in Exploration Wells in
1092491 the Kupe Region, South Taranaki Bight, New Zealand.
Holt, W.E. & Stern, T.A. 1994. Subduction, platform GNS Science Report 2013/48. GNS Science, Lower
subsidence, and foreland thrust loading: the late Ter- Hutt, New Zealand.
tiary development of Taranaki Basin, New Zealand. Morley, C.K. 1995. Developments in the structural geo-
Tectonics, 13, 1068–1092. logy of rifts over the last decade and their impact
Hoolihan, K. & Yang, J.S. 1991. Eocene to Recent struc- on hydrocarbon exploration. In: Lambiase, J.J. (ed.)
tural history of onshore Taranaki, with reference to the Hydrocarbon Habitat in Rift Basins. Geological Soci-
main play styles. In: 1991 New Zealand Oil Explora- ety, London, Special Publications, 80, 1– 32, http://
tion Conference Proceedings. Ministry of Commerce, doi.org/10.1144/GSL.SP.1995.080.01.01
Wellington, 168– 175. Morley, C.K. 1999. Patterns of displacement along large
Imber, J., Holdsworth, R.E., Roberts, A.M. & Lloyd, normal faults: implications for basin evolution and
G.E. 1997. Fault-zone weakening processes along fault propagation, based on examples from East Africa.
the reactivated Outer Hebrides Fault Zone, Scotland. American Association of Petroleum Geologists
Journal of the Geological Society, London, 154, Bulletin, 83, 613 –634.
105–110, http://doi.org/10.1144/gsjgs.154.1.0105 Morley, C.K., Haranya, C., Phoosongsee, W., Pong-
Jackson, C.A.L., Chua, S.T., Bell, R.E. & Magee, C. wapee, S., Kornsawan, A. & Wonganan, N. 2004.
2013. Structural style and early stage growth of inver- Activation of rift oblique and rift parallel pre-existing
sion structures: 3D seismic insights from the Egersund fabrics during extension and their effect on deforma-
Basin, offshore Norway. Journal of Structural Geo- tion style: examples from the rifts of Thailand. Journal
logy, 46, 167–185. of Structural Geology, 26, 1803–1829.
Kelly, P.G., Peacock, D.C.P., Sanderson, D.J. & Murakami, Y. 2002. Metal Fatigue: Effects of Small
McGuirk, A.C. 1999. Selective reverse reactivation Defects and Nonmetallic Inclusions. Elsevier,
of normal faults, and deformation around reverse- Amsterdam.
reactivated faults in the Mesozoic of the Somerset Nicol, A., Watterson, J., Walsh, J.J. & Childs, C.
coast. Journal of Structural Geology, 21, 493– 509. 1996. The shapes, major axis orientations and
C. REILLY ET AL.

displacement patterns of fault surfaces. Journal of 273, 299– 311, http://doi.org/10.1016/j.epsl.2008.


Structural Geology, 18, 235–248. 06.042
Nicol, A., Walsh, J.J., Watterson, J. & Underhill, Schlishe, R.W., Young, S.S., Ackerman, R.V. & Gupta,
J.R. 1997. Displacement rates of normal faults. Nature, A. 1996. Geometry and scaling relations of a popula-
390, 157– 159. tion of very small rift-related normal faults. Geology,
Nicol, A., Walsh, J., Berryman, K. & Nodder, S. 2005. 24, 683– 686.
Growth of a normal fault by accumulation of slip over Schmidt, D.S. & Robinson, P.H. 1989. The structural set-
millions of years. Journal of Structural Geology, 27, ting and depositional history for the Kupe South Field,
327– 342. Taranaki Basin. In: 1989 New Zealand Oil Exploration
Palmer, J.A. & Andrews, P.R. 1993. Cretaceous – Conference Proceedings. Ministry of Commerce,
Tertiary sedimentation and implied tectonic controls Wellington, 151– 172.
on the structural evolution of Taranaki Basin, New Seebeck, H., Nicol, A., Villamor, P., Ristau, J. &
Zealand. In: Ballance, P.F. (ed.) South Pacific Sedi- Pettinga, J. 2014. Structure and kinematics of the
mentary Basins. Sedimentary Basins of the World, 2. Taupo Rift, New Zealand. Tectonics, 33, 1178– 1199.
Elsevier, Amsterdam, 309–328. Sibson, R.H. 1995. Selective fault reactivation during
Panien, M., Schreurs, G. & Pfiffner, A. 2005. Sandbox basin inversion: potential for fluid redistribution
experiments on basin inversion: testing the influence of through fault-valve action. In: Buchanan, J.G. &
basin orientation and basin fill. Journal of Structural Buchanan, P.G. (eds) Basin Inversion. Geological
Geology, 27, 433– 445. Society, London, Special Publications, 88, 3 –19,
Petersen, K., Clausen, O.R. & Korstgard, J.A. 1992. http://doi.org/10.1144/GSL.SP.1995.088.01.02
Evolution of a salt-related listric growth fault near Stagpoole, V.M. & Nicol, A. 2008. Regional structure
the D-1 well, block 5605, Danish North Sea; displace- and kinematic history of a large subduction back thrust:
ment history and salt kinematics. Journal of Structural Taranaki Fault, New Zealand. Journal of Geophysical
Geology, 14, 565– 577. Research: Solid Earth, 113, B01403, http://doi.org/
Pilaar, W.F.H. & Wakefield, L.L. 1978. Structural and 10.1029/2007JB005170
stratigraphic evolution of the Taranaki Basin, offshore Stille, H. 1924. Grundfragen der Vergleichenden
North Island, New Zealand. The APEA Journal, 18, Tektonik. Brontrager, Berlin.
93–101. Taylor, S.K., Nicol, A. & Walsh, J.J. 2008. Displace-
Pilkey, W.D. & Pilkey, D.F. 2008. Peterson’s ment loss on growth faults due to sediment compac-
Stress Concentration Factors, 3rd edn. John Wiley, tion. Journal of Structural Geology, 30, 394–405.
Hoboken, NJ. Thrasher, G.P. 1990. Tectonics of the Taranaki Rift.
Raleigh, C.B., Healy, J.H. & Bredchoeft, J.D. 1972. In: 1989 New Zealand Oil Exploration Conference
Faulting and crustal stress at Rangely, Colorado. In: Proceedings. Ministry of Commerce, Wellington,
Heard, H.C., Borg, I.Y., Carter, N.L. & Raleigh, 124–133.
C.B. (eds) Flow and Fracture of Rocks. American Thrasher, G.P. 1992. Late Cretaceous geology of
Geophysical Union, Geophysical Monograph Series, Taranaki Basin, New Zealand. PhD thesis, Victoria
16, 275–284. University of Wellington, New Zealand.
Reilly, C., Nicol, A., Walsh, J.J. & Seebeck, H. 2015. Thrasher, G.P. & Cahill, J.P. 1990. Subsurface Maps
Evolution of faulting and plate boundary deformation of the Taranaki Basin Region, New Zealand. New Zea-
in the Southern Taranaki Basin, New Zealand. Tecto- land Geological Survey Report G142. New Zealand
nophysics, 651–652, 1 –18, http://doi.org/10.1016/ Geoloical Survey, Lower Hutt.
j.tecto.2015.02.009. Turner, J.P. & Williams, G.A. 2004. Sedimentary basin
Roncaglia, L., Morgans, H.E.G. et al. 2008. Stratigra- inversion and intra-plate shortening. Earth-Science
phy, Well Correlation and Seismic-to-well tie in the Reviews, 65, 277–304.
Upper Cretaceous to Pliocene Interval in the Kupe Uliana, M.A., Arteaga, M.E., Legarreta, L., Cerdán,
Region, Taranaki Basin, New Zealand. Introduction J.J. & Peroni, G.O. 1995. Inversion structures and
to the Stratigraphic Database in PETREL. GNS Sci- hydrocarbon occurrence in Argentina. In: Buchanan,
ence Report 2008/7. GNS Science, Lower Hutt, New J.G. & Buchanan, P.J. (eds) Basin Inversion. Geo-
Zealand. logical Society, London, Special Publications, 88,
Roncaglia, L., Milner, M. et al. 2010. Procedures and 211–233, http://doi.org/10.1144/GSL.SP.1995.088.
Metadata Protocols used in Modelling Taranaki Basin 01.13
Petroleum Systems: Guidelines from a Pilot Case Walcott, R.I. 1978. Present tectonics and late Cenozoic
Study in the Kupe Area, Lower Hutt. GNS Science evolution of New Zealand. Geophysical Journal of
Report 2009/49. GNS Science, Lower Hutt, New the Royal Astronomical Society, 52, 137–164.
Zealand. Walsh, J.J., Childs, C. et al. 2001. Geometric controls
Roncaglia, L., Fohrmann, M., Milner, M., Morgans, on the evolution of normal fault systems. In: Holds-
H.E.G. & Crundwell, M.P. 2013. Well Log Stratigra- worth, R.E., Strachan, R.A., Magloughlin, J.F.
phy in the Central and Southern Offshore Areas of the & Knipe, R.J. (eds) The Nature and Tectonic Signifi-
Taranaki Basin, New Zealand. GNS Science Report cance of Fault Zone Weakening. Geological Society,
2013/27. GNS Science, Lower Hutt, New Zealand. London, Special Publications, 186, 157 –170, http://
Schlagenhauf, A., Manighetti, I., Malavieille, J. & doi.org/10.1144/GSL.SP.2001.186.01.10
Dominguez, S. 2008. Incremental growth of nor- Walsh, J.J., Nicol, A. & Childs, C. 2002. An alternative
mal faults: insights from a laser-equipped analog model for the growth of faults. Journal of Structural
experiment. Earth and Planetary Science Letters, Geology, 24, 1669–1675.
SIZE CONTROLS ON INVERSION

Walsh, J.J., Childs, C., Imber, J., Manzocchi, T., example from the northern Horda Platform, Norwegian
Watterson, J. & Nell, P.A.R. 2003. Strain loca- North Sea. Basin Research, 26, 523– 549.
lisation and population changes during fault sys- White, S.H., Bretan, P.G. & Rutter, E.H. 1986.
tem growth within the Inner Moray Firth, Northern Fault-zone reactivation: kinematics and mechanisms.
North Sea. Journal of Structural Geology, 25, Philosophical Transactions of the Royal Society of
1897–1911. London A, 317, 81– 97.
Wang, G.M., Coward, M.P., Yuan, W., Liu, S. & Wang, Williams, G.D., Powell, C.M. & Cooper, M.A. 1989.
W. 1995. Fold growth during basin inversion – Geometry and kinematics of inversion tectonics.
example from the East China Sea Basin. In: Bucha- In: Cooper, M.A. & Williams, G.D. (eds) Inversion
nan, J.G. & Buchanan, P.J. (eds) Basin Inversion. Tectonics. Geological Society, London, Special Pub-
Geological Society, London, Special Publications, 88, lications, 44, 3–15, http://doi.org/10.1144/GSL.SP.
493–522, http://doi.org/10.1144/GSL.SP.1995.088. 1989.044.01.02
01.26 Withjack, M., Baum, M.S. & Schlische, R.W.
Warren, M.J. 2009. Tectonic inversion and petroleum 2010. Influence of pre-existing fault fabric on inver-
system implications in the rifts of Central Africa. sion-related deformation: a case study of the inverted
Paper presented at the Frontiers and Innovation, Fundy rift basin, southeastern Canada. Tectonics, 29,
CSPG CSEG CWLS Convention, 4– 8 May 2009, 1– 22.
Calgary, Alberta. Worthington, R.P. & Walsh, J.J. 2011. Structure of
Whipp, P.S., Jackson, C.A-L., Gawthorpe, R.L., Lower Carboniferous basins of NW Ireland, and its im-
Dreyer, T. & Quinn, D. 2014. Normal fault array plications for structural inheritance and Cenozoic fault-
evolution above a reactivated rift fabric; a subsurface ing. Journal of Structural Geology, 33, 1285– 1299.

You might also like