a model of active faulting in new...

59
1 A model of active faulting in New Zealand NJ Litchfield a , R Van Dissen a , R Sutherland a , PM Barnes b , SC Cox c , R Norris d , RJ Beavan a , R Langridge a , P Villamor a , K Berryman a , M Stirling a , A Nicol a , S Nodder b , G Lamarche b , DJA Barrell c , JR Pettinga e , T Little f , N Pondard a , JJ Mountjoy b , K Clark a a GNS Science, P.O. Box 30368, Lower Hutt 5040, New Zealand b National Institute of Water and Atmospheric Research, P.O. Box 14901, Wellington 6021, New Zealand c GNS Science, Private Bag 1930, Dunedin 9016, New Zealand d Department of Geology, University of Otago, P.O. Box 56, Dunedin 9054, New Zealand e Department of Geological Sciences, University of Canterbury, Private Bag 4800, Christchurch 8140, New Zealand f School of Geography, Environment and Earth Sciences, Victoria University of Wellington, P.O. Box 600, Wellington, New Zealand Active fault traces are a surface expression of permanent deformation that accommodates the motion within and between adjacent tectonic plates. We present an updated national-scale model for active faulting in New Zealand, summarise the present understanding of fault kinematics in 15 tectonic domains, and undertake some brief kinematic analysis including comparison of fault slip rates with GPS velocities. The model contains 635 simplified faults with tabulated parameters of their attitude (dip and dip-direction) and kinematics (sense of movement and rake of slip vector); net slip rate; and a quality code. Fault density and slip rates are, as expected, highest along the central plate boundary zone, but the model is undoubtedly incomplete, particularly in rapidly eroding mountainous areas and submarine areas with limited data. The active fault data presented are of value to a range of kinematic, active fault and seismic hazard studies. Keywords: active fault, New Zealand, tectonic domain, plate boundary, kinematics, slip rate

Upload: others

Post on 27-Feb-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

1

A model of active faulting in New Zealand NJ Litchfielda, R Van Dissena, R Sutherlanda, PM Barnesb, SC Coxc, R Norrisd, RJ Beavana, R Langridgea, P Villamora, K Berrymana, M Stirlinga, A Nicola, S Nodderb, G Lamarcheb, DJA Barrellc, JR Pettingae, T Littlef, N Pondarda, JJ Mountjoyb, K Clarka a GNS Science, P.O. Box 30368, Lower Hutt 5040, New Zealand b National Institute of Water and Atmospheric Research, P.O. Box 14901, Wellington 6021, New Zealand c GNS Science, Private Bag 1930, Dunedin 9016, New Zealand d Department of Geology, University of Otago, P.O. Box 56, Dunedin 9054, New Zealand e Department of Geological Sciences, University of Canterbury, Private Bag 4800, Christchurch 8140, New Zealand f School of Geography, Environment and Earth Sciences, Victoria University of Wellington, P.O. Box 600, Wellington, New Zealand

Active fault traces are a surface expression of permanent deformation that accommodates the

motion within and between adjacent tectonic plates. We present an updated national-scale

model for active faulting in New Zealand, summarise the present understanding of fault

kinematics in 15 tectonic domains, and undertake some brief kinematic analysis including

comparison of fault slip rates with GPS velocities. The model contains 635 simplified faults

with tabulated parameters of their attitude (dip and dip-direction) and kinematics (sense of

movement and rake of slip vector); net slip rate; and a quality code. Fault density and slip

rates are, as expected, highest along the central plate boundary zone, but the model is

undoubtedly incomplete, particularly in rapidly eroding mountainous areas and submarine

areas with limited data. The active fault data presented are of value to a range of kinematic,

active fault and seismic hazard studies.

Keywords: active fault, New Zealand, tectonic domain, plate boundary, kinematics, slip rate

Page 2: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

2

Introduction

Active fault traces are a surface expression of permanent, generally coseismic, deformation.

Analyses of large datasets of active fault slip data can therefore provide insights into the

kinematics of plate boundary zones. For example, data on the geometry and slip rate of active

faults can be converted into tensor estimates of strain or seismic moment rates for comparison

with seismicity or geodetic data (e.g., Wesnousky et al. 1982; Holt & Haines 1995; Klein et

al. 2009; Stirling et al. 2009). Moreover, such data can be used for developing seismic (e.g.,

Stirling et al. 1998, 2002, 2012; Peterson et al. 2008) and tsunami (e.g., Berryman 2005;

Priest et al. 2009) hazard models. One key to the success of such analyses is having

moderately to highly complete active fault datasets with well constrained fault geometries

and slip rates, and that include reliable estimates in the uncertainty in the fault kinematic

properties.

In New Zealand a rich record of surface or near-surface land and marine active faults

are preserved. Reasons for this include: (i) its position straddling the Pacific-Australia plate

boundary zone (Fig. 1A); (ii) an historical record, albeit a short one (c. 170 years), of ground-

rupturing earthquakes; (iii) good geomorphic and bathymetric expression of faults, reflecting

a balance between deformation, erosion, and burial rates; (iv) relatively easy land access to

most recognised active fault traces; (v) large areas of forest clearance; (vi) extensive high

quality aerial photography coverage; (vii) extensive coverage of marine seismic reflection

and multibeam bathymetric datasets; and (viii) a relatively long history of the geological,

geodetic, and paleoseismological study of active faults (summarised in the next section). This

makes New Zealand an ideal place to study plate boundary kinematics using active fault data.

In this study, we summarise a new nation-wide compilation of active faults that spans

the entire New Zealand plate boundary (Fig. 1) and represents our current best understanding

of the location, style, and deformation rates of New Zealand active faults. We describe the

compilation as a model rather than database because each fault has been mapped at a

regional-scale (no better than 1:250,000) as a highly generalised line, with a single set of fault

characterisation parameters. Each mapped fault is therefore referred to as a ‘fault zone’ which

emphasises the distinction between the simplified fault zones in this model and the more

detailed active fault traces and parameters stored in databases such as the GNS Science

Active Faults Database of New Zealand (http://data.gns.cri.nz/af/). Fault zone names are

often informal. Our model is a significant advance on previous regional or national

compilations (summarised below), benefiting from numerous recent (in the last c. 10 years)

fault-specific studies, the systematic new mapping of active faults for the entire country at

Page 3: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

3

1:250,000 scale for the QMAP programme

(http://www.gns.cri.nz/Home/Products/Maps/Geological-Maps/QMAP.-1-250-000-

Geological-Map-of-New-Zealand), and the parallel development of other models that utilise

active fault data such as the National Seismic Hazard Model for New Zealand (Stirling et al.

2012). The tabulated characteristics of each fault zone in the model are presented in

supplementary file Tables S1 and S2, and comprise: (i) dip; (ii) dip-direction; (iii) sense of

movement; (iv) rake; (v) net slip rate; and a quality code. Also provided are regional-scale

maps of fault zone locations (supplementary file Figures S1-S9), and digital maps

(shapefile,ascii, and kmz files) of the fault zone locations (supplementary files Maps S1,S2

and S3). The derivations of each parameter for individual fault zones are described by

Litchfield et al. (2013).

The purposes of this paper are to: (i) describe the model compilation methods; (ii)

summarise the general nature and current state of knowledge of kinematics of active fault

zones in each of 15 tectonic domains (Fig. 2); (iii) provide some preliminary statistical

analysis of key model parameters; (iv) comment on model completeness; and (v) compare

geological fault slip data with GPS velocities. We expect that this paper will provide a

valuable and up-to-date synthesis upon which more detailed and/or multidisciplinary future

plate boundary kinematic investigations can be founded with confidence.

Previous regional-scale active fault compilations

The first systematic regional active fault compilation in New Zealand was undertaken by

Wellman (1953), who utilised aerial photographs to identify and map Recent and Late

Pleistocene South Island faults, including the Alpine, Wairau, Clarence, Kekerengu, and

Hope faults. Wellman (1955) then presented the first kinematic analysis of New Zealand

active faults in which transcurrent faults were grouped together in a circum-Pacific mobile

belt extending northeast across New Zealand, surrounded by a few normal faults.

Earthquakes and active faults were noted to be directly related to one another.

From the 1960s onwards, Late Quaternary fault traces were shown (as red lines) on

geological maps (e.g., Grindley 1960; Kingma 1962; Lensen 1963), but parameters

describing the kinematics of slip on these faults were only mentioned sparsely, if at all, in the

texts accompanying these published maps. In the 1970s and 1980s five active fault maps

were published as the Late Quaternary Tectonic Map of New Zealand series. This series

included the first, and in fact only, published New Zealand-wide active fault paper map

(Officers of the Geological Survey 1983). That map shows 31 named active faults and 15

Page 4: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

4

named active folds, but was not accompanied by any fault parameters such as dip, slip rate, or

sense of movement.

Over the same period a series of papers summarised the state of knowledge of earth

deformation in New Zealand at that time (Lensen 1975; Berryman 1984a,b; Lewis 1984;

Walcott 1984; Berryman & Beanland 1988). Some of these papers contain fault parameters

(e.g., 35 faults in Berryman & Beanland 1988) or summarise historical surface ruptures

(Berryman 1984b; Berryman & Beanland 1988). Most also describe regional variations in

fault styles and define a number of tectonic domains for the New Zealand region (also

Berryman & Beanland 1991).

The first digital databases of New Zealand active faults were developed in the 1990s.

The GNS Science Active Faults Database of New Zealand (http://data.gns.cri.nz/af/) was first

developed at 1:250,000 scale and nominally contains faults with evidence of activity in the

last 125,000 years (Jongens & Dellow 2003; Litchfield & Jongens 2006). Later, in the late

1990s, it began to be supplemented with data captured at 1:50,000 scale. Much active fault

data have also remained archived, if unpublished, in university theses. Significant submarine

active fault data has been collected by NIWA at a variety of scales from <1:50,000 to

1:200,000. All of these data have been meticulously perused, compiled, and incorporated

into this active fault model.

Since the early 1990s, New Zealand-wide analyses of active faults have been

primarily for the purpose of seismic hazard analysis (Stirling et al. 1998, 2002, 2012; Van

Dissen et al. 2003). The most recent (2010) update of the New Zealand Seismic Hazard

Model (Stirling et al. 2012) was compiled in parallel with the active fault model presented in

this paper. The model in this paper however, contains some new material, and is focused

more on faulting rates and deformation mechanisms, rather than as sources of past and

potentially future earthquakes.

Methods of compilation

Definition of an active fault zone

In the model presented in this paper, a fault zone is defined as active if there is either

evidence or inferred evidence for displacement in the past 125,000 years (late Pleistocene).

We make an exception for the Taupo Rift (Fig. 1A) however, which is evolving so rapidly

that fault zones in the rift are classified as active if there is evidence or inferred evidence for

displacement in the past 25,000 years (Villamor & Berryman 2001, 2006a). Thus only faults

in the inner, youngest part of the rift are included in this model. As well as representing

Page 5: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

5

timeframes over which we infer that current tectonic regime may be considered stable, these

timeframes are convenient in being marked on land by prominent marine or fluvial terraces

(Pillans 1990a, 1991) (125,000 years BP). Circa 25,000 years B.P., moreover, coincides with

the depositional age of nationally-widespread volcanic eruptive (the Kawakawa/Oruanui

Tephra e.g., Wilson 2001; Vandergoes et al. in press) that was deposited across most of

northern and central New Zealand. Offshore, many active fault zones beneath the continental

shelf are constrained by deformation of the transgressive post-last glacial (<20,000 years BP)

marine ravinement surface and associated post-glacial sediments (e.g., Lewis 1971; Nodder

1994; Barnes 1996; Lamarche et al. 2006). Farther offshore, active fault zones can be

recognized on the basis of whether or not they displace surfaces that formed since the last

interglacial period, and/or considering their prominent geomorphic expression in late

Quaternary submarine landscapes (e.g. Barnes 1996; Barnes et al. 2010). In this model we do

not include thrust faults of the frontal wedge of the Hikurangi subduction margin, which have

been considered to have developed aseismically (Barnes et al. 2010; Stirling et al. 2012).

Active fault zone traces

Our compilation maps show the generalised trace (i.e. surface position or projected surface

position) of each delineated active fault zone representative of crustal scale deformation (Fig.

1B; supplementary files Figures S1-S9 and Maps S1-S3). The simplified fault zone

characterisations presented in this paper are intended primarily to assist regional kinematic

and seismic hazard studies, and the more detailed datasets should be consulted for any

purposes relating to land-use or engineering development. The simplified fault zones (e.g.,

Fig. 3), are mostly the same as those used in the 2010 update of the New Zealand National

Seismic Hazard Model (Stirling et al. 2012). Where faults are represented in more detailed

databases by only one or a few short surface traces, the faults were either: (i) extended along

bedrock faults (faults separating different bedrock units, but mapped as inactive) from the

GNS Science 1:250,000-scale Geological Map of New Zealand (QMAP); (ii) extended along

topographic or bathymetric features, such as range-fronts or submarine ridges; or (iii) merged

to reflect minimum threshold earthquake magnitudes for ground-rupturing seismogenic

faults. Consideration was also given to likely fault length (e.g., fault zone 170 in Fig. 3A) and

subsurface connectivity (e.g., fault zones 230 and 229 in Fig. 3B).

On land ten blind fault zones are recognized. These include buried faults imaged on

seismic reflection profiles (Melhuish et al. 1996); those inferred by Jackson et al. (1998) to

exist at depth beneath active anticlines near Palmerston North (Fig. 1B); and the mainly blind

Page 6: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

6

fault inferred to have ruptured in the 1931 Napier Earthquake (Hull 1990). Offshore, no

differentiation is made here between faults imaged that terminate upward beneath anticlines

(i.e., blind faults), and those that break the sea floor to form seafloor scarps.

In the rapidly uplifting (≤ 10 mm/yr) and eroding central Southern Alps (Fig. 1A),

geodetic (Beavan et al. 2010a; Wallace et al. 2007) and geologic (Cox & Findlay 1995; Cox

& Sutherland 2007; Cox et al. 2012) data indicate that active faults must be present, but their

activity cannot be characterised in detail because Late Quaternary scarps and markers are not

preserved as a result of rapid erosion in a steep mountainous environment that has been

subject to high rates of annual rainfall and repeated glaciations. We address this gap in the

data by adding indicative fault zones, typically 5-15 km apart perpendicular to strike, that

approximately correspond to the known faults in bedrock as currently mapped (Cox & Barrell

2007; Rattenbury et al. 2010) and which are considered likely, based on structural or

geomorphic observations, to have experienced late Quaternary activity (Cox et al. 2012). The

names of most of the Southern Alps fault zones (as well as some others in this model,

particularly in offshore Bay of Plenty) are informal names that do not necessarily correspond

to published active or bedrock fault names.

Active fault zone parameters

Active fault zone parameters defined in this model are: (i) dip; (ii) dip-direction; (iii)

sense of movement; (iv) rake; (v) net slip rate; and (vi) quality code. Strike is not explicitly

tabulated, but can be derived from the fault zone traces. Table 1 provides a description of

each parameter and describes in general terms the sources of information used to assign

values for each parameter. The quality codes are further defined in Table 2. For the numerical

parameters dip, rake, and net slip rate, uncertainties are quantified by providing three values

for each, a minimum value, a maximum value, and what is considered to be the best (most

likely) value. The minimum and maximum values are inferred to approximate 95%

confidence bounds, in a qualitative rather than statistically rigorous way. The best value may

in some cases be the mean of several site-specific measurements, or be the median between

maximum and minimum values. In other cases, the best value is calculated from the best-

constrained site-specific offset and age data, and may return a value anywhere between the

maximum and minimum (e.g. for a particular fault zone, a value towards the maximum may

be considered most likely). As a result there is not necessarily a symmetrical distribution of

uncertainty between maximum and minimum values. Specific details of the derivations of

each parameter for individual active fault zones are provided by Litchfield et al. (2013).

Page 7: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

7

Selected Transects of Cumulative Net Slip Rate

To aid with our description of the fault kinematic properties of each tectonic domain in the

next section, we have calculated cumulative net slip rates for representative transects across

each domain (Fig. 2). The transects were selected to: (i) cross fault zones with the best slip

rate data; (ii) be close to perpendicular to the general strike of the fault zones in the domain;

and (iii) be close to parallel within each domain (Fig. 2). The cumulative net slip rate (the

sum of net slip rates for faults crossing each domain irrespective of fault orientation or rake)

is considered to reflect seismic moment release at the surface across the domain (rather than,

for instance, horizontal extension or contraction). The cumulative net slip rate uncertainties

were calculated from the root mean square of individual fault zone slip rate uncertainties. In

addition, the cumulative slip rate was assigned an overall high, medium, or low confidence

ranking, which is calculated from the quality codes assigned to each fault zone (high broadly

corresponds to code 1, medium to 2, and low to 4; codes 3 and 5 fault zones are not

included). We make the assumption that slip rates have been constant over the current period

of activity (125,000 or 25,000 years), unless explicitly stated.

Fault kinematics of the tectonic domains

The New Zealand plate boundary can be divided into three main components from north to

south: (i) oblique westward subduction of the oceanic Pacific Plate beneath the continental

Australian Plate east of the North Island (Hikurangi Margin); (ii) transpression (dextral) and

oblique continent-continent convergence in the South Island (South Island continental

transpression zone); and (iii) northeastward subduction of the oceanic Australian Plate

beneath the continental Pacific Plate southwest of the South Island (Puysegur Margin).

Within this overall framework, 15 tectonic domains have been recognised, based on

subjective geographic groupings of active faults that have similar geometries and kinematics

(Fig. 2). Many of these domains have been defined in previous studies (e.g., Berryman &

Beanland 1991; Stirling et al. 2002, 2012) and in particular, the domains in this study are

slightly modified from those in the 2010 National Seismic Hazard Model (Stirling et al.,

2012). Together, domains 1–6 make up the Hikurangi Subduction Margin, domains 7–12

make up the South Island continental transpression zone; and domains 13–15 make up the

Puysegur - Fiordland Margin. The characteristics of the fault zones in each of these domains

and the present state of knowledge of domain kinematics (mostly from previous studies) are

summarised briefly in the following sections, generally from north to south. Figures 4 and 5

Page 8: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

8

show the fault zones colour-coded by dominant sense of movement and slip rate,

respectively.

Extensional western North Island fault zones (domain 1)

Domain 1 contains relatively few active fault zones. In the north, two fault zones are the only

currently active faults of the Hauraki Rift (Fig. 1B) (Hochstein & Ballance 1993). These

north-northwest striking fault zones have normal senses of movement and net slip rates of

0.1–0.3 mm/yr (e.g., M. Persaud, P. Villamor & K. Berryman unpubl. data 2006).

Northeast-striking active fault zones in western Taranaki (labelled W on Fig. 1B)

appear to have purely normal senses of movement and net slip rates of 0.1–1.5 mm/yr (e.g.,

Nodder 1994; Townsend et al. 2010; Mouslopoulou et al. 2012). These form the southern part

of the mainly submarine Taranaki Basin (Fig. 1A). Farther east, at the western margin of the

Wanganui Basin (labelled E on Fig. 1B), a series of north-northeast striking fault zones also

have normal senses of displacement, although several have an additional minor component of

dextral strike-slip. Net slip rates on individual fault zones are 0.02–0.2 mm/yr (e.g., Pillans

1990b). The Wanganui Basin active faults are interpreted to be bending-moment normal

faults on uplifted basement blocks (Pillans 1990b).

A cumulative net slip rate of 3.2 +0.5/-1.1 mm/yr is calculated along the transect

shown in Fig. 2. This rate is significantly higher than the slip rates on fault zones in the

Hauraki Rift to the north (maximum 0.3 ± 0.1 mm/yr).

The active tectonics in domain 1 indicates back-arc extension distal to the present day

volcanic arc. The Hauraki Rift is largely a relict structure related to the now-extinct

Coromandel Volcanic Arc (Hochstein & Ballance 1993; Fig. 1B), and its relict nature may

account for the low slip rates on faults there. Roll-back of the Pacific Plate subducted slab

and southward motion of the southern termination of subduction and mantle corner flow has

been proposed as a mechanism for relatively large slip rates in the southern part of the

domain (Giba et al. 2010).

Extensional Havre Trough – Taupo Rift fault zones (domain 2)

The southern part of domain 2 is the Taupo Rift (also referred to as the Taupo Fault Belt after

Grindley 1960), which is situated within the active volcanic arc, the Taupo Volcanic Zone

(e.g., Cole et al. 1995; Rowland & Sibson 2001). Offshore from the Bay of Plenty coast, the

Taupo Rift broadens westward into the southern Havre Trough (Fig. 1A) (Wright 1993). In

this study, only faults within the southern 100 km of the Havre Trough have been presented.

Page 9: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

9

Fault zones within the Havre Trough - Taupo Rift mostly strike northeast and are

typically short (c. 4–33 km) and closely spaced. All appear to have a normal sense of

movement (Fig. 4). Net slip rates on individual fault zones are thought to range from 0.05 to

4 mm/yr (Fig. 5), with the highest well-constrained slip rate, at 3.5 mm/yr, in the centre of the

Taupo Rift (Villamor & Berryman 2001). A northward increase in extension rates in this

domain has been previously recognised, from 2.3 ± 1.2 mm/yr south of Mount Ruapehu (Fig.

1B) (Villamor & Berryman 2006b), to 4.4 +2.4/-1.9 mm/yr near Rotorua (Villamor &

Berryman 2001) to 13 ± 6 mm/yr immediately north of the Bay of Plenty coast (Lamarche et

al. 2006). This trend is also expressed by a northward increase in cumulative net slip rate for

the domain 2 transects shown in Fig. 2.

The Havre Trough is interpreted to be in an incipient phase of distributed and

disorganised oceanic crustal accretion and spreading (Wysoczanski et al. 2010). In the Taupo

Rift, two areas devoid of fault zones coincide with the locations of two major volcanic

centres, the Okataina and Taupo Volcanic Centres (Fig. 1B), where much of the back-arc

extension may be locally accommodated by magmatic (e.g., dike intrusion) processes (e.g.,

Seebeck & Nicol 2009; Cole et al. 2010; Rowland et al. 2010). The northward increase in

extension rates is consistent with clockwise rotation of the eastern North Island in response to

slab rollback and pinning of the southern edge of subduction against the Chatham Rise (Fig.

1B) (e.g., Beanland & Haines 1998; Wallace et al. 2004, 2007).

Contractional Kapiti – Manawatu fault zones (domain 3)

The contractional Kapiti – Manawatu fault zones lie southwest of the Taupo Rift, and

northwest of the North Island Dextral Fault Belt. The majority of domain 3 fault zones are

steep reverse structures (Fig. 4) which are likely reactivated Pliocene – early Pleistocene

normal faults (Lamarche et al. 2005). Many also have blind tips associated with hanging wall

anticlines and footwall synclines (Melhuish et al. 1996; Jackson et al. 1998; Lamarche et al.

2005).

Fault zone lengths range from c. 8 to 91 km, but the shortest are likely underestimated

because they transect highly erodible Plio-Pleistocene sediments. Net slip rates range from

0.04 to 3.0 mm/yr (Fig. 5), with the highest occurring on the submarine Mascarin Fault (3

+0.5/-2 mm/yr) (Nodder et al. 2007). Cumulative net slip rates increase northwards from 2.5

+0.3/-0.6 to 5.4 +0.6/-1.6 mm/yr (Fig. 2). The lower rate in the south is thought to be due to

the fault zones there having a greater dextral component than further northeast and

contractional motion partitioned onto reverse faults in the southern Hikurangi subduction

Page 10: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

10

margin forearc (domain 5). This apparent northward increase in slip rate is however

inconsistent with geodetic modelling that predicts a northwards decrease in shortening

(Wallace et al. 2012).

Domain 3 fault zones uplift the eastern margin of the otherwise subsiding Wanganui

Basin (Fig. 1A). The basin has been inferred to be subsiding as a result of high friction on the

plate interface, producing a downward pull and flexure of the overriding Australian Plate

(Stern et al. 1992). Reverse-slip on these fault zones may also have contributed to subsidence

through crustal shortening and lithospheric loading (Lamarche et al. 2005).

North Island Dextral Fault Belt (domain 4)

The North Island Dextral Fault Belt (also referred to as the North Island Fault System,

Mouslopoulou et al., 2007a) is a c. 470 km long series of fault zones within and along the

eastern margin of the North Island’s axial ranges (Fig. 1A). Fault zones at the southern end of

the system strike northeast and are predominantly dextral strike-slip at the ground surface

(Fig. 4). Further north, the faults become progressively more oblique with an increasing

component of extension. Where they intersect and terminate against the eastern margin of the

Taupo Rift (domain 2), the northerly-striking faults in the North Island Dextral Fault Belt

(domain 4) slip with approximately equal components of dextral and normal slip (Fig. 4)

(Mouslopoulou et al. 2007a,b; 2009).

Accompanying the northward change in slip obliquity is a decrease in overall slip rate

(Van Dissen & Berryman 1996; Beanland & Haines 1998; Mouslopoulou et al. 2007b) (Figs.

2 and 5). Net slip rates on individual fault zones are up to 11 mm/yr (dextral) in the south

(Carne et al. 2011), but are no more than 2.5 mm/yr (dextral-normal) in the north

(Mouslopoulou et al. 2007b). The four transects in Fig. 2 show that the cumulative net slip

rate decreases northward from 22.3 +1.8/-3.8 to 14.4 ± 1.7, 7.1 +0.3/-0.6, and 6.2 ± 1.1

mm/yr.

The North Island Dextral Fault Belt accommodates much of the margin-parallel

(dextral) plate motion (Walcott 1987; Cashman et al. 1992; Van Dissen & Berryman 1996;

Beanland & Haines 1998; Mouslopoulou et al. 2007a; Nicol & Wallace 2007). The fault

zones are inferred to splay upwards off the Hikurangi subduction thrust and the 1855 rupture

of the Wairarapa Fault at the southern end of the system may have involved slip on the

subduction interface (Darby & Beanland 1992; Rodgers & Little 2006). The northward

decrease in slip rates compliments the northward increasing rates of normal slip across the

Page 11: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

11

adjacent Taupo Rift to the west (southern domain 2), a displacement transfer that is

associated with clockwise rotation of the eastern North Island (e.g., Wallace et al. 2004).

The southern end of the North Island Dextral Fault Belt is along strike from strike-slip

fault zones of the Marlborough Fault System, but the two are not directly connected beneath

Cook Strait (Fig. 1A) (Pondard & Barnes 2010). The discontinuity in Cook Strait separates

more northerly-striking fault zones in the North Island Dextral Fault Belt from more easterly-

striking fault zones in the northern part of the Marlborough Fault System. This crustal scale

hinge or kink between the two dextral-slip fault systems is thought to transfer some of the slip

on the Marlborough Fault System northward onto the Hikurangi subduction thrust (Wallace

et al. 2012).

Hikurangi subduction margin forearc (domain 5)

The Hikurangi subduction margin forearc is a relatively wide domain, extending westward

from the seafloor trace of the Hikurangi subduction thrust to the North Island Dextral Fault

Belt and the Havre Trough. The majority of the fault zones are thrust faults striking northeast

in the submarine accretionary wedge, and are associated with hanging wall anticlines and

footwall synclines (e.g., Barnes et al. 2002a, 2010). In the south there are some easterly-

striking dextral strike-slip fault zones, while a series of short northeast-striking normal fault

zones occur in the northern onshore portion (Fig. 4).

Fault zone lengths range from c. 6 to 146 km, the shortest being the Raukumara

Peninsula normal fault zones and the longest the submarine thrust and strike-slip fault zones

at the south end of the domain. Net slip rates are estimated to range from c. 0.1 to 5 mm/yr

(Fig. 5), with the lowest being the Raukumara Peninsula (Fig. 1B) normal faults (Berryman et

al. 2009). The highest slip rate occurs on the Lachlan Fault (4.5 ± 2 mm/yr), a major

nearshore thrust fault uplifting Mahia Peninsula (Berryman 1993; Barnes et al. 2002a;

Mountjoy & Barnes 2011). Cumulative net slip rates for three transects which cross the

entire domain are, from north to south, 13.6 +3.4/-2.5, 9.7 ± 1.6 and 14.4 +2.5/-2.9 mm/yr

(Fig. 2).

Uplift of the inner forearc (coastal ranges) is primarily driven by subduction of the

buoyant, oceanic, Hikurangi Plateau (Fig. 1A) (Davy 1992; Kelsey et al. 1995; Litchfield et

al. 2007), with contributions from sediment underplating and localised upper plate faulting

(Walcott, 1987; Reyners et al. 1999, 2006; Nicol et al. 2002; Formento-Trigilio et al. 2003).

The fault zones in this domain are interpreted to splay upwards from the subduction thrust

(Lewis & Pettinga 1993; Henrys et al. 2006; Barker et al. 2009; Barnes et al. 2010; Pedley et

Page 12: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

12

al. 2010). The southern dextral strike-slip fault zones are inferred to facilitate slip transfer

between the northern Marlborough Fault System (domain 8) and either the southern North

Island Dextral Fault Belt (domain 4) or the Hikurangi subduction thrust (domain 6) (Wallace

et al. 2004, 2012). The normal fault zones in Raukumara Peninsula are interpreted to be

shallow, secondary extensional structures formed as a consequence of rapid uplift, and/or

trenchward collapse, of the northern forearc (Thornley 1997; Berryman et al. 2009). An

absence of consistent along-strike variations in the cumulative slip rates indicate that

convergence variability along the margin (Wallace et al. 2004) is largely accommodated by

the subduction interface and associated frontal thrusts.

Hikurangi subduction thrust (domain 6)

The Hikurangi subduction thrust extends southward from the southern end of oceanic-oceanic

crust subduction at the Kermadec Trench (Fig. 1A), where convergence rates exceed 60

mm/yr, to offshore northeastern Marlborough, where plate motion is predominantly

transferred to the Marlborough Fault System. The thrust intersects the seafloor along the

western edge of the Hikurangi Trough (Fig. 1A), and is expressed by frontal thrust-faulted

anticlinal ridges and a proto-thrust zone (Lewis & Pettinga 1993; Barker et al. 2009; Barnes

et al. 2010). Following Wallace et al. (2009) and Stirling et al. (2012), the thrust is divided

into three sections along strike (Fig. 1B), on the basis of margin characteristics, historical

seismicity and the distribution of interseismic coupling and slow slip events (summarised by

Wallace et al. 2009). The section boundaries are only approximately located however,

because of the relatively low resolution of the location of these changes at depth. The model

of Stirling et al. (2012) focused on the seismogenic part of the thrust, for the purposes of

seismic hazard modelling, whereas our model addresses all parts of the thrust down-dip from

its trace in the Hikurangi Trough.

All sections are inferred to be largely dip-slip, as the thrust takes up a significant

proportion of the plate-normal motion (Barnes et al. 1998; Nicol & Beavan 2003; Wallace et

al. 2004, 2009; Nicol & Wallace 2007). The concave (e.g., Ansell & Bannister 1996), and in

places kinked, profile of the thrust in cross-section (Henrys et al. 2006; Barker et al. 2009)

has led us not to assign single dip values to the fault. The full net slip rate on the thrust (i.e.,

including seismogenic and creeping components) ranges from 54 ± 6 mm/yr in the north, 44

± 7 mm/yr in the centre, to 25 ± 5 mm/yr in the south (Wallace et al. 2004; Stirling et al.

2012).

Page 13: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

13

Contractional northwestern South Island fault zones (domain 7)

Contractional northeast-striking fault zones produce a range and basin topography north and

west of the transpressional Alpine Fault. Many of these fault zones are reactivated Mesozoic

or early Cenozoic normal faults (Laird 1968; Lihou 1992; Bishop & Buchanan 1995; Ghisetti

& Sibson 2006) and many have associated hanging wall anticlines and footwall synclines

(Suggate 1987; Nicol & Nathan 2001; Ghisetti & Sibson 2006). Despite their prominent

geomorphic expression, only some of these fault systems show evidence for Late Pleistocene

activity (Suggate 1987, 1989; Nicol & Nathan 2001), implying that more faults could be

potentially active than currently recognised (e.g., Stafford et al. 2008; Sibson & Ghisetti

2010).

Fault zone lengths range from c. 38 to 140 km, based mainly on bedrock geology, and

sense of movement is primarily reverse (Fig. 4). Net slip rates are poorly constrained (e.g.,

Stafford et al. 2008), and generally low (Fig. 5). Cumulative net slip rates across two

transects in Fig. 2 are 0.7 ± 0.3 (north) and 1 ± 0.4 (south) mm/yr.

The northwestern South Island fault zones have, through the Late Cenozoic, taken up

some of the plate-normal component of continental convergence in the South Island.

Deformation is particularly concentrated northwest of a major bend in the Alpine Fault,

where some of the reverse fault zones likely form footwall shortcut thrusts and backthrusts to

the Alpine Fault (Ghisetti & Sibson 2006).

Strike-slip Marlborough Fault System (domain 8)

The Marlborough Fault System is a series of predominantly continental strike-slip fault zones

(Fig. 4) in the northern South Island, which link the Hikurangi Subduction thrust (at depth) to

the Alpine Fault. Although the majority of fault zones are chiefly dextral strike-slip, several

also have measurable dip-slip motion (Fig. 4), most commonly up-to-the-northwest. For

example, reverse faults contribute to uplift of the Kaikoura Ranges (e.g., Van Dissen & Yeats

1991; Nicol & Van Dissen 2002) and parts of the continental shelf (Barnes & Audru 1999). A

normal component of slip on the Wairau and Cloudy faults create transtensional basins in

Cook Strait (Fig. 1A) (Pondard & Barnes 2010), while extension on the Hanmer Fault

contributes to localised subsidence of Hanmer Basin along the Hope Fault (Fig. 1B) (Wood et

al. 1994) (Fig. 4).

Whilst fault zone lengths range from c. 18 to 145 km, the majority of the structures

are elements of longer segmented strike-fault systems (e.g., the Hope Fault, Langridge et al.

2003, which has a total length of c. 300 km as defined here). Net slip rates range from c. 0.1

Page 14: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

14

to 25 mm/yr (Fig. 5), with the highest being the eastern section of the Hope Fault (Langridge

et al. 2003). Cumulative net slip rates along transects (Fig. 2) are 23.5 +3.4/-3.7 (southwest)

and 37.8 ± 3.4 (northeast) mm/yr, showing a northeastward increase in the transfer of slip

from the Alpine Fault to the Marlborough Fault System.

The Wairau Fault, sometimes considered the northern part of the Alpine Fault based

on the continuity of the surrounding bedrock and total offset (e.g., Berryman et al. 1992), is

included in this domain to reflect its current role in the Marlborough Fault System. Domain 8

lies above the subducted Pacific Plate, but its constituent faults likely intersect the underlying

plate interface only in the east, where the interface is shallowest (Reyners & Cowan 1993;

Wannamaker et al. 2009; Eberhart-Phillips & Bannister 2010; Wallace et al. 2012). The age

of the Marlborough Fault System has been interpreted to decrease southeast-ward, probably

reflecting southward migration of the Hikurangi subduction thrust (e.g., Yeats & Berryman

1987; Little et al. 1998; Wallace et al. 2007).

Contractional North Canterbury fault zones (domain 9)

Lying southeast of the Marlborough Fault System, this domain is characterised by mixed

fault zone orientations and slip types (Fig. 4). In the northeastern, offshore part of the

domain, reverse and thrust fault zones with associated hanging wall anticlines and footwall

synclines largely verge eastward. They cut Mesozoic basement and possibly splay upwards

from the Hikurangi subduction thrust (Barnes et al. 1998; Wallace et al., 2012). Onshore, in

the centre of the domain, reverse fault zones with associated hanging wall anticlines and

footwall synclines, verge mainly northwest in the coastal region and southeast further inland

(Al-Daghastani & Campbell 1995; Nicol et al. 1995; Barnes 1996; Pettinga et al. 2001;

Litchfield et al. 2003). Localised northerly and easterly striking fault zones are probably

reactivated Cretaceous normal faults (Nicol & Wise 1992; Nicol 1993; Litchfield et al. 2003).

The fault zones could extend upward from a mid-crustal detachment identified in

microseismicity (Reyners & Cowan 1993; Campbell et al. 2012). Fault zones in the western

part of the domain are inferred to be predominantly transpressional (Cox et al. 2012), but

some northwest-striking fault zones form complex transfer structures between northeast-

striking faults (Abercrombie et al. 2000; Robinson & McGinty 2000). Fault zones in the far

west of the domain may splay upwards from the southeast-dipping Alpine Fault (Pettinga et

al. 2001). The exact position of the southwest boundary between domains 9 and 11 (in the

Southern Alps) is somewhat subjective, but has been defined based on a slight change in

Page 15: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

15

strike (generally more easterly to the north) and sense of movement (some dextral motion to

the north).

Fault zone lengths range from c. 7 to 90 km, with the longest, the Porters Pass – Grey

fault zone, comprising at least two segments (Howard et al. 2005). Slip rates are mostly

poorly constrained; they range from c. 0.1 to 3.5 mm/yr, with the highest rate estimated for

part of the Porters Pass fault (3.5 +0.6/-0.4 mm/yr) (Howard et al. 2005), and the lowest on

submarine faults in the east (Barnes 1996; Barnes et al. 1998) (Fig. 5). Cumulative net slip

rates calculated along transects shown in Fig. 2 decrease northeastward from >5.6 +0.9/-0.6

to 3.1 +1/-0.8 and 2.3 ± 0.7 mm/yr, although the slip rate on the westernmost transect is a

minimum because it includes Southern Alps fault zones with no assigned slip rates.

The North Canterbury fault zones are considered to be relatively youthful (<1 Myr

e.g., Nicol et al. 1994; Barnes 1996; Cowan et al. 1996; Pettinga et al. 2001; Litchfield et al.

2003), and represent an encroachment of plate boundary deformation into the region from the

north. The evolution of this domain means the southeast boundary may be migrating south,

the possibility exists that the faults that ruptured in the 2010-2011 Canterbury Earthquake

Sequence (e.g., Gledhill et al. 2011; Kaiser et al. 2012) could be included in this domain (e.g.,

Campbell et al. 2012) rather than the lower strain domain 11, as proposed here (see below).

Extensional North Mernoo fault zones (domain 10)

The North Mernoo fault zones are entirely submarine (Fig. 4) and comprise east-striking fault

zones on the northwestern edge of the submerged continental Chatham Rise (Fig. 1A).

Although mapping is limited by data coverage, the larger structures are inferred to be

between c. 20 and 65 km long (Barnes 1994a). Sense of movement is inferred to be normal

(Fig. 4) with net slip rates, assigned on the basis of low long-term extension (Pliocene-

Quaternary), of c. 0.1 to 0.25 mm/yr (Fig. 5). The cumulative net slip rate across this domain

is 1 ± 0.4 mm/yr (Fig. 2).

The current phase of activity started in the Late Miocene, and involved a reactivation

of Cretaceous and Eocene (normal) fault systems (Barnes 1994b). Two kinematic models to

explain this extension have been proposed. One is minor buckling at the northern end of the

southern Marlborough Fault System, which results in slivers of the Chatham Rise being

spalled off as the Chatham Rise slides past southern Marlborough (Anderson et al. 1993). The

other model invokes flexural extension (bending-moment faulting) of the edge of the

Chatham Rise at the southern end of the Hikurangi Subduction Zone (Barnes 1994a).

Page 16: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

16

Contractional southeastern South Island fault zones (domain 11)

Domain 11 is the largest and comprises all of the active faults of onland South Island

southeast of the Alpine Fault and south of the North Canterbury fault zones (domain 9) (Fig.

2). The majority of fault zones strike north to northeast, but there is an additional set of

northwest-striking faults within much of the domain, producing an orthogonal fault pattern.

Many of these structures are range-bounding reverse faults (Fig. 4) with hanging wall

anticlines and footwall synclines (e.g., Beanland et al. 1986; Jackson et al. 1996; Markley &

Tikoff 2003). Fault zones close to the Alpine Fault have a significant component of dextral

strike-slip motion (Sutherland 1995; Ballard 1989; Cox et al. 2012) and there are two

predominantly strike-slip fault zones of note. The northwest-striking Wharekuri Fault (Barrell

et al. 2009) in the Waitaki Valley (Fig. 1B) is the only dominantly sinistral fault zone in the

model, and may facilitate eastward widening of the area of contraction east of the Alpine

Fault (see below). The east-striking Greendale Fault in the northeast corner ruptured in the

2010 Darfield Earthquake (e.g., Quigley et al. 2010, 2012; Gledhill et al. 2011) and is likely

to be a reactivated Cretaceous normal fault (Browne et al. 2012; Campbell et al. 2012;

Ghisetti & Sibson 2012; Jongens et al. 2012).

Fault zone lengths range from c. 15 to 90 km, while the range-bounding reverse fault

zones are mostly 20–70 km long. Net slip rates are 0.01 to 2.2 mm/yr (Fig. 5), with the

highest being the Fox Peak Fault (2.2 +2.4/-1.8 mm/yr) (Upton et al. 2004). Cumulative net

slip rates across the three transects shown in Fig. 2 are >3.9 +3.1/-1.9, 3.9 +1.9/-1.1, and 1.2

± 0.6 mm/yr from north to south, although the northern (and possibly the central) transect slip

rate is a minimum because it includes fault zones within the Southern Alps with no assigned

slip rates. Conversely, the central cumulative net slip rate may be a maximum as there is

some evidence for the activity of some of these (Beanland & Berryman 1989; Berryman &

Beanland 1991), as well as other east Otago (Litchfield & Norris 2000), fault zones to be

episodic.

The convergence zone in the north and centre of this domain can be described as an

asymmetric two-sided orogen, with the higher rates of erosion in the west (inboard) focusing

deformation on the southeast-dipping Alpine Fault (Koons 1990; Norris et al. 1990).

Variations in the width of the outboard zone (east) are likely to be a function of lithology,

with the widest part in the centre of the domain corresponding to thicker and weaker

schistose crust (Upton et al. 2009). Northern outboard fault zones (north of Oamaru; Fig. 1B)

likely splay upwards from a mid-crustal shear zone (e.g., Wannamaker et al. 2002; Van

Page 17: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

17

Avendonk et al. 2004), whereas central fault zones (in Otago, the area between approximately

Oamaru and Papatowai in Fig. 1B) may splay upwards from a thick ductile lower crust

(Upton et al. 2009). Many of the outboard fault zones are reactivated Cretaceous normal

faults (Mutch & Wilson 1952; Norris et al. 1987; Ghisetti et al. 2007). Inboard (northwest-

dipping) fault zones in the centre and north of the domain are likely to be back-thrusts off the

southeast-dipping Alpine Fault (Norris et al. 1987, 1990; Cox & Sutherland 2007).

Alpine Fault (domain 12)

The Alpine Fault is a dextral strike-slip fault, defined here to be c. 740 km in length, forming

the western (inboard) side of the South Island continental-convergence zone. The fault links

the Marlborough Fault System with the Puysegur subduction thrust (Fig. 2).

In this paper, we divide the Alpine Fault into 6 sections (Fig. 1B) based on changes in

slip rate and geometry along strike. The southern sections are steeply-dipping to the

southeast, almost pure dextral strike-slip (Barnes et al. 2001, 2005), and include the highest

net slip rate (31 ± 3 mm/yr) found on the fault (Barnes 2009). The rate decreases northwards

to 23 ± 2 mm/yr in south Westland (Sutherland et al. 2006). The central sections, bounding

the western side of the Southern Alps, have a moderate dip to the southeast (Davey et al.

1998). Near Aoraki/Mt Cook (Fig. 1B) the dip-slip rate is locally up to 10 mm/yr (Norris &

Cooper 2001) and inferred uplift rates to the east of the fault are >8 mm/yr (Batt 2001; Little

et al. 2005). The dextral slip rate decreases northwards where slip is transferred onto the

Marlborough Fault System, particularly the Hope Fault (Langridge et al. 2010).

The Alpine Fault is inferred to be inherited from an Eocene rift and passive margin

that formed during the initiation of the Cenozoic plate boundary through New Zealand

(Kamp 1986; Sutherland et al. 2000). It has accrued displacement since the Late Oligocene or

Early Miocene as South Tasman ocean crust was subducted beneath Fiordland (Fig. 1A),

while continental crust of the Australian Plate (Challenger Plateau; Fig. 1A) acted as an

indentor into weaker continental crust to the southeast (Sutherland et al. 2000, 2009). The

absence of active faults west of the central Alpine Fault (southern domain 7) provides

supporting evidence that the Australian Plate is acting as a rigid indentor.

Western Fiordland Margin - Caswell High fault zones (domain 13)

Domain 13 fault zones are entirely submarine and lie west of the southern part of the Alpine

Fault. Fault zones form two parallel sets. In the west are a series of normal fault zones along

Page 18: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

18

the edge of the Caswell High (Fig. 1A) and in the east are west-verging reverse fault zones on

the east side of the Fiordland Basin (Fig. 4).

Fault zone lengths range from c. 25 to 75 km. Net slip rates, although poorly

constrained, are inferred to be low (0.4–3 mm/yr; Fig. 5) on the basis of offset seismic

reflectors tied to dated samples (Barnes et al. 2002b). Cumulative net slip rates show a

northward decrease in slip rate from 7.4 ± 2.2 to 3.8 ± 1.3 mm/yr (Fig. 2). This decrease

continues further north as the zone narrows to just a single fault zone, the Barn Fault, which

forms part of a set of submarine, landward-dipping reverse faults which broadly correspond

to the South Westland Fault Zone and uplifts Pleistocene marine deposits along the west

coast (Sutherland & Norris 1995; Sutherland et al. 2007).

Fault zones in this domain are relatively young, commencing activity in the Pliocene

prior to 3 ± 1 Myr, but mostly developed during the Quaternary (Barnes et al. 2002b). The

thrust faults form part of a narrow wedge accommodating some of the margin-normal

component of plate motion (Delteil et al. 1996; Lebrun et al. 2000; Wood et al. 2000; Barnes

et al. 2002b). The Caswell High fault zones are the result of flexural uplift caused by the tight

bending of the Australian plate as a result of its subduction beneath Fiordland (Fig. 1A)

(Delteil et al. 1996; Malservi et al. 2003). Many of the faults may be reactivated Cretaceous

or Eocene normal faults (Sutherland et al. 2000; Wood et al. 2000).

Puysegur Ridge – Bank strike-slip fault zones (domain 14)

Puysegur Ridge (Fig. 1A), sitting above the Puysegur subduction thrust, has a faulted gully

parallel to the ridge crest that is inferred to be the surface expression of a major strike-slip

fault system, the Puysegur Ridge Fault (Collot et al. 1995). At the northern end of the ridge,

at least two splays can be traced into the Snares Trough (Fig. 1B) (Lamarche & Lebrun

2000). North of the Snares Trough, two fault zones extend across and displace wave

planation surfaces of probable Quaternary age on Puysegur Bank (Fig. 1B) (Melhuish et al.

1999).

Fault zone lengths range from c. 66 to 298 km, although it is likely that the Puysegur

Ridge Fault is segmented. The slip rates on these fault zones are unknown, because the age of

the wave-cut surface is poorly constrained and the amount of strike-slip displacement cannot

be measured on the surface. The rates for these faults are therefore assigned as a portion of

the plate convergence rate and cumulative slip rate transects have not been compiled.

The morphology and tectonic setting of Puysegur Ridge is very similar to Macquarie

Ridge to the south, where a structure analogous to the Puysegur Ridge Fault ruptured with a

Page 19: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

19

dextral strike-slip focal mechanism in the 1989 Macquarie Ridge earthquake (Mw 8.2)

(Anderson 1990). The northern (Snares) fault zones transfer much of the margin-parallel

motion from the Puysegur Ridge Fault back to the subduction thrust at shallow (<10 km)

depths (Lamarche & Lebrun 2000).

Puysegur subduction thrust (domain 15)

The Puysegur subduction thrust is considered as a single fault zone which dips east from the

Puysegur Trench (Fig. 1A). The seafloor trace extends southward from Resolution Ridge

(Fig. 1A), to the Macquarie Ridge Complex (off Fig. 1, to the south) where plate boundary

motion is primarily dextral strike-slip (e.g., Massell et al. 2000).

A series of earthquakes on the northern end of the thrust, including the 2009 Mw7.8

Dusky Sound Earthquake, show that the sense of movement on the Puysegur subduction

thrust has a component of dextral strike-slip, and that the azimuth of the slip vector is

precisely parallel to that predicted by relative plate motions (Doser et al. 1999; Reyners et al.

2003; Beavan et al. 2010b; Fry et al. 2010). The parallelism of the Dusky Sound Earthquake

slip to the plate motion vector further indicates that the thrust takes up the majority of the

plate motion, and a net slip rate of 27 +10/-7 mm/yr has been assigned to reflect this.

Subduction on the Puysegur subduction thrust initiated at c. 10 Ma along a pre-

existing fracture zone within the southeast Tasman oceanic crust, following a major

reorganisation of the overall Pacific-Australia plate boundary (Lamarche et al. 1997; Lebrun

et al. 2000).

Discussion

Fault zone densities and lengths

Statistics have been derived for the numbers, densities and lengths of the fault zones in the

updated nation-wide model. These numbers are considered to be first-order. They are likely,

for example, to comprise fault lengths that are in many cases shorter than the actual lengths

due to erosion or burial. Moreover, some domain boundaries (particularly their outer,

offshore, edges) are necessarily arbitrary and there has been considerable simplification and

interpretation of fault zone traces (e.g., Fig. 3). Nevertheless, we think our compilations

reveal some important kinematic features of the New Zealand plate boundary zone, and also

highlight issues of potentially unidentified faults, which are discussed further in the model

completeness section.

Page 20: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

20

A total of 635 active fault zones or fault zone sections are included in this model

(Figs. 1, 4, 5). The majority (416; 66%) are in North Island domains despite the total

geographic area of North and South Island domains being similar (Table 3). In part this

reflects the large number (196) of closely-spaced fault zones in the Havre Trough - Taupo

Rift domain, which contains 47% of North Island fault zones and 31% of the total fault zones

(Table 3). Many of the fault zones in this domain are relatively short however, and the

domain with the greatest cumulative fault zone length (3265 km) is the contractional

southeastern South Island domain (Table 3). Other domains with high fault zone numbers and

cumulative lengths are the Hikurangi subduction margin forearc and North Island Dextral

Fault Belt (Table 3). Terrestrial and submarine fault zones are similar in abundance to one

another: 299 (47%) fault zones are entirely terrestrial, 292 (46%) are entirely submarine, and

44 (7%) cross the coast. Of the entirely submarine fault zones 146 (50%) are in the offshore

Taupo Rift and southern Havre Trough.

Translating the number and length of fault zones per domain into density (i.e. dividing

the numbers and lengths by domain area), shows that indeed the Havre Trough – Taupo Rift

has the greatest density of fault zones, while the North Island Dextral Fault Belt has the

longest fault zones per area (aside from the Alpine Fault domain). Other domains of high

density are the Kapiti-Manawatu and Marlborough Fault System domains. Not surprisingly,

the least-densely faulted domains are situated further away from the main locus of plate

boundary strain (excluding the Alpine Fault, and the Hikurangi and Puysegur subduction

thrusts). These areas include the western North Island, northwestern South Island,

southwestern South Island, and Puysegur Ridge – Bank domains (Table 3). These domains

may also contain a larger proportion of unidentified active faults (e.g., Sibson & Ghizetti

2010, Ries et al. 2011), as discussed in the model completeness section below.

Fault zone kinematics and slip rates

Figure 4 illustrates that the majority (83%) of fault zones in the model are interpreted to have

a predominantly dip-slip sense of movement. Forty six percent are predominantly normal,

38% are predominantly reverse, 17% are predominantly dextral, and there is only 1

dominantly sinistral fault. Figure 4 also reveals how oblique plate boundary convergence is

partitioned in the upper plate between domains of primarily dip-slip versus strike-slip faulting

that are parallel to one another. However, it should be noted that Figure 4 only shows the

dominant sense of movement, whereas many fault zones have an oblique sense of movement

(supplementary file Tables S1 and S2).

Page 21: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

21

Figure 5 illustrates that the highest slip rates occur along the northern and central

Hikurangi subduction thrust. The transfer of motion from the Alpine Fault onto the Puysegur

subduction thrust and the Puysegur Bank – Ridge fault zones in the south, and into the

Marlborough Fault System, and ultimately the Hikurangi subduction thrust and southern

North Island Dextral Fault Belt, in the north are also clearly shown. Forty-three fault zones

have relatively high slip rates (≥5 mm/yr). The combined length of these fault zones is 3448

km, which is 17% of the total length of fault zones in the model (20,618 km).

A total of 226 fault zones (35% of the total length) have moderate (1-5 mm/yr) slip

rates, and these are generally close to the major tectonic elements of the plate boundary zone

(Fig. 5). Some of the central Southern Alps fault zones may also have moderate slip rates.

Fault zones with low (< 1 mm/yr) slip rates (331; 42% of the total length) dominate in areas

that are distal to the chief plate boundary elements. The majority of low slip rate fault zones

are either reverse or normal (Fig. 4).

The cumulative net slip rates on the transects in Fig. 2 confirm these general patterns

of slip rate distribution across the plate boundary zone with the highest slip rates (not

including the Alpine Fault and the Hikurangi and Puysegur subduction thrusts) occurring in

the Marlborough Fault System, followed by the southern North Island Dextral Fault Belt and

the northern Taupo Rift – Havre Trough. Notable along-strike changes in cumulative slip rate

occur within single domains. These include the rapid northward decrease in cumulative slip

rate along the North Island Dextral Fault Belt, and the parallel northward increase in slip rate

across the Kapiti - Manawatu fault zones and the extensional Havre Trough – Taupo Rift. In

contrast, the slip rate on the Hikurangi subduction thrust decreases strongly to the southwest,

whilst the Alpine Fault rate decreases to the northeast, illustrating a transfer of strain from

these first-order plate boundary structures onto second-order fault zones in between.

Model completeness

Our model has added significant numbers of active faults to previous New Zealand-wide

compilations. However, like other national active fault or fault source databases (e.g.,

Machette et al. 2004; Yoshioka et al. 2005; Basili et al. 2008; Clark et al. 2012), the model is

undoubtedly incomplete. Table 4 summarises our best assessment of the relative level of

completeness of each domain and lists the main areas where we consider the model to likely

be incomplete. Not surprisingly, the two main areas of likely incompleteness are: (i) rapidly

eroding and difficult-to-access mountainous areas; and (ii) some submarine areas where there

is currently insufficient knowledge of likely active faulting.

Page 22: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

22

A more quantitative calculation of model completeness is beyond the scope of this

study, but a preliminary analysis of 532 fault sources in the 2010 version of the New Zealand

National Seismic Hazard Model (Stirling et al. 2012) showed a power-law distribution for

faults with slip rates ≥1 mm/yr (Nicol et al. 2011). Departure from the power-law distribution

can be interpreted to indicate that the seismic hazard model (and therefore the model

presented in this paper given the overlap) is incomplete at slip rates ≤1 mm/yr. The level of

completeness will vary between regions with faults with higher slip rates (e.g., 0.5-1 mm/yr)

most likely to be undersampled where they are exceeded by erosion or deposition rates (e.g.,

the Southern Alps and western Canterbury Plains).

Comparison of geological and geodetic rates

The type of fault parameters presented in this model may be used in studies of seismic hazard

(e.g., Stirling et al., 2012), kinematic deformation (Beavan et al. 2007; Wallace et al. 2012),

and fault interactions (Robinson et al. 2009, 2011; Pondard and Barnes 2010). Comparison of

the permanent deformation, revealed by fault slip rates typically measured over thousands to

tens of thousands of years, with strains recorded by GPS velocities, measured over the last

few decades, places constraints on the spatial distribution, sampling biases and timing of

deformation. We present one such comparison, that of regional horizontal deformation for the

transects shown in Figure 2. Only mean values are calculated here, but considering the

uncertainties in fault zone slip rates in Figure 2, it is only differences of several mm/yr or

more that are considered significant. Only horizontal deformation is compared, as a nation-

wide vertical GPS velocity field is currently unavailable.

Cumulative horizontal fault zone slip rates were obtained by first converting the net

slip rates of individual fault zones to horizontal slip rates, and then into transect parallel and

normal components. These were then summed for each transect and a resultant vector

calculated. Horizontal GPS velocities relative to a fixed Australian plate were calculated from

a velocity field produced using a spline-fitting function to 920 site velocities (Beavan 2012).

Velocity differences were calculated across each transect (i.e., difference between velocities

at the transect endpoints), and were resolved into transect parallel and normal components.

The outer edges of the transects in domain 10 and 13 lie outside the current version (4.0) of

the GPS velocity field, so velocity differences for those transects could not be calculated.

Inspection of Figure 6 and Table 5 shows that for the majority (37 of 47) of transects,

the rates of GPS deformation are higher than the cumulative fault slip rates. This is likely to

be for a variety of reasons, including incompleteness in the fault model in some areas (Table

Page 23: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

23

4), that GPS data includes distributed deformation between the faults in the model and, for

the North Island, because elastic strains in the GPS signal are accruing due to (and in the

future will likely be converted to) slip on the Hikurangi subduction thrust (Nicol and Wallace

2007). Distributed deformation may take the form of large-scale vertical-axis block rotations,

folding associated with fault slip at depth and fault slip that is below the resolution of the

present data (i.e. incomplete sampling of fault slip rates). Incomplete sampling may be

particularly important for transect-normal deformation (i.e. strike-slip faulting) in domains 1,

2, 7, 9 and 11 where fault slip rates are generally low (e.g., <1 mm/yr), fault piercing points

are rare and, for parts of domains 9 and 11, erosion rates are likely to be higher than the slip

rates, removing evidence of fault displacements.

In the predominantly strike-slip North Island Dextral Fault Belt and the Marlborough

Fault System, cumulative fault slip rates are generally significantly higher (e.g., ~20-100%)

than GPS deformation. This difference may partly arise because faults in these domains take

up most, if not all, of the margin-parallel plate motion over geological timescales, while

contemporary deformation is distributed beyond these relatively narrow domains. For

example, geological slip rates are greater than GPS rates in the Marlborough Fault System

(33.2 versus 18.8 mm/yr), while GPS rates significantly exceed geological slip rates in

domains either side (northwestern North Island and North Canterbury). Thetotal transect-

normal deformation across all three domains are similar however, 34.4 (geological slip rates)

and 33.1 mm/yr (GPS), respectively. There is no evidence for episodic activity on any fault

zones in the North Island Dextral Fault Belt and the Marlborough Fault System, so we infer

episodicity is unlikely to account for the higher geological slip rates.

There is a general consistency between the overall sense of motion (contraction or

extension) defined by geological and GPS data for nearly all transects (43 of 47 transects).

This consistency indicates that the contemporary elastic strains will be largely converted to

long-term (thousands to millions of years) permanent deformation in earthquakes (e.g., Nicol

and Wallace 2007). Of the four transects with differing geological and GPS sense of

movement, three (transect M1 in the North Island Dextral Fault Belt, and the southwestern

transects in the Marlborough Fault System and North Canterbury domain) are mainly the

result of sensitivity to the orientation of transects. Both fault and GPS deformation is at a

high angle and therefore a small change in transect orientation results in transect parallel

motion switching from extension to contraction and vice versa. The fourth transect with

differing geological and GPS sense of movement is in the western North Island, where the

Page 24: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

24

GPS velocities indicate transect parallel (WNW-ESE) contraction but the fault zones are all

clearly extensional. There is currently no clear explanation for this difference.

A final general observation is that fault zone slip is inferred to be transect parallel

(i.e., pure normal or reverse) for a number of transects (marked with an * in Figure 6A)

whereas the GPS velocities are always oblique. There may be explanations for some of these

differences, such as the permanent transect normal motion being taken up by block rotations

in the Havre Trough - Taupo Rift, and Alpine Fault locking contributing to the GPS velocities

across southeastern South Island transects. However, in other cases it raises the question as to

whether fault slip is purely dip slip, or if strike-slip components are unrecognised.

There are clearly drawbacks to this 2 dimensional (transect parallel and normal)

comparison, such as the sensitivity of the results to the orientation and endpoints of transects.

To further improve understanding of plate boundary kinematics in the future, the approach of

Wallace et al. (2004, 2007, 2012) could be applied on a New Zealand scale, whereby GPS

velocities, active fault data, and earthquake slip vectors are inverted simultaneously to derive

block rotations and degree of interseismic coupling. Alternatively, both datasets can be

converted to strain rates for direct comparison (e.g., Holt & Haines 1995; Beanland & Haines

1998; Beavan et al. 2007).

Summary and conclusions

A new, regional-scale model of active faulting in New Zealand has been compiled which

contains 635 fault zones, including 631 upper crustal fault zones and 4 subduction thrusts.

Fault zone traces are available for download in shapefile, ascii and kmz formats

(supplementary files Maps S1- S3). One of the biggest additions to this model over previous

compilations is the large number of submarine fault zones, which comprise 52% of the total.

Tabulated fault parameters include: (i) dip; (ii) dip-direction; (iii) sense of movement;

(iv) rake; (v) net slip rate; and (vi) quality code, and are available for download

(supplementary files Table S1 and S2). We endeavoured to obtain, assign, or infer each of

these for all faults, with the main exception being slip rates for the central Southern Alps

faults, where geomorphic markers are generally absent as a result of high rates of erosion.

The fault zones have been grouped into 15 tectonic domains, comprising subjective

geographic groupings of fault zones which have similar kinematics. The current knowledge

of kinematics in each domain have been summarised, primarily from previously published

work.

Page 25: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

25

The majority (83%) of the fault zones have a dip-slip sense of movement; 46% are

normal, 38% are reverse, 17% are dextral, and there is only 1 sinistral fault.

In terms of fault identification and mapping, the model is estimated to be mostly

complete for fault zones with slip rates ≥1 mm/yr. Main areas where the model is likely to be

incomplete are not surprisingly, areas of rapidly eroding and/or difficult to access hills and

mountains, and submarine areas where high resolution seismic and swath data have yet to be

collected. Data is also unavailable from the rapidly eroding central Southern Alps, where

there are mapped faults but unknown slip rates, such that the southeastern South Island

domain has the largest disparities between cumulative fault zone slip rates and differences in

horizontal GPS velocity.

The model should be useful for a range of purposes, and an example is provided

comparing horizontal cumulative fault slip rates with GPS velocity differences parallel and

normal to selected transects crossing each domain. This example has highlighted several

differences which could be explored with more sophisticated kinematic modelling and will

help target future active fault studies.

Supplementary files:

Table S1: Active fault zone parameters and data source references.

Table S2: Microsoft excel file of the active fault zone parameters.

Figures S1–S8: Index and enlarged maps of the active fault zone traces.

Map S1: Shapefile of the active fault zone traces and parameters (attribute table), in NZTM

2000 projection.

Map S2: Ascii file of the active fault zone traces, in WGS84 projection.

Map S3: Kmz file of the active fault zone traces, in Google Earth coordinate system.

Acknowledgements

This work was funded by the New Zealand Science and Technology Contracts C05X0702

and C05X0804 (GNS Science) and C01X0702 (NIWA). We would like to thank Dougal

Townsend and John Begg for their helpful reviews. We would also like to acknowledge

Neville Palmer for the GPS velocity data, Patrick Sweetensen for his work on the references,

and Bella Ansell for assistance with figures.

Page 26: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

26

References

Abercrombie RE, Webb TH, Robinson R, McGinty RJ 2000. The enigma of the Arthur's

Pass, New Zealand, earthquake 1. Reconciling a variety of data for an unusual

earthquake sequence. Journal of Geophysical Research 105(B7): 16119-16137.

Al-Daghastani HSY, Campbell JK 1995. The evolution of the lower Waipara River gorge in

response to active folding in North Canterbury, New Zealand. ITC Journal 3: 245-255.

Anderson HJ 1990. The 1989 Macquarie Ridge Earthquake and its contribution the regional

seismic moment budget. Geophysical Research Letters 17: 1013-1016.

Anderson H, Webb T, Jackson J 1993. Focal mechanisms of large earthquakes in the South

Island of New Zealand: implications for the accommodation of Pacific-Australian plate

motion. Geophysical Journal International 115: 1032-1054.

Ansell JH, Bannister JH 1996. Shallow morphology of the subducted Pacific plate along the

Hikurangi margin, New Zealand. Physics of the Earth and Planetary Interiors 93: 3-20.

Ballard HR 1989. Permian arc volcanism and aspects of the general geology of the Skippers

Range, NW Otago. Unpublished PhD thesis, University of Otago, Dunedin, New

Zealand.

Barker DHN, Sutherland R, Henrys SA, Bannister SC 2009. Geometry of the Hikurangi

subduction thrust and upper plate, North Island, New Zealand. Geochemistry,

Geophysics, Geosystems 10: Q02007. doi:02010.01029/02008GC002153.

Barnes PM 1994a. Continental extension of the Pacific Plate at the southern termination of

the Hikurangi subduction zone: the North Mernoo Fault Zone, offshore New Zealand.

Tectonics 13: 735-754.

Barnes PM 1994b. Inherited structural control from repeated Cretaceous to Recent extension

in the North Mernoo Fault Zone, western Chatham Rise, New Zealand. Tectonophysics

237: 27-46.

Barnes PM 1996. Active folding of Pleistocene unconformities on the edge of the Australian-

Pacific plate Boundary zone, offshore North Canterbury, New Zealand. Tectonics 15:

623-640.

Barnes PM 2009. Postglacial (after 20 ka) dextral slip rate of the offshore Alpine fault, New

Zealand. Geology 37: 3-6.

Barnes PM, Audru JC 1999. Recognition of active strike-slip faulting from high-resolution

marine seismic reflection profiles: Eastern Marlborough fault system, New Zealand.

Geological Society of America Bulletin 111: 538-559.

Page 27: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

27

Barnes PM, Mercier de Lepinay B, Collot J-Y, Delteil J, Audru J-C 1998. Strain partitioning

in the transition area between oblique subduction and continental collision, Hikurangi

margin, New Zealand. Tectonics 17: 534-557.

Barnes PM, Sutherland R, Davy B, Delteil J 2001. Rapid creation and destruction of

sedimentary basins on mature strike-slip faults: an example from the offshore Alpine

Fault, New Zealand. Journal of Structural Geology 23: 1727-1739.

Barnes PM, Nicol A, Harrison T 2002a. Late Cenozoic evolution and earthquake potential of

an active listric thrust complex above the Hikurangi subduction zone, New Zealand.

Geological Society of America Bulletin 114: 1379-1405.

Barnes PM, Davy B, Sutherland R, Delteil J 2002b. Frontal accretion and thrust wedge

evolution under very oblique plate convergence: Fiordland Basin, New Zealand. Basin

Research 14: 439-466.

Barnes PM, Sutherland R, Delteil J 2005. Strike-slip structure and sedimentary basins of the

southern Alpine Fault, Fiordland, New Zealand. Geological Society of America

Bulletin 117: 411-435.

Barnes PM, Lamarche G, Bialans J, Henrys S, Pecher I, Netzeband GL, Grenert J, Mountjoy

JJ, Pedley K, Crutchley G 2010. Tectonic and geological framework for gas hydrates

and cold seeps on the Hikurangi subduction margin, New Zealand. Marine Geology

272: 26-48.

Barrell DJA, Read SAL, Van Dissen RJ, Macfarlane DF, Walker J, Rieser U 2009.

Aviemore: a dam of two halves. In: Turnbull IM ed. Geological Society of New

Zealand and New Zealand Geophysical Society Joint Annual Conference, Oamaru, 23-

27 November 2009: field trip guides. Geological Society of New Zealand

Miscellaneous Publication 128B: 30.

Batt GE 2001. An approuch to steady-state thermochronological distribution following

orogenic development in the Southern Alps of New Zealand. American Journal of

Science 301: 374-384.

Basili R, Valensise G, Vannoli P, Burrato P, Fracassi U, Mariano S, Tiberti MM, Boshi E

2008. The Database of Individual Seismogenic Sources (DISS), version 3: summarizing

20 years of research on Italy’s earthquake geology. Tectonophysics 453: 20-43.

Beanland S, Haines J 1998. The kinematics of active deformation in the North Island, New

Zealand, determined from geological strain rates. New Zealand Journal of Geology and

Geophysics 41: 311-323.

Page 28: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

28

Beanland S, Berryman KR, Hull AG, Wood PR 1986. Late Quaternary deformation at the

Dunstan fault, Central Otago, New Zealand. In: Reilly WI, Harford BE eds. Recent

crustal movements of the Pacific region. Bulletin of the Royal Society of New Zealand

24: 293-306.

Beanland S, Berryman KR 1989. Style and episodicity of late Quaternary activity on the Pisa-

Grandview Fault Zone, Central Otago, New Zealand. New Zealand Journal of Geology

and Geophysics 32: 451-461.

Beavan J 2012. Darfield Earthquake geodetic investigations. GNS Science Consultancy

Report 2012/164.

Beavan J, Haines J, Litchfield N, Van Dissen R, Barnes P, Haines J 2007. Continuous

deformation map of New Zealand from active faulting. In: Mortimer N, Wallace L eds.

Geological Society of New Zealand and New Zealand Geophysical Society Joint

Annual Conference: Tauranga; programme and abstracts. Geological Society of New

Zealand Miscellaneous Publication 123A: 8.

Beavan RJ, Denys P, Denham M, Hager B, Herring T, Molnar P 2010a. Distribution of

present-day vertical deformation across the Southern Alps, New Zealand, from 10 years

of GPS data. Geophysical Research Letters 37: L16305.

doi:16310.11029/12010GL044165.

Beavan RJ, Samsonov S, Denys P, Sutherland R, Palmer NG, Denham M 2010b. Oblique

slip on the Puysegur subduction interface in the 2009 July Mw 7.8 Dusky Sound

earthquake from GPS and InSAR observations: implications for the tectonics of

southwestern New Zealand. Geophysical Journal International 183: 1265-1286. doi:

1210.1111/j.1365-1246X.2010.04798.x.

Berryman K 1984a. Active folding and faulting. In: Speden IG ed. Natural Hazards in New

Zealand. New Zealand Commision for UNESCO, Wellington. Pp. 302-331.

Berryman KR 1984b. Late Quaternary tectonics of New Zealand. In: Walcott RI ed. An

introduction to the Recent crustal movements of New Zealand. Royal Society of New

Zealand Miscellaneous Series 7: 91-107.

Berryman KR 1993. Age, height and deformation of Holocene marine terraces at Mahia

Peninsula, Hikurangi subduction margin, New Zealand. Tectonics 12: 1347-1364.

Berryman KR, Beanland S 1988. Ongoing deformation of New Zealand: rates of tectonic

movement from geological evidence. Transactions of the Institution of Professional

Engineers New Zealand 15: 25-35.

Page 29: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

29

Berryman KR, Beanland S 1991. Variation in fault behaviour in different tectonic provinces

of New Zealand. Journal of Structural Geology 13: 177-189.

Berryman KR, Beanland S, Cooper AF, Cutten HN, Norris RJ, Wood PR 1992. The Alpine

Fault, New Zealand: variation in Quaternary style and geomorphic expression. Annales

Tectonicae 6(supplement): 126-163.

Berryman KR (compiler) 2005. Review of tsunami hazard and risk in New Zealand. Institute

of Geological and Nuclear Sciences client report 2005/104. 139 p.

Berryman KR, Marden M, Palmer A, Litchfield NJ 2009. Holocene rupture of the

Repongaere Fault, Gisborne: implications for Raukumara Peninsula deformation and

impact on the Waipaoa Sedimentary System. New Zealand Journal of Geology and

Geophysics 52: 335-347.

Bishop DJ, Buchanan PG 1995. Development of structurally inverted basins: a case study

from the West Coast, South Island, New Zealand. In: Buchanan JG, Buchanan PG eds.

Basin inversion. Geological Society Special Publication 88: 549-585.

Browne GH, Field BD, Barrell DJA, Jongens R, Bassett KN, Wood RA 2012. The geological

setting of the Darfield and Christchurch earthquakes, New Zealand Journal of Geology

and Geophysics 55: 193-197.

Campbell JK, Pettinga JR, Jongens R 2012. The tectonic and structural setting of the 4

September 2010 Darfield (Canterbury) earthquake sequence, New Zealand, New

Zealand Journal of Geology and Geophysics 55: 155-168.

CANZ 1996. Undersea New Zealand (New Zealand Region Physiography), 1:4 000 000 (2nd

edition). New Zealand Oceanographic Institute (now NIWA) Miscellaneous Chart

Series No. 74.

Carne RC, Little TA, Reiser U 2011. Using displaced river terraces to determine Late

Quaternary slip rate for the central Wairarapa Fault at Waiohine River, New Zealand.

New Zealand Journal of Geology and Geophysics 54: 217-236.

Cashman SM, Kelsey, HM Erdman, CF Cutten, HNC Berryman KR 1992. Strain partitioning

between structural domains in the Forearc of the Hikurangi Subduction Zone. New

Zealand. Tectonics 11: 242-257.

Clark D, McPherson A, Van Dissen R 2012. Long-term behaviour of Australian stable

continental region (SCR) faults. Tectonophysics 566-567: 1-30.

Cole JW, Darby DJ, Stern TA 1995. Taupo Volcanic Zone and Central Volcanic Region:

backarc structures of North Island, New Zealand. In: Taylor B ed. Backarc basins:

tectonics and magmatism. Plenum Press, New York. Pp. 1-28.

Page 30: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

30

Cole JW, Spinks KD, Deering CD, Nairn IA, Leonard GS 2010. Volcanic and structural

evolution of the Okataina Volcanic Centre; dominantly silicic volcanism associated

with the Taupo Rift, New Zealand. Journal of Volcanology and Geothermal Research

190: 123-135.

Collot J-Y, Lamarche G, Wood RA, Delteil J, Sosson M, Lebrun J-F, Coffin MC 1995.

Morphostructure of an incipient subduction zone along a transform plate boundary:

Puysegur Ridge and Trench. Geology 23: 519-522.

Cowan HA, Nicol A, Tonkin P 1996. A comparison of historical and paleoseismicity in a

newly formed fault zone and a mature fault zone, North Canterbury, New Zealand.

Journal of Geophysical Research 101(B3): 6021-6036.

Cox SC, Findlay RF 1995. The Main Divide Fault Zone and its role in formation of the

Southern Alps. New Zealand Journal of Geology and Geophysics 38: 489-499.

Cox SC, Barrell DJA 2007. Geology of the Aoraki area. Institute of Geological & Nuclear

Sciences 1:250,000 geological map 15. 1 sheet + 71 p, GNS Science, Lower Hutt.

Cox SC, Sutherland R 2007. Regional geological framework of South Island, New Zealand,

and its significance for understanding the active plate boundary. In: Okaya DA, Stern

TA, Davey FJ eds. A continental plate boundary: tectonics at South Island, New

Zealand. Geophysical Monograph Series 175: 19-46.

Cox SC, Stirling MW, Herman F, Gerstenberger M, Ristau J 2012. Potentially active faults in

the rapidly eroding landscape adjacent to the Alpine Fault, central Southern Alps, New

Zealand. Tectonics 31: TC2011. doi:10.1029/2011TC003038.

Darby DJ, Beanland S 1992. Possible source models for the 1855 Wairarapa Earthquake,

New Zealand. Journal of Geophysical Research 97(B9): 12,375-312,389.

Davey FJ, Henyey T, Holbrook WS, Okaya D, Stern TA, Mellhuish A, Henrys S, Anderson

H, Eberhart-Philips D, McEvill T, Uhrammer R, Wu R, Jiracek GR, Wannamaker PE,

Caldwell G, Christensen N 1998. Preliminary results from a geophysical study across a

modern continent-continent collisional plate boundary - the Southern Alps, New

Zealand. Tectonophysics 288: 221-235.

Davy BW 1992. The influence of subduction plate-buoyancy of subduction of the Hikurangi-

Chatham plateau beneath the North Island, New Zealand. In: Watkins JS, Feng Z,

McMillan K eds. Geology and geophysics of continental margins. AAPG Memoir 53:

75-91.

Delteil J, Collot J-Y, Wood RA, Herzer RH, Calmant S, Christoffel D, Coffin M, Ferriere J,

Lamarche G, Lebrun J-F, Mauffret A, Pontoise B, Popoff M, Ruellan E, Sosson M,

Page 31: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

31

Sutherland R 1996. From strike-slip faulting to oblique subduction: a survey of the

Alpine Fault-Puysegur Trench transition, New Zealand, results of cruise Geodynz-sud

leg 2. Marine Geophysical Researches 18: 383-399.

Doser DI, Webb TH, Maunder DE 1999. Source parameters of large historical (1918-1962)

earthquakes, South Island, New Zealand. Geophysical Journal International 139: 769-

794.

Eberhart-Phillips D, Bannister SC 2010. 3-D imaging of Marlborough, New Zealand,

subducted plate and strike-slip fault systems. Geophysical Journal International 182:

73-96. doi: 10.1111/j.1365-1246X.2010.04621.x.

Formento-Trigilio ML, Burbank DW, Nicol A, Shulmeister J, Rieser U 2003. River response

to an active fold-and-thrust belt in a convergent margin setting, North Island, New

Zealand. Geomorphology 49: 125-152.

Fry B, Bannister SC, Beavan RJ, Bland L, Bradley BA, Cox SC, Cousins WJ, Gale NH,

Hancox GT, Holden C, Jongens R, Power WL, Prasetya G, Reyners ME, Ristau J,

Robinson R, Samsonov S, Wilson KJ, the GeoNet team 2010. The Mw7.6 Dusky Sound

earthquake of 2009: preliminary report. Bulletin of the New Zealand Society for

Earthquake Engineering 43: 24-40.

Ghisetti FC, Sibson RH 2006. Accommodation of compressional inversion in north-western

South Island (New Zealand): Old faults versus new? Journal of Structural Geology 28:

1994-2010.

Ghisetti FC, Gorman AR, Sibson RH 2007. Surface breakthrough of a basement fault by

repeated seismic slip episodes: the Ostler Fault, South Island, New Zealand. Tectonics

26: TC6004. doi:6010.1029/2007TC002146.

Ghisetti FC, Sibson RH 2012. Compressional reactivation of E–W inherited normal faults in

the area of the 2010–2011 Canterbury earthquake sequence, New Zealand Journal of

Geology and Geophysics 55: 177-184.

Giba M, Nicol A, Walsh JJ 2010. Evolution of faulting and volcanism in a back-arc basin and

its implications for subduction processes. Tectonics 29: TC4020.

doi:4010.1029/2009TC002634.

Gledhill K, Ristau J, Reyners M, Fry B, Holden C 2011. The Darfield (Canterbury, New

Zealand) Mw 7.1 Earthquake of September 2010: A preliminary seismological report.

Seismological Research Letters 82: 378-386.

Grindley GW 1960. Sheet 8 – Taupo. Geological map of New Zealand 1:250 000.

Wellington, New Zealand. Department of Scientific and Industrial Research.

Page 32: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

32

Henrys S, Reyners M, Pecher I, Bannister S, Nishimura Y, Maslen G 2006. Kinking of the

subducting slab by escalator normal faulting beneath the North Island of New Zealand.

Geology 34: 777-780.

Hochstein MP, Ballance PF 1993. Hauraki Rift: a young, active, intra-continental rift in a

back-arc setting. In: Balance PF ed. South Pacific Sedimentary Basins. Sedimentary

Basins of the World 2, Elsevier Science Publishers, B.V., Amsterdam. Pp 295-305.

Holt WE, Haines AJ 1995. The kinematics of northern South Island, New Zealand,

determined from geologic strain rates. Journal of Geophysical Research 100(B9):

17991-18010.

Howard M, Nicol A, Campbell JK, Pettinga JR 2005. Holocene paleoearthquakes on the

strike-slip Porters Pass Fault, Canterbury, New Zealand. New Zealand Journal of

Geology and Geophysics 48: 59-74.

Hull AG 1990. Tectonics of the 1931 Hawkes Bay earthquake. New Zealand Journal of

Geology and Geophysics 33: 309-320.

Jackson J, Norris R, Youngson J 1996. The structural evolution of active fault and fold

systems in central Otago, New Zealand: evidence revealed by drainage patterns. Journal

of Structural Geology 18: 217-234.

Jackson J, Van Dissen R, Berryman K 1998. Tilting of active folds and faults in the

Manawatu region, New Zealand: evidence from surface drainage patterns. New

Zealand Journal of Geology and Geophysics 41: 377-385.

Jongens R, Dellow G 2003. The Active Faults Database of NZ: Data dictionary. Institute of

Geological & Nuclear Sciences Science Report 2003/17.

Jongens R, Barrell DJA, Campbell JK, Pettinga JR 2012. Faulting and folding beneath the

Canterbury Plains identified prior to the 2010 emergence of the Greendale Fault, New

Zealand Journal of Geology and Geophysics 55: 169-176.

Kaiser A, Holden C, Beavan J, Beetham D, Benites R, Celentano A, Collet D, Cousins J,

Cubrinovski M, Dellow G, Denys P, Fielding E, Fry B, Gerstenberger M, Langridge R,

Massey C, Motagh M, Pondard N, McVerry G, Ristau J, Stirling M, Thomas J, Uma

SR, Zhao J 2012. The Mw 6.2 Christchurch earthquake of February 2011: preliminary

report. New Zealand Journal of Geology and Geophysics 55: 67-90.

Kamp PJJ 1986. The mid-Cenozoic Challenger Rift System of western New Zealand and its

implications for the age of Alpine fault inception. Geological Society of America

Bulletin 97: 255-281.

Page 33: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

33

Kelsey HM, Cashman SM, Beanland S, Berryman KR 1995. Structural evolution along the

inner forearc of the obliquely convergent Hikurangi Margin, New Zealand. Tectonics

14: 1-18.

Kingma JT 1962. Sheet 11 – Dannevirke. Geological Map of New Zealand 1:250 000.

Wellington, New Zealand. Department of Scientific and Industrial Research.

Klein EG, Flesch LM, Holt WE, Haines AJ 2009. Evidence of long-term weakness on

seismogenic faults in western North America from dynamic modelling. Journal of

Geophysical Research 114: B03402. doi:10.1029/2007JB005201.

Koons PO 1990. Two-sided orogen: Collision and erosion from the sandbox to the Southern

Alps, New Zealand. Geology 18: 679-682.

Laird MG 1968. The Paparoa Tectonic Zone. New Zealand Journal of Geology and

Geophysics 11: 435-454.

Lamarche G, Collot J-Y, Wood RA, Sosson M, Sutherland R, Delteil J 1997. The Oligocene-

Miocene Pacific-Australia plate boundary south of New Zealand: evolution from

oceanic spreading to strike-slip faulting. Earth and Planetary Science Letters 148: 129-

139.

Lamarche G, Lebrun J-F 2000. Transition from strike-slip faulting to oblique subduction:

active tectonics at the Puysegur Margin, South New Zealand. Tectonophysics 316: 67-

89.

Lamarche G, Proust J-N, Nodder SD 2005. Long-term slip rates and fault interactions under

low contractional strain, Wanganui Basin, New Zealand. Tectonics 24: TC4004.

doi:10.1029/2004TC001699.

Lamarche G, Barnes PM, Bull JM 2006. Faulting and extension rates over the last 20,000

years in the offshore Whakatane Graben, New Zealand continental shelf. Tectonics 25:

TC4005. doi:4010.1029/2005TC001886.

Langridge RM, Campbell J, Hill NL, Pere V, Pope J, Pettinga J, Estrada B, Berryman KR

2003. Paleoseismology and slip rate of the Conway Segment of the Hope Fault at

Greenburn Stream, South Island, New Zealand. Annals of Geophysics 46: 1119-1139.

Langridge RM, Villamor P, Basili R, Almond P, Martinez-Diaz JJ, Canora C 2010. Revised

slip rates for the Alpine Fault at Inchbonnie: Implications for plate boundary kinematics

of South Island, New Zealand. Lithosphere 2: 139-152.

Lebrun J-F, Lamarche G, Collot J-Y, Delteil J 2000. Abrupt strike-slip fault to subduction

transition: The Alpine Fault-Puysegur Trench connection, New Zealand. Tectonics 19:

688-706.

Page 34: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

34

Lensen GJ 1963. Sheet 16 – Kaikoura. Geological Map of New Zealand 1:250 000.

Wellington, New Zealand. Department of Scientific and Industrial Research.

Lensen GJ 1975. Earth-deformation studies in New Zealand. Tectonophysics 29: 541-551.

Lewis KB 1971. Growth rate of folds using tilted wave-planed surfaces: coast and continental

shelf, Hawke’s Bay, New Zealand. In Collins BW, Fraser R eds. Recent Crustal

Movements. Royal Society of New Zealand Bulletin 9: 225-231.

Lewis KB 1984. Offshore active deformation. In: Speden IG ed. Natural Hazards in New

Zealand. New Zealand Comission for UNESCO. Pp. 69-72.

Lewis KB, Pettinga JR 1993. The emerging, imbricate frontal wedge of the Hikurangi

Margin. In: Balance PF ed. South Pacific Sedimentary Basins. Elsevier Science

Publishers, Amsterdam. Pp. 225-250.

Lihou JC 1992. Reinterpretation of seismic reflection data from the Moutere Depression,

Nelson region, South Island, New Zealand. New Zealand Journal of Geology and

Geophysics 35: 477-490.

Litchfield NJ, Norris RJ 2000. Holocene motion on the Akatore Fault, south Otago coast,

New Zealand. New Zealand Journal of Geology and Geophysics 43: 405-418.

Litchfield NJ, Campbell JK, Nicol A 2003. Recognition of active reverse faults and folds in

North Canterbury, New Zealand, using structural mapping and geomorphic analysis.

New Zealand Journal of Geology and Geophysics 46: 563-579.

Litchfield N, Jongens R 2006. Active Faults Database of New Zealand. In: Earthquakes and

urban development. New Zealand Geotechnical Society 2006 Symposium. Institution

of Professional Engineers Proceedings of Technical Groups 31: 327-334.

Litchfield NJ, Ellis SM, Berryman KR, Nicol A 2007. Insights into subduction-related uplift

along the Hikurangi Margin, New Zealand, using numerical modeling. Journal of

Geophysical Research 112(F2): F02021. doi:02010.01029/02006JF000535.

Litchfield NJ, Van Dissen R, Sutherland R, Barnes PM, Cox S, Norris R, Beavan J,

Langridge R, Villamor P, Berryman K, Stirling M, Nicol A, Nodder S, Lamarche G,

Barrell DJA, Pettinga JR, Little T, Pondard N, Mountjoy JJ, Clark K 2013. A model of

active faulting in New Zealand: fault parameter descriptions. GNS Science Report

2012/19.

Little TA, Grapes R, Berger GW 1998. Late Quaternary strike-slip on the eastern part of the

Awatere fault, South Island, New Zealand. Geological Society of America Bulletin

110: 127-148.

Page 35: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

35

Little TA, Cox S, Vry JK, Batt GE 2005. Variations in exhumation level and uplift-rate

related to oblique-slip ramp geometry, Alpine fault, central Southern Alps, New

Zealand, Geological Society of America Bulletin 117: 707-723.

Machette M, Haller K, Wald L 2004. Quaternary Fault and Fold Database for the Nation,

United States Geological Survey Fact Sheet, 2004-3033.

Malservisi R, Furlong KP, Anderson H 2003. Dynamic uplift in a transpressional regime:

numerical model of the subduction area of Fiordland, New Zealand. Earth and

Planetary Science Letters 206: 349-364.

Markley M, Tikoff B 2003. Geometry of the folded Otago peneplain surface beneath Ida

valley, Central Otago, New Zealand, from gravity observations. New Zealand Journal

of Geology and Geophysics 46: 449-456.

Massell C, Coffin MF, Mann P, Mosher, S Frohlich C, Duncan CS, Karner G, Ramsey D,

Lebrun J-F 2000. Neotectonics of the Macquarie Ridge Complex, Australia-Pacific

plate boundary. Journal of Geophysical Research 105(B6): 13,457-413,480.

Melhuish A, Van Dissen R, Berryman K 1996. Mount Stewart - Halcome Anticline: a look

inside a growing fold in the Manawatu region, New Zealand. New Zealand Journal of

Geology and Geophysics 39: 123-133.

Melhuish A, Sutherland R, Davey FJ, Lamarche G 1999. Crustal structure and neotectonics

of the Puysegur oblique subduction zone, New Zealand. Tectonophysics 313: 335-362.

Mountjoy JJ, Barnes PM 2011. Active upper plate thrust faulting in regions of low plate

interface coupling, repeated slow slip events, and coastal uplift: Example from the

Hikurangi Margin, New Zealand, Geochemistry Geophysics Geosystems 12: Q01005,

doi:10.1029/2010GC003326.

Mouslopoulou V, Nicol A, Little TA, Walsh JJ 2007a. Displacement transfer between

intersecting regional strike-slip and extensional fault systems. Journal of Structural

Geology 29: 100-116.

Mouslopoulou V, Nicol A, Little TA, Walsh JJ 2007b. Terminations of large strike-slip

faults: an alternative model from New Zealand. In: Cunningham WD, Mann P eds.

Tectonics of strike-slip restraining and releasing bends. Geological Society of London

Special Publication 290: 387-415.

Mouslopoulou V, Walsh JJ, Nicol A 2009. Fault displacement rates on a range of timescales.

Earth and Planetary Science Lettters 278: 186-197.

Page 36: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

36

Mouslopoulou V, Nicol A, Walsh JJ, Begg JG, Townsend DB, Hristopulos DT 2012. Fault-

slip accumulation in an active rift over thousands to millions of years and the

importance of paleoearthquake sampling. Journal of Structural Geology 36: 71-80.

Mutch AR, Wilson DD 1952. Reversal of movement of the Titri Fault. New Zealand Journal

of Science and Technology B33: 398-403.

Nicol A 1993. Haumurian (c 66-80 Ma) half graben, Mid Waipara, North Canterbury and its

implication for Late Cretaceous deformation in New Zealand (note). New Zealand

Journal of Geology and Geophysics 36: 127-130.

Nicol A, Nathan S 2001. Folding and the formation of bedding-parallel faults on the western

limb of Grey Valley Syncline near Blackball, New Zealand. New Zealand Journal of

Geology and Geophysics 44: 127-135.

Nicol A, Van Dissen RJ 2002. Up-dip partitioning of displacement components on the

oblique-slip Clarence Fault, New Zealand. Journal of Structural Geology 24: 1521-

1535.

Nicol A, Beavan RJ 2003 Shortening of an overriding plate and its implications for slip on a

subduction thrust, central Hikurangi Margin, New Zealand. Tectonics 22: 1070,

doi:10.1029/2003TC001521.

Nicol A, Wallace LM 2007. Temporal stability of deformation rates: Comparison of

geological and geodetic observations, Hikurangi subduction margin, New Zealand.

Earth and Planetary Science Letters 258: 397-413.

Nicol A, Wise DU 1992. Paleostress adjacent to the Alpine Fault of New Zealand: vein and

stylolite data from the Doctors Dome area. Journal of Geophysical Research 97(B12):

17685-17692.

Nicol A, Alloway B, Tonkin P 1994. Rates of deformation, uplift and landscape development

associated with active folding in the Waipara area of North Canterbury, New Zealand.

Tectonics 13: 1327-1344.

Nicol A, Cowan H, Campbell JK, Pettinga JR 1995. Folding and the development of small

sedimentary basins along the New Zealand plate boundary. Tectonophysics 241: 47-54.

Nicol A, Van Dissen R, Vella P, Alloway B, Melhuish A 2002. Growth of contractional

structures during the last 10 m.y. at the southern end of the emergent Hikurangi forearc

basin, New Zealand. New Zealand Journal of Geology and Geophysics 45: 365-385.

Nicol A, Begg J, Mouslopoulou V, Stirling M, Townsend D, Van Dissen R, Walsh J 2011.

Active faults in New Zealand: what are we missing? In: Litchfield NJ, Clark K eds.

Page 37: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

37

Abstract Volume, Geosciences 2011 Conference, Nelson, New Zealand. Geoscience

Society of New Zealand Miscellaneous Publication 130A: 79.

Nodder SD 1994. Characterizing potential offshore seismic sources using high-resolution

geophysical and seafloor sampling programs: an example from Cape Egmont fault

zone, Taranaki shelf, New Zealand. Tectonics 13: 641-658.

Nodder SD, Lamarche G, Proust J-N, Stirling MW 2007. Characterizing earthquake

recurrence parameters for offshore faults in the low-strain, compressional Kapiti-

Manawatu Fault System, New Zealand. Journal of Geophysical Research 112(B12):

B12102. doi:12110.11029/12007JB005019.

Norris RJ, Cooper AF 2001. Late Quaternary slip rates and their significance for slip

partitioning on the Alpine Fault, New Zealand. Journal of Structural Geology 23: 507-

520.

Norris RJ, Koons PO, Landis CA 1987. Aspects of the South Island collision zone, central

Otago and the West Coast. In: Annual Conference of the Geological Society of New

Zealand post conference field trip 2. Geological Society of New Zealand Miscellaneous

Publication vol. 37c: 39-88.

Norris RJ, Koons PO, Cooper AF 1990. The obliquely-convergent plate boundary in the

South Island of New Zealand: implications for ancient collision zones. Journal of

Structural Geology 12: 712-725.

Officers of the New Zealand Geological Survey 1983. Late Quaternary Tectonic Map of New

Zealand 1:2 000 000. 2nd ed. New Zealand Geological Survey Miscellaneous Series

Map 12, Department of Scientific and Industrial Research, Wellington.

Pedley KL, Barnes PM, Pettinga JR, Lewis KB 2010. Seafloor structural geomorphic

evolution of the accretionary frontal wedge in response to seamount subduction,

Poverty Indentation, New Zealand. Marine Geology 270: 119-138.

Petersen MD, Frankel AD, Harmsen SC, Mueller CS, Haller KM, Wheeler, RL, Wesson RL,

Zeng Y, Boyd OS, Perkins DM, Luco N, Field EH, Wills CJ, Rukstales KS 2008.

Documentation for the 2008 Update of the United States National Seismic Hazard

Maps. United States Geological Survey Open-File Report 2008–1128, 61 p.

Pettinga JR, Yetton MD, Van Dissen RJ, Downes G 2001. Earthquake source identification

and characterisation for the Canterbury Region, South Island, New Zealand. Bulletin of

the New Zealand Society of Earthquake Engineering 34: 282-317.

Pillans B 1990a. Pleistocene marine terraces in New Zealand: a review. New Zealand Journal

of Geology and Geophysics 33: 219-231.

Page 38: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

38

Pillans B 1990b. Vertical displacement rates on Quaternary faults, Wanganui Basin. New

Zealand Journal of Geology and Geophysics 33: 271-275.

Pillans B 1991. New Zealand Quaternary stratigraphy: an overview. Quaternary Science

Reviews 10: 405-418.

Pondard N, Barnes PM 2010. Structure and paleoearthquake records of active submarine

faults, Cook Strait, New Zealand: implications for fault interactions, stress loading, and

seismic hazard. Journal of Geophysical Research 115: B12320.

doi:12310.11029/12010JB007781.

Priest GR, Goldfinger C, Wang K, Witter RC, Zhang Y, Baptista AM 2009. Confidence

levels for tsunami-inundation limits in northern Oregon inferred from a 10,000-year

history of great earthquakes on the Cascadia subduction zone. Natural Hazards 54: 27-

73.

Quigley M, Van Dissen RJ, Villamor P, Litchfield NJ, Barrell DJA, Furlong K, Stahl T,

Duffy B, Bilderback E, Noble D, Townsend DB, Begg JG, Jongens R, Ries W,

Claridge J, Klahn A, Mackenzie H, Smith A, Hornblow S, Nicol R, Cox SC, Langridge

RM, Pedley K 2010. Surface rupture of the Greendale Fault during the Darfield

(Canterbury) Earthquake, New Zealand: initial findings. Bulletin of the New Zealand

Society of Earthquake Engineering 43: 236-242.

Quigley M, Van Dissen RJ, Litchfield NJ, Villamor P, Duffy B, Barrell DJA, Furlong K,

Stahl T, Bilderback E, Noble D 2012. Surface rupture during the 2010 Mw 7.1 Darfield

(Canterbury) earthquake: implications for fault rupture dynamics and seismic-hazard

analysis. Geology 40: 55-58.

Rattenbury MS, Jongens R, Cox SC 2010. Geology of the Haast area. Institute of Geological

& Nuclear Sciences 1:250,000 geological map 14. 1 sheet + 58 p, GNS Science, Lower

Hutt.

Reyners M, Cowan H 1993. The transition from subduction to continental collision: crustal

structure in the North Canterbury region, New Zealand. Geophysical Journal

International 117: 1124-1136.

Reyners M, Eberhart-Phillips D, Stuart G 1999. A three-dimensional image of shallow

subduction: crustal structure of the Raukumara Peninsula, New Zealand. Geophysical

Journal International 137: 873-890.

Reyners M, McGinty P, Cox S, Turnbull I, O'Neill T, Gledhill K, Hancox G, Beavan J,

Matheson D, McVerry G, Cousins J, Zhao J, Cowan H, Caldwell G, Bennie S, the

Geonet team 2003. The Mw 7.2 Fiordland earthquake of August 21, 2003: Background

Page 39: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

39

and preliminary results. Bulletin of the New Zealand Society of Earthquake

Engineering 36: 233-248.

Reyners ME, Eberhart-Phillips D, Stuart G, Nishimura Y 2006. Imaging subduction from the

trench to 300 km depth beneath the central North Island, New Zealand, with Vp and

Vp/Vs. Geophysical Journal International 165: 565-583.

Ries WF, de Roiste M, Villamor P 2011. Location of active faults using geomorphic indices

in eroded landscapes, South Taranaki, New Zealand. In: Clark KJ, Litchfield NJ eds.

Geosciences 2011: Annual Conference of the Geoscience Society of New Zealand:

abstract volume. Geoscience Society of New Zealand Miscellaneous Publication 130A:

90-91.

Robinson R, McGinty P 2000. The enigma of the Arthur's Pass, New Zealand, earthquake 2.

The aftershock distribution and its relation to regional and induced stress fields. Journal

of Geophysical Research 105(B7): 16139-16150.

Robinson R, Nicol A, Walsh JJ, Villamor P 2009. Features of Earthquake occurrence in a

complex normal fault network: results from a synthetic seismicity model of the Taupo

Rift, New Zealand. Journal of Geophysical Research 114, B12306,

doi:10.1029/2008JB006231.

Robinson R, Van Dissen R, Litchfield N 2011. Using synthetic seismicity to evaluate seismic

hazard in the Wellington region, New Zealand. Geophysical Journal International 187:

510-528.

Rodgers DW, Little TA 2006. World's largest coseismic strike-slip offset : the 1855 rupture

of the Wairarapa Fault, New Zealand, and implications for displacement/length scaling

of continental earthquakes. Journal of Geophysical Research 111(B12): B12408.

doi:10.1029/2005JB004065.

Rowland JV, Sibson RH 2001. Extensional fault kinematics within the Taupo Volcanic Zone,

New Zealand: soft-linked segmentation of a continental rift system. New Zealand

Journal of Geology and Geophysics 44: 271-283.

Rowland JV, Wilson CJN, Gravley DM 2010. Spatial and temporal variations in magma-

assisted rifting, Taupo Volcanic Zone, New Zealand. Journal of Volcanology and

Geothermal Research 190: 89-108.

Seebeck H, Nicol A 2009. Dike intrusion and displacement accumulation at the intersection

of the Okataina Volcanic Centre and Paeroa Fault zone, Taupo rift, New Zealand.

Tectonophysics 475: 575-585.

Page 40: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

40

Sibson RH, Ghisetti FC 2010. Characterising the seismic potential of compressional

inversion structures, NW South Island. EQC Project 08/547.

Stafford PJ, Pettinga JR, Berrill JB 2008. Seismic source identification and characterisation

for probabilistic seismic hazard analyses conducted in the Buller – NW Nelson Region,

South Island, New Zealand. Journal of Seismology 12: 477-498.

Stern TA, Quinlan GM, Holt WE 1992. Basin formation behind an active subduction zone:

three-dimensional flexural modelling of Wanganui Basin, New Zealand. Basin

Research 4: 197-214.

Stirling MW, Wesnousky SG, Berryman KR 1998. Probability seismic hazard analysis of

New Zealand. New Zealand Journal of Geology and Geophysics 41: 355-375.

Stirling MW, McVerry GH, Berryman KR 2002. A new seismic hazard model for New

Zealand. Bulletin of the Seismological Society of America 92: 1878-1903.

Stirling M, Berryman K, Litchfield N, Wallace L 2009. Multi-disciplinary probabilistic

tectonic hazard analysis. In: Conner CB, Chapman NA, Conner LR eds. Volcanic and

tectonic hazard assessment for nuclear facilities. Cambridge University Press,

Cambridge. Pp. 257-275.

Stirling MW, McVerry GH, Gerstenberger MC, Litchfield NJ, Van Dissen RJ, Berryman KR,

Barnes P, Wallace LM, Villamor P, Langridge RM, Lamarche G, Nodder S, Reyners

ME, Bradley B, Rhoades DA, Smith WD, Nicol A, Pettinga J, Clark KJ, Jacobs K

2012. National seismic hazard model for New Zealand: 2010 update. Bulletin of the

Seismological Society of America 102, 1514-1542.

Suggate RP 1987. Active folding in north Westland, New Zealand. New Zealand Journal of

Geology and Geophysics 30: 169-174.

Suggate RP 1989. The postglacial shorelines and tectonic development of the Barrytown

coastal lowland, north Westland, New Zealand. New Zealand Journal of Geology and

Geophysics 32: 443-450.

Sutherland R 1995. Late Cenozoic tectonics in the SW Pacific and development of the Alpine

Fault through southern South Island, New Zealand. Unpublished PhD thesis, University

of Otago, Dunedin, New Zealand.

Sutherland R, Norris RJ 1995. Late Quaternary displacement rate, paleoseismicity, and

geomorphic evolution of the Alpine Fault: evidence from Hokuri Creek, South

Westland. New Zealand Journal of Geology and Geophysics 38: 419-430.

Page 41: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

41

Sutherland R, Davey FJ, Beavan RJ 2000. Plate boundary deformation in South Island, New

Zealand, is related to inherited lithospheric structure. Earth and Planetary Science

Letters 177: 141-151.

Sutherland R, Berryman K, Norris R 2006. Quaternary slip rate and geomorphology of the

Alpine fault: Implications for kinematics and seismic hazard in southwest New

Zealand. Geological Society of America Bulletin 118: 464-474.

Sutherland R, Kim KJ, Zondervan A, McSaveney MJ 2007. Orbital forcing of mid-latitude

Southern Hemisphere glaciation since 100 ka inferred from cosmogenic nuclide ages of

moraine boulders from the Cascade Plateau, southwest New Zealand. Geological

Society of America Bulletin 119: 443-451.

Sutherland R, Gurnis M, Kamp PJJ, House MA 2009. Regional exhumation history of brittle

crust during subduction initiation, Fiordland, southwest New Zealand, and implications

for thermochronologic sampling and analysis strategies. Geosphere 5: 409-425.

Thornley S 1997. Neogene tectonics of Raukumara Peninsula, northern Hikurangi margin,

New Zealand. Unpublished PhD thesis, Victoria University of Wellington, Wellington,

New Zealand. 291 p.

Townsend D, Nicol A, Mouslopoulou V, Begg JG, Beetham RD, Clark D, Giba M, Heron D,

Lukovic B, McPherson A, Seebeck H, Walsh JJ 2010. Paleoearthquake histories across

a normal fault system in the southwest Taranaki Peninsula, New Zealand. New Zealand

Journal of Geology and Geophysics 53: 375-394.

Upton P, Craw D, James Z, Koons PO 2004. Structure and late Cenozoic tectonics of the

southern Two Thumb Range, mid Canterbury, New Zealand. New Zealand Journal of

Geology and Geophysics 47: 141-153.

Upton P, Koons PO, Craw D, Henderson CM, Enlow R 2009. Along-strike differences in the

Southern Alps of New Zealand: Consequences of inherited variation in rheology.

Tectonics 28: TC2007. doi:2010.1029/2008TC002353.

Van Avendonk HJA, Holbrook WS, Okaya D, Austin JK, Davey FJ, Stern T 2004.

Continental crust under compression: a seismic refraction study of South Island

Geophysical Transect I, South Island, New Zealand. Journal of Geophysical Research

109(B6): B06302. doi:06310.01029/02003JB002790.

Van Dissen R, Yeats YS 1991. Hope Fault, Jordan Thrust and uplift of the Seaward Kaikoura

Range, New Zealand. Geology 9: 393-396.

Page 42: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

42

Van Dissen R, Berryman KR 1996. Surface rupture earthquakes over the last ~1000 years in

the Wellington region, New Zealand, and implications for ground shaking hazard.

Journal of Geophysical Research 101(B3): 5999-6019.

Van Dissen RJ, Berryman KR, Webb TH, Stirling MW, Villamor P, Wood PR, Nathan S,

Nicol A, Begg JG, Barrell DJA, McVerry GH, Langridge RM, Litchfield NJ, Pace B

2003. An interim classification of New Zealand's active faults for the mitigation of

surface rupture hazard. In: Proceedings of the 2003 Pacific Conference on Earthquake

Engineering, 13-15 February 2003, Christchurch, New Zealand. New Zealand Society

of Earthquake Engineering Wellington. 8 p.

Vandergoes MJ, Hogg AG, Lowe DJ, Newnham RM, Denton GH, Southon J, Barrell DJA,

Wilson CJN, McGlone MS, Allan ASR, Almond PC, Petchey, F, Dalbell K,

Dieffenbacher-Krall AC, Blaauw M In press. A revised age for the Kawakawa/Oruanui

tephra, a key marker for the Last Glacial Maximum in New Zealand. Quaternary

Science Reviews.

Villamor P, Berryman KR 2001. A Late Quaternary extension rate in the Taupo Volcanic

Zone, New Zealand, derived from fault slip data. New Zealand Journal of Geology and

Geophysics 44: 243-269.

Villamor P, Berryman KR 2006a. Evolution of the southern termination of the Taupo Rift,

New Zealand. New Zealand Journal of Geology and Geophysics 49: 23-37.

Villamor P, Berryman KR 2006b. Late Quaternary geometry and kinematics of faults at the

southern termination of the Taupo Volcanic Zone, New Zealand. New Zealand Journal

of Geology and Geophysics 49: 1-21.

Walcott RI 1984. The major structural elements of New Zealand. In: Recent crustal

movements of the New Zealand region. Royal Society of New Zealand Miscellaneous

Series 7: 1-6.

Walcott RI 1987. Geodetic strain and the deformational history of the North Island of New

Zealand during the late Cainozoic. Philosophical Transactions of the Royal Society of

London Series A 321: 163-181.

Wallace LM, Beavan RJ, McCaffrey R, Darby DJ 2004. Subduction zone coupling and

tectonic block rotations in the North Island, New Zealand. Journal of Geophysical

Research 109(B12): B12406. doi:12410.11029/12004JB003241.

Wallace LM, Beavan J, McCaffrey R, Berryman K, Denys P 2007. Balancing the plate

motion budget in the South Island, New Zealand using GPS, geological and

seismological data. Geophysical Journal International 168: 332-352.

Page 43: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

43

Wallace LM, Reyners ME, Cochran UA, Bannister SC, Barnes PM, Berryman KR, Downes

GL, Eberhart-Phillips D, Fagereng A, Ellis SM, Nicol A, McCaffrey R, Beavan RJ,

Henrys SA, Sutherland R, Barker DHN, Litchfield NJ, Townend J, Robinson R, Bell

RE, Wilson KJ, Power WL 2009. Characterizing the seismogenic zone of a major plate

boundary subduction thrust: Hikurangi Margin, New Zealand. Geochemistry,

Geophysics, Geosystems 10: Q10006. doi:10010.11029/12009GC002610.

Wallace LM, Barnes P, Beavan RJ, Van Dissen RJ, Litchfield NJ, Mountjoy J, Langridge

RM, Lamarche G, Pondard N 2012. The kinematics of a transition from subduction to

strike-slip: an example from the central New Zealand plate boundary. Journal of

Geophysical Research Q10006: B02405. doi:10.1029/2011JB008640.

Wannamaker PE, Jiracek GR, Stodt JA, Caldwell TG, Gonzalez VM, McKnight JD, Porter

AD 2002. Fluid generation and pathways beneath an active compressional orogen, the

New Zealand Southern Alps, inferred from magnetotelluric data. Journal of

Geophysical Research 107(B6): 2117. doi:10.1029/2001JB000186.

Wannamaker PE, Caldwell TG, Jiracek FR, Maris V, Hill GJ, Ogawa Y, Bibby HM, Bennie

SL, Heise W 2009. Fluid and deformation regmie of an advancing subduction system at

Marlborough, New Zealand. Nature 460: 733-737.

Wellman HW 1953. Data for the study of recent and late Pleistocene faulting in the South

Island of New Zealand. New Zealand Journal of Science and Technology 34: 270-288.

Wellman HW 1955. New Zealand Quaternary tectonics. Geologische Rundschau 43: 248-

257.

Wesnousky SG, Scholz CH, Shimazaki K 1982. Deformation of an island arc: rates of

moment release and crustal shortening in intraplate Japan determined from seismicity

and Quaternary fault data. Journal of Geophysical Research 87(B8): 6829-6852.

Wilson CJN 2001. The 26.5 ka Oruanui eruption, New Zealand: an introduction and

overview. Journal of Volcanology and Geothermal Research 112: 133-174.

Wood RA, Pettinga JR, Bannister S, Lamarche G, McMorran TJ 1994. Structure of the

Hanmer strike-slip basin, Hope Fault, New Zealand. Geological Society of America

Bulletin 106: 1459-1473.

Wood RA, Herzer RH, Sutherland R, Melhuish A 2000. Cretaceous-Tertiary tectonic history

of the Fiordland margin, New Zealand. New Zealand Journal of Geology and

Geophysics 43: 289-302.

Wright IC 1993. Southern Havre Trough - Bay of Plenty (New Zealand): structure and

seismic stratigraphy of an active back-arc basin complex. In: Balance PF ed. South

Page 44: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

44

Pacific sedimentary basins. Sedimentary basins of the world. 2. Elsevier Science

Publishers, Amsterdam. Pp. 195-211.

Wysoczanski RJ, Todd E, Wright IC, Leybourne MI, Hergt JM, Dam A, Mackay K 2010.

Backarc rifting, constructional volcanism and nascent disorganised spreading in the

southern Havre Trough backarc rifts (SW Pacific). Journal of Volcanology and

Geothermal Research 190: 39-57.

Yeats RS, Berryman KR 1987. South Island, New Zealand, and Transverse Ranges,

California: A seismotectonic comparison. Tectonics 6: 363-376.

Yoshioka T, Awata Y, Shimokawa Y, Fusejima Y 2005. Explanatory text of the Rupture

Probability Map of Major Active Faults in Japan. Tectonic Map Series 14. Geological

Survey of Japan, AIST, Japan.

Page 45: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

45

Figure captions

Figure 1 A, Tectonic setting of New Zealand. The Pacific plate motion vectors relative to the

Australian plate are from Wallace et al. (2007) and the coloured bathymetry map is from

CANZ (1996). AF is the Alpine Fault, AR is North Island axial ranges, CH is Caswell High,

ChP is Challenger Plateau, CR is Chatham Rise, CS is Cook Strait, cSA is central Southern

Alps, F is Fiordland, HP is Hikurangi Plateau, HR is Hauraki Rift, HkT is Hikurangi Trough,

HvT is Havre Trough, KR is Kaikoura Ranges, KT is Kermadec Trench, PR is Puysegur

Ridge, PT is Puysegur Trench, RR is Resolution Ridge, RP is Raukumara Peninsula, TB is

Taranaki Basin, TR is the Taupo Rift, WB is Wanganui Basin. B, New Zealand active fault

zones (n = 635) compiled for this study. Enlarged maps, on which fault zones are individually

numbered for correlation with the fault zone parameter table, are contained in supplementary

file Figures S1-S9. Sense of movement is also shown for all faults in supplementary file

Figures S1-S9, but is shown for selected faults here; ticks denote the downthrown side of

normal fault zones; triangles the upthrown side of reverse fault zones; arrows the sense of

movement of strike-slip fault zones. The boundaries between sections of the Hikurangi

subduction thrust and the Alpine Fault are denoted with short cross-lines. CVA is

Coromandel Volcanic Arc, CaP is Campbell Plateau, E is eastern Taranaki Rift, FB is

Fiordland Basin, HB is Hanmer Basin, HR is Hauraki Rift, MtR is Mount Ruapehu, MtC is

Mount Cook, O is Oamaru, OVC is Okataina Volcanic Centre, P is Papatowai, PB is

Puysegur Bank, PN is Palmerston North, ST is Snares Trough, TVC is Taupo Volcanic

Centre, W is western Taranaki Rift.

Figure 2 Major New Zealand tectonic domains, defined primarily by grouping fault zones

with similar sense of movements. Black lines are representative transects along which

cumulative net slip rates (mm/yr) have been calculated from summing net slip rates on

individual fault zones, as labelled. Uncertainties have been calculated from the root mean

square of individual fault zone net slip rate uncertainties. H, M, and L are relative confidence

rankings, broadly H (high) = most slip rates were calculated from field or seismic reflection

data, M (moderate) = most slip rates were assigned from nearby fault zones or consideration

of slip rate budgets, L (low) = most slip rates were inferred.

Figure 3 Examples of active fault traces in the GNS Science Active Faults Database

(http://data.gns.cri.nz/af/) (red) and the simplified active fault zones compiled for the model

presented in this paper (black). A, Simplification of multiple trace normal faults in the central

Page 46: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

46

Taupo Rift. B, Simplification and extension of strike-slip faults, and the addition of an

inferred connecting fault (230) in highly erodible hill country, northeast North Island. The

map locations are shown on Fig. 1B as dotted rectangles. Numbers refer to the fault zone

numbers in supplementary file Table S1.

Figure 4 Fault zones coloured according to their dominant sense of movement. Domain

names are listed in Figure 2.

Figure 5 Fault zones coloured according to their net slip rate (best estimates). Domain names

are listed in Figure 2. The map highlights that the highest slip rates are along the main plate

boundary elements, as well as the transfer of deformation between faults such as the Alpine

Fault and the Marlborough Fault System.

Figure 6 Comparison of cumulative horizontal fault zone slip rates (red) and horizontal GPS

velocity differences (between transect endpoints) (blue) for selected transects across tectonic

domains, shown as vectors (A), transect parallel motion (B), and transect normal motion (C).

In A, where the cumulative horizontal fault zone slip rate vectors are shorter than the size of

the arrow heads, the arrow heads are not shown, and the asterisks mark vectors which are

very close, or exactly parallel to the transects. The vectors are shown relative to a fixed

Australian plate.

Page 47: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

PACIFICPLATE

AUSTRALIANPLATE

A

Figs. 1B & 2

B

40°S

170°E 175°E N

35°S

48

42

41 mm/yr

40

TR

cSAH

kT

AF

PT P

R

KT

RR

HP

CR

WB

TB

CS

RP

HvT

F

CH

AR

AR

40°S

45°S

165°E 170°E 175°E 180°

Fig. 3A

Fig. 3B

MtC

TVC

OVC

O

P

PN

HB

CVA

MtR

FB

ST

PB

CaP

HR

W

E

KR

HR

400 km

ChP

Page 48: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

40°S

45°S

165°E 170°E 175°E 180°

3.2 H

15.2 H5.9 ± 0.9 M4.4 ± 0.7 M

6.2 ± 1.1 M

7.1 M

14.4 ± 1.7 L

5.4 M

13.6 M

9.7 ± 1.6 M

14.4 L

0.7 ± 0.3 L

1.0 ± 0.4 L

37.8 ± 3.4 H

+3.4

-3.7

23.5

M

2.3 ± 0.7 L

1.0

± 0

.4 L

1.2 ± 0.6 L

3.8 ± 1.3 M7.4 ± 2.2 M

+0-2.6

+0.3-0.6

+0.6-1.6

2.5 M

+0.3-0.622.3 M

+1.8-3.8

+3.4-2.5

+2.5

-2.9

3.1

L+1.0

-0.8

>5.6

L+

0.9

-0.6

>3.9 M

+3.1-1.93.9 L

+1.9-1.1

2

3

5

4

10

8

7

9

11

12

1. extensional western North Island fault zones2. extensional Havre Trough - Taupo Rift fault zones3. contractional Kapiti - Manawatu fault zones4. North Island Dextral Fault Belt5. Hikurangi subduction margin forearc6. Hikurangi subduction thrust7. contractional northwestern South Island fault zones8. strike-slip Marlborough Fault System9. contractional North Canterbury fault zones10. extensional North Mernoo fault zones11. contractional southeastern South Island fault zones12. Alpine Fault13. western Fiordland Margin - Caswell High fault zones14. Puysegur Ridge - Bank strike-slip fault zones15. Puysegur subduction thrust

1

Tectonic domains

13

14

6

15

+0.5-1.1

Page 49: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

10 km10 km

170

172

171

173

17417

517

6

177 17

8

A

B

231

10 km

Page 50: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

Reverse

Normal

Dextral

Sinistral

Dominant sense of movement

40°S

45°S

165°E 170°E 175°E 180°

2

3

5

4

10

8

7

9

11

12

1

13

14

6

15

Page 51: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

40°S

45°S

165°E 170°E 175°E 180°

2

3

5

4

10

8

7

9

11

12

1

13

14

6

15

³40

10 - 19.9

5 - 9.9

1 - 4.9

Slip rate (mm/yr)

0.5 - 0.9

unknown

£ 0.4

20 - 39.9

Page 52: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

*

-9.1 -6.1

-2.9 -4.3-2.3 -1.8

1.9 6.00.5 1.5

-1.3 2.8

0.04 -1.1

1.0 5.7

2.4 8.8

9.5 3.8

7.1 8.9

6.7 10.4

0.4 5.9

0.6 4.1

4.0 2.20.7

-3.1

1.2 2.21.4

1.2

>1.2

-1.3

-0.6

>1.9 9.7

2.5 9.2

0.2 8.2

1.8

4.5

B

-1.1 -1.9

Fault zone slip rates

GPS velocity differences

0.8

3.1

N

14

11

13

109

8

7

3

4

6

25

1

12

*

*

*

*

*

*

*

*

*

**

*

*

20 mm/yr

14

11

13

109

8

7

3

4

6

25

1

12

A

0.4

5.2

C

0 7

.4

0 3

.7

0 5

.1

0 6

.90.8

3.1

3.2

6.3

6.9

5.3

13.4

8.0

19.2

12.

8

0.04

8.5

0.2

18.

67.9 1

1.3

0 5

.9

0 4

.6 33.2 18.8

22.5 16.51.2 4

.8

0.7 8.6

>3.4 11.2

0

>1.2 2

0.3

0.2

17.

7

0.5

20.

5

0

0

109

12

3

8

2 5

4

1

6

13

11

14

7

Cumulative horizontal fault zone slip rates (mm/yr)

GPS horizontal velocitydifferences (mm/yr)

Page 53: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

1

Table 1 Descriptions of active fault zone parameters compiled in this model (supplementary files Table S1 and S2) and the main methods of

data compilation. The derivations of each parameter for individual fault zones are described by Litchfield et al. (2013).

Parameter Unit Description Main methods of compilation:

Dip 0 – 90º Downward inclination of the fault plane from the horizontal. All fault zones are considered to be planar and the best estimate of average dip used.

(i) Obtained from direct measurements in shallow natural or trench exposures, seismic reflection profiles, or multibeam bathymetric data. (ii) Inferred from geomorphic expression (e.g., upthrown side) and sense of movement, contiguous fault zones along strike, or from fault zones within the same tectonic domain. Uncertainties: (i) Obtained from the range of dip values in an exposure or a seismic profile line. (ii) Inferred to span reasonable values (generally no more than ±10º-15º).

Dip-direction N, NE, E, SE, S, SW, W, NW

The geographic octant towards which the fault zone dips.

Based on the overall strike of the mapped fault zone, combined with information for dip described above.

Sense of movement

Normal Reverse Dip-slip Sinistral Dextral

Strike-slip

Dominant, and where applicable, secondary sense (type) of relative movement (slip or displacement) on the fault plane.

(i) Obtained from direct measurements in shallow natural or trench exposures, or seismic reflection profiles. (ii) Inferred from geomorphic expression (e.g., topographic or bathymetric scarps, offset channels) and dip, or from more direct data for nearby fault zones of similar strike and/or geological setting.

Rake 0 – 360º, where: 360º = 0º = pure

dextral 90º = pure normal

180º = pure sinistral 270º = pure reverse

The direction of hanging wall slip, relative to a horizontal line on the fault plane.

(i) Obtained directly from field data (e.g., fault plane striations). (ii) Calculated from some combination of H:V ratios (e.g., from offset geomorphic features) and dip, or strike-slip rate and dip-slip rate. (iii) Inferred, generally assuming pure dip-slip or strike-slip motion. Uncertainties: (i) Calculated from the range of field-derived values, H:V ratios, or slip rate components. (ii) Inferred values assumed to be reasonable based on other fault zones in the same tectonic domain (generally no more than ±10º). Note: Except where multibeam bathymetric data can be used, the expression of strike-slip on submarine fault zones cannot generally be quantified unequivocally in marine seismic reflection data.

Slip rate mm/yr Net (rake-parallel) rate of movement averaged over a time period spanning at

(i) Calculated from strike-slip, vertical, and/or dip-slip rate, dip, and/or rake. The calculated slip rates are generally from offset geologic units (e.g., tephras in a trench)

Page 54: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

2

least two, but typically many more, ground-rupturing earthquakes. Where fault zones have associated folds, the net slip rate usually includes displacement accommodated by folding.

or geomorphic features (e.g., fluvial terrace surfaces or channels), or seismic reflectors (e.g., the post-glacial marine transgression surface). (ii) Assigned from long-term (millions of years) slip rates, particularly if the rates are broadly consistent with GPS strain rates or a portion of overall plate boundary rates. (iii) Inferred from:

(a) Geomorphic expression (e.g. scarp height or morphology) in relation to the assumed age of the faulted geomorphic surface.

(b) Slip rates assigned to nearby similar fault zones (e.g., from onshore to offshore fault zones).

(c) Consideration of geomorphic expression combined with a slip rate budget for a particular part of the plate boundary.

Quality code 1-5, where 1 is the high quality

and 5 is poor. See Table 2 for

definitions

Subjective ranking of the relative quantity and type of data available for each fault zone, particularly in regard to calculation of slip rate.

Assigned using the definitions in Table 2.

Page 55: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

3

Table 2 Definitions of the quality code parameter. The codes are particularly weighted

toward the quality of slip rate data. Code number Description

1 Fault zone with high quality field (e.g., trench, dated displaced markers) or marine (e.g., swath, markers with well constrained ages) data.

2 Fault zone with some constraints from field or marine data; slip rate largely estimated from considering regional slip rate budgets.

3 Fault zone in offshore Bay of Plenty (northern domain 2 in Fig. 2) with well constrained geometries from swath and seismic data, but slip rate inferred based on distribution of extension rates derived from modelling of onshore GPS data.

4 Fault zone with few data and slip rates mainly inferred from geomorphic expression (e.g. scarp height and morphology as an indicator of age).

5 Fault zone within the central South Island (northwestern domain 11 in Fig. 2) that does not have demonstrable activity and hence a slip rate has not been inferred.

Page 56: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

4

Table 3 Number and density of faults in each tectonic domain. The highest values for each

are shown in bold. Island Domain

number Number of fault zones

% of total number of fault zones

Cumulative fault zone

length (km)

Domain area

(km2)

Density 1 No. of fault zones per

km2

Density 2 Length (km) of fault zones

per km2 North 1 27 4% 788 110,394 0.0002 0.0071 2 196 31% 3086 42,468 0.0046 0.0727 3 28 4% 684 11,353 0.0025 0.0602 4 63 10% 2224 24,513 0.0025 0.0907 5 100 16% 3142 92,249 0.0011 0.0341 6 3 0.5% 700 15,533 0.0002 0.0451

North Totals

416 66% 10,623 296,520 0.0014 0.0358

South 7 9 1% 569 53,337 0.0002 0.0107 8 41 6% 1536 21,023 0.0020 0.0730 9 43 7% 1461 24298 0.0018 0.0601 10 12 2% 482 8,391 0.0014 0.0574 11 81 13% 3265 151,612 0.0005 0.0215 12 6 1% 741 6,692 0.0009 0.1107 13 21 3% 945 20,298 0.0010 0.0465 14 4 1% 553 21,705 0.0002 0.0255 15 1 0.2% 430 10,079 0.0001 0.0427

South Totals

219 34% 9981 317,497 0.0007 0.0314

NZ totals 635 20,064 614,017 0.0010 0.0336

Page 57: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

5

Table 4 Summary of our best assessment of domain fault zone mapping completeness. Domain number

Relative level of

completeness

Likely main areas where the fault model is incomplete

1 Medium (i) Centre – rapidly eroding areas of soft mudstone (ii) South and North – submarine areas where seismic profile analysis has not been undertaken or is incomplete

2 High (i) Centre – volcanic centres buried beneath lakes and tephra cover

3 Medium (i) North – rapidly eroding areas of soft mudstone and dunefields (ii) Far south – submarine areas of strong seafloor erosion

4 High (i) Centre – difficult access (steep, forest-covered) areas of moderate erosion and few Late Pleistocene deposits

5 Medium (i) North – rapidly eroding areas of soft mudstone and (ii) North – submarine areas with less seismic profile coverage

6 High (i) Additional fault zone segmentation possible

7 Low (i) All onshore areas – difficult access (steep, forest-covered) areas of moderate erosion and few Late Pleistocene deposits

(ii) All submarine areas

8 High (i) Southwest – difficult access and rapidly eroding areas of high rainfall

9 Medium (i) Southwest – difficult access and rapidly eroding areas of high rainfall (ii) Southeast – submarine areas where seismic profile analysis still needs

to be completed

10 Medium (i) Everywhere – inadequate density of high resolution seismic data, swath data absent

11 Medium (i) Northeast – areas of thick, young alluvium (ii) West and southwest – difficult access high rainfall and/or erosion areas

12 High (i) Additional fault zone segmentation possible

13 Medium (i) Far north – close to coast

14 Low (i) Everywhere –absence of high resolution seismic or swath data

15 High (i) Additional fault zone segmentation possible

Page 58: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

6

Table 5 Total fault slip rates and GPS velocities for representative transects across each domain (Fig. 2).

Domain Transect1 Total Fault Slip Rates GPS Velocities Differences

Total Net2 HTP3 HTN4

Total Horizontal5 HTP6 HTN7 HTP8 HTN9

1

3.2 -1.3 0.4 5.9 2.8 5.2 1.5 4.8 2 N 15.2 -9.1 0.00 9.6 -6.1 7.3 -3.0 7.3 2 M 5.9 -2.9 0.00 5.6 -4.3 3.7 1.4 3.7 2 S 4.4 -2.3 0.00 5.5 -1.8 5.1 -0.4 5.1 3 N 5.4 1.9 0.00 9.1 6.0 6.9 4.1 6.9 3 S 2.5 0.5 0.8 3.4 1.5 3.1 1.0 2.2 4 N 6.2 -1.1 3.2 6.6 -1.9 6.3 0.9 3.1 4 M1 7.1 0.04 6.9 5.4 -1.1 5.3 1.1 -1.5 4 M2 14.4 1.0 13.4 9.8 5.7 8.0 4.6 -5.4 4 S 22.3 2.4 19.2 15.5 8.8 12.8 6.4 -6.3 5 N 13.6 9.5 0.04 9.3 3.8 8.5 -5.7 8.4 5 M 9.7 7.1 0.2 20.6 8.9 18.6 1.9 18.3 5 S 14.4 6.7 7.9 15.4 10.4 11.3 3.8 3.3 7 N 0.7 0.4 0.00 8.4 5.9 5.9 5.6 5.9 7 S 1.0 0.6 0.00 6.1 4.1 4.6 3.5 4.6 8 N 37.8 4.0 33.2 19.0 2.2 18.8 -1.8 -14.4 8 S 23.5 0.7 22.5 16.8 -3.1 16.5 2.4 -6.0 9 N 2.3 1.2 1.2 5.3 2.2 4.8 0.9 3.7 9 M 3.1 1.4 0.7 8.7 1.2 8.6 -0.1 7.9 9 S >5.610 >1.210 >3.410 11.2 -1.3 11.2 0.0 7.8 10

1.0 -0.6 0.00 ND

11 N >3.910 >1.910 >1.210 22.5 9.7 20.3 7.8 19.0 11 M 3.9 2.5 0.2 19.9 9.2 17.7 6.7 17.5 11 S 1.2 0.2 0.5 22.1 8.2 20.5 8.0 19.9 13 N 3.8 1.8 0.00 ND 13 S 7.4 4.5 0.00 ND

Page 59: A model of active faulting in New Zealandhypocenter.usc.edu/research/CSEP/canterbury_data/Active... · 2013-09-24 · 1 A model of active faulting in New Zealand NJ Litchfielda, R

7

1 N = Northern, M = Middle, S = Southern. 2 These values, with their uncertainties and a confidence ranking, are plotted in Figure 2. 3 HTP = Horizontal Transect-Parallel, negative = extension, positive = contraction. 4 HTN = Horizontal Transect-Normal, negative = sinistral (none), positive = dextral. 5 ND = No Data, as one end of these transects are outside the velocity field. 6 HTP = Horizontal Transect-Parallel, negative = extension, positive = contraction. 7 HTN = Horizontal Transect-Normal, negative = extension, positive = contraction. 8 HTP = Horizontal Transect-Parallel. Absolute differences, so negative is where cumulative fault slip rates > GPS velocities. 9 HTN = Horizontal Transect-Normal. Absolute differences, so negative is where cumulative fault slip rates > GPS velocities. 10 Minimum because the transect crosses central Southern Alps Faults which have no slip rates.