You are on page 1of 19

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

net/publication/284122077

Definition and classification of fault damage zones: A review and a new


methodological approach

Article  in  Earth-Science Reviews · November 2015


DOI: 10.1016/j.earscirev.2015.11.006

CITATIONS READS

129 8,851

4 authors:

Jin-Hyuck Choi Paul Edwards


Korea Institute of Geoscience and Mineral Resources Pukyong National University
36 PUBLICATIONS   284 CITATIONS    3 PUBLICATIONS   132 CITATIONS   

SEE PROFILE SEE PROFILE

Kyoungtae Ko Young-Seog Kim


Korea Institute of Geoscience and Mineral Resources Pukyong National University
23 PUBLICATIONS   274 CITATIONS    385 PUBLICATIONS   4,818 CITATIONS   

SEE PROFILE SEE PROFILE

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

Nuclear Power Plant Site Investigation View project

Disruption of the Sino-Korean Peninsula View project

All content following this page was uploaded by Young-Seog Kim on 03 November 2017.

The user has requested enhancement of the downloaded file.


Earth-Science Reviews 152 (2016) 70–87

Contents lists available at ScienceDirect

Earth-Science Reviews

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

Definition and classification of fault damage zones: A review and a new


methodological approach
Jin-Hyuck Choi a,b, Paul Edwards a, Kyoungtae Ko a,c, Young-Seog Kim a,⁎
a
Department of Earth & Environmental Sciences, Pukyong National University, Yongso-ro 45, Nam-gu, Busan 608-737, Republic of Korea
b
Lithosphere Tectonics and Mechanics, Institut de Physique du Globe de Paris, 1 rue Jussieu, Paris 75005, France
c
Geologic Research Division, Korea Institute of Geoscience and Mineral Resources, Gwahang-no 124, Yuseong-gu, Daejeon 305-350, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: Although the widths of fault damage zones commonly show a positive correlation with displacements, these re-
Received 30 May 2014 lationships also show a somewhat scattered distribution. We believe that one of the fundamental reasons for this
Received in revised form 6 November 2015 problem is strongly related to subjective definitions and inconsistent uses of the term ‘damage zone’. Thus, firstly
Accepted 7 November 2015
we classify damage zones into along-fault, around-tip and cross-fault damage zones based on descriptive views of
Available online 10 November 2015
an arbitrary fault exposure as well as their tridimensional locations around a segmented fault system. Secondly,
Keywords:
we propose an advanced field technique and data acquisition method to more accurately define a damage zone
Fault scaling relationship using the distribution of cumulative fracture frequency. We tested this method on new field and borehole obser-
Fault damage zone vations as well as previously published data to identify damage zone boundaries, and express them as a change in
Damage zone classification slope gradients of the cumulative distribution of deformation structures. The results show how this slope change
Damage zone width can be a useful criterion in accurately defining the width of damage zones and some internal properties of fault
Cumulative fracture frequency zones. We argue that this damage zone classification and definition method should be adopted and used to pre-
vent discrepancies in field data. This will help us to gain a better understanding of fault damage zone properties
and their scaling with fault displacement.
© 2015 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2. Previous studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.1. Fault zone architectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.2. Scaling relationships between damage zone width and displacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.3. Limitation 1: subjective definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.4. Limitation 2: complexity in 3-D damage zone architectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3. Classification of fault damage zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.1. Along-fault damage zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.2. Around-tip damage zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.3. Cross-fault damage zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4. Estimation of damage zone width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.1. Research method based on cumulative fracture frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.2. Case studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.2.1. A segment of the Moab fault, SE Utah, USA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.2.2. Field and borehole observations at the Gyeongsang Basin, SE Korea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2.3. Reanalysis of the Pajarito fault system, New Mexico, USA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.1. Uncertainties of displacement measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.2. Defining damage zone and estimating its width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.3. Asymmetry of fault damage zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

⁎ Corresponding author.
E-mail address: ysk7909@pknu.ac.kr (Y.-S. Kim).

http://dx.doi.org/10.1016/j.earscirev.2015.11.006
0012-8252/© 2015 Elsevier B.V. All rights reserved.
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 71

6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

1. Introduction et al., 2006; Riley et al., 2010; Savage and Brodsky, 2011; Torabi and
Berg, 2011). Whilst the suggested parameters could account for some
Brittle faults were traditionally considered to be single planar struc- factors in the data scattering, the fundamental reasons of data scattering
tures, but more recently they are described as complex volumetric are still poorly understood (e.g. Blenkinsop, 1989; Evans, 1990; Shipton
zones composed of a variety of internal structures, such as slip surfaces, et al., 2006; Faulkner et al., 2010).
fault rock assemblages, and subsidiary deformation structures. Over the In quantitative fault studies, variations in damage zone width may
last few decades, a number of studies have focused on the fault zone be induced by genetic and evolutionary properties of the damage
architecture to understand fault evolution as well as its impact on zones. These properties include gradual, not drastic, changes in frequen-
fluid flow and mechanical behavior of the Earth's crust (Chester and cy of deformation structures (e.g. Shipton et al., 2006; Savage and
Logan, 1986; Aydin and Schultz, 1990; McGrath and Davison, 1995; Brodsky, 2011), the asymmetry between the hanging-wall and the foot-
Caine et al., 1996; Childs et al., 1996, 2009; Peacock, 2002; Walsh wall damage zone volumes (e.g. White et al., 1986; Knott et al., 1996;
et al., 2003; Kim et al., 2004; Shipton et al., 2006; Faulkner et al., 2010; Aarland and Skjerven, 1998; Berg and Skar, 2005; de Joussineau and
Smith et al., 2013). In particular, it is well known that the damage Aydin, 2007), and complex architectures related to multi-strands or
zones, consisting of subsidiary structures through relatively large braided fault cores (e.g. Childs et al., 1997; Faulkner et al., 2003, 2008,
volume of rock surrounding the fault core, are associated with fault ini- 2010). Uncertainties underlying this analysis may be more common
tiation, propagation and termination as well as its long-term evolution on the large-scale faults, which are generally composed of a number
(Segall and Pollard, 1980; Cowie and Shipton, 1998; Peacock et al., of fault segments, because the above-mentioned factors can affect the
2000; Scholz, 2002; Pachell and Evans, 2002; Kim et al., 2003; Fossen, geometric complexity of the fault zones in multiple ways.
2010; Gudmundsson et al., 2010). Damage zones are regarded as a Damage zone width, in general, has been defined by the frequency
key factor in a variety of geologic fields, such as the deformation pro- distribution of damage structures, such as cracks, fractures and defor-
cesses associated with faulting (e.g. Chester et al., 1993; Schulz and mation bands, which commonly decreases with distance from fault
Evans, 1998; Wilson et al., 2003), strain distribution and deformation core (e.g. Chester and Logan, 1986; Smith et al., 1990; Scholz and
history in a region (e.g. Scholz and Cowie, 1990; Walsh et al., 1991; Anders, 1994; Goddard and Evans, 1995). In other words, the external
Marrett and Allmendinger, 1992), earthquake rupture propagation edge of the damage zones is generally displayed as the point at which
and related seismic hazards (e.g. Sibson, 1989; Kim and Sanderson, the frequency of damage structures drops to a minimum or background
2008; Choi et al., 2012), and fluid permeability in the crust (Caine level (Beach et al., 1999; Cello et al., 2001; Agosta and Kirschner, 2003;
et al., 1996; Zhang and Sanderson, 1996; Childs et al., 1997; Evans Berg and Skar, 2005; Faulkner et al., 2006; de Joussineau and Aydin,
et al., 1997; Townend and Zoback, 2000; Jourde et al., 2002; Shipton 2007; Mitchell and Faulkner, 2009; Gudmundsson et al., 2010; Riley
et al., 2002; Faulkner et al., 2003; Geraud et al., 2006; Kim and et al., 2010). This method offers a useful tool for identifying damage
Sanderson, 2010). The last one is particularly important as it is used in zones, and hence has been broadly used in most of previous studies deal-
practical applications to ground water (e.g. Lopez and Smith, 1995; ing with damage zone width. However, the criteria used to define the
Bense et al., 2008; Cilona et al., 2015), hydrocarbon reservoirs and damage zone boundary have been varied in each field-measurement
ore-deposits (e.g. Aydin, 2000; Brogi, 2011; Rotevatn and Fossen, study, and these will be discussed in detail on Section 2. Inconsistent
2011), and the underground storage of CO2 (e.g. Shipton et al., 2004; calculations of damage zone width, therefore, may in part come from
Dockrill and Shipton, 2010). ambiguity and/or subjectivity of the definition and measurement of
Damage zones have been described in terms of architecture and geo- damage zone, regardless of its genetic properties.
metrical dimensions based on structural maps, field observations, and Furthermore, and perhaps most importantly, as faults are rarely
microstructural analysis (Chester et al., 1993, 2004; Antonellini and completely exposed in 3-D, interpreting the 3-D distribution of damage
Aydin, 1994; Bruhn et al., 1994; McGrath and Davison, 1995; Schulz zones is always challenging. These incomplete observations of fault
and Evans, 1998; Vermilye and Scholz, 1998; Storti et al., 2003; Billi zones have led to two slightly different uses of the term ‘damage zone’
et al., 2003; Micarelli et al., 2003; Di Toro and Pennacchioni, 2005; in modern structural geology (e.g. Schultz and Fossen, 2008; Kim and
Agosta and Aydin, 2006; Johansen and Fossen, 2008; Riley et al., 2010; Sanderson, 2010). The first, which is the more extensively used in a
Bistacchi et al., 2011; Hausegger and Kurz, 2013; Lin and Yamashita, cross-section of the fault zone, is referring to a highly deformed zone
2013; Smith et al., 2013; Walker et al., 2013). Damage zone architecture on both sides of the fault core (Chester and Logan, 1986; Ben-Zion
has been linked to slip accumulation and used in order to understand and Malin, 1991; Caine et al., 1996; Billi et al., 2003; Odling et al.,
fault growth and evolution (Cowie and Shipton, 1998; Shipton and 2004; Bullock et al., 2014). The other use is for local clusters of sub-
Cowie, 2001, 2003; Kim et al., 2003; Fossen et al., 2005; de Joussineau sidiary structures along fault traces, especially at fault step-overs
and Aydin, 2007; Childs et al., 2009). Numerous studies proposed fault and tips (Peacock and Sanderson, 1994; McGrath and Davison,
evolution models using the scaling relationship between damage zone 1995; Kim et al., 2000, 2003, 2004; Flodin and Aydin, 2004; Zhang
width and displacement (Evans, 1990; Scholz et al., 1993; Childs et al., et al., 2008). Note that these different uses are only related to the dif-
1996; Knott et al., 1996; Vermilye and Scholz, 1998; Beach et al., ferences in descriptive views of a fault zone, not physical and/or me-
1999; Fossen and Hesthammer, 2000; Shipton et al., 2006; Mitchell chanical characteristics of damage structures, and can cause an
and Faulkner, 2009; Faulkner et al., 2011; Savage and Brodsky, 2011; uncertainty to conduct a comprehensive and/or comparative study
Torabi and Berg, 2011). Although there is generally a broad positive of the damage zones. Some researchers have mentioned that these
correlation in this relationship, the collected data show a relatively terminological inconsistencies on estimating damage zone width
scattered distribution. Several parameters, such as lithology and associ- could be a major cause of data uncertainty in scaling relationships
ated diagenesis, depth of faulting, tectonic environment, and deforma- between damage zone width and displacement (e.g. Shipton et al.,
tion mechanism, have been suggested as responsible of the scattering 2006; Childs et al., 2009).
(Evans, 1990; Childs et al., 1997; Fossen and Hesthammer, 2000; We argue that damage zone terminology should be clearly defined
Shipton and Cowie, 2001; Di Toro and Pennacchioni, 2005; Shipton and classified based on clear criteria to help improve our understanding
72 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

and application of fault damage zones. For this purpose, we firstly exam- as a type of damage zones (e.g. Aydin and Johnson, 1978; Underhill
ine previous studies on damage zones and their scaling relationships and Woodcock, 1987; Mair et al., 2000; Fossen et al., 2007; Schultz
with displacement, and discuss the limitations as well as the necessity and Fossen, 2008; Rotevatn and Fossen, 2011; Tondi et al., 2012;
of improved research methods. Next, we propose a conceptual model Schueller et al., 2013). Damage zones composed mainly of deformation
showing a new classification of damage zones with three main catego- bands are characterized by some reduction in rock permeability unlike
ries; along-fault, around-tip and cross-fault damage zones, to avoid ter- those where fractures are dominant.
minological ambiguity and to reduce data uncertainty. Then we In some cases, particularly in large-scale faults, it is not easy to define
describe, in quantitative ways, damage structures and their frequency structural domains based on the simple description of fault zones due
distributions from outcrop and borehole data. In particular, we focus to geometrical complexity (e.g. Schulz and Evans, 2000; Shipton and
on changes in the gradient of cumulative frequency as a function of Cowie, 2001). Consequently, some researchers have suggested a need
the distance from fault core in order to define the critical boundaries to classify in detail the first-order fault zone model, which can vary
of a damage zone. In this stage, variations in damage zone width are due to different lithologies and tectonic settings (e.g. Gabrielsen and
also examined based on the above-suggested damage zone classifica- Aarland, 1990; Jones and Knipe, 1996; Braathen and Gabrielsen, 1998;
tion. Finally, we discuss some advantages of the suggested methodolog- Clausen et al., 2003). Most of them have proposed further subdivisions
ical approach as well as some other related issues; e.g. uncertainties in of the fault core and/or the damage zone respectively, such as central
displacement measurement, complexity of the 3-D fault zone architec- and distal cores and/or inner and outer damage zones (Jones and Knipe,
ture, and asymmetry of damage zones, which can supply limitations of 1996; Braathen and Gabrielsen, 1998; Heynekamp et al., 1999; Clausen
field-based data. et al., 2003; Micarelli et al., 2003; Berg and Skar, 2005). In some cases,
transitional or mixed zones are observed on both borders of the damage
2. Previous studies zones and these are characterized by the combination or repetition of
deformational features of different structural domains across a fault
2.1. Fault zone architectures zone (e.g. Evans and Bradbury, 2004).

Brittle faults are mainly classified into two structural domains 2.2. Scaling relationships between damage zone width and displacement
(Fig. 1); a central core and its enveloping damage zones, which can be
distinguished from the surrounding wall rock containing background Although damage zone width is commonly proportional to displace-
deformation (e.g. Chester and Logan, 1986; Chester et al., 1993; Caine ment (Knott, 1994; Knott et al., 1996; Beach et al., 1999; Fossen and
et al., 1996; Cello et al., 2001). The fault core is the result of highly local- Hesthammer, 2000; Shipton and Cowie, 2001; Childs et al., 2009), sever-
ized strain and intense shearing that accommodates the majority of the al researchers have reported that the relationships are scattered over
displacement within the fault zone, and generally consists of a number several orders of magnitude on power-law distributions (e.g. Hull,
of recurring slip surfaces and fault rocks such as gouges, cataclasites, and 1988; Knott, 1994; Shipton et al., 2006; Childs et al., 2009). In this sec-
breccias (e.g. Sibson, 1977; Bruhn et al., 1994; Childs et al., 1996; tion, we review previous studies and examine the reasons of scattering
Wibberley et al., 2008; Bastesen et al., 2009; Bastesen and Braathen, focusing on the various criteria used to define the damage zone. To re-
2010). Damage zones are characterized by relatively low strain and duce the effects of other factors, we reanalyzed previous data only
less intense deformation compared to the fault core, and these zones from normal faults. Also the data were classified into two groups; defor-
generally exhibit several second-order structures such as subsidiary mation bands in porous sandstones and fractures in other tight rocks
faults, fractures, and veins as well as fault-related folds and/or drag (Fig. 2a). Furthermore, when possible, we subdivided the data into
folds in some cases (e.g. Bruhn et al., 1994; Billi et al., 2003; Odling hanging-wall and footwall damage zones (Fig. 2b, c), and separately
et al., 2004; Berg and Skar, 2005; Faulkner et al., 2010). In porous analyzed the widths of them to take into account the well-known strain
rocks, strain localization commonly occurs through the formation of asymmetry between both fault blocks of dip-slip faults (e.g. Mandl,
deformation bands, and these low-displacement bands are also known 1988, 2000; Knott et al., 1996). The collected dataset, therefore, may

Fig. 1. Conceptual block-diagram of fault zone across a fault (modified from Caine et al., 1996; Faulkner et al., 2003). This diagram shows main components of fault zone architecture
(i.e. fault core, damage zone) and structural elements and features in each component. Note that transitional (or mixed) zone is characterized by alterations of fault elements from
fault core and damage zones.
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 73

Fig. 2. Log–log plots of damage zone width against displacement from the previous studies. Damage zone widths measured using joints/fractures and deformation bands are separately
denoted by open and closed circles, respectively; (a) total damage zone width, (b) hanging-wall damage zone (HWDZ) width, and (c) footwall damage zone (FWDZ) width vs. displace-
ment data. (d) Total damage zone width estimated by deformation bands (DB) are plotted separately based on their different criteria in defining damage zone.

represent a reliable scaling relationship unaffected by different tectonic 2.3. Limitation 1: subjective definitions
settings, lithologies, damage structures, and by asymmetric damage
patterns between both fault blocks. To estimate the damage zone width, two boundaries of damage
The datasets display scattering around three orders of magnitude for zones with fault core and wall rock should be carefully defined (see
a certain value of displacement (Fig. 2a), and this scattering seems to be Fig. 1). In field observations, the greatest difference between fault core
independent of the type of deformation structure and lithology. When and damage zone is the existence of original texture and/or fabric
the datasets are separated into hanging-wall and footwall damage of wall rock. In damage zones, even if deformational structures have
zones, the results are characterized by slightly different distributions, developed, they do not remove the original lithological properties of
such as different gradients or magnitudes for a given displacement the deformed wall rock (e.g. Gudmundsson et al., 2010). This remark-
(Fig. 2b, c). However, the scatter is still evident. able difference makes it easy to detect the inner boundary of damage
Fig. 2d shows several relationships between displacement and dam- zones. However, it is relatively difficult to detect precisely the outer
age zone width, on which the width was estimated by the frequency boundary of damage zone, because all physical properties change grad-
distribution of deformation bands but with different criteria of the dam- ually (e.g. Gudmundsson et al., 2010; Schueller et al., 2013).
age zone boundary. Each different criterion is introduced in Fig. 2d. To identify the outer boundary of damage zone, in general, the
There are two remarkable characteristics; the first one is that the rela- changes in fracture/deformation band frequency are examined based
tionships have slightly different distributions, for example, different on scanline survey normal to fault strike. In most of the cases, the
gradients of best-fit trends or different magnitudes with about one or boundary is determined as the distance at which an abrupt decrease
two order-ranges for a given displacement. This indicates that the data or falloff of the frequency occurs (e.g. Du Bernard et al., 2002; Micarelli
scattering of the relationships between damage zone width and dis- et al., 2003; Savage and Brodsky, 2011; Reyer et al., 2012). Due to the
placement may be controlled by the subjective definition of the damage lack of an accepted method to define the boundary, several researchers
zone. Second one is that data scattering still exists in each relationship defined the outer damage zone boundary using their own criteria as
although the scattering is greatly reduced. This means that there is an- follows: 1) frequency of structures equal to a value of background fre-
other fundamental reason which may result in data scattering indepen- quency measured at a location far away from the fault (Knott et al.,
dent on lithology, deformation element, and strain asymmetry between 1996; Beach et al., 1999; Mitchell and Faulkner, 2009; Riley et al.,
both faulted blocks as well as subjective definition. 2010), 2) the first occurrence of an unit interval of no damage structures
74 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

along a scanline (Schueller et al., 2013), 3) the edges of clusters of defor- Much of the previous data for damage zones, meanwhile, has come
mation bands (Fossen and Hesthammer, 2000; Shipton and Cowie, from traverses across faults (1 or 2-D measurements) without the con-
2001), 4) the farthest deformation band, even if it is an occasional isolat- sideration of fault exposure location around a 3-D fault volume. Howev-
ed structure beyond its clusters (Berg and Skar, 2005), and 5) the edges er, Kim et al. (2004) summarized various patterns of locally
of drag folds (Berg and Skar, 2005). concentrated damage structures along faults, and suggested a 3-D
These different criteria for defining the outer boundary of a damage model of fault damage zones showing the variety of damage patterns
zone have been applied in each study, and this difference results in a resulted from 1) different tip modes of propagating fault segments,
variation of damage zone width. Furthermore, using the same criteria, 2) different stages in the evolution of fault systems, and 3) the locations
results can vary greatly depending on the chosen unit interval of the of viewed exposure around the fault. Therefore, it is worth to consider
scanline survey and the background level of damage structures used that clusters of deformation structures, in the absence of interpreting
in each study (Du Bernard et al., 2002). This is well displayed in 3-D fault geometry, may create variations in damage zone width for
Fig. 2d, and therefore, we argue that subjectivity underlying this analy- quantitative analysis of complex fault systems. In short, we argue that
sis is one of the main factors of data uncertainty on the relationship be- complex damage zone architecture around faults, particularly com-
tween damage zone width and displacement. posed of a number of fault segments, is another fundamental reason
for data scattering in the relationship between damage zone width
2.4. Limitation 2: complexity in 3-D damage zone architectures and displacement.

There are a few different approaches to the study of fault zone archi- 3. Classification of fault damage zones
tecture; focusing on subsidiary structures around tips of fault segments,
as fault linkage is an important mechanisms of fault growth and evolu- A clear recognition of limitations on data collection, that is the esti-
tion (e.g. Peacock, 1991; Pachell and Evans, 2002; Kim et al., 2003; mation of damage zone width in one or two dimensions, can help to
Kim and Sanderson, 2005). A number of studies have commented that reduce some error and confusion during damage zone analysis. An ad-
the damage structures are concentrated around fault tips, or process vanced subdivision of damage zones, based on 3-D fault zone architec-
zones, and can be used to interpret the kinematics and dynamics of ture and its effect on strain distribution, should be carefully considered
faults (e.g. Peacock and Sanderson, 1994; McGrath and Davison, 1995; when we estimate damage zone width for quantitative analysis of com-
Kim et al., 2000, 2003; Flodin and Aydin, 2004; Myers and Aydin, plex fault systems. Here we suggest, therefore, that damage zones can be
2004; Zhang et al., 2008). Note that these observations may be classified into three categories; along-fault, around-tip and cross-fault,
completely opposite to the general relationship between damage zone mainly focusing on the concerns and observations of fault exposures in
width and displacement, as fault tips are generally characterized by terms of the complex 3-D fault zone architecture (Fig. 3).
a tapering out of fault displacement. Also, several researchers have
commented that damage zones in fault branch points and linkage 3.1. Along-fault damage zone
zones tend to be wider than the damage zone of isolated, straight or
simple fault segments (Shipton and Cowie, 2001; Shipton et al., 2006; The first category for fault damage zones, along-fault damage zone
Childs et al., 2009; Bastesen and Braathen, 2010; Lin and Yamashita, (AL-FDZ), is used to describe variations in damage structure along
2013; Schueller et al., 2013). Similarly, Knott et al. (1996) pointed out fault traces (Fig. 3). The AL-FDZ can be further subdivided into tip-,
that the width of damage zone, measured normal to the fault plane, is linking-, and wall-damage zones with each zone named using the termi-
greatest in the extensional quadrant of a given fault. Thus, the variations nology by Kim et al. (2004) (Figs. 3, 4a). Tip damage zones are charac-
in damage zone width along the fault are also strongly influenced by the terized by concentration of damage structures at the ends of a fault
spatial evolution and its related complexity of fault zone architectures trace (e.g. Segall and Pollard, 1983; McGrath and Davison, 1995; Kim
(Childs et al., 2009; Gudmundsson et al., 2010). et al., 2003, 2004). It has been proposed that tip damage structures are

Fig. 3. Simple 3-D schematic illustration of the fault damage zones around a segmented left-lateral fault (modified from Kim et al., 2003, 2004). Damage zones can be classified into along-
fault, around-tip, and cross-fault damage zones based on different exposures, concerns, and descriptive concepts. C indicates contractional zone and D indicates dilational zone.
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 75

generated by stress concentrations or large displacement gradients damage structures all around the faults. Therefore, we adapted a new
(Cowie and Scholz, 1992; Cowie and Shipton, 1998; Kim et al., 2003; descriptive concept, around-tip damage zone (AR-FDZ), to describe
Shipton and Cowie, 2003). Linking damage zones are developed by sec- slightly different geometries and kinematics of tip damage structures
ondary deformation between two fault segments through their interac- depending on slip sense and related tip mode (e.g. Kim et al., 2003,
tion and linkage (e.g. Martel et al., 1988; Peacock and Sanderson, 1991, 2004; Dor et al., 2006). For example, fault traces on horizontal surfaces
1995; Kim et al., 2004; Myers and Aydin, 2004). On slip-parallel fault can be parallel (strike-slip sense) or perpendicular (dip-slip sense)
traces, there are two main categories: dilational and contractional jogs, to the main slip sense, so their tips can belong to mode II or mode III, re-
depending on the direction of fault step-over and slip sense, that is, spectively (see Fig. 3). Note that mode II and mode III tips are two end-
the stress conditions in the linkages (Pakiser, 1960; Clayton, 1966; members of slip mode around a fault, and the majority of tip mode
Crowell, 1974; Suppe, 1985; Kim et al., 2000; Kim and Sanderson, around a fault would be mixed slip modes (Fig. 3).
2010). Wall damage zones are the secondary fractures developed in Mode II tip damage zones are defined as secondary deformations at
the wall along a fault trace, and may have developed by the abandon- slip-perpendicular tips, and have been generally described in outcrop-
ment of tip damage zone or newly generated damage structures associ- scale normal faults in cross-section (Peacock and Sanderson, 1994)
ated with fault evolution (Kim et al., 2004). and strike-slip faults in map-view as in Figs. 4a and 5 (e.g. McGrath
Note that the AL-FDZ has typical patterns, which are dependent on and Davison, 1995; Kim et al., 2000, 2003; Flodin and Aydin, 2004;
their location along a segmented fault and on the stage of the evolution Zhang et al., 2008). Whereas, mode III tip damage zones can be repre-
of a fault (see Kim et al., 2004). Although the AL-FDZ has been generally sented by a zone where secondary structures develop, often without a
described for slip-parallel fault traces based on outcrop-scale faults through-going main fault trace, around a slip-parallel fault tip (Fig. 5).
(Peacock and Sanderson, 1994; McGrath and Davison, 1995; Kim et al., Linking damage zones can be regarded as the combinations of the
2000, 2003; Flodin and Aydin, 2004; Zhang et al., 2008), its geometric two tips of previously isolated fault segments, which can occur at the
and kinematic characteristics were also presented on map-scale faults mode II, mode III or mixed mode tips of each fault segment (Walsh
(Storti et al., 2003; Kim and Sanderson, 2006) and on earthquake surface et al., 1999). In some cases, therefore, linking damage zones also can
ruptures (Ambraseys and Tchalenko, 1969; Lin et al., 2001, 2004; be described as mode II and mode III end-member linkages, and the
Choi et al., 2012). Note that the suggested descriptions belonging to most well-known example of the latter case is a relay ramp along nor-
the AL-FDZ can be used regardless of the angular relationships between mal faults (see Fig. 4b).
fault trace and slip mode, and this is the main difference from the Tridimensional tip damage structures can provide useful informa-
around-tip damage zones, which are introduced in the next section. tion on fault evolution, and their linkage is also regarded as an impor-
tant factor in fault growth (Peacock, 1991; Childs et al., 1995; Kim
3.2. Around-tip damage zone et al., 2004; Soliva and Schultz, 2008). For example, fault evolution
studies based on scaling relationships between fault trace length and
Although the concept of AL-FDZ is useful to describe various types displacement can vary depending on dominant slip components; dip-
of damage structures, there are some limitations for description of tip and strike-slips, because different tip modes at fault discontinuities

Fig. 4. Examples of along-fault damage zones. (a) Left-lateral fault system developed on a granitic rock in Uljin area, Korea. Along-fault damage zone can be subdivided into tip-, linking-,
and wall-damage zones depending on the locations of damage structures along fault traces (Kim et al., 2004). (b) Relay ramp formed at a linking damage zone between two normal fault
segments in Arches National Park, Utah, USA.
76 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

Fig. 5. Right-lateral fault system developed on a granitic rock in Tondosa-temple area, Korea. Damage patterns are various depend on their locations along the segmented fault as well as tip
modes.

may result in different linkage-related extension of fault length. We two limitations; resolution of any frequency change and noise for ran-
suggest that when damage zones are described as AL-FDZ (tip-, linking-, dom fluctuation due to small change depending on location of an indi-
and wall-damage zones), it is useful to consider their relationships vidual fracture, producing uncertainty in the damage zone width.
with slip mode and location around the main fault, based on the AR- Analysis using the cumulative frequency of damage structures across
FDZ concept, because we could easily understand tridimensional dam- fault zone was plotted in Fig. 6d. Each structure is plotted at its true
age patterns and their kinematics. distance along the traverse and the slope of the cumulative frequency indi-
cates changes in the density of damage structures (Fig. 6d). It preserves
3.3. Cross-fault damage zone the individual location of each fracture and often allows a more accurate
determination of similar density zones in frequency. In this example,
When we study across a fault, it is not easy to recognize the entire the density in the wall rock is ~ 2.4 m−1 and that in the damage zone
3-D fault zone architecture in a given fault exposure. In other words, is ~ 6 m− 1, and their boundary locates at distance of 6.7 m from the
particularly in large-scale faults, it is very hard to classify damage fault core. These results come from the intersection point of the two
zones into the previously suggested categories (i.e. AL- and AR-FDZs). slopes with different gradients, and we suggest that this method can
Thus, in fact, most of the studies on damage zones have generally de- reduce inconsistency in definition of damage zones.
scribed damage structures in sections orthogonal to fault strike with- A curve of cumulative frequency as well as its shape and gradient can
out considerations of 3-D fault zone geometry. It could actually be an be very useful to represent and to define different degrees of deforma-
important reference to understand the deformation and fluid flow tion across the fault zone (Genter et al., 1997; Berg and Skar, 2005).
characteristics across a fault zone away from the main slip surface. We believe that the cumulative fracture frequency method can be
Here we suggest, therefore, that the data acquired in fault-cross sections used at all scales; micro fractures, field observations, and lineament
should be introduced as another category of fault damage zones, cross- analysis, but it is necessary to clearly define fracture resolution in each
fault damage zone (CR-FDZ) to avoid terminological confusions (Fig. 3). study.
Based on the CR-FDZ concept, several models have been suggested
with a further subdivision of the damage zone into different domains, 4.2. Case studies
such as transition zone, mixed zone, intensely deformed zone, and
inner/outer damage zones depending on local strain distribution (Jones We carried out a comparison analysis of the ‘classic’ interval and
and Knipe, 1996; Heynekamp et al., 1999; Micarelli et al., 2003; Berg ‘newly suggested’ cumulative fracture frequency methods on field and
and Skar, 2005). We argue that although damage zone width under borehole data. Furthermore, if possible, we examined the measurement
the CR-FDZ concept is the most easily accessible quantitative data, its location along a fault trace to consider the effect of 3-D fault zone
further classification should be, if possible, carefully examined based architecture.
on the 3-D model of fault damage zones to reduce variations and uncer-
tainty of data. 4.2.1. A segment of the Moab fault, SE Utah, USA
The Moab fault is a well exposed normal fault, with a 45 km long
4. Estimation of damage zone width trace and maximum throw of about 960 m, which runs approximately
NW–SE in SE Utah (Fig. 7a). The northern part of the Moab fault is com-
4.1. Research method based on cumulative fracture frequency posed of numerous splay faults, and one of these splays, the Bartlett
fault, is the focus of this study (Fig. 7b). A sedimentary sequence of
To reduce data uncertainty associated with measuring damage zone Jurassic to Cretaceous age is cut by this fault segment, and the estimated
width, methodological limitations related to various criteria for the throw are in the range between 250 m and 285 m (e.g. Davatzes and
boundary of damage zones should be mitigated by consistent use of Aydin, 2005). The Bartlett fault is well exposed in several canyons
damage zone definitions in future studies. An analytical method, (Fig. 7c), and this has led a number of studies into fault zone architec-
which is most commonly used to define damage zone using the number ture at this location to understand fault zone properties and the effect
of damage structures per unit length along traverses across the fault of lithology on fault growth (e.g. Foxford et al., 1998; Davatzes and
zone, is schematically represented in Fig. 6a, b. This is a hypothetical dis- Aydin, 2005; Fossen et al., 2005, 2007).
tribution of damage structures across a fault zone, where the density is Here, we carried out scanline analysis at five sites along the Bartlett
~2 m−1 outside the damage zone and increases from ~4 m−1 to ~8 m−1 fault to estimate damage zone width and to investigate its variations
through the damage zone. Fracture frequency is shown by bar diagrams along fault strike (Fig. 7b). To reduce several factors of data uncertainty
on Fig. 6b, and damage zones can induce different widths; ~8 and ~9 m related to damage zone width, such as lithology and strain asymmetry
respectively, depending on the bin size. Fig. 6c showing the damage between both fault blocks, we focused only on the distribution of defor-
structures around 9 m indicates that the damage zone boundary can mation bands in the uppermost layer of the footwall; an eolian sand-
also be in the range of 8.8–9.3 m. As a result, bar diagrams can involve stone belonging to the Jurassic Moab Member (Fig. 7b, c).
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 77

Fig. 6. Schematic diagrams comparing different analytical methods to estimate the damage zone width. (a) Images showing the fracture distribution across a hypothetical section. (b) Bar
diagrams showing the fracture frequency based on unit intervals of 2 and 1 m. (c) Inset for fracture distribution around the estimated damage zone boundary. Note that different damage
zone widths can be measured by different unit intervals as well as researchers. (d) Plots of the cumulative fracture frequency and a distribution of its slope gradients. The intersection point
between two different gradients indicates outer boundary of damage zone.

The cumulative frequency of deformation bands is plotted as a dot, on slope gradient, and this difference indicates a different strain distri-
and these results are compared to the normal frequency distribution bution depending on the distance from the fault core.
of 1-m-interval, which are displayed as bar diagrams (Fig. 8). Along One of the most interesting points is that the slope gradients in the
a 20 m long scanline at Site.1, deformation bands are developed in- outermost sections of each site are similar; the density of deformation
tensely within 4 m from the fault core, although frequency data are bands in the range of between 1.5 and 3 m− 1. We suggest that this
collected from three partially exposed parts (Fig. 8a). It is, however, value represents background level of deformation band frequency in
not easy to accurately define a damage zone boundary based on bar this study area. Consequently, this value can be used to determine the
diagrams because incomplete data indicates the minimum frequency damage zone boundary where a single slope fitting for the damage
of developed deformation bands. This limitation can be overcome zone is inappropriate, such as Sites.1 and 5 (Fig. 8a, e). In the case
using slope gradients of cumulative frequency, that is, frequency data of Site.1, damage zone boundary can be determined as 8.73 m based
for the unexposed parts can be inferred using surrounding slope gradi- on the outermost background density. As an abrupt change in slope gra-
ents. Corrected data of cumulative frequency indicates that three sec- dient is visible at the distance of 4.63 m from fault core, we argue that
tions with different slope gradients have two intersection boundaries the inner section near the fault core can be regarded as intensely de-
at 4.63 and 8.73 m. formed zone and the outer section as transition zone between damage
It is relatively hard to determine damage zone width using a bar zone and wall rock. In a similar way, at Site.5, although the damage
diagram along a 17 m long scanline at Site.2 due to its irregular changes zone boundary can be determined as 26.84 m, the density of deforma-
as well as partial exposures (Fig. 8b). However, the slope profile of tion bands from 12.52 m to 21.25 m is almost the same with the inferred
corrected data using cumulative frequency is characterized by two dis- background density. Thus, we argue that the damage zone width can be
tinctive sections with different slope gradients. Thus, the damage zone estimated as 12.52 m. Note that the local steep gradient may result from
boundary can be determined to be at 10.96 m where the two slopes in- other controlling factors such as the effects of a subsidiary fault and its
tersect. The damage zone boundaries at Sites.3 and 4 can be determined related deformation band clustering outside of the first-order damage
in the same way at 14.37 and 2.57 m, respectively (Fig. 8c, d). Note that, zone.
at Site.3, the result is based on reanalysis of previous data from Berg and Based on the cumulative frequency of deformation bands, the esti-
Skar (2005), that is, the cumulative frequency is replotted based on the mated damage zone widths are in the range of between 2.57 and
unit interval data (Fig. 8c). Along the longest scanline of 40 m at Site.5 14.37 m along the Bartlett fault (Fig. 8). Berg and Skar (2005) have ex-
(Fig. 8e), the bar diagram shows an abrupt decrease in frequency at amined the frequency of deformation bands within footwall damage
13 m and a convex pattern from 20 to 25 m. Similarly, the most visible zone at the same locations; Sites.1–3, and their results also show that
change in slope gradient on the cumulative frequency distribution the density of damage structures varies in a range of between 3.3 and
is marked at 12.52 m, other intersection points are also observed at 8.9 m− 1. These variations are very marked because the estimated
21.25 and 26.84 m. throws are almost the same at the five sites (e.g. Foxford et al., 1998;
Although the bar diagrams show a general decrease in deformation Davatzes and Aydin, 2005). The difference in damage zone width, in
band frequency away from the fault core, these changes are not always particular, is visible at Site.4 which is part of a relay zone near Bartlett
systematic; e.g. Sites.2 and 5 (Fig. 8b, e). These local variations can result Wash (see location of the Site.4 on Fig. 7b). Note that the stress field
from missing data due to partially covered outcrops or the effects around a relay zone can be perturbed by relay geometry and evolutional
of clustering. The cumulative frequency plots display a broken-line- stage (Segall and Pollard, 1980; Crider and Pollard, 1998; Long and
graph pattern at all sites (Fig. 8). The frequency distribution at each Imber, 2011). On the other hand, in general, the wider the damage
site can be classified into two (or more than two) sections depending zone, the lower the slope gradient, and thus the variations in damage
78 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

Fig. 7. (a) Simplified structural and locality map showing the Moab fault system. (b) Local geology and fault geometry around the study area. (c) Cross-sectional view of the fault zone at
Bartlett Wash (Site.3). Deformation band frequencies are counted based on scanline analysis at five sites in Moab Member sedimentary layer. Collected data to estimate damage zone
width came from the top layer in footwall side of the fault.

zone width may be related to differential localization of deformation in several minor faults and dykes (Fig. 10a). Sedimentary strata on this
each damage zone. Thus, we argue that the variation of damage zone slope belong to Chilgok Formation, which is composed of tens-of-
width along the Bartlett fault may result from fault linkage or irregular- meter-thick massive tuffaceous sandstones intercalated with mostly
ity in deformation concentration around faults. reddish and purplish mudstones. The distribution of fracture frequency
is analyzed to examine the scale of damage zone associated with neigh-
4.2.2. Field and borehole observations at the Gyeongsang Basin, SE Korea boring fault zones or fault-lineaments.
The Gyeongsang Basin in southeast Korea contains a Cretaceous se- To investigate the fracture distribution, the number of fractures per
ries of mainly lacustrine siliciclastic and volcanic rocks (e.g. Chough meter as well as their cumulative frequency was measured along a
and Shon, 2010). The basement rocks have been intruded by Creta- 93 m long scanline. About 1100 fractures traversing the scanline from
ceous and Tertiary igneous rocks; mainly granite batholiths and both the east to the west were plotted (Fig. 10b). The results show that frac-
mafic and felsic dikes (Fig. 9). The NNE–SSW trending Yangsan Fault ture density as well as gradient of cumulative distribution generally in-
and NNW–SSE trending Ulsan Fault are the major structural features creases towards the east along the slope. Thus, numerous second-order
in the Gyeongsang Basin, and several Tertiary basins are locally devel- structures, such as minor faults and fractures, on this slope may belong
oped within the eastern block of the Yangsan–Ulsan fault system. to a portion of a fault damage zone (Ko et al., 2012).
In this study, we focus on fracture distributions on a road-cut slope We argue that although the general trend of fracture frequency is
and two borehole samples in the western part of the Yangsan fault easily recognized on interval plots, two section boundaries of the frac-
(Fig. 9). ture frequency (17.7 and 39.3 m) can be more clearly defined by slope
changes in cumulative distribution (Fig. 10). In this outcrop scale, there-
4.2.2.1. Field observation at a road-cut slope in Daegu. Due to rapid expan- fore, the lowest slope gradient on the farthest west section may indicate
sion of the road network in South Korea, some of numerous road-cut a background level of the fractures around the study area. Note that the
slopes have been exposed on well-bedded Cretaceous sedimentary minimum width of the damage zone can be estimated as 39.3 m, and
rocks in the Daegu area (see location in Fig. 9). Although there is no doc- two distinguishable sections based on slope gradient within damage
umented large-scale fault around the study area, a number of fracture zone may indicate a different strain distribution depending on the dis-
zones and fault-lineaments are reported from a slope stability analysis tance from the fault core (e.g. inner and outer damage zones).
(Ko et al., 2012). Here, we focused on an E–W trending road-cut slope, Local clusters of fractures are remarkable on the bar diagram, and
which is highly unstable due to exposed dense vertical fractures with this variation is mainly observed around minor faults and individual
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 79

Fig. 8. Deformation band distributions; normal frequency as bar diagrams and cumulative frequency as dots, against distance from the fault core at each site. Note that it is relatively easy to
detect the outer boundary of damage zone using the slope gradients of cumulative frequency data. At Site.3, the slope is drawn by cumulative numbers of previously sampled data based on
interval frequency from Berg and Skar (2005).

dykes (Fig. 10). This means that strain distribution within a damage gradient (Fig. 10). In other words, interferences of local effects can be re-
zone is controlled by local lithological or structural variations. One inter- moved when describing fracture frequency related to a main fault.
esting point is that this variation is also displayed as a step-like pattern Therefore, this case study demonstrates another advantage of cumula-
in the cumulative frequency distribution. However, it is relatively hard tive frequency as an analytical method for, particularly large-scale,
to detect these local variations because of the smooth changes in slope fault damage zone.
80 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

Fig. 9. Simplified regional geologic map of the Gyeongsang Basin, SE Korea (modified from Chough and Shon, 2010), and locations of field and borehole sites.

4.2.2.2. Borehole observations. Fractures create zones of high permeabil- Firstly, we analyzed a 1200 m deep vertical borehole data showing
ity within a rock mass, and hence their spatial distributions at depth a total core recovery of 99.8% (data from Geotech Consultant Co., Ltd.).
have been analyzed by borehole analysis for petroleum, geothermal res- Core sample is acquired at the northern part of the Gaeum fault sys-
ervoirs, and carbon capture and storages (e.g. Genter et al., 1997). In re- tem, which is characterized by WNW–ESE trending left-lateral faults,
cent years, as drilling technologies are advanced, it is possible to take around the Euiseong area (Fig. 9). It is composed mainly of alternating
borehole samples across a whole fault zone with high core recovery, layers of mudstone and sandstone, with numerous subvertical frac-
and to observe fault zone architecture in detail (e.g. Hung et al., 2009; tures observed through the core, even though the borehole is vertical
Li et al., 2013; Olierook et al., 2014). In this study, we carried out fault (Fig. 11a).
zone analysis based on fracture frequency from two borehole samples The number of fractures within each interval of 3 m is plotted against
around map-scale faults in the Gyeongsang Basin, SE Korea (Fig. 9). depth as bars (Fig. 11b), and the cumulative number of fractures is also
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 81

Fig. 10. (a) Road-cut section characterized by the existence of minor faults, dyke, and dense factures in Daegu area, SE Korea. (b) Distribution of fracture frequency based on both interval
and cumulative methods. The results indicate that the damage zone extent is at least 39.3 m based on the intersection point of different slope gradients.

drawn as dots (Fig. 11c). The depths of fault and fracture zones, which distributions is relatively arbitrary. We argue that this problem can be
are confirmed based on ATV (Acoustic Tele-Viewer) observation, are solved by using gradient changes in the slopes of cumulative distribu-
marked as cross bars to compare with fracture density distribution. tion. Thus, each section that shows different fracture characteristics
Although changes in fracture density with depth are visible on the can be classified and defined by the same cumulative distribution gradi-
bar graph, evaluating the boundaries of each sections using fracture ent. The boundary between each section can also be clearly detected by

Fig. 11. Borehole observation focusing on fracture distributions with depth in Euiseong area, SE Korea. (a) Examples of subvertical fault and fracture zones in core sample, with white ar-
rows indicating traces of fault gouge zones. (b) Bar diagrams showing fracture frequency per unit interval of 3 m. (c) Slope distribution based on cumulative fracture frequency. At least
four damage zones and their apparent thicknesses can be defined based on cumulative frequency data of fractures.
82 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

the intersection points between two different slope gradients (see fracture zones. Next, all fractures are classified into four sets depending
dashed cross-lines in Fig. 11c). on the dip angle of fractures; Set.1 for sub-horizontal, Set.2 for low
We recognized four sections that can be defined as fault damage angle, Set.3 for high angle, and Set.4 for sub-vertical fractures. The
zones, which are detected on both ATV observation and cumulative step-like pattern on cumulative distribution is much clearer from the
fracture frequency (Fig. 11c). Note that we excluded the upper part data of Sets.2 and 3, whereas a straight-line pattern is shown from the
with an extremely high fracture density as it is most likely associated data Set.1. Note that the number of data for Set.4 is very small due to
with drilling induced fracturing. One interesting point is that the both vertical borehole and fractures. Consequently, we can assume
lower two damage zones show about 100 m of thickness in the bore- that the fractures with sub-horizontal dip angle (Set.1) may be devel-
hole sample. We argue that this can be associated with a lithological oped due to artificial effects related to drilling and that these should
variation or due to the effect of the Gaeum fault. Although there are be removed to evaluate the distribution of natural fractures. Eventually,
analytical limitations since the borehole sample is effectively 1-D all data except Set.1 are plotted again for interval and cumulative distri-
data, this case study clearly shows that changes in fracture distribu- butions (Fig. 12b).
tion can be more clearly displayed by the cumulative distribution of The result shows that slope gradient is much more gradual (i.e.
fractures. higher fracture density) at the five sections (damage zones), and their
Secondly, a similar approach is examined on 300 m deep vertical depths are very similar to the locations of the fault and fracture zones.
borehole data sampled near the southern Yangsan Fault, one of the We argue, therefore, that at least five damage zones exist within the
major right-lateral faults with a total displacement of about 30 km, in borehole of 300 m depth, and this may indicate that the studied site
the Eonyang area (Fig. 9). The Yangsan Fault is geometrically character- probably belongs to the western damage zone of the Yangsan Fault
ized by pull-apart basin in a right step of the fault at the study area, and (see location in Fig. 9). One other interesting point is that some damage
its damage zone is mainly composed of second-order normal faults and zones are characterized by a slightly curved pattern in the cumulative
extensional fractures (Choi et al., 2009). The drilling site is located with- distribution (see FDZs 1 and 5 on Fig. 12b). This implies that the number
in Cretaceous granite, and hence we assume that fracture frequency is of fractures gradually increases downward in the damage zones, and
not controlled by variations in lithology or structural discontinuities abruptly decreases beyond the density peak. In fact, this feature is
such as bedding. Fracture frequency is measured along the borehole more visible in the bar diagram as an asymmetric pattern (Fig. 12b).
sample, and the locations of fault and fracture zone are marked by As plotted data are almost inclined fractures, one possible interpretation
cross bars (Fig. 12). Here, fault zones are defined by the existence of a for this result is asymmetric distribution of fault-related fractures be-
slip surface and/or fault gouge, and fracture zones mean sections com- tween the hanging wall and footwall. This result is consistent with
posed of rock fragments due to dense fracturing. other studies where the width of the hanging wall damage zone is
The number of fractures within a unit depth of 1 m is plotted as bars wider than that of the footwall (Knott et al., 1996; Mandl, 2000; Berg
and the cumulative distribution is plotted as dots (Fig. 12a). About 1200 and Skar, 2005).
fractures were measured and the average fracture density is 3.6 m−1 for
the entire borehole. Note that some sections that show high fracture 4.2.3. Reanalysis of the Pajarito fault system, New Mexico, USA
density, which is marked by gentle gradients of cumulative fracture fre- Riley et al. (2010) studied the Bandelier Tuff cut by the Pajarito
quency, are relatively well matched with the locations of fault and normal fault system to estimate damage zone width and to

Fig. 12. Fracture distributions; normal frequency as bar diagrams and cumulative frequency as dots, with depth on a borehole into granitic bedrock in Eonyang area, SE Korea. (a) Four
fracture sets are classified by dip angle of fracture and their cumulative frequencies are separately plotted with depth. Note that there is no variation in slope gradient for sub-horizontal
fractures (Set.1). (b) Replotting of the fractures except Set.1, and the result shows that at least five sections are related with fault damage zones (FDZ).
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 83

understand its controlling factors. They examined the influence of li- Here, the reanalyzed data show that the cumulative frequency
thology, pre-existing structures (or anisotropy), and strain asymmetry distributions from different sites have almost consistent points. Firstly,
between the hanging-wall and the footwall on damage zone properties. in all cases, they can be subdivided into three sections based on slope
These authors displayed fracture distribution using the interval fre- gradient. Secondly, although each innermost section has a different
quency method and suggested more than 10 fractures/10 m to be the width, their slope gradients are almost the same; fracture densities
lower limit of the damage zone related with faulting. Here, we of 1.98, 2.16, and 2.15 m−1, respectively. This may imply that the defor-
reanalyzed the same data using the cumulative frequency method to mation mechanisms (or strain localization) associated with faulting
examine the advantages of each method by comparing the results of is almost the same along the fault. Lastly, the slope gradients of the out-
the different analytical methods. most sections are similar; fracture densities of 0.58, 0.63 and 0.61 m−1,
The Pajarito fault system was mapped by Lewis et al. (2009), and they respectively. This is associated with the degree of background fracturing
estimated throw profiles along segmented fault traces (Fig. 13a, b). within the wall rock. We argue that, therefore, cumulative fracture fre-
In this study, we focus on only the fracture data distributed on the quency method can also be used to define damage zones and measure
hanging-wall block in welded ignimbrites to reduce the influence of their widths using previous data with some modification. Note that
other controlling factors (see section 4.1 on Riley et al., 2010). These the part with a relatively lower slope gradient than inner damage
data were measured at three exposures along the canyon perpendicular zone may be related with differential strain distribution within the
to the fault trace; Sawyer, Bland and Frijoles (Fig. 13a), which show dif- damage zone, identified as a transition zone between the damage
ferent fault throws in three canyons; 30 m, 115 m and at least 160 m, zone and the wall rock (Micarelli et al., 2003). The estimated damage
respectively (Fig. 13b). In terms of fault geometry, Bland and Frijoles zone widths are not directly proportional to the fault throw, and this
Canyons are located around branch points of the main fault traces. may result from the complex 3-D geometry of the Pajarito fault system,
Also, Frijoles and Sowyer Canyons are located near a linking damage such as fault branches and step-overs (Fig. 13). In fact, the variation in
zone between the MPF (Main Pajarito Fault) and NPF (Northern Pajarito damage zone width due to fault zone complexity can be a good target
Fault) segments. for further studies, and we believe that the new damage zone categories
Riley et al. (2010) have estimated the damage zone width to be 80 m must be helpful to improve future research results.
at Sawyer Canyon and 160–170 m at Bland and Frijoles Canyons (see
the bar diagrams on Fig. 13c). We used the total number of fractures 5. Discussion
within each 10 m transect to draw the slope of cumulative fracture fre-
quency. This alternative method can be applied for previously reported 5.1. Uncertainties of displacement measurement
data based on interval frequency analysis. In this stage, sections with no
data are corrected based on the slope gradient around them, and we de- In many cases of fault scaling relationships, variations in data uncer-
fined the damage zone width using the intersection points of different tainties related to displacement have been considered as one of the main
slope gradients (Fig. 13c). As a result, the damage zone boundaries sources in comparing different data sets (e.g. Walsh and Watterson,
based on the intersection points of different slope gradients can be de- 1988; Kim and Sanderson, 2005). The most common factor affecting
termined as 202.2, 179.1 and 135.4 m, respectively. Our estimated dam- data uncertainties may be 1 or 2-D measurements of apparent displace-
age zone boundaries are slightly different from the results of Riley et al. ment along a fault trace, i.e. separation (e.g. Walsh and Watterson,
(2010), and the differences reach up to 42.2 m at Bland Canyon and 1988). Fault ‘throw (or vertical separation)’, which we discussed in
55.4 m at Sawyer Canyon. two case studies for normal faults, can be one example of data

Fig. 13. Reanalysis of fracture distributions at three canyons across the Pajarito fault system. (a) Overview map showing detailed geometry of the Pajarito fault system around the study
area (Lewis et al., 2009). (b) Throw variations along the segmented fault traces (Lewis et al., 2009). (c) Comparison analysis for the distributions of fracture frequency between interval and
cumulative methods. Fracture data is collected against distance from the fault core to the hanging-wall side at three sites. Note that the estimated damage zone widths are slightly different
depending on the research methods. See the text for further description.
84 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

limitations related to displacement measurements. Furthermore, the propagation associated with earthquakes (e.g. Di Toro et al., 2005;
estimated displacement near fault linkage zones can be various and/ Griffith et al., 2009; Bhat et al., 2010) and unilateral fault/rupture prop-
or abruptly change depending on different evolutional stages of seg- agation (e.g. Kim and Sanderson, 2008; Choi et al., 2012). Note that as
ment linkage (e.g. Kim and Sanderson, 2005). For example, displace- these asymmetric damage zones can be a considerable variation in sta-
ment can be almost zero near early fault segment linkage, when two tistical analysis related to cross-fault damage zone width, it is necessary
fault tips begin to interact. However, more hard linked segments to take into account the details of the 3-D fault zone architectures and
tend to accumulate greater displacement around the linkage zone as related asymmetries.
the two segments are acting as one through-going fault.

5.2. Defining damage zone and estimating its width 6. Conclusions

In all case studies, we have tried to compare analyzed results of both Although fault damage zone is one of the most important issues
interval and cumulative distributions of fracture frequency to define the in modern structural geology, the term has been used in a variety of
damage zones across a fault. Firstly, the analytical method using interval ways depending on the concerns of researchers. This has led to inconsis-
frequency is more representative to variation in the data. For instance, it tencies in summarizing several data sets from different researchers. In
may be useful to detect the local effects, such as the existence of second- this paper, therefore, we classified the broad terminology ‘damage zone’
order faults or asymmetric damage patterns, within the first-order dam- into three main categories; along-fault, around-tip, and cross-fault dam-
age zones (see Figs. 10b, 12b). However, it is relatively difficult to detect age zones, and suggested an advanced method to define the boundary
the boundary of the damage zone, and this problem mainly results from of cross-fault damage zone using the distribution of cumulative fracture
largely inconsistent criteria for defining the limit of the damage zone density. The following main conclusive points can be drawn from our
depending on the authors' criteria in each study. We argue that this study.
limitation can be partly overcome by a unification of unit intervals and
by the establishment of damage zone criterion using background frac-
(1) In general, damage zone widths have been estimated based on
ture density based on cumulative frequency in a given study area.
the distribution of fracture frequency per unit interval across a
We have confirmed several advantages of using the cumulative
given fault, and this method is suitable to detect data variations,
fracture frequency and its slope gradients to define damage zone
such as local effects by second-order faults. Data uncertainties
boundaries; 1) obvious detection for the outer edges of the damage
related to this analytical tool, however, can occur due to incon-
zones based on the intersection point between different slope gradients,
sistent unit intervals as well as subjectivity of researchers in de-
2) less influence of local lithological and structural effects on general
fining background fracture density and/or damage zone width.
distribution, and 3) correlation of the unexposed parts based on sur-
(2) The outer boundary of damage zones can be detected by the
rounding slope gradients. We suggest, therefore, that cumulative fre-
intersection points of different slope gradients on cumulative
quency method can be used to reduce some uncertainty, which can be
fracture distribution. Sampling constraints related to partly un-
derived by subjectivity and inconsistent research methods, in damage
exposed outcrops can be covered based on neighboring slope
zone analysis. Furthermore, it is possible to subdivide the estimated
gradients, and the damage zone can easily be subdivided into
damage zones depending on the steepness of slope gradient, that is, in-
subsections showing different deformation intensities.
tensity of deformation. Eventually, we argue that a combination of both
(3) In this study, we have suggested an analytical method for defin-
interval and cumulative frequency methods can be a critical tool in
ing the widths of CR-FDZ because the data from many previous
defining damage zone and estimating its width, and can reduce data
studied faults are inconsistently collected without a clear defini-
scattering in scaling relationships between damage zone width and
tion of the damage zone boundary and a classification of 3-D
fault displacement.
fault zone architecture. As a result, the estimated damage zone
widths vary depending on not only well-known fundamental fac-
5.3. Asymmetry of fault damage zones
tors, such as displacement, but also complex fault geometry, such
as fault branches and linkages. We firmly believe, therefore, that
Generally, two separate damage zones develop on both sides across
further classification of the damage zone, into along-fault and
a fault and asymmetric widths of these pairs has frequently been
around-tip damage zones based on the interpretation of the 3-D
reported in many previous publications (Aydin and Johnson, 1978;
fault zone geometry, will be helpful to reduce data scattering on
Koestler and Ehrman, 1991; Antonellini and Aydin, 1994; Knott et al.,
the relationship between cross-fault damage zone width and dis-
1996; Aarland and Skjerven, 1998; Braathen and Gabrielsen, 1998;
placement.
Nelson et al., 1999; Mitra and Ismat, 2001; Clausen et al., 2003;
(4) Damage zone width is commonly asymmetric between the
Doughty, 2003; Berg and Skar, 2005; Dor et al., 2006; Mitchell et al.,
hanging-wall and the footwall, as well as on both tips along a
2011). Asymmetric damage zone widths are partly related to different
fault trace and both sides of a fault tip. Therefore, complexity
rock properties across the fault (e.g. Berg and Skar, 2005). For dip-slip
and asymmetry of damage zones should be carefully considered
faults, in general, asymmetric patterns also result from different stress
when applying different sets of data to scaling relationships of
conditions in both the hanging-wall and the footwall during faulting
fault attributes.
(Mandl, 1988, 2000; Knott et al., 1996).
Asymmetric patterns are also observed in along-fault damage zones,
and this has been reported in many previous studies (e.g. Pollard
and Segall, 1987; Kim et al., 2003; Flodin and Aydin, 2004; Kim and Acknowledgments
Sanderson, 2008; de Joussineau and Aydin, 2007; Molli et al., 2010).
Damage structures are commonly more dominant in dilational quad- We thank Antonino Cilona, Michele Fondriest, and Giulio Di Toro for
rants rather than contractional ones especially at linking and tip their constructive reviews that greatly improved the manuscript. We
zones of faults. Kim et al. (2000) suggested that this difference is relat- also thank David J. Sanderson for reviewing an early version of this
ed to the different slip nature on both sides of a fault tip associated with paper, and members of GSGR Lab. in Pukyong National University
rock behavior to stress (in general, brittle rocks are broken more easily for their help with data analysis and fruitful discussions. This work
under tension than compression at shallow depth). Asymmetric tip was supported by the National Research Foundation of Korea (NRF-
damage patterns can also be developed by dynamic rupture 2013R1A1A2010910).
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 85

References of the 1957 MW 8.1 Gobi–Altay earthquake rupture along the Bogd fault, Mongolia.
J. Geophys. Res. 117. http://dx.doi.org/10.1029/2011JB008676.
Aarland, R.K., Skjerven, J., 1998. Fault and fracture characteristics of a major fault zone in Chough, S.K., Shon, Y.K., 2010. Tectonic and sedimentary evolution of a Cretaceous conti-
the northern North Sea: analysis of 3D seismic and oriented cores in the Brage Field nental arc–backarc system in the Korean peninsula: new view. Earth Sci. Rev. 101,
(Block 31/4). Geol. Soc. Lond. Spec. Publ. 127, 209–229. 225–249.
Agosta, F., Aydin, A., 2006. Architecture and deformation mechanism of a basin-bounding Cilona, A., Aydin, A., Johnson, N.M., 2015. Permeability of a fault zone crosscutting a
normal fault in Mesozoic platform carbonates, central Italy. J. Struct. Geol. 28, sequence of sandstones and shales and its influence on hydraulic head distribution
1445–1467. in the Chatsworth Formation, California, USA. Hydrogeol. J. 23, 405–419.
Agosta, F., Kirschner, D.L., 2003. Fluid conduits in carbonate-hosted seismogenic Clausen, J.A., Gabrielsen, R.H., Johnsen, E., Korstgard, J.A., 2003. Fault architecture and clay
normal faults of central Italy. J. Geophys. Res. 108, B4. http://dx.doi.org/10.1029/ smear distribution. Examples from field studies and drained ring-shear experiments.
2002JB002013. Nor. J. Geol. 83, 131–146.
Ambraseys, N.N., Tchalenko, J.S., 1969. The Dasht-e Bayaz (Iran) earthquake of August 31, Clayton, L., 1966. Tectonic depressions along the Hope fault, a transcurrent fault in North
1968: a field report. Bull. Seismol. Soc. Am. 59, 1751–1792. Canterbury, New Zealand. N. Z. J. Geol. Geophys. 9, 95–104.
Antonellini, M., Aydin, A., 1994. Effect of faulting on fluid flow in porous sandstones: Cowie, P.A., Scholz, C.H., 1992. Physical explanation for the displacement–length relation-
petrophysical properties. Am. Assoc. Pet. Geol. Bull. 78, 355–377. ship of faults, using a post-yield fracture mechanics model. J. Struct. Geol. 14,
Aydin, A., 2000. Fractures, faults and hydrocarbon entrapment, migration and flow. Mar. 1133–1148.
Pet. Geol. 17, 797–814. Cowie, P.A., Shipton, Z.K., 1998. Fault tip displacement gradients and process zone dimen-
Aydin, A., Johnson, A.M., 1978. Development of faults as zones of deformation bands and sions. J. Struct. Geol. 20, 983–997.
as slip surfaces in sandstone. Pure Appl. Geophys. 11b, 931–942. Crider, J.G., Pollard, D.D., 1998. Fault linkage: three-dimensional mechanical interaction
Aydin, A., Schultz, R.A., 1990. Effect of mechanical interaction on the development of between echelon normal faults. J. Geophys. Res. 103 (B10), 24373–24391.
strike-slip faults with echelon patterns. J. Struct. Geol. 12, 123–129. Crowell, J.C., 1974. Origin of Late Cenozoic basins in southern California. In: Dickinson,
Bastesen, E., Braathen, A., 2010. Extensional faults in fine grained carbonates — analysis of W.R. (Ed.), Tectonics and Sedimentation. Society of Economic Paleontologists and
fault core lithology and thickness–displacement relationships. J. Struct. Geol. 32, Mineralogists, Special Publication 22, pp. 109–204.
1609–1628. Davatzes, N.C., Aydin, A., 2005. Distribution and nature of fault architecture in a layered sand-
Bastesen, E., Braathen, A., Nøttveit, H., Gabrielsen, R.H., Skar, T., 2009. Extensional fault stone and shale sequence: an example from the Moab Fault, Utah. In: Sorkhabi, R., Tsuji,
cores in micritic carbonates, a case study from Gulf of Corinth, Greece. J. Struct. Y. (Eds.), Faults, Fluid Flow, and Petroleum Traps. AAPG Memoir 85, pp. 153–180.
Geol. 31, 403–420. De Joussineau, G., Aydin, A., 2007. The evolution of damage zone with fault growth in
Beach, A., Welbon, A.I., Brockbank, P.J., McCallum, J.E., 1999. Reservoir damage around sandstone and its multiscale characteristics. J. Geophys. Res. 112, B12401. http://dx.
faults: outcrop examples from the Suez Rift. Pet. Geosci. 5, 109–116. doi.org/10.1029/2006JB004711.
Bense, V., Person, M., Chaudhary, K., You, Y., Cremer, N., Simon, S., 2008. Thermal anom- Di Toro, G., Pennacchioni, G., 2005. Fault plane processes and mesoscopic structures of a
alies indicate preferential flow along faults in unconsolidated sedimentary aquifers. strong-type seismogenic fault in tonalities (Adamello batholith, Southern Alps).
Geophys. Res. Lett. 35. http://dx.doi.org/10.1029/2008GL036017. Tectonophysics 402, 55–80.
Ben-Zion, Y., Malin, P., 1991. San Andreas fault zone head waves near Parkfield, California. Di Toro, G., Han, R., Hirose, T., De Paola, N., Nielsen, S., Mizoguchi, K., Ferri, F., Cocco, M.,
Science 251, 1592–1594. Shimamoto, T., 2005. Fault lubrication during earthquake. Nature 471, 494–498.
Berg, S.S., Skar, T., 2005. Controls on damage zone asymmetry of a normal fault zone: out- Dockrill, B., Shipton, Z.K., 2010. Structural controls on leakage from a natural CO2 geologic
crop analyses of a segment of the Moab fault, SE Utah. J. Struct. Geol. 27, 1803–1822. storage site: Central Utah, U.S.A. J. Struct. Geol. 32, 1768–1782.
Bhat, H.S., Biegel, R.L., Rosakis, A.J., Sammis, C.G., 2010. The effect of asymmetric dam- Dor, O., Ben-Zion, Y., Rockwell, T.K., Brune, J., 2006. Pulverized rocks in the Mojave section
age on dynamic shear rupture propagation II: with mismatch in bulk elasticity. of the San Andreas Fault Zone. Earth Planet. Sci. Lett. 245, 642–654.
Tectonophysics 493, 263–271. Doughty, P.T., 2003. Clay smear seals and fault sealing potential of an exhumed growth
Billi, A., Salvini, F., Storti, F., 2003. The damage zone–fault core transition in carbonate fault, Rio Grande Rift, New Mexico. Am. Assoc. Pet. Geol. Bull. 87, 427–444.
rocks: implications for fault growth, structure and permeability. J. Struct. Geol. 25, Du Bernard, X., Labaume, P., Darcel, C., Davy, P., Bour, O., 2002. Cataclastic slip band distri-
1779–1794. bution in normal fault damage zones, Nubian sandstones, Suez rift. J. Geophys. Res.
Bistacchi, A., Griffith, W.A., Smith, S.A.F., Di Toro, G., Jones, R., Nielsen, S., 2011. Fault 107 (B7), 2141. http://dx.doi.org/10.1029/2001JB000493.
roughness at seismogenic depths from LIDAR and photogrammetric analysis. Pure Evans, J.P., 1990. Thickness-displacement relationships for fault zones. J. Struct. Geol. 12,
Appl. Geophys. 168, 2345–2363. 1061–1065.
Blenkinsop, T.G., 1989. Thickness–displacement relationships for deformation zones: Evans, J.P., Bradbury, K.K., 2004. Faulting and fracturing of nonwelded Bishop tuff, eastern
discussion. J. Struct. Geol. 11, 1051–1054. California deformation mechanisms in very porous materials in the Vadose zone.
Braathen, A., Gabrielsen, R.H., 1998. Lineament Architecture and Fracture Distribution Vadose Zone J. 3, 602–623.
in Metamorphic and Sedimentary Rocks, With Application to Norway. Geological Evans, J.P., Forster, C.B., Goddard, J.V., 1997. Permeability of fault-related rocks, and impli-
Survey of Norway, Trondheim, p. 78. cations for hydraulic structure of fault zones. J. Struct. Geol. 19, 1393–1404.
Brogi, A., 2011. Bowl-shaped basin related to low-angle detachment during continental Faulkner, D.R., Lewis, A.C., Rutter, E.H., 2003. On the internal structure and mechanics of
extension: the case of the controversial Neogene Siena Basin (central Italy, Northern large strike-slip fault zones: field observations of the Carboneras fault in southeastern
Apennines). Tectonophysics 499, 131–148. Spain. Tectonophysics 367, 147–156.
Bruhn, R.L., Parry, W.T., Yonkee, W.A., Thompson, T., 1994. Fracturing and hydrothermal Faulkner, D.R., Mitchell, T.M., Healy, D., Heap, M.J., 2006. Slip on ‘weak’ faults by the rota-
alteration in normal fault zones. Pure Appl. Geophys. 142, 609–644. tion of regional stress in the fracture damage zone. Nature 444, 922–925.
Bullock, R.J., De Paola, N., Holdsworth, R.E., Trabucho-Alexandre, J., 2014. Lithological Faulkner, D.R., Mitchell, T.M., Rutter, E.H., Cembrano, J., 2008. On the structure and me-
controls on the deformation mechanisms operating within carbonate-hosted faults chanical properties of large strike-slip faults. Geol. Soc. Lond. Spec. Publ. 299, 139–150.
during the seismic cycle. J. Struct. Geol. 58, 22–42. Faulkner, D.R., Jackson, C.A.L., Lunn, R.J., Schlische, R.W., Shipton, Z.K., Wibberley, C.A.J.,
Caine, J.S., Evans, J.P., Forster, C.B., 1996. Fault zone architecture and permeability struc- Withjack, M.O., 2010. A review of recent developments concerning the structure,
ture. Geology 24, 1025–1028. mechanics and fluid flow properties of fault zones. J. Struct. Geol. 32, 1557–1575.
Cello, G., Tondi, E., Micarelli, L., Invernizzi, C., 2001. Fault zone fabrics and geofluid proper- Faulkner, D.R., Mitchell, T.M., Jensen, E., Cembrano, J., 2011. Scaling of fault damage zones
ties as indicators of rock deformation modes. J. Geodyn. 32, 543–565. with displacement and the implications for fault growth processes. J. Geophys. Res.
Chester, F.M., Logan, J.M., 1986. Composite planar fabric of gouge from the Punchbowl 116, B05403. http://dx.doi.org/10.1029/2010JB007788.
fault zone, California. J. Struct. Geol. 9, 621–634. Flodin, E., Aydin, A., 2004. Faults with asymmetric damage zones in sandstone, Valley of
Chester, F.M., Evans, J.P., Biegel, R.L., 1993. Internal structure and weakening mechanisms Fire State Park, southern Nevada. J. Struct. Geol. 26, 983–988.
of the San-Andreas fault. J. Geophys. Res. 98 (B1), 771–786. Fossen, H., 2010. Structural Geology. Cambridge University Press, Cambridge.
Chester, F.M., Chester, J.S., Kirschner, D.L., Schulz, S.E., Evans, J.P., 2004. Structure of large- Fossen, H., Hesthammer, J., 2000. Possible absence of small faults in the Gullfaks Field,
displacement, strike-slip fault zones. In: Karner, G.D., Taylor, B., Driscoll, N.W., northern North Sea: implications for downscaling of faults in some porous sandstone.
Kohlstedt, D.L. (Eds.), Rheology and Deformation in the Lithosphere at Continental J. Struct. Geol. 22, 851–863.
Margins. Columbia Univ. Press, New York. Fossen, H., Johansen, S.E., Hesthammer, J., Rotevatn, A., 2005. Fault interaction in porous
Childs, C., Watterson, J., Walsh, J.J., 1995. Fault overlap zones within developing normal sandstone and implications for reservoir management; examples from southern
fault systems. J. Geol. Soc. 152, 535–549. Utah. AAPG Bull. 89, 1593–1606.
Childs, C., Nicol, A., Walsh, J.J., Watterson, J., 1996. Growth of vertically segmented normal Fossen, H., Schultz, R.A., Shipton, Z.K., Mair, K., 2007. Deformation bands in sandstone: a
faults. J. Struct. Geol. 18, 1389–1397. review. J. Geol. Soc. 164, 1–15.
Childs, C., Walsh, J.J., Watterson, J., 1997. Complexity in fault zone structure and implica- Foxford, K.A., Walsh, J.J., Watterson, J., Garden, I.R., Guscott, S.C., Burley, S.D., 1998. Struc-
tions for fault seal prediction. In: Møller-Pedersen, P., Koestler, A.G. (Eds.), Hydrocar- ture and content of the Moab Fault Zone, Utah, USA, and its implications for fault seal
bon Seals: Importance for Exploration and Production. Norwegian Petroleum Society prediction. In: Jones, G., Fisher, Q.J., Knipe, R.J. (Eds.), Faulting, Fault Sealing, and Fluid
(NPF), Special Publication 7, pp. 61–72. Flow in Hydrocarbon Reservoirs. Geological Society of London, Special Publications
Childs, C., Manzocchi, T., Walsh, J.J., Bonson, C.G., Nicol, A., Schopfer, M.P.J., 2009. A geo- 147, pp. 87–103.
metric model of fault zone and fault rock thickness variations. J. Struct. Geol. 31, Gabrielsen, R.H., Aarland, R.-K., 1990. Characteristics of pre- and syn-consolidation struc-
117–127. tures and tectonic joints and microfaults in fine to medium-grained sandstone. In:
Choi, J.-H., Yang, S.-J., Kim, Y.-S., 2009. Fault zone classification and structural characteris- Stephansson, B. (Ed.), Rock Joints. Balkema, Rotterdam, pp. 45–50.
tics of the southern Yangsan fault in the Sangcheon-ri area, SE Korea. J. Geol. Soc. Genter, A., Castaing, C., Dezayes, C., Tenzer, H., Traineau, H., Villemin, T., 1997. Compara-
Korea 45, 9–28 (in Korean with English abstract). tive analysis of direct (core) and indirect (borehole imaging tools) collection of frac-
Choi, J.-H., Jin, K., Enkhbayar, D., Davvasambuu, B., Bayasgalan, A., Kim, Y.-S., 2012. Rup- ture data in the Hot Dry Rock Soultz reservoir (France). J. Geophys. Res. 102 (B7),
ture propagation based on damage patterns, slip distribution, and fault segmentation 15419–15431.
86 J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87

Geraud, Y., Diraison, M., Orellana, N., 2006. Fault zone geometry of a mature active normal Marrett, R., Allmendinger, R.W., 1992. Amount of extension on “small” faults: an example
fault: a potential high permeability channel (Pirgaki fault, Corinth rift, Greece). from the Viking graben. Geology 20, 47–50.
Tectonophysics 246, 61–76. Martel, S.J., Pollard, D.D., Segall, P., 1988. Development of simple strikeslip fault
Goddard, J.V., Evans, J.P., 1995. Chemical changes and fluid–rock interaction in faults of zones, Mount Abbot quadrangle, Sierra Nevada, California. Geol. Soc. Am. Bull. 100,
crystalline thrust sheets, northwestern Wyoming, U.S.A. J. Struct. Geol. 17, 533–547. 1451–1465.
Griffith, W.A., Rosakis, A., Pollard, D.D., Ko, C.W., 2009. Dynamic rupture experiments elu- McGrath, A.G., Davison, I., 1995. Damage zone geometry around fault tips. J. Struct. Geol.
cidate tensile crack development during propagating earthquake ruptures. Geology 17, 1011–1024.
37, 795–798. Micarelli, L., Moretti, I., Daniel, J.M., 2003. Structural properties of rift-related normal
Gudmundsson, A., Simmenes, T.H., Larsen, B., Philipp, S.L., 2010. Effects of internal struc- faults: the case study of the Gulf of Corinth, Greece. J. Geodyn. 36, 275–303.
ture and local stresses on fracture propagation, deflection, and arrest in fault zones. Mitchell, T.M., Faulkner, D.R., 2009. The nature and origin of off-fault damage surrounding
J. Struct. Geol. 32, 1643–1655. strike-slip fault zones with a wide range of displacements: a field study from the
Hausegger, S., Kurz, W., 2013. Cataclastic faults along the SEMP fault system (Eastern Alps, Atacama fault system, northern Chile. J. Struct. Geol. 31, 802–816.
Austria) — a contribution to fault zone evolution, internal structure and paleo- Mitchell, T.M., Ben-Zion, Y., Shimamoto, T., 2011. Pulverized fault rocks and damage
stresses. Tectonophysics 608, 237–251. asymmetry along the Arima-Takatsuki Tectonic Line, Japan. Earth Planet. Sci. Lett.
Heynekamp, M.R., Goodwin, L.B., Mozley, P.S., Haneberg, W.C., 1999. Controls on fault- 308, 284–297.
zone architecture in poorly lithified sediments, Rio Grande Rift, New Mexico: implica- Mitra, G., Ismat, Z., 2001. Microfracturing associated with reactivated fault zones and
tions for fault-zone permeability and fluid flow. In: Goodwin, L.B., Mozley, P.S., Moore, shear zones: what can it tell us about deformation history? In: Holdsworth, R.E.,
J.M., Haneberg, W.C. (Eds.), Faults and Subsurface Fluid Flow in the Shallow Crust Strachan, R.A., Magloughlin, J.F., Knipe, R.J. (Eds.), The Nature and Tectonic Signifi-
Geophysical Monograph 113. American Geophysical Union, Washington, pp. 27–49. cance of Fault Zone Weakening. Geological Society of London, Special Publications
Hull, J., 1988. Thickness–displacement relationships for deformation zones. J. Struct. Geol. 186, pp. 113–140
10, 431–435. Molli, G., Cortecci, G., Vaselli, L., Ottria, G., Cortopassi, A., Dinelli, E., Mussi, M., Barbieri, M.,
Hung, J.-H., Ma, K.-F., Wang, C.-Y., Ito, H., Lin, W., Yeh, E.-C., 2009. Subsurface structure, 2010. Fault zone structure and fluid–rock interaction of a high angle normal fault in
physical properties, fault-zone characteristics and stress state in scientific drill holes Carrara marble (NW Tuscany, Italy). J. Struct. Geol. 32, 1334–1348.
of Taiwan Chelungpu Fault Drilling Project. Tectonophysics 466, 307–321. Myers, R., Aydin, A., 2004. The evolution of faults formed by shearing across joint zones in
Johansen, S.E., Fossen, H., 2008. Internal geometry of fault damage zones in siliclastic rocks. sandstone. J. Struct. Geol. 26, 947–966.
In: Kurz, W., Wibberley, C.A.J., Imber, J., Collettini, C., Holdsworth, R.E. (Eds.), The Inter- Nelson, E.P., Kullman, A.J., Gardner, M.H., 1999. Fault-fracture networks and related fluid
nal Structure of Fault Zones: Fluid Flow and Mechanical Properties. Geological Society flow and sealing, Brushy Canyon Formation, West Texas. In: Goodwin, L.B., Mozley,
of London, Special Publications 299, pp. 35–56. P.S., Moore, J.M., Haneberg, W.C. (Eds.), Faults and Subsurface Fluid Flow in the Shal-
Jones, M.A., Knipe, R.J., 1996. Seismic attribute maps; application to structural interpreta- low Crust Geophysical Monograph 113. American Geophysical Union, Washington,
tion and fault seal analysis in the North Sea Basin. First Break 14, 449–461. pp. 69–81.
Jourde, H., Flodin, E.A., Aydin, A., Durlofsky, L.J., Wen, X.-H., 2002. Computing permeability Odling, N.E., Harris, S.D., Knipe, R.J., 2004. Permeability scaling properties of fault damage
of fault zones in eolian sandstone from outcrop measurements. Am. Assoc. Pet. Geol. zones in siliclastic rocks. J. Struct. Geol. 26, 1727–1747.
Bull. 86, 1187–1200. Olierook, H.K.H., Piane, C.D., Timms, N.E., Esteban, L., Rezaee, R., Mory, A.J., Hancock, d.L.,
Kim, Y.-S., Sanderson, D.J., 2005. The relationship between displacement and length of 2014. Facies-based rock properties characterization for CO2 sequestration: GSWA
faults: a review. Earth Sci. Rev. 68, 317–334. Harvey 1 well, Western Australia. Mar. Pet. Geol. 50, 83–102.
Kim, Y.-S., Sanderson, D.J., 2006. Structural similarity and variety at the tips in a wide Pachell, M.A., Evans, J.P., 2002. Growth, linkage, and termination processes of a 10-km-long
range of strike-slip faults: a review. Terra Nova 18, 330–344. strike-slip fault in jointed granite: the Gemini fault zone, Sierra Nevada, California.
Kim, Y.-S., Sanderson, D.J., 2008. Earthquake and fault propagation, displacement and J. Struct. Geol. 24, 1903–1924.
damage zones. In: Landowe, S.J., Hammler, G.M. (Eds.), Structural Geology: New Pakiser, L.C., 1960. Transcurrent fault and volcanism in Owens Valley. Geol. Soc. Am. Bull.
Research. Nova Science Publishers, Hauppauge, New York, pp. 99–117. 71, 153–160.
Kim, Y.-S., Sanderson, D.J., 2010. Inferred fluid flow through fault damage zones Peacock, D.C.P., 1991. Displacements and segment linkage in strike-slip fault zones.
based on the observation of stalactites in carbonate caves. J. Struct. Geol. 32, J. Struct. Geol. 13, 1025–1035.
1305–1316. Peacock, D.C.P., 2002. Propagation, interaction and linkage in normal fault systems. Earth
Kim, Y.-S., Andrews, J.R., Sanderson, D.J., 2000. Damage zones around strike-slip fault sys- Sci. Rev. 58, 121–142.
tems and strike-slip fault evolution, Carckington Haven, southwest England. Geosci. J. Peacock, D.C.P., Sanderson, D.J., 1991. Displacement, segment linkage and relay ramps in
4, 53–72. normal fault zones. J. Struct. Geol. 13, 721–733.
Kim, Y.-S., Peacock, D.C.P., Sanderson, D.J., 2003. Mesoscale strike-slip faults and damage Peacock, D.C.P., Sanderson, D.J., 1994. Geometry and development of relay ramps in nor-
zones at Marsalforn, Gozo Island, Malta. J. Struct. Geol. 25, 793–812. mal fault systems. Am. Assoc. Pet. Geol. Bull. 78, 147–165.
Kim, Y.-S., Peacock, D.C.P., Sanderson, D.J., 2004. Fault damage zones. J. Struct. Geol. 26, Peacock, D.C.P., Sanderson, D.J., 1995. Strike-slip relay ramps. J. Struct. Geol. 17, 1351–1360.
503–517. Peacock, D.C.P., Knipe, R.J., Sanserson, D.J., 2000. Glossary of normal faults. J. Struct. Geol.
Knott, S.D., 1994. Fault zone thickness versus displacement in the Permo-Triassic 22, 291–305.
sandsones of NW England. J. Geol. Soc. 151, 17–25. Pollard, D.D., Segall, P., 1987. Theoretical displacements and stresses near fractures in rock:
Knott, S.D., Beach, A., Brockbank, P.J., Brown, J.L., McCallum, J.E., Welbon, A.I., 1996. Spatial with applications to faults, joints, veins, dykes and solution surfaces. In: Atkinson, B.K.
and mechanical controls on normal fault populations. J. Struct. Geol. 18, 359–372. (Ed.), Fracture Mechanics of Rock. Academic Press, London, pp. 277–349.
Ko, K., Choi, J.-H., Kim, Y.-S., 2012. Effects of geological structures on slope stability: an ex- Reyer, D., Bauer, J.F., Philipp, S.L., 2012. Fracture systems in normal fault zones crosscut-
ample from the northwestern part of Daegu, Korea. J. Eng. Geol. 22, 1–13 (in Korean ting sedimentary rocks, Northwest German Basin. J. Struct. Geol. 45, 38–51.
with English abstract). Riley, P.R., Goodwin, L.B., Lewis, C.J., 2010. Controls on fault damage zone width, structure,
Koestler, A.G., Ehrman, W., 1991. Description of brittle extensional features in chalk and symmetry in the Bandelier Tuff, New Mexico. J. Struct. Geol. 32, 766–780.
on the crest of a salt ridge (NW Germany). In: Roberts, A., Yielding, G., Freeman, B. Rotevatn, A., Fossen, H., 2011. Simulating the effect of subseismic fault tails and process
(Eds.), The Geometry of Normal Faults. Geological Society of London, Special Publica- zones in a siliciclastic reservoir analogue: implications for aquifer support and trap
tions 56, pp. 113–123. definition. Mar. Pet. Geol. 28, 1648–1662.
Lewis, C.J., Gardner, J.N., Schultz-Fellenz, E.S., Lavine, A., Reneau, S.L., Olig, S., 2009. Fault Savage, H.M., Brodsky, E.E., 2011. Collateral damage: evolution with displacement of
interaction and along-strike variation in throw in the Pajarito fault system, Rio fracture distribution and secondary fault strands in fault damage zones. J. Geophys.
Grande rift, New Mexico. Geosphere 5, 252–269. Res. 116, B03405.
Li, H., Wang, H., Xu, Z., Si, J., Pei, J., Li, T., Huang, Y., Song, S.-R., Kuo, L.-W., Sun, Z., Chevalier, Scholz, C.H., 2002. The Mechanics of Earthquakes and Faulting. Cambridge University
M.-L., Liu, D., 2013. Characteristics of the fault-related rocks, fault zones and the prin- Press, Cambridge.
cipal slip zone in the Wenchuan Earthquake Fault Scientific Drilling Project Hole-1 Scholz, C.H., Anders, M.H., 1994. The permeability of faults: in the mechanical involve-
(WFSD-1). Tectonophysics 584, 23–42. ment of fluids in faulting. U.S. Geol. Surv. Open File Rep. 94-228, 247–253.
Lin, A., Yamashita, K., 2013. Spatial variations in damage zone width along strike-slip Scholz, C.H., Cowie, P.A., 1990. Determination of total strain from faulting using slip mea-
faults: an example from active faults in southwest Japan. J. Struct. Geol. 57, 1–15. surements. Nature 346, 837–839.
Lin, A., Ouchi, T., Chen, A., Maruyama, T., 2001. Co-seismic displacements, folding and Scholz, C.H., Dawers, N.H., Yu, J.Z., Anders, M.H., Cowie, P., 1993. Fault growth and fault
shortening structures along the Chelungpu surface rupture zone occurred during scaling laws: preliminary results. J. Geophys. Res. 98 (B12), 21951–21961. http://dx.
the 1999 Chi-Chi (Taiwan) earthquake. Tectonophysics 330, 225–244. doi.org/10.1029/93JB01008.
Lin, A., Guo, J., Fu, B., 2004. Co-seismic mole track structures produced by the 2001 Ms 8.1 Schueller, S., Braathen, A., Fossen, H., Tveranger, J., 2013. Spatial distribution of deforma-
Central Kunlun earthquake, China. J. Struct. Geol. 26, 1511–1519. tion bands in damage zones of extensional faults in porous sandstones: statistical
Long, J.J., Imber, J., 2011. Geological controls on fault relay zone scaling. J. Struct. Geol. 33, analysis of field data. J. Struct. Geol. 52, 148–162.
1790–1800. Schultz, R.A., Fossen, H., 2008. Terminology for structural discontinuities. Am. Assoc. Pet.
Lopez, D.L., Smith, L., 1995. Fluid flow in fault zones: analysis of the interplay of convec- Geol. Bull. 92, 853–867.
tive circulation and topographically driven groundwater flow. Water Resour. Res. Schulz, S.E., Evans, J.P., 1998. Spatial variability in microscopic deformation and composi-
31, 1489–1503. tion of the Punchbowl fault, southern California: implications for mechanisms, fluid–
Mair, K., Main, I., Elphick, S., 2000. Sequential growth of deformation bands in the labora- rock interaction, and fault morphology. Tectonophysics 295, 223–244.
tory. J. Struct. Geol. 22, 25–42. Schulz, S.E., Evans, J.P., 2000. Mesoscopic structure of the Punchbowl Fault, Southern
Mandl, G., 1988. Mechanics of Tectonic Faulting. Models and Basic Concepts. Elsevier, California and the geologic and geophysical structure of active strike-slip faults.
Amsterdam. J. Struct. Geol. 22, 913–930.
Mandl, G., 2000. Faulting in Brittle Rocks. An Introduction to the Mechanics of Tectonic Segall, P., Pollard, D.D., 1980. Mechanics of discontinuous faults. J. Geophys. Res. 85,
Faults. Springer, Berlin. 4337–4350.
J.-H. Choi et al. / Earth-Science Reviews 152 (2016) 70–87 87

Segall, P., Pollard, D.D., 1983. Nucleation and growth of strike slip faults in granite. Underhill, J.R., Woodcock, N.H., 1987. Faulting mechanisms in high-porosity sand-stones:
J. Geophys. Res. 88, 555–568. New Red Sandstone, Arran, Scotland. In: Jones, M.E., Preston, R.M.F. (Eds.), Deforma-
Shipton, Z.K., Cowie, P.A., 2001. Damage zone and slip-surface evolution over mm to km tion of Sediments and Sedimentary Rocks. Geological Society of London, Special
scales in high-porosity Navajo sandstone, Utah. J. Struct. Geol. 23, 1825–1844. Publications 29, pp. 91–105.
Shipton, Z.K., Cowie, P.A., 2003. A conceptual model for the origin of fault damage zone Vermilye, J.M., Scholz, C.H., 1998. The process zone: a microstructural view. J. Geophys.
structures in high-porosity sandstone. J. Struct. Geol. 25, 333–344. Res. 103, 12223–12237.
Shipton, Z.K., Evans, J.P., Robeson, K.R., Forster, C.B., Snelgrove, S., 2002. Structural hetero- Walker, R.J., Holdsworth, R.E., Imber, J., Faulkner, D.R., Armitage, P.J., 2013. Fault zone ar-
geneity and permeability in eolian sandstone: implications for subsurface modeling chitecture and fluid flow in interlayered basaltic volcaniclastic–crystalline sequences.
of faults. Am. Assoc. Pet. Geol. Bull. 86, 863–883. J. Struct. Geol. 51, 92–104.
Shipton, Z., Evans, J., Kirchner, D., Kolesar, P., Williams, A., Heath, J., 2004. Analysis of CO2 Walsh, J.J., Watterson, J., 1988. Analysis of the relationship between the displacements
leakage through low-permeability faults from natural reservoirs in the Colorado and dimensions of faults. J. Struct. Geol. 10, 239–247.
Plateau, southern Utah. In: Baines, S., Worden, R. (Eds.), Geological Storage of Carbon Walsh, J.J., Watterson, J., Yielding, G., 1991. The importance of small-scale faulting in
Dioxide. Geological Society of London, Special Publications 233, pp. 43–58. regional extension. Nature 351, 391–393.
Shipton, Z.K., Soden, A.M., Kirkpatrick, J.D., Bright, A.M., Lunn, R.J., 2006. How thick is Walsh, J.J., Watterson, J., Bailey, W.R., Childs, C., 1999. Fault relays, bends and branch-lines.
a fault? Fault displacement-thickness scaling revisited. In: Abercrombie, R. (Ed.), J. Struct. Geol. 21, 1019–1026.
Earthquake: Radiated Energy and the Physics of Faulting, pp. 193–198. Walsh, J.J., Bailey, W.R., Childs, C., Nicol, A., Bonson, C.G., 2003. Formation of segmented
Sibson, R.H., 1977. Fault rocks and fault mechanisms. J. Geol. Soc. 133, 191–213. normal faults: a 3-D perspective. J. Struct. Geol. 25, 1251–1262.
Sibson, R.H., 1989. Earthquake faulting as a structural process. J. Struct. Geol. 11, 1–14. White, S.H., Bretan, P.G., Rutter, E.H., 1986. Fault-zone reactivation: kinematics and mech-
Smith, L., Porster, C.B., Evans, J.P., 1990. Interaction of fault zones, fluid flow, and heat anisms. Philos. Trans. R. Soc. Lond. Ser. A 317, 81–97.
transfer at the basin scale: in hydrogeology of permeability environments. Int. Wibberley, C.A.J., Yielding, G., Di Toro, G., 2008. Recent advances in the under-standing of
Assoc. Hydrogeol. 2, 41–67. fault zone internal structure: a review. In: Wibberley, C.A.J., Kurz, W., Imber, J.,
Smith, S.A.F., Bistacchi, A., Mitchell, T.M., Mittempergher, S., Di Toro, G., 2013. The structure Holdsworth, R.E., Collettini, C. (Eds.), The Internal Structure of Fault Zones: Implica-
of an exhumed intraplate seismogenic fault in crystalline basement. Tectonophysics tions for Mechanical and Fluid-flow Properties. Geological Society of London, Special
599, 29–44. Publication 299, pp. 5–33.
Soliva, R., Schultz, R.A., 2008. Distributed and localized faulting in extensional settings: Wilson, J.E., Chester, J.S., Chester, F.M., 2003. Microfracture analysis of fault growth and
insights from the North Ethiopian Rift–Afar transition area. Tectonics 27, TC2003. wear processes, Punchbowl fault, San Andreas system, California. J. Struct. Geol. 25,
Storti, F., Holdsworth, R.E., Salvini, F., 2003. Intraplate strike-slip deformation belts. In: 1855–1873.
Storti, F., Holdsworth, R.E., Salvini, F. (Eds.), Intraplate Strike-slip Deformation Belts. Zhang, X., Sanderson, D.J., 1996. Numerical modelling of the effects of fault slip on fluid
Geological Society of London, Special Publication 210, pp. 1–14. flow around extensional faults. J. Struct. Geol. 18, 109–119.
Suppe, J., 1985. Principles of Structural Geology. Prentice-Hall, Englewood Cliffs, New Zhang, Y., Schaubs, P.M., Zhao, C., Ord, A., Hobbs, B.E., Barnicoat, A.C., 2008. Fault related
Jersey (280 pp.). dilation, permeability enhancement, fluid flow and mineral precipitation patterns:
Tondi, E., Cilona, A., Agosta, F., Aydin, A., Rustichelli, A., Renda, P., Giunta, G., 2012. Growth numerical models. In: Wibberley, C.A.J., Kurz, W., Imber, J., Holdsworth, R.E.,
processes, dimensional parameters and scaling relationships of two conjugate sets of Collettini, C. (Eds.), The Internal Structure of Fault Zones: Implications for Mechanical
compactive shear bands in porous carbonate grainstones, Favignana Island, Italy. and Fluid-flow Properties. Geological Society of London, Special Publication 299,
J. Struct. Geol. 37, 53–64. pp. 225–239.
Torabi, A., Berg, S.S., 2011. Scaling of fault attributes: a review. Mar. Pet. Geol. 28, 1444–1460.
Townend, J., Zoback, M.D., 2000. How faulting keeps the crust strong. Geology 28,
399–402.

View publication stats

You might also like