You are on page 1of 29

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

net/publication/308165914

Glossary of fault and other fracture networks

Article  in  Journal of Structural Geology · September 2016


DOI: 10.1016/j.jsg.2016.09.008

CITATIONS READS

132 8,560

5 authors, including:

David Peacock Casey W. Nixon

92 PUBLICATIONS   5,443 CITATIONS   
University of Bergen
55 PUBLICATIONS   1,044 CITATIONS   
SEE PROFILE
SEE PROFILE

Atle Rotevatn David J. Sanderson


University of Bergen University of Southampton
142 PUBLICATIONS   2,902 CITATIONS    201 PUBLICATIONS   9,392 CITATIONS   

SEE PROFILE SEE PROFILE

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

Tectono-Stratigraphic evolution of the syn-rift system on and around the Frøya High, Mid-Norwegian Continental Margin (Sub-Project of Syn-Rift System project) View
project

IODP Expedition 381: Corinth Active Rift Development View project

All content following this page was uploaded by David Peacock on 30 July 2018.

The user has requested enhancement of the downloaded file.


Glossary of fault and other fracture networks

D.C.P. Peacock1, C.W. Nixon1, A. Rotevatn1 D.J. Sanderson2 & L.F. Zuluaga1
1
Department of Earth Science, University of Bergen, Allégaten 41, 5007 Bergen, Norway
2
Engineering and the Environment, University of Southampton, Highfield, Southampton, SO17 1BJ, UK

Abstract
Increased interest in the two- and three-dimensional geometries and development of faults and other types of fractures in
rock has led to an increasingly bewildering terminology. Here we give definitions for the geometric, topological, kinematic and
mechanical relationships between geological faults and other types of fractures, focussing on how they relate to form networks.

Introduction
The large amount of work on faults, joints, veins and other fractures over the last few decades has led to the introduction
of many new terms and some confusing terminology. Even the simple term fracture has itself become confusing and misused,
tending to be used so generally that geological significance can disappear, as discussed by Manda and Horsman (2015). Referring
to a fracture in rock can be like saying a person has a pet mammal – strictly true, but so vague as to be meaningless. Is the
fracture a fault, joint, vein or dyke? Is the mammal a cat, capybara, honey badger or marmoset?
In this glossary, we focus on the relationships between different fractures and how they interact with each other and form
fracture networks (Fig. 1). We give terms related to: a) different fracture types, b) relationships between fractures, c) fracture
geometries, kinematics and mechanics. We attempt to show how the different terms are related to each other and to highlight key
differences between terms. Although we suggest that the term fracture is too vague for geological analysis, we use fracture to
denote where a term is applicable to a range of fracture types, e.g., faults, joints, veins. We give “fault or other type of fracture” in
the definitions to avoid misunderstanding. This is to highlight similarities between different structures. In these cases, appropriate
references are provided for each fracture type. For some terms, however, use of fracture is appropriate, e.g., conductive fracture.
We also focus on structural discontinuities that are not fractures per definition (e.g., stylolites and deformation bands), but that are
included for completeness. Where usage has changed through time, we provide examples of current usage.
The glossary is intended to be useful for structural geologists and other scientists working on fault or other type of
fracture networks. The terms in this glossary cover three main categories (Table 1):
a) Terms used to describe individual fractures and their associated structures. A review of fracture terminology is
provided by Schultz and Fossen (2008), whilst similar glossaries for strike-slip faults are given by Biddle and Christie-
Blick (1985), for thrusts by Butler (1982) and McClay (1992), and for normal faults by Peacock et al. (2000). We do not
cover all words relating to the faults and other types of fracture, both for brevity and to avoid duplicating existing
glossaries, with general terms being available in standard textbooks (e.g., Ramsay and Huber, 1987; Price and Cosgrove,
1990; Twiss and Moores, 1992; Fossen, 2010), as are general terms used in structural geology (such as strain and stress).
Instead, this glossary focusses on terms relating to networks of faults and other types of fracture. We have attempted to
find the first usage, or the earliest reasonable usage, of terms.
b) Terms used to describe the associations and interactions between two (or more) fractures. Here we seek to establish a
consistent and rational set of terms to describe how two or more faults or other types of fractures interact geometrically,
kinematically or mechanically, covering a range of structures from relays, through approaching and abutting fractures, to
cross-cutting fractures (Fig. 2). In this section we attempt to rationalise geometric terms dealing with fractures in general
and the kinematic interactions of faults (Fig 3). We use the term interaction to describe any relationship whereby the
development of one fault or other type of fracture affects others. Two adjacent faults or other types of fractures are said to
intersect if their surfaces are connected (geometric linkage), whereas kinematic linkage occurs where displacement is
distributed between the faults without physical intersection, as in relay ramps (Peacock and Sanderson 1991). We
generalise the concept of coherency, introduced for fault arrays by Walsh and Watterson (1991), to describe systematic
interactions (either geometric or kinematic) in networks of faults.
c) Terms used to describe networks and populations of fractures. Populations of fractures can be described in terms of
their frequency, intensity, spacing, and size (length), but networks can also be described by their topology (Sanderson and
Nixon 2015) and through associated graph theory, and the application of percolation concepts. Fig. 4 summarises the node
and branch model used by Sanderson and Nixon (2015) to characterise fracture networks with increasing interaction as
they pass through the percolation threshold to form a connected network.

We attempt to divide terms into those that are geological, geometric, topological, kinematic and/or mechanical (Fig. 1
and Table 1). This is to show that different terms have particular uses in each step of an analysis.

Peacock et al., Glossary of fault and fault and other fracture networks Page 1
Geological: here we use the adjective geological for nouns that describe features observed in rock that do not necessarily have
direct geometric, kinematic or mechanical significance. Examples include fault, fracture, joint, vein, etc.
Geometric: Geometry is defined by OxfordDictionaries.com (2015,
http://www.oxforddictionaries.com/definition/english/geometry) as “the branch of mathematics concerned with the
properties and relations of points, lines, surfaces, solids, and higher dimensional analogues”. In the context of faults or
other types of fractures, geometry refers to the size, shape and orientation of individual faults or other types of fractures
and to the patterns formed by different faults or other types of fractures. Here we use the adjective geometric to describe
the shapes or patterns of geological features.
Topological: Topology describes the arrangement of and geometric relationships between spatial objects such as faults and other
fractures. Topological characterisation generally involves using components, such as nodes and branches, and
dimensionless parameters that are invariant to scale, strain and continuous transformations (Jing and Stephansson, 1997;
Sanderson and Nixon, 2015). We use the adjective topological to help describe the geometric arrangements of faults and
other fractures. Whilst topological terms are not currently in common usage within structural geology, we include them
because they represent useful ways for describing the characteristics of networks.
Kinematic: kinematics is defined by OxfordDictionaries.com (2015,
http://www.oxforddictionaries.com/definition/english/kinematics) as “the branch of mechanics concerned with the motion
of objects without reference to the forces which cause the motion”. Here we use the adjective kinematic to refer to terms
that relate to displacement, and strains associated with the development of geological features.
Mechanical: here we use the adjective mechanical to refer to terms that describe the processes by which geological features have
formed or to the mechanical properties of the rock mass produced by fractures.

Glossary
An alphabetic list of the terms used, along with other relevant terms and approximate synonyms, are given here. Our
preferred terms are shown in Fig. 1. Each of these terms are geological structures that can form networks, or are related to the
geometries, topologies, kinematics or mechanics of fault and other fracture networks in rock.

A
Abutting fault or other type of fracture [geometric]: a fault or other type of fracture that meets another at an intersection line or
point (Fig 2a). This is a geometric term and does not signify a developmental relationship between the faults or other types
of fractures, e.g., they could have originally been two different faults or other types of fractures that met or intersected, or
one may have splayed off the other. Abutting relationships have been described for faults (Nixon et al., 2014) and joints
(Rives et al., 1994).
Alteration zone [geological]: an area of mineralisation, commonly associated with hydrothermal fluids, faults and economic
deposits (e.g., Lee and Wilkinson, 2002; Sutherland et al., 2012). The term chemical alteration zone is used by
Thienprasert et al. (1991) and by Kristensen et al. (accepted).
Anastomosing [geometric]: based on the definition of anastomosis by OxfordDictionaries.com (2015,
http://www.oxforddictionaries.com/definition/english/anastomosis), anastomosing would be the geometry whereby
adjacent channels, tubes, fibres, or other parts of a network are cross-connected. Anastomosing geometries have been
described for faults (Rowe et al., 2013), joints (Singh, 1992), veins (Philipp, 2008) and dykes (Valentine and van Wyk de
Vries, 2014).
Anticrack [geological and mechanical]: see stylolite.
Antithetic fault [geometric and kinematic]: originally defined as a minor fault that dips in the opposite direction to the dip
direction of the beds they displace (Cloos, 1928; Hills, 1940, fig. 41). Antithetic fault is now commonly used for a fault that
has the opposite shear sense to a related dominant fault or fault set (e.g., Gibbs, 1984). See synthetic fault and conjugate.
Aperture [geometric]: the width of a fracture measured perpendicular to the walls, i.e., the amount of extension across the
fracture. Synonymous with thickness of veins (Vermilye and Scholz, 1995) and dykes (e.g., Jolly and Sanderson, 1995).
Joint apertures are discussed by Hooker et al. (2009).
Approaching fault or other type of fracture [geometric]: where a fault or other type of fracture has propagated towards, and
interacts with, another fault or other type of fracture, but the two faults or other types of fractures do not intersect (Fig. 2a)
(Peacock et al., in review). This term is valid for both parallel faults and faults at angles to each other. Examples of
approaching faults are shown by Flodin and Aydin (2004, fig. 1), approaching (and connected) joints are shown by
Noroozi et al. (2015, fig. 8), and approaching veins are shown by Virgo et al. (2014, fig. 10).
Approaching damage zone [kinematic and mechanical]: defined by Peacock et al. (in review) as the area of deformation related to
interaction between two or more faults that do not intersect (Fig. 5). It is a type of interaction damage zone and is intended
to be a more general term than the linking damage zone of Kim et al. (2004). Also see damage zone.
Architecture [geological]: defined by OxfordDictionaries.com (2015,
http://www.oxforddictionaries.com/definition/english/architecture) as “the art or practice of designing and constructing
Peacock et al., Glossary of fault and fault and other fracture networks Page 2
buildings... The style in which a building is designed and constructed, especially with regard to a specific period, place, or
culture... The complex or carefully designed structure of something... The conceptual structure and logical organisation of a
computer or computer-based system”. The term is now used in sedimentology to denote the geometry of sedimentary rocks
and facies (e.g., Marzo et al., 1988). The term is being used increasingly in structural geology, apparently to denote the
geometries of structures (e.g., O'dea and Lister, 1995).
Arrangement, network [topological]: the spatial layout of lines or planes in a network, that may or may not be connected to each
other. Topological description considers the network as a system of topological components (e.g., nodes and branches),
e.g., which indicate the relationships and connectivity of faults and other fractures within a network (e.g., Jing and
Stephansson, 1997; Adler and Thovert, 1999). Also see topology, node and branch.
Array, fault or other type of fracture [geometric and kinematic]: a set of stepping or en echelon fault or other type of fracture
segments or fault or other type of fracture zones. Arrays are common geometries for faults (e.g., Segall and Pollard, 1980),
joints (e.g., La Pointe and Hudson, 1985), veins (e.g., Rothery, 1988) and dykes (e.g., Brown et al., 2007, fig. 1). Vein
arrays are synonymous with tension gashes.

B
Backbone [topological]: the connected lines or planes within a network, i.e., that are composed entirely of doubly-connected
branches or C-C branches (Fig. 4), e.g., a system of connected faults or other fractures. The backbone has the potential to
allow flow or transport across a spanning cluster and thus is commonly used as a qualitative and quantitative descriptor in
percolation theory (e.g., Stanley, 1977; Sahimi, 1994; Aizenman, 1997; Manzocchi, 2002; Hunt et al., 2014). Also see
doubly-connected branch, cluster, spanning cluster, percolation theory.
Bed-bound fault or other type of fracture [geometric and mechanical]: see layer-bound fault or other type of fracture. Also see
mechanical stratigraphy.
Bimodal [geometric]: fault or other type of fracture pattern forming two sets of planes with dip directions or strikes that exhibit a
two-fold symmetry (e.g., Woodcock and Underhill, 1987; Healy et al., 2015). This pattern is commonly associated with
faults and other fractures that formed under a plane strain (Anderson, 1951; Reches, 1978). Synonymous with conjugate.
Block [geometric]: a body of rock bounded entirely, or in plane-view, by faults (e.g., Versey, 1927) or other fractures (e.g.,
Rawnsley et al., 1998). Also see horse.
Branch [topological]: a line that is bound by nodes at each end (Fig. 4), e.g., a fault trace. There are three types of branch:
isolated branches (I-I), singly-connected branches (I-C), and doubly-connected branches (C-C) (e.g., Sanderson and
Nixon, 2015).
Branch [geometric]: verb for where one or more faults or other types of fractures extend off a larger fault or other type of
fracture, typically used for faults (e.g., Nevin, 1931; Butler, 1982).
Branch line [geometric]: the line along which two fault planes connect in when viewed in 3D (Butler, 1982). It appears not to be
commonly used for other types of geological fractures. Whilst we use intersection line to keep our terminology consistent
and applicable to other fracture types, we do not suggest that branch line be replaced as a term for faults.
Branch point [geometric]: the point at which two fault traces branch or splay when viewed in 2D. Thrust branch points are
described by McClay (1992). To keep our terminology consistent and applicable to other fracture types, we use
intersection point. Synonymous with connecting node.
Breccia [geological]: defined by OxfordDictionaries.com (2015, http://www.oxforddictionaries.com/definition/english/breccia) as
a “rock consisting of angular fragments of stones cemented by finer calcareous material”. Other minerals can form
cements, including quartz and gypsum. Non-sedimentary breccias consist of vein networks, including hydrothermal
breccias (e.g., Jébrak, 1997) and fault breccias (e.g., Woodcock and Mort, 2008).
Bridge, fault or other type of fracture [geometric and kinematic]: block of rotated rock between stepping and interacting extension
fractures (e.g., Kemeny, 2005). Bridges have been described for joints (e.g., Yan et al., 2007), veins (e.g., Laing, 2004) and
dykes (e.g., Platten, 2000). Also see broken bridge.
Brittle fault or other type of fracture [mechanical]: defined by OxfordDictionaries.com (2015,
http://www.oxforddictionaries.com/definition/english/brittle-fracture) as a “fracture of a metal or other material occurring
without appreciable prior plastic deformation”. Brittle structures and the geometries and mechanics of brittle deformation
are described in detail in various structural geology textbooks (e.g., Ramsay, 1967; Ramsay and Huber, 1987). Note that
“semi-brittle faults” also occur (e.g. Chester, 1989).
Broken bridge [geometric and kinematic]: a bridge across which two extension fractures have geometrically linked (e.g.,
Schofield et al., 2012).

C
Cataclasis [geological and mechanical]: process involving brittle fragmentation of minerals or grains, with rotation of grain
fragments, grain boundary sliding, grain flaking and fracturing, grain size reduction, and commonly volume change
(Sibson, 1977). Cataclasis typically occurs along fault planes, within fault zones, or within deformation bands.
Peacock et al., Glossary of fault and fault and other fracture networks Page 3
Clastic dyke [geological]: used for fractures filled by sedimentary rocks (e.g., Jolly et al., 1998). Some clastic dykes infill from
the surface (termed neptunian dyke), whilst others form by injection of fluidised sediment fed from below(e.g., Kenkmann,
2003). Also known as injectite or sedimentary dyke.
Closed fracture [geological]: used to describe fractures in the sub-surface which have zero porosity (e.g., Wu and Pollard, 2002),
e.g., because they are sealed by minerals. See non-conductive fracture.
Cluster [topological]: a group of interconnected lines or planes within a network, usually made up of a backbone (doubly-
connected branches) and dangling ends (singly-connected branches), e.g., a system of link faults. These can be either an
isolated cluster or a spanning cluster (Stanley, 1977; Aizenman, 1997; Odling et al., 1999). Cluster has also been applied
geologically to a group of faults or other type of fractures that produce a certain spatial distribution or arrangement, e.g.,
that are closely-spaced or have the same orientation (Bour and Davy, 1999; Manzocchi, 2002). Also see isolated cluster
and spanning cluster.
Coherence, geometric [geometric]: the existence of regular and systematic displacement patterns in arrays or networks of faults.
The concept was originally applied to zones of normal faults with splays and relays (Walsh and Watterson, 1991, fig. 4),
but can be extended to domains within fault networks, including cases where different types of faults combine to produce
systematic patterns (Fig. 3).
Coherence, kinematic [kinematic]: the existence of displacement distributions that are arranged such that geometric coherence is
maintained (Walsh and Watterson, 1991).
Compartment [geometric]: a block in which the fluids are sealed from adjacent blocks by faults (e.g., Knipe, 1997), commonly
identified adjacent compartments having distinct fluid pressures or chemistries. Also see block, horse and lens.
Compartmentalisation [geometric]: the geometry whereby a rock mass is divided into compartments, or the process by which
compartments develop (e.g., Clark et al., 1996).
Conductive fracture [geological]: a fracture within which fluids flow (Olsson et al., 1992), or that is identified by electrical
conductivity on borehole image logs (e.g., Samantray et al., 2010). Note that some open fractures may be electrically non-
conductive (e.g., Boutt et al., 2010). See open fault or other type of fracture. Antonym of non-conductive fracture.
Conjugate [geometric and kinematic]: for faults, conjugate refers to the relationship between two intersecting sets of faults (or
joints) that each formed under the same stress field (Daubrée, 1878, cited by Dennis, 1967). Two conjugate faults have the
opposite shear sense and the same angle (generally ~ 30) to the maximum principal stress direction (Anderson, 1951).
Interacting conjugate faults may move alternately (Freund, 1974; Horsfield, 1980), or may move synchronously to form an
area of high strain near the intersection (Peacock, 1991; Odonne and Massonnat, 1992). Pairs of conjugate faults can
consist of a larger fault and a smaller antithetic fault. Also see antithetic fault and synthetic fault.
Connecting fault or other type of fracture [geometric and kinematic]: connects two stepping faults or other types of fractures (e.g.,
Peacock and Sanderson, 1994). Synonymous with connecting splay (Ramsay and Huber, 1987, fig. 23.50b); for faults, the
term breaching fault is synonymous (Miller, 1999; Peacock and Parfitt, 2002). Connecting geometries have been described
for faults (e.g., Cruikshank et al., 1991) and dykes (e.g., Žák et al., 2012).
Connecting node [topological]: the point of connection between two lines, i.e., X-nodes and Y-nodes (Manzocchi, 2002; Nixon et
al., 2012; Sanderson and Nixon, 2015). Synomynous with intersection point. See also Nodes.
Connectivity [geometric and topological]: the degree to which faults or other types of fractures are connected within a network,
which depends on the size, frequency, orientation, spatial correlation, scaling and topology of the (Berkowitz et al., 2000;
Manzocchi, 2002). Connectivity has been described for a range of structures, including faults (e.g., Cherpeau et al., 2011),
veins (e.g., Virgo et al., 2014), joints (e.g., Alghalandis et al., 2015), stylolites (Ben-Itzak et al., 2014) and deformation
bands (Sternlof et al., 2004). Also see critical connectivity.
Corridor, fracture [geometric]: tabular zone of significantly-increased fracture intensity (e.g., Questiaux et al., 2010; Gabrielsen
and Braathen, 2014). Also see swarm, fault or other type of fracture.
Coupling [kinematic]: defined by OxfordDictionaries.com (http://www.oxforddictionaries.com/definition/english/coupling) as
“the pairing of two items”.
Coupling, geometric [geometric]: where one fault or other type of fracture intersects or crosses another fault or other type of
fracture, i.e., where two faults or other types of fractures are physically connected -synonymous with geometrical linkage
(cf. Chambon and Rudnicki, 2001).
Coupling, kinematic [kinematic]: where the displacements, stresses and strain of one fault or other type of fracture influences
those of another fault or other type of fracture. Synonymous with kinematic linkage.
Crack [geological]: defined by the Oxford English Dictionary (1989) as being “a fissure or opening formed by the cracking,
breaking, or bursting of a hard substance”, with first usage attributed to Palsgrave (1530). The term appears to be used
synonymously with fault or other type of fracture (e.g., Zhao et al., 2007), particularly in the engineering literature. The
term crack has no particular mechanical implications, so includes any type of fracture.
Critical connectivity [topological]: the average number of intersections per fracture needed to reach the percolation threshold, at
which a spanning cluster first forms within a network (e.g., Robinson, 1983, 1984; Balberg and Binenbaum, 1983;
Manzocchi, 2002). Also see percolation threshold and spanning cluster.
Peacock et al., Glossary of fault and fault and other fracture networks Page 4
Crossing fault or other type of fracture [geometric]: a fault or other type of fracture that cuts and displaces another fault (e.g.,
Chen, 2013) or fracture (Fig. 2a).

D
Damage zone [kinematic and mechanical]: location at which a major change in fabric (e.g., caused by pressure solution) is readily
visible (Chester and Logan, 1986; Wu and Groshong, 1991), or the volume of deformed wall rocks around a fault surface
that results from the initiation, propagation, interaction and build-up of slip along faults (Cowie and Scholz, 1992; McGrath
and Davison, 1995; Kim et al., 2004; Choi et al., in press). Synonymous with deformation front and texture zone (Wu and
Groshong, 1991).
Dangling end [topological]: a line or cluster of lines that is connected to the rest of the cluster only at one end, the other end being
a tip or I-node (Fig. 4) (Stanley, 1977; Hunt et al., 2014). Thus these are unable to carry flow through the network unlike
the backbone (Sahimi, 1994; Hunt et al., 2014). Also see singly-connected branch, backbone, cluster, isolated cluster and
spanning cluster.
Deformation band [geological]: a mm-scale tabular zone of localised but non-discrete strain (Friedman and Logan, 1973;
Engelder, 1974), typically involving grain reorganisation (grain sliding and rolling), cataclasis (grain flaking, fracturing
and crushing) and/or dissolution or precipitation (e.g., Aydin, 1978, Knott, 1994). They typically occur in porous granular
rock types. Note that there are wide range of terms related to deformation bands, including compactional shear band
(Fossen et al., 2007), compactive shear band (Aydin et al., 2006), cataclastic deformation band (Wibberley et al., 2000),
cataclastic slip band (Fowles and Burley, 1994), compaction band (Mollema and Antonellini, 1996; Rotevatn et al., 2016),
dilatant shear band (Aydin et al., 2006), dilation band (Du Bernard et al., 2002), disaggregation band (Fossen et al.,
2003), disaggregation zone (Knipe, 1986), framework-phyllosilicate structures/band (involving 10-40% phyllosilicate
minerals; Fisher and Knipe, 1998), granulation seam (e.g., Pittman, 1981), isochoric shear band (Aydin et al., 2006),
Lüder’s bands (Friedman and Logan, 1973), phyllosilicate/clay smear (involving >40% phyllosilicate minerals; Fisher and
Knipe, 2001), shear band (Menendéz et al., 1996), shear-enhanced compaction band (Eichhubl et al., 2010), slipped
deformation band (Rotevatn et al., 2008), solution band (Gibson, 1998), solution and cementation band (Fossen et al.,
2007).
Density, fault or other type of fracture [geometric]: see intensity, fault or other type of fracture.
Desiccation crack [geological]: an opening mode fracture formed by contraction of soft sediment, usually clay or mud, as it dries
(e.g., Lister and Secrest, 1985). Also called mud crack.
Diatreme [geological]: pipe-shaped plug of breccia and igneous rock produced by explosive intrusion (e.g., White and Ross,
2011).
Dimensionless intensity [geometric and topological]: a dimensionless measure derived from the product of the fault or other type
of fracture intensity and the average line or branch length (Robinson, 1983; Balberg and Binenbaum, 1983; Manzocchi,
2002; Sanderson and Nixon 2015). It is scale invariant and provides a useful indication of connectivity (Sanderson and
Nixon, in review), with critical values that indicate the percolation threshold, at which a spanning cluster first forms within
a network. Also see percolation threshold and spanning cluster.
Dip-linkage [geometric and kinematic]: the along-dip, or vertical, linkage of two faults or other types of fractures that were
initially geometrically uncoupled (Mansfield and Cartwright, 1996).
Dip-slip [kinematic]: the component of displacement parallel to the dip of the fault plane (cf. strike-slip) (Reid et al., 1913).
Normal, reverse and thrust faults are all dip-slip faults.
Discontinuity [geological and mechanical]: a surface across which there is a variation in the mechanical behaviour of rocks, such
as bedding planes or fractures (e.g., Fookes and Parrish, 1969).
Domain [geological, geometric and kinematic]: in this context, a volume, area or region of rock with distinct geometries,
kinematics or mechanics of geological structures (e.g., Schlische and Withjack, 2009; Nixon et al., 2011; Zu et al., 2013).
Domino faults [geometric and kinematic]: a system of planar rotating normal faults, in which the dip of a particular stratigraphic
datum is approximately constant (e.g., Twiss and Moores, 1992). Synonymous with bookshelf faults (e.g., Ramsay and
Huber, 1987, fig. 23.19) and with rotational planar normal faults of Wernicke and Burchfiel (1982).
Doubly-connected branch [topological]: a line with connecting nodes at both ends, thus often referred to as a C-C branch
(Sanderson and Nixon, 2015; Morley and Nixon, 2016), e.g. a fault trace that connects with other faults at both ends. Also
see topology, isolated branch, singly-connected branch, connecting node. Synonymous with backbone. Also see lens.
Duplex [geological and geometric]: an arrangement of faults with the same shear sense that are geometrically- and kinematically-
linked to form a series of lensoidal or sigmoidal horses, i.e., each pair of faults has two intersection points in map- or
section-view. Duplexes have been described for thrust (e.g., Butler, 1982), normal (e.g., Gibbs, 1984) and strike-slip
(Woodcock and Fischer, 1986) faults.
Dyke (or dike) [geological]: defined by the Oxford English Dictionary (1989) as “a mass of mineral matter, usually igneous rock,
filling up a fissure in the original strata”, with first use accredited to Playfair (1802). The term is now most commonly used
for a fault or other type of fracture at a high angle to bedding that is filled by intrusive igneous rock (e.g., Paquet et al.,
Peacock et al., Glossary of fault and fault and other fracture networks Page 5
2007), not necessarily following a pre-existing fault or other type of fracture (e.g., Delaney and Pollard, 1981). Dykes and
sills open because the internal pressure of the magma exceeds the least compressive stress in the surrounding rocks (e.g.,
Pollard, 1987), with the least compressive stress being sub-horizontal for a sub-vertical dyke to form. The fault or other
type of fracture utilised by the magma may be a pre-existing feature or a new fault or other type of fracture created by the
overpressured magma (Pollard, 1987). Dyke networks are described by Jolly and Sanderson (1995).

E
En echelon [geometric]: defined by Biddle and Christie-Blick (1985) as “a stepped arrangement of relatively short consistently
overlapping or underlapping structural elements such as faults or folds that are approximately parallel to each other but
oblique to the linear or relatively narrow zone in which they occur”. En echelon patterns have been described for faults
(e.g., Hempton and Neher, 1986), joints (e.g., Bahat, 1986), veins (e.g., Beach, 1975), dykes (e.g., Weinberger et al., 2000)
etc.
Extension(al) fault [kinematic]: see normal fault.
Extension(al) fracture [kinematic]: a fracture that has opened up at a high angle to the fracture plane, with the fracture usually
inferred to be perpendicular to the least, and parallel to the maximum and intermediate, compressive stress axes (e.g.,
Smith, 1997). Extension fractures can therefore include dykes, sills, joints and veins. Extension fractures usually form
when the fluid pressure exceeds the compressive stress perpendicular to the fracture plane (e.g., Jaeger, 1969). Also known
as tension fracture (Bucher, 1920; Gudmundson, 1995). Also see mode I. See Pollard and Aydin (1988, p. 1186).

F
Fault linkage [kinematic and mechanical]: see linkage.
Fault [geological, kinematic and mechanical]: defined by the Oxford English Dictionary (1989) as being “a dislocation or break
in continuity of the strata or vein”, with first usage being attributed to Bakewell (1813). Price (1966) defines a fault as “a
plane of fracture which exhibits obvious signs of differential movement of the rock mass on either side of the plane. Faults
are therefore planes of shear failure”, i.e., they are planes along which there has been movement parallel to the plane. More
recent work has shown that faults are usually zones, having a volume that includes interacting and linked segments,
breccias and fault rocks (e.g., Davatzes and Aydin, 2003; Childs et al. 2009). See Zone, fault.
Fault zone [geometric]: see zone, fault.
Finite cluster [topological]: see isolated cluster.
Fissure [geological]: defined by the Oxford English Dictionary (1989) as “a cleft or opening (usually rather long and narrow)
made by splitting, cleaving, or separation of parts”. With reference to geology, the Oxford English Dictionary define
fissure-vein as “a fissure in the earth's crust filled with mineral”, with Geikie (1882) being cited. Fissure is used in
volcanology for an open fracture developed at the Earth's surface above a dyke (e.g., Pollard et al., 1983), for fault-related
open fractures developed at the Earth’s surface in basalt (e.g., Martel and Langley, 2006), and to describe open fractures in
volcanic rocks at the Earth’s surface (e.g., Paquet et al., 2007). We suggest that the term be limited to open fractures of
unknown origin that reach the Earth’s surface.
Flower structure [geometric]: a system of faults that splay upward within a strike-slip fault zone (Harding and Lowell, 1979).
Negative flower structures involve steepening upwards extensional splay faults, whereas positive flower structures involve
shallowing upwards contractional splay faults (Harding, 1983, 1985).
Fracture [geological and mechanical]: defined by the Oxford English Dictionary (1989) as “the action of breaking or fact of
being broken… The result of breaking; a crack, division, split; a broken part, a splinter… The characteristic appearance of
the fresh surface in a mineral, when broken irregularly by the blow of a hammer”. The first usage for the fresh surface of a
mineral is attributed by the Oxford English Dictionary (1989) to Sullivan (1794). Pollard and Aydin (1988) suggest
fracture should be used as a general term for structure without preserved evidence for the mode of fracturing, i.e., it is
applicable to structures formed by extension or shear. Geological fractures can include such approximately planar
discontinuities as dykes, faults, joints and veins, and the term appears to be synonymous with discontinuity (e.g., Priest and
Hudson, 1976, 1981). Pollard and Aydin (1988) state that, lacking information about the mode of displacement, “geologists
should avoid speculative interpretations and refer to the structures simply as fractures”. We agree, but suggest that fracture
is only used when the origin or style of fracture is unknown.
Frequency, fault or other type of fracture [geometric]: see intensity, fault or other type of fracture.

G
Geometric coherence [geometric]: the existence of regular and systematic displacement patterns in a family of faults (Walsh and
Watterson, 1991, fig. 4).
Geometric coupling [geometric]: where two fracture planes share an intersection line (or branch line), i.e., the two are physically
connected (e.g., Jackson and Rotevatn, 2013). Also see hard-linkage.

Peacock et al., Glossary of fault and fault and other fracture networks Page 6
Geometric linkage [geometric]: where two faults or other types of fractures are connected at an intersection line. The faults or
other types of fractures need not have kinematic linkage, e.g., if one fault or other type of fracture post-dates and cuts the
other without effects on the geometry or displacement pattern.
Granulation seam [geological]: see deformation band.

H
Hard-linkage [geometric and kinematic]: the geometry or process whereby two faults are connected by one or more (usually
smaller) faults which are visible at the scale of observation (Walsh and Watterson, 1991). See connecting faults, hard-
linkage and soft-linkage.
Horse [geometric]: a fault-bound body of rock (e.g., Butler, 1982). Also see block.
Horsetail [geometric and kinematic]: Biddle and Christie-Blick (1985) define a horsetail splay as “one of a set of curved fault
splays near the end of a strike-slip fault that merge with that fault. The set forms an array that crudely resembles a horse’s
tail”. The term need not be restricted to strike-slip faults. Also see pinnate.
Hybrid fracture [kinematic and mechanical]: a fracture that developed by synchronous Mode I and shear modes (e.g., Ramsay
and Chester, 2004).
Hydrofracture [mechanical]: fracture created by fluid pressures that exceed the applied stresses to cause “effective tension” (e.g.,
Price and Cosgrove, 1990). The process of “hydraulic fracturing” (also called “hydrofracking” or “fracking”) is commonly
used to enhance recovery of hydrocarbons (e.g., Mandl and Harkness, 1987). Also see induced fracture.
Hydrothermal [geological]: relating to the activities hot fluids (i.e., higher than the expected geothermal gradient) in rock,
including alteration (e.g., Uglow, 1913), metamorphism (e.g., Naboko, 1963), mineralisation (e.g., Saprykin, 1967) and
brecciation (e.g., Katz et al., 2006).

I
I-node [topological]: the point at the end of a line, e.g., the isolated (blind) tip of a fault or other type of fracture (Fig. 2)
(Manzocchi, 2002; Sanderson and Nixon, 2015).
Imbricate [geometric]: an arrangement of faults that splay off a common fault but that do not meet again to form a duplex.
Imbricate systems have been described for thrust (e.g., Butler, 1982), normal (e.g., Gibbs, 1984) and strike-slip (Woodcock
and Fischer, 1986) faults. Also see duplex and horsetail.
Induced fracture [mechanical]: fracture created when the stresses concentrated around a borehole cause failure of the well walls
(e.g., Aadnoy, 1990). Antonym of natural fracture. Also see hydro-fracture.
Infinite cluster [topological]: see spanning cluster.
Injectite [geological]: see clastic dyke.
Intensity, fault or other type of fracture [geometric]: the fracture area per unit volume of rock, or fracture length per unit area on a
map (Wheeler and Dixon, 1980). Commonly used more loosely to mean the number of faults or other types of fractures per
unit area (or per unit length or unit volume), or the strain that those faults or other types of fractures represent, as used for
deformation bands (e.g., Okubo and Schultz, 2005), faults (e.g., Jamison, 1989) and veins (e.g., Gomez-Rivas et al., 2014).
Also see density, fault or other type of fracture and frequency, fault or other type of fracture.
Interaction damage zone [kinematic and mechanical]: we define this as a general term for the area of deformation caused by the
interaction between two or more faults. Interaction damage zones can be divided into approaching- and intersection-
damage zones (Fig. 5). It is a more general term than linking-damage zone of Kim et al. (2004) because it specifically
includes deformation between faults of any orientation that interact with each other. A later fault may interact with an
earlier fault that represents a mechanical barrier to the propagation of the later fault. Also see damage zone.
Interaction, fault or other type of fracture [kinematic and mechanical]: any relationship whereby the development of one fault or
other type of fracture influences the geometry or kinematics of other faults or other types of fractures. The interacting
faults or other types of fractures can have the same or different orientations, sizes, slip directions, ages, etc. This
relationship has been described for dykes (e.g., Zhang et al., 2014), faults (e.g., Aydin and Schultz, 1990), joints (e.g.,
Whitaker and Engelder, 2005), and veins (e.g., Virgo et al., 2014). See interaction, mechanical.
Interaction, mechanical [mechanical]: where the behaviour and development of one fault or other type of fracture is influenced
by another fault or other type of fracture or faults or other types of fractures (e.g., Segall and Pollard, 1980; Aydin and
Schultz, 1990; Bürgmann et al., 1994). See interaction, fault or other type of fracture.
Intersection damage zone [geometric, kinematic and mechanical]: we define this as the area of deformation around the
intersection point of two or more faults. Peacock et al. (in review). Also see approaching damage zone, damage zone, and
linking damage zone.
Intersection line [geometric]: the line along which two fault or other type of fracture planes meet. Approximately synonymous
with branch line. Used for deformation bands (e.g., Stel, 1991), faults (e.g., Maerten et al., 2000), etc. We prefer
intersection line to branch line because it is based on geometry and implies no origin, i.e., it can be used if it is not known
whether one fault or other type of fracture branched off another, or whether they were originally unconnected.
Peacock et al., Glossary of fault and fault and other fracture networks Page 7
Intersection point [geometric]: the point at which two fault or other type of fracture traces meet. Approximately synonymous with
branch point, which is used for faults. Used for faults (e.g., Kelly et al., 1998), lineaments (e.g., Bagheri, 2015), etc. We
prefer intersection point to branch point because it is based on geometry and implies no origin, i.e., it can be used if it is
not known whether one fault or other type of fracture branched off another, or whether they were originally unconnected.
Intersection, fault or other type of fracture [kinematic]: the process by which, condition in which, or location at which, two
originally separate faults or other types of fractures become physically connected.. It is therefore synonymous with the
term linkage as used by Pollard and Aydin (1984). Intersection is preferred to linkage, however, because the term linkage
appears to have become limited to refer to connection between sub-parallel faults or other types of fractures (e.g., Kim et
al., 2004), e.g., linkage across relay ramps (e.g., Peacock and Sanderson, 1991). Described for dykes (e.g., Caldwell and
Young, 2013), faults (Spotila and Anderson, 2004), fractures (e.g., Reks and Gray, 1982) and veins (e.g., Anderson and
Nash, 1972).
Isolated branch [topological]: a line with two unconnected ends (I-nodes), e.g., a fault or other type of fracture trace that has
isolated tips at both ends (Fig. 4) (Ortega et al., 2006; Sanderson and Nixon, 2015). Synonymous with isolated fault or
other type of fracture. Also called finite cluster. Also see singly-connected branch, doubly-connected branch, node,
connecting node and topology.
Isolated cluster [topological]: a group of interconnected faults or other types of fractures that form a cluster that is not large
enough to provide a pathway across a sample area (Roberts et al., 1998; Sahimi, 2011). They are often referred to as finite
clusters and do not have the potential to percolate, unlike a spanning cluster (Sahimi, 1994; Aizenman, 1997; Hunt et al.,
2014). Also see cluster and spanning cluster.
Isolated fault or other type of fracture [geometric and kinematic]: a fault or other type of fracture that was not affected by
interaction with other faults or other types of fractures during its propagation (for isolated faults, see Muraoka and
Kamata, 1983; Walsh and Watterson, 1987). An isolated fault is usually characterised by a displacement maximum near
the centre of the fault trace, with displacement decreasing approximately linearly towards the tips (Barnett et al., 1987;
Walsh and Watterson, 1987). Harding and Lowell (1979) describe solitary structures that are isolated, singular features.
Nicol et al. (1996) describe unrestricted faults, that are blind (do not intersect the free surface) and that have not interacted
with other faults or with substantial bodies of incompetent rock. The traces of such isolated faults or other types of
fractures would show two I-nodes.

J
Joint [geological]: defined by the Oxford English Dictionary (1989) as “a crack or fissure intersecting a mass of rock; usually
occurring in sets of parallel planes, dividing the mass into more or less regular blocks”, with first usage attributed to
Holland (1601). Price (1966) defines joints as “cracks and fractures in rock along which there has been extremely little or
no movement”, with some authors describing shear joints having a shear displacement (e.g., Hancock, 1985). Pollard and
Segall (1987) show that joints have two parallel surfaces that meet at the fracture front, that these surfaces are
approximately planar, and that the relative displacement of originally adjacent points across the fracture is small compared
to fracture length. Pollard and Aydin (1988) suggest that joints are generally associated with the opening mode, whereas
faults are associated with the shearing modes, and propose that the word “joint” be restricted to those fractures with field
evidence for dominantly opening displacements. We suggest that fractures with dominantly opening displacements but that
have mineral fills should be called veins. Pollard and Aydin (1988) criticise definitions of joints as having no discernible
displacement, because they would not exist without some displacement. They are particularly critical of the concept of
“shear joints”, which they regard as small faults. Pollard and Aydin (1988) define, describe and illustrate various joint
geometries (e.g., cross, diagonal, dip, ladder, normal, polygonal and strike joints), mechanisms (e.g., contraction, cooling,
shrinkage and unloading joints) and surface patterns (e.g., hackle and rib marks and plumose structures).

K, L
Kinematic coupling [kinematic]: where the displacements or strains of two or more faults or other types of fractures are related
(Tvedt et al., 2013), with or without geometric coupling (e.g., Thomas and Woodcock, 2015). Also see soft linkage.
Kinematic linkage [kinematic]: where displacements or strains of two or more (usually coeval) faults or other types of fractures
are related (e.g., Webb and Johnson, 2006). The faults or other types of fractures need not have geometric linkage.
Layer-bound fault or other type of fracture [geometric and kinematic]: fault or other type of fracture that is restricted to a single
bed or mechanical layer. Synonymous with bed-bound fault or other type of fracture. See mechanical stratigraphy. Used
for deformation bands (e.g., Antonellini and Mollema, 2015), faults (e.g., Cartwright and Lonergan, 1996), joints (e.g.
Kettermann et al., 2016), stockworks (e.g., de Vries and Touret, 2007) etc.
Length [geometric]: the distance between the tips of a fracture, usually measured along a fracture trace and in the horizontal plane
(e.g., Scholz and Cowie, 1990). Walsh and Watterson (1988) define fault length as the maximum horizontal distance of a
fault plane observed in 3D. A component of connectivity. Joint lengths are described by Odling (1997).

Peacock et al., Glossary of fault and fault and other fracture networks Page 8
Lens [geometric]: fault-bound zone of relatively weakly-deformed host rock within a fault zone (e.g., Swanson, 2005). Also called
shear lens (e.g. Swanson, 1990). Also see horse.
Lineament [geological]: defined by Biddle and Christie-Blick (1985) as “a linear topographic feature of regional extent that is
thought to reflect crustal structure”.
Linkage, fault or other type of fracture [kinematic and mechanical]: a general term for the situation in which two originally
separate (usually coeval) faults or other types of fractures become connected (Pollard and Aydin, 1984). The connection
may be through physical intersection of the fracture surfaces (geometric linkage) or by sharing displacement
characteristics (kinematic linkage) in relay ramps or approaching fracture tips (Fig. 3).
Linking damage zone [kinematic and mechanical]: area of deformation at steps between two sub-parallel (usually coeval) faults
(Fig. 5) (Kim et al., 2004). This term is therefore more restricted than our term interaction damage zone. Also see damage
zone.

M, N
Mechanical stratigraphy [geometric and mechanical]: where the stratigraphy is defined in terms of rheology, such that different
mechanical units control the types or frequencies of structures (e.g., Corbett et al., 1987). Also see bed-bound fault or other
type of fracture and layer-bound fault or other type of fracture.
Microcrack [geometric and mechanical]: fractures with microscopic-scale lengths, apertures, etc. (e.g., Robertson, 1960).
Mode I [mechanical]: fracture propagation by extension (e.g., Atkinson, 1987, fig. 1.1). Also see extension fracture, tension
fracture and tensile fracture.
Mode II [mechanical]: fault propagation by in-plane shearing or sliding, i.e., perpendicular to the fault tip line (e.g., Atkinson,
1987, fig. 1.1). Also see fault.
Mode III [mechanical]: fault propagation by anti-plane shearing or tearing, i.e., parallel to the fault tip line (e.g., Atkinson, 1987,
fig. 1.1). Also see fault.
Mud crack [geological]: see desiccation crack.
Natural fracture [geological]: fracture resulting from naturally occurring stresses and fluid pressures, i.e., not related to human
activities. Antonym of induced fracture.
Neptunian dyke [mechanical]: fracture filled with sediments from the Earth’s surface, e.g., Strachan et al. (1948). Also see clastic
dyke and sedimentary dyke.
Network, fault or other type of fracture [geometric]: a system of linked and interacting faults or other types of fractures, more
diffuse than a fault or other type of fracture zone (e.g., Stočes and White, 1935). When more than two fault or other type of
fracture sets are present, the network represents a triaxial strain system (Reches, 1978; Krantz, 1988). Networks have been
described for faults (e.g., Nixon et al., 2014), veins (e.g., Garofalo et al., 2002), joints (e.g., Zhang and Sanderson, 1996)
and dykes (Reichardt and Weinberg, 2012, fig. 17).
Node [topological]: a point where a line ends (I-node) or intersects with another line (X-node or Y-node) (Figs. 2 and 4), e.g., the
tip or intersection point of a fault or other type of fracture. Three main types of node can be recognised in fault or other
type of fracture networks, representing isolated tips (I-nodes), abutments or splays (Y-nodes), and crossing faults or other
types of fractures (X-nodes) (Manzocchi, 2002; Nixon et al., 2012; Sanderson and Nixon, 2015). The proportions of the
different node types can be used to characterise the network (Fig. 4d) and calculate topological measures that describe the
networks connectivity (e.g., Manzocchi, 2002; Mäkel, 2007; Sanderson and Nixon, 2015).
Non-conductive fracture [geological]: a fracture within which fluids do not flow (e.g., Fransson, 2007), or that in non-conductive
to electricity on borehole image logs (e.g., Samantray et al., 2010). Note that some open fractures may be electrically non-
conductive (e.g., Boutt et al., 2010). See closed fracture. Antonym of conductive fracture.
Normal fault [geological and kinematic]: fault in which the hanging-wall (the fault block sitting above the fault plane) are
displaced downwards relative to the footwall (rocks below the fault) (e.g., Dennis, 1967).

O
Oblique [geometric]: lines or planes that are not parallel to each other.
Oblique-slip fault [geological and kinematic]: fault with displacement involving components of both dip-slip and strike-slip (e.g.,
Speed and Cogbill, 1979).
Open fracture [geological]: commonly used in the hydrocarbon industry for fluid-filled fractures (e.g., Wennberg et al., 2009).
See conductive fracture.
Opening-mode [mechanical]: see Mode I.
Organisation, network [geometric]: geometric description of a fault or other type of fracture network that involves determining
the sets, orientations and spacing of fault and other fractures.
Orthogonal [geometric]: lines or planes that are at 90° to each other.
Overstep, fault or other type of fracture [geometric]: area between two sub-parallel, non-collinear faults or other types of
fractures. Synonymous terms for fault oversteps include jog (Sibson, 1989), offset, overlapping faults and stepover (Aydin
Peacock et al., Glossary of fault and fault and other fracture networks Page 9
and Nur, 1982). Oversteps have been described for faults (e.g., Biddle and Christie-Blick, 1985), stylolites (Ben-Itzhak et
al., 2014), etc.

P, Q
Partly-open fracture [geological]: commonly used in the hydrocarbon industry to denote a fracture that has an aperture partly-
filled by minerals and partly by fluid (e.g., Wennberg et al., 2016).
Percolation threshold [geometric and topological]: the critical point at which a fault or other type of fracture network becomes
connected enough to produce a spanning cluster (PC in Fig. 4) that could potentially allow flow of fluids through a rock
mass (Stauffer and Aharony, 1991; Sahimi, 1994).
Percolation [geometric and topological]: describes the phenomena of fluid flow through a medium (Broadbent and Hammersley,
1957; Sahimi, 1993). For example, a fault or other type of fracture network is said to be percolating when it is
macroscopically open and connected to allow flow of fluids through a rock mass (Sahimi, 1994; Adler and Thovert, 1999).
Pinnate [geometric and kinematic]: extension fracture developed near the tips of shear fractures (e.g., Segall and Pollard, 1983;
Crider and Peacock, 2004) to accommodate displacement variations along the shear fractures. Also see horsetail and wing
crack.
Polygonal faults [geometric]: a network of faults with no regionally consistent preferred strike orientation (e.g., Lonergan et al.,
1998. Layer-bound polygonal faults in the North Sea are described by Cartwright and Lonergan (1996).
Polymodal [geometric and kinematic]: pattern of faults or other types of fractures where three, four or more sets form and slip
simultaneously (Healy et al., 2015). Such a pattern of faulting is associated with a triaxial strain rather than a plane strain
(Reches, 1978; Healy et al., 2015). Also see bimodal and conjugate.
Population [geometric]: the fractures of all scales that exist in an area or region, formed during one or more deformation events.
Fracture population analysis involves the study of the scaling relationships and strain of the fractures, usually divided into
the different deformation events (e.g., Cowie et al., 1996).
Pull-apart [geometric and kinematic]: a rhomb-shaped opening along an extensional bend on a fault or at an extensional step
between two faults (e.g., Peacock and Anderson, 2012).
Process zone [mechanical]: used in material science for an area of micro-cracking at the tip of a propagating fault or other type of
fracture, involving non-linear behaviour. Process zones in rock are generally dilational areas (Atkinson, 1987, fig. 1.4),
although process zones in normal fault zones involving compactive shear bands have also been described by Cowie and
Shipton (1998) and Rotevatn and Fossen (2012). Also see tip damage zone (Kim et al., 2004).
Propagation [mechanical]: increase in length and area of a fault or other type of fracture or fault or other type of fracture zone,
commonly involving an increase in displacement. Propagation has been described for dykes (e.g., Okamura et al., 1988),
faults (e.g., Walsh and Watterson, 1988; Reches and Lockner, 1994), veins (e.g., Grimm and Orange, 1997) etc.

R
Reactivation [kinematic]: renewed displacement on a fault that has undergone a prolonged period of inactivity (e.g., Shephard-
Thorn et al., 1972; Sibson, 1985). The different displacement events may or may not be of the same sense. Also see
trailing.
Relay pattern [geometric]: an arrangement of overlapping or underlapping sub-parallel faults (e.g., Harding and Lowell, 1979;
Biddle and Christie-Blick, 1985, fig. 2). Relay faults are described by Goguel (1952) and relay ramps between normal
faults are described by Larsen (1988). The term does not appear to be applied to other types of fracture.
Relay ramp [geometric and kinematic]: zone of kinematic linkage between overlapping, geometrically uncoupled, sub-parallel
faults, where strain is relayed from one fault to the other (Larsen, 1988; Peacock and Sanderson, 1991, Rotevatn et al.,
2007). If the faults bounding the relay ramp become geometrically coupled, a breached relay (Trudgill and Cartwright,
1994) is formed.
Relay zone/structure [geometric and kinematic]: zone of geometric or kinematic linkage between sub-parallel (usually coeval)
fault segments (Fossen and Rotevatn, 2016). Encompasses both relay ramps and breached relay ramps.
Reverse fault [geological and kinematic]: steeply-dipping fault (usually considered to have a dip of more than 30°), in which the
hanging-wall (rocks above the fault) are displaced upwards in relation to the footwall (rocks below the fault) (e.g., Brown,
1984, fig. 3). Also see thrust fault.
Roughness, fault or other type of fracture [geometricl]: measure of the degree to which the fault or other type of fracture surface
deviates from planar (e.g., Tsang and Witherspoon, 1983).

S
Saturation, joint [mechanical]: where the spacing of a joint set in a bed reaches a narrow unimodal distribution (e.g., Becker and
Gross, 1996, fig. 10).
Seal [geological]: a barrier to fluid flow in rock, including rock layers (e.g., Downey, 1984) and faults (e.g., Yielding et al., 1997;
Reilly et al., 2016).
Peacock et al., Glossary of fault and fault and other fracture networks Page 10
Sedimentary dyke [geological]: see clastic dyke and neptunian dyke.
Segment [geometric]: an individual fault or other type of fracture plane that is part of a set of sub-parallel faults or other types of
fractures that together form a fault zone or fracture zone (e.g., Segall and Pollard, 1980).
Set, fault or other type of fracture [geometric]: a group of the same type of faults or other types of fractures in a region with the
similar orientations and usually inferred to be genetically-related. Used for faults (e.g., Nieto-Samaniego, 1999), joints
(e.g., Bai et al., 2002), veins (e.g., Stowell et al., 1999), etc.
Shear fracture [mechanical]: fracture formed by modes II and III propagation (e.g., Belayneh and Cosgrove, 2010, fig. 1).
Identified by displaced markers, fault rocks, slickensides, etc. (e.g., Pallister et al., 2013). See fault.
Shear joint [geometric and kinematic]: a joint that displays and was formed by a “minor” shear displacement (e.g., Qidong and
Peizhen, 1982). The term is rejected by Pollard and Aydin (1988) because any shear displacement would make the fracture
a fault.
Sill [geological]: defined by the Oxford English Dictionary (1989) as “a bed, layer, or stratum of rock, esp. of an intrusive igneous
rock. In mod. use, a tabular igneous intrusion lying parallel to the surrounding strata”, with first usage attributed to
Hutchinson (1794). The term is now most commonly used for a horizontal fracture (or fracture sub-parallel to layering)
filled by intrusive igneous rock (e.g., de Sitter, 1964).
Singly-connected branch [topological]: a line with one connecting node and one isolated node (I-node), thus often referred to as
an I-C branch (Sanderson and Nixon, 2015; Morley and Nixon, 2016), e.g. a fault trace that intersects another fault in one
direction but with a blind tip in the other direction. Also see topology, branch, isolated branch, doubly-connected branch,
connecting node. Synonymous with dangling end.
Slickenside [geological]: a striation along a fault plane caused by displacement along the fault (e.g., Bates and Jackson 1980).
Also called slickenside lineations.
Slickolite [geological]: plane made pressure solution along a fault plane, consisting of ridges and groove or half columns that fit
into the ridges and groove on the opposite face of the plane (e.g., Nitecki, 1962). They commonly form along pre-existing
discontinuities orientated obliquely to the maximum compressive stress direction (e.g., Sinha-Roy, 2002). Also see
stylolite.
Slip plane [geological]: used in material science for a plane along which displacement may occur (e.g., Dennis, 1967). Used in
geology for a surface along which displacement has occurred (e.g., Stewart and Hancock, 1991). Synonymous with fault
plane.
Slip surface [geological and mechanical]: see fault plane and slip plane.
Slip vector [kinematic]: the direction of motion of one wall of a fault relative to the other wall of that fault (e.g., Reid et al., 1913).
Soft-linkage [kinematic]: where there is a kinematic coherency between faults that are geometrically uncoupled, achieved by
distributed strain of the wall-rocks, i.e., there is no linkage by faults visible at the scale of observation (Walsh and
Watterson, 1991). A relay ramp (e.g., Larsen, 1988) is an example of soft-linkage, across which strain is relayed between
adjacent (but geometrically uncoupled) faults. Also see hard-linkage.
Spacing, fault or other type of fracture [geometric]: the distance between faults or other type of fracture traces or planes,
normally measured perpendicular to the faults or other types of fractures. Studies on spacing have been carried out on
faults (e.g., Brooks et al., 1996; Sharples et al., in press), joints (e.g., Narr and Suppe, 1991; Rabinovitch et al., 2012) and
veins (e.g., Van Noten and Sintubin, 2010).
Spanning cluster [topological]: a group of interconnected lines that together form a cluster large enough to span a sample area
(Fig. 4), e.g., faults or other types of fractures that are connected such that they cross an area of interest (Roberts et al.,
1998; Sahimi, 2011). They are often referred to as infinite clusters and have the potential to percolate (Sahimi, 1994;
Aizenman, 1997; Hunt et al., 2014). Also see cluster and isolated cluster.
Splay fault [geometric]: one or more smaller faults that join a larger fault to which it is related (e.g., Biddle and Christie-Blick,
1985, fig. 2). The larger fault “splays” if it is connected with one or more splay faults (de Sitter, 1956).
Step, fault or other type of fracture [geometric]: see overstep, fault or other type of fracture.
Stockwork [geological]: a network of veins formed from several sets or randomly orientated veins (e.g., Linnel and Williams-
Jones, 1987).
Strand [geological]: defined by Biddle and Christie-Blick (1985) as “an individual fault of a set of closely spaced parallel or
subparallel faults of a fault system”.
Strike-linkage [kinematic]: the across-dip, or horizontal, linkage of two (usually coeval) faults or other types of fractures that were
initially geometrically uncoupled (cf. Mansfield and Cartwright, 1996).
Strike-slip fault [geological and kinematic]: fault in which the displacement is in the horizontal plane, parallel to the strike of the
fault (e.g., Biddle and Christie-Blick, 1985).
Stylolite [geological]: surface on which insoluble residues are concentrated as pressure solution removes such soluble minerals as
calcite or quartz (e.g., Stockdale, 1926). The teeth of stylolites point in the shortening direction, and, assuming the
stylolites initiated as flat planes and did not propagate out of plane, their amplitudes represent a minimum estimate of the
amount of shortening (compaction) that has occurred (e.g., Rispoli, 1981). Stylolites have been called anticracks (Fletcher
Peacock et al., Glossary of fault and fault and other fracture networks Page 11
and Pollard, 1981) because contraction occurs across the plane. “Bedding-parallel stylolites” are usually assumed to be
formed by the weight of the overburden (e.g., Safaricz and Davison, 2005), while “tectonic stylolites” are at a high angle to
bedding or sub-horizontal and formed by tectonic stresses (e.g., LetouzeyTrémolières, 1980). Single or multiple sets of
stylolites can form stylolite networks (e.g., Ben-Itzhak et al., 2014). Also see slickolite.
Surface, fault or other type of fracture [geometric]: a surface created by brittle failure of rock. Applied to dykes (e.g., Kattenhorn
and Watkeys, 1999), faults (e.g., Gephart, 1990), joints (e.g., Pollard and Aydin, 1988), etc.
Swarm, fault or other type of fracture [geometric]: cluster of closely spaced faults or other types of fractures (e.g., Belayneh et al.,
2007). Also see corridor, fault or other type of fracture.
Synthetic fault [geometric and kinematic]: originally defined by Cloos (1928) as a minor fault that dips in the same direction as
dipping beds. Synthetic fault is now used for a minor fault that has the same shear sense and a similar orientation to a
related major fault (e.g., Gibbs, 1984). Also see conjugate.

T, U
Tensile fracture [kinematic and mechanical]: see extension(al) fracture (e.g., McGrath and Davison, 1995) and mode I. Also
called tension fracture.
Tension gash [geological]: see array, fault or other type of fracture.
Tension(al) fracture [kinematic and mechanical]: see extension(al) fracture (e.g., McGrath and Davison, 1995) and mode I. Also
called and tensile fracture. See Pollard and Aydin (1988, p. 1186).
Thrust fault [geological and kinematic]: gently-dipping fault (usually considered to have a dip of less than 30°), in which the
hanging-wall (rocks above the fault) are displaced upwards in relation to the footwall (rocks below the fault) (e.g., Willis,
1935). Also see reverse fault.
Tip damage zone [kinematic and mechanical]: area of deformation formed in response to stress concentration at a fault tip (Fig. 5)
(Kim et al., 2004); see also process zone.
Tip line [kinematic]: line of zero displacement around the ends of a fracture, e.g., around a fault (e.g., Walsh and Watterson,
1987). Also see tip, fault or other type of fracture.
Tip point, fault or other type of fracture [mechanical]: the point in 2D view at which displacement on a fault or other type of
fracture trace decreases to zero (e.g., Atkinson, 1987; Walsh and Watterson, 1988; Cowie and Shipton, 1998). Also see I
node.
Trace length [geometric]: see length.
Trace [geometric]: the line made by a fault or other type of fracture as in intersects the plane of observation (e.g., Bridwell, 1975;
Muraoka and Kamata, 1983).
Trailing [geometric and kinematic]: where two new faults or other types of fractures are connected via an older fault or other type
of fracture, on which renewed displacement occurs to connect the two later faults or other types of fractures. Trailing
faults are illustrated by Nixon et al. (2014) and trailing veins shown by Virgo et al. (2013, fig. 12c).
Transfer fault [kinematic]: fault that allows kinematic-linkage between two other faults, commonly at a high angle to those faults
(e.g., Karson and Rona, 1990).
Transform fault [geological]: defined by Biddle and Christie-Blick as “a strike-slip fault that acts as a lithospheric plate boundary
and terminates at both ends against major tectonic features such as oceanic ridges subduction zones or rarely other
transform faults that are also plate boundaries”, i.e., it enables geometric- and kinematic-linkage between plate-scale
structures.
Triple junction [geological]: where three fracture traces at approximately 120º to each other meet at a point, commonly used for
the boundaries of three tectonic plates (e.g., McKenzie and Morgan, 1969). Also see polygonal faults, cooling joints.

V, W
Vein [geological]: defined by the Oxford English Dictionary (1989) as “a deposit of metallic or earthy material having an
extended or ramifying course under ground; a seam or lode; spec. a continuous crack or fissure filled with matter (esp.
metallic ore) different from the containing rock”. First usage is credited to Trevisa (1387). A wide range of minerals can
fill veins, with quartz, calcite and anhydrite being especially common. See Bons et al. (2012) for a review.
Vertex [topological]: the midpoint of a line, e.g., the centre fault or other type of fracture trace (Andresen et al., 2013; Morley and
Nixon, 2016). See also edge, equivalent network and vertex degree.
Wall damage zone [kinematic and mechanical]: area of deformation resulting from the propagation of faults through rock (Fig. 5),
or from damage associated with the increase in slip on a fault (Kim et al., 2004). See damage zone.
Wing crack [kinematic]: extension fracture developed near the tips of shear fractures (e.g., Wilkins et al., 2001) to accommodate
displacement variations along the shear fractures. Also see horsetail and pinnate.

X, Y, Z

Peacock et al., Glossary of fault and fault and other fracture networks Page 12
X-node [topological]: the intersection point between two crossing faults or other types of fractures, with the traces thereby
forming an X pattern (Fig. 2) (Manzocchi, 2002; Sanderson and Nixon, 2015). They are rare in fault networks due to the
difficulty in preservation with increased displacements (Nixon et al., 2012). See topology, node, connecting.
Y-node [topological]: where one fault or other type of fracture ends at another fault or other type of fracture, with the traces
thereby forming a Y pattern (Fig. 2). Such a geometric relationship can be produced in a number of ways including
splaying, abutting and cross-cutting relationships (Manzocchi, 2002; Sanderson and Nixon, 2015). See topology, node,
connecting node.
Zone, fault or other type of fracture [geometric]: a system of related fracture segments that interact and link, and are restricted to a
relatively narrow band or volume (e.g., Nevin, 1931). Childs et al. (2009) define fault zone thickness as “the distance
between synthetic slip-surfaces (same dip direction and sense of offset) that can be demonstrated at outcrop to be
kinematically related and which each accommodate at least a few percent of the total offset”.

Acknowledgements
We are grateful to Ian Alsop, Alex Manda and an anonymous reviewer for their careful reviews. Kristin Hool and Torill
Jordal at NAV are thanked for facilitating DCPP working on a placement at UiB. Nixon is supported by a VISTA scholarship
from the Norwegian Academy of Science and Letters. We acknowledge financial support for the ANIGMA project from the
Research Council of Norway (project no. 244129/E20) through the ENERGIX program, and from Statoil ASA through the
Akademia agreement.

References
Aadnoy, B.S., 1990. Inversion technique to determine the in-situ stress field from fracturing data. Journal of Petroleum Geology
and Engineering 4, 127-141.
Adler, P.M., Thovert, J.F.,1999. Fractures and Fracture Networks (Vol. 15). Dordrecht: Kluwer Academic Publishers.
Aizenman, M., 1997. On the number of incipient spanning clusters. Nuclear Physics B 485, 551–582.
Alghalandis, Y.F., Dowd, P.A., Xu, C., 2015. Connectivity field: a measure for characterising fracture networks. Mathematical
Geosciences 47, 63-83.
Anderson, C.A., Nash, J.T., 1972. Geology of the massive sulfide deposits at Jerome, Arizona-a reinterpretation. Economic
Geology 67, 845-863.
Anderson, E.M., 1951. The Dynamics of Faulting. Oliver and Boyd, Edinburgh.
Andresen, C.A., Hansen, A., Le Goc, R., Davy, P., Hope, S.M., 2013. Topology of fracture networks. Frontiers in Physics 1, 1-5.
Antonellini, M., Mollema, P.N., 2015. Polygonal deformation bands. Journal of Structural Geology 81, 45-58.
Atkinson, B.K., 1987. Introduction to fracture mechanics and its geophysical applications. In: Atkinson, B.K. (Ed.), Fracture
Mechanics of Rock. Academic Press, London, 1-26.
Aydin, A. 1978. Small faults formed as deformation bands in sandstone. Pure and Applied Geophysics, 116, 913-930.
Aydin, A., Nur, A., 1982. Evolution of pull-apart basins and their scale-independence. Tectonics 1, 91-105.
Aydin, A., Schultz, R.A., 1990. Effect of mechanical interaction on the development of strike-slip faults with echelon patterns.
Journal of Structural Geology 12, 123-129.
Aydin, A., Borja, R. I., Eichhubl, P., 2006. Geological and mathematical framework for failure modes in granular rock. Journal of
Structural Geology, 28, 83-98.
Bagheri, H., 2015. Crustal lineament control on mineralization in the Anarak area of Central Iran. Ore Geology Reviews 66, 293-
308.
Bahat, D., 1986. Joints and en échelon cracks in Middle Eocene chalks near Beer Sheva, Israel. Journal of Structural Geology 8,
181-190.
Bai, T., Maerten, L., Gross, M.R., Aydin, A., 2002. Orthogonal cross joints: do they imply a regional stress rotation? Journal of
Structural Geology 24, 77-88.
Bakewell, R. 1813. An Introduction to Geology, Illustrative of the General Structure of the Earth: Comprising the Elements of the
Science, and an Outline of the Geology and Mineral Geography of England. Richard Taylor and Co., London (cited by
Oxford English Dictionary, 1989).
Balberg, I., Binenbaum, N. 1983. Computer study of the percolation threshold in a two-dimensional anisotropic system of
conducting sticks Physica Review B, 28, 3799-3812.
Barnett, J.A.M., Mortimer, J., Rippon, J.H., Walsh, J.J., Watterson, J., 1987. Displacement geometry in the volume containing a
single normal fault. Bulletin of the American Association of Petroleum Geologists 71, 925-937.
Bates R.L., Jackson J.A. (eds), 1980. Glossary of Geology. American Geological Institute 749 p.
Beach, A., 1975. The geometry of en-echelon vein arrays. Tectonophyics 28, 245-263.
Becker, A., Gross, M.R., 1996. Mechanism for joint saturation in mechanically layered rocks: An example from southern Israel.
Tectonophysics 257, 223-237.

Peacock et al., Glossary of fault and fault and other fracture networks Page 13
Belayneh, M., Cosgrove, J.W., 2010. Hybrid veins from the southern margin of the Bristol Channel Basin, UK. Journal of
Structural Geology 32, 192-201.
Belayneh, M., Matthäi, S.K., Cosgrove, J.W., 2007. The implications of fracture swarms in the Chalk of SE England on the
tectonic history of the basin and their impact on fluid flow in high-porosity, low-permeability rocks. In: Ries, A.C., Butler,
R.W.H., Graham, R.H. (eds). Deformation of the Continental Crust: the Legacy of Mike Coward. Geological Society,
London, Special Publications, 272, 499-517.
Ben-Itzhak, L.L., Aharonov, E., Karcz, Z., Kaduri, M., Toussaint, R., 2014. Sedimentary stylolite networks and connectivity in
limestone: large-scale field observations and implications for structure evolution. Journal of Structural Geology 63, 106-
123.
Berkowitz, B., Bour, O., Davy, P., Odling, N., 2000. Scaling of fracture connectivity in geological formations. Geophysical
Research Letters 27, 2061-2064.
Biddle, K.T., Christie-Blick, N., 1985. Glossary - strike-slip deformation, basin formation, and sedimentation. In: Biddle, K.T.,
Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. Society of Economic
Mineralogists Special Publication 37, 375-386.
Bons, P.D., Elburg, M.A., Gomez-Rives, E., 2012. A review of the formation of tectonic veins and their microstructures. Journal
of Structural Geology 43, 33-62.
Bour, O., Davy, P., 1999. Clustering and size distributions of fault patterns: theory and measurements. Geophysical Research
Letters 26, 2001-2004.
Boutt, D.F., Diggins, P., Mabee, S.B., 2010. A field study (Massachusetts, USA) of the factors controlling the depth of
groundwater flow systems in crystalline fractured-rock terrain. Hydrogeology Journal 18, 1839-1854.
Bridwell, R.J., 1975. Sinuosity of strike-slip fault traces. Geology 3, 630-632.
Broadbent, S.R., Hammersley, J.M., 1957. Percolation processes. Mathematical Proceedings of the Cambridge Philosophical
Society 53, 629-641.
Brooks, B.A., Allmendinger, R.W., De La Barra, I.G., 1996. Fault spacing in the El Teniente Mine, central Chile: Evidence for
nonfractal fault geometry. Journal of Geophysical Research B: Solid Earth 101, 13633-13653.
Brown, R.J., Kavanagh, J., Sparks, R.S.J., Tait, M., Field, M., 2007. Mechanically disrupted and chemically weakened zones in
segmented dike systems cause vent localization: evidence from kimberlite volcanic systems. Geology 35, 815-818.
Brown, W.G., 1984. A reverse fault interpretation of Rattlesnake Mountain anticline, Big Horn Basin, Wyoming (USA).
Mountain Geologist 21, 31-35.
Bucher, W.H., 1920. The mechanical interpretation of joints. The Journal of Geology, 28, 707-730.
Bürgmann, R., Pollard, D.D., Martel, S.J., 1994. Slip distributions on faults: effects of stress gradients, inelastic deformation,
heterogeneous host-rock stiffness, and fault interaction. Journal of Structural Geology 16, 1675-1690.
Butler, R.W.H., 1982. Terminology of structures in thrust belts. Journal of Structural Geology 4, 239-245.
Caldwell, W.G.E., Young, G., 2013. The Cumbrae Islands: a structural Rosetta Stone in the western offshore Midland Valley of
Scotland. Scottish Journal of Geology 49, 117-132.
Cartwright, J.A., Lonergan, L., 1996. Volumetric contraction during compaction of mudrocks: a mechanism for the development
of regional-scale polygonal fault systems. Basin Research 8, 183-193.
Chambon, G., Rudnicki, J.W., 2001. Effects of normal stress variations on frictional instability of a fluid-infiltrated fault. Journal
of Geophysical Research 106, 11353-11372.
Chen, J., 2013. Normal fault patterns and their rationality analysis. Geophysical Prospecting for Petroleum 52, 201-206.
Cherpeau, N., Caumon, G., Caers, J., Lévy, B., 2011. Assessing the impact of fault connectivity uncertainty in reservoir studies
using explicit discretization. Society of Petroleum Engineers - SPE Reservoir Characterisation and Simulation Conference
and Exhibition 2011, Abu Dhabi, United Arab Emirates, 526-534.
Chester, F.M., 1989. Dynamic recrystallization in semi-brittle faults. Journal of Structural Geology 11, 847-858.
Chester, F.M., Logan, J.M., 1986. Implications for mechanical-properties of brittle faults from observations of the Punchbowl
fault zone, California. Pure and Applied Geophysics 124, 79-106
Childs, C., Manzocchi, T., Walsh, J.J., Bonson, C.G., Nicol, A., Schöpfer, M.P., 2009. A geometric model of fault zone and fault
rock thickness variations. Journal of Structural Geology 31, 117-127.
Choi, J.H., Edwards, P., Ko, K., Kim, Y.S., in press. Definition and classification of fault damage zones: a review and a new
methodological approach. Earth-Science Reviews.
Clark, M.S., Beckley, L.M., Crebs, T.J., Singleton, M.T. 1996. Tectono-eustatic controls on reservoir compartmentalisation and
quality - an example from the upper Miocene of the San Joaquin basin, California. Marine and Petroleum Geology 13, 475-
491.
Cloos, H., 1928. Über antithetische Bewegungen. Geologische Rundschau 19, 246-251.
Corbett, K., Friedman, M., Spang, J., 1987. Fracture development and mechanical stratigraphy of Austin Chalk, Texas. American
Association of Petroleum Geologists Bulletin 71, 17-28.

Peacock et al., Glossary of fault and fault and other fracture networks Page 14
Cowie, P.A., Scholz, C.H., 1992. Growth of faults by accumulation of seismic slip. Journal of Geophysical Research 97, 11085-
11095.
Cowie, P.A., Shipton, Z.K., 1998. Fault tip displacement gradients and process zone dimensions. Journal of Structural Geology
20, 983-997.
Cowie, P.A., Knipe, R.J., Main, I.G., 1996. Special issue. Scaling laws for fault and other fracture populations: analyses and
applications. Introduction. Journal of Structural Geology 18, R5-R11.
Crider, J.G., Peacock, D.C.P., 2004. Initiation of brittle faults in the upper crust: a review of field observations. Journal of
Structural Geology 26, 691-707.
Cruikshank, K.M., Zhao, G., Johnson, A.M., 1991. Duplex structures connecting fault segments in Entrada Sandstone. Journal of
Structural Geology 13, 1185-1196.
Daubrée, G.A., 1878. Recherches expérimentales sur les cassures qui traversent l’écorce terrestre, particulièrement celles qui sont
connues sous les noms de joints et failles. Academe Science (Paris), Comptes Rendus 86, 77-83, 283-289 (cited by Dennis,
1967).
Davatzes, N.C., Aydin, A., 2003. Overprinting faulting mechanisms in high porosity sandstones of SE Utah. Journal of Structural
Geology 25, 1795-1813.
de Sitter, L.U., 1956. Structural Geology. McGraw-Hill, New York.
de Sitter, L.U., 1964. Structural Geology. 2nd edition. McGraw-Hill, New York.
de Vries, S.T., Touret, J.L.R., 2007. Early Archaean hydrothermal fluids; a study of inclusions from the ∼ 3.4 Ga Buck Ridge
Chert, Barberton Greenstone Belt, South Africa. Chemical Geology 237, 289-302.
Delaney, P.T., Pollard, D.D., 1981. Deformation of host rocks and flow of magma during growth of minette dikes and breccia-
bearing intrusions near Ship Rock, New Mexico. United States Geological Survey Professional Paper 1202, 61 pages.
Dennis, J.G. (Ed.), 1967. International Tectonic Dictionary. American Association of Petroleum Geologists, Memoir 7.
Downey, .W., 1984. Evaluating seals for hydrocarbon accumulations. American Association Association of Petroleum Geologists
Bulletin 68, 1752-1763.
Du Bernard, X., Eichhubl, P., Aydin, A., 2002. Dilation bands: a new form of localized failure in granular media. Geophysical
Research Letters 29, 1-4.
Eichhubl, P., Hooker, J.N., Laubach, S.E., 2010. Pure and shear-enhanced compaction bands in Aztec Sandstone. Journal of
Structural Geology 32, 1873-1886.
Engelder, J.T., 1974. Cataclasis and the generation of fault gouge. Geological Society of America Bulletin 85, 1515-1522.
Fisher, Q.J., Knipe, R.J., 1998. Fault sealing processes in siliciclastic sediments. In: Jones, G., Fisher, Q.J., Knipe, R.J. (eds)
Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs. Geological Society, London, Special Publications 147,
117-134.
Fisher, Q.J., Knipe, R.J., 2001. The permeability of faults within siliciclastic petroleum reservoirs of the North Sea and Norwegian
Continental Shelf. Marine and Petroleum Geology 18, 1063-1081.
Fletcher, R.C., Pollard, D.D., 1981. Anticrack model for pressure solution surfaces. Geology 9, 419-424.
Flodin, E.A., Aydin, A., 2004. Evolution of a strike-slip fault network, Valley of Fire State Park, southern Nevada. Geological
Society of America Bulletin 116, 42-59.
Fookes, P.G., Parrish, D.G., 1969. Observations on small-scale structural discontinuities in the London Clay and their relationship
to regional geology. Quarterly Journal of Engineering Geology 1, 217-240.
Fossen, H., 2010. Structural Geology. Cambridge University Press, 480 p.
Fossen, H., Rotevatn, A., 2016. Fault linkage and relay structures in extensional settings - a review. Earth-Science Reviews 154,
14-28.
Fossen, H., Hesthammer, J., Johansen, T.E.S., Sygnabere, T.O., 2003. Structural geology of the Huldra Field, northern North Sea -
a major tilted fault block at the eastern edge of the Horda Platform. Marine and Petroleum Geology 20, 1105-1118.
Fossen, H., Schultz, R.A., Shipton, Z.K., Mair, K., 2007. Deformation bands in sandstone: a review. Journal of the Geological
Society 164, 755-769.
Fowles, J., Burley, S. 1994. Textural and permeability characteristics of faulted, high porosity sandstones. Marine and Petroleum
Geology, 11, 608-623.
Fransson, Å., 2007. A case study to verify methods for estimating transmissivity distributions along boreholes. Hydrogeology
Journal 15, 307-313.
Freund, R., 1974. Kinematics of transform and transcurrent faults. Tectonophysics 21, 93-134.
Friedman, M., Logan, J.M., 1973. Lüders' bands in experimentally deformed sandstone and limestone. Geological Society of
America Bulletin 84, 1465-1476.
Gabrielsen, R.H., Braathen, A., 2014. Models of fracture lineaments - Joint swarms, fracture corridors and faults in crystalline
rocks, and their genetic relations. Tectonophysics, 628, 26-44.
Garofalo, P., Mattha, S.K., Heinrich, C.A., 2002. Three-dimensional geometry, ore distribution and time-integrated mass transfer
through the quartz-tourmaline-gold vein network of the Sigma deposit (Abitibi belt, Canada). Geofluids 2, 217-232.
Peacock et al., Glossary of fault and fault and other fracture networks Page 15
Geikie, A., 1882. Text-Book of Geology. MacMillan, London (cited by Oxford English Dictionary, 1989).
Gephart, J.W., 1990. Stress and the direction of slip on fault surfaces. Tectonics 9, 845-858.
Gibbs, A.D., 1984. Structural evolution of extensional basin margins. Journal of the Geological Society of London 141, 609-620.
Gibson, R.G., 1998. Physical character and fluid-flow properties of sandstone-derived fault zones. In: Coward, M.P., Daltaban,
T.S., Johnson, H. (eds) Structural Geology in Reservoir Characterization. Geological Society, London, Special
Publications, 127, 83-97.
Goguel, J., 1952. Traité de Tectonique. Masson, Paris. Translated by Thalmann, H. E. (1962), Tectonics. Freeman, San Francisco.
Gomez-Rivas, E., Bons, P.D., Koehn, D., Urai, J.L., Arndt, M., Virgo, S., Laurich, B., Zeeb, C., Stark, L., Blum, P., 2014. The
Jabal Akhdar Dome in the Oman mountains: Evolution of a dynamic fracture system. American Journal of Science 314,
1104-1139.
Grimm, K.A., Orange, D.L., 1997. Synsedimentary fracturing, fluid migration, and subaqueous mass wasting: intrastratal
microfractured zones in laminated diatomaceous sediments, Miocene Monterey Formation, California, U.S.A. Journal of
Sedimentary Research 67, 601-613.
Gudmundsson, A., 1995. Infrastructure and mechanics of volcanic systems in Iceland. Journal of Volcanology and Geothermal
Research 64, 1-22.
Hancock, P.L., 1985, Brittle microtectonics: principles and practice. Journal of Structural Geology 7, 437-457.
Harding, T.P., 1983. Divergent wrench fault and negative flower structure, Andamen Sea. In: Bally, A.W. (Ed.), Seismic
Expression of Structural Styles. Volume 3. American Association of Petroleum Geologists, Studies in Geology Series 15,
4.2-1 to 4.2-8.
Harding, T.P., 1985. Seismic characteristics and identification of negative flower structures, positive flower structures, and
positive structural inversion. Bulletin of the American Association of Petroleum Geologists 69, 582-600.
Harding, T.P., Lowell, J.D., 1979. Structural styles, their plate-tectonic habitats, and hydrocarbon traps in petroleum provinces.
Bulletin of the American Association of Petroleum Geologists 63, 1016-1058.
Healy, D., Blenkinsop, T.G., Timms, N.E., Meredith, P.G., Mitchell, T.M., Cooke, M.L., 2015. Polymodal faulting: time for a
new angle on shear failure. Journal of Structural Geology 80, 57-71.
Hempton, M.R., Neher, K., 1986. Experimental fracture, strain and subsidence patterns over en échelon strike-slip faults:
implications for the structural evolution of pull-apart basins. Journal of Structural Geology 8, 597-605.
Hills, E.S., 1940. Outlines of Structural Geology. Methuen, London.
Holland, P. (translator), 1601. Pliny’s Historie of the World, Commonly Called the Natural Historie (cited by Oxford English
Dictionary, 1989).
Hooker, J.N., Gale, J.F.W., L.A. Gomez, L.A., Laubach, S.E., Marrett, R., Reed, R.M., 2009. Aperture-size scaling variations in a
low-strain opening-mode fracture set, Cozzette Sandstone, Colorado. Journal of Structural Geology 31, 707-718.
Horsfield, W.T., 1980. Contemporaneous movement along crossing conjugate normal faults. Journal of Structural Geology 2, 305-
310.
Hunt, A., Ewing, R., Ghanbarian, B., 2014. Percolation Theory for Flow in Porous Media. Springer.
Hutchinson, W., 1794. The history of the county of Cumberland, and some places adjacent, from the earliest accounts to the
present time: comprehending the local history of the county; its antiquities, the origin, genealogy, and present state of the
principal families, with biographical notes; its mines, minerals, and plants, with other curiosities, either of nature or of art.
F. Jollie, Carlisle. (cited by Oxford English Dictionary, 1989).
Jackson, C.A.L., Rotevatn, A., 2013. 3D seismic analysis of the structure and evolution of a salt-influenced normal fault zone: a
test of competing fault growth models. Journal of Structural Geology 54, 215-234.
Jaeger, J.C., 1969. Elasticity, Fracture and Flow with Engineering and Geological Applications. 3rd editions. Chapman and Hall,
London.
Jamison, W.R., 1989. Fault-fracture strain in Wingate Sandstone. Journal of Structural Geology 11, 959-974.
Jébrak, M., 1997. Hydrothermal breccias in vein-type ore deposits: a review of mechanisms, morphology and size distribution.
Ore Geology Reviews 12, 111-134.
Jing, L., Stephansson, O., 1997. Network topology and homogenization of fractured rocks. In Jamtveit, B. Yardley, B.W. (Eds.),
Fluid Flow and Transport in Rocks. Mechanisms and Effects. Oxford: Chapman & Hall, 191–202.
Jolly, R.J.H., Sanderson, D.J., 1995. Variation in the form and distribution of dykes in the Mull swarm, Scotland. Journal of
Structural Geology 17, 1543-1557.
Jolly, R.J.H., Cosgrove, J.W., Dewhurst, D.N., 1998. Thickness and spatial distributions of clastic dykes, northwest Sacramento
Valley, California. Journal of Structural Geology 20, 1663-1672.
Karson, J.A., Rona, P.A., 1990. Block-tilting, transfer faults, and structural control of magmatic and hydrothermal processes in the
TAG area, Mid-Atlantic Ridge 26°N. Geological Society of America Bulletin 102, 1635-1645.
Kattenhorn, S.A., Watkeys, M.K., 1999. Blunt-ended dyke segments. Journal of Structural Geology 17, 1535.1542.

Peacock et al., Glossary of fault and fault and other fracture networks Page 16
Katz, D.A., Eberli, G.P., Swart, P.K., Smith, L.B., 2006. Tectonic-hydrothermal brecciation associated with calcite precipitation
and permeability destruction in Mississippian carbonate reservoirs, Montana and Wyoming. American Association of
Petroleum Geologists Bulletin 90, 1803-1841.
Kelly, P.G., Sanderson, D.J., Peacock, D.C.P., 1998. Linkage and evolution of conjugate strike-slip fault zones in limestones of
Somerset and Northumbria. Journal of Structural Geology 20, 1477-1493.
Kemeny, J., 2005. Time-dependent drift degradation due to the progressive failure of rock bridges along discontinuities.
International Journal of Rock Mechanics and Mining Sciences 42, 35-46.
Kenkmann, T., 2003. Dike formation, cataclastic flow, and rock fluidization during impact cratering: an example from the
Upheaval Dome structure, Utah. Earth and Planetary Science Letters 214, 43-58.
Kettermann, M., von Hagke, C., van Gent, H.W., Grützner, C., Urai, J.L., 2016. Dilatant normal faulting in jointed cohesive
rocks: a physical model study. Solid Earth 7, 843-856.
Kim, Y.S., Peacock, D.C.P., Sanderson, D.J., 2004. Fault damage zones. Journal of Structural Geology 26, 503-517.
Knipe, R. J., 1986. Faulting mechanisms in slope sediments: examples from Deep Sea Drilling Project cores. Geological Society
of America Memoirs, 166, 45-54.
Knipe, R.J., 1997. Juxtaposition and seal diagrams to help analyze fault seals in hydrocarbon reservoirs. American Association of
Petroleum Geologists Bulletin 81, 187-195.
Knott, S.D., 1994. Fault zone thickness versus displacement in the Permo-Triassic sandstones of NW England. Journal of the
Geological Society of London 151, 17-25.
Krantz, R.W., 1988. Multiple fault sets and three-dimensional strain: theory and application. Journal of Structural Geology 10,
225-237.
Kristensen, T.B., Rotevatn, A., Peacock, D.C.P., Ksienzyk, A.K., Henstra, G.A., Midtkandal, I., Grundvåg, S.A., Wemmer, K.,
accepted. The evolution and deformational style of a syn-rift border fault (East Greenland rift system): implications for
seal and leakage from fault-trapped syn-rift plays. Journal of Structural Geology.
Laing, W.P., 2004.Tension vein arrays in progressive strain: complex but predictable architecture, and major hosts of ore deposits.
Journal of Structural Geology 26, 1303-1315.
La Pointe, P.R., Hudson, J.A., 1985. Characterization and interpretation of rock mass joint patterns. Geological Society of
America Special Papers 199, 1-38.
Larsen, P-H., 1988. Relay structures in a Lower Permian basement-involved extension system, east Greenland. Journal of
Structural Geology 10, 3-8.
Lee, M.J., Wilkinson, J.J., 2002. Cementation, hydrothermal alteration, and Zn-Pb mineralization of carbonate breccias in the Irish
midlands: textural evidence from the Cooleen zone, near Silvermines, county Tipperary. Economic Geology 97, 653-662.
Letouzey, J., Trémolières, P., 1980. Paleo-stress fields around the Mediterranean since the Mesozoic from microtectonics.
Comparison with plate tectonic data. In: Scheidegger, A.E. (ed.), Tectonic Stresses in the Alpine-Mediterranean Region,
Volume 9 of the series Rock Mechanics / Felsmechanik / Mécanique des Roches, 173-192.
Linnen, R.L., Williams-Jones, A.E., 1987. Tectonic control of quartz vein orientations at the Trout Lake stockwork molybdenum
deposit, southeastern British Columbia: implications for metallogeny in the Kootenay arc (Canada). Economic Geology 82,
1283-1293.
Lister, L.A., Secrest, C.D., 1985. Giant desiccation cracks and differential surface subsidence, Red Lake Playa, Mohave County,
Arizona. Bulletin of the Association of Engineering Geologists 22, 299-314.
Lonergan, L., Cartwright, J., Jolly, R., 1998. The geometry of polygonal fault systems in Tertiary mudrocks of the North Sea.
Journal of Structural Geology 20, 529-548.
Maerten, L., Pollard, D.D., Karpuz, R., 2000. How to constrain 3-D fault continuity and linkage using reflection seismic data: a
geomechanical approach. American Association of Petroleum Geologists Bulletin 84, 1311-1324.
Mäkel, G.H., 2007. The modelling of fractured reservoirs: constraints and potential for fracture network geometry and hydraulics
analysis. In: Jolley, S.J., Barr, D., Walsh, J.J., Knipe, R.J., (eds.) Structurally Complex Reservoirs. Geological Society of
London, Special Publications 292, 375-403.
Manda, A.K., Horsman, E., 2015. Fracturesis Jointitis: causes, symptoms, and treatment in groundwater communities.
Groundwater 53, 836-840.
Mandl, G., Harkness, R.M., 1987. Hydrocarbon migration by hydraulic fracturing. In: Jones, M.E., Preston, R.M.F. (eds),
Deformation of Sediments and Sedimentary Rocks. Geological Society Special Publication 29, 39-53.
Mansfield, C.S., Cartwright, J.A., 1996. High resolution fault displacement mapping from three-dimensional seismic data;
evidence for dip linkage during fault growth. Journal of Structural Geology 18, 249-263.
Manzocchi, T., 2002. The connectivity of two-dimensional networks of spatially correlated fractures. Water Resources Research
38, 1162.
Martel, S.J., Langley, J.S., 2006. Propagation of normal faults to the surface in basalt, Koae fault system, Hawaii. Journal of
Structural Geology 28, 2123-2143.

Peacock et al., Glossary of fault and fault and other fracture networks Page 17
Marzo, M., Nijman, W., Puigdefabregas, C., 1988. Architecture of the Castissent fluvial sheet sandstones, Eocene, South
Pyrenees, Spain. Sedimentology 35, 719-738.
McClay, K.R., 1992. Glossary of thrust tectonics terms. In: McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, London,
419-433.
McGrath, A.G., Davison, I., 1995. Damage zone geometry around fault tips. Journal of Structural Geology 17, 1011-1024.
McKenzie, D.P., Morgan, W.J., 1969. Evolution of triple junctions. Nature 224, 125-133.
Menéndez, B., Zhu, W., Wong, T.F., 1996. Micromechanics of brittle faulting and cataclastic flow in Berea sandstone. Journal of
Structural Geology 18, 1-16.
Miller, M.G., 1999. Active breaching of a geometric segment boundary in the Sawatch Range normal fault, Colorado, USA.
Journal of Structural Geology 21, 769-776.
Mollema, P.N., Antonellini, M.A., 1996. Compaction bands: a structural analog for anti-mode I cracks in aeolian sandstone.
Tectonophysics 267, 209-228.
Morley, C.K., Nixon, C.W., 2016. Topological characteristics of simple and complex normal fault networks. Journal of Structural
Geology 84, 68-84.
Muraoka, H., Kamata, H., 1983. Displacement distribution along minor fault traces. Journal of Structural Geology 5, 483-495.
Naboko, S.I., 1963. Conditions of the present-day hydrothermal metamorphism of volcanic rocks. International Geology Review
5, 850-858.
Narr, W., Suppe, J., 1991. Joint spacing in sedimentary rocks. Journal of Structural Geology 13, 1037-1048.
Nevin, C.M., 1931. Principles of Structural Geology. Wiley, London.
Nicol, A., Watterson, J., Walsh, J.J., Childs, C., 1996. The shapes, major axis orientations and displacement patterns of fault
surfaces. Journal of Structural Geology 18, 235-248.
Nieto-Samaniego, Á.F., 1999. Stress, strain and fault patterns. Journal of Structural Geology 21, 1065-1070.
Nitecki, M.H., 1962. Observations on slickolites. Journal of Sedimentary Petrology 32, 435-439.
Nixon, C.W., Sanderson, D.J., Bull, J.M., 2011. Deformation within a strike-slip fault network at Westward Ho!, Devon U.K.:
domino vs conjugate faulting. Journal of Structural Geology 292, 309-336.
Nixon, C.W., Sanderson, D.J., Bull, J.M., 2012. Analysis of a strike-slip fault network using high resolution multibeam
bathymetry, offshore NW Devon U.K. Tectonophysics 541-543, 69-80.
Nixon, C.W., Sanderson, D.J., Dee, S., Bull, J.M., Humphreys, R., Swanson, M., 2014. Fault interactions and reactivation within a
normal fault network at Milne Point, Alaska. American Association of Petroleum Geologists Bulletin 98, 2081-2107.
Noroozi, M., Kakaie, R., Jalali, S.E., 2015. 3D geometrical-stochastical modeling of rock mass joint networks: case study of the
right bank of Rudbar Lorestan Dam plant. Journal of Geology and Mining Research 7, 1-10.
O'dea, M.G., Lister, G.S., 1995. The role of ductility contrast and basement architecture in the structural evolution of the Crystal
Creek block, Mount Isa Inlier, NW Queensland, Australia. Journal of Structural Geology 17, 949-960.
Odling, N.E., 1997. Scaling and connectivity of joint systems in sandstones from western Norway. Journal of Structural Geology,
19, 1257-1271.
Odling, N.E., Gillespie, P., Bourgine, B., Castaing, C., Chiles, J.P., Christensen, N.P., Watterson, J., 1999. Variations in fracture
system geometry and their implications for fluid flow in fractures hydrocarbon reservoirs. Petroleum Geoscience 5, 373-
384.
Odonne, F., Massonnat, G., 1992. Volume loss and deformation around conjugate fractures: comparison between a natural
example and analog experiments. Journal of Structural Geology 14, 963-972.
Okamura, A.T., Dvorak, J.J., Koyanagi, R.Y., Tanigawa, W.R., 1988. Surface deformation during dike propagation. United States
Geological Survey Professional Paper 1463, 165-181.
Okubo, C.H., Schultz, R.A. 2006. Near-tip stress rotation and the development of deformation band stepover geometries in mode
II. Geological Society of America Bulletin 118, 343-348.
Olsson, O., Falk, L., Forslund, O., Lundmark, L., Sandberg, E., 1992. Borehole radar applied to the characterization of
hydraulically conductive fracture zones in crystalline rock. Geophysical Prospecting 40, 109-142.
Ortega, O.J., Marrett, R.A., Laubach, S.E., 2006. A scale-independent approach to fracture intensity and average spacing
measurement. American Association of Petroleum Geologists Bulletin 90, 193-208.
Oxford English Dictionary, 1989. Second edition. Oxford University Press, Oxford.
OxfordDictionaries.com, 2015. Oxford University Press.
Pallister, J.S., Cashman, K.V., Hagstrum, J.T., Beeler, N.M., Moran, S.C., Denlinger, R.P., 2013. Faulting within the Mount St.
Helens conduit and implications for volcanic earthquakes. Bulletin of the Geological Society of America 125, 359-376.
Palsgrave, J., 1530. L'éclaircissement de la langue française (cited by Oxford English Dictionary, 1989).
Paquet, F., Dauteuil, O., Hallot, E., Moreau, F., 2007. Tectonics and magma dynamics coupling in a dyke swarm of Iceland.
Journal of Structural Geology 29, 1477-1493.
Peacock, D.C.P., 1991. Displacements and segment linkage in strike-slip fault zones. Journal of Structural Geology 13, 1025-
1035.
Peacock et al., Glossary of fault and fault and other fracture networks Page 18
Peacock, D.C.P., Anderson, M.W., 2012. The scaling of pull-aparts and implications for fluid flow in areas with strike-slip faults.
Journal of Petroleum Geology 35, 389-400.
Peacock, D.C.P., Parfitt, E.A., 2002. Active relay ramps and normal fault propagation on Kilauea Volcano, Hawaii. Journal of
Structural Geology 24, 729-742.
Peacock, D.C.P., Sanderson, D.J., 1991. Displacements, segment linkage and relay ramps in normal fault zones. Journal of
Structural Geology 13, 721-733.
Peacock, D.C.P., Sanderson, D.J., 1994. Geometry and development of relay ramps in normal fault systems. Bulletin of the
American Association of Petroleum Geologists 78, 147-165.
Peacock, D.C.P., Knipe, R.J., Sanderson, D.J., 2000. Glossary of normal faults. Journal of Structural Geology 22, 291-305.
Peacock, D.C.P., Nixon, C.W., Rotevatn, A., Sanderson, D.J., in review. Interacting faults. Journal of Structural Geology.
Philipp, S.L., 2008. Geometry and formation of gypsum veins in mudstones at Watchet, Somerset, SW England. Geological
Magazine 145, 831-844.
Pittman, E.D., 1981. Effect of fault-related granulation on porosity and permeability of quartz sandstones, Simpson Group
(Ordovician), Oklahoma. American Association of Petroleum Geologists Bulletin 65, 2381-2387.
Platten, I.M., 2000. Incremental dilation of magma filled fractures: evidence from dykes on the Isle of Skye, Scotland. Journal of
Structural Geology 22, 1153-1164.
Playfair, J., 1802. Illustrations of the Huttonian Theory of the Earth. Cadell and Davies, Edinburgh (cited by Oxford English
Dictionary, 1989).
Pollard, D.D., 1987. Elementary fracture mechanics applied to the structural interpretation of dykes. In: Halls, H.C., Fahrig, W.F.
(Eds.) Mafic Dyke Swarms. Geological Society of Canada Special Paper 34, 5-24.
Pollard, D.D., Aydin, A., 1984. Propagation and linkage of oceanic ridge segments. Journal of Geophysical Research 89, 10017-
10028.
Pollard, D.D., Aydin, A., 1988. Progress in understanding jointing over the past century. Geological Society of America Bulletin
100, 1181-1204.
Pollard, D.D., Segall, P., 1987. Theoretical displacements and stresses near fractures in rock: with applications to faults, joints,
veins, dikes, and solution surfaces. In: Atkinson, B.K. (Ed.), Fracture Mechanics of Rock. Academic Press, London, 277-
349.
Pollard, D.D., Delaney, P.T., Duffield, W.A., Endo, E.T., Okamura, A.T., 1983. Surface deformation in volcanic rift zones.
Tectonophysics 94, 541-584.
Price, N.J., 1966. Fault and Joint Development in Brittle and Semi-Brittle Rock. Pergamon, Oxford.
Price, N.J., Cosgrove, J.W., 1990. Analysis of Geological Structures. Cambridge University Press, Cambridge.
Priest, S.D., Hudson, J.A., 1976. Discontinuity spacing in rock. International Journal of Rock Mechanics and Mining Science, and
Geomechanics Abstracts 13, 135-148.
Priest, S.D., Hudson, J.A., 1981. Estimation of discontinuity spacing and trace length using scanline surveys. International Journal
of Rock Mechanics and Mining Science, and Geomechanics Abstracts 18, 183-197.
Qidong, D., Peizhen, Z., 1982. Research on the geometry of shear fracture zones. Journal of Geophysical Research 89, 5699-5710.
Questiaux, J.M., Couples, G.D., Ruby, N., 2010. Fractured reservoirs with fracture corridors. Geophysical Prospecting 58, 279-
295.
Rabinovitch, A., Bahat, D., Greenberg, R., 2012. Statistics of joint spacing in rock layers. Geological Magazine 149, 1065-1076.
Ramsay, J.G., 1967. The Folding and Fracturing of Rocks. New York, McGraw-Hill.
Ramsay, J.G., Huber, M.I., 1987. The Techniques of Modern Structural Geology. Volume 2: Folds and Fractures. Academic
Press, London.
Ramsey, J.M., Chester, F.M., 2004. Hybrid fracture and the transition from extension fracture to shear fracture. Nature 428, 63-66.
Rawnsley, K.D., Peacock, D.C.P., Rives, T., Petit, J.-P., 1998. Joints in the Mesozoic sediments around the Bristol Channel Basin.
Journal of Structural Geology 20, 1641-1661.
Reches, Z., 1978. Analysis of faulting in three-dimensional strain field. Tectonophysics 47, 109–129.
Reches, Z., Lockner, D.A., 1994. Nucleation and growth of faults in brittle rocks. Journal of Geophysical Research 99, 18159-
18173.
Reichardt, H., Weinberg, R.F., 2012. The dike swarm of the Karakoram shear zone, Ladakh, NW India: linking granite source to
batholith. Geological Society of America Bulletin 124, 89-103.
Reid, H.F., Davis, W.M., Lawson, A.C., Ransome, F.L., 1913. Report on the Committee on the Nomenclature of Faults.
Geological Society of America Bulletin 24, 163-186.
Reilly, C., Nicol, A., Walsh, J.J., Kroeger, K.F., 2016. Temporal changes of fault seal and early charge of the Maui gas-
condensate field, Taranaki Basin, New Zealand. Marine and Petroleum Geology 70, 237-250.
Reks, I.J., Gray, D.R., 1982. Pencil structure and strain in weakly deformed mudstone and siltstone. Journal of Structural Geology
4, 161-176.
Rispoli, R., 1981. Stress fields about strike-slip faults inferred from stylolites and tension gashes. Tectonophysics 75, T29-T36.
Peacock et al., Glossary of fault and fault and other fracture networks Page 19
Rives, T., Rawnsley, K.D., Petit, J.-P, 1994. Analogue simulation of natural orthogonal joint set formation in brittle varnish.
Journal of Structural Geology 16, 419-429.
Roberts, S., Sanderson, D.J., Gumiel, P., 1998. Fractal analysis of Sn-W mineralization from Central Iberia: insights into the role
of fracture connectivity in the formation of an ore deposit. Economic Geology 93, 360-365.
Robertson, E.C., 1960. Chapter 8: creep of Solenhofen Limestone under moderate hydrostatic pressure. Geological Society of
America Memoirs 79, 227-244.
Robinson, P., 1983. Connectivity of fracture systems-a percolation theory approach. Journal of Physics A: Mathematical and
General 16, 605-614.
Robinson, P., 1984. Numerical calculations of critical densities for lines and planes. Journal of Physics A: Mathematical and
General 17, 2823–2830.
Rotevatn, A., Fossen, H., 2012. Soft faults with hard tips: magnitude-order displacement gradient variations controlled by strain
softening versus hardening; implications for fault scaling. Journal of the Geological Society, London, 169, 123-126.
Rotevatn, A., Fossen, H., Hesthammer, J., Aas, T.E., Howell, J.A., 2007. Are relay ramps conduits for fluid flow? Structural
analysis of a relay ramp in Arches National Park, Utah. In: Lonergan, L., Jolly, R.J.H., Rawnsley, K. & Sanderson, D.J.
(eds) Fractured Reservoirs. Geological Society, London, Special Publications, 270, 55-71.
Rotevatn, A., Torabi, A., Fossen, H., Braathen, A., 2008. Slipped deformation bands: a new type of cataclastic deformation bands
in Western Sinai, Suez rift, Egypt. Journal of Structural Geology 30, 1317-1331.
Rotevatn, A., Thorsheim, E., Bastesen, E., Fossmark, H.S., Torabi, A., Sælen, G., 2016. Sequential growth of deformation bands
in carbonate grainstones in the hangingwall of an active growth fault: Implications for deformation mechanisms in different
tectonic regimes. Journal of Structural Geology, 90, 27-47.
Rothery, E., 1988. En échelon vein array development in extension and shear. Journal of Structural Geology 10, 63-71.
Rowe, C.D., Moore, J.C., Remitti, F., 2013. The thickness of subduction plate boundary faults from the seafloor into the
seismogenic zone. Geology 41, 991-994.
Safaricz, M., Davison, I., 2005. Pressure solution in chalk. American Association of Petroleum Geologists Bulletin 89, 383-401.
Sahimi, M., 1993. Flow phenomena in rocks: from continuum models to fractals, percolation, cellular automata, and simulated
annealing. Review of Modern Physics 65, 1393-1534.
Sahimi, M. 1994. Applications of Percolation Theory. Taylor & Francis, London, 258pp.
Sahimi, M., 2011. Flow and Transport in Porous Media and Fractured Rock: From Classical Methods to Modern Approaches.
John Wiley & Sons.
Samantray, A., Kraaijveld, M., Bulushi, W., Spring, L., 2010 Interpretation and application of borehole image logs in a new
generation of reservoir models for a cluster of fields in southern Oman. In: Poppelreiter, M., Garcia-Carballido, C.,
Kraaijveld, M. (Eds.), Dipmeter and Borehole Image Log Technology. American Association of Petroleum Geologists
Memoir 92, 343-357.
Sanderson, D.J., Nixon, C.W., 2015. The use of topology in fracture network characterization. Journal of Structural Geology 72,
55-66.
Sanderson, D.J., Nixon, C.W., in review. Topology, connectivity and percolation in fracture networks. Journal of Structural
Geology.
Saprykin, Y.P., 1967. Origin of hydrothermal mineralization in cassiterite-sulfide ore deposits of kavalerovo district. International
Geology Review 9, 660-668.
Schlische, R.W., Withjack, M.O., 2009. Origin of fault domains and fault-domain boundaries (transfer zones and accommodation
zones) in extensional provinces: result of random nucleation and self-organized fault growth. Journal of Structural Geology
31, 910-925.
Schofield, N., Heaton, L., Holford, S.P., Archer, S.G., Jackson, C.A.L., Jolley, D.W., 2012. Seismic imaging of ‘broken bridges’:
linking seismic to outcrop-scale investigations of intrusive magma lobes. Journal of the Geological Society, London 169,
421-426.
Scholz, C.H., Cowie, P.A., 1990. Determination of total strain from faulting using slip measurements. Nature 346, 837-839.
Schultz, R.A., Fossen, H., 2008. Terminology for structural discontinuities. American Association of Petroleum Geologists
Bulletin 92, 853-867.
Segall, P., Pollard, D.D., 1980. Mechanics of discontinuous faults. Journal of Geophysical Research 85, 4337-4350.
Segall, P., Pollard, D.D., 1983. Nucleation and growth of strike-slip faults in granite. Journal of Geophysical Research 88, 555-
568.
Sharples, W., Moresi, L.-N., Jadamec, M.A., Revote, J., in press. Styles of rifting and fault spacing in numerical models of crustal
extension. Journal of Geophysical Research B: Solid Earth.
Shephard-Thorn, E.R., Lake, R.D., Atitullah, E.A., 1972. Basement control of structures in the Mesozoic rocks in the Strait of
Dover region, and its reflexion in the certain features of the present land and submarine geology. Philosophical
Transactions of the Royal Society of London 272, 99-113.
Sibson, R.H., 1977. Fault rocks and fault mechanisms. Journal of the Geological Society of London 133, 191-213.
Peacock et al., Glossary of fault and fault and other fracture networks Page 20
Sibson, R.H., 1985. A note on fault reactivation. Journal of Structural Geology 7, 751-754.
Sibson, R.H., 1989. Earthquake faulting as a structural process. Journal of Structural Geology 11, 1-14.
Singh, Y., 1992. Study on the cave passages in Precambrian Bhander limestone of the Vindhyan Supergroup, India. Zeitschrift fur
Geomorphologie, Supplementband 85, 115-126.
Sinha-Roy, S., 2002. Kinetics of differentiated stylolite formation. Current Science 82, 1038-1046.
Smith, J.V., 1997. Initiation of convergent extension fracture vein arrays by displacement of discontinuous fault segments. Journal
of Structural Geology 19, 1369-1373.
Speed, R.C., Cogbill, A.H., 1979. Candelaria and other left-oblique slip faults of the Candelaria region, Nevada. Bulletin of the
Geological Society of America 90, 149-163.
Spotila, J.A., Anderson, K.B., 2004. Fault interaction at the junction of the Transverse Ranges and Eastern California shear zone:
a case study of intersecting faults. Tectonophysics 379, 43-60.
Stanley, H.E., 1977. Cluster shapes at the percolation threshold: and effective cluster dimensionality and its connection with
critical-point exponents. Journal of Physics A: Mathematical and General 10, L211–L220.
Stauffer, D., Aharony, A., 1991. Introduction to Percolation Theory. Taylor & Francis, London, 181pp.
Stel, H., 1991. Linear dilatation structures and syn-magmatic folding in granitoids. Journal of Structural Geology 13, 625-634.
Sternlof, K.R., Chapin, J.R., Pollard, D.D., Durlofsky, L.J., 2004. Permeability effects of deformation band arrays in sandstone.
American Association of Petroleum Geologists Bulletin 88, 1315-1329.
Stewart, I.S., Hancock, P.L., 1991. Scales of structural heterogeneity within neotectonic normal fault zones in the Aegean region.
Journal of Structural Geology 13, 191-204.
Stočes, B., White, C. H., 1935. Structural Geology with Special Reference to Economic Deposits. McMillan, London.
Stockdale, P.B., 1926. The stratigraphic significance of solution in rocks. Journal of Geology 34, 399-414.
Stowell, J.F.W., Watson, A.P., Hudson, N.F.P. 1999. Geometry and population systematics of a quartz vein set, Holy Island,
Angelsey, North Wales. In: McCaffrey, K.J.W., Lonergan, L., Wilkinson, J.J. (eds.) Fractures, Fluid Flow and
Mineralization. Geological Society of London, Special Publication 155, 17-33.
Strachan. I., Temple, J., Williams, A., 1948. The age of the neptunian dyke at Hazler Hill. Geological Magazine 85, 276-278.
Sullivan, R.J., 1794. A View of Nature, in Letters to a Traveller Among the Alps: With Reflections on Atheistical Philosophy,
Now Exemplified in France. T. Becket, London (cited by Oxford English Dictionary, 1989).
Sutherland, R., Toy, V.G., Townend, J., Cox, S.C., Eccles, J.D., Faulkner, D.R., Prior, D.J., Norris, R.J., Mariani, E., Boulton, C.,
Carpenter, B.M., Menzies, C.D., Little, T.A., Hasting, M., De Pascale, G.P., Langridge, R.M., Scott, H.R., Reid Lindroos,
Z., Fleming, B., Kopf, A.J., 2012. Drilling reveals fluid control on architecture and rupture of the Alpine fault, New
Zealand. Geology 40, 1143-1146.
Swanson, M.T., 1990. Extensional duplexing in the York Cliffs strike-slip fault system, southern coastal Maine. Journal of
Structural Geology 12, 499-512.
Swanson, M.T., 2005. Geometry and kinematics of adhesive wear in brittle strike-slip fault zones. Journal of Structural Geology
27, 871-887.
Thienprasert, A., Raksaskulwong, M., Surinkum, A., 1991. Exploration of San Kamphaeng Geothermal Field, Northern Thailand.
Journal of the Geothermal Research Society of Japan Vol.13, No.1(1991) P.15-P.30
Thomas, C.W., Woodcock, N.H., 2015. The kinematic linkage of the Dent, Craven and related faults of Northern England.
Proceedings of the Yorkshire Geological Society 60, 258-274.
Trevisa, J. de, 1387. Polychronicon Ranulphi Higden (cited by Oxford English Dictionary, 1989).
Trudgill, B., Cartwright, J., 1994. Relay-ramp forms and normal-fault linkages, Canyonlands National Park, Utah. Geological
Society of America Bulletin, 106, 1143-1157.
Tsang, Y.W., Witherspoon, P.A., 1983. The dependence of fracture mechanical and fluid flow properties on fracture roughness
and sample size. Journal of Geophysical Research 88, 2359-2366.
Tvedt, A.B.M., Rotevatn, A., Jackson, C.A.-L., Fossen, H., Gawthorpe, R.L., 2013. Growth of normal faults in multilayer
sequences: A 3D seismic case study from the Egersund Basin, Norwegian North Sea. Journal of Structural Geology 55, 1-
20.
Twiss, R.J., Moores, E.M., 1992. Structural Geology. Freeman, New York.
Uglow, W.L., 1913. Hydrothermal alteration. Economic Geology 8, 797-800.
Van Noten, K. , Sintubin, M., 2010. Linear to non-linear relationship between vein spacing and layer thickness in centimetre- to
decimetre-scale siliciclastic multilayers from the High-Ardenne slate belt (Belgium, Germany). Journal of Structural
Geology 32, 377-391.
Valentine, G.A., van Wyk de Vries, B., 2014. Unconventional maar diatreme and associated intrusions in the soft sediment-hosted
Mardoux structure (Gergovie, France). Bulletin of Volcanology 76, 1-16.
Vermilye, J.M., Scholz, C.H., 1995. Relation between vein length and aperture. Journal of Structural Geology 17, 423-434.
Versey, H.C., 1927. Post-Carboniferous movements in the Northumbrian fault block. Proceedings of the Yorkshire Geological
Society 21, 1-16.
Peacock et al., Glossary of fault and fault and other fracture networks Page 21
Virgo, S., Abe, S., Urai, J.L., 2013. Extension fracture propagation in rocks with veins: insight into the crack-seal process using
discrete element method modelling. Journal of Geophysical Research Solid Earth 118, 5236-5251.
Virgo, S., Abe, S., Urai, J.L., 2014. The evolution of crack seal vein and fracture networks in an evolving stress field: insights
from discrete element models of fracture sealing. Journal of Geophysical Research Solid Earth 119, 8708-8727.
Walsh, J.J., Watterson, J., 1987. Distributions of cumulative displacement and seismic slip on a single normal fault. Journal of
Structural Geology 9, 1039-1046.
Walsh, J.J., Watterson, J., 1988, Analysis of the relationship between displacements and dimensions of faults. Journal of
Structural Geology 10, 239-247.
Walsh, J.J., Watterson, J., 1991. Geometric and kinematic coherence and scale effects in normal fault systems. In: Roberts, A.M.,
Yielding, G., Freeman. B. (Eds.), The Geometry of Normal Faults. Geological Society Special Publication 56, 193-203.
Webb, L.E., Johnson, C.L., 2006. Tertiary strike-slip faulting in southeastern Mongolia and implications for Asian tectonics. Earth
and Planetary Science Letters 241, 323-335.
Weinberger, R., Lyakhovsky, V., Baer, G., Agnon, A., 2000. Damage zones around en echelon dike segments in porous
sandstone. Journal of Geophysical Research: Solid Earth 105, 3115-3133.
Wennberg, O.P., Rennan, L., Basquet, R., 2009. Computed tomography scan imaging of natural open fractures in a porous rock:
geometry and fluid flow. Geophysical Prospecting 57, 239-249.
Wennberg, O.P., Casini, G., Jonoud, S., Peacock, D.C.P., 2016. The characteristics of open fractures in carbonate reservoirs and
their impact on fluid flow: a discussion. Petroleum Geoscience 22, 91-104.
Wernicke, B., Burchfiel, B.C., 1982. Modes of extensional tectonics. Journal of Structural Geology 4, 105-115.
Wheeler, R.L., Dixon, J.M., 1980. Intensity of systematic joints: methods and application. Geology 8, 230-233.
Whitaker, A.E., Engelder, T., 2005. Characterizing stress fields in the upper crust using joint orientation distributions. Journal of
Structural Geology 27, 1778-1787.
White, J.D.L., Ross, P.-S., 2011. Maar-diatreme volcanoes: a review. Journal of Volcanology and Geothermal Research 201, 1-29.
Wibberley, C. A., Petit, J. P., Rives, T., 2000. Mechanics of cataclastic ‘deformation band’ faulting in high-porosity sandstone,
Provence. Comptes Rendus de l'Académie des Sciences-Series IIA-Earth and Planetary Science 331, 419-425.
Wilkins, S.J., Gross, M.R., Wacker, M., Eyal, Y., Engelder, T., 2001. Faulted joints: Kinematics, displacement-length scaling
relations and criteria for their identification. Journal of Structural Geology 23, 315-327.
Willis, R., 1935. Development of thrust faults. Bulletin of the Geological Society of America 46, 409-424.
Woodcock, N.H., Fischer, M., 1986. Strike-slip duplexes. Journal of Structural Geology 8, 725-735.
Woodcock, N.H., Mort, K., 2008. Classification of fault breccias and related fault rocks. Geological Magazine 145, 435-440.
Woodcock, N.H., Underhill, J.R., 1987. Emplacement-related fault patterns around the Northern Granite, Arran, Scotland.
Geological Society of America Bulletin 98, 515-527.
Wu, H., Pollard, D.D., 2002. Imaging 3-D fracture networks around boreholes. American Association of Petroleum Geologists
Bulletin 86, 593-604.
Wu, S., Groshong, R.H., 1991. Low-temperature deformation of sandstone, southern Appalachian fold-thrust belt. Bulletin of the
Geological Society of America 103, 861-875.
Yan, M., Elmo, D., Stead, D., 2007. Characterization of step-path failure mechanisms: a combined field based- numerical
modeling study. American Rock Mechanics Association, 1st Canada - U.S. Rock Mechanics Symposium, 27-31 May,
Vancouver, Canada.
Yielding, G., Freeman, B., Needham, D.T., 1997. Quantitative fault seal prediction. American Association of Petroleum
Geologists Bulletin 81, 897-917.
Žák, J., Verner, K., Johnson, K., Schwartz, J.J, 2012. Magma emplacement process zone preserved in the roof of a large
Cordilleran batholith, Wallowa Mountains, northeastern Oregon. Journal of Volcanology and Geothermal Research 227-
228, 61-75.
Zhang, X., Sanderson, D.J., 1996. Numerical modelling of the effects of fault slip on fluid flow around extensional faults. Journal
of Structural Geology 18, 109-119.
Zhang, X., Bunger, A.P., Jeffrey, R.G., 2014. Mechanics of two interacting magma-driven fractures: a numerical study. Journal of
Geophysical Research B: Solid Earth 119, 8047-8063.
Zhao, C., Hobbs, B.E., Ord, A., Robert, P.A., Hornby, P., Peng, S., 2007. Phenomenological modelling of crack generation in
brittle crustal rocks using the particle simulation method. Journal of Structural Geology 29, 1034-1048.
Zu, K., Zeng, L., Gong, L., 2013. Fractures domains in conceptual models of fault-related folds. Scientia Geologica Sinica 48,
1140-1147.

Peacock et al., Glossary of fault and fault and other fracture networks Page 22
Figures

Fig. 1. Schematic illustration of the range of fault and other fracture networks defined in this paper, emphasising our preferred
terms. (a) Individual fractures. (b) Interaction between pairs of fractures. (c) Fracture networks.

Peacock et al., Glossary of fault and fault and other fracture networks Page 23
Fig. 2. (a) Different types of intersections between fractures and (b) related topological terms. Green circles represent I-nodes, red
triangles represent the connections between two fractures (i.e., between three branches, topologically), while blue squares
represent the connections between crossing fractures (i.e., between four branches, topologically).

Fig. 3. Diagram representing different types of interaction between fractures.

Peacock et al., Glossary of fault and fault and other fracture networks Page 24
Fig. 4. The node and branch model used by Sanderson and Nixon (2015) to characterise fracture networks with increasing
interaction as they pass through the percolation threshold (Pc) to form a connected network.

Fig. 5. Different types of damage zones, including our terms approaching damage zone and intersection damage zone. Modified
from Kim et al. (2004).

Peacock et al., Glossary of fault and fault and other fracture networks Page 25
Geological Geometric Topological Kinematic Mechanical

Alteration zone Aperture Branch Dip-slip Anticrack


Anticrack Bed-bound fault or fracture I node Extension(al) fault Bed-bound fault or fracture
Breccia Branch Isolated branch Extension(al) fracture Brittle fault or fracture
Cataclasis Isolated fault or fracture Node Fault Cataclasis
Clastic dyke Layer-bound fault or fracture Vertex Hybrid fracture Discontinuity
Closed fracture Length Y node Isolated fault or fracture Fault
Conductive fault or fracture Lens X node Normal fault Fracture
Crack Roughness, fault or fracture Oblique-slip fault Hybrid fracture
Deformation band Segment Reactivation Hydrofracture
Desiccation crack Shear joint Reverse fault Induced fracture
Diatreme Slip vector Shear joint Layer-bound fault or fracture
Discontinuity Surface, fault or fracture Strike-slip fault Microcrack
Dyke (or dike) Tip line Tensile fracture Mode I
Fault Tip point Tension(al) fracture Mode II
Fissure Trace Thrust fault Mode III
Fracture Trace length Transform fault Opening-mode
Individual fractures and associated structures

Granulation seam Propagation


Hydrothermal Shear fracture
Injectite Tensile fracture
Joint Tension(al) fracture
Lineament
Microcrack
Mud crack
Natural fracture
Neptunian dyke
Non-conductive fracture
Normal fault
Oblique-slip fault
Open fracture
Partly-open fracture
Reverse fault
Seal
Sedimentary dyke
Sill
Slickenside
Slickolite
Slip plane
Slip surface
Strand
Strike-slip fault
Stylolite
Tension gash
Thrust fault
Transform fault
Vein

Peacock et al., Glossary of fault and fault and other fracture networks Page 26
Horse Abutting fault or fracture Branch Antithetic fault Fault linkage
Triple junction Antithetic fault Connecting node Bridge, fault or fracture Interaction, fault or fracture
Approaching fault or fracture Dangling end Broken bridge Interaction, mechanical
Branch Doubly-connected branch Coherence, kinematic Linkage, fault or fracture
Bi-modal Singly-connected branch Conjugate
Branch line Connecting fault or fracture
Branch point Coupling
Bridge, fault or fracture Coupling, kinematic
Broken bridge Dip-linkage
Coherence, geometric Fault linkage
Interaction between two (or more) fractures

Conjugate Hard-linkage
Connecting fault or fracture Interaction, fault or fracture
Coupling, geometric Intersection, fault or fracture
Crossing fault or fracture Kinematic coupling
Dip-linkage Kinematic linkage
Geometric coherence Linkage, fault or fracture
Geometric coupling Pull-apart
Geometric linkage Relay ramp
Horse Relay zone/structure
Hard-linkage Soft-linkage
Intersection line Strike-linkage
Intersection point Synthetic fault
Oblique Trailing
Orthogonal Transfer fault
Overstep, fault or fracture Wing crack
Pull-apart
Relay ramp
Relay zone/structure
Splay fault
Step, fault or fracture
Synthetic fault
Trailing

Peacock et al., Glossary of fault and fault and other fracture networks Page 27
Architecture Anastomosing Arrangement, network Approaching damage zone Approaching damage zone
Array, fault or fracture Block Backbone Array, fault or fracture Damage zone
Domain Bi-modal Cluster Coherence, kinematic Interaction damage zone
Duplex Block Connectivity Damage zone Intersection damage zone
Imbricate Coherence, geometric Critical connectivity Domain Linking damage zone
Stockwork Compartment Dimensionless intensity Domino faults Mechanical stratigraphy
Compartmentalisation Finite cluster Horsetail Process zone
Connectivity Infinite cluster Interaction damage zone Saturation, joint
Corridor, fracture Isolated cluster Intersection damage zone Tip damage zone
Density, fault or fracture Percolation Mechanical stratigraphy Wall damage zone
Domain Percolation threshold Linking damage zone
Domino faults Spanning cluster Pinnate
Fracture networks and populations

Duplex Polymodal
En echelon Reactivation
Flower structure Tip damage zone
Frequency, fault or fracture Wall damage zone
Horsetail Wing crack
Imbricate
Intensity, fault or fracture
Intersection damage zone
Network, fault or fracture
Organisation, network
Percolation
Percolation threshold
Pinnate
Polygonal faults
Polymodal
Population
Relay pattern
Set, fault or fracture
Spacing, fault or fracture
Swarm, fault or fracture
Zone, fault or fracture

Table 1. Lists of geological, geometric terms, topological, kinematic and mechanical terms.

Peacock et al., Glossary of fault and fault and other fracture networks Page 28

View publication stats

You might also like