You are on page 1of 13

Journal of Structural Geology 106 (2018) 41–53

Contents lists available at ScienceDirect

Journal of Structural Geology


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

Relationships between fractures T


a,∗ b a
D.C.P. Peacock , D.J. Sanderson , A. Rotevatn
a
Department of Earth Science, University of Bergen, Allégaten 41, 5007 Bergen, Norway
b
Engineering and the Environment, University of Southampton, Highfield, Southampton, SO17 1BJ, UK

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

Keywords: Fracture systems comprise many fractures that may be grouped into sets based on their orientation, type and
Fractures relative age. The fractures are often arranged in a network that involves fracture branches that interact with one
Interaction another. Interacting fractures are termed geometrically coupled when they share an intersection line and/or ki-
Geometry nematically coupled when the displacements, stresses and strains of one fracture influences those of the other.
Kinematics
Fracture interactions are characterised in terms of the following. 1) Fracture type: for example, whether they
Chronology
have opening (e.g., joints, veins, dykes), closing (stylolites, compaction bands), shearing (e.g., faults, de-
formation bands) or mixed-mode displacements. 2) Geometry (e.g., relative orientations) and topology (the
arrangement of the fractures, including their connectivity). 3) Chronology: the relative ages of the fractures. 4)
Kinematics: the displacement distributions of the interacting fractures. It is also suggested that interaction can be
characterised in terms of mechanics, e.g., the effects of the interaction on the stress field. It is insufficient to
describe only the components of a fracture network, with fuller understanding coming from determining the
interactions between the different components of the network.

1. Introduction for stress concentration and perturbation (e.g., Crider and Pollard,
1998; Kattenhorn et al., 2000; Maerten, 2000; Bourne and Willemse,
Fractures are generally common within a rock volume, and we refer 2001), and therefore for structural development (e.g., Kim et al., 2004;
to the fractures that occur as the fracture system (Fig. 1). The individual Choi et al., 2016), fluid flow (e.g., Knipe, 1992; Zhang and Sanderson,
fractures may have similar properties (fracture type, mineral fill, etc.) 2002) and mineralisation (e.g., Curewitz and Karson, 1997; Cox, 2005;
and a specific range of orientations, in which case it is convenient to Johansen et al., 2005; Xiao-shuang et al., 2005). In spite of their im-
group then into fracture sets (Fig. 1a). The spatial arrangement of the portance, no thorough characterisation and classification scheme for
fractures forms a fracture network, in which the position, orientation fracture interactions appears to have been published. Peacock et al.
and relationships between the individual fractures is mapped in 2D or (2017) provide a characterisation scheme for interacting faults, but not
3D space, where they intersect at intersection points (nodes) (Fig. 1b) or for other types of fractures. Here, we aim to generalise the character-
intersection lines. Such networks can be relatively simple, being made up isation scheme of Peacock et al. (2017) to classify interaction between
of a single set of fractures (Fig. 1c), or they can be made up of many sets all types of fractures, giving a precise methodology for describing the
with a wide range of orientations (Fig. 1d). Fracture sets may develop as different ways in which fractures of any type, orientation and relative
part of a single deformation event, in which case they are likely to have age can interact.
interrelated kinematics and mechanics, or during a sequence of events We focus on describing field examples, many from the Mesozoic
under different boundary conditions. It is common for a set of fractures sedimentary rocks of Somerset, UK, where a history of different tectonic
to be active during different events, with a set established during an regimes provides a rich array of fracture interactions (e.g., Dart et al.,
earlier event being reactivated during a later event. To understand such 1995; Peacock and Sanderson, 1999; Glen et al., 2005; Peacock et al.,
systems, and to evaluate the resulting rock-mass properties, involves 2017). Examples from the literature are also used (e.g., Table 1). In this
both an analysis of the individual fractures and of the network as a paper, we emphasise interactions between different fracture types, thus
whole. In this paper, we discuss the interactions between pairs of building on and generalising the study fault interactions presented by
fractures in terms of their geometry and topology, relative age and ki- Peacock et al. (2017) to include interactions between all fracture types.
nematics. The aim is to present a scheme for the characterisation of the re-
Areas of fracture interaction and intersection can be important sites lationships between any two fractures, of any type, based on their


Corresponding author.
E-mail address: hermangedge@gmail.com (D.C.P. Peacock).

https://doi.org/10.1016/j.jsg.2017.11.010
Received 29 March 2017; Received in revised form 12 September 2017; Accepted 19 November 2017
Available online 21 November 2017
0191-8141/ © 2017 Elsevier Ltd. All rights reserved.
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 1. Fracture networks and interactions. (a) Fractures (1… n) can be divided into different sets (e.g., based on fracture types, orientations, mineral fills, etc.), which may have been
formed in a series of different events. Together, these fractures, sets and events constitute a fracture network. (b) Fractures meet at nodes (intersection points in 2D or intersection lines in
3D), between which the fractures can be divided into branches. (c) Calcites on a bedding plane of Liassic limestone (Kilve, Somerset). A single stress field produced a set of sub-parallel
veins. (d) Network of calcite veins in a limestone bed (Lilstock, Somerset). Note that the block was not in situ. The vein network formed by the overprinting/superposition of two or more
stress fields, producing interactions between veins.

Table 1
Examples of the different types of fracture interactions described in this paper, including possible large examples. In this context, “small” means exposure-scale examples of the type
illustrated (e.g., Figs. 3–9), while “large” means examples tens of metres across or larger, e.g., observable on seismic data.

Relationship Figure in this manuscript Other small examples Possible large examples

Opening mode fractures 3a Hendry et al. (2015, Fig. 8) Ernst et al. (2001, Fig. 1)
Shearing mode fractures 3b Misra et al. (2009, Fig. 12) Gourmelen et al. (2011, Fig. 1)
Compactional structures 3a Katsman (2010, Fig. 1) Lin et al. (2009, Fig. 1)
Hybrid structures 3c Belayneh and Cosgrove (2010, Fig. 5) Morley et al. (2009, Fig. 15)
Isolated fractures 4a Cobbold et al. (2013, Fig. 3c) Walsh and Watterson (1987)
Approaching fractures 4b Omosanya et al. (2013, Fig. 9); Morley (2014, Fig. 12) Gökten et al. (2011, Fig. 1)
Abutting fractures 4c Mughieda and Alzo'ubi (2004, Fig. 5) Khajavi et al. (2014, Fig. 1)
Branching 4d Bons and Montenari (2005, Fig. 5c) Marchal et al. (2003, Fig. 5)
Cross-cutting fractures 4b Lecumberri-Sanchez et al. (2013, Fig. 4) Bogdanoff et al. (2000, Fig. 2)
Synchronous fractures 5a, 6a Katz et al. (2006, Fig. 5) Mathieu et al. (2008, Fig. 9)
Different ages 5c, 6b Pati et al. (2007, Fig. 4) Singh and Slabunov (2014, Fig. 2)
Trailing fractures 6c Wise (2010, Fig. 8) Nixon et al. (2014, Fig. 15)
Reactivation 6d Dabi et al. (2011, Fig. 3) Busquets et al. (2005, Fig. 3)
Branch line parallel to displacement 7a Tewksbury et al. (2014, Fig. 2) Yin and Groshong (2007, Fig. 3)
Branch line normal to displacement 7c Arndt et al. (2014, Fig. 4) Taylor and Peltzer (2006, Fig. 12)
Different displacement directions 7e Micarelli and Benedicto (2008, Fig. 2) Dini et al. (2008, Fig. 13)
Antithetic relationship 8a Choi et al. (2016, Fig. 5) Kim and Sanderson (2006, Fig. 14)
Synthetic relationship 8b Morley (2014, Fig. 20) Kim and Sanderson (2006, Fig. 8)
Neutral intersection 8c, 8d Peacock and Sanderson (1994, Fig. 7) Bonson et al. (2007, Fig. 11)
Parallel displacements 8e Petit et al. (1999, Fig. 2b) Fort and Brun (2012, Fig. 16a)
Perpendicular displacements 8f Zulauf et al. (2011, Fig. 5) Trautwein-Bruns et al. (2010, Fig. 1)
Strains apparently unrelated 9a Peacock (2001, Fig. 8) Macdonald et al. (2015, Fig. 1)
Strains apparently related 9b Petit et al. (1999, Fig. 2a) Nicoll (2010, Fig. 3)
Non-uniaxial strain 9c Weinberger (1999, Fig. 1) Tuckwell et al. (2003, Fig. 1)

geometry and topology, chronology and kinematics, thereby providing 2. Fracture types
an observational framework for the analysis of interactions within
fracture networks. Detailed analysis of the mechanics of fracture in- The first step in characterising fractures in rock is to determine what
teraction is beyond the scope of this paper, but a brief discussion is type of fractures are being analysed. The term fracture is commonly
given of mechanical implications. used in the geological literature to cover a range of structures, including

42
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 2. (a) Fractures can be divided on the basis of the displacement


directions relative to the fracture planes. (b) When the fractures are
discrete surfaces, the shearing fractures include slip-planes, the
opening fractures include joints, and the closing structures include
stylolites. (c) When the fractures are bands or zones, the shearing
fractures include fault zones, the opening fractures include vein
systems, and the closing structures include compaction bands.

faults, joints, veins, dykes and sills, deformation bands, stylolites, etc., include deformation bands, which are mm-scale tabular zones of lo-
as reviewed by Schultz and Fossen (2008) and Peacock et al. (2016). calised but non-discrete strain in porous, granular rocks/sediments
Many fracture networks comprise a range of fracture types, and it is (e.g., Friedman and Logan, 1973; Engelder, 1974), typically invol-
important in many areas of geology to recognise that different types of ving grain reorganisation, cataclasis, dissolution and/or precipita-
fractures will contribute differently to the behaviour of the rock-mass tion (e.g., Aydin, 1978, Knott, 1994).
(e.g., Manda and Horsman, 2015; Peacock et al., 2016). For example, a • Combined fractures: where there are measurable components of two
network of open joints or high permeability sandstone dykes may en- of the above modes (Fig. 3c). Such hybrid or mixed-mode fractures
hance fluid flow, whereas a network of veins or igneous dykes tend to include those developed by synchronous opening and shear modes
be barriers to fluid flow. (e.g., Ramsey and Chester, 2004) and slickolites, where insoluble
Here we group fracture types according to their dominant mode of residue created by pressure solution develops along fault planes
wall rock movement. This extends the work of Pollard and Aydin (e.g., Nitecki, 1962). Compactional shear bands, which record a
(1988), who distinguish joints (generally opening mode) and faults combination of closing- and shearing-modes, are also included (e.g.,
(shearing modes), by including closing-mode and hybrid fractures Aydin et al., 2006; Fossen et al., 2007).
(Fig. 2a).
We apply these terms to both discrete surfaces (Fig. 2b) and narrow
• Extension fractures: where the fracture walls move apart (Fig. 3a), bands or zones of deformation (Fig. 2c).
i.e., opening-mode. These include: joints, unfilled dominantly Many fractures show evidence for combination of two (or more)
opening-mode fractures (e.g., Pollard and Aydin, 1988); veins, i.e., processes that may either operate synchronously (such as shear frac-
mineral-filled fractures (e.g., Bons et al., 2012); dykes and sills, i.e., turing and pressure solution forming a slickolite) or sequentially. For
fractures filled with igneous rocks (e.g., de Sitter, 1964; Paquet example, a joint may be mineralised to become a vein, and then un-
et al., 2007) or remobilised sediments (e.g., Jolly et al., 1998); di- dergo pressure solution to form a stylolite (e.g., Srivastava and
lation bands, i.e., a deformation band recording localised dilation or Engelder, 1990). Similarly, joints may be reactivated to accommodate
porosity increase, typically in highly porous, poorly-consolidated shear displacement, thus forming a fault (e.g., Moore and Schultz,
sand (Du Bernard et al., 2002). 1999).
• Contractional fractures: where the fracture walls converge (Fig. 3a),
i.e., closing-mode. They include stylolites, which are surfaces on 3. Geometric and topological relationships between interacting
which insoluble residues are concentrated as pressure solution re- fractures
moves such soluble minerals as calcite or quartz (e.g., Stockdale,
1926; Fletcher and Pollard, 1981). They can also include compaction Any 2D fracture network can be considered as a system of nodes and
bands (Mollema and Antonellini, 1996; Fossen et al., 2011; Rotevatn branches (e.g., Sanderson and Nixon, 2015), where the branches re-
et al., 2016), which are deformation bands that exhibit closing-mode present the fractures and the nodes the tips or intersection points of the
displacements. fractures (Fig. 1b). Manzocchi (2002) and Sanderson and Nixon (2015)
• Shearing-mode fractures: a general term for a discontinuity along show that natural fracture networks are almost exclusively made up of
which the wall rocks move sub-parallel to the plane of the fracture three types of nodes: an I-node is a tip with a single branch, a Y-node
(Fig. 3b). These include faults and fault zones, defined by Price has three branches, and an X-node has four branches. In this paper, we
(1966) as “a plane of fracture which exhibits signs of differential propose that (1) the interactions between fractures occur at or near the
movement of the rock mass on either side of the plane”. They also nodes, and (2) that the style of the interaction between fractures can be

43
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 3. Photographs of different fracture types, from Somerset, UK, showing a classification based on the displacement direction relative to the fracture plane. (a). Extension fractures
(calcite veins), with the extension direction almost perpendicular to the vein walls, and contractional structures (stylolites) within a zone of en echelon calcite veins (Lilstock). Contraction
is approximately perpendicular to the stylolite planes. (b) Normal faults interacting across relay ramps and cut by joints (Kilve). (c) Combined fractures (pull-aparts linked by slickolites),
with net displacement neither in the plane of the fracture or perpendicular to the fracture (Lilstock). (d) Example of intersections between different fracture types. Gypsum vein abutting
another gypsum vein, which itself abuts a sandstone dyke (Watchet).

understood from the types of fractures that interact. later fracture being arrested at the earlier fracture. Fig. 5(c) shows
As a first step in the analysis of interacting fractures, we analyse the joints that curve to abut another joint at ∼ 90°.
geometric and topological relationship between adjacent fractures. We • Branching or splaying: this also produces Y-nodes, but is where the
recognise the following relationships between any two interacting branches can be demonstrated to have been active at the same time
fractures: (Fig. 4d). Where the branches develop by bifurcation of a propa-
gating fracture (as in many opening- and shearing-mode fractures),
• Isolated: where the fractures are not geometrically (physically) con- two of the branches generally form at a low angle (< 30°) to each
nected to each other (at least in the plane in which they are being another.
observed), nor do they kinematically interact (Fig. 4a). Such fractures • Cross-cutting: where a later fracture cuts another fracture to form an
would have tips in the rock matrix (i.e., I-nodes; Sanderson and X-node (Fig. 4b). Where two fracture sets mutually cross-cut each
Nixon, 2015). other, they may have developed synchronously (e.g., Nicol et al.,
• Approaching: where two fracture tips are close enough to each other 1995, Fig. 5; Ferrill et al., 2009; Morley, 2014, Fig. 20). Fig. 5(d)
to kinematically interact, either through their tip damage zones appears to show an intersection point between two mutually-
overlapping (i.e., linking damage zones of Kim et al., 2004, or the crossing joints, but it is possible that this example actually consists
approaching damage zone of Peacock et al., 2017) or the displace- of three joints, comprising an older joint against which two other
ments interacting (Fig. 4b). Stepping normal faults that overlap joints abut at the same location. Filled extension fractures (e.g.,
without touching each other form relay ramps (e.g., Peacock and veins and dykes) can cross-cut each other if they do not represent
Sanderson, 1991), while bridges occur between stepping extension mechanical barriers (e.g., Polito et al., 2001, Fig. 9).
fractures (e.g., Lisle, 2013). The tips may or may not be connected
by smaller fractures (damage) of the same or differing type. Ap- Abutting and branching relationships are topologically similar, both
proaching fractures therefore have I-nodes that are close enough to producing Y-nodes, and can be impossible to distinguish where the
be kinematically linked but not geometrically linked (i.e., physically fractures are simply discrete surfaces. The two are more easily dis-
connected). For example, Fig. 4(b) shows two veins that approach tinguished where the fractures are filled with other material (as in
each other but do not meet on the bedding plane, on which they are veins, dykes, etc.) or are associated with changes of displacement.
exposed, with interaction indicated by rotation of the bridge be-
tween the veins. 4. Relative age relationships of interacting fractures
• Abutting: where one fracture links with another to form a T- or Y-
shaped node (Fig. 4c; Sanderson and Nixon, 2015). The fractures The geometric and topological relationships discussed above pro-
may have originally been unconnected (at least in the plane in vide a basis for evaluating the age relationships between fractures, with
which they are being observed) or one may have splayed off the key information being located at fracture tips (I nodes), and at abutting
other (see branching). The fractures may have formed in the same (Y) and crossing (X) nodes. Fig. 5 shows examples of different topolo-
deformation event or be from different events. Where a later gical relationships. Rotation of bridges between stepping fractures
opening-mode fracture abuts an earlier one, it commonly curves suggests they were synchronously active (Fig. 5a). A step between sub-
towards it and intersects at a high angle (typically 60°–90°), with the parallel fractures does not, however, necessarily prove that the

44
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 4. Veins from Liassic limestones and shales at Lilstock, Somerset, UK, as examples of different geometric relationships. (a) A partly-filled calcite vein that is isolated (in the plane of
view), which dies out as it passes from a limestone bed into the underlying shales. (b) Vein array, the segments in which interact across bridges. The array is cross-cut by later veins (c)
Abutting veins. (d) Branching (splaying) veins.

Fig. 5. Examples of joints and veins from the Liassic limestones of Somerset, UK, showing different topological relationships. (a) I-nodes at the tips of veins that interact across a bridge,
which has started to break (Lilstock). (b) I-nodes at the tips of joints that approach each other and interact but not actually meet (Lilstock). Interaction is indicated by the curvature of one
of the joints towards the other. (c) Y-node created by a joint that curves towards an earlier joint and abuts it at approximately 90° (Kilve). (d) X-node created by apparently mutually cross-
cutting joints (Kilve).

fractures were synchronously active. For example, the curvature at the only where one can reasonably interpret an abutting fracture to be
tip of one of the joints in Fig. 5(b) towards the other joint indicates only arrested at a through-going one. Fig. 5(d) shows joints that appear to
that it was synchronous with or later than the other joint. For joints mutually-crosscut.
(discrete, unfilled opening mode fractures) the negligible displacement Fig. 6 shows examples of intersecting fractures that can be char-
precludes the use of X-nodes and one must rely on abutting (Y) nodes acterised by their different age relationships, from the Mesozoic sedi-
(Fig. 5c). This generally works well (e.g., Sanderson, 2015, Fig. 9), but mentary rocks of Somerset, UK. Fig. 6(a) shows a network of possible

45
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 6. Examples of fractures from the Liassic rocks of Somerset, UK, showing interacting fractures with different age relationships. (a) Network of possible syneresis cracks on a limestone
bedding plane that appear to have formed synchronously (Kilve). (b) Calcite veins cut by later joints (Lilstock). Alteration around the joints illustrate they were conduits for fluid flow at a
time when the calcite veins were not conduits. (c) Calcite veins on a bedding plane of Liassic limestone (Lilstock). Two later veins intersect and utilise an earlier, with the trailing
geometry causing an increase in aperture of the earlier vein between the later veins. (d) Set of en echelon calcite veins partly reactivated as a later joint (Lilstock).

syneresis cracks on a limestone bedding plane that appear to have • The intersection line is parallel to the displacement direction of one
formed synchronously, and belong to the same deformation event. fracture and perpendicular to the displacement direction of the
Fig. 6(b) shows calcite veins cut by later joints, with alteration zones other fracture (Fig. 7e and f). Fig. 7(e) shows intersecting faults,
around the joints indicating they formed a network for fluid flow. veins and joints, in which the intersection lines between the faults
Fig. 6(c) shows two sets of calcite veins, the later set showing a trailing and joints are approximately perpendicular to the opening direc-
geometry through the earlier vein set, i.e., two younger veins are tions of the joints and parallel to the displacement directions of the
connected via an older vein, on which renewed displacement has oc- faults. Whilst these fractures are not the same age, they still display
curred (Peacock et al., 2016, 2017, Fig. 12c). Fig. 6(d) shows a set of a form of kinematic interaction. Synchronous fractures that may
calcite veins that has been partially reactivated by (or followed by) a display this relationship include normal and strike-slip faults in a
later joint. pull-apart basin (e.g., Karakhanian et al., 2002, Fig. 1).

5. Kinematic relationships between interacting fractures Peacock et al. (2017) characterise antithetic, synthetic and “neutral”
relationships between interacting faults. A similar approach may be
Kinematic relationships can be identified between interacting frac- applied to other fracture types, with shear displacements on shearing-
tures, although fractures may interact whether or not they are kine- model fractures corresponding with displacement directions on exten-
matic related. We define three end-member cases for the kinematic sion, contractional or combined fractures. We therefore suggest that
relationships between interacting fractures based on the angles between any two fractures can interact in the following ways:
the directions of wall rock displacement and the intersection line, i.e., the
line along which fractures meet (Fig. 7). This is used in the same sense • “Antithetic” relationship, where the fractures have opposite senses
as defined by Peacock et al. (2016) and is preferred over the term of displacement and usually occur at a high angle (> 45°) to one-
branch line (e.g., Butler, 1982). Note that intersection lines may be another, with slip approximately perpendicular to the intersection
curved (e.g., Butler, 1982, Fig. 7a). Fracture traces on bedding planes line (Fig. 8a).
and other exposure surfaces (e.g., Figs. 5 and 6) will generally meet • “Synthetic” relationships, where the faults have the same senses of
each other at intersection points (X- or Y-nodes). displacement and usually occur at a low angle (< 30°) to one-an-
We recognise three end-member situations, where: other, with displacement approximately perpendicular to the in-
tersection line. The example shown in Fig. 8(b) shows a shear
• The intersection line is perpendicular to the displacement directions fracture and a vein with displacement directions at a low angle to
of both intersecting fractures (Fig. 7a and b). The veins and stylolites each other, which we characterise as “synthetic”.
in Fig. 7(a) have displacement in the plane of the exposure surface, • “Neutral” relationship, where the fractures have the same senses of
while their intersection lines are approximately normal to the veins, displacement (Fig. 8c) or different sense of displacement (Fig. 8d),
stylolites and the exposure surface. that is approximately parallel to the intersection line. This is
• The intersection line is approximately parallel to the displacement common in normal fault systems that branch and intersect in map
direction (Fig. 7c and d). Fig. 7(d) shows intersecting normal faults, view, and between the detachment and lateral ramps of thrusts. We
with the intersection lines being sub-parallel to the displacement use the term “neutral” because there is no tendency for either ex-
directions of the faults. tension or contraction in the acute bisector, with little or no space

46
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 7. Examples of end-member relationships between the fracture


planes, displacement directions and intersection lines. (a) Veins and
stylolites in the Carboniferous Limestone at Ogmore-by-Sea, Vale of
Glamorgan, Wales. (b) Schematic diagram for the structure shown
in Fig. 7(a). The displacement directions of both structures are
approximately perpendicular to the intersection line. (c) Faults on a
steeply-dipping Liassic limestone bedding plane, Watchet,
Somerset, UK. Intersection points are marked by Xs. (d) Schematic
diagram for the structure shown in Fig. 7(c). The displacement di-
rection of the faults are approximately parallel to the intersection
lines. (e). Normal faults and joints on a gently-dipping Liassic
limestone bedding plane, East Quantoxhead, Somerset, UK. (f)
Schematic diagram for the structure shown in Fig. 7(e). The dis-
placement direction of the fault is approximately parallel to the
intersection line between the fault and the joint, while the dis-
placement direction of the joint is approximately perpendicular to
the intersection line.

problem caused by interaction between the faults (Peacock et al., directions of contraction, extension etc. (e.g., Marrett and Peacock,
2017). 1999). Note that some fracture networks form under conditions where
the principal strain axes are non-unique. For example, Fig. 9(c) shows a
Interacting opening and closing mode fractures may show a range of network of calcite veins in a septarian nodule that represent extension
relative displacement directions, from having sub-parallel displacement in all horizontal directions. Desiccation cracks (e.g., Weinberger, 1999)
directions (Fig. 8e) to perpendicular displacement directions (Fig. 8f). and polygonal faults (e.g., Lonergan et al., 1998) are analogous ex-
When considering any fracture network, the kinematic relationships amples.
between different fracture sets can be considered to be incompatible if
they could have formed under different stress or strain boundary con- 6. Discussion
ditions (Fig. 9a). Conversely we consider the fractures to be compatible
if the sets of fractures imply the same stress or strain boundary condi- 6.1. Mechanics of fracture interaction
tions (Fig. 9b). Since we will generally know little about the stress field,
or of the material properties and deformation mechanisms operating Field-based analysis of fractures and their interactions, such as those
when the fractures formed, the kinematic relationships are usually best described above, give evidence of the geometries, kinematics and re-
expressed in terms of the implied principal incremental strain axes, i.e., lative age relationships of the fractures. This information may then be

47
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 8. Different displacement relationships between interacting fractures from Liassic limestones, Somerset, UK. (a) Antithetic faults, with the faults having different displacement senses,
which are at a high angle to the intersection line (Watchet). (b) Synthetic relationship, with the opening of the veins being at < 90° to the later dextral fault, with these displacements
being at a high angle to the intersection lines (East Quantoxhead). (c) “Neutral” relationship, with the displacement directions of the synthetic faults being parallel to the intersection line
(Kilve). (d) “Neutral” relationship, with the displacement directions of the different faults being parallel to the intersection line (Watchet). Also see Peacock et al. (2017, Fig. 8d). (e)
Intersecting veins with the same displacement direction, obtained from mineral fibres (Lilstock). (f) Approximately perpendicular joint sets with opening directions at ∼ 90° to each other
(East Quantoxhead).

used as a basis for the analysis or modelling of the mechanics of in- characterised in terms of the orientations and magnitudes of the stresses
teraction between fractures, for example of the interpretation of the between and around the fractures.
stresses involved in fracturing. There have been many papers about While mechanical interpretation may give a deeper understanding
modelling the stress distributions around interacting fractures, in- of the development of structures than does kinematic analysis (e.g.,
cluding faults (e.g., Worum et al., 2005; Ganas et al., 2006; Pollard, 2000), we note that there are various problems involved in
Micklethwaite and Cox, 2006), dykes (e.g., Hamling et al., 2010; Hou using mechanics to characterise interaction between fractures, in-
et al., 2010; Re et al., 2015), and joints (e.g., Eberhardt, 2001; Moir cluding:
et al., 2010). While detailed analysis of mechanical relationships be-
tween fractures is beyond the scope of this paper, we suggest that in- • Proper interpretation of the stresses between interacting fractures
formation about mechanics between interacting fractures may be used depends on the correct interpretation of the geometries, kinematics
to characterise the interaction. For example, the interaction may be and age relationships of the interacting fractures. For example,

48
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

rates, rheology and the boundary conditions at the time or times of


fracturing.

6.2. Description and classification scheme

Peacock et al. (2017, Fig. 15) shows a classification scheme for in-
teracting faults based on their geometric, kinematic and age relation-
ships. In Fig. 10, we adapt the fault characterisation scheme of Peacock
et al. (2017) to show the relationships between any two interacting
fractures based on fracture types, geometric and topological relation-
ships, relative age relationships and kinematics (e.g., angles between
the intersection lines and displacement directions). This scheme is ap-
plicable to any two fractures, being independent of such factors as li-
thology, tectonic setting and scale. We suggest that this scheme is a
useful tool for analysing fracture systems because it emphasises the
geometric, temporal and kinematic relationships between the compo-
nents of a network. It therefore enables progress beyond analyses of the
orientations and displacements to enable the geometries, topology,
chronology and kinematics to be understood.

6.3. Scaling of fracture interactions

It is possible that some of the fracture relationships shown in the


examples (e.g., Figs. 3–9) may occur between much larger fractures. For
example, faults show self-similar geometries (Tchalenko, 1970;
Turcotte, 1989; Main, 1996), with fault lengths and displacements
commonly exhibiting power-law scaling relationships (e.g.,
Mandlebrot, 1982; Gillespie et al., 1993). Similarly, dykes can be
hundreds of kilometres long and hundreds of metres wide (e.g., Bunger
et al., 2013). Other structures, however, do not show such a wide range
of sizes. For example, whilst populations of stylolites (e.g., Peacock and
Azzam, 2006) and veins (e.g., Clark et al., 1995) have also been shown
to obey power-law scaling relationships, they rarely show displace-
ments of more than a metre. The description and classification scheme
shown in Fig. 10 may, therefore, be of limited use at larger scales, for
example in the interactions between faults and dykes. The possible
large examples shown in Table 1 include orogenic belts as large-scale
contractional structures.

6.4. The three-dimensional character of fracture interaction

The examples of fracture interactions presented in Figs. 3–9 are


exposed on rock surfaces, so are two-dimensional. In these cases, frac-
ture intersections lines are represented by intersection points where
they meet the exposed rock surface. The three-dimensional geometry of
fracture intersections is illustrated in Fig. 7. Whilst fractures are com-
monly studied in two-dimensions, especially in the field, it must be
remembered that they interact in three-dimensions. For example,
stepping joint segments may be linked out of the plane of observation
Fig. 9. Relationships between different fracture sets and strain from Mesozoic sedimen- (e.g., Delaney and Pollard, 1981, Fig. 29).
tary rocks, Somerset, UK. (a) Normal fault zone and calcite veins on a Liassic limestone
bedding plane (East Quantoxhead). Whilst the normal faults and the veins can be related 7. Conclusions
to the same strain regime, the network of joints that cut them was formed in an appar-
ently incompatible strain regime. (b) A zone of calcite-filled veins and pull-aparts linked
by slickolites in Liassic limestone (Lilstock). These structures all appear to have formed in
Various criteria can be used to evaluate the interactions between
the same strain regime. (c) Septarian nodule, representing extension in all horizontal any two fractures within a network. Different fracture types are re-
directions (Blue Anchor Bay). cognised (e.g., faults, deformation bands, joints, veins, dykes) and their
kinematics rationalised into opening-, closing- and shearing modes, that
can be applied to both discrete surfaces (e.g., joints) and banded
stresses between fractures of the same age are likely to be sig-
structures (e.g., fault zones and deformation bands). Fracture segments
nificantly different than stresses between fractures of different ages.
• Stresses will evolve as the fracture interaction progresses (e.g.,
are the branches of such networks and the nodes represent the location
of intersections with adjacent fractures (Y- and X-nodes). A fracture that
Nabavi et al., 2017), so a characterisation scheme would have to
terminates in the rock volume has an I-node at its tip and has no direct
incorporate the variations in the orientations and magnitudes of
geometrical linkage to other fractures.
stresses through time.
• Various aspects of the mechanics may be difficult to determine with
The geometry and topology of the network allows us to recognise
isolated, approaching, abutting, branching and cross-cutting relation-
any degree of accuracy, including the temperatures, pressures, strain
ships that characterise the interactions and the relative ages of the

49
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

(caption on next page)

50
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Fig. 10. Classification of intersecting fractures, generalising the scheme for interacting faults presented by Peacock et al. (2017, Fig. 15). (a) to (d) Classification of fracture types based on
the displacement direction relative to the fracture plane. (a) Extension fracture, with extension approximately perpendicular to the fracture plane (e.g., joints, some veins, dykes, sills). (b)
Shear fracture, with displacement along the plane of the fracture (e.g., faults, some deformation bands). (c). Contractional, with contraction approximately perpendicular to the fracture
plane (e.g., stylolites). (d) Combined, with net displacement neither in the plane of the fracture or perpendicular to the fracture (e.g., slickolites, some veins and some compaction bands).
(e) to (i) Classification based on the geometric relationships between the fractures. (e) Isolated fracture. (f) The fractures interact as they approach each other, but are not connected. (g)
One fracture abuts the other. (h) One fracture (earlier) is cut by the other (later). (i) The fractures mutually cross-cut each other. (j) to (l) Classification based on topographic relationships.
(j) A geometrically isolated fracture with two I-nodes. (k) One fracture abuts another to form a Y-node. (l) Two fractures cut each other to form an X-node. (m) to (p) Classification based
on relative ages. (m) The fractures are synchronous. (n) The fractures are different ages. (o) “Trailing” geometry, where part of an earlier fracture is reactivated by interaction with later
fractures. (p) One or both fractures have been reactivated, in this case a vein wall has been reactivated as a stylolite. (q) to (s) Classification based on the relationship between the
intersection line and the displacement direction. (q) The intersection line is parallel to the displacement directions or both interacting fractures. This is only likely for some types of fault
interactions. (r) The intersection line is perpendicular to the displacement directions of both fractures, as would be typical of interacting extension fractures. (s) The intersection line is
parallel to the displacement direction of one fracture (typically a fault) and perpendicular to the displacement direction of the other fracture. (t) to (x) Classification based on the
displacements and strains between the interacting fractures. (t) “Antithetic” relationship, in which the displacement directions on the two fractures are at > 90° in their acute bisector.
(u) “Synthetic” relationship, in which the displacement directions on the two fractures are at < 90° in their acute bisector. (v) Neutral relationship, where the fractures have the same
senses of slip, approximately parallel to the intersection line. (w) The fractures have the same displacement direction. (x) The displacement directions of the fractures are at 90° to each
other.

fractures. From this information we can determine sets of fractures that 70–87.
could have formed during different events, representing different stages Clark, M.B., Brantley, S.L., Fisher, D.M., 1995. Power-law vein-thickness distributions and
positive feedback in vein growth. Geology 23, 975–978.
in the geological evolution of the network. Geometrical linkage need Cobbold, P.R., Zanella, A., Rodrigues, N., Løseth, H., 2013. Widespread bedding-parallel
not imply kinematic interaction (e.g., cross-cutting fractures of different veins of fibrous calcite (“beef”) in a mature source rock (Vaca Muerta Fm, Neuquén
age), and vice versa (e.g., at bridges between stepping veins). Basin, Argentina): evidence for overpressure and horizontal compression. Mar.
Petrol. Geol. 43, 1–20.
Fracture interactions can also be characterised by the relationship Cox, S.F., 2005. Coupling between deformation, fluid pressures, and fluid flow in ore-
between the intersection line and the displacement direction. We producing hydrothermal systems at depth in the crust. Econ. Geol. 100, 39–75.
identify three end-member cases: whether the intersection line is par- Crider, J.G., Pollard, D.D., 1998. Fault linkage: three-dimensional mechanical interaction
between echelon normal faults. J. Geophys. Res. 103, 24373–24391.
allel or perpendicular to the displacement directions of both fractures,
Curewitz, D., Karson, J.A., 1997. Structural settings of hydrothermal outflow: fracture
or whether the fractures have different displacement directions. Such permeability maintained by fault propagation and interaction. J. Volcanol.
kinematic interactions do not need to indicate that the fractures were Geotherm. Res. 79, 149–168.
Dabi, G., Siklósy, Z., Schubert, F., Bajnóczi, B., Tóth, T.M., 2011. The relevance of vein
active at the same time, either through synchronous formation or re-
texture in understanding the past hydraulic behaviour of a crystalline rock mass:
activation. reconstruction of the palaeohydrology of the Mecsekalja Zone, south Hungary.
The scheme outlined in this paper provides a basis for describing Geofluids 11, 309–327.
and understanding the different ways various fractures interact, and Dart, C.J., McClay, K., Hollings, P.N., 1995. 3D analysis of inverted extensional fault
systems, southern Bristol Channel basin, UK. In: Buchanan, J.G., Buchanan, P.G.
provides a basis for describing the geometries and development of (Eds.), Basin Inversion. Geological Society, London, pp. 393–413 Special
fracture networks. Publications 88.
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.
Acknowledgements U. S. Geol. Surv. Prof. Pap. 1202, 61.
de Sitter, L.U., 1964. Structural Geology, second ed. McGraw-Hill, New York.
Rodolfo Carosa, Alex Manda and an anonymous reviewer are Dini, A., Westerman, D.S., Innocenti, F., Rocchi, S., 2008. Magma emplacement in a
transfer zone: the Miocene mafic Orano dyke swarm of Elba Island, Tuscany, Italy. In:
thanked for their helpful comments. We thank Casey Nixon for useful Thomson, K., Petford, N. (Eds.), Structure and Emplacement of High-level Magmatic
discussions about this work. Systems. Geological Society, London, pp. 131–148 Special Publications 302.
Du Bernard, X., Eichhubl, P., Aydin, A., 2002. Dilation bands: a new form of localized
failure in granular media. Geophys. Res. Lett. 29, 1–4.
References Eberhardt, E., 2001. Numerical modelling of three-dimension stress rotation ahead of an
advancing tunnel face. Int. J. Rock Mech. Min. Sci. 38, 499–518.
Arndt, M., Virgo, S., Cox, S.F., Urai, J.L., 2014. Changes in fluid pathways in a calcite vein Engelder, J.T., 1974. Cataclasis and the generation of fault gouge. Geol. Soc. Am. Bull. 85,
mesh (Natih Formation, Oman Mountains): insights from stable isotopes. Geofluids 1515–1522.
14, 391–418. Ernst, R.E., Grosfils, E.B., Mège, D., 2001. Giant dike swarms: earth, Venus, and Mars.
Aydin, A., 1978. Small faults formed as deformation bands in sandstone. Pure Appl. Annu. Rev. Earth Planet. Sci. 29, 489–534.
Geophys. 116, 913–930. Ferrill, D.A., Morris, A.P., McGinnis, R.N., 2009. Crossing conjugate normal faults in field
Aydin, A., Borja, R.I., Eichhubl, P., 2006. Geological and mathematical framework for exposures and seismic data. Am. Assoc. Petrol. Geol. Bull. 93, 1471–1488.
failure modes in granular rock. J. Struct. Geol. 28, 83–98. Fletcher, R.C., Pollard, D.D., 1981. Anticrack model for pressure solution surfaces.
Belayneh, M., Cosgrove, J., 2010. Hybrid veins from the southern margin of the Bristol Geology 9, 419–424.
channel Basin, UK. J. Struct. Geol. 32, 192–201. Fort, X., Brun, J.P., 2012. Kinematics of regional salt flow in the northern Gulf of Mexico.
Bogdanoff, S., Michard, A., Mansour, M., Poupeau, G., 2000. Apatite fission track analysis In: Alsop, G.I., Archer, S.G., Hartley, A.J., Grant, N.T., Hodgkinson, R. (Eds.), Salt
in the Argentera massif: evidence of contrasting denudation rates in the External Tectonics, Sediments and Prospectivity. Geological Society, London, pp. 265–287
Crystalline Massifs of the Western Alps. Terra Nova 12, 117–125. Special Publications 363.
Bons, P., Montenari, M., 2005. The formation of antitaxial calcite veins with well-de- Fossen, H., Schultz, R.A., Shipton, Z.K., Mair, K., 2007. Deformation bands in sandstone: a
veloped fibres, Oppaminda Creek, South Australia. J. Struct. Geol. 27, 231–248. review. J. Geol. Soc. 164, 755–769.
Bons, P.D., Elburg, M.A., Gomez-Rives, E., 2012. A review of the formation of tectonic Fossen, H., Schultz, R.A., Torabi, A., 2011. Conditions and implications for compaction
veins and their microstructures. J. Struct. Geol. 43, 33–62. band formation in the Navajo Sandstone, Utah. J. Struct. Geol. 33, 1477–1490.
Bonson, C.G., Childs, C., Walsh, J.J., Schöpfer, M.P.J., Carboni, V., 2007. Geometric and Friedman, M., Logan, J.M., 1973. Lüders' bands in experimentally deformed sandstone
kinematic controls on the internal structure of a large normal fault in massive and limestone. Geol. Soc. Am. Bull. 84, 1465–1476.
limestones: the Maghlaq Fault, Malta. J. Struct. Geol. 29, 336–354. Ganas, A., Sokos, E., Agalos, A., Leontakianakos, G., Pavlides, S., 2006. Coulomb stress
Bourne, S.J., Willemse, E.J.M., 2001. Elastic stress control on the pattern of tensile triggering of earthquakes along the Atalanti Fault, central Greece: two April 1894
fracturing around a small fault network at Nash Point. J. Struct. Geol. 23, 1753–1770. M6+ events and stress change patterns. Tectonophysics 420, 357–369.
Bunger, A.P., Menand, T., Cruden, A., Zhang, X., Halls, H., 2013. Analytical predictions Gillespie, P.A., Howard, C.B., Walsh, J.J., Watterson, J., 1993. Measurement and char-
for a natural spacing within dyke swarms. Earth Planet. Sci. Lett. 375, 270–279. acterization of spatial distributions of fractures. Tectonophysics 226, 113–141.
Busquets, P., Colombo, F., Heredia, N., Sole de Porta, N.L.R., Rodriguez Fernandez, L.R., Glen, R.A., Hancock, P.L., Whittaker, A., 2005. Basin inversion by distributed deforma-
Alvarez Marron, J., 2005. Age and tectonostratigraphic significance of the upper tion: the southern margin of the Bristol Channel Basin, England. J. Struct. Geol. 27,
Carboniferous series in the basement of the Andean Frontal Cordillera: geodynamic 2113–2134.
implications. Tectonophysics 399, 181–194. Gökten, E., Demirtaş, R., Özaksoy, V., Herece, E., Varol, B., Temiz, U., 2011. Faulting and
Butler, R.W.H., 1982. The terminology of structures in thrust belts. J. Struct. Geol. 4, stress distribution in the Bolu Pull-apart Basin (North Anatolian fault zone, Turkey):
239–245. the significance of new dates obtained from the basin fill. Turkish J. Earth Sci. 20,
Choi, J.H., Edwards, P., Ko, K., Kim, Y.S., 2016. Definition and classification of fault 1–26.
damage zones: a review and a new methodological approach. Earth-Sci. Rev. 152, Gourmelen, N., Dixon, T.H., Amelung, F., Schmalzle, G., 2011. Acceleration and evolution

51
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

of faults: an example from the Hunter Mountain-Panamint Valley fault zone, Eastern National Park, Utah. Geol. Soc. Am. Bull. 111, 808–822.
California. Earth Planet. Sci. Lett. 301, 337–344. Morley, C.K., 2014. Outcrop examples of soft-sediment deformation associated with
Hamling, I.J., Wright, T.J., Calais, E., Bennati, L., Lewi, E., 2010. Stress transfer between normal fault terminations in deepwater, Eocene turbidites: a previously undescribed
thirteen successive dyke intrusions in Ethiopia. Nat. Geosci. 3, 713–717. conjugate fault termination style? J. Struct. Geol. 69, 189–208.
Hendry, J.P., Gregg, J.M., Shelton, K.L., Somerville, I.D., Crowley, S.F., 2015. Origin, Morley, C.K., Kongwung, B., Julapour, A.A., Abdolghafourian, M., Hajian, M., Waples, D.,
characteristics and distribution of fault-related and fracture-related dolomitization: Warren, J., Otterdoom, H., Srisuriyon, K., Kazemi, H., 2009. Structural development
insights from Mississippian carbonates, Isle of Man. Sedimentology 62, 717–752. of a major late Cenozoic basin and transpressional belt in central Iran: the Central
Hou, G., Kusky, T.M., Wang, C., Wang, Y., 2010. Mechanics of the giant radiating Basin in the Qom-Saveh area. Geosphere 5, 325–362.
Mackenzie dyke swarm: a paleostress field modeling. J. Geophys. Res. 115, B02402. Mughieda, O., Alzo’ubi, A.K., 2004. Fracture mechanisms of offset rock joints – a la-
http://dx.doi.org/10.1029/2007JB005475. boratory investigation. Geotechnical Geol. Eng. 22, 545–562.
Johansen, T.E.S., Fossen, H., Kluge, R., 2005. The impact of syn-faulting porosity re- Nabavi, S.T., Alavi, S.A., Mohammadi, S., Ghassemi, M.R., Frehner, M., 2017. Analysis of
duction on damage zone architecture in porous sandstone: an outcrop example from transpression within contractional fault steps using finite-element method. J. Struct.
the Moab Fault, Utah. J. Struct. Geol. 27, 1469–1485. Geol. 96, 1–20.
Jolly, R.J.H., Cosgrove, J.W., Dewhurst, D.N., 1998. Thickness and spatial distributions of Nicol, A., Walsh, J.J., Watterson, J., Bretan, P.G., 1995. Three-dimensional geometry and
clastic dykes, northwest Sacramento Valley, California. J. Struct. Geol. 20, growth of conjugate normal faults. J. Struct. Geol. 17, 847–862.
1663–1672. Nicoll, K., 2010. Geomorphic and hazard vulnerability assessment of recent residential
Karakhanian, A., Djrbashian, R., Trifonov, V., Philip, H., Arakelian, S., Avagian, A., 2002. developments on landslide-prone terrain: the case of the Traverse Mountains, South
Holocene-historical volcanism and active faults as natural risk factors for Armenia Salt Lake Valley, Utah, USA. J. Geogr. Regional Plan. 3, 126–141.
and adjacent countries. J. Volcanol. Geotherm. Res. 113, 319–344. Nitecki, M.H., 1962. Observations on slickolites. J. Sediment. Petrol. 32, 435–439.
Katsman, R., 2010. Extensional veins induced by self-similar dissolution at stylolites: Nixon, C.W., Sanderson, D.J., Dee, S., Bull, J.M., Humphreys, R., Swanson, M., 2014.
analytical modelling. Earth Planet. Sci. Lett. 299, 33–41. Fault interactions and reactivation within a normal fault network at Milne Point,
Kattenhorn, S.A., Aydin, A., Pollard, D.D., 2000. Joints at high angles to normal fault Alaska. Am. Assoc. Petrol. Geol. Bull. 98, 2081–2107.
strike: an explanation using 3-D numerical models of fault-perturbed stress fields. J. Omosanya, K.O., Akinbodewa, E.A., Mosuro, G.O., Kaigama, U., 2013. Alteration of σ3
Struct. Geol. 22, 1–23. field during the evolution of a polycyclic basement complex. J. Geol. Min. Res. 5,
Katz, D.A., Eberli, G.P., Swart, P.K., Smith, L.B., 2006. Tectonic-hydrothermal brecciation 23–37.
associated with calcite precipitation and permeability destruction in Mississippian Paquet, F., Dauteuil, O., Hallot, E., Moreau, F., 2007. Tectonics and magma dynamics
carbonate reservoirs, Montana and Wyoming. Am. Assoc. Petrol. Geol. Bull. 90, coupling in a dyke swarm of Iceland. J. Struct. Geol. 29, 1477–1493.
1803–1841. Pati, J.K., Patel, S.C., Pruseth, K.L., Malviya, V.P., Arima, M., Raju, S., Pati, P., Prakash,
Khajavi, N., Quigley, M., Langridge, R.M., 2014. Influence of topography and basement K., 2007. Geology and geochemistry of giant quartz veins from the Bundelkhand
depth on surface rupture morphology revealed from LiDAR and field mapping, Hope Craton, central India and their implications. J. Earth Syst. Sci. 116, 497–510.
Fault, New Zealand. Tectonophysics 630, 265–284. Peacock, D.C.P., 2001. The temporal relationship between joints and faults. J. Struct.
Kim, Y.S., Sanderson, D.J., 2006. Structural similarity and variety at the tips in a wide Geol. 23, 329–341.
range of strike-slip faults: a review. Terra Nova 18, 330–344. Peacock, D.C.P., Azzam, I.N., 2006. Development and scaling relationships of a stylolite
Kim, Y.S., Peacock, D.C.P., Sanderson, D.J., 2004. Fault damage zones. J. Struct. Geol. 26, population. J. Struct. Geol. 28, 1883–1889.
503–517. Peacock, D.C.P., Sanderson, D.J., 1991. Displacements, segment linkage and relay ramps
Knipe, R.J., 1992. Faulting processes and fault seal. In: In: Larsen, R.M., Brekke, H., in normal fault zones. J. Struct. Geol. 13, 721–733.
Larsen, B.T., Tallerass, E. (Eds.), Structural and Tectonic Modelling and its Peacock, D.C.P., Sanderson, D.J., 1994. Geometry and development of relay ramps in
Application to Petroleum Geology, vol. 1. Norwegian Petroleum Society Special normal fault systems. Am. Assoc. Petrol. Geol. Bull. 78, 147–165.
Publications, pp. 325–342. Peacock, D.C.P., Sanderson, D.J., 1999. Deformation history and basin-controlling faults
Knott, S.D., 1994. Fault zone thickness versus displacement in the Permo-Triassic sand- in the Mesozoic sedimentary rocks of the Somerset coast. Proc. Geol. Assoc. 110,
stones of NW England. J. Geol. Soc. Lond. 151, 17–25. 41–52.
Lecumberri-Sanchez, P., Newton, M.C., Westman, E.C., Kamilli, R.J., Canby, V.M., Peacock, D.C.P., Nixon, C.W., Rotevatn, A., Sanderson, D.J., Zuluaga, L.F., 2016. Glossary
Bodnar, R.J., 2013. Temporal and spatial distribution of alteration, mineralization of fault and other fracture networks. J. Struct. Geol. 92, 12–29.
and fluid inclusions in the transitional high-sulfidation epithermal-porphyry copper Peacock, D.C.P., Nixon, C.W., Rotevatn, A., Sanderson, D.J., Zuluaga, L.F., 2017.
system at Red Mountain, Arizona. J. Geochem. Explor. 125, 80–93. Interacting faults. J. Struct. Geol. 97, 1–22.
Lin, C.C., Lin, A.T.S., Liu, C.S., Chen, G.Y., Liao, W.Z., Schnurle, P., 2009. Geological Petit, J.P., Wibberley, C.A.J., Ruiz, G., 1999. ‘Crack-seal’ slip: a new fault valve me-
controls on BSR occurrences in the incipient arc-continent collision zone off south- chanism? J. Struct. Geol. 21, 1199–1207.
west Taiwan. Mar. Petrol. Geol. 26, 1118–1131. Polito, P.A., Bone, Y., Clarke, J.D.A., Mernagh, T.P., 2001. Compositional zoning of fluid
Lisle, R., 2013. Shear zone deformation determined from sigmoidal tension gashes. J. inclusions in the Archaean Junction gold deposit, Western Australia: a process of fluid
Struct. Geol. 50, 35–43. - wall-rock interaction? Aust. J. Earth Sci. 48, 833–855.
Lonergan, L., Cartwright, J., Jolly, R., 1998. The geometry of polygonal fault systems in Pollard, D.D., 2000. Strain and stress: discussion. J. Struct. Geol. 22, 1359–1367.
Tertiary mudrocks of the North Sea. J. Struct. Geol. 20, 529–548. Pollard, D.D., Aydin, A., 1988. Progress in understanding jointing over the past century.
Macdonald, R., Fettes, D.J., Bagiński, B., 2015. The Mull Paleocene dykes: some insights Geol. Soc. Am. Bull. 100, 1181–1204.
into the nature of major dyke swarms. Scott. J. Geol. 51, 116–124. Price, N.J., 1966. Fault and Joint Development in Brittle and Semi-brittle Rock.
Maerten, L., 2000. Variation in slip on intersecting normal faults: implications for pa- Pergamon, Oxford.
leostress inversion. J. Geophys. Res. 105 25,565–25,565. Ramsey, J.M., Chester, F.M., 2004. Hybrid fracture and the transition from extension
Main, I., 1996. Statistical physics, seismogenesis, and seismic hazard. Rev. Geophys. 34, fracture to shear fracture. Nature 428, 63–66.
433–462. Re, G., White, J.D.L., Or, M.H., 2015. Dikes, sills, and stress-regime evolution during
Mandlebrot, B.B., 1982. The Fractal Geometry of Nature. W.H. Freeman, San Francisco. emplacement of the jagged rocks complex, Hopi Buttes Volcanic field, Navajo Nation,
Manda, A.K., Horsman, E., 2015. Fracturesis Jointitis: causes, symptoms, and treatment in USA. J. Volcanol. Geotherm. Res. 295, 65–79.
groundwater communities. Groundwater 53, 836–840. Rotevatn, A., Thorsheim, E., Bastesen, E., Fossmark, H.S., Torabi, A., Sælen, G., 2016.
Manzocchi, T., 2002. The connectivity of two dimensional networks of spatially corre- Sequential growth of deformation bands in carbonate grainstones in the hangingwall
lated fractures. Water Resour. Res. 38. http://dx.doi.org/10.1029/2000WR000180. of an active growth fault: implications for deformation mechanisms in different
Marchal, D., Guiraud, M., Rives, T., 2003. Geometric and morphologic evolution of tectonic regimes. J. Struct. Geol. 90, 27–47.
normal fault planes and traces from 2D to 4D data. J. Struct. Geol. 25, 135–158. Sanderson, D.J., 2015. Field-based structural studies as analogues to sub-surface re-
Marrett, R., Peacock, D.C.P., 1999. Strain and stress. J. Struct. Geol. 21, 1057–1063. servoirs. In: In: Bowman, M., Smyth, H.R., Good, T.R., Passey, S.R., Hirst, J.P.P.,
Mathieu, L., van Wyk de Vries, B., Holohan, E.P., Troll, V.R., 2008. Dykes, cups, saucers Jordan, C.J. (Eds.), The Value of Outcrop Studies in Reducing Subsurface Uncertainty
and sills: analogue experiments on magma intrusion into brittle rocks. Earth Planet. and Risk in Hydrocarbon Exploration and Production Geological Society, London, pp.
Sci. Lett. 271, 1–13. 207–217 Special Publications 436.
Micarelli, L., Benedicto, A., 2008. Normal fault terminations in limestones from the SE- Sanderson, D.J., Nixon, C.W., 2015. The use of topology in fracture network character-
Basin (France): implications for fluid flow. In: Wibberley, C.A.J., Kurz, W., Imber, J., ization. J. Struct. Geol. 72, 55–66.
Holdsworth, R.E., Collettini, C. (Eds.), The Internal Structure of Fault Zones: Schultz, R.A., Fossen, H., 2008. Terminology for structural discontinuities. Am. Assoc.
Implications for Mechanical and Fluid-flow Properties. Geological Society, London, Petrol. Geol. Bull. 92, 853–867.
pp. 123–138 Special Publications 299. Singh, V., Slabunov, A., 2014. The Central Bundelkhand Archaean greenstone complex,
Micklethwaite, S., Cox, S.F., 2006. Progressive fault triggering and fluid flow in after- Bundelkhand craton, central India: geology, composition, and geochronology of su-
shock domains: examples from mineralized Archaean fault systems. Earth Planet. Sci. pracrustal rocks. Int. Geol. Rev. 57, 1349–1364.
Lett. 250, 318–330. Srivastava, D.C., Engelder, T., 1990. Crack-propagation sequence and pore-fluid condi-
Misra, S., Mandal, N., Dhar, R., Chakraborty, C., 2009. Mechanisms of deformation lo- tions during fault-bend folding in the Appalachian Valley and Ridge, central
calization at the tips of shear fractures: findings from analogue experiments and field Pennsylvania. Geol. Soc. Am. Bull. 102, 116–128.
evidence. J. Geophys. Res. 114, B04204. http://dx.doi.org/10.1029/2008JB005737. Stockdale, P.B., 1926. The stratigraphic significance of solution in rocks. J. Geol. 34,
Mollema, P.N., Antonellini, M.A., 1996. Compaction bands: a structural analog for anti- 399–414.
mode I cracks in aeolian sandstone. Tectonophysics 267, 209–228. Taylor, M., Peltzer, G., 2006. Current slip rates on conjugate strike-slip faults in central
Moir, H., Lunn, R.J., Shipton, Z.K., Kirkpatrick, J.D., 2010. Simulating brittle fault evo- Tibet using synthetic aperture radar interferometry. J. Geophys. Res. 111
lution from networks of pre-existing joints within crystalline rock. J. Struct. Geol. 32, B12402–B12402.
1742–1753. Tchalenko, J.S., 1970. Similarities between shear zones of different magnitudes. Geol.
Moore, J.M., Schultz, R.A., 1999. Processes of faulting in jointed rocks of Canyonlands Soc. Am. Bull. 81, 1625–1640.

52
D.C.P. Peacock et al. Journal of Structural Geology 106 (2018) 41–53

Tewksbury, B.J., Hogan, J.P., Kattenhorn, S.A., Mehrtens, C.J., Tarabees, E.A., 2014. district of central Peru. Bol. Soc. Geol. del Perú 104, 59–80.
Polygonal faults in chalk: insights from extensive exposures of the Khoman Worum, G., Michon, L., van Balen, R.T., van Wees, J.D., Cloetingh, S., Pagnier, H., 2005.
Formation, Western Desert, Egypt. Geology 42, 479–482. Pre-Neogene controls on present-day fault activity in the West Netherlands Basin and
Trautwein-Bruns, U., Schulze, K.C., Becker, S., Kukla, P.A., Urai, J.L., 2010. In situ stress Roer Valley Rift System (southern Netherlands): role of variations in fault orientation
variations at the Variscan Deformation Front - results from the deep Aachen geo- in a uniform low-stress regime. Quat. Sci. Rev. 24, 475–490.
thermal well. Tectonophysics 493, 196–211. Xiao-shuang, X., Jing-ru, T., Hua, K., Shao-xun, H., 2005. Control of Relay Structure on
Tuckwell, G.W., Lonergan, L., Jolly, R.J.H., 2003. The control of stress history and flaw Mineralization of Sedimentary-exhalative Ore Deposit in Growth Faults of Graben
distribution on the evolution of polygonal fracture networks. J. Struct. Geol. 25, Systems. Journal of Central South University of Technology 12, pp. 340–345.
1241–1250. Yin, H.W., Groshong, R.H., 2007. A three-dimensional kinematic model for the de-
Turcotte, D.L., 1989. Fractals in geology and geophysics. Pure Appl. Geophys. 131 formation above an active diapir. Am. Assoc. Petrol. Geol. Bull. 91, 343–363.
171–19. Zhang, X., Sanderson, D.J., 2002. Numerical Modelling and Analysis of Fluid Flow and
Walsh, J.J., Watterson, J., 1987. Distributions of cumulative displacement and seismic Deformation of Fractured Rock Masses. pp. 288 Pergamon.
slip on a single normal fault. J. Struct. Geol. 9, 1039–1046. Zulauf, G., Gutiérrez-Alonso, G., Kraus, R., Petschick, R., Potel, S., 2011. Formation of
Weinberger, R., 1999. Initiation and growth of cracks during desiccation of stratified chocolate-tablet boudins in a foreland fold and thrust belt: a case study from the
muddy sediments. J. Struct. Geol. 21, 379–386. external Variscides (Almograve, Portugal). J. Struct. Geol. 33, 1639–1649.
Wise, J.M., 2010. Evaluation of conjugate vein formation in the Huachocolpa base-metal

53

You might also like