Sie sind auf Seite 1von 24

Society of Economic Geologists

Reviews v. 14, 2001, p. 1–24

Chapter 1

Principles of Structural Control on Permeability and


Fluid Flow in Hydrothermal Systems
S. F. COX,†
Centre for Advanced Studies of Ore Systems, Department of Geology and Research School of Earth Sciences,
The Australian National University, Canberra, ACT 0200, Australia

M. A. KNACKSTEDT,
Research School of Physical Sciences and Engineering, The Australian National University, Canberra, ACT 0200, Australia

AND J. BRAUN
Research School of Earth Sciences, The Australian National University, Canberra, ACT 0200, Australia

Abstract

Fluid pathways between metal sources and sites of ore deposition in hydrothermal systems are gov-
erned by fluid pressure gradients, buoyancy effects, and the permeability distribution. Structural controls
on ore formation in many epigenetic systems derive largely from the role that deformation processes and
fluid pressures play in generating and maintaining permeability within active faults, shear zones, associ-
ated fracture networks, and various other structures at all crustal levels.
In hydrothermal systems with low intergranular porosity, pore connectivity is low, and fluid flow is typ-
ically controlled by fracture permeability. Deformation-induced fractures develop on scales from microns
to greater than hundreds of meters. Because mineral sealing of fractures can be rapid relative to the life-
times of hydrothermal systems, sustained fluid flow occurs only in active structures where permeability is
repeatedly renewed.
In the brittle upper crust, deformation-induced permeability is associated with macroscopic fracture ar-
rays and damage products produced in episodically slipping (seismogenic) and aseismically creeping faults,
growing folds, and related structures. In the more ductile mid- to lower crust, permeability enhancement is
associated with grain-scale dilatancy (especially in active shear zones), as well as with macroscopic hydraulic
fracture arrays. Below the seismic–aseismic transition, steady state creep leads to steady state permeability
and continuous fluid flow in actively deforming structures. In contrast, in the seismogenic regime, large
cyclic changes in permeability lead to episodic fluid flow in faults and associated fractures.
The geometry and distribution of fracture permeability is controlled fundamentally by stress and fluid
pressure states, but may also be influenced by preexisting mechanical anisotropies in the rock mass. Frac-
ture growth is favored in high pore fluid factor regimes, which develop especially where fluids discharge
from faults or shear zones beneath low-permeability flow barriers. Flow localization within faults and shear
zones occurs in areas of highest fracture aperture and fracture density, such as damage zones associated
with fault jogs, bends, and splays. Positive feedback between deformation, fluid flow, and fluid pressure
promotes fluid-driven growth of hydraulically linked networks of faults, fractures, and shear zones.
Evolution of fluid pathways on scales linking fluid reservoirs and ore deposits is influenced by the rel-
ative proportions of backbone, dangling, and isolated structures in the network. Modeling of the growth
of networks indicates that fracture systems reach the percolation threshold at low bulk strains. Just above
the percolation threshold, flow is concentrated along a small proportion of the total fracture population,
and favors localized ore deposition. At higher strains, flow is distributed more widely throughout the frac-
ture population and, accordingly, the potential for localized, high-grade ore deposition may be reduced.

Introduction monly localized within faults, shear zones, or associated frac-


ture systems. Additionally, mineralization can be restricted
THE FORMATION of many types of epigenetic ore deposits to particular parts of faults or shear zones such as jogs or
involves some form of control by structures produced dur- bends (Hulin, 1929; Newhouse, 1942; McKinstry, 1948; Sib-
ing crustal deformation. In particular, ore deposition is com- son, 1987), to fold-related structures such as saddle reefs
(Hulin, 1929; Chace, 1949; Cox et al., 1991), or related to
†Corresponding author: e-mail, sfcox@geology.anu.edu.au deformation around heterogeneities such as competent

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
2 COX ET AL.

rock masses (Oliver et al., 2001). Epigenetic mineralization permeability is generally high enough that pore spaces are
also is typically localized along low displacement faults and highly interconnected and fluid pressures are close to
fracture systems (e.g., Robert et al., 1995; Cox, 1999). hydrostatic values (i.e., owing to the weight of the fluid col-
Hydrothermal mineral deposits are produced by focused umn). In this regime, topographic relief is a major factor
flow of large volumes of fluids (Henley et al., 1985; Fyfe, establishing hydraulic gradients that drive subsurface flow
1987; Cox, 1999). Mass balance calculations, based for downward from regions of high topography towards
example on silica and metal solubility, indicate that time- regions of lower topography. Depths of topographically
integrated fluid volumes in many types of hydrothermal ore driven fluid penetration are typically up to several kilome-
systems are typically greater than several cubic kilometers. ters (Forster and Smith, 1990).
Crustal deformation processes influence fluid flow in Transient, deformation-induced pore pressure changes
hydrothermal systems via controls on rock permeability and that set up vertical and lateral hydraulic gradients can be
the driving forces for fluid flow. Understanding how defor- important at all crustal depths. These changes arise from
mation processes and structures influence permeability evo- compaction of intergranular pore spaces during burial or
lution is a key aspect to understanding the architecture of regional deformation, grain-scale dilatation associated with
fluid pathways in hydrothermal systems, and the factors con- granular flow, or from microscopic to macroscopic crack
trolling where ore deposits form along these pathways. This growth and closure during deformation. Such deforma-
review explores how active deformation controls the perme- tion-induced pore pressure gradients are associated with
ability of rock masses through the formation of macroscopic both small elastic responses (poroelastic effects) or larger
fracture systems, damage zones, and grain-scale microcrack inelastic deformations of intergranular pore spaces and
permeability within actively deforming structures. The role fractures owing to stress changes associated with the seismic
of fluid pressures and stress regimes in driving permeability cycle and heterogeneous deformation (Muir-Wood and
enhancement and growth of permeable fault/fracture/ King, 1993). In particular, macroscopic dilatancy associated
shear networks is highlighted. We outline how stress regimes with episodic fault slip exerts a powerful control on fluid
and variations in fluid pressure in hydrothermal systems gov- migration around active faults in the crustal seismogenic
ern the location and geometry of fluid pathways between regime (Sibson, 1987, 1993, 2001). Transient fluid migra-
metal sources, fluid sources, and sites of ore deposition. We tion in response to pressure gradients due to deformation-
also examine how linkages among structures develop during induced grain-scale dilatancy during regional deformation
progressive deformation, and how they control the architec- has been discussed by Cox and Etheridge (1989) and
ture of fluid pathways from deposit to crustal scales. McCaig and Knipe (1990), and modeled by Ord and Oliver
(1997) and Oliver et al. (2001).
Principles of Fluid Flow in Porous and Fractured Rocks Driving pressure-gradients within the crust are also gen-
Fluid migration through the Earth’s crust occurs in erated by development of suprahydrostatic-pressured fluid
response to various driving forces. The fluid flux and geom- reservoirs. For example, suprahydrostatic fluid pressuriza-
etry of flow is fundamentally controlled by permeability tion can be associated with emplacement of magmas into a
variations in the crust. In this section, we outline the forces cooler, porous, fluid-saturated rock-mass. Transient driving
driving crustal fluid flow and examine the dependence of pressure-gradients are also associated with metamorphic
fluid flux on rock permeability. We also discuss how the evo- fluid-production and fluid-absorption reactions (Walther,
lution of grain-scale to macroscopic permeability in 1990; Rumble, 1994), as well as by fluid expulsion associ-
hydrothermal systems is influenced by stress regimes and ated with crystallization of hydrous silicate melts. Modeling
fluid pressure regimes during both brittle and ductile styles of two-dimensional advective flow regimes associated with
of deformation. Two central points are highlighted: firstly, drainage of suprahydrostatic reservoirs has been conducted
the geometry of fluid pathways is strongly dependent on by Matthai and Roberts (1997). In these flow regimes, spa-
the geometry and style of deformation; and secondly, per- tial variations in permeability (for example, owing to the
meability is a transient rock property that can be rapidly presence of active and permeable faults or shear zones)
reduced by porosity-destruction processes such as com- establish hydraulic gradients that focus fluid flow and gov-
paction and sealing of intergranular pores and fractures. ern flow pathways (Fig. 1). Temporal and spatial variations
Maintenance of permeability and fluid flow in hydrother- in permeability induced by episodic fault rupture particu-
mal systems is, therefore, dependent upon active deforma- larly influence the dynamics of fluid flow in pressure-driven
tion repeatedly regenerating permeability. Accordingly, flow regimes in the seismogenic upper crust (Sibson et al.,
fluid flow is localized within structures that were active dur- 1988; Cox, 1999; Sibson, 2001).
ing the operation of hydrothermal systems. Buoyancy drive for crustal fluid flow arises from vertical
variations in fluid density, either through effects of tempera-
Driving forces for fluid flow ture gradients or variable concentrations of dissolved species
Fluid flow occurs in response to two major classes of dri- (e.g., salinity) in pore fluids. Decreasing density of hydro-
ving forces: (1) pressure-driven flow, and (2) buoyancy- thermal fluid with increasing temperature and depth for
driven flow. many pore fluid compositions leads to gravitational instabil-
Pressure-driven flow arises from a number of causes. Par- ity and convection of fluid through porous media in near-
ticularly at crustal depths less than several kilometers, rock hydrostatic fluid pressure regimes. An excellent example of

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 3

this is provided by thermally driven convection around hot


intrusive complexes, as modeled by Norton and Knight
(1977), Forster and Smith (1990), and Cathles et al (1997).
Flow in porous media
In addition to a driving force, crustal fluid flow also
requires a rock mass with pore spaces to contain fluid, as
well as connectivity between pore spaces. The pore spaces
can be intergranular pores or grain-scale to macroscopic
fractures. The one-dimensional macroscopic flux of a sin-
gle-phase fluid in a horizontal plane through an isotropic,
porous medium is described by a simplified form of Darcy’s
law, which states that

q = Q/At = k/η (dP/dx), (1)

where
q is the fluid flux (i.e., the volume Q of fluid traversing
cross-sectional area A perpendicular to the x axis, per
unit time t);
k is the permeability of the porous medium;
a η is the kinematic viscosity of the pore fluid; and
dP/dx is the horizontal fluid pressure gradient that drives
flow (Bear, 1972; Fig. 2).

The volume flux has dimensions of velocity, and is some-


times referred to as the Darcian velocity. Note, however,
that the Darcian velocity is not the actual velocity at which
the fluid moves through the pore space. Rather, it is a mea-
sure of the fluid volume per unit time moving through unit
cross-sectional area of the rock. It is related to the real aver-
age fluid velocity, v, by the relationship,

q = v·Q. (2)

Particularly in low-porosity rocks, the real fluid velocity at


the pore scale can be much larger than the Darcian veloc-
ity. Darcy’s law assumes laminar flow, and applies in porous
rocks at flow rates up to about 1 m/s. At higher flow rates,
turbulence and high inertial forces lead to breakdown of
Darcy’s law (Guéguen and Palciauskas, 1994).
Permeability has units m2 (1 darcy = 10–12 m2), and is an
intrinsic rock property quantifying the capacity of fluids to
pass through rock. Permeability is influenced by the con-
nectivity between intergranular pore spaces or fractures in
b a rock, and is particularly sensitive to the minimum throat
size between connecting pores. Natural rocks have perme-
FIG. 1. Numerical (finite element) models simulating steady state, pres- abilities that range over more than ten orders of magnitude
sure-driven fluid flow patterns around permeable faults or shear zones (Brace, 1990; Manning and Ingebritsen, 1999). For exam-
embedded in a less permeable medium. A vertical lithostatic fluid pressure ple, porous sandstone (porosity, φ > 15%) and poorly com-
gradient is maintained in the medium away from the fault. Length of flow pacted tuffs can have permeabilities as high as 10–12 m2,
vectors corresponds to flow velocity; flow vectors within faults not shown.
Contours indicate departures of fluid pressure from lithostatic values: light whereas unfractured, “tight” metamorphic or igneous rocks
areas are below and dark areas are above lithostatic pressure. a. Simple pla- can have permeabilities less than 10–22 m2. Fluid fluxes and
nar fault or shear zone with a permeability 103 times that of the surround- the geometry of flow are, therefore, particularly dependent
ing host rock matrix. Note fluid focusing at the upstream (lower) part of on spatial and time variations in permeability within the
the structure, and fluid discharge around the downstream (upper) part of
the structure. b. Fluid flow patterns associated with the presence of a fault
crust. We examine below the critical effect of deformation
stepover region. Permeability and imposed fluid pressure gradients are processes in generating large, although commonly tran-
the same as in (a). sient, changes in rock permeability.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
4 COX ET AL.

pressure, whereas the ρ g(∂ z/∂ xj) term accounts for gravity
DARCY’S LAW and also buoyancy effects associated with spatial variations
in fluid density. This full Darcian relationship illustrates
that the driving force for fluid flow is the difference
Q/At = k/ η (dP/dx) between the fluid pressure gradient and the hydrostatic
component of the fluid pressure gradient. This result is
illustrated by considering fluid pressure variations in a ver-
tical column of fluid in a porous medium with fully con-
nected porosity. The depth dependence of fluid pressure Pf
in a column of fluid at rest is given by
fluid volume, Q
Pf = ρ gz, (4)

area, A
where z is the depth below the surface.
driving pressure
In this case, the vertical fluid pressure gradient, ρ g,
FLUID PRESSURE

gradient,dP/dx
P1 owing to the weight of the fluid column does not drive flow,
and is known as the hydrostatic fluid pressure gradient.
Deviations in dP/dz from ρ g are required to drive vertical
P2 flow. Accordingly, for vertical flow of a uniform density
fluid, Darcy’s law reduces to

DISTANCE, x q = (k/η )·(dP/dz – ρ g). (5)


FIG. 2. Darcy’s law and control of horizontal fluid flow rate by driving
pressure gradient, rock permeability, and fluid viscosity. Where dP/dz is greater than ρ g, upwards flow occurs,
whereas if dP/dz is less than ρ g, downwards flow occurs in a
fluid of uniform density. For general pressure-driven flow in
The Darcy equation also illustrates how fluid flux is rocks with isotropic permeability, flow is parallel to the non-
inversely proportional to the kinematic viscosity of the pore hydrostatic component of the fluid pressure gradient. Per-
fluid. At temperatures between 100° and 800°C, and pres- meability anisotropy may result in flow that is not parallel to
sures between 50 and 300 MPa, the viscosity of water ranges the driving pressure gradient.
over about one order of magnitude (40–400 µ Pa·s),
decreasing with increasing temperature and decreasing Types of porosity
pressure (Haar et al., 1984). Viscosity increases with increas- Porosity, φ , is the ratio of the volume of void space to the
ing salinity (Garven and Freeze, 1984). The extreme range bulk volume of a porous medium,
in permeability in rocks compared with the much smaller
range in viscosity in typical ore fluids, means that variations φ = Vpores/(Vpores + Vsolids). (6)
in permeability usually have a greater impact on fluid flux
than variations in viscosity in simple water-dominated Permeability is not always simply related to porosity. It is
hydrothermal systems. The dynamics of two-phase flow in dependent on various factors, including pore diameters
porous media can lead to complex effects that will not be and shapes, diameters of pore throats (which connect large
treated in this review (Sahimi, 1994). intergranular pores), and connectivities between pores.
For fluid flow with a vertical component of motion (z From the standpoint of flow through a porous medium, the
axis), the driving force for flow is not the absolute fluid interconnected pore volume is of interest.
pressure gradient, but the nonhydrostatic component of Porosity usually comprises intergranular porosity and
the gradient. An expanded expression for Darcy’s law is fracture porosity. Intergranular porosity, such as primary
given by the relationship, pore spaces in poorly cemented or weakly compacted clas-
tic sediments and pyroclastic materials, can control fluid
qi = (kij/η )·(∂ P/∂ xj – ρ g(∂ z/∂ xj)), (3) flow in shallow, low-temperature hydrothermal systems.
Examples include some carbonate-hosted Pb-Zn systems
where (Garven, 1985) and some parts of shallow, volcanic-related
qi is the fluid flux in the coordinate direction i; geothermal systems. Intergranular porosity, although usu-
kij is the permeability tensor (allowing for anisotropy of per- ally very low, may also be present in metamorphic regimes
meability); (Holness, 1997). Reaction-enhanced porosity is associated
ρ is the fluid density; and with volume changes during metamorphic reactions at ele-
g is the gravitational acceleration (Forster and Smith, vated temperatures and pressures (Rumble et al., 1982;
1990); Zhang et al., 2000). This effect is particularly important in
∂ P/∂ xj represents the driving force due to gradients in fluid controlling grain-scale fluid infiltration in skarn environ-

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 5

ments, and pervasive hydrothermal alteration associated


with fluid discharge from fracture-controlled hydrothermal a
systems into relatively low-permeability wall-rocks. Volume
change between the solid reactants and reaction products
tends to produce transient changes in intergranular poros-
ity and permeability in these regimes. Macroscopically duc-
tile deformation in shear zones can be associated with grain
translation and grain-scale cracking, which also enhances
intergranular porosity. Similarly, in low-temperature brittle
deformation regimes, cataclasis in active fault zones may
modify the intergranular porosity of the host rocks.
In rocks with low, or poorly connected intergranular
porosity, fluid transport is controlled by fracture porosity.
Fractures can be present at scales ranging from grain-scale
to macroscopic (> tens of meters long). Fracture formation backbone pores
is usually controlled by the stress and fluid pressure states
pore throats
during deformation of a rock mass (see “Principles of dangling pores
Macroscopic Fracture Formation” below), but may also be
influenced by thermal shocking. isolated pores

Porosity-permeability relationships
Low temperature regimes—Clastic rocks and fault rocks: In low-
temperature crustal regimes where primary intergranular
porosity can survive for geologically long periods, perme-
ability exhibits a simple relationship with porosity, espe- b 10-11

cially at high porosities. For porosity greater than about 6 to


10 percent, pores are generally fully interconnected and 10-12 k∝φ 3
porosity is related to permeability by a relationship of the
PERMEABILITY, m 2

form
10-13
k ∝ φ n, (7)

where n is approximately 3. In this regime, permeability 10-14


decrease with decreasing porosity is controlled by pore
shrinkage. At lower porosities, progressive loss of connectiv-
10-15
ity between pores occurs through closure of narrow throats
between the larger intergranular pores. Pores that form part
of a connected network that spans the sample/system com-
10-16
prise what is known as the backbone porosity (Fig. 3a). All percolation threshold
flow is localized along this backbone pore network; accord-
ingly, the backbone porosity controls the permeability. The 10-17
remaining porosity comprises two elements: (1) isolated 0 0.05 0.10 0.15 0.20 0.25
pores, which do not connect with the backbone porosity, TOTAL POROSITY
even though they may form localized clusters of pores that
are connected to each other; and (2) dangling, or dead-end FIG. 3. a. Schematic illustration of classes of intergranular pores. Flow
pores, which connect to the backbone porosity from one rate is controlled by the backbone porosity, rather than by total porosity or
side only. Although they are part of the interconnected total connected (i.e., backbone + dangling) porosity. b. Relationship
between total porosity and permeability in Fontainebleau sandstone (after
porosity and contain pore fluid, they do not contribute to Bourbié and Zinszner, 1985). The percolation threshold porosity and the
the flow. At low porosities, permeability changes are directly relationship k ∝ φ 3 are indicated.
related to progressive changes in the relative proportions of
backbone, dangling, and isolated pores, as well as the total
porosity of a rock. threshold. Where permeability is controlled by approxi-
The relationship, k ∝ φ 3, breaks down where the back- mately equant intergranular pores, the percolation thresh-
bone porosity becomes less than the total porosity. At lower old typically occurs at total porosities in the range 3 to 6
porosity, permeability typically becomes more sensitive to percent in many types of clastic rocks.
small changes in porosity (Fig. 3b). The porosity at which Above the percolation threshold, permeability associated
all pore connectivity is lost (i.e., φ backbone = 0, φ isolated = 1) with intergranular porosity is influenced only moderately
and permeability vanishes, is known as the percolation by confining pressure (David and Darot, 1989). Permeabil-

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
6 COX ET AL.

ity decreases with increasing confining pressure (Fig. 4) in a grain B


fluid
response to elastic and inelastic deformation, causing pore
shrinkage and collapse. Note that at constant confining Θ
pressure, changes in pore fluid pressure (i.e., changes in fluid
effective confining pressure) lead to relatively small
changes in permeability. Poroelastic effects on total inter- grain A grain C grain A grain C
granular pore volume, driven by changes in stress states
associated with fault rupture events and interseismic stress
recovery, may play a role in episodic fluid redistribution
around active fault zones (Muir-Wood and King, 1993; Sib-
son, 1993). Such effects are likely to be a small part of the
overall fluid budget in a high-flux hydrothermal system. b c
The relationships between porosity, permeability, and
effective stresses, found for clastic sedimentary rocks, will
apply also to granular damage products (gouge, cataclasite,
and breccia) produced in fault zones.

High temperature regimes—metamorphic rocks: In high-


temperature metamorphic regimes, atomic diffusion,
creep, grain boundary migration, and chemical reactions
are fast on geologic timescales. In this case, intergranular
porosity and pore connectivity in rock-fluid systems is con-
FIG. 5. Schematic illustration of pore geometries in polycrystalline grain
trolled principally by interfacial surface energy effects, pro- aggregates at elevated temperatures where pore shapes are controlled by
vided that the polycrystalline aggregate is not deforming. surface energy minimization. a. Cross section through a grain-edge chan-
Where pore shapes are controlled by surface energy mini- nel; dihedral wetting angle, Θ , is indicated. b. For dihedral wetting angles
mization, there is a balance of surface tension forces along ≤ 60°, fluid forms isolated pockets on two-grain interfaces, and fluid chan-
the surfaces where two solid grains and a fluid phase meet, nels along three-grain edges may connect fluid pockets at four-grain cor-
ners. c. For dihedral wetting angles >60°, fluid on two-grain interfaces,
and we can define a dihedral wetting angle, Θ , which is the three-grain edges, and four-grain corners occurs as isolated pockets (after
angle between the two solid–fluid interfaces (Fig. 5a). The Watson and Brenan, 1987).
magnitude of the wetting angle is controlled by the balance
of interfacial forces expressed by the relation,
The three-dimensional connectivity of intergranular
Θ = 2 arcos(γ s–s/2γ s–fl), (8) pores in isotropic mineral aggregates is dictated by interfa-
cial wetting angles and total porosity. We distinguish
where γ s–s and γ s–fl are the solid–solid and solid–fluid sur- between pores that form at two-grain interfaces, along
face energies per unit area, respectively (Smith, 1964). three-grain edges, and at grain corners (Fig. 5). For dihe-
dral wetting angles greater than 60°, pores at two grain
interfaces tend to form isolated pockets. At low porosity,
grain-edge channels pinch off to form discontinuous beads
along grain edges, and pores at grain corners tend to be
2 isolated (Fig. 5c). In contrast, at dihedral wetting angles less
PERMEABILITY, x 10 -15 m2

than or equal to 60°, connectivity between pores at grain


corners may be provided to low porosities by continuous
1.8
channels along grain edges (Fig. 5b). For Θ greater than 0°,
the equilibrium fluid distribution on two-grain interfaces is
disconnected bubbles. Only in the extreme case of Θ equal
1.6
to 0° can a continuous fluid film exist stably on two-grain
40 MPa interfaces. Experimental studies of interfacial wetting
30 MPa
20 MPa angles for common minerals and pore fluids indicate that
1.4
10 MPa for some common mineral–fluid systems, dihedral wetting
angles are greater than 60° (Holness, 1997). For example,
1.2 in the quartz-H2O-CO2 system at elevated pressures and
10 20 30 40 50 60 70 temperatures in the range 950° to 1,150°C, dihedral wet-
ting angles are typically greater than 60° unless high solute
CONFINING PRESSURE, MPa concentrations are present in the pore fluid (Watson and
FIG. 4. Relationship between permeability and confining pressure, at Brenan, 1987).
various pore fluid pressures, in Fontainebleau sandstone with porosity of The evolution of permeability with decreasing porosity at
0.06 (after David and Darot, 1989). elevated temperatures, where surface energy effects control

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 7

pore geometry, is illustrated by high-temperature isostatic


compaction experiments on calcite aggregates (Zhang et
a 10-15

al., 1994a). The overall porosity-permeability relationship is 10-16


similar to that for clastic sedimentary rocks (Fig. 6a). For

PERMEABILITY, m 2
the calcite-argon system, the dihedral wetting angle is 10-17
greater than 60°, so the situation is analogous to many min-
eral–fluid systems at elevated temperatures. The porosity- 10-18
permeability relationship approximately follows the cube
law (i.e., n = 3 in equation 7) to porosities down to about 10-19
0.06, where full connectivity between pores starts to be lost
(Fig. 6b). Below this porosity, permeability is much more 10-20
strongly dependent on total porosity. The critical porosity,
φ c, at which pores completely lose connectivity (i.e., the per- 10-21
colation threshold) occurs at a porosity of 0.04. For min- 0 0.05 0.1 0.15 0.2
eral–fluid systems with wetting angles less than 60°, the per- TOTAL POROSITY
colation threshold will occur at porosities less than 0.04.
Near φ c, the permeability of the intergranular pore network
obeys the scaling law (Knackstedt and Cox, 1995),
b 0.16
experimentally measured
k ∝ (φ – φ 2
c) . (9)

ACCESSIBLE POROSITY
model backbone porosity
0.12
For high-temperature, isostatically stressed (i.e., all prin-
cipal stresses equal) mineral–fluid systems, where pore
geometry is controlled by minimization of interfacial sur-
0.08
face energy, dihedral wetting angles have a profound effect
on fluid transport. For wetting angles greater than about
60°, pore connectivity is lost at porosities of several per-
0.04
cent. Accordingly, many metamorphic fluid–rock systems,
which have porosities less than one percent, will be below
the percolation threshold and effectively impermeable,
0
unless deformation processes actively generate fracture 0 0.04 0.08 0.12 0.16
networks.
TOTAL POROSITY
Fracture-controlled fluid flow
FIG. 6. a. Porosity–permeability relationships for calcite grain aggre-
In hydrothermal systems where compaction, pore sealing gates, isostatically hot-pressed at temperatures between 360° and 560°C.
processes, or diffusion-controlled surface energy effects Confining pressures ranged from 200 to 300 MPa and argon pore fluid
have driven intergranular porosity below the percolation pressures ranged from 100 to 250 MPa (after Zhang et al., 1994a). b. Rela-
tionship between total porosity and connected (or accessible) porosity
threshold, fluid migration is dependent on the generation experimentally determined during isostatic hot-pressing of calcite aggre-
of fracture porosity. gates. Also shown are numerically modeled values of the backbone poros-
ity. Modified after Zhang et al. (1994a) and Knackstedt and Cox (1995).
Macroscopic to microscopic fractures at low temperatures: For
steady state, laminar, incompressible flow in an ideal, hori-
zontal, parallel-sided plane fracture, the volume flow rate In reality, relationships between fracture apertures and
(m3/s) is given by the relationship, flow rates are more complicated owing to the effects of frac-
ture roughness leading to aperture changes and tortuous
Q = wb3/12η ·(dP/dx), (10) flow paths. Experimental studies provide clear evidence of
surface roughness causing departures from the cubic law in
where natural rough fractures (Witherspoon et al., 1980; Tsang
b is the fracture aperture; and Witherspoon, 1981; Brown, 1987). Numerical model-
w is the width of the fracture (measured in the fracture ing of flow through rough fractures also confirms that frac-
plane); and ture roughness and tortuosity lead to substantial departures
dP/dx is the fluid pressure gradient along the fracture from the cubic law (Waite et al., 1999).
(Guéguen and Dienes, 1989). Because cracks have high aspect ratios, elastic opening or
By analogy with Darcy’s law, for a fracture with cross- closure of cracks in response to changes in effective stress
sectional area wb, the equivalent fracture permeability is states has important effects on crack aperture and perme-
given by ability (Walsh, 1981). The dependence of fracture perme-
ability on effective confining pressure (i.e., confining pres-
k = b2/12. (11) sure – fluid pressure), illustrated in Figure 7, shows that the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
8 COX ET AL.

60 tems, and is discussed below in “Evolution of Flow Pathways


PERMEABILITY, x10 -14 m2

during Deformation.”
40 In summary, for flow in fracture-controlled hydrothermal
systems, the highest flux will occur where and when (1)
fracture apertures are highest, (2) fracture density is high-
20 est, and (3) fracture connectivity is highest.

Permeability during macroscopically ductile deformation:


10 Experimental results, together with field observations of
8 fluid-rock interaction in exhumed shear zones, indicate
that permeability enhancement is commonly associated
6 with macroscopically ductile deformation. Experimental
4 studies indicate that macroscopically ductile deformation at
0 50 100 150 200 elevated temperatures and confining pressures is only asso-
ciated with permeability enhancement when a component
EFFECTIVE CONFINING PRESSURE, MPa of strain (albeit small) is accommodated by microfractur-
ing, which generates grain-scale fluid pathways.
FIG. 7. Dependence of crack permeability on effective confining pres-
sure (after Guéguen and Palciauskas, 1994). The evolution of permeability during deformation
involving both intragranular plastic deformation and grain-
scale crack growth is illustrated by experimental studies on
pressure dependence is substantially greater than for mate- calcite rocks (Zhang et al., 1994b). This work demonstrates
rials where permeability is controlled by approximately that for low effective confining pressures, permeability
equant intergranular pores (Fig. 4). Accordingly, low poro- increase with increasing strain can be very rapid and large
elastic and inelastic strains in cracked rocks will be more (Fig. 8a, b). For example, at an effective confining pressure
effective than deformation of intergranular pores in chang- of 30 MPa, permeability increases by two orders of magni-
ing permeability, and driving fluid redistribution in tude with three percent shortening, and increases by a fur-
response to stress changes (e.g., around active faults). ther order of magnitude after ten percent shortening. Only
For a cracked rock with fracture porosity φ f, and where small increases in permeability occur with higher strains,
penny-shaped cracks are fully interconnected and ran- however.
domly oriented, permeability k is given by Major increase in permeability, at strains as low as a few
percent, is associated with growth of microcracks and
k = a2φ f/3, (12) rapid development of connectivity in grain-scale crack net-
works (Fig. 9a). Such behavior persists well into the domi-
where a is the average crack half-aperture. For an average nantly crystal plastic deformation regime, provided that
crack radius, r, and average crack spacing, l, fluid pressures are high enough to facilitate some defor-
mation by microcracking (Fischer and Paterson, 1992;
φ f = 2π r2a/l3, (13) Stormont and Daemen, 1992; Zhang et al., 1994b; Peach
and Spiers, 1996). As both temperature and effective con-
and so (Guéguen and Palciauskas, 1994), fining pressure increase, brittle/frictional processes are
impeded and intracrystalline plasticity is favored. So, crack
k = 2π r2a3/3l3. (14) growth rates reduce, and the critical strain required to
develop a well-connected, high-permeability crack network
This relationship holds true for fractured media where increases.
fractures are developed at microscopic to macroscopic The experimental studies demonstrate that where high
scales. It illustrates how both fracture density and fracture fluid pressures produce low effective confining pressures,
connectivity play a critical role in controlling permeability. grain-scale crack growth significantly increases the perme-
For randomly distributed fractures that are not fully con- ability of active shear zones relative to their host-rocks,
nected, permeability is given by even though most displacement may occur by microscop-
ically ductile deformation mechanisms such as dislocation
k = (4π /15)·fa3r2/l3, (15) flow and dissolution-precipitation creep. A significant
aspect of the experimental work is that fracture networks
where connectivity, f, is 0 ≤ f ≤ 1 (Guéguen and Dienes, can develop high crack connectivity and high permeabil-
1989). As in the case for a single fracture, this relationship ity at very low strains. One implication of this result is that
illustrates how the permeability of a cracked medium is very low strain deformation, especially when localized in net-
sensitive to average fracture aperture. The evolution of con- works of faults, shear zones, and associated fracture arrays,
nectivity among elements of networks of fractures, faults, may have a big impact on the localization of fluid flow
and shear zones, and its influence on flow architecture, is (see “Evolution of Flow Pathways during Deformation”
an important aspect of the evolution of hydrothermal sys- below).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 9

a 0.08
50 MPa
relationships in veins in shallow epithermal systems and
deeper mesothermal systems commonly indicate multiple
CONNECTED POROSITY

0.06
episodes of fracturing and vein sealing (Fig. 9c). Experi-
30 MPa
mental studies of fracture healing (Hickman and Evans,
1987; Brantley et al., 1990; Zhang et al., in press), also
0.04 demonstrate that, at temperatures greater than approxi-
100 MPa
mately 300°C, diffusional crack healing and associated loss
0.02 of crack connectivity in minerals such as quartz and calcite
can be fast, even on laboratory timescales.
Permeability evolution in faults in the seismogenic upper
0 crust can be influenced strongly by porosity changes associ-
ated with deformation during repeated, episodic slip events
-0.02 (over several seconds), or during aseismic creep. In rocks
0 5 10 15 20 with high initial intergranular porosity, fault slip is associated
STRAIN, % with reduction in porosity and permeability owing to the
production of fine-grained, compact, low-porosity gouges
b 10-16 (Knipe, 1998; Zhang et al., 1999). In this case, faults become
seals or aquitards in hydrothermal systems. However, in
30 MPa intrinsically low-porosity, tight rocks (e.g., metamorphic and
PERMEABILITY (m2)

10-17
igneous rocks), faulting leads to porosity and permeability
increase during slip, owing to fracture growth and cataclasis.
50 MPa
10-18 In contrast, porosity evolution during interseismic intervals is
governed by progressive porosity destruction by pore col-
10-19
lapse and hydrothermal sealing (Angevine et al., 1982; Cox
100 MPa
and Paterson 1991; Cox, 1995). Episodic slip and interseis-
mic sealing may result in large, time-dependent changes in
10-20 permeability in fault zones (Fig. 10). Particularly in fracture-
controlled hydrothermal systems, flow within permeable
10-21 faults and associated fracture arrays (Fig. 1) promotes rapid
0 5 10 15 20 pore sealing by mineral deposition. As soon as porosity drops
STRAIN, %
below the percolation threshold, fluid flow will shut off. An
important conclusion then, is that in the immediate post-
FIG. 8. Relationship between (a) porosity and strain, and (b) perme- rupture phase, faults in low-permeability host rocks are
ability and strain, as a function of effective confining pressure during highly permeable structures that act as fluid conduits. How-
deformation of Carrara marble at room temperature and in the presence ever, with progressive interseismic sealing and loss of perme-
of argon pore fluid. Relationships at effective confining pressures of 30,
50, and 100 MPa are illustrated. Confining pressure = 300 MPa; nominal ability, faults can become aquitards until a later slip event.
strain rate = 1.2 × 10–4 s–1 (after Zhang et al., 1994b). Accordingly, in active hydrothermal regimes, permeability
is rapidly destroyed unless ongoing deformation regenerates
permeability. Additionally, for permeable fluid pathways to
Competition between deformation-induced permeability develop, the rate of deformation-induced permeability
enhancement and permeability reduction processes enhancement must be greater than the rate of permeability
reduction owing to closure and sealing of pores and fractures.
At depth in the Earth’s crust, and especially at elevated
temperatures in active hydrothermal systems, porosity- Episodic versus continuous flow
destruction processes such intergranular cementation, Competition between porosity-creation processes and
compaction, and healing and sealing of fractures can cause porosity-destruction processes in actively deforming rocks
permeability to decrease on timescales that are short rela- results in a contrast between flow regimes in the upper
tive to the lifetimes of hydrothermal systems. Permeability crustal seismogenic regime and the lower crustal aseismic
evolution is, therefore, controlled directly by competition regime. At depths below the seismic–aseismic transition
between deformation-induced porosity-creation processes (typically 10–20 km deep), where steady state creep
and various porosity-destruction processes. processes usually dominate, a balance between rates of
The structure of veins in ore systems provide spectacular porosity destruction and porosity creation in creeping
evidence of repeated fracturing and fracture sealing in shear zones is expected to generate quasi-steady state per-
hydrothermal systems. For example, crack-seal microstruc- meabilities that are higher than those of the surrounding,
tures in veins (Ramsay, 1980; Cox, 1995) indicate that less rapidly deforming rock mass. This leads to essentially
macroscopic fractures in some deep hydrothermal regimes continuous fluid flow along actively creeping elements of
can open and seal up to several thousand times in the life- shear networks. Creeping faults and shear zones probably
time of one active vein (Fig. 9b). Similarly, overprinting play a key role in focusing the migration of deeply sourced

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
10 COX ET AL.

FIG. 9. a. Grain-scale crack networks produced in Carrara marble during macroscopically ductile deformation at room
temperature and an effective pressure of 100 MPa. The shortening direction is parallel to the micrograph long axis. Inci-
dent light micrograph. b. Quartz-rich extension vein with dark, crack-seal inclusion bands composed of tourmaline. This
texture indicates multiple episodes of extensional fracturing and fracture-sealing associated with gold mineralization,
Sigma Mine, Val d’Or, Quebec. Transmitted light micrograph. c. Laminated fault-fill veins indicating multiple episodes
of fault slip and dilation during gold mineralization. Revenge gold deposit, St Ives goldfield, Eastern Goldfields Province,
Western Australia.

fluids upwards to the base of the seismogenic regime. In these structures is inherently influenced by the orientations
contrast, above the seismic–aseismic transition, potentially and relative magnitudes of stresses in hydrothermal systems.
large, cyclic changes in fault permeability cause episodic The distribution of these structures is also controlled by fluid
fluid flow. Episodic flow is associated with fluid pressure pressure regimes. This section outlines the critical influence
cycling and episodic fluid redistribution around active of both fluid pressures and stress states in controlling macro-
faults, and has important implications for the dynamics of scopic fracture growth and associated permeability genera-
flow and reaction in hydrothermal systems in the seismo- tion in hydrothermal systems. Importantly, macroscopic frac-
genic regime (Cox, 1999; Sibson, 2001). ture systems can develop not only in the brittle upper crust,
but also in the more ductile deeper crust, provided fluid
Principles of Macroscopic Fracture Formation pressures are high enough.
The development of faults, shear zones, and associated
fracture arrays plays a key role in controlling the local per- Types of fractures and orientation relationships
meability distributions and the macroscopic architecture of with stress fields
fluid pathways in hydrothermal systems. The geometry of Three classes of macroscopic fractures may form during

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 11

regimes during crustal deformation produce distinctive ori-


a entations of shear fractures (faults) and spatially associated
extension fractures (and veins; Fig. 13). The relative orien-
tations of faults and extension veins provide one of the
DISPLACEMENT

most powerful ways of determining shear sense on struc-


tures that were active during hydrothermal mineralization
(Robert and Poulsen, 2001).
Faults and associated fracture arrays exhibit heteroge-
neous distributions of fracture apertures and fracture den-
sities at mesoscopic to macroscopic scales. Areas with the
highest apertures and/or fracture densities produce the
highest permeability sites, which localize most fluid migra-
EQ EQ EQ EQ
tion if they connect to fluid reservoirs. High fracture aper-
tures and densities are typically associated with dilatant
TIME fault bends, stepover regions, or jogs that link approxi-
mately planar segments of faults and shear zones (Figs. 11d
b and 14a). High fracture densities and apertures can also
develop in areas of competence contrast (Oliver et al.,
LOG PERMEABILITY

2001) and during dilation at fold hinges during flexural


slip folding (Cox et al., 1991). High fracture permeability
can also localize around fault termination zones, where
fault splays, wing cracks, or brecciated regions develop (Fig.
14b). Contractional jogs, as well as dilational jogs, can be
sites of high fracture density, which localize fluid flow and
ore deposition.
The orientation relationships between shear fractures,
jogs, and slip directions result in the long axis of jogs devel-
oping approximately parallel to the σ 2 orientation (Fig.
EQ EQ EQ EQ
14c). This produces a permeability anisotropy favoring flow
parallel to the jog axis. For reverse and normal faulting
TIME regimes, extensional and contractional jogs have subhori-
zontal plunges (Fig. 14d). This produces good horizontal
FIG. 10. Schematic illustration of time-dependent changes in (a) poros-
ity and (b) permeability in fault rocks during the seismic cycle. Sudden fracture connectivity within faults, and may cause ore shoots
permeability enhancement is associated with episodic fault slip (EQ). to have gentle plunges. In contrast, strike-slip regimes pro-
Interseismic pore sealing and compaction reduces permeability between duce jogs with good vertical connectivity that can control
slip events. After Cox (1995). the geometry of steeply plunging ore shoots (Fig. 14d).
Stress magnitudes and fracture formation
brittle deformation: (1) pure extension fractures; (2) shear The types of fractures that control permeability in hydro-
fractures; and (3) hybrid extensional-shear fractures. Mineral thermal systems in the brittle regime are governed by the mag-
filling in each of these fracture types produces veins (Fig. 11). nitudes of stress differences (σ 1 – σ 3). The stress states in rocks,
Any stress field can be resolved into three mutually per- and the relationships between stress magnitudes, stress differ-
pendicular components, which are known as the maximum ences, fracture type, and fracture orientation can be illustrated
principal stress (σ 1), the intermediate principal stress (σ 2), by two-dimensional Mohr circle constructions.
and the minimum principal stress (σ 3). Pure extension The stresses acting on a plane inclined to σ 1 at an angle
fractures form perpendicular to σ 3 and open perpendicu- α can be resolved into a normal stress (σ n) acting perpen-
lar to the fracture wall (Fig. 12a). In intact, isotropic rock, dicular to the plane, and a shear stress (τ s) acting parallel to
shear fractures (i.e., faults) tend to form at angles typically the plane (Fig. 15a). The magnitudes of the normal stress
between 20° and 35° to the orientation of σ 1, such that the and shear stress are given by the relationships,
fault plane contains the orientation of σ 2 (Fig. 12b). Two
conjugate sets of shear fractures can form, and where devel- σ n = 1/2(σ 1 + σ 3) + 1/2(σ 1 – σ 3)·cos 2α (16)
oped, their intersection is subparallel to σ 2. Hybrid exten- and
sional-shear fractures have components of displacement
both parallel and perpendicular to the fracture plane. They τ s = 1/2(σ 1 – σ 3)·sin 2α . (17)
form at angles between 0° and approximately 25° to the σ 1
direction. A Mohr diagram plots σ n against τ s, and has the geomet-
The orientations of the maximum principal stresses asso- ric property that the normal and shear stresses acting on a
ciated with contractional, extensional, and wrench tectonic plane inclined at angle α to σ 1 are given by the coordinates

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
12 COX ET AL.

FIG. 11. Examples of vein types classified according to fracture modes. a. Pure extension veins; opening direction is
perpendicular to the fracture walls. Mt Lyell, Tasmania. Coin is 2 cm in diameter. b. Hybrid extensional-shear vein, with
opening direction inclined to the fracture wall indicated by orientation of displacement-controlled crack-seal quartz
fibers. This vein is associated with smaller, subhorizontal extension veins. Mt Lyell, Tasmania. c. Gently left-dipping exten-
sion veins spatially associated with gently right-dipping fault-fill veins, within dilatant segments of two faults associated
with gold mineralization. The orientation relationship between the subhorizontal extension veins and the fault veins
indicates that mineralization formed during reverse faulting. Victory gold mine, St Ives goldfield, Western Australia.
Field of view is 6 m wide. d. Calcite-filled dilatant jogs developed on stepovers between small, steeply dipping, sinistral
wrench faults. Les Matelles, Languedoc, France.

of the point P on the perimeter of the circle having diame- line BC in Figure 15a. This line specifies the brittle shear
ter (σ1 – σ3) and center at (σ1 + σ3)/2 on the normal stress failure strength of the medium as a function of normal
axis (Fig. 15a). This circle is called the stress circle (or stress. So, in dry, intact rock, brittle shear failure occurs
Mohr circle). Note that P is on a radius inclined at 2α to the where and when the Mohr circle (describing the shear
normal stress axis. stress and normal stress state in a rock) contacts the failure
envelope BC. Shear failure may, therefore, be induced by
Brittle shear failure: Under fluid-absent conditions, the decreasing σ3 and/or increasing the value of σ1 (stress cir-
stress state that causes compressional brittle shear failure cle A in Fig. 15b). Because the angle between the Mohr cir-
(i.e., faulting) in an intact, isotropic medium is approxi- cle radius (AP) and the normal stress axis is 2α, the angle α
mated by the relationship, between the shear failure plane and the orientation of the
maximum principal stress is given by the relationship,
τ = C + µσn, (18)
α = (90° – arctan µ)/2. (19)
where
C is the cohesive strength of the medium; and For typical friction coefficients of approximately 0.75,
µ is the coefficient of friction. shear fractures are inclined at approximately 27° to σ1.
This Coulomb shear failure criterion is indicated by the Shear fractures formed in intact rock and obeying rela-

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 13

a EXTENSION FRACTURES a REVERSE FAULT

σ 3

σ 1 σ 1

σ 1

subhorizontal extension veins


σ 3

b SHEAR FRACTURES σ 2
b NORMAL FAULT

σ 1
subvertical extension
veins

σ 3

FIG. 12. Orientations of fracture types with respect to principal stress c WRENCH FAULT
directions, σ 1, σ 2, and σ 3. a. Pure extension fractures. b. Conjugate shear
fractures, with sense of shear indicated. σ 3
subvertical extension
veins
tionship (19) are termed “optimally oriented” faults (Sib-
son, 1985). Activation, or re-activation, of misoriented
structures may also be important in controlling fracture
permeability in hydrothermal systems. For example, mis- σ 1
oriented faults may develop by rotation away from optimal
orientations with respect to the stress field, due either to
rotation of stress fields and/or rock masses during pro-
gressive deformation. Additionally, preexisting faults, or
σ 3
other mechanical anisotropies such as bedding or folia-
tion, can also be (re-)activated. Shear (re-)activation
occurs when the shear stress on the misoriented plane sat- FIG. 13. Orientations of faults and associated extension veins relative to
isfies the Mohr-Coulomb failure criterion (Fig. 15c). Fig- the principal stress directions for contractional (a), extensional (b), and
ure 15c has been constructed with the cohesive strength, strike-slip (c) tectonic regimes.
C, of the preexisting fault or anisotropy less than that of
intact rock. Note, however, that rapid sealing or com- and shear stress state (Mohr circle) contacts the nonlinear
paction of gouge on fault slip patches in hydrothermal sys- part of the failure envelope at negative (i.e., tensile) nor-
tems may cause some faults to quickly regain cohesive mal stresses. Pure extension failure occurs where the nor-
strength (Kanagawa et al., 2000). mal stress equals the tensile strength (T) of the rock (stress
circle C in Fig. 15b). Note that the angle 2α is zero, as
Extension and hybrid extensional-shear failure: Extension fail- expected for extension failure along the σ 1-σ 2 plane, per-
ure and extensional shear failure occur if the normal stress pendicular to σ 3. Hybrid extensional shear failure occurs

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
14 COX ET AL.

a c long axis of dilatant site


dilational jog slip direction

jog

contractional jog σ 2

σ 1
σ 3

d normal fault
b wing cracks
reverse fault

contractional splays
wrench fault

FIG. 14. a. Geometry of contractional and dilatant jogs. b. Wing cracks and contractional splays developed around
fault terminations. c. Geometry of a dilatant jog relative to fault slip direction. d. Orientations of jogs in contractional,
extensional, and wrench tectonic regimes. High fracture density, aperture, and connectivity along jogs favors high fluid
flux along the jog axis.

where the stress circle contacts the failure envelope In terms of the Mohr circle representation of stress
between –T and C (stress circle B in Fig. 15b). states, the role of fluid pressure is to move the stress circle
Pure extension failure only occurs at relatively small to the left (Fig. 16). Note that although fluid pressure mod-
stress differences, typically less than approximately 4T. For ifies normal stress, it does not influence shear stress.
4T < (σ 1 – σ 3) < 6T, failure occurs in extensional-shear In hydrothermal regimes, changes in stresses and/or
mode. Shear failure occurs at stress differences greater fluid pressures can induce brittle failure. For example, at
than approximately 6T. Because rock tensile strengths are low stress differences, pure extension failure is induced by
typically less (sometimes substantially less) than about 10 increase in fluid pressure (Fig. 16a), provided the effective
MPa, the occurrence of pure extension veins in hydrother- minimum principal stress becomes negative and equal in
mal systems indicates stress differences less than 40 MPa magnitude to the tensile strength of the rock. This fluid-dri-
during vein opening (Etheridge, 1983). ven extension fracturing is known as hydraulic extension
fracture. The hydraulic fracture criterion is, thus,
The role of fluids
Fluid pressure modifies stress states at depth in the Pf = σ 3 + T. (21)
Earth’s crust. The effect of pore fluid pressure (Pf) is to
reduce the effective normal stress (σ n′ ) according to the The abundance of mineral-filled extension fractures in
relationship, many epigenetic ore deposits indicates that tensile effective
stress states, and, thus, fluid pressures greater than σ 3, are
σ n′ =σ n – Pf. (20) common in hydrothermal systems. Importantly, equation

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 15

a a

SHEAR STRESS
SHEAR STRESS, τs
B
σ3
σn

σ3
C
τ=
σn σn
τ τ
α
σ1 σ1
θ
P
τ C
C Pf

2α σ 3′ σ1 ′ σ3 σ1
-T σ3 σn A σ1
EFFECTIVE NORMAL STRESS
NORMAL STRESS, σn

b b
SHEAR STRESS, τs

SHEAR STRESS
A

B Pf

C
2α 2α

-T
NORMAL STRESS, σn -T σ3′ σ3 σ1′ σ1
EFFECTIVE NORMAL STRESS
c K
SHEAR STRESS, τs

OC FIG. 16. Mohr circle constructions illustrating the effects of increasing


C TR
TA fluid pressure, by an amount Pf, on effective stress states and conditions
IN
B for failure in intact rock. a. Increasing fluid pressure at low stress differ-
ences results in extension failure. b. Increasing fluid pressure at high stress
differences induces shear failure.

(21) illustrates how the opening of hydraulic extension


fractures limits the maximum fluid pressures that may
A 2β develop in hydrothermal systems.
An important aspect of the role of fluid pressure in frac-
NORMAL STRESS, σn ture processes is that brittle failure may be induced at any
depth, provided fluid pressures are high enough. Perme-
FIG. 15. a. Mohr diagram and inset illustrating relationship between
principal stresses, shear stress, and normal stress. The normal stress (σn) ability enhancement by fracture growth can occur in
and shear stress (τs) on a plane inclined at angle α to the orientation of response to changes in stress difference (i.e., through
the maximum principal stress are given by the coordinates of the point P changes to σ1 and/or σ3), and also by changes in fluid
on a Mohr circle having diameter (σ1 – σ3) and center at (σ1 + σ3)/2 on pressure. This concept is illustrated in Figure 17, which
the normal stress axis. The curved line –TCB is a typical brittle failure
envelope (modified Griffith-Mohr-Coulomb failure criterion). T is the ten-
plots failure modes as a function of stress difference and
sile strength of the rock; the shear failure envelope (CB) is approximated fluid pressure. Here, fluid pressure is expressed as the
by the relationship τ = C + µσn, where τ is the shear strength, C is the cohe- pore fluid factor (λv), the ratio of fluid pressure to over-
sive strength, µ is the coefficient of friction, and σn is the normal stress. burden stress (σv). The pore fluid factors and stress differ-
Note that shear strength is dependent on normal stress. Stress states to the ences leading to failure in extensional, extensional-shear,
right of the failure envelope do not result in brittle failure. b. Mohr dia-
gram illustrating Mohr circles for (A) shear failure (faulting), (B) hybrid and shear modes are plotted for a strike-slip regime at a
extensional shear, and (C) pure extension failure of intact rock. c. Mohr depth of 3 km. From some ambient stress state, failure can
diagram illustrating stress conditions for reactivation of a nonoptimally be induced either by increasing the pore fluid factor along
oriented, low cohesion, shear failure plane (e.g., preexisting fault, or foli- the trajectories such as Q and S (Fig. 17) without changing
ation/bedding anisotropy) inclined at angle β to the maximum principal
stress. Note that when the shear failure envelope (AB) for a preexisting
the absolute stress state (an example of purely fluid-driven
mechanical weakness lies below the failure envelope for intact rock, the failure), or by maintaining a constant pore fluid factor and
preexisting structure may be preferentially reactivated. increasing the stress difference (e.g., trajectory P), or some

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
16 COX ET AL.

strike-slip faulting, depth = 3 km (Nguyen et al., 1998) that cannot be ascribed to repeated
fluctuations in temperature and depth. They are instead
related to fluctuations in shear stress and fluid pressure asso-
1.5 ciated with episodic fault slip and fault-valve behavior in
PORE FLUID FACTOR, λ v

EXTENSION
shear zones (Sibson et al., 1988; Sibson, 2001). Changes in
EXTENSIONAL SHEAR
A deformation style, caused by changes in fluid pressure states,
1.0 lithostatic are a key factor influencing deformation style, the nature of
permeability, and pervasiveness of flow and alteration.
Q S R For example, in the shear-hosted Revenge gold deposit in
P the St. Ives Goldfield (Eastern Goldfields Province, Yilgarn
0.5
hydrostatic craton, Western Australia), Au-mineralized shear zones
SHEAR B exhibit initial brittle shear failure, subsequent mixed brit-
tle–ductile behavior, and finally, fully brittle behavior during
progressive hydrothermal alteration (Nguyen et al., 1998).
0 20 40 60 80 100 Initial fluid flow was controlled by macroscopic shear frac-
STRESS DIFFERENCE, MPa ture at pore factors near one, that is, near-lithostatic fluid
pressures (Fig. 18a). Progressive potassic alteration of
FIG. 17. Failure mode diagram illustrating how changes in stress differ- metabasalt and metadolerite host-rocks produced weaker,
ence and fluid pressure (expressed as pore fluid factor; see text for expla- biotite-rich alteration assemblages. Reduced plastic shear
nation) lead to brittle failure. Fields for pure extension, extensional-shear,
and shear failure are indicated for a strike-slip regime at a depth of 3 km. strength, due to reaction-weakening, induced a transition to
A failure envelope (AB) is indicated for intact rock with cohesive strength ductile deformation (Fig. 18b). In this deformation regime,
of 10 MPa. Failure occurs when the pore fluid factor and/or stress differ- fluid flow was controlled by permeability enhancement asso-
ence increase to touch the failure envelope AB. Trajectory P leads to shear ciated with grain-scale dilatancy and reaction-enhanced
failure at constant fluid pressure with increasing shear stress. Trajectory R
leads to shear failure by increase in both fluid pressure and stress differ-
porosity. Repeated brittle slip events occurred during ongo-
ence. Trajectories S and Q lead to shear failure and extension failure, ing ductile deformation, and were associated with formation
respectively, by pore fluid pressure increase alone. of breccias and laminated fault-fill veins in dilatant bends
and jogs. Episodic fast seismic(?) slip events are interpreted
to have been driven by increasing pore fluid factors during
combination of both stress and fluid pressure increase individual fault-valve cycles. Each brittle shear failure event
(e.g., trajectory R). was preceded by an interval of interseismic ductile shear at
Fluid-driven failure is extremely important in generating lower pore fluid factors (Fig. 18b). Sodic metasomatism late
fault/fracture networks in hydrothermal systems in both in the hydrothermal history produced localized albite-rich
the upper and lower crust, provided fluid pressures are assemblages, which resulted in reaction-strengthening and a
high enough (Sibson, 1996; Cox, 1999). A significant late transition back to fully brittle behavior and macroscopic
aspect of fluid-driven fracturing is that, even without fracture-controlled fluid flow (Fig. 18c).
changes in the stresses acting on a previously nondeform-
ing rock medium, infiltration of high-pressure fluids can Implications of stress and fluid pressure regimes for
drive the spontaneous growth of fractures. The orientations localization of fracturing, fluid flow, and mineralization
of these fractures are controlled by the orientation of the Fluid pressure regimes in the crust are controlled to a
principal stresses, magnitudes of the stress differences, and first order by permeability and fluid driving pressures. In
geometry of any preexisting mechanical anisotropies in the the upper few kilometers of the continental crust, where
rock mass. fracture and pore connectivity is high, pore fluid pressures
are generally close to hydrostatic. At deeper levels, where
Effects of stress and fluid pressure on transitions pore sealing and collapse are more rapid, loss of long-term
between failure modes pore connectivity results in fluid pressures increasing
Particularly in hydrothermal systems formed at mid- towards lithostatic levels. Deeper level fluid reservoirs are
crustal depths (e.g., some mesothermal gold systems), fluid accordingly suprahydrostatically pressured. Tapping of
flow, hydrothermal alteration, and mineralization are con- these reservoirs by active and permeable faults, fractures,
trolled by faults and shear zones that exhibit mixed brit- and shear zones provides transient pathways for rapid
tle–plastic behavior. First order transitions between brittle upwards migration of fluids under high driving pressure
and plastic behavior in the crust are commonly ascribed to gradients. Fluid discharge from the upper levels of these
the effects of increasing temperature and confining pres- structures can generate overpressured domains at these
sure (with increasing depth) inhibiting brittle deformation sites (Fig. 1), especially if the fluid discharge zone is capped
and promoting the operation of thermally activated defor- by a low-permeability domain. Fluid-driven growth of
mation processes such as dislocation creep and diffu- fault/fracture systems is more favored as pore fluid factors
sional/advective mass transfer processes. However, shear progressively increase above hydrostatic levels (Fig. 19).
zones in some mesothermal gold systems exhibit evidence An example of localized fluid pressurization and associ-
for repeated transitions between brittle and plastic behavior ated mineralization is provided by the development of the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 17

swarm of subhorizontal extension veins which hosts tung-


a P Q R sten mineralization at Panasqueira in Portugal (Foxford et
PORE FLUID FACTOR
1
• •• lithostatic Pf
al., 2000). The vein swarm in this deposit developed above a
brittle shear failure granitoid stock in a subhorizontal, lenticular domain which
B is less than 200 m thick. The veins are interpreted to have
grown where fluid migrating up from the stock became
0.9 •S trapped below a low permeability zone and developed a
supralithostatic-pressured domain of hydraulic fracturing
(i.e., Pf = σ 3 + T). Because extension veining at Panasqueira
A ductile shear failure is not associated with faulting, stress differences are inferred
0.8
0 50 100
•T 150
to have been low. The thickness of the veined domain is con-
trolled by the limited depth interval over which supralitho-
STRESS DIFFERENCE, MPa static fluid pressures developed above the stock.
The Bendigo goldfield in central Victoria, Australia, pro-
vides an example of where the structure of low permeabil-
b P Q ity stratigraphic units may have localized the distribution of
• • •R lithostatic Pf
PORE FLUID FACTOR

1 high pore fluid factor sites and associated vein gold


deposits. The Bendigo goldfield occurs in the regional cul-
•S mination zone of a series of doubly-plunging folds. Addi-
ductile shear failure tionally, most fault-related and vein-related gold deposits in
0.9 the area are located in and near saddle reefs in the hinges
B of anticlines. High pore fluid factors, associated hydraulic
fracturing, and veining at Bendigo are interpreted to have
A been controlled by fluid migration to the structurally high-
0.8
0 50
•T 100 150
est levels beneath folded, low-permeability stratigraphic
seals (Cox et al., 1991).
STRESS DIFFERENCE, MPa At the Porgera gold deposit in Papua New Guinea, early
hydrothermal alteration, fluid flow, and low-grade, dis-
persed mineralization was associated with pervasive grain-
c P Q scale and fracture-controlled flow around a mafic intrusive
• • •R lithostatic Pf
PORE FLUID FACTOR

1 complex (Munroe, 1995). Early in the development of the


hydrothermal system, stress differences were not high
enough to generate faults, but high pore fluid factors pro-
•S duced extension veins. Subsequent growth of a fault across
0.9 the active, intrusive-related hydrothermal system was asso-
ciated with a rotation of the stress field and increase in
stress difference (Cox and Munroe, 1999). The increase in
stress difference, together with high pore fluid factors,
0.8
0 50 100
•T 150
drove the growth of a fault and its associated high-perme-
ability damage zone, which tapped deeper level hydrother-
STRESS DIFFERENCE, MPa mal fluids. Localization of fluid flow within the fault and its
FIG. 18. Failure mode diagrams illustrating brittle and plastic failure
damage zone led to the formation of a very high grade gold
envelopes as a function of pore fluid factor and stress difference. The dia- deposit around the fault. This contrasts with the early min-
gram is constructed for optimally oriented reverse faulting at a depth of 12 eralization which is more dispersed and not spatially associ-
km in rock with a cohesive strength of 10 MPa, tensile strength of 5 MPa, ated with faulting.
and friction coefficient 0.75. Fluid pressure and stress changes associated Competence contrasts within rock masses also control
with fault-valve behavior may produce episodic transitions between no
deformation, ductile shear failure, brittle shear failure, and brittle exten- localization of fracturing (see also Oliver et al., 2001). For
sion failure at various phases of the seismic cycle. The failure envelope is example, where strong rock masses occur within a weaker
indicated by the curve PQRST. Ductile shear failure occurs for fluid pres- matrix, with boundaries oriented at a high angle to the short-
sure and stress states in the interval ST; brittle shear failure occurs between ening direction, viscous or frictional drag along the contacts
S and R; extensional-shear occurs between R and Q; and extensional fail-
ure occurs in the interval PQ. a. Brittle shear failure at point B induced by
reduces the minimum principal stress in the competent rock
increase in stress difference and fluid pressure. b. Reaction-weakening unit. As fluid pressures build-up, hydraulic extension frac-
decreases the ductile shear strength (ST), so that increase in stress differ- ture or shear failure (depending on stress difference) occurs
ence and fluid pressure leads firstly to ductile shear failure (aseismic first in the zone of decreased minimum effective principal
creep) at point B, followed by brittle shear failure at point S, in response to stress in the competent rock unit. Examples include the
progressive increase in pore fluid factor. c. Late-stage reaction-hardening
increases ductile shear strength, so that failure occurs by brittle shear fail- development of mesothermal Au-mineralized extension vein
ure at high stress differences, or extension failure occurs at low stress dif- arrays in dolerite in the Mt Charlotte orebodies at Kalgoorlie
ferences and high pore fluid factors. (Clout et al., 1990); ladder vein systems in dikes, such as at

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
18 COX ET AL.

a b
surface FLUID PRESSURE
X

near-hydrostatic
fluid pressure high permeability
regime

low low permeability


permeability seal zone
domain

zone of
fluid discharge zone Pf ≥ σ 3 + T hydrofracture veining

DEPTH
and fluid-driven shear
failure
E
ON

focussed fluid
TZ

migration
UL
FA

Pf = σ 3 + T
overpressured hydrostatic Y
fluid reservoir (λ v ≈ 0.38)

FIG. 19. Localization of fluid-driven failure controlled by depth-dependence of stresses and fluid pressure. a. Upwards,
head-driven fluid flow in a fault, which taps an overpressured fluid reservoir. Fluid discharge at the downstream end of
the fault drives fluid pressure build-up beneath a low permeability leaky seal zone. Where stress differences are low,
hydraulic extension fracture arrays can form in the domain where Pf ≥ σ 3 + T. For higher stress differences, brittle shear
failure can occur where Pf < σ 3 + T. b. Schematic profile of fluid pressure as a function of depth (solid line XY) for the
fault-controlled flow system illustrated in (a). Depth dependence of fluid pressures equivalent to hydrostatic and (σ 3 +
T) levels shown for comparison. For low stress differences (σ 1 – σ 3 < 4T), hydraulic extension fracturing occurs where Pf
= σ 3 + T. At higher stress differences, shear failure may occur at lower fluid pressures.

Wood’s Reef, Victoria (Clappison, 1954; Edwards, 1954), and For τ f greater than or equal to 0, shear failure will occur;
at Lamaque, Quebec (Robert, 1990); and gold-bearing vein for τ f less than 0, shear failure will not occur. This parame-
systems hosted within dike-like, competent felsic intrusions ter is the same as the Coulomb stress parameter used to
in the St Ives goldfield in Western Australia. predict locations of aftershocks triggered by stress transfer
In the seismogenic regime, fluid pressure and shear after major earthquakes (King et al., 1994; Stein, 1999).
stress cycling associated with fault-valve behavior potentially Similarly, the growth of hydraulic extension fractures will
lead to complex, but cyclically repeated changes to pore occur where and when
fluid factors and stresses (Sibson, 2001). Time and space
variations in pore factors, shear stress, and normal stress are Pf ≥ σ 3 + T. (24)
the main parameters influencing the timing and location of
episodic fracture, and distribution of high-permeability That is, where and when fluid pressure is highest, σ 3 is
zones within hydrothermal systems. least, or tensile strength is least (Cox et al., 1987).
According to Coulomb failure criteria, brittle shear fail- Understanding how stress and fluid pressure states vary in
ure occurs where and when shear stress τ s is greater than a rock mass has important implications for predicting local-
the rock shear strength; that is, where ization of fluid flow and ore deposition in hydrothermal sys-
tems. The use of Coulomb failure criteria to predict where
τ s ≥ C + µ (σ n – Pf). (22) deformation-induced flow paths can form requires a knowl-
edge of the variation of shear stress, normal stress, fluid pres-
Proximity to shear failure can, thus, be described by a sure, friction coefficients, and cohesive strengths in rock
time-dependent parameter, τ f, which is the difference masses in three dimensions, over time. Numerical modeling
between the shear stress and the shear strength. Thus, approaches such as that by Holyland and Ojala (1997) used
two-dimensional and three-dimensional linear elastic defor-
τ f = τ s – C – µ (σ n – Pf). (23) mation modeling to predict where minimum mean stresses

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 19

or effective minimum principal stresses occur. However, cient connectivity to create networks that link fluid source
equation (23) illustrates that these approaches are highly rocks and ore deposition sites. Percolation theory (Sahimi,
simplified and might not adequately treat all the important 1994) provides useful insights about the evolution of
parameters controlling “proximity to failure.” In particular, hydraulic connectivity and the partitioning of fluid flow
potential spatial variations in fluid pressures need to be mod- amongst elements of a network of relatively permeable
eled. Modeling strategies, used to assist prediction of where structures in an otherwise low-permeability medium
fluid flow and mineralization may be localized, need to treat (Berkowitz, 1995; Cox and Knackstedt, 1999).
coupling between stress and fluid pressure, as well as poten- In the same way as for grain-scale pore networks, macro-
tial time and space variations in stress states associated with scopic percolation networks, comprising faults, fractures,
stress relief, stress transfer, and fluid pressure fluctuations and shear zones, can be described in terms of three types of
during the seismic cycle. This is a fertile field for further elements: backbone, dangling, and isolated elements (Fig.
development. 20a). Backbone elements provide a direct connection from
An example of potential feedbacks between deformation, one side of the system to the other (e.g., metal source reser-
effective stress states, and permeability in controlling ore voir to ore deposit) and carry the bulk of the fluid flux. Dan-
localization is provided by the distribution of some gling elements, also known as dead-end elements, branch
fault/shear-hosted mesothermal gold deposits around shear from the flow backbone and act as fluid feeders to the back-
systems. Archean mesothermal deposits are typically located bone in the upstream part of the system, or as distributary or
within low displacement faults, shear zones, and associated discharge structures in the downstream part of the system.
fracture systems adjacent to larger, crustal-scale fault systems Isolated elements are disconnected from both the backbone
(Robert et al., 1995; Cox, 1999). Timing, spatial, and kine- and dangling elements in the network, and are, therefore,
matic relationships between low-displacement and high-dis- low-flux structures not connected to fluid reservoirs.
placement structures indicate that many of the low-displace- At very low bulk crustal strains, most of the faults or shears
ment faults, shear zones, and fracture networks that host Au in a deforming domain will be short, isolated structures. With
deposits probably formed as aftershock structures in response increasing deformation, active faults and shears increase in
to slip events on major faults. Co-seismic stress transfer is a length and surface area; new structures also nucleate and
major factor controlling the location of aftershock activity grow so that fault connectivity increases with strain. A critical
(Stein, 1999). Accordingly, fluid invasion of slipped fault seg- point, known as the percolation threshold (Sahimi, 1994), is
ments, the consequent decrease in effective stresses, and reached when enough elements connect to allow fluid flow
fluid-driven failure after major slip events, may act together across the entire width of the network. Reaching the percola-
with mainshock-related stress transfer in localizing growth of tion threshold corresponds to the sudden onset of flow. The
low-displacement structures, high fluid flux, and mineraliza- point at which the percolation threshold is reached is depen-
tion that produces mesothermal gold systems. dent on several factors including strain accommodated by
growth of permeable faults, fractures, and shear zones, as well
Evolution of Flow Pathways during Deformation as by fracture geometries and relative rates of fracture growth
and nucleation (Zhang and Sanderson, 1994; An and Sam-
Localization of deformation and flow mis, 1996; Roberts et al., 1998).
The strain distribution associated with crustal deformation Partitioning of flow among elements of a fracture/shear
is typically very heterogeneous, with higher strains and asso- network is dependent on the relative proportions of back-
ciated permeability enhancement being localized along bone, dangling, and isolated elements (Cox and Knackst-
structures such as faults and shear zones. The significance of edt, 1999). Just above the percolation threshold, the flow
active fracture networks in controlling fluid migration is illus- backbone is a very small fraction of the total network, but
trated by analytical solutions and numerical modeling stud- most of the flow is localized along this part of the system
ies of two-dimensional flow patterns around high-permeabil- (Fig. 20b). At higher strains the proportion of elements
ity zones in a less permeable matrix (Phillips, 1991; Matthai that are part of the backbone progressively increases, so
and Roberts, 1997; Taylor et al., 1999). For permeable struc- that flow becomes progressively more evenly distributed
tures inclined at low angles to the regional gradient of across the system as more faults or shears become con-
hydraulic head, fluid focusing occurs around the higher nected to each other and to fluid reservoirs.
pressure (upstream) levels of shear zones, whereas fluid dis- At low displacements, all fracture segments are isolated
charge occurs in the lower pressure (downstream) levels of from one another and the fluid reservoir. The percolation
shear zones (Fig. 1). The dimensions of fluid charge regions threshold is reached when approximately 30 percent of the
are comparable to the dimensions of the high-permeability sites are occupied. For the three-dimensional fault configu-
segments of faults and shear zones. Importantly, shears ration illustrated in Figure 20b, and using typical
inclined at high angles to the regional hydraulic head gradi- length/displacement scaling relationships for faults
ent are much less effective at focusing fluid flow. (Scholz, 1990), the percolation threshold is reached at bulk
strains of only a few percent. For three-dimensional perco-
Development of percolation networks in hydrothermal systems lation, flow pathways can be very tortuous and produce
Fracture/shear-controlled hydrothermal systems develop point-like distribution of high fluid flux sites along
where and when linked permeable structures develop suffi- restricted segments of faults and shear zones (Fig. 21). This

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
20 COX ET AL.

tures, which link fluid sources and sinks. This situation max-
a HYDRAULIC HEAD GRADIENT
imizes fluid/rock ratios in the fluid-accessible parts of perco-
lation networks, and therefore maximizes the potential for
generation of “giant” ore deposits. In contrast, for hydrother-
mal systems well above the percolation threshold, fluid flow
is distributed over a larger proportion of the fracture popu-
lation. This more dispersed flow is likely to produce more
distributed, potentially lower grade mineral deposits.
An example of this control could be provided by the typ-
ical distribution of major mesothermal gold deposits, or
clusters of deposits, along a small proportion of the total
population of kinematically related (and thereby broadly
backbone dangling isolated co-active) faults and shear zones (Cox, 1999). This distribu-
elements elements elements tion indicates that most fluid flow is localized along a small
number of active structures, and, therefore, that the
fault/shear networks that successfully generate major
mesothermal lode gold systems have operated near the per-
colation threshold during ore genesis. Some porphyry-type
b 1
systems, where flow occurs through dense and highly inter-
connected fracture networks (e.g., Haynes and Titley,
FRACTION OF OCCUPIED

0.8 1980), may be deposit-scale examples of flow systems well


backbone
above the percolation threshold.
SITE TYPES

0.6 isolated At any instant, hydraulic connectivity in a network is


dependent upon which parts of the system are actively frac-
0.4 turing (and hence permeable). For example, as deforma-
tion migrates through a shear system, not all parts of a geo-
dangling metrically connected and kinematically linked network of
0.2
faults and shear zones are active. So, the locus of fluid flow
will be transient and migrates with active deformation
0
0 0.2 0.4 0.6 0.8 1 through the system. In thrust systems, where new thrusts
FRACTION OF TOTAL SITES commonly nucleate progressively deeper into the footwall
OCCUPIED of previously active thrust sheets, the locus of fluid flow will
migrate towards the foreland with time.
FIG. 20. a. Schematic two dimensional representation of a fault/fracture Particularly in the middle to upper crustal seismogenic
network consisting of isolated elements, dangling elements, and the back-
bone structure. Most flow is localized along the flow backbone. Dangling
regime, rapid changes in permeability in fault networks on
elements in the upstream (lower, in this case) part of the system feed fluid timescales of seismic slip recurrence can also lead to sud-
to the backbone of the system, whereas dangling elements of the network den changes in the location and architecture of flow back-
in the downstream (upper) part of the system act as fluid discharge sites. b. bones, unless most displacement is also localized on the
Fraction of isolated, dangling, and backbone sites as a function of total fluid flow backbone.
number of sites for the three-dimensional case of conjugate fractures or
faults inclined at 45° to the bulk flow direction (after Cox and Knackstedt, Self-organization of flow networks
1999).
It was noted earlier that the distribution of mesothermal
gold deposits along a small proportion of the total popula-
factor will act together with variations in fracture densities tion of kinematically related and co-active faults or shear
and apertures to control flow distribution. zones may indicate that these systems develop most effec-
Several points emerge directly from percolation model- tively in percolation networks that operated close to their
ing. Firstly, the result that the percolation threshold is percolation threshold. This raises a question: do some
reached at bulk strains of only a few percent indicates that processes maintain fault/fracture/shear networks near the
fracture-controlled hydrothermal systems and associated percolation threshold? Such behavior is a manifestation of
ore deposits can develop at very low strains. High-displace- “self-organized criticality,” a term used to describe processes
ment structures and large regional strains are not necessary which hold complex or disordered systems near their critical
to facilitate high fluid fluxes. In this regard, small strains point (e.g., percolation threshold). Many natural processes
during the initial stages of changes in plate kinematics may such as slope failure in sandpiles and earthquake rupturing
be important in developing new fracture systems that tap events exhibit self-organized behavior (Bak, 1997).
into previously inaccessible fluid reservoirs. In the seismogenic regime, interseismic fault sealing can
Secondly, for networks just above the percolation thresh- repeatedly drive networks back below the percolation
old, where the backbone is a very small proportion of the threshold after rupture events, so that on the lifetime of the
total fault population, flow is localized on relatively few struc- hydrothermal system the network may oscillate about the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 21

FLUID OUT

FLUID IN
FIG. 21. Three-dimensional flow paths associated with flow through part of a network of intersecting fractures or faults
just above the percolation threshold. Fluid enters the network at the bottom and follows a tortuous path (shown in
black) to higher structural levels. Three-dimensional flow near the percolation threshold produces a point-like distrib-
ution of high flux in map view.

percolation threshold. However, other factors may also lead Such extreme fluid focusing generates high potential for
to self-organization in hydrothermal systems. formation of giant ore deposits.
The network modeling described above assumed that,
although elements of a percolation network nucleate ran- Ore deposition in percolation networks
domly, all elements grow at the same rate. In cases where In many cases, ore deposition is controlled by processes
fracture systems connect to an overpressured fluid reser- such as (1) fluid mixing, (2) fluid-rock reaction, (3) phase
voir, invasion of high-pressure fluids along fluid-accessing separation driven by sudden fluid pressure drop, and (4)
elements of the network may preferentially weaken these gradients in pressure and temperature (Skinner, 1997).
elements and dramatically enhance their growth rates rela- Ore deposition is dependent on structurally controlled
tive to elements isolated from fluid reservoirs. This occurs delivery of fluids to appropriate reaction sites. These
through the effects of high fluid pressures promoting fail- processes operate with different effectiveness in different
ure and reaction-softening processes. Accordingly, con- parts of percolation networks.
nected elements can slip and grow at much faster rates than Isolated elements of percolation networks have the low-
isolated elements. The fluid-driven growth, or “self-genera- est fluid flux and, therefore, have minimal ore potential in
tion” (Sibson, 1996), of percolation networks in response hydrothermal systems. The ore potential of backbone and
to invasion of overpressurized fluid, therefore, provides a dangling elements depends on the nature of the ore depo-
positive feedback between fluid access and fracture growth sition reactions, and whether the site is in the upstream or
rate. This may localize repeated slip and fluid flow on the downstream part of the system.
flow backbone that forms at the percolation threshold. Most flow occurs along the network backbone. Even
Fluid-driven network growth may well lead some hydrother- though they are high-flux sites, flow backbones are unlikely
mal systems to self-organize near the percolation threshold. to be sites of substantial fluid-rock interaction or fluid mix-

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
22 COX ET AL.

ing away from their upstream or downstream terminations. fault termination zones (wing cracks and splays) and fault
However, backbone sites do provide potential for ore depo- intersections, form structural sites with high potential for
sition in response to temperature and pressure gradients. localizing fluid flow.
Backbone elements also have the potential for sudden fluid The macroscopic fluid pathways in fracture-controlled
pressure drops (and associated phase separation) caused by hydrothermal systems are influenced by the evolving con-
co-seismic dilatancy associated with large slip on these struc- nectivity among elements of fracture/fault/shear networks
tures in the seismogenic regime. The potential for ore during progressive deformation. The most extreme flow
deposition involving fluid-rock reaction is dependent on localization and greatest potential for ore localization
the grain-scale pervasiveness of flow. Especially during flow occurs at the onset of system-wide fluid flow, when the sys-
along macroscopic fractures, fluid migrating along the flow tem reaches the percolation threshold. Percolation thresh-
backbone has limited interaction with the fracture walls. olds may be reached at bulk strains as low as a few percent.
However, where flow is controlled by grain-scale dilatancy, Systems well above the percolation threshold are associated
such as in cataclasite or ductile shear zone materials, fluid- with more dispersed fluid flow, and may provide lower
rock reaction can occur along the flow path. potential for ore formation.
High potential for fluid-rock reaction is provided in the
downstream parts of percolation networks, where fluids dis- Acknowledgments
charge from active structures into the surrounding, perme- This work was funded partly by a grant from the Aus-
able rock medium. Fluid discharge from the downstream tralian Research Council, as well as by support from the
termination of the flow backbone (Fig. 1), or from down- Research School of Earth Sciences and WMC Resources
stream dangling members of the network, promotes fluid- Ltd. R. Henley, S. Munroe, P. Nguyen, K. Ruming, R. Sib-
rock interaction. Similarly, this part of the flow network son, W. Stone, J. Streit, and S. Zhang all provided valuable
generates potential for fluid mixing if deeply derived fluids discussion about ore systems and fluid flow in rock media.
discharging from the network interact with shallow-level K.A. Foxford is thanked for providing a preprint. The con-
fluid reservoirs already present in the discharge regime, as structive review comments of M. Jébrak and S. Temperley,
may be the case for many types of porphyry-type and as well as the editorial contributions of J. Richards, are also
epithermal systems. greatly appreciated.
Potential for ore deposition by fluid mixing reactions
also occurs in the upstream parts of networks. The REFERENCES
upstream regions of flow backbones are dominated by An, L.-J., and Sammis, C.G., 1996, A cellular automaton for the develop-
focusing of fluids from dispersed reservoirs and host rocks, ment of crustal shear zones: Tectonophysics, v. 253, p. 247–270.
into the permeable backbone network. Mixing of fluids Angevine, C.L., Turcotte, D.L., and Furnish, M.D., 1982, Pressure solution
derived from different reservoirs will be particularly effec- lithification as a mechanism for stick-slip behavior of faults: Tectonics,
v. 1, p. 151–160.
tive where dangling (fluid feeder) elements of networks Bak, P., 1997, How nature works—the science of self-organized criticality:
intersect the upstream segments of backbone structures. Oxford, Oxford University Press, 212 p.
Bear, J., 1972, Dynamics of fluids in porous media: New York, Elsevier, 764 p.
Conclusions Berkowitz, B., 1995, Analysis of fracture network connectivity using perco-
Active deformation is necessary to generate and maintain lation theory: Mathematical Geology, v. 27, p. 467–483.
Bourbié, T., and Zinszner, B., 1985, Hydraulic and acoustic properties as a
permeability, and to sustain large-scale fluid flow in many function of porosity in Fontainebleau sandstone: Journal of Geophysical
hydrothermal systems. At upper crustal levels, and in rocks Research, v. B13, p. 11,524–11,532.
with low intergranular porosity, flow is controlled by per- Brace, W.F., 1990, Permeability of crystalline and argillaceous rocks—sta-
meability associated with grain-scale cataclasis and macro- tus and problem: International Journal of Rock Mechanics, Mining Sci-
ence and Geomechanics Abstracts, v. 17, p. 876–893.
scopic fracture growth. At deeper crustal levels, fluid flow in Brantley, S.L., Evans, B., Hickman, S.H., and Crerar, D.A., 1990, Healing
active shear zones typically is controlled by deformation- of microcracks in quartz—implications for fluid flow: Geology, v. 18,
induced, grain-scale microcrack permeability. However, p. 136–139.
macroscopic fracture permeability still develops in high Brown, S.R., 1987, Fluid flow through rock joints—the effect of surface
pore fluid factor regimes at elevated temperatures. roughness: Journal of Geophysical Research, v. 92, p. 1337–1347.
Cathles, L.M., Erendi, A.H.J., and Barrie, T., 1997, How long can a
The distribution and geometry of fracture-controlled hydrothermal system be sustained by a single intrusive event?: Economic
fluid pathways is governed by fluid pressure and stress state. Geology, v. 92, p. 766–771.
High pore fluid factors are important in driving growth of Chace, F.M., 1949, Origin of the Bendigo saddle reefs, with comments on
fracture permeability at all scales. High pore fluid factor the formation of ribbon quartz: Economic Geology, v. 44, p. 561–597.
Clappison, R.J.S., 1954, The Morning Star Mine, Wood’s Point, in
regimes tend to develop above leaking, suprahydrostatic- Edwards, A.B., ed., Geology of Australian ore deposits: Melbourne, Aus-
pressured fluid reservoirs, and below low-permeability tralasian Institute of Mining and Metallurgy, p 1077–1082.
domains. Faults and shear zones play a key role in tapping Clout, J.M.F., Cleghorn, J.H., and Eaton, P.C., 1990, Geology of the Kalgo-
fluids from deeper level fluid reservoirs. orlie goldfield, in Hughes, F.E., ed., Geology of the mineral deposits of
The highest fluid flux in individual faults or shear zones Australia and Papua New Guinea: Australasian Institute of Mining and
Metallurgy Monograph Series, v. 14, p. 411–431.
is localized in sites with the highest fracture apertures Cox, S.F., 1995, Faulting processes at high fluid pressures: An example of
and/or fracture density. Thus, both dilational and contrac- fault-valve behavior from the Wattle Gully Fault, Victoria, Australia: Jour-
tional bends and jogs on faults and shear zones, as well as nal of Geophysical Research, v. 100, p. 841–859.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
PRINCIPLES OF STRUCTURAL CONTROL IN HYDROTHERMAL SYSTEMS 23

——1999, Deformational controls on the dynamics of fluid flow in Hulin, C.D., 1929, Structural control of ore deposition: Economic Geol-
mesothermal gold systems, in McCaffrey, K., Lonergan, L., and Wilkin- ogy, v. 24, p. 15–49.
son, J., eds., Fractures, fluid flow, and mineralization: Geological Society Kanagawa, K., Cox, S.F., and Zhang, S., 2000, Effects of dissolution-precip-
of London Special Publication, v. 155, p. 123–140. itation processes on the strength and mechanical behavior of quartz
Cox, S.F., and Etheridge, M.A., 1989, Coupled grain-scale dilatancy and gouge at high-temperature hydrothermal conditions: Journal of Geo-
mass transfer during deformation at high fluid pressures, Mt Lyell area, physical Research, v. 105, p. 11,115–11,126.
Tasmania: Journal of Structural Geology, v. 11, p. 147–162. King, G.C.P., Stein, R.S., and Lin, J., 1994, Static stress changes and the
Cox, S.F., and Knackstedt, M.A., 1999, Ore genesis in fracture-controlled triggering of earthquakes: Bulletin of the Seismological Society of Amer-
hydrothermal systems: Percolation theory approaches: PACRIM ’99, ica, v. 84, p. 935–953.
Bali, Indonesia, October 10–13, 1999, Proceedings, p. 639–642. Knackstedt, M., and Cox, S.F., 1995, Percolation and pore geometry of
Cox, S.F., and Munroe, S.M., 1999, Fluid-driven faulting processes in an crustal rocks: Physical Review E, v. 51, p. R5181–R5184.
intrusive-related hydrothermal system, Porgera, Papua New Guinea Knipe, R.J., 1998, Faulting, fault sealing and fluid flow in hydrocarbon
[abs.]: Eos, v. 80 (46), p. F728–F729. reservoirs—an introduction, in Jones, G., Fisher, Q.J., and Knipe, R.J.,
Cox, S.F., and Paterson, M.S., 1991, Experimental dissolution-precipitation eds., Faulting, fault sealing, and fluid flow in hydrocarbon reservoirs:
creep in quartz aggregates at high temperatures: Geophysical Research Geological Society of London Special Publications, v. 147, p. vii–xxi.
Letters, v. 18, p. 1401–1404. Manning, C.E., and Ingebritsen, S.E., 1999, Permeability of the continen-
Cox, S.F., Etheridge, M.A., and Wall, V.J., 1987, The role of fluids in syn- tal crust—the implications of geothermal data and metamorphic sys-
tectonic mass transport, and the localization of metamorphic vein-type tems: Reviews in Geophysics, v. 37, p. 127–150.
ore deposits: Ore Geology Reviews, v. 2, p. 65–86. Matthai, S.K., and Roberts, S.G., 1997, Transient versus continuous fluid
Cox, S.F., Wall, V.J., Etheridge, M.A., and Potter, T.F., 1991, Deformational flow in seismically active faults: An investigation by electric analogue and
and metamorphic processes in the formation of mesothermal vein- numerical modelling, in Jamtveit, B., and Yardley, B.W.D., eds., Fluid
hosted gold deposits—examples from the Lachlan Fold Belt in central flow and transport in rocks—mechanisms and effects: London, Chap-
Victoria, Australia: Ore Geology Reviews, v. 6, p. 391–423. man and Hall, p. 263–292.
David, C., and Darot, M., 1989, Permeability and conductivity of sand- McCaig, A.M., and Knipe, R.J., 1990, Mass transport mechanisms in
stone, in Maury, V., and Fourmaintreaux, D., eds., Rocks at great depth: deforming rocks—recognition using microstructural and microchemical
Amsterdam, Balkema, p. 203–210. criteria: Geology, v. 18, p. 824–827.
Edwards, A.B., 1954, Mines of the Walhalla-Wood’s Point auriferous belt, McKinstry, H.E., 1948, Mining geology: Englewood Cliffs, N.J., Prentice-
in Edwards, A.B., ed., Geology of Australian ore deposits: Melbourne, Hall Inc., 680 p.
Australasian Institute of Mining and Metallurgy, p. 1061–1076. Muir-Wood, R., and King, G.C.P., 1993, Hydrological signatures of earth-
Etheridge, M.A., 1983, Differential stress magnitudes during regional quake strain: Journal of Geophysical Research, v. 98, p. 22,035–22,068.
deformation and metamorphism: Upper bound imposed by tensile frac- Munroe, S.M., 1995, The Porgera gold deposit, Papua New Guinea: The
turing: Geology, v. 11, p. 231–234. influence of structure and tectonic setting on hydrothermal fluid flow
Fischer, G.J., and Paterson, M.S., 1992, Measurements of permeability and and mineralisation at a convergent margin: PACRIM ’95, Auckland,
storage capacity in rocks during deformation at high temperature and N.Z., November 19–22, 1995 Proceedings, p. 413–416.
pressure, in Evans, B., and Wong, T.-f., eds., Fault mechanics and trans- Newhouse, W.H., 1942, Ore deposits as related to structural features:
port properties of rocks: San Diego, Academic Press, p. 213–252. Princeton, N.J., Princeton University Press, 280 p.
Forster, C., and Smith, L., 1990, Fluid flow in tectonic regimes: Mineralog- Nguyen, P.T., Cox, S.F., Powell, C.M., and Harris, L., 1998, Fault-valve
ical Association of Canada Short Course Handbook, v. 18, p. 1–47. behaviour in optimally-oriented shear zones at Revenge gold mine,
Foxford, K.A., Nicholson, R., Polya, D.A., and Hebblethwaite, R.P.B., 2000, Kambalda, Western Australia: Journal of Structural Geology, v. 20,
Extensional failure and hydraulic valving at Minas da Panasqueira, Por- p. 1625–1640.
tugal—evidence from vein spatial distributions, displacements, and Norton, D., and Knight, J., 1977, Transport phenomena in hydrothermal sys-
geometries: Journal of Structural Geology, v. 22, p. 1065–1086. tems: Cooling plutons: American Journal of Science, v. 277, p. 937–981.
Fyfe, W.S., 1987, Tectonics, fluids, and ore deposits—mobilization and Ord, A., and Oliver, N.H.S., 1997, Mechanical controls on fluid flow dur-
remobilization: Ore Geology Reviews, v. 2, p. 21–36. ing regional metamorphism: some numerical models: Journal of Meta-
Garven, G., 1985, The role of regional fluid flow in the genesis of the Pine morphic Geology, v. 15, p. 345–359.
Point deposit, Western Canada sedimentary basin: Economic Geology, Oliver, N.H.S., Ord, A., Valenta, R.K., and Upton, P., 2001, Deformation,
v. 80, p. 307–324. fluid flow, and ore genesis in heterogeneous rocks, with examples and
Garven, G., and Freeze, R.A., 1984, Theoretical analysis of the role of numerical models from the Mount Isa district, Australia: Reviews in Eco-
groundwater flow in the genesis of stratabound ore deposits. 1. Mathe- nomic Geology, v. 14, p. 51–74.
matical and numerical model: American Journal of Science, v. 284, Peach, C.J., and Spiers, C.J., 1996, Influence of crystal plastic deformation
p. 1085–1124. on dilatancy and permeability development in synthetic salt rock:
Guéguen, Y., and Dienes, J., 1989, Transport properties of rocks from sta- Tectonophysics, v. 256, p. 101–128.
tistics and percolation: Mathematical Geology, v. 21, p. 1–13. Phillips, O.M., 1991, Flow and reactions in permeable rocks: Cambridge,
Guéguen, Y., and Palciauskas, V., 1994, Introduction to the physics of Cambridge University Press, 285 p.
rocks: Princeton, Princeton University Press, 294 p. Ramsay, J.G., 1980, The crack-seal mechanism of rock deformation:
Haar, L., Gallagher, J.S., and Kell, G.S., 1984, Steam tables: Washington Nature, v. 284, p. 135–139.
D.C., Hampshire Publishing Corporation, 319 p. Robert, F., 1990, Structural setting and control of gold-quartz veins of the
Haynes, F.M., and Titley, S.R., 1980, The evolution of fracture-related per- Val d’Or area, southeastern Abitibi Subprovince, in Ho, S.E., Robert, F.,
meability within the Ruby Star granodiorite, Sierrita porphyry copper and Groves, D.I., eds., Gold and base metal mineralisation in the Abitibi
deposit, Pima County, Arizona: Economic Geology, v. 75, p. 673–683. Subprovince, Canada, with emphasis on the Quebec segment: Nedlands,
Henley, R., Truesdell, A.H., and Barton, P.B., Jr., 1985, Fluid-mineral equi- University of Western Australia Publication 24, p. 167–209.
libria in hydrothermal systems: Reviews in Economic Geology, v. 1, Robert, F., and Poulsen, K.H., 2001, Vein formation and deformation in
p. 1–267. greenstone gold deposits: Reviews in Economic Geology, v. 14,
Hickman, S.H., and Evans, B., 1987, Influence of geometry upon crack p. 111–155.
healing in calcite: Physics and Chemistry of Minerals, v. 15, p. 91–102. Robert, F., Boullier, A.-M., and Firdaous, K., 1995, Gold-quartz veins in
Holness, M.B., 1997, The permeability of non-deforming rock, in Holness, metamorphic terranes and their bearing on the role of fluids in faulting:
M.B., ed., Deformation-enhanced fluid transport in the Earth’s crust and Journal of Geophysical Research, v. 100, p. 12,861–12,881.
mantle: London, Chapman and Hall, p. 9–39. Roberts, S., Sanderson, D.J., and Gumiel, P., 1998, Fractal analysis of Sn-W
Holyland, P.W., and Ojala, V.J., 1997, Computer-aided structural targeting mineralization from central Iberia—insights into the role of fracture
in mineral exploration: Australian Journal of Earth Sciences, v. 44, connectivity in the formation of an ore deposit: Economic Geology, v. 93,
p. 421–432. p. 360–365.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020
24 COX ET AL.

Rumble, D., 1994, Water circulation in metamorphism: Journal of Geo- Tsang, Y.W., and Witherspoon, P.A., 1981, Hydromechanical behavior of a
physical Research, v. 99, p. 15,499–15,502. deformable rock fracture subject to normal stress: Journal of Geophysi-
Rumble, D., Ferry, J.M., Hoering, T.C., and Boucot, A.J., 1982, Fluid flow cal Research, v. 86, p. 9287–9298.
during metamorphism at the Beaver Brook fossil locality, New Hamp- Waite, M.E., Ge, S., and Spetzler, H., 1999, A new conceptual model for
shire: American Journal of Science, v. 282, p. 886–919. fluid flow in discrete fractures—an experimental and numerical study:
Sahimi, M., 1994, Applications of percolation theory: London, Taylor and Journal of Geophysical Research, v. 104, p. 13,049–13,059.
Francis, 258 p. Walsh, J.B., 1981, Effect of pore pressure and confining pressure on per-
Scholz, C.H., 1990, The mechanics of earthquakes and faulting: Cam- meability: International Journal of Rock Mechanics, v. 18, p. 429–435.
bridge, Cambridge University Press, 439 p. Walther, J.V., 1990, Fluid dynamics during progressive regional metamor-
Sibson, R.H., 1985, A note on fault reactivation: Journal of Structural Geol- phism, in Bredehoft, J.D., and Norton, D.L., eds., The role of fluids in
ogy, v. 7, p. 751–754. crustal processes: Washington, D.C., National Academy of Sciences,
——1987, Earthquake rupturing as a mineralizing agent in hydrothermal p. 64–71.
systems: Geology, v. 15, p. 701–704. Watson, E.B., and Brenan, J.M., 1987, Fluids in the lithosphere. I. Experi-
——1993, Load-strengthening versus load-weakening faulting: Journal of mentally-determined wetting characteristics of CO2-H2O fluids and their
Structural Geology, v. 15, p. 123–128. implications for fluid transport, host rock physical properties, and fluid
——1996, Structural permeability of fluid-driven fault-fracture meshes: inclusion formation: Earth and Planetary Science Letters, v. 85, p. 497–515.
Journal of Structural Geology, v. 18, p. 1031–1042. Witherspoon, P.A., Wang, J.S.Y., Iwai, K., and Gale, J.E., 1980, Validity of
——2001, Seismogenic framework for hydrothermal transport and ore the cubic law in a deformable rock fracture: Water Resources Research,
deposition: Reviews in Economic Geology, v. 14, p. 25–50. v. 19, p. 1016–1024.
Sibson, R.H., Robert, F., and Poulsen, K.H., 1988, High-angle reverse Zhang, S., Paterson, M.S., and Cox, S.F., 1994a, Porosity and permeability
faults, fluid-pressure cycling, and mesothermal gold deposits: Geology, evolution during hot isostatic pressing of calcite aggregates: Journal of
v. 16, p. 551–555. Geophysical Research, v. 99, p. 15,741–15,760.
Skinner, B.J., 1997, Hydrothermal mineral deposits—what we do and Zhang, S., Cox, S.F., and Paterson, M.S., 1994b, The influence of room
don’t know, in Barnes, H.L., ed., Geochemistry of hydrothermal ore temperature deformation on porosity and permeability in calcite aggre-
deposits, 3rd ed.: New York, Wiley, p. 1–26. gates: Journal of Geophysical Research, v. 99, p. 15,761–15,775.
Smith, C.S., 1964, Some elementary principles of polycrystalline Zhang, S., Tullis, T.E., and Scruggs, V.J., 1999, Permeability anisotropy and
microstructures: Metallurgical Reviews, v. 9, p. 1–47. pressure dependence of permeability in experimentally sheared gouge
Stein, R.S., 1999, The role of stress transfer in earthquake occurrence: materials: Journal of Structural Geology, v. 21, p. 795–806.
Nature, v. 402, p. 605–609. Zhang, S., FitzGerald, J.G., and Cox, S.F., 2000, Reaction-enhanced per-
Stormont, J.C., and Daemen, J.K., 1992, Laboratory study of gas perme- meability during decarbonation of calcite + quartz → wollastonite + car-
ability changes in rock salt during deformation: International Journal of bon dioxide. Geology, v. 28, p. 911–914.
Rock Mechanics and Mining Science Geomechanics Abstracts, v. 29, Zhang, S., Cox, S.F., and Paterson, M.S., in press, Microcrack growth and
p. 325–342. healing in deformed calcite aggregates: Tectonophysics.
Taylor, W.I., Pollard, D.D., and Aydin, A., 1999, Fluid flow in discrete joint Zhang, X., and Sanderson, D.J., 1994, Fractal structures and deformation
sets—field observations and numerical simulations: Journal of Geophys- of fractured rock masses, in Kruhl, J.H., ed., Fractals and dynamic sys-
ical Research, v. 104, p. 28,983–29,006. tems in geoscience: Berlin, Springer-Verlag, p. 37–52.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3808278/9781629490212_ch01.pdf


by guest
on 12 June 2020

Das könnte Ihnen auch gefallen