Subido por yuder espiritu

McCuaig Hronsky 2014 SPEC

Anuncio
©2014 Society of Economic Geologists, Inc.
Special Publication 18, pp. 153–175
Chapter 8
The Mineral System Concept: The Key to Exploration Targeting
T. Campbell McCuaig1,† and Jon M. A. Hronsky1,2
1 Centre
for Exploration Targeting and Australian Research Council Centre of Excellence for Core to Crust Fluid Systems,
School of Earth and Environment, University of Western Australia 6009, Australia
2 Western
Mining Services, Suite 26, 17 Prowse Street, West Perth, Western Australia 6005, Australia
Abstract
To aid conceptual targeting, the past two decades have seen the emergence of the mineral systems concept,
whereby ore deposits are viewed as small-scale expressions of a range of earth processes that take place at
different temporal and spatial scales. The mineral systems approach has been spurred by three main drivers:
the recognition of patterns of mineralization in increasingly available large geoscience datasets; advances in
geographic information system (GIS) technologies to spatially query these datasets; and marked advances in
understanding the evolution of earth systems and geodynamics that provide context for mineralization patterns.
An understanding of mineral systems and the scale-dependent processes that form them is important for guiding exploration strategies and further research efforts.
Giant ore deposits are zones of focused mass and energy flux. Advances in understanding of the physics of
complex systems—self organized critical systems—leads to a new understanding of how fluid flow is organized
in the crust and how high-quality orebodies are formed. Key elements for exploration targeting include understanding and mapping threshold barriers to fluid flow that form extreme pressure gradients, and mapping the
transient exit pathways in which orebodies form.
It is proposed that all mineral systems comprise four critical elements that must combine in nested scales
in space and time. These include whole lithosphere architecture, transient favorable geodynamics, fertility,
and preservation of the primary depositional zone. Giant mineral deposits have an association with large, longlived deeply penetrating and steeply dipping structures that commonly juxtapose distinctly different basement
domains. These structures are vertically accretive in nature, often having limited or subtle expressions at or
above the level of ore deposition.
Three transient geodynamic scenarios are recognized that are common to many mineral systems: anomalous
compression, initial stages of extension, and switches in the prevailing far-field stress. In each of these scenarios,
”threshold barriers” are established which produce extreme energy and fluid/magma pressure gradients that
trigger self-organized critical behavior and ore formation.
Fertility is defined as the tendency for a particular geologic region or time period to be better endowed than
otherwise equivalent geologic regions. Fertility comprises four major components: secular Earth evolution
(variations in the Earth’s atmosphere-hydrosphere-biosphere-lithosphere through geologic history that result
in formation of deposits), lithospheric enrichment, geodynamic context, and paleolatitude (in specific mineral
systems).
The primary depositional zone is usually within the upper 10 km of the Earth’s surface, where large P-T-X
gradients can be established over short distances and time scales. The variable preservation of this zone through
subsequent orogeny explains the secular distribution of many ore deposit types.
The mineral system approach has advantages in exploration targeting compared to approaches that use
deposit models. Emphasizing common ore-forming processes, it links many large ore systems (e.g., VMS-epithermal, porphyry-orogenic gold) that are currently considered disparate deposit models and relates these ore
systems in a predictable way to their large-scale geodynamic context. Moreover, it focuses mineral exploration
strategies on incorporating primary datasets that can map the critical elements of mineral systems at a variety of
scales, and particularly the regional to camp scales needed to make exploration decisions.
Introduction
For centuries the characteristics of mineral deposits have
been studied with two main objectives: to better understand
their geometry and nature to enable their efficient exploitation, and to better understand their genesis to aid prediction
of the location of additional mineral resources (Agricola, 1556;
Loughlin and Behre, 1933). The outcomes of these research
efforts are currently embedded in the literature as deposit
models, and current studies tend to assign deposits to major
groups or subgroups (Hedenquist et al., 2005, and references
† Corresponding
author: e-mail, [email protected]
therein). These models serve as useful frameworks in which
we can compare and contrast deposits and deposit styles, better understand common links in ore genesis, particularly at
the site of deposition, and design detection techniques to find
further analog mineral concentrations (e.g., Hedenquist et al.,
2005).
One important measure of the utility of a model is its ability
to predict the location and quality of undiscovered resources.
Most economic geologists would agree that our understanding of deposit-scale controls on mineral deposition and ore
genesis, albeit incomplete, has improved substantially concomitant with the large amount of research effort directed
153
154
MCCUAIG AND HRONSKY
toward understanding them. A corollary should be that we
are now better at finding new high-quality mineral deposits.
Yet recent reviews of exploration success indicate that this is
not the case, with exploration effectiveness and the quality
of the resource project pipeline declining (e.g., McKeith et
al., 2010). Furthermore, credit for the role of predictive geologic targeting concepts in mineral discoveries has been quite
limited, with most discoveries described as the result of surface prospecting (Sillitoe, 2004). It appears that there is a gap
between our current understanding of ore deposits and our
ability to translate this knowledge into a predictive framework
to find further high-quality resources (McCuaig et al., 2010).
A theme that has emerged over the past two decades to aid
conceptual targeting is the mineral system concept, whereby
mineral systems are viewed as small expressions of a much
larger set of geologic processes that align to concentrate
minerals (Wyborn et al., 1994; Knox-Robinson and Wyborn,
1997). The approach places the mineral deposit within the
rapidly increasing understanding of evolving earth systems
that has advanced substantially over the past two decades
(Groves et al., 2005; Kerrich et al., 2005; Reddy and Evans,
2009; Goldfarb et al., 2010; Cawood and Hawkesworth,
2013). This multiscale approach to mineral systems has substantial predictive power compared to the standard deposit
model concept (McCuaig et al., 2010).
The purpose of this paper is twofold: to outline the current understanding of the mineral system as an organizational
framework to both understand ore genesis and predict the
location of high-quality deposits, and to establish a coherent
language of mineral systems for economic geologists. First,
the evolution of the understanding of mineral systems is
reviewed. Secondly, the principle of ore deposits as the loci
of large fluxes of mass and energy is discussed in terms of
the physics of fluid flow and self-organizing critical systems.
Thirdly, the principal elements of a mineral system, here
defined as whole lithosphere architecture, transient geodynamic triggers, fertility and preservation of the primary
depositional zone, are reviewed with examples taken from
many mineral districts globally across a range of commodities. Finally, links between currently disparate ore deposit
models are illustrated to highlight the potential power of the
mineral system framework over the traditional application of
the deposit model framework.
Evolution of Understanding of Mineral Systems
Deposit models
A traditional tool for understanding ore deposit genesis is
the deposit model. No matter what mineral concentration is
studied, inevitably it is compared to one or more “models”
of deposit styles. Deposit models commonly refer to ”type”
deposits, e.g., Carlin-type Au (Cline et al., 2005), Kambaldatype NiS (Gresham and Loftus-Hills, 1981), Bushveld-type
PGE (Barnes and Lightfoot, 2005), and Witswatersrand-type
Au (Frimmel et al., 2005; Law and Phillips, 2005). These
models are built dominantly from deposit-scale observations.
In the past two decades, more sophisticated deposit models
have included links between different deposit types to make
more comprehensive ”unified” deposit models or model spectrums, e.g., the epithermal deposit spectrum (Simmons et al.,
2005), the broader porphyry-related ore environment (Fig. 1;
Sillitoe, 2010), the crustal continuum model for Archean gold
deposits (Groves, 1993), orogenic gold (Groves et al., 1998;
Goldfarb et al., 2005), and the clastic-dominated (CD), Mississippi Valley-type (MVT) spectrum of lead-zinc deposits
(Leach et al., 2005, 2010).
The above mentioned models are excellent summaries of
variations between deposit types and can be excellent syntheses of deposit-scale processes. From these models has
come an understanding of analog structural, chemical, and
mineralogical footprints of mineralization. Structural analogs
have been useful in mine-scale targeting of similar ore shoots
(e.g., Archean orogenic gold at Norseman, Western Australia; Campbell, 1990), whereas the chemical and mineralogical understanding of alteration halos has been very important
in helping discover further resources in known mineralized
provinces (e.g., porphyry deposits in the American Cordillera,
Fig. 1; Lowell and Guilbert, 1970; Sillitoe, 2000; Sillitoe and
Perelló, 2005; Sillitoe and Thompson, 2006).
The way deposit models are traditionally applied to regional
exploration is to search for geologic situations analogous to
those defined by the deposit model. However, there are several challenges associated with this.
1. Too strict a focus on the analog can result in deposits
being missed because their geologic setting does not have all
the features of the deposit model. However, those missing
features may not be fundamental to the process of ore formation. For example, in the Eastern Goldfields of the Archaean
Yilgarn craton of Western Australia, the earliest gold discoveries in the 1890s were dominantly in mafic rock types. As a
result, exploration targeting up until the 1980s in this region
was heavily biased toward this rock type. This mafic host rockcentric targeting model blinded many explorers to significant
gold concentrations in other rock types. However, in the Yilgarn, the major discoveries in the past three decades have
been in other settings such as late conglomerate sequences
(Wallaby, 3 Moz; Kanowna Belle, 8.5 Moz), sedimentary/volcaniclastic basins (Sunrise, 7 Moz), and intrusive stocks (Boddington, 35 Moz; data from Guj et al., 2011)—all in previously
low-ranked regions. Clearly, the presence of a mafic host rock
is not a fundamental requirement for a major gold deposit.
2. Conversely, targeting focused on the analog characteristics of the model may generate many “false positives” (McCuaig et al., 2009). For example, in the case of targets in mafic
rocks for gold discussed in the example above, there are many
more targets without gold than with gold.
3. For some commodities, there is an overabundance of
defined deposit models. Commonly this develops over time,
as every major new discovery that differs significantly from
previously established analogs results in the definition of
a new variation of the deposit model. Uranium models are
an excellent example of this. Until recently, uranium deposit
classification had 14 models and 22 submodels (International
Atomic Energy Agency, 2000), almost a different model for
each major deposit. In such a classification, there are often
too many variations on a theme for practical application to
exploration (Kreuzer et al., 2010).
4. Most importantly, deposit models struggle to differentiate between large or high-quality mineral concentrations and
155
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
High sulfidation
epithermal disseminated
Au+ Ag + Cu
Phreatic breccia
V
V
V
V
V
Rocks
V
V
V
V
KEY
Intermediate
sulfidation
Au-Ag
V
Volcanic rocks
Carbonate rocks
V
V
V
V
V
V
High sulfidation lode
Cu-Au+ Ag
V
V
+
+
+
+
+
1 Km
+
Porphyry intrusion
Alteration
Steam heated
Intermediate argillic
Quartz-kaolinite
Quartz pyrophyllite
Sericitic
Chloritic
+
Skarn
Potassic
+
+
+
Vuggy quartz + silicification
+
Porphyry
Cu+ Au+ Mo
+
Quartz-alunite
+
1 Km
V
V
Marble
front
+
Marble front
V
V
V
V
V
V
V
V
V
V
V
V
V
V
V
Propylitic
+
Fig. 1. Example of a deposit model showing linkages between porphyry Cu ± Au ± Mo deposits, epithermal deposit types,
and peripheral vein or skarn systems depicting the spatial variation in deposit style with depth, host-rock composition, and
distance from the intrusion. Also depicted is the traditional view of deposit-scale footprints delineated by chemical and mineralogical changes caused by fluid-rock interaction around the deposit. Note that although porphyry Cu deposits are known
to have some of the largest footprints, their halos of alteration are only detectable on the deposit to camp scale. Modified
from Sillitoe (2012).
small or low-quality mineral concentrations. The past half
century of intensive ore deposit research has demonstrated
that big deposits look very similar to small deposits at the
deposit scale—they are just bigger. Small deposits have the
same deposit-scale structural and lithologic settings, fluid or
sequences of fluids, alteration, and metal anomalism as large
metal concentrations.
5. Deposit models also focus on describing how and why
mineralization occurs, but rarely on the spatial prediction of
large, high-quality deposits. Therefore they are commonly of
little use in early-stage exploration environments with little
geologic data (the norm for most undercover exploration).
Therefore, there is a disconnection between deposit models
and their successful practical application to mineral exploration. It is proposed that a large portion of this disconnection
relates to the issue of scale and particularly the identification
of elements of the deposit models that are relevant at different scales of exploration targeting. The deposit model framework often focuses on features at the deposit scale, whereas
many of the critical decisions in mineral exploration occur at
larger scales, prior to the employment of detection technology
at the deposit scale (McCuaig and Hronsky, 2000; Hronsky
and Groves, 2008; McCuaig et al., 2010). At these larger
scales, the deposit models have had limited predictive power
(Sillitoe, 2004; Simmons et al., 2005; Sillitoe and Thompson,
2006).
This scale consideration is best illustrated from the perspective of footprints of mineralization. For the purposes
of this paper, footprints are defined as the recognizable and
mappable expressions of mineral deposits, and it is from
the understanding of deposit footprints that the prediction
and detection methods employed by the exploration industry are derived. Traditional deposit footprints are formed by
the circulation of fluids in surrounding wall rocks that leave a
chemical and mineralogical expression and, in certain cases, a
geophysical expression that can be used to vector toward mineralization. These traditionally defined footprints are commonly small, on the order of tens of square kilometers, and
often much smaller (Fig. 1; Large et al., 2001; Seedorf et al.,
2005; Sillitoe, 2010), and even when a footprint is recognized,
one cannot determine subeconomic or barren footprints from
well-mineralized and economic counterparts.
The search for new high-quality mineral districts is increasingly focused in technically challenging areas below cover
where traditional detection technology is both expensive and
156
MCCUAIG AND HRONSKY
less effective. In these areas, the traditional footprints discussed above will be of limited use in exploration targeting.
It is therefore imperative to understand and recognize the
larger scale footprints of the entire mineral system in order to
effectively target mineral exploration. The largest scale footprint of a mineral system is commonly at the continental (Fig.
2) or transcrustal scale (e.g., Drummond et al., 2006; Snyder,
2013). Such large-scale footprints are never defined by traditional deposit-scale studies.
Therefore, there is a need for an organizing framework to
aid exploration targeting for new high-quality mineral deposits. The framework must encompass the expression of mineral
systems at a range of scales appropriate to the range of scales
of critical exploration targeting decisions.
industry-driven approach, as academics did not initially have
easy access to these datasets that were proprietary and closely
guarded by industry when first available. Second, advances
in GISs and the ability to systematically interrogate large
datasets spurred the need to find common characteristics of
mineral deposits that could be queried to produce conceptual targets alongside empirically driven approaches (Wyborn
et al., 1994; Knox-Robinson and Wyborn, 1997). Finally,
advances in understanding the evolution of earth systems and
geodynamics provided a multiscale context for understanding various expressions of mineralization (Groves et al., 2005;
Kerrich et al., 2005; Goldfarb et al., 2010; Leach et al., 2010).
A mineral system was defined by Wyborn et al. (1994, p. 109)
as “all the geologic factors that control the generation and preservation of mineral deposits [stressing] the processes that are
involved in mobilizing ore components from a source, transporting and accumulating them in more concentrated form,
and then preserving them throughout the subsequent geologic
history.” This seminal paper recognized that ore deposits were
the expression of a large number of geologic processes aligning
to trigger ore accumulation. Furthermore, they emphasized
A systems approach to understanding mineralization
The systems approach to understanding mineralization
emerged as a result of three concurrent factors. First, the
advent in the 1980s of large-scale image-processed geophysical datasets allowed recognition of regional patterns correlating with mineralization (Woodall, 1984, 1994). This was an
Canada
0
Galena-Decorah-Platteville
formations
500
1000
km
Outline of
Upper Miss
District
Zn in deep
oil wells
Viburnum Trend
Old Lead Belt
Zn in deep
oil wells
Bonneterre
Fm
Knox Group
NE
Arkansas
Davis
K
x
no
Gr
o
East
Middle
Tennessee
Tennessee District
District
up
Sinking
Valley
Timberville
Freidensville
Atlantic Ocean
Dolomite Fronts and associated
Zn/Pb mineralization mid-continent USA
Gulf of Mexico
Dolomite Front (hachures on
dolomitized side of front)
Dolomitized rock behind
Dolomite Front
Ore occurrences/significant
mineralization
Fig. 2. Example of a continental-scale footprint of a mineral system. All MVT Pb-Zn deposits in the Phanerozoic basins of
North America occur at the dolomite front (Harper and Borrok, 2007). Therefore, the dolomite front is an example of a highorder mappable expression of the mineral system footprint that can be used to narrow the search space for Pb-Zn deposits at
an early stage of exploration strategy.
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
that by understanding the entire system, one had the potential
to expand the mappable footprint of a deposit.
The mineral systems approach was originally driven by
industry. The earliest documented successful application of a
mineral systems approach (although not called that at the time)
was the discovery by Western Mining Corporation (WMC) of
the Yeelirrie uranium deposit in Western Australia in 1972
(Woodall, 1994), followed by the discovery of Olympic Dam,
South Australia in 1975 (Woodall, 1994; Haynes, 2006). These
conceptually driven exploration programs were actually aimed
at finding existing orebody types—sandstone-hosted uranium
in the case of Yeelirrie, and sedimentary-hosted copper in the
case of Olympic Dam. In the Yeelirrie case, the concept of
having a source of U-rich granites and a sedimentary system
with redox gradients to precipitate U was transferred to the
Archean Yilgarn craton of Western Australia. This program
resulted in the discovery of paleochannel uranium, the first
deposit of this kind discovered globally. In the case of Olympic Dam, WMC geologists had broken down the critical elements of models for sedimentary rock-hosted copper deposits
to a required source of Cu (in this case weathered basalts),
large structural pathways to transport the fluids through the
crust (lineaments recognized in large geophysical datasets),
and an overlying sedimentary basin that could host sequences
of reduced rocks to cause Cu deposition. What the exploration program discovered was an entirely different orebody
type—and a style never before discovered—an iron-oxide
copper gold uranium orebody in the basement beneath the
sedimentary rocks. These examples highlight a key strength
of the mineral systems concept. Because the mineral systems
concept focuses on the critical processes required to generate
a large concentration of metal, it sees through deposit-style
variations to common links between ore systems, in this case
lithosphere-scale architecture, metal source regions, and depositional zones. No evidence of significant metal concentration was required, nor entered into, in the targeting process.
Therefore, this approach to targeting mineralization has the
potential to lead exploration geoscientists into regions with
no empirical evidence of mineralization and to find the yet
undiscovered styles of mineral systems (e.g., Woodall, 1994).
The mineral system approach gained favor in parallel with
advances in GISs, which allowed systematic querying of
datasets for conceptual and empirical expressions of mineral
systems in GIS platforms (Wyborn et al., 1994; Knox-Robinson and Wyborn, 1997; Kreuzer et al., 2010). The original
approach mimicked that used by the petroleum industry since
the 1970s (Magoon and Beaumont, 1991). Three main aspects
of ore formation were considered: sources (of ligands, metals,
fluid, heat, magma), pathways (permeable strata or actively
deforming structures that create permeability and transport
ore-forming fluids/magmas through the crust), and traps
(areas of metal deposition). The key aspects of this approach
were that it was process based (rather than descriptive based),
the processes considered were critical for ore formation, and
the processes were independent of each other. Recognizing
the difference between petroleum systems (a mass trapping
system) and mineral systems (a mass scrubbing process),
McCuaig et al. (2010) modified the mineral system framework
to consider the critical elements of source, pathway, physical throttle (putting a large amount of fluid/magma through
157
a small volume of rock), and mass scrubbing (triggering
metal deposition from the transporting medium, but allowing
flowthrough of the fluid medium). Contrast this mineral system approach to that of the mafic-hosted model for Au in the
Yilgarn, as discussed previously. The latter focused on geologic
characteristics of known deposits, but failed to articulate that
the differentiated dolerites were just one (albeit very good)
example of a physical throttle (brittle dolerite attracts fracturing and permeability) and a chemical scrubber (high Fe/Fe
+ Mg ratio triggers sulfidation and Au deposition; McCuaig
et al., 2010). Yet there are many other favorable depositional
sites in the Yilgarn, as shown by discoveries of the past three
decades noted previously.
There have been many versions of mineral systems proposed (cf. Wyborn et al., 1994; Knox-Robinson and Wyborn,
1997; Lord et al., 2001; Price and Stoker, 2002; Kreuzer et
al., 2008, 2010; McCuaig et al., 2010; Murphy et al., 2011),
all aimed at compositing observations from mineral deposit
models to develop a set of broad encompassing principles
that could be used for conceptual targeting. The challenge
in application of the mineral system approach is typically
scale. McCuaig et al. (2010) outlined the scale dependency
of targeting parameters, highlighting that processes critical
to mineralization on the regional scale have little relevance
on the deposit scale and vice versa. Yet this understanding is
rarely carried through in targeting exercises. In practice, targeting exercises on the regional scale almost always become
a structural-kinematic analysis, intersecting with the spatial
distribution of favorable host rocks, as these are usually the
only elements we can map with any confidence in our regional
datasets. On the belt to prospect scale, targeting exercises
also become a chemical (host rock, alteration, metal anomalism) and structural targeting exercise. Human nature draws
attention to areas with abundant data and where evidence for
mineralization exists (anomalies), irrespective of whether they
are likely to host large concentrations of minerals, and biases
against data-poor areas with no historic evidence of mineralization (McCuaig et al., 2009). Yet at the scale of a craton,
the processes at the site of deposition are not relevant for
exploration targeting (McCuaig et al., 2010). Conceptual targeting is difficult at the larger scale because the processes of
interest may occur within the deep crust or lithosphere, cannot be easily observed, and are much more uncertain. Therefore, a review of the critical elements of a mineral system is
warranted.
Giant Ore Systems as Zones of Focused Mass
and Energy Transfer
What are the critical elements that must come together to
form a mineral system? The original definition by Wyborn et
al. (1994) emphasized that a mineral system requires sources
of the mineralizing fluids and transporting ligands, sources of
the metals and other ore components, a migration pathway
for the fluids, a thermal gradient to drive fluid flow, an energy
source, a mechanical and structural focusing mechanism at
the trap site, and chemical or physical traps for mineralization. However, although this is a good description of the various components of a mineral system, this description does not
explain why these components (which collectively are common in orogens) have sometimes interacted to form a giant
158
MCCUAIG AND HRONSKY
mineral deposit. Because high-quality orebodies are very rare,
the coincident conditions that create them also must be rare.
with the hottest magma and largest magma channel recorded
on Earth (Barnes et al., 1988; Barnes, 2006). The giant Olympic Dam deposit is hosted by a brecciated rock volume of
>20 km3 (Ehrig et al., 2012), which must rank it as one of the
largest breccia complexes globally.
There are some basic physical constraints on what can constitute an ore-forming fluid. First, the fluid needs to have low
viscosity to enable significant mass and energy transfer over
ore-forming time scales (104−105 yrs, Arribas et al., 1995;
Repetski and Narkiewicz, 1996; McInnes et al., 2005; Hickey
et al., 2014). There are three important low-viscosity fluids in
the crust: water, mafic-ultramafic magmas, and hydrocarbons.
Currently economic metal deposits are only known to form
from the first two of these, even though the capacity of hydrocarbons to transport metals has been demonstrated (Emsbo
et al., 2009).
Second, the fluid needs to be available in large quantities
over these geologically short time frames. This places significant constraints on what constitutes a viable process to drive
ore-forming fluids for many systems. For example, it is now
widely accepted that compaction-driven basin dewatering
cannot move fluid at high enough rates to form sedimentary
basin-hosted metal deposits and that instead topographically driven flow of fluid, sourced from the hydrosphere, is
Physical constraints on ore fluids
The most essential constraint on the ore formation process
is that it must take elements at low concentration in large volumes of source rock and deposit them at high concentration
in small volumes of rock (Kerrich, 1983). The only plausible
mechanism to do this is through large-scale advective mass
flux (Fig. 3). This in turn requires the presence of a transporting medium (hydrothermal fluid or magma, both generically
referred to here as “fluid”). This implies that giant ore deposits must be the foci of large-scale systems of energy and mass
flux.
In many cases of magmatic and hydrothermal deposits, an
association between giant ore deposits and large-scale energy
flux can be very clearly demonstrated. For example, Norilsk
is not only the world’s largest nickel deposit, it occurs within
the Permian-Triassic Siberian Traps which is the world’s largest recognized continental flood basalt province and is associated with the largest mass extinction event in the geologic
record (Rampino and Strothers, 1988). The Perseverance
nickel deposit in the Yilgarn, Western Australia, is not only the
largest komatiite-hosted nickel-sulfide deposit, it is associated
A
B
FLUID SINK
Energy Sink
Deposit Scale
Energy Flux released in transient “Avalanches”
Threshold Barrier
Potential
Energy
Gradient
Self-Organised
System
Camp Scale
Focused
fluid exit
conduit
FLUID
RESERVOIR
Fluid
flow
barrier
Entropy
(exported to
environment
as diffuse heat)
Fluid
delivery
pathway
Energy Flux fed into system at a slow rate
Regional
Scale
Energy Source
1. Map
architecture
PRIMARY FLUID SOURCE REGION
Fig. 3. A. Simple model for a self-organized system. Keys to the generation of self-organized critical behavior are that the
energy is continually added to the system and that a threshold barrier is present that prevents the energy from dissipating
to the sink. The energy is added continually and slowly over long time frames, whereas the energy release is via transient
dynamic “avalanches” of much shorter duration. See text for discussion. B. Translation of the self-organized critical system
concept to ore genesis as the focus of a scale-hierarchical mass concentrative fluid (including magma) advection system. Different sets of processes are important at different scales, an aspect that must be taken into account in research and exploration
strategy. See text for discussion.
2. Then map
chemistry
onto
architecture
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
required (Ingebritsen and Appold, 2012, and references
therein). Metamorphism is another process that generates
large amounts of fluid, but does so at a significantly lower rate
(ca. 10−10 ms−1; Skelton, 2011) than required for mineralization (Cox, 1999), and contrasts with the faster processes of
magmatic devolatilization (10−8 ms−1; Cathles and Shannon,
2007; Simmons and Brown, 2008) or topographically driven
basinal flow (10−6 −10−5 ms−1; Cathles, 1990).
Third, fluid-flow systems must be highly organized to produce the required extreme concentrations of metal. Organized
fluids are those that are delivered in a highly focused manner
in both space and time. These conditions cannot be common
because most crustal fluid flow and magma systems do not
produce ore. Although ore fluids may contain metal concentrations that are one to three orders of magnitude greater than
background fluids (Yardley, 2005; Kouzmanov and Pokrovski,
2012), they show a continuum in chemical composition with
background crustal fluids that do not form ore (Yardley, 2005;
Simmons and Brown, 2008). Thus orebodies must represent
moments in time when fluid flow was highly organized (and
potentially transiently supersaturated?) and released as transient pulses to dramatically focus the mass and energy through
the rock mass and trigger ore deposition. Such extreme fluid
flow is characteristic of earthquakes, where transient extreme
permeability increases of five orders of magnitude above
background have been recorded after seismic events (Miller
et al., 2004). The topic of the dynamics of how ore-fluid flow
systems organize is addressed below.
Mineralization as a product of self-organized critical systems
A seminal paper by Bak et al. (1987) proposed a model
to explain the physics of complex natural systems that show
order and organization, despite a wide range of initial starting
conditions. These systems were termed self-organizing critical systems. One of the key examples used by Bak (1996) and
Bak et al. (1987) was a geologic one: the patterns of earthquake occurrence. Hronsky (2011) proposed that ore-forming
fluid-flow systems can also be considered as self-organized
critical systems.
Key factors that control self-organized critical behavior
are that energy is added slowly and continually to the system and that this energy is prevented from dispersing into
a sink by a threshold barrier, thereby creating an enhanced
energy gradient (Fig. 3A). This energy gradient builds until it
reaches a critical point and overwhelms the threshold barrier.
Then energy is released in rapid transient pulses termed avalanches, with the potential size of the avalanches proportional
to the energy gradient it seeks to disperse (Bak et al., 1987;
Schneider and Dorion, 2005). In Bak’s earthquake example,
the energy input is the slow but continual buildup of stress
due to the motion of the earth’s tectonic plates. The threshold
barrier is the brittle upper crust, which prevents stress release
through ductile deformation. The avalanches are earthquakes
that release energy through brittle failure of rock, with the size
distribution of earthquakes proportional to the stress gradient
the system seeks to disperse. As long as the energy gradient
and threshold barrier remain intact, the system will remain
organized around the critical point (in the earthquake example, the brittle-ductile transition). The avalanches in which
energy is released show a distinct set of characteristics: they
159
are multiple events with heavy-tailed size-frequency distributions following scale-invariant power law behavior, are fractal
in geometry (in that they show similar patterns over a range
of scales), and occur on a vastly shorter time scale than the
energy input to the system (Bak et al., 1987; Jensen, 1998).
Ore deposits also demonstrate the characteristics of selforganized critical systems. They show power-law size-frequency relationships (Folinsbee, 1977; Schodde and Hronsky,
2006; Guj et al., 2011) and their spatial distributions exhibit
fractal geometries (Carlson, 1991). Many mineral deposits
present evidence that their formation involved multiple transient pulses of intense fluid flow. This is commonly recorded
as multiple-overprinting generations of veins and brecciation, or layers within a single vein (e.g., Sibson et al., 1988;
Cathles and Adams, 2005; Cox, 2005). Cathles and Smith
(1983) calculated that volumetric flux of brine pulses that
formed Mississippi Valley-type deposits were more than three
orders of magnitude greater than those which could be produced by steady-state basin dewatering, with the most likely
mechanism for these pulses being the rupture of an overpressured fluid reservoir. Consistent with increased fluid flow,
some fault-related damage zones show direct evidence that
transient, extreme fluid permeability, and fluid flow followed
seismic events. For example, in the 1997 Umbria Marche
earthquake sequence in northern Italy, transient, postseismic permeability for a localized zone in the hanging wall of a
major rupture has been estimated as 4 × 10−11m2; this is 105 to
106 times greater than background crustal permeability at that
depth (Miller et al., 2004). At a more regional scale, mineral
deposits form in narrow time frames within the longer lasting
magmatic-deformation-hydrothermal history (energy input)
of the region (Goldfarb et al., 2005, 2014; Sillitoe and Perello,
2005; Sillitoe and Mortensen 2010; Maydagán et al., 2014).
If we consider ore-forming systems to be examples of selforganized critical systems, they represent the particular case
where energy flux occurs primarily as advective fluid (including magma) flux. Self-organized critical behavior is likely to be
essential for ore formation in most magmatic-hydrothermal
systems because it is the only viable physical mechanism in
the crust to produce the concentrated fluid fluxes generally
considered to be required for major metal accumulations
(e.g., Cathles and Adams, 2005; Cox, 2005). Background permeability is far too low for convective fluid flow throughout
most of the crust (Manning and Ingebritsen, 1999).
Implications of self-organized critical systems
for mineral exploration
Fundamentally, the above discussion implies that generation of extreme and anomalous fluid-pressure gradients is
important for producing more focused and larger fluxes of
mass and energy. Such regions will have larger size-frequency
distributions of energy release events and the greatest opportunity for the formation of giant ore deposits.
The most important implication of these ideas for economic
geology is that a localized threshold barrier to fluid flow is an
essential element of ore-forming systems (Fig. 3B). Although
intuition suggests that ore deposits will form in the most dilational parts of the crust, this is not supported by observational
data. Instead, evidence from a range of deposit types including
orogenic gold, porphyry copper, and magmatic Ni-Cu-PGE
160
MCCUAIG AND HRONSKY
sulfide deposits commonly indicates exactly the opposite, with
anomalous localized compressional geodynamics and/or barriers to fluid flow playing an important role (e.g., Sibson et
al., 1988; Czamanske et al., 1995; Cathles and Adams, 2005;
Rohrlach and Loucks, 2005; Begg et al., 2010; Sillitoe, 2010).
The threshold barriers in mineral systems can be physical
features distal to fluid source regions, such as the crystallizing carapaces of intrusions, antiformal culminations, impermeable stratigraphic layers in basins, and the steep-dipping
structural margins of sedimentary basins. The threshold
barriers can also be geodynamic, such as an arc undergoing
transient pulses of increased compression, causing magmas
to pond at depth (Fig. 4; e.g., Rohrlach and Loucks, 2005).
In many cases, this threshold barrier will have a clear spatial identity and enclose an overpressured fluid or magma
reservoir that is episodically ruptured. An important component of any targeting strategy then becomes the identification of potential paleo-overpressured fluid or magma
reservoir sites. These are likely to be much larger targets
than the ore environment itself. It is proposed here that the
scale of these overpressured reservoirs defines the scale of
associated mineral deposit camps (i.e., a cluster of closely
related deposits, see below). For example, the typical scale
of clusters of porphyry copper deposits is 5 to 30 km, similar to the scale of underlying magmatic reservoirs inferred
to drive these systems (Tosdal et al., 2009; Sillitoe, 2010;
Fig. 4).
The energy-release events in ore-forming self-organized
critical systems comprise pulses of overpressured fluid. The
self-organized critical system concept predicts that these
focused pulses of ore fluid will nucleate at sites where the
structural architecture is most conducive to failure driven by
fluid pressure build-up (i.e., “the weakest link” in the barrier).
These pulses will create their own conduits (although usually, but not always, utilizing existing structural weaknesses)
and form pipe-like networks between their source and sink.
Thus, the permeability creation and fluid flow is not driven by
deformation on active structures (Cox, 2005) but by extreme
fluid pressures opening up weaknesses in preexisting architecture beyond the threshold barrier. Depending on the details
of the chemistry of ore deposition in a particular system, ore
deposits will form either within the conduit or where the fluid
discharges into its fluid sink (Figs. 3, 4).
There are consistent spatial patterns for large mineral
deposits that are commonly observed across a range of
mineral deposit types. These include a regional camp-scale
spatial association with zones of localized complexity along
long-lived, large-scale structures and a direct association
between ore-fluid conduits and low bulk-strain fracture
networks (Ridley, 1993). It is also quite common for ore
deposits widely separated in time to form in the same volume of rock (e.g., Kambalda NiS and St. Ives Au, Miller
et al., 2010; Mt. Isa Pb-Zn and Cu, Swager, 1985; Carlin
district Au, Bettles, 2002). These coincidences must have
an underlying explanation and it is proposed that only certain structural architectures may be favorable for the generation of an ore-forming self-organized critical system.
Determining the nature of such architectures should be a
target for further research.
In summary, some key implications for exploration targeting emerge from this self-organized critical concept, with
different relevance at different scales. At the broadest scale,
Limit of lithocap
v
v
v
v
v
v
v
v
v
v
v
Base of degraded
volcanic edifice
Paleosurface
v
v
KEY
v
Late-mineral
Intermineral
Early
Internal Threshold
Barrier (carapace)
Fluid Exit Conduit
Porphyry
Stock
Parental pluton
Composite
precursor
pluton
v
v
Comagmatic volcanic rocks
Subvolcanic basement
5 Km
5 Km
Fluid Reservoir
External Threshold Barrier (transient anomalous compression)
Fig. 4. Diagram illustrating how porphyry deposits may form in exit conduits above a threshold barrier. In this example
the threshold barrier is a combination of a local physical barrier (the carapace of previous crystallized phases of intrusion)
and geodynamic (the transient anomalous compression due to increased far-field stress that clamps the vertical permeability
of the system, shutting off volcanism and causing magmas to pond at depth). Modified from Sillitoe (2010); see also Figure 9.
161
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
recognition of transient geodynamic episodes that create
threshold barriers is critical. At a camp scale, the ability to
define these threshold barriers in space and time becomes
an important element in mineral exploration. Finally, at the
deposit scale, understanding the 3-D architecture is critical
for predicting possible exit conduits. Although ore is formed
within exit conduits, it is not formed everywhere within them.
Therefore we need to recognize pathways when they are not
ore. An important exploration strategy becomes first locating
the fluid exit pathways, then finding the ore within them.
Critical Elements of a Mineral System
McCuaig et al. (2010) proposed a pragmatic methodology
for applying the mineral system concept to exploration targeting. The purpose of this method is to relate the highest
order, generic elements of a mineral system (termed “critical elements”) to practical observations that can be made
in available exploration data sets. The choice of critical elements is important because it organizes the entire process of
applying mineral system concepts to exploration targeting. As
discussed above, different combinations of critical elements
have been applied by various authors and mineral explorationists, including the classic trinity of “source-transport-trap”
borrowed from the petroleum industry (Lord et al., 2001;
Kreuzer et al., 2008, 2010; McCuaig et al., 2010). Although
these are valid, we advocate that the most practical and useful
set of critical elements are lithosphere architecture, transient
favorable geodynamics, fertility, and preservation of primary
depositional zone (Table 1; Fig. 5). In this framework, ore formation results from the conjunction of these critical independent elements.
Fertility
Favorable
Whole-Lithosphere
Architecture
+
Preservation (of
Primary
Depositional
Zone)
Favorable
(Transient)
Geodynamics
Ore Genesis
Fig. 5. Critical elements of a mineral system. Ore deposition occurs as a
conjunction of whole lithosphere architecture, favorable transient geodynamics, and fertility. Postmineralization preservation of the primary depositional
zone is a critical element for ore deposit discovery.
Table 1. Critical and Constituent Elements in the Formation of Gold Deposits1
Critical elements
—————————————————— Scale —————————————————
Fertility
Favorable architecture
Primary depositional zone
Ore-shoot
N/A at this scale
N/A at this scale
Localized dilatant zone in
conduit-hosting structure
Deposit
N/A at this scale
Pipelike rock volume more
favorable for fracturing by
fluid-exit pulse (either local
structural complexity or pipe
of more competent rock)
2nd order—pressure drops;
3rd order—favorable
substrate (chemical
reaction)
Camp
N/A at this scale
Period of low active tectonic
strain, e.g., stress switch causing
transient neutral stress state
causing fluid system to selforganize; areas of greatest uplift
favored (provides stress switch
and high rates of energy and
mass transfer)
Major heterogeneity (e.g.,
cross-structure intersection)
along trend of inverted
rift-axial (or rift-marginal)
fault with associated physical
seal (e.g., antiformal
culmination or unconformity)
Province
Discrete Au-enriched upper
lithospheric domain, particularly
near its margins; potentially
mantle lithosphere enriched by
small volume partial melts prior
to termination of orogeny
Terminal phase of synore
orogenic event (e.g., the transition to incipient extension
associated with the termination of collision and locus of
subduction retreating
oceanward)
Inverted retroarc rift;
preferably developed at a
continental margin, or margin
of deep mantle lithosphere
root; long-lived “vertically
accretive” structure
Continental
A major collisional orogenic
event within the history of an
evolving accretional orogen;
the major collision that actually
terminates a long-lived
(>200 Ma) accretionary orogen
is most prospective and usually
associated with a peak of
supercontinent formation
Major subcontinental scale
lineament (representing longlived zone of transverse
dislocation within accretionary
orogen); long-lived “vertically
accretive” structure
Currently unclear but the
occurrence of the western
US gold superprovince
suggests that some control at
this scale exists
Favorable geodynamics
1st order—upper 10 km
of crust at the time of
mineralizing event where
fluid pressure (+T, X)
gradients are greatest;
preserved through
multiple orogenic cycles
Notes: Note the scale-dependent nature of the constituent elements; these elements require translation into features that can be mapped directly in existing or obtainable geoscience datasets at the appropriate scale to generate targets (e.g., McCuaig et al, 2010)
162
MCCUAIG AND HRONSKY
Favorable whole lithosphere architecture
The control of structure on hydrothermal mineral deposits has long been recognized, from the scale of ore shoots
(Conolly, 1936) to the broad regional scale (Billingsley and
Locke, 1935, 1941; O’Driscoll, 1986). Traditionally, in the academic economic geology community, the focus on structure
has been at the ore shoot-orebody-camp scales, and regional
studies have focused on upper-crustal aspects of major fault
systems and their role as permeability pathways or structural
traps (e.g., anticlines).
The most important and consistent structural pattern in
giant mineral deposit targeting, however, is a spatial relationship between large deposits and inferred fundamental basement structures. This pattern has been recognized since at
least the 1930s and widely applied in the industry (O’Driscoll,
1986; Woodall, 1994). These structures tend to be difficult to
recognize in surface geologic mapping, and historically were
often only recognized as linear arrays of mineral deposits associated with subtle linear patterns of structural discontinuities
(i.e., “lineaments”). Therefore, this association tended to be
treated with skepticism by the academic community. However, this changed in the last decade with the availability to the
academic community of large-scale geophysical datasets that
could image these features to great depth (e.g., Chernicoff et
al., 2002; Crafford and Grauch, 2002; Murphy et al., 2008).
The giant polymetallic skarn deposit at Antamina, Peru, is
an excellent example of deep architecture with only cryptic
surface expression (Fig. 6). Antamina formed at the intersection of an apparently minor thrust and a cryptic transfer fault
in the western Andes at ca. 9 Ma as a result of multiple stress
switches during the Quechuan orogeny (Love et al., 2004;
Fig. 6A, B). However, as shown by Love et al. (2004), the
faults reactivated a thrust architecture that was established
at ca. 42 Ma during the Incaian orogeny. Furthermore, the
Incaian orogeny involved reactivation of structures that had
been established in the Jurassic, at the time when the host
sedimentary sequences were deposited in a back-arc basin off
the western coast of South America. Recent work in the Pataz
region of the eastern Andes by Witt et al. (2013) indicates
that these E-NE-trending orogen-transverse structures were
present on the western margin of South America at least as
early as 300 m.y. ago. Therefore, the transfer fault is a fundamental lithospheric structure that was active for a long period
of time in geologic history, yet is marked only as a series of
brittle joints at the scale of the deposit, and by subtle changes
in structural trends and thickness of stratigraphic units on a
regional scale (Figs. 6C, D, 7).
The Motherlode orogenic gold province of California is
another excellent example. At surface, the host orogen-parallel fault network is an anastomosing series of moderate displacement high-angle reverse faults. However, this network
overlies a fundamental feature in the deep lithosphere as
imaged by Bouguer gravity and seismic tomography data, corresponding to the margin between older Precambrian mantle
lithosphere to the east and Phanerozoic mantle lithosphere to
the west (Bierlein et al., 2008; Griffin et al., 2013).
In the northern Chilean porphyry Cu belt, the important
ore-controlling Domeyko (or West) fault can be imaged by
magnetotellurics as penetrating through the mantle lithosphere (Lezaeta, 2001). It is a small displacement structure at
surface, yet marks a fundamental translithospheric structure
that has been active periodically since at least the Mesozoic
(Padilla-Garza et al., 2001).
Whole-lithosphere architecture also exerts control on the
location of many deposits traditionally viewed as upper-crustal
fluid systems. For example, unconformity uranium deposits in
the Athabasca basin align along a reactivated basement structure corresponding to the margin of the Paleoproterozoic
Trans-Hudson orogeny (Mercadier et al., 2013). The Century Zn deposit and Kupfershiefer Cu-Pb-Zn deposits have
also been demonstrated to be associated with deep-seated
structures (Murphy et al., 2008; Borg et al., 2012). Even Precambrian banded iron-hosted Fe oxide deposits, which show
evidence for hypothermal upgrades before supergene modification (Angerer et al., 2014), correlate with early lithospherescale architecture (Fig. 8; Mole et al., 2013).
More recently there has been the recognition in the academic community (although long known in industry) that the
spatial and genetic relationship of deposits to lithospheric
structures does not just apply to hydrothermal deposits but
also to orthomagmatic deposits at a range of scales (Fig. 8;
Begg et al., 2009, 2010; McCuaig et al., 2010; Griffin et al.,
2013; Mole et al., 2013, 2014). The clearest example of the
control of mantle lithosphere architecture on ore systems is
kimberlites. Kimberlites are notorious for having little correlation with upper-crustal structure, but a strong correlation
with deep structures, as suggested by tomographic and isotopic images of the mantle lithosphere (Begg et al., 2009; Jelsma
et al., 2009).
There are a number of key features that these fundamental ore-controlling structures share that are important for
targeting: (1) they are both strike and depth extensive, usually penetrating into or through the lithospheric mantle as
imaged by geophysical data or interpreted from changes in
isotopic composition of magmatic rocks (e.g., Carlin Trend,
Crafford and Grauch, 2002; Domeyko fault zone in northern
Chile, Lezaeta, 2001; Yilgarn gold and nickel, Mole et al.,
2013, 2014; Fig. 8); (2) they are relatively difficult to trace in
map patterns (at least at the structural level of ore formation)
and are not the obvious structures at or above the level of
mineralization (e.g., Antamina, Love et al., 2004; Fig. 6; St.
Ives, Western Australia, Miller et al., 2010; Fig. 7); (3) they
are never classic flat-dipping thrust zones even though such
structures are some of the biggest crustal faults on Earth, and
usually not the continuous major shear zones that are obvious in regional maps; (4) they have an anomalously low ratio
of displacement to strike length, in comparison to the typical
scaling relationships in neoformed faults (cf. Peacock, 2003),
implicating formation by reactivation of underlying ancestral
structures; (5) they can be demonstrated to be multiply-reactivated (commonly with variable senses of movement) faults
with a very long (commonly hundreds of millions of years or
more) history (Love et al., 2004; Mole et al., 2014; Miller et al.,
2010); and (6) they commonly juxtapose distinctly different
basement domains as imaged by magma chemistry (Loucks,
2014) or isotope chemistry (Mole et al., 2013; Fig. 8). These
important translithospheric ore-controlling structures have
what is termed here as ”vertically accretive growth” histories
(Fig. 7). Major structures at depth can be overlain by younger
volumes of rock that are deposited or obducted into place;
163
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
Deposit
scale
N
A
B
Transfer Zone
1 km
1 km
Regional scale
D
Camp scale
10 km
C
81°W
Peru
8°S
Pasto Bueno
Magistral
Trujillo
LIMA
PACIFIC
OCEAN
9° 30'S
18°S
RD
Basin
Margin
BL
A
CA
ER
AN
ILL
Chimbote
Mineral Deposits
RD
9°S
A
CO
ER
ILL
Fig. 6A,B
69°W
CO
ANTAMINA
0°S
GR
Casma
NE
9°40'S
Pierina
Fig. 6C
Antamina
Huaraz
A
10°S
77°20'W
77°10'W
77°00'W
Pliocene - Quaternary clastic sediments
Middle Eocene - Upper Miocene granitic rocks
Middle Eocene - Middle Miocene volcanic rocks
Albian - Upper Cretaceous carbonate rocks
Upper Jurassic - Lower Cretaceous siliciclastic
rocks
pre-Ordovician Marañón Metamorphic Complex
anticline
syncline
plunge direction
thrust fault
normal fault
strike-slip fault
approx. locus of fold
plunge changes, and
strike of folds and faults
0
100 Huarmey
km
79 °W
78° W
Cenozoic clastic sediments
Lower - Upper Miocene granitic rocks
Eocene - Pliocene volcanic rocks
Cretaceous - Paleogene intrusive rocks
77°W
Albian - Cretaceous carbonate
rocks
Cretaceous volcanic rocks
Mississippian - Cretaceous
siliciclastic rocks
pre-Ordovician rocks
Fig. 6. Multiscale architectural controls on the Antamina polymetallic skarn deposit, Peru. Formed at ca. 9 Ma during a
period of multiple stress switches, the deposit lies at the intersection of an inverted margin of a basin-controlling orogenyparallel fault with a transfer zone that has only a cryptic expression at surface but has been active since at least the Mesozoic.
A. Premineralization deposit-scale architecture established by 10 Ma (McCuaig et al., 2003). B. Synmineralization intrusion
follows preexisting architecture of thrusts and fold axes at mine scale (McCuaig et al., 2003). C. Camp-scale architecture
showing cross orogen, vertically accretive structure that has only cryptic expression at surface expressed as discontinuities
in structural patterns across it (number of thrusts and folds, bends in fold axes); after Love et al. (2004). D. Regional-scale
expression of cross orogen, vertically accretive structures showing control on the distribution of stratigraphic units through
time, indicating that the structures have been active at least since deposition of host rocks in the Mesozoic-Cenozoic; after
Love et al. (2004).
reactivation of the underlying shear zone initially produces
complex anastomosing fractures in the overlying rock volume
as the structure propagates through the overlying strata. Thus
they may have only a subtle en echelon or apparently unconnected (soft-linked) brittle fracture pattern at the current surface of the Earth.
Transient favorable geodynamics
The increasing availability of high-resolution geochronological data over the last decade or so has led to what we
consider one of the most significant scientific discoveries in
the field of economic geology: the realization that ore deposit
formation occurred during very narrow time intervals within
164
MCCUAIG AND HRONSKY
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
V
V
+
+
+
V
+
+
+
+
+
+
+
+
Re-activation and
Upward Propagation
of Basement Structure
Newly Deposited Volcanic/
Sedimentary Strata
Fundamental Basement Structure
V
V
V
V
+
+
+
V
V
+
+
V
+
+
V
+
+
V
V
+
+
+
+
+
V
V
+
+
+
+
+
+
V
V
+
+
+
+
V
+
+
+
+
GEOLOGICAL TIME
Fig. 7. Schematic diagram illustrating the concept of vertically accretive structures. Deeply rooted structures in the lithosphere may have only cryptic expressions at or above the level of ore formation. See text for discussion.
Sm-Nd (εNd):
-9.3 - -6.2
-6.2 - -3.9
-3.9 - -2.2
-2.2 - -1.0
-1.0 - -0.2
-0.2 - 0.5
0.5 - 0.9
0.9 - 1.6
1.6 - 2.4
2.4 - 3.6
Gold deposit
Iron deposit
Nickel deposit
Fig. 8. Epsilon neodymium isotopic map of the Yilgarn craton of Western Australia at 2.7 to 2.6 Ga, showing distribution
of ca. 2.7 Ga nickel deposits, ca. 2.65 Ga gold deposits, and BIF-hosted iron oxide deposits. A strong N-NW-trending gradient in the center of the image is interpreted as representing the margin of the paleocraton at ca. 2.7 Ga, with evolved crust to
the west and juvenile crust to the east. Other areas of more juvenile crust in the western portion of the craton are indicated.
Gold deposits are largely restricted to the margins of more juvenile domains. Komatiite-hosted nickel-sulfide deposits are also
controlled by this whole lithosphere architecture, where high melt volumes were channeled around the SCLM lithospheric
root to the palecraton margin. BIF-hosted iron-oxide deposits are restricted to the margins of stable cratonic roots (domains
of evolved crust), where thicker BIF sequences were present and fluids were channeled to produce hypogene upgrades of the
BIF. Such isotopic maps enable imaging of whole lithosphere architecture through time. After Mole et al. (2013).
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
the much broader evolution of their host terranes. Even more
remarkably, in many cases it can be demonstrated that disparate deposit types, commonly separated by large distances
(100s to, in some cases, 1,000s of kilometers) form in the same
narrow time windows. For example, dating of alteration minerals has narrowed the age of formation of all deposits in the
Juneau orogenic gold system in Alaska to a ~1-m.y. interval
at about 55 Ma (Goldfarb et al., 1991; Fig. 9E, F). Within
the Carlin Trend, Hickey et al. (2014) have shown through
apatite fission track dating that the Betze-Post gold deposits
(ca. 40 Moz of Au) formed in less than 15 to 45 k.y. Yamada
and Yoshida (2011) showed that the Kuroko VHMS deposits
formed within a ~1-m.y. interval at about 14 Ma.
There are now also several good examples of major oreforming events occurring simultaneously over broad areas. In
southeast Australia, orogenic Au mineralization in the western Lachlan orogen of Victoria took place at the same time
(440 Ma) as the formation of the porphyry Au-Cu deposits of
the eastern Lachlan orogen in New South Wales (Squire and
Miller, 2003), which also correlates with the timing of major
gold mineralization in northern Kazakhstan (Goldfarb et al.,
2014). In the southwestern U.S., all of the Au deposits in the
Carlin trend, the Cripple Creek alkalic intrusion-associated
Au deposit and the Bingham Canyon porphyry Cu deposits,
generally considered to represent very different deposit types,
formed in a narrow window of ca. 40 to 30 Ma (Presnell, 1992;
Kelley and Ludington, 2002; Muntean et al., 2011). Furthermore, major porphyry deposits in northern Chile (e.g.,
Chuquicamata, Escondida; Sillitoe and Perelló, 2005) and in
southwest China (Lu et al., 2013) also formed in this narrow
time period, suggesting that this important event was global
in nature.
There is only one plausible explanation for these observations: these critical time horizons must reflect unusual
regional-scale geodynamic settings that are favorable for mineralization and that these favorable geodynamic settings must
be transient, lasting for only short periods of geologic time.
Furthermore, the apparently generic link (i.e., independent
of deposit type) between periods of favorable geodynamics
and ore formation implies that fundamental physical processes must be involved.
As predicted, there is an association between these favorable metallogenic time periods and major periods of geodynamic reorganization. In some cases, this can be shown to be
at a global scale. For example, in the case of the late Eocene
to early Oligocene (40−30 Ma) epoch in the southwestern
U.S., the period of ore formation corresponds closely with
a major geodynamic transition in this region from a long
period of compressional tectonics (the Laramide orogeny)
to a period of slab roll-back and extensional tectonics that
ultimately produces the Basin and Range province (e.g.,
Muntean et al., 2011). However, the simultaneous timing of
mineralization in northern Chile and southwest China and
the fact that this period coincides in time with formation of
the bend in the Emperor Seamount-Hawaian Island chain
(O’Connor et al., 2013) suggest a more global-scale control.
Another example is provided by the Jiaodong gold province
of the North China craton and Motherlode orogenic gold
province of California, both of which coincide with the emergence of the Ontong-Java Plateau and plate reorganization
165
in the Pacific at ca. 110 Ma (Goldfarb et al., 2014). Rohrlach
(2002) elegantly demonstrated that the formation of the giant
Neogene Tampakan porphyry and high-sulfidation Cu-Au
deposit on Mindanao formed within a few 100-Kyr period
during an anomalous peak of compression that was the result
of a subduction zone flip. Saunders et al. (2008) showed that
the formation of the Miocene northern Nevada Bonanza gold
province (including important deposits such as Sleeper) was
closely associated with the initial impact of the Yellowstone
hot spot; no mineralization is associated with the subsequent
hot-spot track. Goldfarb et al. (2005) demonstrated that the
formation of the Paleocene Juneau orogenic gold deposits was associated with a large-scale tectonic switch from
orthogonally convergent to transpressional tectonics (Fig.
9E, F). The geodynamic context of the period of formation
of the Miocene Kuroko VHMS deposit is a tectonic reorganization that terminates back-arc rifting in the Kuroko rift
(Yamada and Yoshida, 2011). Jelsma et al. (2009) showed that
kimberlite swarms correlate with periods of major tectonic
plate reorganization. Rosenbaum et al. (2005) correlated the
temporal and spatial distribution of ore deposit formation in
the central Andes mountains to the oblique subduction of the
Nazca ridge (Fig. 9C, D).
It is easier to recognize the precise geodynamic context of
these metallogenically favorable events in young rocks. However, Idnurm (2000) showed that the periods of formation
of the giant Mount Isa and Olympic Dam Proterozoic metal
deposits in Australia corresponded with periods of major
“bends” in the apparent polar wander paths for Australia that
might plausibly be interpreted to represent major periods of
tectonic reorganization. Robert et al. (2005) demonstrated
that Archean orogenic gold deposits tend to occur close in
time to the termination of their host orogenic belts.
What are the processes that link periods of tectonic reorganization with ore formation? This is a critical question and
we propose a hypothesis that relates these observations to the
self-organized critical system concept discussed above. We
propose that these favorable periods are times when the prevailing geodynamic conditions impose strong threshold barriers to fluid flow, causing fluid-flux systems to become highly
organized (e.g., Fig. 3; Hronsky, 2011). These are cases where
there is both a strong fluid supply and a lack of active, pervasive vertical permeability, the latter of which is due either to
tectonic strain that clamps vertical permeability, or simply to
the absence of active pervasive fracturing of the rock mass to
generate many possible fluid escape paths (Fig. 9B, D, F).
In these scenarios, the only way fluids can escape is by the
buildup of extreme fluid pressure and subsequent organized
and focused fluid expulsion through transient exit pathways,
using preexisting weaknesses in the rock mass. This situation
leads to the extreme fluid focusing over short timeframes
(104−105 yrs) required to form ore deposits and is likely quite
rare in the history of orogens.
Central Italy provides a good example for the regional-scale
relationship between fluid-flux patterns and tectonic regime.
This region is characterized by a significant background flux of
mantle-derived CO2. Chiodini et al. (2004) presented the first
map of regional-scale variation in this CO2 flux and showed
that a broad, diffuse flux characterized the extensional setting
of the Tyrrhenian back-arc basin. The anomalous flux of CO2
166
MCCUAIG AND HRONSKY
A
B
Bismarck Sea Spreading Ridge
Inactive Outer
Melanesian Arc
Transient zone of
incipient rifting
(determines
extent of “camp”)
Established rifting limited high-quality
ore formation
Epithermal (e.g. Lihir)
Sea level
Lihir
Grasberg
Ok Tedi
Porgera
VMS
(e.g. Solwara)
Limited active
permeability
creation
Extreme energy
and fluid pressure
gradients produced
System selforganises and
high-quality ore
bodies produced
Solwara-1
Crust
SCLM
Melt zone
Small volume alkalic melts
Asthenosphere
C
80 W
300
Inverted orogen-parallel vertical Orogen-parallel limit of
accretive structure controls ‘camp’ anomalous compression
Fig 1
limits extent of ‘camp’
Anomalous
transient
compression
Crust
Moho
Magma ponds at
depth: volatiles +
Oc
ea
metal content
nic
increases
pl
Large Hg deposit
tion
10 S
D
Large Cu deposit
duc
Km
70 W
KEY
Large Zn deposit
sub
0
8 - 6 Ma
e
zon
Arc volcano
ate
/ ri
dg
e
au
Ri
OLM
e
LAB
Melt zone
zc
Asthenosphere
Moho
LAB
Na
E
Vertical permeability
‘clamped’
Extreme vertical
energy gradient
produced
SCLM
a
15 S
dg
Fig 4
F
Fig. 9. Examples of common transient dynamic settings that trigger ore formation. A. Incipient rifting in the Bismark
Sea triggering both Au-rich VMS deposits (Solwara) and epithermal Au (Lihir). After Hronsky et al. (2012). B. A schematic
section showing how incipient rifting causes self organization of the fluid-flow system. C. Reconstruction of the subduction
history of the Nazca Ridge (thick black outline) at 8 to 6 Ma imposing transient anomalous compression in the arc, cessation
of volcanism, and the formation of ore deposits. Symbols represent volcanoes and ore deposits formed at this time. Note
that this wave of anomalous compression and accompanying mineralization has migrated along the arc due to the oblique
subduction of the ridge. From Rosenbaum et al. (2005). D. Schematic cross section illustrating how transient anomalous
compression causes self organization of the fluid flow system. Relative position of Figures 1 and 4 are shown. LAB = lithosphere-aesthenosphere boundary, OLM = oceanic lithospheric mantle, SCLM = subcontinental lithospheric mantle. E and F.
Example of a stress switch controlling Au mineralization in the Juneau belt in Alaska, after Goldfarb et al. (2005). During the
stress switch, active permeability ceases, high fluid pressure gradients form, and the system self organizes to form ore. After
the stress switch, active permeability creation resumes in the new stress regime, and ore formation ceases. Common to all of
these situations is a transient period where vertical permeability is clamped or active permeability creation ceases, yet energy
and fluid is still supplied to the system. The result is the build-up of extreme fluid pressure and energy gradients, triggering
extreme self-organized critical behavior and the formation of ore during energy avalanches. The areal extent of the threshold
barrier intersecting vertically accretive structures determines the scale of camps.
167
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
Diffuse Low-Grade
Metal Anomalism
Ore Formation
Diffuse Low-Grade
Metal Anomalism
SINK
SINK
SINK
FLUID FLUX
Transient focused
ore-forming flux
FLUID FLUX
is absent in the adjacent Apennine orogenic belt even though
it is the cause of most of the seismicity which characterizes this
part of Italy. This is because at depth, the CO2 accumulates
in crustal traps, generating overpressurized reservoirs which
induce seismicity. This is analogous to the self-organized critical process discussed above.
At least three physical scenarios are relevant to this discussion (Fig 9): (1) the initial stages of extensional events and/
or plume/hot-spot impacts (e.g., Fig. 9A, B; Kelley and Ludington, 2002; Saunders et al., 2008) are settings where pervasive crustal-scale permeability has not yet been established;
(2) transient anomalous compression, which effectively seals
an active system (e.g., Fig. 9C, D; Rohrlach, 2002; Rohrlach
and Loucks, 2005), is a setting that is important in forming
porphyry deposits; and (3) switches in the prevailing far-field
stress that transiently result in a neutral stress field and result
in the absence of pervasive deformation-induced permeability (Richard Tosdal; pers. commun., 2009). All of these scenarios are likely to occur in the context of much larger scale
patterns of geodynamic reorganization.
The significance of threshold barriers may also explain the
general absence of significant mineralization from strongly
extensional environments, which lack the large threshold
barriers compared to those established in compressional settings. A well-documented example is the Andean copper belt,
where the three most productive belts with the largest and
highest grade deposits correlate with brief periods of anomalously strong compression. The deposits associated with periods of more extensional and less compressional tectonics are
smaller and of lower grade (Sillitoe and Perelló, 2005). In this
context, it is important to distinguish between actively extending and previously extended settings. Many deposits form in
rifts but this does not necessarily indicate they form during
periods of significant active rifting. For example, although
VHMS deposits are considered to be rift associated, deposits
within the classic Neogene Kuroko VHMS province (arguably among the best constrained because of their young age)
formed during a period of geodynamic reorganization that
actually marks the termination of the extensional history of its
host rift (Yamada and Yoshida, 2011).
This hypothesis of transient geodynamic triggers for ore formation predicts that for most of the history of a potential oreforming system, ore deposits do not form and that the only
manifestation of the system might be broad regional alteration
and metal anomalism (Fig. 10). When the geodynamic threshold barrier is established, extreme fluid pressure gradients are
produced, and the system self organizes to form ore (Figs. 3,
9B, D). When the barrier is removed, ore formation ceases
and the system reverts to broad regional fluid flow, alteration,
and metal anomalism (Fig. 10). This punctuated history of ore
formation within a broader background context is best documented for the Central Andean porphyry province. Although
this region has continuously been an active site of subduction
and arc magmatism since at least the Mesozoic, porphyry Cu
mineralization in Chile is restricted to several brief windows
on the order of millions of years, preceded and succeeded by
longer periods of barren magmatism, deformation, fluid flow,
and alteration (Sillitoe and Perelló, 2005). Importantly, this
model for fluid flow implies that broad regional alteration and
metal anomalism, currently considered a vector to ore, may
THRESHOLD
BARRIER
SOURCE
SOURCE
SOURCE
Temporal Evolution of System
Fig. 10. Schematic diagram showing that ore deposition occurs during
brief intervals within much longer episodes of deformation, magmatism, fluid
flow, and alteration that is largely barren. Mineral systems only transiently self
organize to form ore.
have no direct relationship to the formation of high-quality
ore deposits. The system needs to have undergone self organization to form high-quality ore. Therefore, exploration geologists must focus on identifying at various scales the factors
that represent geodynamic threshold barriers and the resulting highly organized fluid flow that forms high-quality ore.
The key to applying the geodynamic concept in regional
targeting is to identify geologic proxies for metallogenically
favorable geodynamic epochs. The most practical way to do
this is to determine the spatial distribution of rocks that most
closely correlate in time with the mineralizing event of interest. For example, Robert et al. (2005) discussed the spatial
and temporal relationship between formation of late tectonic
sedimentary basins and Archean orogenic gold deposits.
Although these sedimentary rocks are not necessarily host
rocks to the gold, they represent the unit that formed closest
in time to the gold deposits. In this example, these deformed
late tectonic sedimentary rocks are the best available proxy for
synmineralization tectonic activity. This concept implies that
improved precision in geochronology will increase our predictive targeting capability. Within compressional tectonic scenarios, there is commonly an empirical correlation between
uplift and the location of large mineral systems. For example,
within the young porphyry belt of West Papua New Guinea,
the Grasberg deposit (largest Cu deposit in the belt) is located
near the highest elevation. Similarly, the Miocene porphyry
systems of Central Chile are located near the highest elevations within the Andes Mountains. These areas of highest elevation may indicate areas of stronger anomalous compression.
Ancient deposits may show similar relationships; for example,
the Kalgoorlie and St. Ives goldfields in Western Australia are
located along the axis of maximum uplift, as measured by the
exposure of basal stratigraphy in the middle of the Kalgoorlie terrane (Robert et al., 2005). Similarly, the Timmins gold
camp in the Superior province of Canada is located where the
lowermost stratigraphy is uplifted and structurally repeated
in the center of the Abitibi southern volcanic zone, indicating
an area of major uplift along the translithospheric PorcupineDestor fault (Robert et al., 2005). In these ancient cases, areas
of relatively higher uplift within the primary depositional zone
168
MCCUAIG AND HRONSKY
may have corresponded with areas of higher elevation and
stronger anomalous compression.
Fertility
It has become increasingly obvious that there are specific
tectonic scenarios that are triggers for major ore-forming
events. Yet during these events, not all regions contain worldclass mineralization. Therefore, another critical element must
be present, and that is—fertility. Fertility is defined here as
the tendency for a particular geologic region or geologic time
period to be systematically better endowed than otherwise
equivalent geologic environments. Fertility is usually the
highest order (largest scale) control on endowment potential.
Four factors may contribute to fertility: secular Earth evolution, lithosphere enrichment, large-scale geodynamic processes, and paleolatitude (in specific cases).
Secular earth evolution: Secular patterns in ore deposit
distribution have long been recognized (Turneaure, 1955;
Meyer, 1981, 1985, 1988; Barley and Groves, 1992; Goldfarb
et al., 2010; Cawood and Hawkesworth, 2013; O’Neill et al.,
2013b) and are summarized in a special issue of Economic
Geology (Goldfarb et al., 2010). Key drivers of secular variation in mineral systems include cooling of the earth, evolution
of the atmosphere-hydrosphere, evolution of the biosphere,
and evolution of global geodynamics and supercontinent
cycles. The evolution of the Earth’s lithosphere-hydrospherebiosphere-atmosphere controls the availability or mobility of
metals (Hazen et al., 2008; Hazen, 2010), and the geodynamic
triggers for fluid/magma flow and mineralization.
Although there are conflicting arguments, planetary physical models and the rock record suggest that the Earth has
been cooling over time; for example, since the Archean, there
has been a decrease in the abundance and calculated melting temperature of ultramafic magmas reaching the earth’s
surface (Davies, 1995). The mineral systems most directly
affected by this cooling are orthomagmatic deposits such as
NiS and PGEs. For example, komatiite-hosted NiS deposits
first enter the rock record at ca. 3.0 Ga, peak at 2.7 Ga, and
are not found in rocks younger than 1.9 Ga (Naldrett, 2010).
The atmosphere and hydrosphere have undergone dramatic and rapid changes through Earth’s history. A rise in oxygen occurred at the end of the Archean (the Great Oxidation
Event) and again at the end of the Neoproterozoic (Kump,
2008; Farquhar et al., 2010). This oxygenation of the atmosphere had three major effects on mineral systems, including
influencing (1) the available valence states for metals of interest, (2) the stability of host minerals, and (3) the availability of
ligands (Hazen et al., 2008; Hazen, 2010). For example, uranium is generally mobile in +6 valence state, and immobile in
+4 valence state. Prior to the Great Oxidation Event, primary
uranium minerals (e.g., pitchblende) were stable in the relatively reducing hydrosphere and atmosphere, and as a consequence, survived in placer deposits (e.g., Witswatersrand,
Elliot Lake). After the Great Oxidation Event, primary uranium minerals were subject to weathering and oxidation,
and uranium with +6 valence became available for transport
in an oxidizing hydrosphere (Cuney, 2010). It is therefore
no coincidence that the highest grades of uranium globally
(unconformity uranium deposits) occur in Paleoproterozoic
basins. This was the first time that major large sedimentary
basins were stable (at the terminal stage of a supercontinental
cycle) with U-rich detritus and basement rocks that had never
before been leached of their uranium.
The oxidation of the hydrosphere and particularly its effect
on the redox state of sulfur had a major impact on the availability of metal transporting ligands (Farquhar et al., 2010).
Before the Great Oxidation Event, the atmosphere and hydrosphere were relatively reduced and the hydrosphere was Fe
rich, with sulfur largely sequestered as sulfide in shales. After
the Great Oxidation Event, the oceans were scrubbed of
their iron and there was enhanced oxidation of sulfides in the
crust that provided sulfate to the hydrosphere. This chemical transition had a major effect on some mineral systems.
For example, an important factor in the development of large
clastic-dominated or SEDEX Pb-Zn deposits after the Great
Oxidation Event was the compositional and redox gradients in
the ocean and the availability of metals and sulfate provided
by the oxidative weathering of the crust (Leach et al., 2010).
Evolution of the biosphere was coupled to the evolution of the atmosphere and was characterized by periodic
“spurts” of life. Examples include the organic bloom at 2.2
to 2.0 Ga, marked by widespread generation of organic-rich
shales (Papineau, 2010), and the emergence of reef-building
carbonate-secreting organisms in the Phanerozoic (Leach et
al., 2010). The former had an effect on the expression of sediment-hosted deposits, forming thick sequences of reduced
shale. Moreover, these organic-rich rocks may have preferentially sequestered metals from the seawater column (Tomkins, 2013), subsequently forming preferential source rocks
for metals when buried and metamorphosed (Tomkins, 2013).
The latter formed host carbonate sequences for MVT-style
deposits (Leach et al., 2010).
Lithosphere enrichment: Some areas of the lithosphere
appear to be intrinsically enriched in a target metal. This
concept has long been applied empirically in the mineral
exploration industry. It is clearest in systems where metal is
sequestered from nearby protoliths, such as a uranium-rich
hinterland for sediment-hosted uranium mineralization, or
substantial sequences of banded iron formation for sedimenthosted Fe oxide deposits. Some researchers have also speculated on lithosphere enrichment for more complex mineral
systems. For example, Hodgson (1993) suggested that gold
endowment in Archean greenstone belts was related to the
abundance of komatiites that preenriched the succession in
gold. The first rigorous analysis of the lithosphere enrichment
concept was provided by Titley (2001). Further strong support for this idea was provided by Sillitoe (2008), who showed
that the bulk of the gold endowment of the American (north
and south) Cordillera, regardless of age, is restricted to a few
segments of this long-lived orogenic system of approximately
20,000-km strike length. Hronsky et al. (2012) developed
these ideas further, proposing an integrated process model
that related regions of long-term metasomatic enrichment of
gold in the lithospheric mantle to regions of persistent upper
crustal Au endowment. Pettke et al. (2010) have also convincingly demonstrated that lithosphere enrichment may be relevant to the genesis of the giant Bingham Canyon porphyry
Cu deposit. Zhang et al. (2008) and Griffin et al. (2013) have
suggested that lithospheric mantle enrichment may also play
a key role in the formation of orthomagmatic Ni-Cu-PGE
169
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
deposits, although this currently remains controversial (cf.
Arndt, 2013). Whatever the precise mechanisms, it appears
that large-scale lithosphere control on fertility is real. This
enrichment is independent of deposit type, in that different
styles of mineralization are generated from the same seemingly enriched volume of lithosphere in many cases, e.g.,
southwest U.S., eastern Papua New Guinea (Fig. 9A; Sillitoe, 2008; Hronsky et al., 2012). The enrichment process
also transcends association with any one ore-forming event,
as some terranes show multiple periods of mineralization in
the same location (Robert et al., 2005; Hronsky et al., 2012).
The implications of preferential lithosphere fertility are
significant for our understanding and targeting of ore deposits. For regions where lithosphere fertility may be a factor,
metallogenic potential is not solely a property of a particular
magmatic or tectonostratigraphic event. Figure 11 shows this
concept schematically. For these regions, we can no longer
rank an entire belt or province as prospective or nonprospective. Instead, areas along strike of a poorly endowed segment
of a belt might host large ore systems, and vice versa, if there
is a significant change in the fertility of the underlying lithosphere. These zones of differing fertility will be bounded and
defined by the whole lithosphere architecture, hence the
importance of constructing maps of the deep lithosphere to
aid mineral potential assessment.
Fertility and large-scale geodynamic context: It is instructive to view the critical element of fertility in a geodynamic
context. Concomitant with the cooling of the Earth has been
a change in the mechanism by which the Earth transfers its
heat to space. It has been increasingly accepted that the early
Earth consisted of a relatively stagnant lithosphere periodically breached by plume tectonics and affected by meteorite
bombardment, which eventually gave way to more steady-state
Infertile
Lithospheric Domain
subduction and plate tectonics (Debaille et al., 2013; O’Neill
et al., 2013a, b). Within this evolution the earth has undergone
major geodynamic cycles, often referred to as supercontinent
cycles, although their exact nature prior to 2.7 Ga is strongly
debated (O’Neill et al., 2013a). The link between mineral
deposits and such cycles has been recognized and gave rise
to the first empirical observations of the association between
different mineral deposit types and specific tectonic regimes
(Meyer, 1981, 1985, 1988; Barley and Groves, 1992; Groves
et al., 2005; Kerrich et al., 2005; O’Neill et al., 2013b). Examples include the association of global ”orogenic” Au formation with the terminal stages of accretionary orogenic events
and supercontinent formation (Goldfarb et al., 2005; Groves
et al., 2005; Kerrich et al., 2005; Robert et al., 2005) and the
association of mafic-hosted NiS deposits with peaks in supercontinent formation (Begg et al., 2010). The most important
period of orthomagmatic sulfide mineralization and associated plume activity (e.g., Norilsk, Russia, and the Emeishan
flood-basalt associated deposits of southwest China) and the
most important period of orogenic gold deposit formation
(e.g., the Altaids province of Central Asia) in the Phanerozoic occurred broadly coeval with the final assembly of Pangea (ca. 300−250 Ma). In the case of orthomagmatic deposits,
this association was most likely related to the accumulation of
subducted lithosphere on the core-mantle boundary during
supercontinent assembly, which was subsequently reactivated
as a series of major plumes (Begg et al., 2009, 2010). In the
case of orogenic gold provinces, the precise reason for this
particular geodynamic association remains unclear.
An example of a long-lived fertile geodynamic setting is the
western margin of South America, beginning in the mid-Cretaceous. Rapid opening of the Atlantic ocean in the mid Cretaceous (~100 Ma) resulted in the Andean margin switching
Strongly Fertile
Lithospheric Domain
Weakly Fertile
Lithospheric Domain
Orogen-parallel vertically
accretive structure
BARRE
N
Outline of Metallogenic Province
(eg, Magmatic Arc)
STRONGLY ENDOWED
WED
Y ENDO
WEAKL
BARRE
Orogen-transverse vertically accretive structures
N
STRONGLY ENDOWED
WED
Y ENDO
WEAKL
Cover
Fig. 11. Schematic diagram illustrating how fertility affects targeting strategy.
170
MCCUAIG AND HRONSKY
from an extensional (i.e., formation of the Tarapaca back-arc
basin during the Jurassic) to an increasingly compressional
mode (e.g., Chen et al., 2013). The western margin of South
America is currently characterized by the largest extent of flat
subduction on Earth (Lallemand et al., 2005; Syracuse and
Abers, 2006), which exemplifies the extent of anomalous compression. Following this tectonic switch, a series of arc-related
Cu-Au metallogenic events occurred along this margin (e.g.,
Sillitoe and Perelló, 2005). Each of the metallogenic events
correlates with a brief local period of even stronger compression due to factors such as subduction of anomalously thick or
buoyant lithosphere (e.g., Fig. 9C; Rosenbaum et al., 2005).
Therefore, the metallogenic events may be controlled by
nested scales of geodynamic control: (1) a broadly anomalous
compressional plate margin, and (2) local, transient increased
compression resulting in clamping of vertical permeability,
extreme self-organized critical behavior, and high-quality ore
formation (Fig. 9C, D).
Paleolatitude: Paleolatitude may have an effect on fertility
for some mineral systems, particularly those that require highsalinity fluids. For example, Leach et al. (2010) showed that
Phanerozoic clastic-dominated Pb-Zn deposits, when spatially reconstructed to time of formation, dominantly formed
between 5° to 30° of the equator—the ”arid zone” with high
evaporation rates conducive to the formation of evaporites and
high-salinity brines. These dense residual brines penetrate
deep into the crust and leach metals along their flow paths
until expelled by tectonic triggers where they deposit metals
upon encountering reducing horizons (e.g., organic matterrich shales). High-salinity ore fluids responsible for deposition
of other deposit types may have formed by leaching of evaporites by meteoric fluids after evaporite formation. Regardless
of how such brines formed, there is increasing evidence of a
role for evaporate-related brines in other near-surface metal
accumulations such as BIF-hosted Fe deposits (Evans et al.,
2013), or sediment-hosted uranium systems (Kish and Cuney,
1981; Richard et al., 2011).
Preservation of primary depositional zone
There are factors about ore deposition that are common for
a range of deposit styles, within and among commodity types.
In both hydrothermal and magmatic systems, metal deposition involves destabilization of the metal-ligand complexes in
the transporting fluid. This destabilization is generally triggered by a physical or chemical change, for example, pressure
drop, cooling, or rapid interaction with other rocks or fluids
out of equilibrium with the transporting fluid (e.g., M
­ cCuaig
and Kerrich, 1998). The focus of metal deposition rapidly into
a small volume of rock involves large P-T-X gradients over
relatively short distances and times (e.g., Knox-Robinson and
Wyborn, 1997). Therefore, the primary depositional zone
for most metal systems resides in the upper portions of the
Earth’s crust, broadly defined as the upper 10 km, as this is the
zone with both the highest geothermal and pressure gradients
(and in particular the zone where phase separation becomes
more common in magmas and fluids to drive dramatic chemical changes in the transporting media).
Within this primary depositional zone, high-quality ore formation may be a natural consequence of the self-organized
critical nature of the system. In other words, if enough mass
and energy is focused in avalanches above a threshold barrier, metal will probably deposit somewhere along the exit
pathway. Thus, determining sites of self-organized critical
behavior is the critical goal in regional targeting. Preferred
depositional sites are only pertinent at the camp to prospect
scales (Table 1; McCuaig et al., 2010).
In addition to predicting potential sites of mass and energy
transfer, self-organized critical behavior, and potential formation of high-quality ore deposits, it is necessary to know if the
potential deposits are exhumed close to the current surface of
the Earth but not exhumed to the extent that they are eroded
and not preserved. Currently available exploration technology dictates that, outside the immediate vicinity of major
mines, deposits can only be located if they are present within
a few hundred meters of the land surface. Given that most
metal deposits form at significantly greater depths below their
paleosurface, this implies that mineral discovery also requires
a favorable postdepositional geologic history.
A general empirical observation is that different deposit
styles have different preservation potentials. For example,
Kessler and Wilkinson (2006) demonstrated that an increasing
statistical probability of erosional removal with age is consistent with the temporal distribution of porphyry and epithermal deposits (which are largely restricted to the Cenozoic).
However, this does not explain the temporal distribution of
orogenic gold and VHMS deposits, which instead may be
related to the different tectonic settings of formation of these
different deposit types (see summary in Hronsky et al., 2012).
Those deposits that form in actively uplifting regions (e.g.,
porphyries) have low-preservation potential, whereas those
deposits that form in rifts (e.g., VHMS deposits) or at significant depths late in the history of the orogen (e.g., orogenic
gold deposits) have relatively higher preservation potential
(Groves et al., 2005; Kerrich et al., 2005). Lithospheric architecture is also an important factor in deposit preservation.
Deposits that either form within thick stable cratonic lithosphere or are incorporated into such lithosphere at the end
of their host orogenic cycle, without exhumation and erosion,
are likely to have a high probability of long-term preservation
(e.g., Groves et al., 2005; Kerrich et al., 2005).
Implications for Grouping of Ore Deposits
to Aid Targeting
One important implication of the mineral system concept is
how it highlights connections between different deposit types.
Of critical importance is to recognize that apparently disparate
deposit types can exist as a continuum, related in a predictable way based on the processes controlled by the host geodynamic environment. For example, it has become increasingly
clear that a continuum exists between rift-related epithermal
deposits and VHMS deposits, relating to degree of extension
and water depth in the host rift (Fig. 9A, B; see summary in
Hronsky et al., 2012). The progression from epithermal to
VMS-epithermal hybrid to classic VMS mineralization with
increasing crustal thinning and water depth is exemplified by
the Alexander Triassic metallogenic belt in Alaska and British
Columbia (Taylor et al., 2008). An important pragmatic implication of understanding this particular continuum is that Aurich hybrid VMS-epithermal systems, such as Henty (Halley
and Roberts, 1997) or Eskay Creek (Sherlock et al., 1999), are
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
likely to be associated with the transition from the epithermal
to VMS environment.
Viewing ore deposits within a broader geodynamic context
allows large-scale metallogenic patterns to be discerned. For
example, MVT Pb-Zn deposits and orogenic gold deposits
would normally be considered very different ore types and
almost certainly have very different fluid sources. However,
because MVT deposits form in orogenic forelands, which represent the more distal parts of orogens (Bradley and Leach,
2003), and orogenic gold deposits form in the more proximal
parts of orogens (e.g., Goldfarb et al., 2005), it is plausible
that both deposit types could have formed broadly synchronously during the same metallogenic event. Possible examples
of such are MVT deposits of the Earaheedy basin of Western
Australia that have been dated at about 1.79 Ga, which is also
the main period of Paleoproterozoic orogenic gold mineralization in Western Australia (Pirajno, 2004). Similarly, the Solwara Au-rich sea-floor VMS deposit and the giant epithermal
Au Lihir deposit, two very different deposit types, are potentially related to the same volume of enriched lithosphere and
same transient geodynamic trigger (Hronsky et al., 2012).
Conclusions
The understanding of mineral deposits has evolved from
describing ore specimens at hand-sample scale, to identifying
specific favorable host rocks, to determining structural control on fluid flow (Conolly, 1936; McKinstry, 1941, 1955), and
to understanding broad metasomatic processes (Korzhinskii,
1968; Bohlke, 1989). This evolution led to the development
of analog deposit models that incorporate the processes of
ore formation. More recently, deposit models have evolved
into a systems-based understanding of mineralization processes within the larger context of planetary geodynamics and
secular Earth evolution (Groves et al., 2005; Kerrich et al.,
2005; Goldfarb et al., 2010; Leach et al., 2010; Cawood and
Hawkesworth, 2013; O’Neill et al., 2013b). Thus, the understanding of the mineral system has evolved from observation
to taxonomical classification to a predictive understanding of
processes, much like that for many other natural systems (e.g.,
tectonics, biology).
The mineral systems framework presented here builds
on decades of mineral deposit research. This framework
views ore-forming processes as a conjunction of the critical
elements of metal fertility, whole lithosphere architecture,
and transient geodynamic events to trigger organization of
extreme fluid-flow and metal deposition, followed by preservation of the primary depositional zone to allow discovery
(Table 1; Fig. 5). These critical elements operate in nested
scales of both time and space. This mineral systems approach
has significant predictive power compared to the traditional
deposit model paradigm, particularly as mineral exploration
moves into regions under cover and to greater depths.
It is hoped that the mineral system concept will inform
both future exploration strategies and research efforts. From
a research perspective, the mineral systems approach allows
identification of the highest value fields on which to focus
limited intellectual and financial resources. Key areas that
emerge are (1) better integration of geology and geophysics
to image multiscale whole lithosphere architecture; (2) better methodologies for determining the time scales of mineral
171
systems (geochronological methods)—and the geodynamic
triggers to ore-forming episodes that transect commodities
and deposit types; (3) a better understanding of terrane fertility, that is, the regions and time periods of Earth’s geologic
record that are more conducive for mobilizing large volumes
of metals in fluids; and (4) distinct methods for identifying
self-organizing critical systems, and for understanding how
fluid flow is organized in the crust to generate ore deposits. A
particularly fruitful avenue of research is on the use of detrital
minerals at an early stage of mineral exploration to identify
transient geodynamic events, lithosphere character, and terrane fertility.
From an exploration perspective, emphasis should be
placed on identification of multiscale footprints of mineral
systems that are nested in space and time, which can be
mapped in geoscience datasets at early stages of exploration
and in covered terranes where traditional deposit-scale footprints are not obvious with current detection technologies.
These mineral system footprints will have different expressions at differing scales, and therefore require the collation
of primary datasets that can map the critical elements of the
mineral system at the scales appropriate to the exploration
decision being considered.
Acknowledgments
The authors wish to acknowledge input and discussions
with colleagues over the years, including Graham Begg, Nick
Hayward, Steve Beresford, John Miller, Marco Fiorentini, the
past and current staff and students of the Centre for Exploration Targeting, the ARC Centre of Excellence for Core to
Crust Fluid Systems, CSIRO Minerals Down Under Flagship,
and the predictive mineral discovery Cooperative Research
Centre. In particular, the authors acknowledge the influence
of Roy Woodall, David Groves, and Rob Kerrich in the application of multiscale mineral-systems thinking to ore deposit
geology. David Leach and Neil Williams are thanked for helpful reviews of earlier versions of the manuscript, and formal
reviews from Karen Kelley, Dick Tosdal, and Noel White
greatly improved the manuscript. McCuaig acknowledges
receipt of ARC Linkage grant LP110100667. This is contribution 467 from the ARC Centre of Excellence for Core to
Crust Fluid Systems (http://www.ccfs.mq.edu.au).
REFERENCES
Agricola, G., 1556, De Re Metallica: Translation from the original in Latin
into English by Hoover, H.C., and Hoover, L.H., 1950: New York, Dover
Publications, 638 p.
Angerer, T., Duuring, P., Hagemann, S.G., Thorne, W., and McCuaig, T.C.,
2014, A mineral system approach to iron ore in Archaean and Palaeoproterozoic BIF of Western Australia: Geological Society, London, Special
Publication 393, DOI 10.1144/SP393.X.
Arribas, A., Jr., Hedenquist, J.W., Itaya, T., Okada, T., Concepcion, R.A., and
Garcia, J.S., Jr., 1995, Contemporaneous formation of adjacent porphyry
and epithermal Cu-Au deposits over 300 ka in northern Luzon, Philippines:
Geology, v. 23, p. 337–340.
Arndt, N., 2013, The lithospheric mantle plays no active role in the formation
of orthomagmatic ore deposits: Economic Geology, v. 108, p. 1953–1970.
Bak, P., 1996, How nature works: the science of self-organized criticality:
Copernicus, New York, Springer-Verlag, 212 p.
Bak, P., Tang, C., and Weisenfeld, K., 1987, Self-organized criticality: An
explanation of 1/f noise: Physical Review Letters, A38, p. 364–374.
Barley, M.E., and Groves, D.I., 1992, Supercontinent cycles and the distribution of metal deposits through time: Geology, v. 20, p. 291–294.
172
MCCUAIG AND HRONSKY
Barnes, S.J., 2006, Komatiite-hosted nickel sulfide deposits: Geology, geochemistry, and genesis: Society of Economic Geologists Special Publication
13, p. 51–97.
Barnes, S.J., and Lightfoot, P.C., 2005, Fformation of magmatic nickel sulfide
ore deposits and processes affecting their copper and platinum group element contents: Economic Geology 100th Anniversary Volume, p. 179–214.
Barnes, S.J., Hill, R.E.T., and Gole, M.J., 1988, The Perseverance ultramafic
complex, Western Australia: Product of a komatiite lava river: Journal of
Petrology, v. 29, p. 305–331.
Begg, G.C., Griffin, W.L., Natapov, L.M., O’Reilly, S.Y., Grand, S., O’Neill,
C.J., Hronsky, J.M.A., Poudjom Djomani, Y., Deen, T., and Bowden, P.,
2009, The lithospheric architecture of Africa: Seismic tomography, mantle
petrology and tectonic evolution: Geosphere, v. 5, p. 23–50.
Begg, G.C., Hronsky, J.M.A., Arndt, N.T., Griffin, W.L., OReilly, S.Y., and
Hayward, N., 2010, Lithospheric, cratonic and geodynamic setting of NiCu-PGE sulfide deposits: Economic Geology, v. 105, p. 1057–1070.
Bettles, K., 2002, Exploration and geology, 1962 to 2002, at the Goldstrike
Property, Carlin Trend, Nevada: Society of Economic Geologists Special
Publication 9, p. 275–298.
Bierlein, F.P., Northover, H.J., Groves, D.I., Goldfarb, R.J., and Marsh, E.E.,
2008, Controls on mineralization in the Sierra Foothills gold province, central California, USA: a GIS-based reconnaissance prospectivity analysis:
Australian Journal of Earth Sciences, v. 55, p. 61–78.
Billingsley, P., and Locke, A., 1935, Tectonic position of ore districts in the
Rocky Mountain region: American Institute of Mining and Metallurgical
Engineers Transactions, v. 115, p. 59–68.
——1941, Structure of ore districts in the continental framework: American
Institute of Mining and Metallurgical Engineers Transactions, v. 144, p. 9–64.
Bohlke, J.K., 1989, Comparison of metasomatic reactions between a common
CO2-rich vein fluid and diverse wall rocks: Intensive variables, mass transfers, and Au mineralization at Alleghany, California: Economic Geology, v.
84, p. 291–327.
Borg, G., Piestrzyński, A., Bachmann, G.H., Püttmann, W., Walther, S., and
Fiedler, M., 2012, An overview of the Kupferschiefer deposits: Society of
Economic Geologists Special Publication 16, p 455–486.
Bradley, D.C., and Leach, D.L., 2003, Tectonic controls of Mississippi Valleytype lead-zinc mineralization in orogenic forelands: Mineralium Deposita,
v. 38, p. 652–667.
Campbell, J.D., 1990, Hidden gold: the Central Norseman story: Australian
Institute of Mining and Metallurgy Monograph 16, 68 p.
Carlson, C.A., 1991, Spatial distribution of ore deposits: Geology, v. 19, p.
111–114.
Cathles, L.M., 1990, Scale and effects of fluid flow in the upper crust: Science, v. 248, p. 322–329.
Cathles, L.M., and Adams, J.J., 2005, Fluid flow and petroleum and mineral resources in the upper (<20-km) continental crust: Economic Geology
100th Anniversary Volume, p. 77–110.
Cathles, L.M., and Shannon, R., 2007, How potassium silicate alteration suggests the formation of porphyry ore deposits begins with the nearly explosive but barren expulsion of large volumes of magmatic water: Earth and
Planetary Science Letters, v. 262, p. 92–108.
Cathles, L.M., and Smith, A.T., 1983, Thermal constraints on the formation
of Mississippi Valley-type lead-zinc deposits and their implications for episodic basin dewatering and deposit genesis: Economic Geology, v. 78, p.
983–1002.
Cawood, P.A., and Hawkesworth, C.J., 2013, Temporal relations between mineral deposits and global tectonic cycles: Geological Society, London, Special
Publications, first published October 24, 2013; DOI 10.1144/SP393.1
Chen, H., Cooke, D.R., and Baker M., 2013, Mesozoic iron oxide coppergold mineralization in the Central Andes and the Gondwana Supercontinent breakup: Economic Geology, v. 108, p. 37–44.
Chernicoff, C.J., Richards, J.P., and Zappettini, E.O., 2002, Crustal lineament
control on magmatism and mineralization in northwestern Argentina: Geological, geophysical, and remote sensing evidence: Ore Geology Reviews,
v. 21, p. 127–155.
Chiodini, G., Cardellini, C., Amato, A., Boschi, E., Caliro, S., Frondini, F.,
and Ventura, G., 2004, Carbon dioxide Earth degassing and seismogenesis
in central and Southern Italy: Geophysical Research Letters, v. 31, L07615,
DOI 10.1029/2004GL019480.
Cline, J.S., Hofstra, A.H., Muntean, J.L., Tosdal, R.M., and Hickey, K.A.,
2005, Carlin-type gold deposits in Nevada: Critical geologic characteristics and viable models: Economic Geology 100th Anniversary Volume, p.
451–484.
Conolly, H.J.C., 1936, A contour method of revealing some ore structures:
Economic Geology, v. 31, p. 259–271.
Cox, S.F., 1999, Deformational controls on the dynamics of fluid flow in
mesothermal gold systems: Geological Society, London, Special Publication 155, p. 123–140.
——2005, Coupling between deformation, fluid pressures, and fluid flow in
ore-producing hydrothermal systems at depth in the crust: Economic Geology 100th Anniversary Volume, p. 39–75.
Crafford, A.E.J., and Grauch, V.J.S., 2002, Geologic and geophysical evidence
for the influence of deep crustal structures on Paleozoic tectonics and the
alignment of world-class gold deposits, north-central Nevada, USA: Ore
Geology Reviews, v. 21, p. 157–184.
Cuney, M., 2010, Evolution of uranium fractionation processes through time:
Driving the secular variation of uranium deposit types: Economic Geology,
v. 105, p. 553–569.
Czamanske, G.K., Zen’ko, T.E., Fedorenko, V.A., Calk, L.C., Budahn, J.R.,
Bullock, J.H., Fries, T.L., King, B.S., and Siems, D.F., 1995, Petrography
and geochemical characterization of ore-bearing intrusions of the Noril’sk
type, Siberia, with discussion of their origin: Resource Geology Special
Issue 18, p. 1–48.
Davies, G.F., 1995, Punctuated tectonic evolution of the Earth: Earth and
Planetary Science Letters, v. 136, p. 363–379.
Debaille, V., O’Neill, C., Brandon, A. D., Haenecour, P., Yin, Q. Z., Mattielli,
N., and Treiman, A.H., 2013, Stagnant-lid tectonics in early Earth revealed
by 142Nd variations in late Archean rocks: Earth and Planetary Science Letters, v. 373, p. 83–92.
Drummond, B., Lyons, P., Goleby, B., and Jones, L., 2006, Constraining models of the tectonic setting of the giant Olympic Dam iron oxide-copper-gold
deposit, South Australia, using deep seismic reflection data: Tectonophysics, v. 420, p. 91–103.
Ehrig, K., McPhie, J., and Kamenetsky, V., 2012, Geology and mineralogical
zoning of the Olympic Dam iron oxide Cu-U-Ag deposit, South Australia:
Society of Economic Geologists Special Publication 16, p. 237–267.
Emsbo, P., William Jones, A.E., Koenig, A.E., and Wilson, S.A., 2009, Petroleum as an agent of metal transport: Metallogenic and exploration implications: Smart Science for Exploration and Mining, Biennial SGA Meeting,
10th, Townsville, 2009, Proceedings, v. A1, p. 99–101.
Evans, K.A., McCuaig, T.C., Leach, D., Angerer, T., and Hagemann, S.G.,
2013, Banded iron formation to iron ore: A record of the evolution of Earth
environments: Geology, v. 41, p. 99–102.
Farquhar, J., Wu, N., Canfield, D.E., and Oduro, H., 2010, Connections
between sulfur cycle evolution, sulfur isotopes, sediments, and base metal
sulfide deposits: Economic Geology, v. 105, p. 509–533.
Folinsbee, R.E., 1977, World’s view—from Alph to Zipf: Bulletin of the Geological Society of America, v. 88, p. 897–907.
Frimmel, H.E., Groves, D.I., Kirk, J., Ruiz, J., Chesley, J., and Minter,
W.E.L., 2005, The formation and preservation of the Witwatersrand goldfields, the world’s largest gold province: Economic Geology 100th Anniversary Volume, p. 769–788.
Goldfarb, R.J., Snee, L.W., Miller, L.D., and Newberry, R.J., 1991, Rapid
dewatering of the crust deduced from ages of mesothermal gold deposits:
Nature, v. 354, p. 296–298.
Goldfarb, R.J., Baker, T., Dubé, D., Groves, D.I., Hart, C.J.R., and Gosselin,
P., 2005, distribution, character, and genesis of gold deposits in metamorphic terranes: Economic Geology 100th Anniversary Volume, p. 407–450.
Goldfarb, R.J., Bradley, D., and Leach, D.L., 2010, Secular variation in economic geology: Economic Geology, v. 105, p. 459–465.
Goldfarb, R.J., Taylor, R.D., Collins, G.S., Goryachev, N.A., and Orlandini,
O.F., 2014, Phanerozoic continental growth and gold metallogeny of Asia:
Gondwana Research, v. 25, p. 48–102.
Gresham, J.J., and Loftus-Hills, G.D., 1981, The geology of the Kambalda
nickel field, Western Australia: Economic Geology, v. 76, p. 1373–1416.
Griffin, W.L., Begg, G.C., and O’Reilly, S.Y., 2013, Continental-root control
on the genesis of magmatic ore deposits: Nature Geoscience, DOI 10.1038/
NGEO1954.
Groves, D.I., 1993, The crustal continuum model for late-Archaean lode-gold
deposits of the Yilgarn Block, Western Australia: Mineralium Deposita, v.
28, p 366–374.
Groves, D.I., Goldfarb, R.J., Gebre-Mariam, M., Hagemann, S.G., and Robert, F., 1998, Orogenic gold deposits: A proposed classification in the context of their crustal distribution and relationship to other gold deposit types:
Ore Geology Reviews, v. 13, p. 7–27.
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
Groves, D.I., Condie, K.C., Goldfard, R.J., Hronsky, J.M.A., and Vielreicher,
R.M., 2005, Secular changes in global tectonic processes and their influence on the temporal distribution of gold-bearing mineral deposits: Economic Geology, v. 100, p. 203–224.
Guj, P., Fallon, M., McCuaig, T.C., and Fagan, R., 2011, A time-series audit
of Zipf’s law as a measure of terrane endowment and maturity in mineral
exploration: Economic Geology, v. 106, p. 241–259.
Halley, S.W., and Roberts, R.H., 1997, Henty: A shallow-water gold-rich volcanogeneic massive sulfide deposit in western Tasmania: Economic Geology, v. 100, p. 203–224.
Harper, D.D., and Borrok, D.M., 2007, Dolomite fronts and associated zinclead mineralization, USA: Economic Geology, v. 102, p. 1345–1352.
Haynes, D., 2006, The Olympic Dam ore deposit discovery—a personal view:
Society of Economic Geologists Newsletter 66, p. 1, 8–15.
Hazen, R.M., 2010, Evolution of minerals: Scientific American, v. 302, p.
58–65.
Hazen, R.M., Papineau, D., Bleeker, W., Downs, R.T., Ferry, J.M., McCoy,
T.J., and Yang, H., 2008, Mineral evolution: American Mineralogist, v. 93,
p. 1693–1720.
Hedenquist, J.W., John, F.H., Thompson, J.F.H., Goldfarb, R.J., and Richards, J.P., 2005, eds., Economic Geology 100th Anniversary Volume, 1,136 p.
Hickey, K.A., Barker, S.L., Dipple, G.M., Arehart, G.B., and Donelick, R.A.,
2014, The brevity of hydrothermal fluid flow revealed by thermal halos
around giant gold deposits: Implications for Carlin-type gold systems: Economic Geology, v. 109, p. 1461–1487.
Hodgson, C.J., 1993, Mesothermal lode-gold deposits: Geological Association
of Canadian Special Paper 40, p. 635–678.
Hronsky, J.M.A., 2011, Self-organized critical systems and ore formation: The
key to spatial targeting?: Society of Economic Geology Newsletter 84, p.
14–16.
Hronsky, J.M.A., and Groves, D.I., 2008, Science of targeting: Definition,
strategies, targeting and performance measurement: Australian Journal of
Earth Science, v. 55, p. 3–12.
Hronsky, J.M.A., Groves, D.I., Loucks, R.R., and Begg, G.C., 2012, A unified
model for gold mineralization in accretionary orogens and implications for
regional scale targeting methods: Mineralium Deposita, v. 47, p. 339–358.
Idnurm, M., 2000, Towards a high resolution Late Palaeoproterozoic earliest
Mesoproterozoic apparent polar wander path for northern Australia: Australian Journal of Earth Science, v. 47, p. 405–429.
Ingebritsen, S.E, and Appold, M.S., 2012, The physical hydrogeology of ore
deposits: Economic Geology, v. 107, p. 559–584.
International Atomic Energy Agency, 2000, Methods of exploitation of different types of uranium deposits: Vienna, International Atomic Energy
Agency IAEA-TECDOC-1174, 75 p.
Jelsma, H., Barnett, W., Richards, S., and Lister, G., 2009, Tectonic setting of
kimberlites: Lithos, v. 112, p. 155–165.
Jensen, H.J., 1998, Self-organized criticality: Emergent complex behavior in
physical and biological systems: Cambridge Lecture Notes in Physics, Cambridge University Press, v. 10, 152 p.
Kelley, K.D., and Ludington, S., 2002, Cripple Creek and other alkalinerelated gold deposits in the southern Rocky Mountains, USA: Influence of
regional tectonics: Mineralium Deposita, v. 37, p. 38–60.
Kerrich, R., 1983, Geochemistry of gold deposits in the Abitibi greenstone
belt: Canadian Institute of Mining and Metallurgy Special Paper 27, 75 p.
Kerrich, R., Goldfarb, R.J., and Richards, J.P., 2005, Metallogenic provinces
in an evolving geodynamic framework: Economic Geology 100th Anniversary Volume, p. 1097–1136.
Kesler, S.E., and Wilkinson, B.H., 2006, The role of exhumation in the temporal distribution of ore deposits: Economic Geology, v. 101, p. 919–922.
Kish, L., and Cuney, M., 1981, Uraninite-albite veins from the Mistamisk
Valley of the Labrador trough, Québec: Mineralogical Magazine, v. 44, p.
471–483.
Knox-Robinson, C.M., and Wyborn, L.A.I., 1997, Towards a holistic exploration strategy: Using geographic information systems as a tool to enhance
exploration: Australian Journal of Earth Sciences, v. 44, p. 453–463.
Korzhinskii, D.S., 1968, The theory of metasomatic zoning: Mineralium
deposita, v. 3, p. 222–231.
Kouzmanov, K., and Pokrovski, G.S., 2012, Hydrothermal controls on metal
distribution in porphyry Cu (-Mo-Au) systems: Society of Economic Geologists Special Publication 16, p. 573–618.
Kreuzer, O.P., Etheridge, M.A., Guj, P., McMahon, M.E., and Holden, D.,
2008, Linking mineral deposit models to quantitative risk analysis and decision-making in exploration: Economic Geology, v. 103, p. 829–850.
173
Kreuzer, O.P., Markwitz, V., Porwal, A.K., and McCuaig, T.C., 2010, A continent-wide study of Australia’s uranium potential: Part I: GIS-assisted manual prospectivity analysis: Ore Geology Reviews, v. 38, p. 334–366.
Kump, L.R., 2008, The rise of atmospheric oxygen: Nature, v. 451, p. 277–278.
Lallemand, S., Heuret, A., and Boutelier, D., 2005, On the relationships
between slab dip, back-arc stress, upper plate absolute motion, and crustal
nature in subduction zones: Geochemistry Geophysics Geosystems, v. 6,
Q09006, DOI 10.1029/2005GC000917.
Large, R.R., McPhie, J., Gemmell, J.B., Herrmann, W., and Davidson, G.J.,
2001, The spectrum of ore deposit types, volcanic environments, alteration
halos, and related exploration vectors in submarine volcanic successions:
Some examples from Australia: Economic Geology, v. 96, p. 913–938.
Law, D.M., and Phillips, G.N., 2005, Hydrothermal replacement model
for Witwatersrand gold: Economic Geology 100th Anniversary Volume, p.
799–812.
Leach, D.L., Sangster, D.F., Kelley, K.D., Large, R.L., Garven, G.G., Allen,
C.,R., Gutzmer, J., and Walters, S., 2005, Sediment-hosted lead-zinc deposits: a global perspective: Economic Geology 100th Anniversary Volume, p.
561–608.
Leach, D.L., Bradley, D.C., Huston, D., Pisarevsky, S.A., Taylor, R.D., and
Gardoll, S. J., 2010, Sediment-hosted lead-zinc deposits in Earth history:
Economic Geology, v. 105, p. 593–625.
Lezaeta, P., 2001, Distortion analysis and 3-D modeling of magnetotelluric
data in the southern Central Andes: Ph.D. dissertation, Fachbereich Geowissenschaften, Freie Universität Berlin, 196 p.
Lord, D., Etheridge, M.A., Willson, M., Hall, G., and Uttley, P.J., 2001, Measuring exploration success: An alternative to the discovery-cost-per-ounce
method of quantifying exploration success: Society of Economic Geologists
Newsletter 45, p. 10–16.
Loucks, R.R., 2014, Distinctive composition of copper-ore-forming arcmagmas: Australian Journal of Earth Sciences, p. 1–13, DOI 10.1080/08120099.
2013.865676.
Loughlin, G.F., and Behre, C.H., Jr., 1933, Classification of ore deposits,
in Finch, J.W., Bateman, A.M., Butler, B.S., Loughlin, G.F., McLaughlin, D.H., and Singewald, J.T., Jr., eds., Ore deposits of the western states
(Lindgren volume): New York, American Institute Mining and Metallurgical Engineers, p. 17–55.
Love, D.A., Clark, A.H., and Glover, J.K., 2004, The lithologic, stratigraphic,
and structural setting of the giant Antamina copper-zinc skarn deposit,
Ancash, Peru: Economic Geology, v. 99, p. 887–916.
Lowell, J., and Guilbert. J.M., 1970, Lateral and vertical alteration mineralization zoning in porphyry ore deposits: Economic Geology, v. 65, p. 373–406.
Lu, Y.J., Kerrich, R., Kemp, A.I., McCuaig, T.C., Hou, Z.Q., Hart, C.J., and
Evans, N.J., 2013, Intracontinental Eocene-Oligocene porphyry Cu mineral systems of Yunnan, Western Yangtze craton, China: Compositional
characteristics, sources, and implications for continental collision metallogeny: Economic Geology, v. 108, p. 1541–1576.
Magoon, L.B., and Beaumont, E.A., 1991, Petroleum systems, in Beaumont,
E.A., and Foster, N.H., eds., Exploring for oil and gas traps: Treatise of
Petroleum Geology, Handbook of Petroleum Geology, p. 3-1–3-34.
Manning, C.E., and Ingebritsen, S.E., 1999, Permeability of the continental crust: The implications of geothermal data and metamorphic systems:
Reviews of Geophysics, v. 37, p. 127–150
Maydagán, L., Franchini, M., Chiaradia, M., Dilles, J., and Rey, R., 2014,
The Altar porphyry Cu-(Au-Mo) deposit (Argentina): A complex magmatichydrothermal system with evidence of recharge processes: Economic Geology, v. 109, p. 621–641.
McCuaig, T.C., and Hronsky, J.M.A., 2000, The current status and future
of the interface between the exploration industry and economic geology
research: Reviews in Economic Geology, v. 13, p. 553–559.
McCuaig, T.C., and Kerrich, R., 1998, P-T-t-deformation-fluid characteristics
of lode gold deposits: Evidence from alteration systematic: Ore Geology
Reviews, v. 12, p. 381–453.
McCuaig, T.C., Guibal, D., Roberts, M., and Cowan, J., 2003, Structural
review of the Antamina mine, Peru: SRK report (unpub.) for Compania
Minera Antamina, April, 2003, 37 p.
McCuaig, T.C., Porwal, A., and Gessner, K., 2009, Fooling ourselves: Recognizing uncertainty and bias in exploration targeting: Centre for Exploration
Targeting Newsletter, March 2009, p. 1, 5–8.
McCuaig, T.C., Beresford, S., and Hronsky, J., 2010, Translating the mineral
systems approach into an effective exploration targeting system: Ore Geology Reviews, v. 38, p. 128–138.
174
MCCUAIG AND HRONSKY
McInnes, B.I., Evans, N.J., Fu, F.Q., and Garwin, S., 2005, Application of
thermochronology to hydrothermal ore deposits: Reviews in Mineralogy
and geochemistry, v. 58(1), p. 467–498.
McKeith, T.D., Schodde, R.C., and Baltis, E.J., 2010, Gold discovery trends:
Society of Economic Geologists Newsletter 81, p. 1, 20–26.
McKinstry, H.E., 1941, Structural control of ore deposition in fissure veins:
American Institute of Mining Engineers Transactions, v. 144, p. 65–95.
——1955, Structure of hydrothermal ore deposits: Economic Geology, v. 50,
p. 170–225.
Mercadier, J., Annesley, I.R.J., McKechnie, C.L., Bogdan T.S., and Creighton, S., 2013, Magmatic and metamorphic uraninite mineralization in the
western margin of the Trans-Hudson orogen (Saskatchewan, Canada): A
uranium source for unconformity-related uranium deposits?: Economic
Geology, v. 108, p. 1037–1065.
Miller, J., Blewett, R., Tunjic, J., and Connors, K., 2010, The role of early
formed structures on the development of the world class St Ives goldfield,
Yilgarn, WA: Precambrian Research, v. 183, p. 292–315.
Miller, S.A., Colletini, C., Chiaraluce, L., Cocco, M., Barchi, M., and Klaus,
B.J.P., 2004, Aftershocks driven by a high-pressure CO2 source at depth:
Nature, v. 247, p. 724–727.
Meyer, C., 1981, Ore-forming processes in geologic history: Economic Geology, v. 75, p. 6–41.
——1985, Ore metals through geologic history: Science, v. 227, p. 1421–1428.
——1988, Ore deposits as guides to geologic history of the Earth: Annual
Reviews of Earth and Planetary Science, v. 16, p. 147–171.
Mole, D.R., Fiorentini, M.L., Cassidy, K.F., Kirkland, C.L., Thebaud, N.,
McCuaig, T.C., and Miller, J., 2013, Crustal evolution, intra-cratonic architecture and the metallogeny of an Archaean craton: Geological Society,
London, Special Publication 393, SP393-8; DOI 10.1144/SP393.8.
Mole, D.R., Fiorentini, M.L., Thebaud, N., Cassidy, K.F., McCuaig, T.C.,
Kirkland, C.L., Romano, S.S., Doublier, M.P., Belousova, E.A., Barnes, S.J.,
and Miller, J., 2014, Archean komatiite volcanism controlled by the evolution of early continents: Proceedings of the National Academy of Sciences
(in press).
Muntean, J.L., Cline, J.S., Simon, A.C., and Longo, A.A., 2011, Magmatic
hydrothermal origin of Nevada’s Carlin-type gold deposits: Nature Geoscience, v. 4, p. 122–127.
Murphy, F.C., Cull, J., Lee, T.J., Lee, S.K., and Song, Y., 2008, Magnetotelluric soundings and crustal architecture at Century mine, northern Australia:
Australian Journal of Earth Sciences, v. 55, p. 5, 655–667.
Murphy, F.C., Hutton, L.J., Walshe, J.L., Cleverley, J.S., Kendrick, M.A.,
McLellan, J., Rubenach, J., Oliver, N.H.S., Gessner, K., Bierlein, F.P.,
Jupp, B., Ailleres, L., Laukamp, C., Roy, I.G., Miller, J.McL., Keys, D., and
Nortje, G.S., 2011, Mineral system analysis of the Mt Isa-McArthur River
region, northern Australia: Australian Journal of Earth Sciences, v. 58, p.
849–873.
Naldrett, A.J., 2010, Secular variation of magmatic sulfide deposits and their
source magmas: Economic Geology, v. 105, p. 669–688.
O’Connor, J.M., Steinberger, B., Regelous, M., Koppers, A.A., Wijbrans, J.R.,
Haase, K.M., Stoffers, P., Jokat, W., and Garbe-Schönberg, D., 2013, Constraints on past plate and mantle motion from new ages for the HawaiianEmperor Seamount Chain: Geochemistry, Geophysics, Geosystems, v. 14,
p. 4564–4584.
O’Driscoll, E.S.T., 1986, Observations of the lineament-ore relation: London,
Philisophical Transactions of the Royal Society, A317, p. 195–218.
O’Neill, C., Debaille, V., and Griffin, W., 2013a, Deep earth recycling in the
Hadean and constraints on surface tectonics: American Journal of Science,
v. 313, p. 912–932.
O’Neill, C., Lenardic, A., and Condie, K.C., 2013b, Earth’s punctuated tectonic evolution: Cause and effect: Geological Society, London, Special Publication, DOI 10.1144/SP389.4.
Padilla-Garza, R.A., Titley, S.R., and Pimental, F.B., 2001, Geology of the
Escondida porphyry copper deposit, Antofagasta region, Chile: Economic
Geology, v. 96, p. 307–324.
Papineau, D., 2010, Global biogeochemical changes at both ends of the Proterozoic: Insights from phosphorates: Astrobiology, v. 10, p. 165–181.
Peacock, D.C.P., 2003, Scaling of transfer zones in the British Isles: Journal of
Structural Geology, v. 25, p. 1561–1567.
Pettke, T., Oberli, F., and Heinrich, C.A., 2010, The magma and metal source
of giant porphyry-type ore deposits, based on lead isotope microanalysis
of individual fluid inclusions: Earth and Planetary Science Letters, v. 296,
p.267–277.
Pirajno, F., 2004, Metallogeny in the Capricorn orogen, Western Australia,
the result of multiple ore-forming processes: Precambrian Research, v. 128,
p. 411–439.
Presnell, R.D., 1992, Local and regional geology of the Oquirrh Mountains:
Utah Geological Survey Miscellaneous Publications, v. 3, p. 293–306.
Price, G.P., and Stoker, P., 2002, Australian Geodynamics Cooperative
Research Centre’s integrated research program delivers a new minerals
exploration strategy for industry: Australian Journal of Earth Sciences, v.
49, p. 595–600.
Rampino, M.R., and Stothers, R.B., 1988, Flood basalt volcanism during the
past 250 million years: Science, v. 241, p. 663–668.
Reddy, S.M., and Evans, D.A.D., 2009, Palaeoproterozoic supercontinents
and global evolution: Correlations from core to atmosphere: Geological
Society, London, Special Publication 323, p. 1–26.
Repetski, J.E., and Narkiewicz, M., 1996, Conodont color and surface textural alteration in the Muschelkalk (Triassic) of the Silesian-Cracow Zn-Pb
district, Poland: Prace Państwowego Instytutu Geologicznego, v. 154, p.
113–121.
Richard, A., Banks, D.A., Mercadier, J., Boiron, M.C., Cuney, M., and Cathelineau, M., 2011, An evaporated seawater origin for the ore-forming brines
in unconformity-related uranium deposits (Athabasca Basin, Canada): Cl/
Br and δ37Cl analysis of fluid inclusions: Geochimica et Cosmochimica
Acta, v. 75, p. 2792–2810.
Ridley, J., 1993, The relations between mean rock stress and fluid flow in
the crust: With reference to vein-and lode-style gold deposits: Ore Geology
Reviews, v. 8, p. 23–37.
Robert, F., Poulsen, K.H., Cassidy, K.F., and Hodgson, C.J., 2005, Gold
metallogeny of the Superior and Yilgarn craton: Economic Geology 100th
Anniversary Volume, p. 1001–1033.
Rohrlach, B.D., 2002, Tectonic evolution, petrochemistry, geochronology and
palaeohydrology of the Tampakan porphyry and high sulphidation epithermal Cu-Au deposit, Mindanao, Philippines: Unpublished Ph.D. thesis,
Canberra, Australia, Australian National University, 499 p.
Rohrlach, B.D., and Loucks, R.R., 2005, Multi-million year cyclic ramp-up of
volatiles in a lower crustal magma chamber trapped below the Tampakan
copper-gold deposit by Mio-Pliocene crustal compression in the southern
Philippines, in Porter T.M., ed., Super porphyry copper and gold deposits:
A global perspective: Adelaide, PGC Publishing, v. 2, p. 270–313.
Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., and Duboz,
C., 2005, Subduction of the Nazca Ridge and the Inca Plateau: Insights into
the formation of ore deposits in Peru: Earth and Planetary Science Letters,
v. 239, p. 18–32.
Saunders, J.A., Unger, D.L., Kamenov, G.D., Fayek, M., Hames, W.E., and
Utterback, W.C., 2008, Genesis of middle Miocene Yellowstone hotspotrelated bonanza epithermal Au-Ag deposits, northern Great Basin, USA:
Mineralium Deposita, v. 43, p.715–734.
Schneider, E.D., and Dorion, S., 2005, Into the cool: Energy flow, thermodynamics and life: Chicago, IL, University of Chicago Press, 362 p.
Schodde, R.C., and Hronsky, J.M.A., 2006, The role of world-class mines in
wealth creation: Society of Economic Geologists Special Publication 12, p.
71–90.
Seedorff, E., Dilles, J.H., Proffett, J.M., Jr., Einauidi, M.T., Zurcher, L.,
Stavast, W.J.A., Johnson, D.A., and Barton, M.D., 2005, Porphyry deposits:
Characteristics and origin of hypogene features: Economic Geology 100th
Anniversary Volume, p. 251–298.
Sherlock, R.L., Roth, T., Spooner, E.T.C., and Bray, C.J., 1999, Origin of
the Eskay Creek precious metal-rich volcanogenic massive sulfide deposit:
Fluid inclusion and stable isotope evidence: Economic Geology, v. 94, p.
803–824.
Sibson, R.H., Robert, F., and Poulsen, K.H., 1988, High-angle reverse faults,
fluid-pressure cycling and mesothermal gold deposits: Geology v. 14, p.
551–555.
Sillitoe, R.H., 2000, Gold-rich porphyry deposits: Descriptive and genetic
models and their role in exploration and discovery: Society of Economic
Geologists Reviews in Economic Geology, v. 13, p. 315–345.
——2004, Musings on future exploration targets and strategies in the Andes:
Society of Economic Geologists Special Publication 11, p. 1–14.
——2008, Major gold deposits and belts of the North and South American
Cordillera: Distribution, tectonomagmatic settings, and metallogenic considerations: Economic Geology, v. 103, p. 663–688.
——2010, Porphyry copper systems: Economic Geology, v. 105, p. 3–41.
——2012, Copper provinces: Society of Economic Geologists Special Publication 16, p. 1–18.
THE MINERAL SYSTEM CONCEPT: THE KEY TO EXPLORATION TARGETING
Sillitoe, R.H., and Mortensen, J.K., 2010, Longevity of porphyry copper formation at Quellaveco, Peru: Economic Geology, v. 105, p. 1157–1162.
Sillitoe, R.H., and Perelló, J., 2005, Andean copper province: Tectonomagmatic settings, deposits types, metallogeny, exploration and discovery: Economic Geology 100th Anniversary Volume, p. 845–890.
Sillitoe, R.H., and Thompson, J.F.H., 2006, Changes in mineral exploration
practice: Consequences for discovery: Society of Economic Geologists Special Publication 12, p. 193–219.
Simmons, S.F., and Brown, K.L., 2008, Precious metals in modern hydrothermal solutions and implications for the formation of epithermal ore deposits:
Society of Economic Geologists Newsletter 72, p. 1, 9–12.
Simmons, S.F., White, N.C., and John, D.A., 2005, Geological characteristics
of epithermal precious and base metal deposits: Economic Geology 100th
Anniversary Volume, p. 485–522.
Skelton, A., 2011, Flux rates for water and carbon during greenschist facies
metamorphism: Geology, v. 39, p. 43–46.
Snyder, D.B., 2013, Imaging Archaean-age whole mineral systems: Precambrian Research, v. 229, p. 125–132.
Squire, R.J., and Miller, J.M.M., 2003, Synchronous compression and extension in East Gondwana: Tectonic controls on world-class gold deposits at
440 Ma: Geology, v. 31, p. 1073–1076.
Swager, C.P., 1985, Syndeformational carbonate-replacement model for the
copper mineralization at Mount Isa, Northwest Queensland: A microstructural study: Economic Geology, v. 80, p. 107–125.
Syracuse, E.M., and Abers, G.A., 2006, Global compilation of variations in
slab depth beneath arc volcanoes and implications: Geochemistry, Geophysics, Geosystems, v. 7, DOI 10.1029/2005GC001045.
Taylor, C.D., Premo, W.R., Meier, A.L., and Taggart, J.E., 2008, The metallogeny of Late Triassic rifting of the Alexander terrane in southeastern
Alaska and northwestern British Columbia: Economic Geology, v. 103, p.
89–115.
175
Titley, S.R., 2001, Crustal affinities of metallogenesis in the American Southwest: Economic Geology, v. 96, p. 1323–1342.
Tomkins, A.G., 2013, A biogeochemical influence on the secular distribution
of orogenic gold: Economic Geology, v. 108, p. 193–197.
Tosdal, R.M., Dilles, J.H., and Cooke, D.R., 2009, From source to sinks in
auriferous magmatic-hydrothermal porphyry and epithermal deposits: Elements, v. 5, p. 289–295.
Turneaure, F.S., 1955, Metallogenetic provinces and epochs: Economic
Geology, v. 50, p. 38.
Witt, W.K., Hagemann, S.G., Villanes, C., and Zeng, Q., 2013, New geochronological results and structural evolution of the Pataz gold mining district:
Implications for the timing and origin of the batholith-hosted veins: Ore
Geology Reviews, v. 50, p. 143–170.
Woodall, R., 1984, Success in mineral exploration: Confidence in science and
ore deposit models: Geoscience Canada, v. 11, p. 127–133.
——1994, Empiricism and concept in successful mineral exploration: Australian Journal of Earth Sciences, v. 41, p. 1–10.
Wyborn, L.A.I., Heinrich C.A., and Jaques A.L., 1994, Australian Proterozoic
mineral systems: Essential ingredients and mappable criteria: Australian
Institute of Mining and Metallurgy Annual Conference, Melbourne, Proceedings, p. 109–115.
Yamada, R., and Yoshida, T., 2011, Relationships between Kuroko volcanogenic massive sulfide (VMS) deposits, felsic volcanism, and island arc
development in the northeast Honshu arc, Japan: Mineralium Deposita, v.
46, p. 431–448.
Yardley, B.W.D., 2005, Metal concentrations in crustal fluids and their relationship to ore formation: Economic Geology, v. 100, p. 613–632.
Zhang, M., O’Reilly, S.Y., Wang, K-L., Hronsky, J., and Griffin, W.L., 2008,
Flood basalts and metallogeny: The lithospheric mantle connection: Earth
Science Reviews, v. 86, p. 145–174.
Descargar