Sie sind auf Seite 1von 22

2005 Society of Economic Geologists, Inc.

Economic Geology, v. 100, pp. 203224

100th Anniversary Special Paper:


Secular Changes in Global Tectonic Processes and Their Influence on the
Temporal Distribution of Gold-Bearing Mineral Deposits
DAVID I. GROVES,
Centre for Global Metallogeny, School of Earth and Geographical Sciences, University of Western Australia,
Crawley, Western Australia 6009, Australia

KENT C. CONDIE,
Department of Earth and Environmental Science, New Mexico Institute of Mining and Technology, Soccorro, New Mexico 87801

RICHARD J. GOLDFARB,
U.S. Geological Survey, Box 25046, Mail Stop 964, Denver Federal Center, Denver, Colorado 80225-0046

JONATHAN M.A. HRONSKY,


WMC Resources Ltd., 191 Great Eastern Highway, Belmont, Western Australia 6104, Australia
AND

RICHARD M. VIELREICHER

Centre for Global Metallogeny, School of Earth and Geographical Sciences, University of Western Australia,
Crawley, Western Australia 6009, Australia

Abstract
Mineral deposit types commonly have a distinctive temporal distribution with peaks at specific periods of
Earth history. Deposits of less redox-sensitive metals, such as gold, show long-term temporal patterns that
relate to first-order changes in an evolving Earth, as a result of progressively declining heat production and
attendant changes in global tectonic processes. Despite abundant evidence for plate tectonics in the early Precambrian, it is evident that plume events were more abundant in a hotter Earth.
Episodic growth of juvenile continental crust appears to have been related to short-lived (<100 m.y.) catastrophic mantle-plume events and formation of supercontinents, whereas shielding mantle-plume events correlated with their breakup. Different mineral deposit types are associated with this cycle of supercontinent formation and breakup. Broadly synchronous with juvenile continental crust formation was the development of
subcontinental lithospheric mantle, which evolved due to progressively declining heat flow and decreasing
plume activity. Archean subcontinental lithospheric mantle has a distinct mineralogical composition and is
buoyant, whereas later lithosphere was progressively more dense. Changes in the buoyancy of both oceanic
lithosphere and subcontinental lithospheric mantle led to evolution of tectonic scenarios in which buoyant,
roughly equidimensional, early Precambrian cratons were rimmed by Proterozoic or Phanerozoic linear elongate belts of neutral to negative buoyancy.
Orogenic gold deposits, which formed over at least 3.4 b.y., had the highest preservation potential of any gold
deposit type. The pattern of formation and preservation, from episodic to more cyclic, broadly mirrors that of
crustal growth. Early Precambrian (mostly ca. 2.7 and 2.01.8 Ga) deposits, protected from uplift and erosion
in the centers of buoyant cratons, are rare between ca. 1.7 Ga and 600 Ma due to the change to more modernstyle plate tectonic processes, with nonpreservation of deposits of this age due to uplift and erosion of more
vulnerable orogenic belts. Volcanic-hosted massive sulfide (VHMS) deposits were accreted into the convergent
margin terranes in which orogenic gold deposits were forming. Their temporal distribution, from strongly
episodic to more cyclic peaks, also supports a model of selective preservation.
The first appearance of iron-oxide copper-gold (IOCG) deposits at ~2.55 Ga closely follows development of
early Precambrian subcontinental lithosphere mantle. Their genesis involved melting of metasomatized subcontinental lithosphere mantle, so they could not form until such metasomatized mantle evolved below cratons with
buoyant lithosphere. Giant Precambrian paleoplacer gold deposits probably formed by effective fluvial sorting
under extreme climatic conditions but were largely preserved due to early buoyant subcontinental lithospheric
mantle below hosting foreland basins. Unequivocal intrusion-related gold deposits are related to complex felsic
intrusions with a mixed mantle-crustal signature, which intruded deformed shelf sedimentary sequences close to
but outside craton margins. Given that post-Paleoproterozoic uplift and erosion is likely in vulnerable orogenic
belts with negatively buoyant lithosphere, this deposit type is likely to be rare in Paleozoic and older terranes.
Gold-bearing deposit types thus display distinctive temporal distributions related to change from a more
buoyant plate tectonic style in the early hotter Earth to a modern plate tectonic style typical of the
Corresponding

author: e-mail, dgroves@segs.uwa.edu.au

0361-0128/01/3493/203-22 $6.00

203

204

GROVES ET AL.

Phanerozoic. Late Archean formation of buoyant subcontinental lithospheric mantle was particularly important in the anomalous preservation of some earlier formed deposit types located inboard of craton margins and
in providing critical conditions for the formation of others. Development of negatively buoyant subcontinental
lithospheric mantle can explain the lack of preservation of some deposit types that formed in the later Proterozoic. A single fundamental concept of coupled episodic crustal growth and preservation in the Archean and
Paleoproterozoic, evolving to decoupled episodes of growth and preservation from the Mesoproterozoic onward, can thus explain the temporal distribution of a number of gold-bearing mineral deposit types.

Introduction
THE ADVENT of the theory of plate tectonics generated considerable interest in the relationship between mineral deposits and their global tectonic setting (e.g., Sawkins, 1972,
1990; Sillitoe, 1972; Mitchell and Garson, 1981; Brimhall,
1987; Hutchinson, 1993; Titley, 1993; Kesler, 1997; Barley et
al., 1998; Kerrich et al., 2000; Blundell, 2002). This has led to
the realization that different mineral deposit types are related
to specific convergent, divergent, or anorogenic settings and
that the recognition of these settings is useful in exploration.
At the same time, there has been a growing recognition that
the various mineral deposit types have a heterogeneous temporal distribution (Fig. 1), with characteristic peaks in their
abundance at specific times in Earth history (e.g., Meyer,
1981, 1988; Barley and Groves, 1992; Goldfarb et al., 2001a).
The uneven temporal distribution can be explained by
changes in the processes that combine to produce the mineral
deposits and/or the preservation potential of their depositional environments. In turn, temporal changes in the process

of mineral deposit formation can be ascribed to (1) the evolution of atmosphere-hydrosphere-biosphere systems (e.g.,
Holland, 1984), (2) the widely accepted secular decrease in
global heat flow (e.g., as indicated by the virtual restriction of
komatiite-associated nickel-copper-sulfide deposits to the
early Precambrian; Lesher, 1989), and (3) long-term changes
in tectonic processes (e.g., Windley, 1984; Meyer, 1988;
Veizer et al., 1989; Barley and Groves, 1992). As discussed
below, secular decrease in global heat flow directly affects irreversible changes in tectonic processes, and both affect the
long-term preservation potential of terranes of different age
and, hence, represent important first-order controls on the
temporal distribution of mineral deposit types. Whether or
not the temporal distribution of near-surface mineral deposits, particularly sediment-hosted deposits, for which metal
transport and deposition are highly affected by redox conditions, is controlled by an evolving atmosphere-hydrospherebiosphere system is not discussed here. There is considerable
current controversy on whether there was a sudden appearance of an oxygenated world (e.g., Lasaga and Ohmoto, 2002)

A)

B)
3Ga

2Ga

1Ga

Ore type

Cyprus-type

Uranium in
weathered profile

Abitibi-type

Kiruna-type

IOCG

VHMS

Ore type

3Ga

2Ga

1Ga

Olympic Dam-type

Kuroko

Porphyry

Ilmenite-anorthosite
Lead-zinc in
clastic sediments

Paleoplacer
and placer gold

Lead-zinc in carbonates: Mississippi


Valley-type

SEDEX

Orogenic gold

Porphyry Cu

Copper in clastic
sediments

Porphyry Mo
FIG. 1. Distribution through time of the number of preserved specific mineral deposits ascribed to (A) orogenic-convergent margin settings and (B) anorogenic or continental-basin settings. Peaks in abundance of anorogenic and continentalbasin metal deposits appear to correspond to (1) possible breakup, or incipient breakup, of a Paleoproterozoic supercontinent, (2) formation of Rodinia at about 1 Ga, and (3) formation and breakup of Gondwana and Pangea. Adapted from Barley
and Groves (1992). Peaks for deposit types in (A) are better defined and discussed in the text. Note that the temporal pattern for orogenic gold deposits, in particular (see also Fig. 8), have evolved as better dating techniques have become employed. IOCG = iron-oxide copper-gold deposits, SEDEX = sedimentary-exhalative deposits, VHMS = volcanic-hosted massive sulfide deposits.

0361-0128/98/000/000-00 $6.00

204

205

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

Juvenile Continental Crust and the Supercontinent Cycle


In order to understand the temporal evolution of mineral
deposits that are preserved within continental crust, it is important to understand the temporal evolution of that crust
and its relationship to the configuration of continental masses
through timethe supercontinent cycle. The evolution of the
Earth is intrinsically linked to a gradual decay of heat production, lowering of mantle temperatures and viscosity, and
thickening of the subcontinental lithosphere (e.g., Pollack,
1986). Some authors consider this to have been more or less
a continuous evolution (e.g., Pollack, 1997), but the episodic
distribution of radiogenic isotopic ages, first recognized by
Gastil (1960), has led to models of more episodic growth of
juvenile continental crust (Taylor and McLennan, 1985; Stein
and Hoffmann, 1994; Condie, 1998, 2000). The distribution
of U-Pb zircon ages, coupled with Nd and Hf isotope data,
suggest that significant crustal growth commenced at or before 3.0 Ga with two major peaks in juvenile crustal production rate, one at ca. 2.7 Ga and another at ca. 1.9 Ga (Fig. 2).
The global dominance of greenstone belts and granitoids that
formed in the interval from ca. 2.75 to 2.60 Ga, in particular,
cannot be a sampling artifact, as such terranes are known
from all continents (e.g., Yilgarn craton of Australia, southern
Superior and Slave provinces of Canada, Zimbabwe craton,
Tanzania craton, and Sao Francisco craton of Brazil), and robust geochronologic techniques have been employed increasingly in these terranes in the past decade. In addition, although it is evident that the precise timing of peaks in crustal
growth is progressively evolving as more data become available (compare Condie, 1998, with Condie, 2001a), with a
probable uncertainty of 50 Ma, the relative patterns are robust. Thus, in addition to the two major peaks in crustal production rate, smaller peaks may be present at ca. 2.8, 2.5, 2.1,
0361-0128/98/000/000-00 $6.00

14

Volume percent growth

or more gradual evolution to a present-day atmosphere (e.g.,


Farquhar et al., 2000) and discussion of these issues is beyond
the scope of this paper.
This paper examines current evidence for the evolution of
the continental crust, the supercontinent cycle, the nature of
mantle plumes and their influence on the supercontinent
cycle, and the transition from Archean- to modern-style plate
tectonics in order to develop a model that explains secular
variations in metallogeny in terms of changes in tectonic
processes and their preservational consequences. In particular, these consequences are examined with respect to the
global metallogeny of gold deposits. The paper deals with
global secular patterns of mineral deposits in terms of
changes in tectonic processes but does not deal in any detail
with the environmental consequences (e.g., Titley, 1993;
Kesler, 1997) of those tectonic processes. The best possible
temporal patterns for >90 percent of the resources of each
deposit type are presented within the constraints on compatibility of resource data and uncertainties of precise timing for
some deposits. The paper attempts to integrate data and concepts across several disciplines and focuses only on the most
recent overviews. The interested reader is referred to these
reviews for exhaustive lists of references within the different
disciplines. The review by Kerrich et al. (2000) is a particularly complete source of references that relate to the geodynamics of world-class gold deposits.

12

2.7 Ga

10

1.9 Ga

8
6
4
2

3.8 3.4

3.0 2.6 2.2 1.8 1.4 1.0 0.6 0.2

Age (Ga )
FIG. 2. Frequency distribution of juvenile continental crust based on a
total volume of continental crust of 7.177 109 km3. Juvenile crust ages are
U/Pb zircon ages used in conjunction with Nd isotope data and lithologic associations. Modified after Condie (1998, 2000).

1.7, 0.48, 0.28, and 0.1 Ga (Condie, 2001b). As discussed


below, the pattern of these peaks and the peaks themselves
correspond broadly with those of orogenic gold provinces and
hence are an important control on global metallogeny. It is
still not clear if, and how, these peaks in the rate of crustal
production relate to the supercontinent cycle, as defined, for
example, by Hoffman (1988) and Murphy and Nance (1992),
and to mantle overturn events (e.g., Condie, 1998). They may,
or may not, correlate with the accumulation or breakup phase
of supercontinents. In the last 1 b.y., the formation and
breakup of three such major supercontinents (Rodinia,
Gondwana, and Pangea), and a possible short-lived supercontinent (Pannotia) in the latest Proterozoic, have been recognized (Unrug, 1997; Condie, 2002a). Geologic data support
the existence of at least two earlier supercontinents, one at
the end of the Archean and one in the Paleoproterozoic
(Hoffman, 1989; Rogers, 1996; Aspler and Chiarenzelli, 1998;
Pesonen et al., 2003).
Ages from Archean cratons suggest that the first supercontinent (or supercontinents; Aspler and Chiarenzelli, 1998)
formed during frequent collisions and suturing of older continental blocks and juvenile oceanic terranes (principally arcs
and oceanic plateaus) between 2750 and 2650 Ma (Fig. 2).
The Late Archean peak in the rate of juvenile crust production is also centered at 2700 50 Ma (Fig. 2), further suggesting a correlation between supercontinent formation and
juvenile continental crust production. Major gold provinces
developed in the southern Superior, Slave, Yilgarn, Zimbabwe, Tanzania, and Sao Francisco cratons at about this time
(as summarized by Goldfarb et al., 2001a).
The final breakup of the Late Archean supercontinent(s)
occurred in the Paleoproterozoic between 2200 and 2100 Ma.
Subsequently, a new supercontinent formed between 1900
and 1800 Ma, with most collisions in Laurentia, Baltica, and
Siberia occurring ca. 1850 Ma (Condie, 2002b; Pesonen et al.,
2003; Fig. 2). The giant Homestake gold deposit formed at
around this time (Dahl et al., 1999).

205

206

GROVES ET AL.

In the Phanerozoic, crustal growth peaks are indicated at


ca. 480, 280, and 100 Ma (Condie, 2001b). The first two peaks
broadly correlate with the rapid growth of Pangea between
480 and 250 Ma and the third with the growth of a new supercontinent beginning at ca. 150 Ma. Importantly, these
peaks broadly coincide with major gold events in the Tasman
orogen, Central Asia, and circum-Pacific (Goldfarb et al.,
2001a, b). Increased rates of production of juvenile crust appear to correlate with formation and not breakup phases of
the supercontinents over the last 1,200 m.y. (Condie, 2004).
Mantle-Plume Events and the Supercontinent Cycle
The following section discusses the relationship between
crustal growth and mantle-plume events and how the latter
impact on the supercontinent cycle, which, in turn, impacts
on the temporal distribution of mineral deposits.
Major peaks in the rate of production of juvenile crust are
interpreted to be caused by mantle overturn events involving
numerous mantle plumes in a short time interval (Condie,
1998, 2000). Mantle plume events are short-lived episodes
(100 m.y.) during which many plumes impact on or affect
the base of the lithosphere, and they appear to have been important throughout Earth history (Condie, 1998; Isley and
Abbott, 1999). During such events, plume activity may be
concentrated in one or more mantle upwellings, as during the
mid-Cretaceous (ca. 100 Ma) superplume event focused in
the Pacific basin (Larson, 1991).
Mantle plume events can be identified by the recognition
of igneous rocks associated with mantle plumes, commonly
called plume proxies (Ernst and Buchan, 2002; Isley and Abbott, 2002). Using a combination of plume proxies, including
flood basalts, komatiites, and high MgO lavas, giant dike
swarms, and layered intrusions, Abbott and Isley (2002) recognized 36 mantle-plume events in the last 3.8 b.y. (Fig. 3). A
weighted time series analysis shows that most mantle-plume
events, regardless of age, lasted about 10 m.y., with major
Precambrian events at ca. 2.75, 2.45, 1.8, 1.75, and 1.65 Ga,
and several events in the Phanerozoic. Peaks in the abundance of plume proxies at ca. 2.75, 1.8, 0.25, and 0.12 Ga are
close to calculated peaks in the production of juvenile crust,
although not all such events show such correlations (Condie,
2004).
There may be two types of mantle-plume events; one associated with supercontinent formation and one with supercontinent breakup (Condie, 2004). It is commonly believed that
1.5

Height

1.2
0.9
0.6
0.3
4.0

3.5

3.0

2.5
2.0
Age (Ga)

1.5

1.0

0.5

FIG. 3. Distribution of mantle plume events deduced from time series


analysis of plume proxies (i.e., igneous rocks associated with mantle plumes)
from Abbott and Isley (2002). Peak height depends on the number of plume
proxies and the errors of the age, the latter set at 5 m.y.
0361-0128/98/000/000-00 $6.00

plumes associated with mantle upwellings are responsible for


fragmenting supercontinents (Condie, 2001b). Computer
models suggest that it takes 200 to 400 m.y. for shielding of a
large supercontinent to cause a mantle upwelling beneath it
(Lowman and Jarvis, 1996). This is followed by development
of numerous mantle plumes within the upwelling and, finally,
by supercontinent breakup over a period of up to approximately 200 m.y. (Fig. 4). If, in a mantle-plume event (Condie,
2004), the supercontinent is too small to provide sufficient
mantle shielding to produce an upwelling, the small supercontinent may move laterally with the downwelling (e.g., Trubitsyn et al., 2003) or may not completely fragment, as appears to be the case with the 1.9-Ga supercontinent (Condie,
2002c). Thus, there may be long-term survival of the roots of
relatively small cratonic blocks such as the Yilgarn and Pilbara
cratons of Western Australia (Trubitsyn et al., 2003). Although plume heads can have diameters of up to 2,500 km in
such a shielding mantle-plume event, the volume of juvenile
mafic crust associated with a given plume (e.g., estimated by
the volume of Phanerozoic flood basalts and mafic underplates) is probably relatively small, as is the volume of oceanic
plateaus accreted to the continents, at least since the end of
the Archean (Condie, 2001b).
There must be, therefore, something unique about the
mantle-plume events associated with peaks in the production
of juvenile crust at about 2.7 and 1.9 Ga. These events, called
catastrophic mantle-plume events (Condie, 2004), must be
triggered by some process other than plate shielding. They
are also short lived (less than 100 m.y.) in contrast to shielding events (>200 m.y.) and must be more intense and perhaps
more widespread. The cause of the catastrophic mantle
plumes is unclear. Peltier et al. (1997) and Condie (1998) suggested that the breakup of Precambrian supercontinents triggered slab avalanches at the 660-km discontinuity in the mantle, resulting in catastrophic plume production in a (D)
layer just above the core. However, this model does not appear to work for the 2.7-Ga event, since there is no evidence
for significant earlier supercontinent fragmentation.
The most voluminous banded iron formations (BIF) were
deposited in intracratonic, passive-margin or platform basins
during stands of high sea level in the Late Archean and Paleoproterozoic (Simonson and Hassler, 1997). The iron and silica enrichments in these rocks appear to have been derived
chiefly from hydrothermal vents on the deep sea floor (Klein
and Beukes, 1992; Isley, 1995; Barley et al., 1999). Two major
peaks in BIF deposition at 2.7 and 1.9 Ga correlate well with
and can be explained by mantle-plume events at these times
(Klein and Beukes, 1992; Isley and Abbott, 1999). The enhanced submarine volcanism and hydrothermal venting associated both with ocean-ridge and oceanic plateau volcanism
during a mantle-plume event may be the source of iron and
silica. Also, the elevated sea level at 1.9 Ga caused by a plume
event could provide the extensive shallow marine basins along
stable continental platforms necessary to preserve BIF (cf. Titley, 1993).
Crustal growth associated with catastrophic mantle-plume
events is chiefly by addition of magmatic arcs directly within
continental margins or by accretion of oceanic arcs to the
edge of the continents (Rudnick, 1995; Condie, 2001b), although at least in the last 2.5 b.y., the latter played a relatively

206

207

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

Super continents

Superplume Events

Breakup

Juvenile
Crust
Formation

G P

R
3

Age (Ga)

FIG. 4. Formation and breakup of supercontinents during the past 3.0 b.y, based on available paleogeographic
reconstructions. Also shown are times of maximum production rate of juvenile continental crust and proposed catastrophic
mantle-plume events. Data from Condie (1998, 2001b, 2002a-c). G = Gondwana, N = new supercontinent, P = Pangea, R =
Rodinia.

The Transition from Archean to


Modern-Style Plate Tectonics
If the Earth has steadily cooled with time, as is widely accepted (e.g., Pollack, 1986), there is good reason to suspect
that tectonic processes were different in the Archean (e.g.,
Fyfe, 1978). For example, a hotter mantle in the Archean
would have produced more melt at ocean ridges and hence
thicker oceanic crust (e.g., Bickle, 1990). Model calculations
indicate that the Archean oceanic crust should have been
about 20 km thick, and perhaps even thicker in the Early
Archean, compared to the present thickness of about 7 km
(Sleep and Windley, 1982). It is also possible that there were
more numerous, but smaller, plates in the Archean (e.g., Pollack, 1997; deWit, 1998) and that there was a coupled reduction in the number of plates and increase in the size of plates
with time, perhaps leading to a progression to larger supercontinents with time.
Modern oceanic lithosphere reaches neutral buoyancy in
about 20 m.y. (Davies, 1992), after which it becomes negatively buoyant. The observed mean age of oceanic crust
when it begins to subduct is 60 (Parsons, 1982) to 80 m.y.
(Sprague and Pollack, 1980). In the presence of a hotter
mantle, it would take oceanic plates longer to become neutrally buoyant, and thus, neutral buoyancy would not be
reached until about 80 m.y. in the Late Archean (Fig. 5). In
0361-0128/98/000/000-00 $6.00

addition, given a hotter mantle, it is probable that plates


would move faster and, thus in the Late Archean, spreading
crust may have subducted sooner, possibly in about 35 m.y.
and perhaps as rapidly as 20 m.y. (Bickle, 1986). Whereas the
time to reach neutral buoyancy in modern oceans is less than
the time it takes for spreading crust to reach an active trench,
oceanic crust in the Archean may have encountered active
subduction zones while it was still buoyant. The crossover in
these two ages depends on which values are accepted for
subduction rates but is likely to be between 2.5 and 2.0 Ga
(Fig. 5), the time that Archean buoyant plate tectonics
(deWit, 1998) may have started to evolve into modern plate
tectonics.
The Archean and probably Paleoproterozoic plate tectonic
dynamics were likely controlled by various types of plume
events, the largest of which were responsible for voluminous
continental growth and accretion of buoyant oceanic lithosphere. The voluminous komatiites and related basalts in the

Time to reach neutral buoyancy


or for subduction to begin (m.yr.)

minor role. That supercontinent formation occurred nearly


simultaneously with the 1.9 Ga plume event may not be coincidental. Perhaps breakup of the Late Archean supercontinent at 2.2 to 2.1 Ga served as a trigger for the 1.9 Ga plume
event, and, in this sense, a catastrophic event provided positive feedback for crustal growth that began during a shielding
plume event. A growing supercontinent also may contribute
to the preservation of juvenile crust by trapping it in collisional and accretionary orogens.
Broadly, the mineral deposits shown in Figure 1 have temporal distributions related, at least in part, to the supercontinent cycle (e.g., Barley et al., 1998). As shown in Figure 1A,
both Abitibi-type volcanic-hosted massive sulfide (VHMS)
and orogenic gold deposits correlate with a possible catastrophic mantle-plume event at 2.7 Ga, and ore types in Figure 1B may correlate with shielding-type plume events.

Subduction
Neutral Buoyancy

80
70
60
50
40
30
20
10
4.0 3.5

3.0

2.5

2.0 1.5

1.0

0.5

Age (Ga)
FIG. 5. Effect of mantle cooling on the time needed for a plate to reach
neutral buoyancy and the time needed for subduction to commence. Modified after Davies (1992).

207

208

GROVES ET AL.

Archean and, to a lesser extent, Paleoproterozoic granitoidgreenstone terranes attest to the widespread influence of
plume tectonics (e.g., Campbell et al., 1989; Pirajno, 2000).
There have been arguments against Archean plate tectonics
to produce the magmatic and deformational features of
greenstone belts (e.g., Hamilton, 1998). However, the
nonkomatiitic volcanic and volcano-sedimentary successions
in the greenstone belts are remarkably similar in their compositional range and geochemistry to volcanic successions in
modern convergent settings, such as arcs, backarcs, and interarc rifts, with differences readily explained by contrasts
between shallow subduction of hot, old, Archean oceanic
lithosphere versus steeper subduction of colder, younger
oceanic lithosphere (e.g., Barley et al., 1998). This implicates
the existence of Archean-Paleoproterozoic plume-influenced, modified plate tectonics. This is strongly supported
by the recognition of fossil subduction zones in the Abitibi
belt of Canada during the LITHOPROBE project (e.g., Ludden and Hynes, 2000). The similarity in structures and
chronology of deformation events in the Precambrian greenstone belts and modern convergent margins and the abundance in both of orogenic gold deposits (Groves et al., 1998)
and VHMS deposits (e.g., Franklin et al., 1981) is discussed
further below.
Evolution of the Subcontinental Lithospheric Mantle
The higher heat flow and greater plume activity in the early
Earth should also be reflected in the mantle lithosphere
below the continental crust. There is evidence for this in the
unusual, broadly equidimensional (~1,0001,500-km diam)
shapes of the Archean and Paleoproterozoic cratons and the
anomalously high abundance of granitoids that make these
cratons readily identifiable in maps or remote sensing images
at the regional scale. Although individual greenstone belts are
elongate in a pattern similar to that of modern orogenic belts,
the largely granitic cratons in which they are encased are
equidimensional. Each craton (i.e., granitoid-greenstone terrane) has its own volcanic history and specific substrate, and
its own distinctive metallogenic associations. For example,
only the Yilgarn craton contains both giant gold and nickel deposits such as those of the Eastern Goldfields province, and
only the Abitibi belt of the Superior craton contains both
world-class gold and VHMS deposits, each hosted in ca. 2.72
to 2.65 Ga greenstone belts. The different metallogeny of
these cratons may partly reflect the smaller size of Archean
plates and certainly relates to the nature of the crust (and
lithosphere) in the various provinces at the time of mineralization. For example, the VHMS deposits of the Abitibi belt
were mostly formed on primitive crust (Ayer et al., 2002),
whereas the komatiite-hosted nickel deposits in the Yilgarn
craton formed on rifted continental crust (Krapez et al.,
2000), as also shown by the predominance of granitoids and
tonalites in the Abitibi belt compared with monzogranites in
the Yilgarn craton (Champion and Sheraton, 1997). DeWit
and Thiart (2003) further demonstrate that individual cratons
at a global scale have their own distinctive metallogenic associations regardless of the timing of mineralization (pre-, syn-,
or postcratonization). The postcratonization example suggests
not only that each craton is unique but also that it has unique
subcontinental lithospheric mantle reflected by the nature of
0361-0128/98/000/000-00 $6.00

the mineral deposits (e.g., PGE, diamonds, Fe oxide Cu-Au)


that form after craton development. It is thus important from
a metallogenic viewpoint to examine the evolution of the subcontinental lithospheric mantle (e.g., Richter, 1985; Jordan,
1988) and its potential effect on the temporal distribution of
mineral deposits.
The density of the subcontinental lithospheric mantle beneath stabilized continental crust of varying ages has been determined using data on composition, thermal state, and petrological thickness (e.g., OReilly and Griffin, 1996; Poudjom
Djomani et al., 2001). Data from mantle-derived peridotite
xenoliths and garnet xenocrysts provide a pattern of secular
evolution of such lithosphere, with progressively less depletion in Al and Ca and lower Mg no. and Fe/Al from the
Archean to the Phanerozoic. Thermobarometric data from
the xenolith and xenocryst suites show that paleogeotherms
were lower in Archean than in Phanerozoic subcontinental
lithospheric mantle and that the typical thickness of the
lithosphere, defined as a chemical boundary layer, was greater
in the Archean (250180 km) than in the Proterozoic
(180150 km) and Phanerozoic (14060 km). From these
data, and data on mineral end members that constitute the
subcontinental lithospheric mantle, Poudjom Djomani et al.
(2001) calculate mean densities (at 20C) for Archean subcontinental lithospheric mantle of 3.31 0.016 Mg/m3 compared to Proterozoic and Phanerozoic equivalents of 3.35
0.02 and 3.36 0.02 Mg/m3, respectively. Thus, not only are
there significant variations in the compositions of subcontinental lithospheric mantle with time (Fig. 6), but these translate, in an analogous way to oceanic lithosphere discussed
above, into temporal variations in density, and hence buoyancy, of the subcontinental lithospheric mantle (Fig. 7). This
secular evolution of the subcontinental lithospheric mantle
implies broadly synchronous formation of crust and its lithospheric root and their linkage through their subsequent history
(Griffin et al., 2003; Sleep, 2003). This elegantly explains the
individual metallogenic signatures of Precambrian cratons
(deWit and Thiart, 2003), as each craton may have a lithospheric root of different depth extent and composition, dependent on degree of mantle melt extraction, mantle incompatible-element metasomatism, and degree of modification
along craton margins. Griffin et al. (2003) also argue that the
unique Archean subcontinental lithospheric mantle roots represent residues and/or cumulates from deep high-degree
melting related to the unique abundance of major mantleplume events of varying type, as discussed above. The exact
processes by which subcontinental lithospheric mantle
formed and was coupled to the crust are beyond the scope of
this paper, but various models are discussed by Arndt et al.
(2002), and a schematic model of lithosphere accretion involving plume-arc interaction is presented in Kerrich et al.
(2000). Whatever the process, massive melting events in the
mantle, and subsequently in the lower and middle crust to
produce many of the voluminous granitoids that typify the
overlying terranes (e.g., Champion and Sheraton, 1997),
probably were the crustal-scale cause of the broadly equidimensional surface expressions of early Precambrian cratons.
There are major implications from these studies in terms of
tectonics and metallogeny. As summarized by Poudjom
Djomani et al. (2001), Archean subcontinental lithospheric

208

209

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

Archean
Kaapvaal
> 90 Ma

Pr ojected Depth (km )

50

Proterozoic
Slave craton
Canada

Siberia

Phanerozoic

NE Siberia

100

E China

50

50

100

100

LAB
Depleted
Harzburgite

150

150

Depleted
Lherzolite

150

LAB

Fertile
Lherzolite
Lherzolite (LRE E
metasomatized)

200

LAB

200

200

Peridotite (Meltmetasomatized)

LAB
0

50

50

50

100

50

100

50

100

Cumulative % of each rock type


FIG. 6. Depth and depth variation in composition of Archean, Proterozoic, and Phanerozoic lithospheric sections, including changes in the composition of subcontinental lithospheric mantle with time. Profiles for each age of lithosphere show
estimated proportions of components (see legend) at various depths in the lithosphere. LAB = lithosphere-asthenosphere
boundary. From Griffin et al. (2003).

mantle would have been buoyant relative to asthenosphere in


any reasonable geologic scenario. As this buoyancy reduces
stress and viscosity, Archean subcontinental lithospheric mantle could not have been delaminated solely by gravitational
processes and would have been preserved unless disrupted by
rifting and replacement by more fertile asthenosphere. Thus,
mineral deposits formed late in Archean craton evolution, or
in anorogenic settings after Archean cratonization, should
have enhanced preservation potential.
The calculations by Poujom Djomani et al. (2001) indicate
that Proterozoic subcontinental lithospheric mantle was
denser than the Archean equivalent, but that typical 150- to

A)

Proterozoic
2
(40mW/m )

100
Archean
2
(35mW/m )

LAB

150

LAB

200

3.30

3.34

3.38

Crust
Upper Mantle

50

Depth (km)

Depth (km)

B)

Crust
Upper Mantle

50

250
3.26

180-km-thick subcontinental lithospheric mantle was likely


moderately buoyant nonetheless, and thus also unlikely to
have been delaminated. As in Archean cratons, Paleoproterozoic mineral deposits that formed late in the cratonization
process should still have had relatively high preservation potential.
In contrast, the commonly <100-km-thick Phanerozoic
subcontinental lithospheric mantle should be negatively
buoyant and readily delaminated, allowing the generation of
wide, gravitationally unstable, orogenic belts. Thus, preservation of mineral deposits formed above such lithosphere is less
likely than for those above Precambrian lithosphere, and

LAB

100
150

Phanerozoic
2
(50mW/m )

Southeastern
Australia

200
250
3.26

3.42

3.30

3.34
3.38
3
Cumulative density (g/cm )

FIG. 7. Density profiles and average heat flow for (A) Archean and Proterozoic and (B) Phanerozoic subcontinental lithospheric mantle and for the asthenosphere. LAB = lithosphere-asthenosphere boundary. Adapted from Poudjom Djomani et
al. (2001). The stars show where the isotherms for each lithosphere cross the LAB.
0361-0128/98/000/000-00 $6.00

209

3.42

210

GROVES ET AL.

older Phanerozoic and Neoproterozoic deposits are least


likely to have been preserved.
The temporal evolution of subcontinental lithospheric
mantle, which directly reflects the secular change from
strongly plume-influenced Archean tectonics to Phanerozoicstyle plate tectonics, must be considered a major factor influencing secular variations in both mineral deposit types and
the present abundance of specific deposit styles.
Tectonic Evolution and Secular Change in
Metallogeny: Useful Markers
To determine links between secular tectonic and metallogenic evolution, it is imperative to consider mineral deposit
styles that are at least only mildly affected by other evolutionary changes. Thus, it is important to avoid deposits of strongly
redox-sensitive metals (e.g., sediment-hosted deposits), the
temporal evolution of which may be linked to any temporal
change in the atmosphere-hydrosphere-biosphere system as
well as tectonic environment.
Epigenetic gold deposits are potentially useful for testing
secular changes in tectonic processes because most such deposits formed below the influence of surficial processes and,
hence, were unlikely to have been influenced by any secular
variation in atmosphere-hydrosphere-biosphere systems. Porphyry Cu-Au deposits, some of which are significant gold deposits in their own right (e.g., Kesler et al., 2002) and epithermal Au-Ag Cutype deposits have a strong tectonic
control in convergent margin settings (e.g., Sillitoe, 1997),
particularly in tectonic settings with anomalous, high K magmatism (e.g., Barley et al., 2002). However, they form at shallow crustal levels (<3 and <1 km, respectively) in arc and
backarc environments with high to extremely high uplift
rates. Thus, epithermal Au-Ag Cu deposits are mostly Tertiary or younger, with some examples of Mesozoic and fewer
examples of Paleozoic age. Porphyry Cu-Au deposits are
mostly younger than Mesozoic, although significant Paleozoic
examples are increasingly being recognized in Central Asia,
Mongolia, and China (e.g., Perello et al., 2001; Yakubchuk et
al., 2002), with older examples apparently selectively preserved by accretion or collision of hosting arcs with continental blocks. The virtual absence of porphyry Cu-Au and Mo deposits from the mid-Paleozoic to the Late Paleoproterozoic,
but their appearance there and in the Archean (Fig. 1A), is an
interesting temporal pattern that must be related to preservation rather than formation processes, as there were suitable
arc environments throughout Earth history.
In contrast to the porphyry and epithermal deposits, orogenic gold deposits, although formed in similar convergent
margin settings, were deposited in a wide range of crustal environments from a 3- to 20-km depth (as summarized by
Groves, 1993, from previous literature, particularly Hodgson
and MacGeehan, 1982, and Colvine et al., 1984) during the
late compressional to transpressional deformation that stabilized their host orogens (e.g., as summarized by Kerrich and
Cassidy, 1994; Groves et al., 1998). Orogenic deposits formed
over more than 3.4 b.y. of Earth history (e.g., Goldfarb et al.,
2001a, b) are abundant and geographically widespread and
hence are potentially sensitive tracers of temporal changes in
tectonic processes. For this reason, they are used here as the
primary example to test the correlation between tectonic
0361-0128/98/000/000-00 $6.00

evolution and secular changes in metallogeny. VHMS deposits, which formed over at least 3.25 b.y. of Earth history
(e.g., Barrie and Hannington, 1999), are the only other deposit type to show such a long-lived distribution. They formed
in deep marine basins and were preserved beneath subsequent, rapidly erupted volcanic or rapidly deposited sedimentary sequences. They were either accreted into the same
convergent margin settings in which orogenic gold deposits
were forming (e.g., Titley, 1993) or were actually formed
there (e.g., Solomon and Quesada, 2003). A brief discussion
of their secular distribution, plus that of intrusion-related
gold, IOCG and paleoplacer gold, follows the discussion of
orogenic gold deposits.
Tectonic Evolution and Secular Change in
Metallogeny: Messages from Orogenic Gold Deposits
As discussed by Goldfarb et al. (2001a, b), most Mesozoic
to Tertiary orogenic gold deposits coincide with external
ocean margins, where accretion of juvenile crust took place in
environments in which large thermal anomalies were associated with thrust-related thickening of radiogenic crust (e.g.,
Jamieson et al., 1998) or upwelling of asthenosphere due to
ridge subduction (e.g., Haeussler et al., 1995), subduction
rollback (Landefeld, 1988), or lithospheric delamination or
erosion (e.g., Griffin et al., 1998). Precambrian deposits appear to have formed also in similar tectonic settings of anomalously high thermal energy (e.g., Kerrich and Cassidy, 1994;
Qiu and Groves, 1999; Wyman et al., 1999). Thus, orogenic
gold deposits of all ages record orogen-wide fluxes of deeply
sourced heat and fluid as part of the orogenic process in convergent margins (e.g., Fyfe and Kerrich, 1985), almost certainly in response to global-scale tectonic events.
Figure 8 shows a comparison of the timing of orogenic gold
deposit formation (Goldfarb et al., 2001a) and periods of
crustal growth (Condie, 2000). It must be noted that not all
orogenic gold provinces are well dated using robust
geochronology and that their temporal distribution in Figure
8 is a best current estimate only. Similarly, as discussed above,
the interpretation of ages of crustal growth is evolving and
displayed ages have an uncertainty of 50 Ma. The earliest
orogenic gold deposits formed in the Middle Archean in the
Pilbara and Kaapvaal cratons, from about 3.4 Ga (e.g., Zegers
et al., 2002), with the economically most important Barberton
deposits having formed at around 3.1 Ga (e.g., de Ronde et
al., 1991). The earliest orogenic gold deposits in the Pilbara
craton broadly correspond to the earliest significant growth of
continental crust, and the Barberton deposits correspond to
the start of a broad peak in crust formation (Figs. 2, 8). If the
Witwatersrand deposits are of paleoplacer origin (Kirk et al.,
2001) and derived from the erosion of a large source area following the collision of relatively small Archean crustal blocks
(Frimmel et al., 2005), then potentially both Middle Archean
juvenile crust and orogenic gold abundance would have been
significantly greater than currently exposed.
The formation ages of subsequent orogenic gold deposits
define two major Precambrian peaks at 2.75 to 2.55 Ga (centered at about 2.65 Ga) and 2.1 to 1.75 Ga (probably two
peaks centered approximately at 2.0 and 1.8 Ga). There is a
marked lack of deposits between 1.7 Ga and 600 Ma and a
cyclic distribution from about 600 to 50 Ma (Fig. 8A). The

210

211

Russia - East China

Central Asia - Tien Shan


Mother Lode

Victoria

Urals

Baikal

Arabian - Nubian Shield

SW Siberia

West Africa

Homestake

10

Amazonian

30

Kolar

100
50

Quadrilatero Ferrifero

Superior

200

Barberton

Gold Resource (Moz)

A)

Yilgarn

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

B)
2.7 Ga

12
10

50 %

1.9 Ga

25 %

Volume per cent growth

14

6
4
2
3

0.5

Age (Ga)
FIG. 8. Timing of orogenic gold deposits vs. periods of crustal growth. A. Distribution of major orogenic gold provinces
with time. Adapted from Goldfarb et al. (2001a). See Groves et al. (2003) for updated version. B. Temporal evolution of continental crustal growth. Note, y-axis shows relative crustal growth. Adapted from Condie (2000). Variable bar width in (A)
due to varying uncertainties in age of mineralization, and in (B) to better illustrate major periods of crustal growth. The annotations 50 and 25 percent refer to approximate percentages of recorded crustal growth in the Late Archean and mid to
late Paleoproterozoic, respectively.

Late Archean and Paleoproterozoic peaks broadly reflect the


major periods of episodic growth of Precambrian continental
crust, centered at ca. 2.7 and 1.9 Ga (Fig. 8B), although
peaks in the formation of orogenic gold deposits clearly flank
the latter peak in crustal growth. Similarly, the Phanerozoic
orogenic gold deposits broadly mimic the more continuous,
shorter wavelength distribution of Phanerozoic Cordillerantype orogenic events and associated periods of crustal
growth, although accurate correlations are limited by uncertainties, particularly in the ages of some gold provinces. Despite these uncertainties, the secular evolution of orogenic
gold deposits clearly reflects the proposed evolutionary trend
from strongly episodic, plume-influenced plate tectonics in
the Archean to more cyclic modern-style plate tectonics.
Archean and, to a lesser extent, Paleoproterozoic orogenic
gold provinces related to the catastrophic mantle-plume
events of Condie (2004) have equivalent gold production
(Fig. 8A), despite their age, deep erosion, and common deep
regolith cover. This, combined with evidence for relatively
thin lithosphere at the time of gold deposit formation and
0361-0128/98/000/000-00 $6.00

prior to cratonization for some of the larger Archean gold


provinces (e.g., Abitibi belt and Kalgoorlie terrane; Groves et
al., 2003) and Paleoproterozoic gold provinces (e.g., Ashanti
belt; Pigois et al., 2003), suggests that crustal-scale thermal
input may have played a critical role in the formation of giant
orogenic gold provinces in the youngest and most juvenile
crust preserved.
Despite the strong Mesoproterozoic to Neoproterozoic
record of crustal growth from 1.8 to 1.2 Ga (Fig. 8B), in part
correlating with the formation of Rodinia at ca 1.3 to 1.0 Ga
through a number of continental collisions (e.g., Dalziel,
1991; Hoffman, 1991), there is a lack of gold deposits between about 1.7 Ga and 600 Ma (Fig. 8A). The Olympiada
deposit of southwestern Siberia (ca. 820 Ma; Safonov, 1997),
as shown in Figure 8A, may be an important exception
although it could be as young as 600 Ma (Konstantinov et al.,
1999). This suggests that an additional factor, other than just
the growth of juvenile crust, was important in determining
the temporal distribution of orogenic gold deposits preserved
in the geologic record. Clues to the nature of this factor lie in

211

212

GROVES ET AL.

the contrasting regional-scale shapes of the major gold


provinces of different age. Archean and Paleoproterozoic orogenic gold deposits are distributed along elongate
supracrustal belts, but these are hosted within and commonly
in the central part of (cf. fig. 5.2 of Solomon and Groves,
1994) roughly equidimensional cratons as noted above. In
contrast, Phanerozoic deposits are distributed in elongate
belts along the margins of these cratons or along the margins
of older, marginal Phanerozoic belts (fig. 5 of Goldfarb et al.,
2001a). The time of transition between these styles of orogenic belts is unclear, but reconstructions of Rodinia (Fig. 9)
suggest that modern-style orogenic belts were in existence
prior to 1.0 Ga. These observations, combined with evidence
for timing of changes in buoyancy of the oceanic lithosphere
(Fig. 5) and subcontinental lithospheric mantle (Figs. 67),
suggest that an important time in the transition from plumeinfluenced to modern-style plate tectonics was somewhere
near the end of the Archean to early Paleoproterozoic.
The unusual lack of Mesoproterozoic to Neoproterozoic
orogenic gold deposits (Fig. 8A) can be explained largely in
terms of the preservation potential of the hosting terranes.
Deposits embedded in Archean and Paleoproterozoic cratons would be underlain by relatively buoyant subcontinent
lithospheric mantle, which would be difficult to delaminate.
Hence, there would be a very high chance of preservation except adjacent to craton margins or in relatively small cratonic
blocks, as probably existed in the Middle Archean, where

later orogeny could cause uplift and erosion. In contrast,


modern-style, highly elongate, metasedimentary and
metavolcanic rock-dominated accretionary belts along the
margins of the cratons, with their negatively buoyant subcontinent lithospheric mantle, would be much more prone to
uplift and erosion. Thus, it is likely that Mesoproterozoic to
Neoproterozoic orogenic gold deposits did form during
major continental crust-forming events in the period from
ca. 1.7 Ga to 600 Ma, but most were removed by long-lived
erosion of the narrow continental margin orogens down to
their high metamorphic grade root zones; such zones are
below the crustal depths that typically would contain orogenic deposits (Groves et al., 1998). The reappearance of
abundant orogenic gold deposits at ca. 600 Ma suggests that
this is an approximate threshold for preservation, or lack of
complete destruction, of deposits in modern-style orogenic
belts. The abundance of gold placers spatially associated with
many Phanerozoic orogenic gold provinces attests to their
progressive erosion such that, within 500 to 600 m.y. of initial unroofing, an orogenic gold province and associated placers may be totally lost from the geologic record under conditions of Cordilleran-style plate tectonics. Similarly, the lack
of any economically significant orogenic gold provinces in
rocks younger than ca. 50 Ma suggests that 50 m.y. may be
the minimum period to uplift and expose a productive
province (Goldfarb et al., 2001a). Uplift rates were lower
than those for the arc-related porphyry and epithermal

Tibet
Madagascar

Rodinia

INDI A

1.3 - 0.9 Ga

S China
Kalahari

Malaysia

Orogens
Juvenile Crust

N China

S Australia
N Australia
SIBERIA
LAURENTIA

Mawson

Berentsia
AMAZONIA

La Plata
West
Africa

BALTICA

FIG. 9. Schematic representation of Rodinia, a supercontinent formed between 1.3 and 0.9 Ga and fragmented between
750 and 600 Ma. The Kalahari, La Plata (Rio de La Plata) and West Africa cratons were probably never part of Rodinia. Reconstruction modified after Tohver et al. (2002) and Pisarevsky et al. (2003). Note the location of the linear, elongate, Mesoproterozoic cratonic belts (orogens; shaded medium gray) around relatively equidimensional Archean-Paleoproterozoic cratonic blocks.
0361-0128/98/000/000-00 $6.00

212

213

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

2000
1000

W Tasmania
(Mt Lyell)

Age (Ga)

0.5

Turkey (Murgal)
Japan (Hokuroko)

Manitoba (Flin Flon)

Abitibi (Kidd Creek)

Mt (ore)

3000

Iberia (Rio Tinto)


NW British Columbia
(Windy Craggy)

Volcanic-hosted massive sulfide deposits


VHMS deposits, many of which are gold rich (Huston,
2000), although commonly subdivided into subtypes based on
geographic settings (e.g., Meyer, 1981; Fig. 1A) or metal ratios (e.g., Large, 1992), are a coherent class of deposits
formed at or below the sea floor by circulating hot seawater
(e.g., Barrie and Hannington, 1999). Their distribution

Urals (Gai)

Examples of Tectonically Induced Selective Preservation


Other mineral deposit types that are little influenced by
near-surface redox conditions are examined below. Most
show temporal patterns that relate to preservation processes
(e.g., VHMS, paleoplacer gold), whereas the distribution of
others can be related to the development of Archean or Paleoproterozoic subcontinental lithospheric mantle (e.g., IOCG,
intrusion-related gold).

through time has been examined by Meyer (1981), Titley


(1993), and Barrie and Hannington (1999), among others.
Production data for many deposits, particularly those in the
former Soviet Union and other Asian and Eastern Block
countries, are incomplete or contradictory, as, importantly,
are accurate ages for some deposits. Modern VHMS systems
are present at sea-floor spreading ridges, such as the East Pacific Rise, and also in backarc basins, such as the Lau basin.
Where the former eventually reach the continental margin,
they may be incorporated into the accreting margin in tectonic environments similar to those in which younger orogenic gold deposits were generated (e.g., Goldfarb, 1997).
Solomon and Quesada (2003) point out that some deposits
such as those of the Iberian pyrite belt may form in terranes
after their accretion to a margin. Where VHMS deposits and
orogenic gold deposits occur in the same terrane, the former
are always older (e.g., Spooner and Barrie, 1993), possibly
leading to overprinting of VHMS deposits by orogenic gold
mineralization, as summarized by Groves et al. (2003), although this appears to be rare (e.g., Hannington et al., 1999).
The VHMS deposits (Fig. 10) show a preservation history
that is broadly similar to that of orogenic gold deposits, with
episodic development in the early Precambrian and a more
cyclic pattern in the Paleozoic. The oldest, but noneconomic
VHMS deposits in the east Pilbara and Barberton terranes, at
ca. 3.5 to 3.25 Ga (e.g., Vearncombe et al., 1995), correspond
broadly with the oldest orogenic gold events globally (e.g.,
Zegers et al., 2002) and with the earliest record of formation
of significant continental crust (Fig. 2). VHMS deposits are
not uniformly distributed in greenstone belts globally, with
most belts being devoid of major deposits. The economically
significant ca. 2.7 Ga VHMS deposits are mainly from the
Abitibi belt. The ca. 1.85 Ga VHMS deposits are in the Flin
Flon district, Manitoba, Canada (Syme and Bailes, 1993) and
Wisconsin. These occurrences broadly coincide with hypothesized mantle-plume events at 2.7 and 1.9 Ga, respectively
(Fig. 8B). Both regions are composed of terranes displaying
evidence for thin, dominantly oceanic lithosphere at the time
of VHMS mineralization (e.g., Stern et al., 1999; Ayer et al.,
2002; Hart et al., 2004). Similarly, VHMS deposits appear to
have formed and been incorporated into continental crust almost continuously since the latest Neoproterozoic, although

Kazakhstan (Maikain)

provinces of the present-day circum-Pacific, perhaps because the orogenic gold deposits formed largely during transpressional, rather than compressional, tectonics (e.g., Goldfarb et al., 2001a).
Thus, it appears that orogenic gold deposits are inherent
products of crust-forming events throughout Earth history.
The processes responsible for the formation of this gold deposit type were broadly similar through time, reflecting moderate- to high-temperature tectonic events superimposed for
the first time on growing, volatile-rich (i.e., H2O, CO2, H2S)
juvenile continental crust. However, the change from a
plume-influenced plate tectonic style in the latest Archean to
early Paleoproterozoic to a modern tectonic style affected the
preservation potential of terranes of different age. Hence, the
temporal distribution of the deposits (Fig. 8A) reflects a combination of processes of formation and preservation rather
than a fundamental change due to secular changes in tectonics. In the Archean and Paleoproterozoic, crustal growth and
preservation processes worked together to produce the abundant deposits in those time periods, whereas preservation was
not favored during the critical tectonic transition in the
Mesoproterozoic to Neoproterozoic; that is, crustal growth
and associated metallogenesis and preservation were coupled
in the Archean and Paleoproterozoic but largely decoupled
thereafter.

FIG. 10. Temporal distribution of volcanic-hosted massive sulfide deposits. Data derived from Barrie and Hannington
(1999)examples of specific deposits given in parentheses. Due to some uncertainties in age, deposits are grouped in places
into broad time groups rather than absolute time periods. Note expanded time scale from 0 to 0.5 Ga.
0361-0128/98/000/000-00 $6.00

213

214

GROVES ET AL.

there are distinct peaks in the early to middle Paleozoic and


in the Mesozoic (e.g., Titley, 1993), as there are for orogenic
gold deposits. Similarly, although VHMS deposits are recognized between ca. 1.7 Ga (e.g., Jerome; McCandless et al.,
1999) and 600 Ma (e.g., Barrie and Hannington, 1999), they
too are markedly under-represented in this age range. Again,
it appears most likely that the VHMS deposits were largely
lost from the geologic record because the linear belts in
which they formed occur above less buoyant subcontinental
lithospheric mantle and were uplifted and deeply eroded to
their roots. The youngest exposed VHMS deposits and/or
prospects are of Tertiary age, for example Ufuo in PNG (Corlett and Akiro, 1999), deposits in Prince William Sound of
southern Alaska (Goldfarb, 1997), and deposits in Ecuador
(Chiaradia and Fontbot, 2001), corresponding broadly to the
age of the youngest exposed, orogenic gold deposits.
The most logical conclusion is that, although VHMS deposits formed throughout Earth history, their temporal distribution, as for orogenic gold deposits, largely represents a pattern of preservation due to changes in the lithosphere caused
by changing global tectonic processes.
Placer and paleoplacer gold
Most giant placer gold deposits were deposited in foreland
(commonly retroarc) basins in Mesozoic-Cenozoic convergent margins of the circum-Pacific (e.g., New Zealand, California, Alaska) through the erosion of Paleozoic to Mesozoic
orogenic gold deposits (e.g., Henley and Adams, 1979; Edwards and Atkinson, 1986; Goldfarb et al., 1998). Similar
Paleozoic margins were the source for additional large placer
fields in Victoria (Hughes et al., 2004) and the Eastern
Cordillera of South America (e.g., Haeberlin et al., 2003), although much of the central Asian Paleo-Tethyan margin was

preserved by subsequent Himalayan continent-continent collision. In addition, some significant gold placers formed in
permafrost regions (e.g., Eastern Russia, Siberia) or in areas
of deep tropical weathering of Precambrian orogenic gold deposits, as for example in the Ashanti belt of Ghana and the
Tapajos region of the Amazon (e.g., Santos et al., 2001). Most
placer deposits were mined from Recent river systems and
beaches, although some Tertiary paleoplacers were preserved
by overlying volcanic and volcaniclastic rocks (e.g., Ballarat
and Bendigo, Victoria, Australia). The fact that most worldclass placers are associated with source lodes older than ca.
100 Ma (i.e., Fairbanks, Nome, Eastern Russia) suggests that
~50 m.y. is an approximate threshold between unroofing of a
gold province in a Cordilleran-style orogen and loss of a significant percentage of gold to the secondary environment.
Paleoplacer gold deposits are rare in the geologic record
before the Tertiary (Fig. 11), yet the giant Late Archean Witwatersrand deposits represent the largest gold province on
Earth. Their origin has been controversial, with both modified placer (e.g., Minter et al., 1993) and hydrothermal (e.g.,
Phillips and Myers, 1989; Barnicoat et al., 1997) models being
proposed. However, a variety of recent evidence, particularly
from Re-Os dating of both gold and associated rounded
pyrites, which yields ca. 3.0 Ga ages, which are presedimentation of the host conglomerates (e.g., Kirk et al., 2001, 2002),
suggests that the Witwatersrand gold ores are modified paleoplacers, as summarized by Frimmel et al. (2005).
Significantly, the Witwatersrand gold was deposited in the
Central Rand Group in a retroarc foreland basin setting
(Kositcin and Krapez, 2004) that is similar to many modern
depositional settings of placer gold. Extreme environmental
conditions, specific to the early Earth, including a potentially
more acidic and chemically aggressive atmosphere (Holland,
3000

100,000
Witwatersrand

300

10,000

15
Tarkwa

400

Moz (Au)

t (Au)

500
Modern Placers

300
200

100

Jacobina and Roraima

3.0

2.5

2.0

1.5

1.0

0.5

Age (Ga)
FIG. 11. Temporal distribution of placer and paleoplacer gold deposits. Main sources of data are Boyle (1979), Bache
(1987), Goldfarb et al. (1998), Milesi et al. (2002), and Frimmel et al. (2005). The Tertiary-Recent placer total production is
a minimum as a significant proportion of placer gold production by small-scale miners may have been unrecorded in official
production records. The total production is given for Tertiary to Recent placers as a single bar as there are considerable uncertainties in the age of some placers.
0361-0128/98/000/000-00 $6.00

214

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

1984), the lack of vegetation and other organisms, and hence


the predominance of braided stream environments and potential for effective wind erosion and sorting (e.g., Minter,
1999), can potentially explain the giant size of the Witwatersrand deposits. Some present-day placers are derived from
adjacent orogenic gold deposits (e.g., Hughes et al., 2004).
However, many modern placers are not derived from unusually large primary orogenic deposits (i.e., Nome, Alaska;
Klondike, Yukon) but owe their origin to tectonic, erosional,
and sedimentary processes that combine to produce extremely effective sorting and detrital gold concentration. The
source of the Witwatersrand gold is unclear. Frimmel et al.
(2005) suggest that the Witwatersrand gold, with anomalously
high Re and Os concentrations and 187Os/188Os of 0.108,
equivalent to the Os isotope composition of the 3.0 Ga mantle (e.g., Kirk et al., 2002), was likely derived from dispersed
magmatic phases in mantle-derived mafic and/or ultramafic
rocks of now eroded granitoid-greenstone terranes in the hinterland to the Central Rand basin. A source region with a bulk
concentration between about 0.5 and 7 ppb Au, in agreement
with a value of 4 ppb Au for Barberton-type granitoid-greenstone crust (Robb and Meyer, 1990), is all that is required to
form the Witwatersrand deposits (Loen, 1992), provided that
concentration mechanisms were very effective (Frimmel et
al., 2005). High background gold values in the early Archean
komatiites may have correlated with mantle plumes derived
from the mantle boundary with the gold-rich core (Kerrich,
1999). Alternatively, the gold could have been derived from
gold deposits equivalent in age to, or slightly younger than,
the orogenic Barberton gold deposits, in the proposed hinterland of the Witwatersrand basin, for example in the AmaliaKraaipan, Murchison, Pietersburg and Giyani belts, but the
age of this gold is poorly defined (e.g., Barton, 1984; Beattie
and Barton, 1992). A source of some gold from epithermal or
epizonal orogenic gold deposits is also reasonable because
some of the gold has low fineness and high mercury contents
(Feather and Koen, 1975). Interestingly, orogenic deposits in
the Murchison greenstone belt of the potential hinterland
also have gold with high mercury contents (cf. Poujol et al.,
1996). As stated above, the precise source or sources of the
gold are currently unclear, but mass-balance arguments
(Frimmel et al., 2005) clearly show that there was sufficient
gold in the hinterland of the Witwatersrand basin to produce
the recorded gold concentrations.
The question remains as to why these giant Witwatersrand
deposits are preserved. On a larger scale, the Witwatersrand
ores almost certainly owe their preservation to their location
in old subcontinental lithospheric mantle (e.g., Shirey et al.,
2003), where buoyancy protected them from subsequent destruction. Whereas the host Kaapvaal craton is not unique in
terms of the antiquity of its subcontinental lithospheric mantle (e.g., Richardson et al., 2004), the age of its earliest well
preserved basins (ca. 3.02.8 Ga) and the remarkable preservation of its sedimentary and volcanic basins from ca. 3.5 Ga
to the Mesoproterozoic attests to its long-term stability and
high preservation potential. Importantly, the Re-Os age of the
detrital gold and pyrite in the Witwatersrand paleoplacers is
ca. 3.0 Ga, 300 m.y. earlier than the first significant crustforming event at 2.7 Ga (Figs. 2, 8B), so that early subcontinental lithospheric mantle could have been in place at this
0361-0128/98/000/000-00 $6.00

215

time, as also indicated by the occurrence of detrital diamonds


in the Witwatersrand rocks. The detrital gold was presumably
derived from the upthrusted margins of relatively small continental blocks to the north and west, which would have been
susceptible to destruction despite potentially buoyant subcontinental lithospheric mantle. The Central Rand Group of
the Witwatersrand basin was probably deposited in the foreland of a collision between the amalgamated Witwatersrand
block to the south (Schmitz et al., 2004) and the Pietersburg
block to the north and was not related to the major interaction between the larger Zimbabwe and Kaapvaal cratons during Limpopo orogeny, more than 100 m.y. after Witwatersrand Supergroup sedimentation (Frimmel et al., 2005, and
references therein). The involvement of relatively small Middle Archean crustal blocks, rather than large cratons, can explain the widespread uplift and erosion of the now eroded
terranes that are the suggested source of Witwatersrand gold.
There are also clearly regional-scale processes that assisted
preservation. The most important was the outpouring of the
Klipriversberg Group basaltic lavas over the Witwatersrand sedimentary basin, but there was also the possible formation of a
resistant veneer of impact melt over at least part of the basin as
a result of the Vredefort meteorite impact in the center of the
Witwatersrand basin, as summarized by Frimmel et al. (2005).
The other significant gold paleoplacer provinces are Paleoproterozoic in age (Fig. 11), with Tarkwa in Ghana at ca. 2.1
Ga (Pigois et al., 2003) being the largest, and Jacobina (ca. 2.0
0.1 Ga) and Roraima (ca. 1.96 Ga) representing smaller
provinces (Frimmel et al., 2005). The gold production from
these probable foreland basin settings is two orders of magnitude lower than for the Witwatersrand (Fig. 11). These deposits are 100 to 200 m.y. older than the 1.9-Ga peak in crust
production and presumably formed from unroofing and exposure of early orogenic gold source provinces (see above). In
Ghana, at least, some orogenic gold deposits overprint preexisting Tarkwa-type paleoplacers (Pigois et al., 2003). As suggested for the Witwatersrand, it is, thus, most likely that the
paleoplacer deposits owe their preservation to the buoyancy
of the underlying Paleoproterozoic subcontinental lithospheric mantle, developed during the protracted orogenic
events in which the primary gold deposits formed.
It is suggested that placers and paleoplacers, like orogenic
gold and VHMS deposits, display a temporal pattern largely
dictated by preservation. Presumably, they formed throughout Earth history whenever orogenic gold provinces were uplifted and eroded, particularly after 1.7 Ga, but deposits
formed before the Tertiary only survived where regional-scale
processes assisted their preservation above buoyant subcontinental lithospheric mantle.
Examples of Tectonically Induced Selective
Formation and Preservation
Iron-oxide copper-gold (IOCG) deposits
The IOCG deposit type (e.g., Hitzman et al., 1992) has been
expanded to include many different styles of iron-oxiderich
mineralization that formed in a variety of tectonic settings (cf.
Hitzman, 2000), but only those deposits with significant copper- and iron-bearing sulfides and gold resources are considered here. As summarized by Groves and Vielreicher (2001),

215

216

GROVES ET AL.

their origin is equivocal. However, there is a clear connection


to alkaline magmatism, particularly if the giant Palabora FeP-REE-Cu-Au-PGE deposit of South Africa is included.
Neodymium isotope data for the giant Olympic Dam deposit,
South Australia, combined with gravity and magnetic data,
also implicate a mafic alkaline body at depth (e.g., Campbell
et al., 1998) that could be the feeder to lamprophyre dikes in
the deposit. Dating of the giant Salobo (Requia et al., 2003)
and Igarap Bahia (Tallarico et al., 2005) deposits at Carajs,
Brazil, also suggests a temporal association with A-type
granitic magmatism, as does the presence of smaller, younger
IOCG (Bi Sn W) and Au-PGE deposits associated with
two subsequent events of alkaline magmatism in the same region (Groves et al., 2004). Importantly, all significant Precambrian IOCG deposits are sited within about 100 km of the
margins of Archean cratons (e.g., Palabora, Carajs, Olympic
Dam) or close to the boundary of Archean and Paleoproterozoic lithosphere as interpreted from remote sensing geophysical data (e.g., Cloncurry, Queensland, Australia).
The temporal distribution of economically significant Precambrian IOCG deposits (Fig. 12) shows major peaks in the
latest Archean (ca. 2.57 Ga; e.g., Carajs deposits), Paleoproterozoic (ca. 2.05 Ga; e.g., Palabora), and Mesoproterozoic
(ca. 1.59 Ga; e.g., Olympic Dam) that are significantly offset
from the main periods of crustal growth at ca. 2.7 and 1.9 Ga
(Fig. 2). The oldest Carajs deposits are sited in a region of
Late Archean postcratonization platformal sequences of similar age to those in the Pilbara and Kaapvaal cratons, rather
than greenstone sequences (e.g., Grainger et al, 2002). The
IOCG deposits formed in the Amazon craton, however, at approximately the same time as late-orogenic gold deposits in
the eastern Dharwar block of the Indian Shield (Hamilton
and Hodgson, 1986; Balakrishnan et al., 1999), emphasizing
the diachroneity of Archean cratonization despite the major
peak in crustal growth at about 2.7 Ga. Presumably, large

2800

IOCG deposits could also have formed in the Pilbara and


Kaapvaal cratons in the Late Archean if alkaline magmatism
had occurred at that time. The Olympic Dam deposit formed
at ca. 1.59 Ga, very soon after amalgamation of Archean and
Paleoproterozoic blocks to form proto-Australia at ca. 1.75 to
1.70 Ga (e.g., Betts et al., 2002), in association with outpouring of the Gawler lavas and associated intrusions of alkaline
affinity. In each case, some form of anorogenic tectonism and
alkaline magmatism is implied.
The association of the giant Precambrian IOCG deposits
with alkaline magmatism and their location near craton margins suggests that the magmas were derived by small degrees
of post-tectonic partial melting of subcontinental lithospheric mantle previously metasomatized (K, U, Th, Au,
LREE) during and after the events related to cratonization.
This could explain the offset between the ca. >2.6 Ga ages of
crustal growth and/or cratonization and the younger periods
of alkaline magmatism and associated IOCG mineralization.
Thus, there is a major tectonic control on the formation of
large Precambrian IOCG deposits close to the margins of
preexisting cratons during periods of magmatism of alkaline
affinity. However, the temporal distribution of IOCG deposits (Fig. 12) also reflects preservation, with giant and
world-class Precambrian examples of the deposit type selectively formed and preserved in buoyant subcontinental
lithospheric mantle.
There are Neoproterozoic deposits of IOCG affinity at
Khetri in India (probably ca. 850 Ma in age; Knight et al.,
2002) and in the Lufilian intracratonic rift basin of Zambia, although large examples of the latter have low copper grades
(<0.9% Cu) and very low gold (<0.1g/t Au) and some contain
insignificant iron oxides (e.g., Nisbet et al., 2000; Hitzman,
2001). There are no significant economic examples of the deposit type until the Mesozoic when the Candelaria deposit
(ca. 115 Ma) was formed (Mathur et al., 2002). The deposit

Carajs Province

Mt (ore)

Olympic Dam

2000

1200

Palabora
Lufilian Arc
Candelaria

400
Khetri

Age (Ga)

FIG. 12. Temporal distribution of economically significant, unequivocal IOCG deposits through time. Data from Williams
and Skirrow (2000), Haynes (2002), and references therein.
0361-0128/98/000/000-00 $6.00

216

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

shows gross similarities to the Precambrian deposits, although


the sulfide assemblage is more sulfur rich, zinc and silver concentrations are anomalous, and the REE concentrations, despite similar LREE enrichment, are more erratic. Candelaria
clearly formed in a different tectonic setting, being related to
transpression and basin inversion in a long-lived, arc-parallel
strike-slip fault system linked to subduction in the coastal
batholith of Chile, outboard from the South American Shield
(e.g., Marschik and Fontbot, 2001). The ore-forming fluid is
considered to have been magmatic, and spatially related subalkaline to alkaline granitoids have Sr, Nd, and Pb isotope ratios that indicate derivation of the parent magmas from a subduction-modified mantle source (Marschik et al., 2003a, b),
as in Precambrian analogues. The shape of the orebody is
mantolike and grossly conformable to layering, rather than
pipelike, as for the Precambrian examples (Marschik and
Fontbot, 2001). This suggests emplacement of the ore into
more permeable horizons rather than brecciation under high
fluid pressures and may indicate a different genetic process.
Intrusion-related gold deposits
During the past decade, there has been an emphasis on the
diversity in gold deposit types within metamorphic belts (e.g.,
Robert and Poulsen, 1997), with increasing interest in a group
of deposits termed intrusion-related gold deposits (e.g., Sillitoe and Thompson, 1998). The original definition of this deposit type included a large number of different styles of deposits with different metal associations (e.g., Sillitoe, 1991),
but only the intrusion-related gold systems in the sense of
Lang and Baker (2001) are considered here. According to
Thompson et al. (1999) and Lang and Baker (2001), they are
characterized by (1) spatial association with relatively reduced
granitoids, (2) carbonic hydrothermal fluids, (3) gold as the
dominant economic element, but with anomalous Bi, W, Mo,
Te, and/or Sb, (4) a low sulfide content, with no iron oxides,
(5) spatially restricted and weak hydrothermal alteration in
mesozonal examples, (6) a tectonic setting well inboard of
recognized convergent plate boundaries, with complex granite petrogenesis, and (7) an association with tungsten and tin
provinces. There may also be regional metal zoning, with
more silver and/or base metal-rich deposits distal to the intrusion-related gold systems (e.g., Lang et al., 2000). As summarized by Groves et al. (2003), there is general acceptance
that the Fort Knox deposit in Alaska, various small and as yet
uneconomic deposits in the eastern part of the Tintina gold
province of Yukon, Canada, and disseminated gold prospects
at Timbarra, NSW, Australia, are intrusion-related gold deposits. A notable difference compared to orogenic-type gold
deposits is the order of magnitude lower gold grades. All
other large deposits that are sometimes classified as intrusionrelated gold deposits by different authors are of uncertain origin and have many characteristics more akin to orogenic gold
deposits (Groves et al., 2003). Hence, no temporal pattern is
presented here.
An important difference between deposit types is the tectonic setting of the undoubted intrusion-related gold deposits
that occur well inboard from the convergent margins and
orogenic gold deposits that form in the margins. The intrusion-related gold deposits are located in deformed and metamorphosed shelf sequences adjacent to cratonic margins,
0361-0128/98/000/000-00 $6.00

217

rather than in the turbiditic or mafic volcanic rock sequences


in commonly forearc settings that characterize orogenic gold
deposits. This also explains their spatial association with
broadly contemporaneous tin and tungsten deposits, which
also form adjacent to older continental crust well inboard
from subduction zones (e.g., Solomon and Groves, 1994). The
suites of felsic intrusions associated with intrusion-related
gold systems are unusual. Lang and Baker (2001) mention coeval intrusions of alkalic, metaluminous calc-alkalic, and peraluminous compositions. Mair et al. (2003) show that, in the
Tombstone belt of the Yukon, the intrusion-related gold depositassociated intrusions have mixed mantle and crustal signatures. They show that the complex array of intrusions is
likely due to mantle-derived mafic alkaline magmas impinging at the base of the crust, consequent melting of the crust
to produce felsic magmas, and contamination of these magmas by the basement and shelf sequences they intrude. These
factors may combine to produce fertile magmas with suitable
redox switchovers for the concentration of gold and its subsequent release in relatively reducing magmatic-hydrothermal
fluids.
The conjunction of anomalous granitic magmas and unusual inboard tectonic setting for their generation is likely to
have been rare in geologic history. Such settings would have
been extremely rare, or absent, in the early Precambrian
when the earliest cratons formed, as also shown by the lack of
tin and tungsten deposits related to fractionated granites in
similar settings. The formation of the deposits in crust above
younger subcontinental lithospheric mantle near the margins
of cratons also would have limited their preservation. Thus
their rarity and the restriction of undoubted examples to the
late Phanerozoic (Timbarra, Permo-Triassic; Tombstone deposits, Late Cretaceous) are expected.
Synthesis
Based on the synthesis of observations presented above, it
appears that Archean, and probably Paleoproterozoic, tectonic processes were dominated by relatively buoyant oceanic
lithosphere and plume events, with the largest, short-lived,
catastrophic events being responsible for voluminous continental growth. Plume-influenced to -dominated plate tectonics operated rather than a modern-style of plate tectonics.
The progressive decline of heat flow and decrease in plume
activity also affected the nature of the subcontinental lithospheric mantle, which records a distinctive temporal evolution
(e.g., Poudjom Djomani et al., 2001; Griffin et al., 2003)
linked to the synchronous formation of crust and its lithospheric root. Archean subcontinental lithospheric mantle, representing residues and/or cumulates from deep high-degree
melting related to the early abundance of mantle-plume
events, had a relatively low density and hence was buoyant.
The more or less similarly sized, broadly equidimensional
early Precambrian cratons were probably a result of these
processes and consequent melting of the lower and middle
crust. Subsequent Proterozoic lithosphere was more dense
and had positive to neutral buoyancy, whereas Phanerozoic
lithosphere was even more dense and negatively buoyant
(Poudjom Djomani et al., 2001). These changes in buoyancy
produced global tectonic patterns in which post-Mesoproterozoic orogenic belts surrounded early Precambrian nuclei.

217

218

GROVES ET AL.

Gold-bearing deposits that form at shallow crustal depths in


convergent margins (e.g., porphyry-skarn-epithermal Cu-MoAu-Ag systems) are highly susceptible to tectonic uplift and
erosion and are only sporadically and selectively preserved
from times older than the Mesozoic by accretion or collision
of their host arcs with continental blocks. In contrast, orogenic gold deposits formed over a much larger range of
crustal depths in deforming convergent margins undergoing
deformation for a period of at least 3.4 b.y. (Goldfarb et al.,
2001a, b). They have a temporal distribution that broadly mirrors that of juvenile crustal growth, particularly the change
from episodic to more cyclic growth with time, although they
are rare between 1.7 Ga and 600 Ma. The exceptional endowment of Archean and Paleoproterozoic provinces, despite
their age, suggests that plume-influenced plate tectonics produced large gold deposits due to associated high thermal flux.
They were incorporated into crust above buoyant subcontinental lithospheric mantle and preserved, particularly in the
centers of cratons. In the Mesoproterozoic, plume activity declined, tectonics akin to modern plate tectonics evolved, and
the subcontinental lithospheric mantle became negatively
buoyant. It appears that uplift and erosion in these newly
formed orogenic belts along the margins of early Precambrian
cratons destroyed the majority of any orogenic gold deposits
that formed between 1.7 Ga and 600 Ma. From 600 until
about 50 Ma, the temporal distribution of the deposits grossly
represents that of the cycle of orogenesis and crustal growth.
Thus, formation of the orogenic gold deposits was broadly
controlled by the timing and intensity of crust-forming
events, but their temporal distribution was strongly affected
by the buoyancy of the subcontinental lithospheric mantle
below their host terranes. VHMS deposits, which have a similar, almost 3.5-b.y. history of formation, have a very similar
temporal distribution, particularly their contrasting Precambrian and Phanerozoic patterns, suggesting that preservation
was a dominant factor in their present distribution. High
thermal flux related to plume-influenced plate tectonics produced highly endowed Archean and Paleoproterozoic VHMS
provinces, as for orogenic gold.
Gold placers have probably formed by erosion of orogenic
gold deposits since the Middle Archean, but paleoplacers and
placers display a highly anomalous temporal distribution with
peaks in the early Precambrian and in the Tertiary to Recent.
The giant, extraordinarily gold rich palaeoplacers of the Witwatersrand probably owe their formation to effective fluvial
sorting under extreme climatic conditions and their preservation to generation of early buoyant subcontinental lithospheric mantle under the Kaapvaal craton. It is also possible
that greenstone source rocks in the hinterland to the host
foreland basin were enriched in gold due to mantle-plume activity in the early Earth. Giant Precambrian IOCG deposits
also appear to have required the preexistence of buoyant
Archean (and/or Paleoproterozoic) subcontinental lithospheric mantle for their formation and subsequent preservation. A concentration of deposits in the Late Archean and late
Paleoproterozoic to early Mesoproterozoic appears to have
formed near craton margins during alkaline magmatism derived from previously metasomatized mantle lithosphere.
The intrusion-related gold deposits are rare because they
require the conjunction of near-craton setting and shelf
0361-0128/98/000/000-00 $6.00

sedimentary sequences to provide the correct mantle and


crustal ingredients for generation of the causitive granites.
Such settings were unlikely in the early Precambrian, and
younger intrusion-related gold deposits that formed near the
margins of cratons would have had only limited chances of
preservation.
A summary diagram (Fig. 13) shows the relative locations
of deposit types discussed in the text in terms of their tectonic setting at the time of formation and where they might
have been preserved in the context of underlying subcontinental lithospheric mantle. It is concluded that the major
factor affecting the temporal distribution of many goldbearing deposits, in terms of evolution of tectonic processes,
is the nature of subcontinental lithospheric mantle formed
at different times in Earth history. Archean (and/or Paleoproterozoic)-style lithosphere favored the formation of
IOCG deposits and hence their temporal distribution. The
switchover from plume-influenced buoyant plate tectonics
to modern-style plate tectonics, with the shift from buoyant
to negatively buoyant subcontinental lithospheric mantle,
strongly influenced the patterns of preservation of other deposits, for example, orogenic gold and VHMS deposits, examples of which started to form before the widespread occurrence of Archean cratons. The intrusion-related gold
deposits may represent a special case in that they require
near-craton settings, but form outside the cratons in negatively buoyant lithosphere, and hence are rare and largely
restricted to the Phanerozoic.
The temporal distribution of most gold-bearing deposits
discussed here reflects the first-order evolution from mantleplumeinfluenced plate tectonics to a modern style of plate
tectonics in a cooling Earth. Coupled crustal growth and
preservation in the Archean and Paleoproterozoic evolved to
decoupled episodes of growth and preservation from the
Mesoproterozoic onward as a result of irreversible changes to
the subcontinental lithospheric mantle with time.
Future Research and Exploration Significance
In the past few years, the global mining and exploration industry has changed dramatically, with amalgamation into
larger global companies and consequent loss of explorationfocused medium-sized companies. The large companies have
become more risk averse at the very time when near-surface
targets in relatively well exposed terranes are nearing exhaustion, particularly in mature mineral provinces. Future significant discoveries will have to be made at depth in covered terranes, perhaps in remote locations, posing a daunting
challenge to the industry. Clearly, conceptual targeting will be
required and, in turn, will necessitate even greater integration
of theoretical and empirical geoscience into mineral exploration (e.g., McCuaig and Hronsky, 2000). In a risk-averse environment, relatively small, but potentially highly productive,
segments of the globe will have to be selected for intensive
conceptually based exploration, rather than continuation of
the presently less focused exploration effort, often highly reliant on near-surface geochemical anomalies, particularly in
the case of gold exploration.
Over the past few decades, economic geology research has
become very deposit centric and forensic, with much of the
research seeking to understand ore genesis through the use of

218

219

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION

A)

BACK ARC
OCEANIC ARC

CONTINENT

CONTINENTAL
CRATO N
ARC
MARGI N
BACK-ARC

ACCRETED
TERRANES
VHMS
Cu-Au

Epithermal
Au-Ag
Porphyry Cu-Au
(skarns)

Epithermal
x Carlin-style Au
Au
Porphyry Cu-AuIntrusionMo (skarns )
related Au

Orogenic
Au

FORELAND BASIN

*
*

Paleoplacer Au
(Witwatersrand)

Accretionary wedge

Continental crust

Older craton

Extensional fault

Granitoids

Oceanic crust

Subcrustal
lithosphere

Compressional
fault/thrust
Deformed shelf sequence

Asthenosphere

B)
- 20

Young negatively
buoyant SCLM
- moderate uplift

- 10
VHM S

OGD

Old negatively
buoyant SCLM
- major uplift
Minor
IRGD OGD

10
20
30

km

200

Moderately buoyant
SCLM - minor uplift

Buoyant SCLM
- minor uplift
OGD

IOCG

CONTINENTAL CRUST
NEOPROTEROZOIC
LITHOSPHERE
PHANEROZOIC
LITHOSPHERE

IOCG

Shelf sequences

Volcanic rocks

Conglomerates

Metamorphic belts

Sedimentary rocks

Fold belts

VHM S

OGD

Witwatersrand - type
Paleoplacer

ARCHEAN
LITHOSPHERE

100
150

PALEOPROTEROZOIC
LITHOSPHERE

FIG. 13. Schematic lithosphere-scale sections showing (A) the formational environments of gold-bearing deposit types discussed in the text (modified from Groves et al., 1998, with palaeoplacer, intrusion-related, and Carlin-style gold deposits
added), and (B) the environments of preservation of the same deposit types. Note that only the spatial positions of the environments are shown, whereas it is evident that the deposit types formed and/or were preserved in those environments at different times in the evolutionary history of the hosting terranes. Both sections are of necessity generalized and simplified to
include all environments and deposit types. IOCG = iron-oxide copper-gold deposits, IRGD = intrusion-related gold
deposits, OGD = orogenic gold deposits, VHMS = volcanic-hosted massive sulfide deposits.

sophisticated mineralogical, geochemical, isotopic, and fluid


inclusion techniques. This has resulted in a much better
knowledge of ore fluids and their sources, metal transport and
deposition, and deposit-forming processes. However, these
studies mostly help understand the mechanism of ore deposit
formation (the how) but not the specific location of the
0361-0128/98/000/000-00 $6.00

deposit (the where). Thus, very few of the researched parameters can be used directly to select specific terranes within
specific segments of the Earth that can be intensively
explored for world-class to giant mineral deposits. This is confirmed by research studies which demonstrate that giant hydrothermal deposits form essentially from the same fluids and

219

220

GROVES ET AL.

by similar processes as smaller deposits (see, e.g., papers in


Whiting et al., 1993, and Cooke and Pongratz, 2002). It appears much more likely that it is the conjunction of provincescale characteristics of a terrane, rather than deposit-scale parameters, that dictate whether a giant deposit will be present
or not, as summarized, for example, by Groves et al. (2003)
for gold deposits in metamorphic belts and noted by Richards
(2003) for porphyry Cu deposits.
In order to understand these province-scale controls on
world-class to giant deposits, and utilize this in predictive
mineral discovery under cover, it will be necessary to understand the four-dimensional evolution of potentially prospective terranes globally. To achieve this will not only require
government or multiclient state of the art remote sensing and
airborne geophysical databases, such have become routinely
available in many mature exploration terranes globally (e.g.,
Australia, Canada, southern Africa, southwestern United
States), but integrated research at the global to province
scale. For example, there needs to be more research on the
specific tectonic settings of mineral deposits. This includes
knowledge of the evolution of the geometries of plates, subduction slabs and transform faults in convergent margin settings, and their structural and petrogenetic signals in the rock
record of ancient terranes. The role that mantle plumes play
in modifying tectonic processes and regimes, and in accentuating global metallogenic epochs, also needs to be better understood, particularly in the early Precambrian when they
may have been the dominant control. There is also a need to
understand the nature and thickness of subcontinental lithospheric mantle beneath prospective terranes at different times
during their evolution, for example by using a combination of
seismic tomology with petrology, mineralogy and age constraints on mantle xenoliths and inclusions in diamonds (e.g.,
Shirey et al., 2003). This will not only provide an indication of
the prospectivity of the terranes at different periods in Earth
history but also provide strong indications of the likelihood of
their preservation.
For most mineral deposit types (e.g., Meyer, 1988; Barley
and Groves, 1992; this study), there are particular times in
Earth history when world-class to giant deposits formed and
were preserved at a global scale. Thus, not only do the tectonic settings and lithospheric structures of potentially
prospective terranes need to be understood, but the ages of
their component units and metamorphic, deformational, and
intrusive events need to be established by robust geochronology. In concert with this, there needs to be more emphasis on
accurate dating of mineral deposits, using robust geochronological methods on clearly ore related minerals, to better define the critical age peaks of major mineral deposit types. Finally, in order to trace prospective, but now dispersed,
Paleozoic to Precambrian terranes across the globe, there will
be need for more research on the paleogeographic relationships between terranes, particularly the timing and geometrical arrangement of supercontinent assembly, using appropriate geochronology and paleomagnetic methods, as times of
assembly and breakup appear particularly significant in global
metallogeny. It will also be important to link these data to the
patterns of global mantle convection.
In summary, in order to meet changing exploration requirements, there is a need to change the emphasis of the
0361-0128/98/000/000-00 $6.00

scale of economic geology research with a shift to integrate


current deposit-centric and forensic research into more multidisciplinary studies that emphasize global tectonics and
metallogeny. This should bring attendant advances in conceptual targeting that will lead to world-class to giant discoveries
to satisfy the resource demands of the next generation.
Acknowledgments
This paper was inspired by the pioneering academic concepts of Chuck Meyer and the global exploration vision of
Roy Woodall. We are grateful to colleagues at the Centre
for Global Metallogeny, University of Western Australia,
particularly Noreen Vielreicher, the U.S. Geological Survey, and Centre for Ore Deposit and Exploration Studies
(CODES), University of Tasmania, particularly Mike
Solomon, for useful discussions on this topic. The paper
was improved by the useful reviews of Dallas Abbott, Phil
Brown, Rob Kerrich, Steve Kesler, and Henry Pollack, and
incisive editorial comments by Steve Kesler and Mark Hannington. This paper is a contribution to the Centre for
Global Metallogeny and Tectonics Special Research Centre
publication 304.
August 30, October 25, 2004
REFERENCES
Abbott, D.H., and Isley, A.E., 2002, The intensity, occurrence and duration
of superplume events and eras over geological time: Journal of Geodynamics, v. 34, p. 265307.
Arndt, N.T., Lewin, E., and Albarde, F., 2002, Strange partners: Formation
and survival of continental crust and lithospheric mantle: Geological Society of London Special Publication, v. 199, p. 91103.
Aspler, L.B., and Chiarenzelli, J.R., 1998, Two Neoarchean supercontinents?
Evidence from the Paleoproterozoic: Sedimentary Geology, v. 120, p.
75104.
Ayer, J., Amelin, Y., Corfu, F., Kamo, J., Ketchum, J., Kwok, K., and Trowell,
N., 2002, Evolution of the southern Abitibi greenstone belt based on U-Pb
geochronology: Autochtonous volcanic construction followed by plutonism,
regional deformation and sedimentation: Precambrian Research, v. 115, p.
6395.
Bache, J.J., 1987, World gold deposits: A geological classification: New York,
Elsevier, 178 p.
Balakrishnan, S., Rajamani, V., and Hanson, G.N., 1999, U-Pb ages for zircon
and titanite from the Ramagiri area, southern India: Evidence for accretionary origin of the eastern Dharwar craton during the late Archean: Journal of Geology, v. 107, p. 6986.
Barley, M.E., and Groves, D.I., 1992, Supercontinent cycles and the distribution of metal deposits through time: Geology, v. 20, p. 291294.
Barley, M.E., Krapez, B., Groves, D.I., and Kerrich, R., 1998, The Late Archaean bonanza: Metallogenic and environmental consequences of the interaction between mantle plumes, lithospheric tectonics and global cyclicity: Precambrian Research, v. 91, p. 6590.
Barley, M.E., Pickard, A.L., Hagemann, S.G., and Folkert, S.L., 1999, Hydrothermal origin for the 2 billion year old Mount Tom Price giant iron ore
deposit, Hamersley province, Western Australia: Mineralium Deposita, v.
34, p. 784789.
Barley, M.E., Rak, P., and Wyman, D., 2002, Tectonic controls on magmatichydrothermal gold mineralization in the magmatic arcs of SE Asia: Geological Society Special Publication, v. 204, p. 3947.
Barnicoat, A.C., Henderson, I.H.C., Knipe, R.J., Yardley, B.W.D., Napier,
R.W., Fox, N.P.C., Kenyon, A.K., Muntingh, D.J., Strydom, D., Winkler,
K.S., Lawrence, S.R., and Cornford, C., 1997, Hydrothermal gold mineralization in the Witwatersrand basin: Nature, v. 386, p. 820824.
Barrie, C.T., and Hannington, M.D., 1999, Classification of volcanic-associated massive sulfide deposits based on host-rock composition: Reviews in
Economic Geology, v. 8, p. 111.
Barton, J.M., Jr., 1984, Timing of ore emplacement and deformation,
Murchison and Sutherland greenstone belts, Kaapvaal craton: Geological
Society of Zimbabwe Special Publication, v. 1, p. 629644.

220

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION


Beattie, M., and Barton, J.M., Jr., 1992, Alteration and occurrence of gold
mineralization in the Roodepoort goldfield, Pietersburg granite-greenstone
terrane: South African Journal of Geology, v. 95, p. 131140.
Betts, P.G., Giles, D., Lister, G.S., and Frick, L.R., 2002, Evolution of the
Australian lithosphere: Australian Journal of Earth Sciences, v. 49, p.
661695.
Bickle, M.J., 1986, Implications of melting for stabilisation of lithosphere and
heat loss in the Archaean: Earth and Planetary Science Letters, v. 80, p.
314324.
1990, Mantle evolution, in Hall, R.P., and Hughes, D.J., eds., Early Precambrian basic magmatism: Glasgow, Blackie and Sons, p. 111135.
Blundell, D.J., 2002, The timing and location of major ore deposits in an
evolving orogen: The geodynamic context: Geological Society Special Publication, v. 204, p. 112.
Boyle, R.W., 1979, The chemistry of gold and its deposits: Geological Survey
of Canada, Bulletin 280, 584 p.
Brimhall, G.H, Jr., 1987, Metallogenesis: Reviews in Geophysics, v. 25, p.
10791088.
Campbell, I.H., Griffith, R.W., and Hill, R.I., 1989, Melting in an Archaean
mantle plume: Heads its basalt tails its komatiites: Nature, v. 339, p. 697699.
Campbell, I.H., Compston, D.M., Richards, I.P., Johnson, J.P., and Kent,
A.R., 1998, Reviews of the application of isotopic studies to the genesis of
Cu-Au mineralization at Olympic Dam and Au mineralization at Porgera,
Tennant Creek district and Yilgarn craton: Australian Journal of Earth Sciences, v. 45, p. 201218.
Champion, D.C., and Sheraton, J.W., 1997, Geochemistry and Nd isotope
systematics of Archaean granites of the Eastern goldfields, Yilgarn craton,
Australia: Implications for crustal growth processes: Precambrian Research, v. 83, p. 109132.
Chiaradia, M., and Fontbot, L., 2001, Mineralogy, lead isotope and metal
geochemistry of gold-rich Tertiary VHMS deposits of the Western
Cordillera of Ecuador: Society for Geology Applied to Mineral Deposits International Meeting, 6th Biennial, Proceedings, p. 269272.
Colvine, A.C., Andrews, A.J., Cherry, M.E., Durocher, M.E., Fyon, J.A., Lavigne, M.J., Macdonald, A.J., Marmont, S., Poulsen, K.H., Springer, J.S., and
Troop, D.G., 1984, An integrated model for the origin of Archean lode-gold
deposits: Ontario Geological Survey Open-File Report, v. 5524, 98 p.
Condie, K.C., 1998, Episodic continental growth and supercontinents: A
mantle avalanche connection?: Earth and Planetary Science Letters, v. 163,
p. 97108.
2000, Episodic continental growth models: Afterthoughts and extensions: Tectonophysics, v. 322, p. 153162.
2001a, Rodinia and continental growth: Gondwana Research, v. 4, p.
154155.
2001b, Mantle plumes and their record in Earth history: Cambridge,
Cambridge University Press, 305 p.
2002a, The supercontinent cycle: Are there two patterns of cyclicity?:
Journal of African Earth Sciences, v. 35, p. 179183.
2002b, Continental growth during a 1.9-Ga superplume event: Journal
of Geodynamics, v. 34, p. 249264.
2002c, Breakup of a Paleoproterozoic supercontinent: Gondwana Research, v. 5, p. 4143.
2004, Supercontinents and superplume events: Distinguishing signals in
the geologic record: Physics of the Earth and Planetary Interiors, v. 146, p.
319332.
Cooke, D.R., and Pongratz, J., eds., 2002, Giant ore deposits: Characteristics,
genesis and exploration: Centre for Ore Deposit and Exploration Studies
(CODES), University of Tasmanis Special Publication, v. 4, 269 p.
Corlett, G.J., and Akiro, K., 1999, VHMS mineralisation in NW Papua New
Guinea, in Weber G., ed., Proceedings PACRIM 99 Congress, Australasian
Institute of Mining and Metallurgy, v. 4/99, p. 243245.
Dahl, P.S., Holm, D.K., Gardner, E.T., Habacher, F.A., and Foland, K.A.,
1999, New constraints on the timing of Early Proterozoic tectonism in the
Black Hills (South Dakota), with implications for docking of the Wyoming
province with Laurentia: Geological Society of America Bulletin, v. 111, p.
13351349.
Dalziel, I.W.D., 1991, Pacific margin of Lauirentia and East Antarctica
Australia as a conjugate rift pair. Evidence and implications for an Eocambrian supercontinent: Geology, v. 19, p. 598601.
Davies, G.F., 1992, On the emergence of plate tectonics: Geology, v. 20, p.
963966.
deRonde, C.E.J., Kamo, S., Davis, D.W., deWit, M.J., and Spooner, E.T.C.,
1991, Field, geochemical and U-Pb isotopic constraints from hypabyssal
0361-0128/98/000/000-00 $6.00

221

felsic intrusions within the Barberton greenstone belt, South Africa: Implications for tectonics and the timing of gold mineralization: Precambrian
Research, v. 49, p. 261280.
deWit, M.J., 1998, On Archean granites, greenstones, cratons and tectonics:
Does the evidence demand a verdict?: Precambrian Research, v. 91, p.
181226.
deWit, M.J., and Thiart, C., 2003, Metallogenic scents of Archaean cratons:
Changing patterns of mineralization during earth evolution: Transactions of
the Institution of Mining and Metallurgy, v. 112, p. B114B116.
Edwards, R., and Atkinson, K., 1986, Ore deposit geology and its influence
on mineral exploration: London, New York, Chapman and Hall, p.
175214.
Ernst, R.E., and Buchan, K.L., 2002, Maximum size and distribution in time
and space of mantle plumes: Evidence from large igneous provinces: Journal of Geodynamics, v. 34, p. 309342.
Farquhar, J., Bao, H., and Thiemens, M., 2000, Atmospheric influence of
Earths earliest sulphur cycle: Science, v. 289, p. 756759.
Feather, C.E., and Koen, G.M., 1975, The mineralogy of the Witwatersrand
reef: Minerals in Science and Engineering, v. 7, p. 189224.
Franklin, J.M., Lydon, J.W., and Sangster, D.F., 1981, Volcanic-associated
massive sulfide deposits: ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME,
p. 485627.
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 worlds largest gold province: ECONOMIC GEOLOGY 100TH ANNIVERSARY VOLUME, in press.
Fyfe, W.S., 1978, The evolution of Earths crust: Modern plate tectonics to
ancient hot spot tectonics?: Chemical Geology, v. 23, p. 89114.
Fyfe, W.S., and Kerrich, R., 1985, Fluids and thrusting: Chemical Geology, v.
49, p. 353362.
Gastil, G., 1960, The distribution of mineral dates in time and space: American Journal of Science, v. 258, p. 135.
Goldfarb, R.J., 1997, Metallogenic evolution of Alaska: ECONOMIC GEOLOGY
MONOGRAPH 8, p. 434.
Goldfarb, R.J., Phillips, G.N., and Nockleberg, W.J., 1998, Tectonic setting of
synorogenic gold deposits of the Pacific Rim: Ore Geology Reviews, v. 13,
p. 185218.
Goldfarb, R.J., Groves, D.I., and Gardoll, S., 2001a, Orogenic gold and geologic time: A global synthesis: Ore Geology Reviews, v. 18, p. 175.
2001b, Rotund versus skinny orogens: Well-nourished or malnourished
gold?: Geology, v. 29, p. 539542.
Grainger, C.J., Groves, D.I., and Costa, C.H.C., 2002, The epigenetic sediment-hosted Serra Pelada Au-PGE deposit and its potential genetic association with Fe-oxide Cu-Au mineralization within the Carajas mineral
province, Amazon craton, Brazil: Society of Economic Geologists Special
Publication 9, p. 4764.
Griffin, W.L., Zhang, A., OReilly, S.Y., and Ryan, C.G., 1998, Phanerozoic
evolution of the lithosphere beneath the Sino-Korean craton, in Flower,
M.F.J., Chung, S.L., Lo, C.H., and Lee, T.Y., eds., Mantle dynamics and
plate interactions in East Asia: Washington DC, American Geophysical
Union Geodynamics Series, p. 107126.
Griffin, W.L., OReilly, S.Y., Abe, N., Aulbach, S., Davies, R.M., Pearson,
N.J., Doyle, B.J., and Kivi, K., 2003, The origin and evolution of Archean
lithospheric mantle: Precambrian Research, v. 127, p. 1941.
Groves, D.I., 1993, The crustal continuum model for late-Archaean lodegold deposits of the Yilgarn block, Western Australia: Mineralium Deposita, v. 28, p. 366374.
Groves, D.I., and Vielreicher, N.M., 2001, The Phalabowra (Palabora) carbonatite-hosted magnetite-copper sulphide deposit, South Africa: An endmember of the iron oxide copper-gold-rare earth element group?: Mineralium Deposita, v. 36, p. 189194.
Groves, D.I., Goldfarb, R.J., Gebre-Mariam, M., Hagemann, S.G., and
Robert, F., 1998, Orogenic gold depositsa proposed classification in the
context of their crustal distribution and relationship to other gold deposit
types: Ore Geology Reviews, v. 13, p. 727.
Groves, D.I., Goldfarb, R.J., Robert, F., and Hart, C.J.R., 2003, Gold deposits in metamorphic belts: Overview of current understanding, outstanding problems, future research, and exploration significance: ECONOMIC GEOLOGY, v. 98, p. 129.
Groves, D.I., Grainger, C.J., and Tallarico, F.H.B., 2004, Repeated alkaline
granitoid magmatism and formation of Fe-oxide Cu-Au and related deposits in the giant Carajas mineral province of Brazil [abs.]: Geological Society of Australia Abstracts, v. 73, p. 83.

221

222

GROVES ET AL.

Haeberlin, Y., Moritz, R., and Fontbot, L., 2003, Paleozoic orogenic gold
deposits in the eastern Central Andes and its foreland, South America: Ore
Geology Reviews, v. 22, p. 4159.
Haeussler, P.J., Bradley, D., Goldfarb, R.J., Snee, L.W., and Taylor, C.D.,
1995, Link between ridge subduction and gold mineralization in southern
Alaska: Geology, v. 23, p. 995998.
Hamilton, J.V., and Hodgson, C.J., 1986, Mineralization and structure of the
Kolar gold field, India, in Macdonald, A.J., ed., Gold 86; An international
symposium on the geology of gold deposits, Proceedings volume: Willowdale, Ontario, Konsult International, p. 270283.
Hamilton, W.B., 1998, Archean magmatism and deformation were not products of plate tectonics: Precambrian Research, v. 91, p. 143179.
Hannington, M.D., Poulsen, K.H., Thompson, J.F.H., and Sillitoe, R.H.,
1999, Volcanogenic gold in the massive sulfide environment: Reviews in
Economic Geology, v. 14, p. 319350.
Hart, T.R., Gibson, H.L., and Lesher, C.M., 2004, Trace element geochemistry and petrogenesis of felsic volcanic rocks associated with volcanogenic
massive Cu-Zn-Pb sulfide deposits: ECONOMIC GEOLOGY, v. 99, p.
10031014.
Haynes, D.W., 2002, Giant iron oxide-copper-gold deposits: Are they in
distinctive geological settings: Centre for Ore Deposit and Exploration
Studies (CODES), University of Tasmania Special Publication, v. 4, p.
5777.
Henley, R.W., and Adams, J., 1979, On the evolution of giant gold placers:
Transactions of the Institution for Mining and Metallurgy, v. 88, p.
B41B51.
Hitzman, M.W., 2000, Iron oxide Cu-Au deposits: What, where, when, and
why, in Porter, T.M., ed., Hydrothermal iron oxide copper-gold and related
deposits: A global perspective: Adelaide, Australian Mineral Foundation, p.
925.
2001, Fe oxide-Cu-Au systems in the Lufilian orogen of southern
Africa [abs.]: Geological Society of America Abstracts with Program, v.
33, p.1.
Hitzman, M.W., Oreskes, N., and Einaudi, M.T., 1992, Geological characteristics and tectonic setting of Proterozoic iron-oxide (Cu-U-Au-REE) deposits: Precambrian Research, v. 58, p. 241287.
Hodgson, C.J., and MacGeehan, P.J., 1982, A review of the geological characteristics of gold only deposits in the Superior province of the Canadian
Shield: Canadian Institute of Mining and Metallurgy Special Volume 24, p.
211229.
Hoffman, P.E., 1988, The United Plates of America, the birth of a craton: Annual Review of Earth and Planetary Sciences, v. 16, p. 543604.
1989, Speculations on Laurentias first gigayear (2.01.0 Ga): Geology, v.
17, p. 135138.
1991, Did the breakout of Laurentia turn Gondwanaland inside-out?:
Science, v. 252, p. 14091414.
Holland, H.D., 1984, Chemical evolution of the atmosphere and oceans:
Princeton, New Jersey, Princeton University Press, 582 p.
Hughes, M.J., Phillips, G.N., and Carey, S.P., 2004, Giant placers of the Victorian gold province: Society of Economic Geology Newsletter 56, p. 1 and
1117.
Huston, D.L., 2000, Gold in volcanic-hosted massive sulfide deposits: Distribution, genesis and exploration: Reviews in Economic Geology, v. 13, p.
401426.
Hutchinson, R.W., 1993, Some broad processes and affects of evolutionary
metallogeny: Resource Geology Special Issue, v. 15, p. 4554.
Isley, A.E., 1995, Hydrothermal plumes and the delivery of iron to BIF: Journal of Geology, v. 103, p. 169185.
Isley, A.E., and Abbott, D.H., 1999, Plume-related mafic volcanism and the
deposition of banded iron formation: Journal of Geophysical Research, v.
104, p. 1546115477.
2002, Implications of the temporal distribution of high-Mg magmas for
mantle plume volcanism through time: Journal of Geology, v. 110, p.
141158.
Jamieson, R.A., Beaumont, C., Fullshock, P., and Lee, B., 1998, Barrovian regional metamorphism: Wheres the heat?: Geological Society of London
Special Publications, v. 138, p. 2351.
Jordan, T.H., 1988, Composition and development of the continental
tectonosphere: Nature, v. 274, p. 544548.
Kerrich, R., 1999, Natures gold factory: Science, v. 284, p. 21012102.
Kerrich, R., and Cassidy, K.F., 1994, Temporal relationships of lode gold
mineralization to accretion, magmatism and deformationArchean to presenta review: Ore Geology Reviews, v. 9, p. 263310.
0361-0128/98/000/000-00 $6.00

Kerrich, R., Goldfarb, R.J., Groves, D.I., and Garwin, S., 2000, The geodynamics of world class gold deposits: Characteristics, space-time distribution, and origins: Reviews in Economic Geology, v. 13, p. 501551.
Kesler, S.E., 1997, Metallogenic evolution of convergent margins: Selected
ore deposit models: Ore Geology Reviews, v. 12, p. 153171.
Kesler, S.E., Chryssoulis, S.L., and Simon, G., 2002, Gold in porphyry copper
deposits: Its abundance and fate: Ore Geology Reviews, v. 21, p. 103124.
Kirk, J., Ruiz, J., Chesley, J., Titley, S., and Walshe, J., 2001, A detrital model
for the origin of gold and sulfides in the Witwatersrand basin based on ReOs isotopes: Geochimica et Cosmochimica Acta, v. 65, p. 21492159.
Kirk, J., Ruiz, J., Chesley, J., Walshe, J., and England, G., 2002, A major
Archean gold and crust-forming event in the Kaapvaal craton, South Africa:
Science, v. 297, p. 18561858.
Klein, C., and Beukes, N.J., 1992, Proterozoic iron formations in Condie,
K.C., ed., Proterozoic crustal evolution: Amsterdam, Elsevier, p. 139146.
Knight, J., Lowe, J., Joy, S., Cameron, J., Merrillees, J., Nag, S., Shah, N.,
Dua, G., and Jhala, K., 2002, The Kheti copper belt, Rajasthan: Iron oxide
copper-gold terrane in the Proterozoic of NW India, in Porter, T.M., ed.,
Hydrothermal iron oxide copper-gold and related deposits: A global perspective, v. 2: Adelaide, PGC Publishing, p. 321341.
Konstantinov, M., Dankovtsev, R., Simkin, G., and Cherkasov, S., 1999, Deep
structure of the north Enisei gold district (Russia) and setting of ore deposits: Geology of Ore Deposits, v. 41, p. 425436.
Kositcin, N., and Krapez, B., 2004, Relationship between detrital zircon agespectra and the tectonic evolution of the Late Archaean Witwatersrand
basin, South Africa: Precambrian Research, v. 129, p. 141168.
Krapez, B., Brown, S.J.A., Hand, J., Barley, M.E., and Cas, R.A.F., 2000, Age
constraints on recycled crustal and supracrustal sources of Archaean
metasedimentary sequences, Eastern Goldfields province, Western Australia: Evidence from SHRIMP zircon dating: Tectonophysics, v. 322, p.
89133.
Landefeld, L.A., 1988, The geology of the Mother Lode gold belt, Sierra
Nevada Foothills metamorphic belt, California [ext. abs.], in Goode,
A.D.T., and Bosma, L.I., eds., Bicentennial Gold 88, Extended Abstracts:
Melbourne, Geological Society of Australia, p. 167172.
Lang, J.R., and Baker, T., 2001, Intrusion-related gold systems: The present
level of understanding: Mineralium Deposita, v. 36, p 477489.
Lang, J.R., Baker, T., Hart, C.J.R., and Mortensen, J.K., 2000, An exploration
model for intrusion-related gold systems: Society of Economic Geology
Newsletter 40, p. 1, 615.
Large, R.R., 1992, Australian volcanic-hosted massive sulfide deposits: Features, styles, and genetic models: ECONOMIC GEOLOGY, v. 87, p. 471510.
Larson, R.L., 1991, Geological consequences of superplumes: Geology, v. 19,
p. 963-966.
Lasaga, A.C., and Ohmoto, H., 2002, The oxygen geochemical cycle: Dynamics and stability: Geochimica et Cosmochimica Acta, v. 66, p. 361381.
Lesher, C.M., 1989, Komatiite-associated nickel sulfide deposits: Reviews in
Economic Geology, v. 4, p. 45101.
Loen, J. S., 1992, Mass balance constraints on gold placers: Possible solutions
to source area problems: ECONOMIC GEOLOGY, v. 87, p. 16241634.
Lowman, J.P., and Jarvis, G.T., 1996, Continental collisions in wide aspect
ratio and high Rayleigh number two-dimensional mantle convection models: Journal of Geophysical Research, v. 101, p. 2548525497.
Ludden, J.N, and Hynes, A., 2000, The lithoprobe Abitibi-Grenville transect;
two billion years of crust formation and recycling in the Precambrian Shield
of Canada: Canadian Journal of Earth Sciences, v. 37, p. 459476.
Mair, J.L., Hart, C.J.R., Groves, D.I., and Goldfarb, R.J., 2003, The nature of
Tombstone Plutonic Suite rocks at Scheelite Dome, Tintina gold province:
Evidence for an enriched mantle contribution, in Belvin, P., Jones, M., and
Chappell, B.W., eds., Magmas to Mineralizationthe Ishihara Symposium:
Granites and Associated Metallogenesis: National Key Centre for Geochemical and Metallogenic Evolution of Continents (GEMOC), Sydney,
New South Wales, Australia, Macquarie University, p. 9396.
Marschik, R., and Fontbot, L., 2001, The Candelaria-Punta del Cobre iron
oxide Cu-Au (-Zn-Ag) deposits, Chile: ECONOMIC GEOLOGY, v. 96, p.
17991826.
Marschik, R., Chiarada, M., and Fontbot, L., 2003a, Implications of Pb isotope signatures of rocks and iron oxide Cu-Au ores in the Candelaria-Punta
del Cobre district, Chile: Mineralium Deposita, v. 38, p. 900912.
Marschik, R., Fontagnie, D., Chiarada, M., and Voldet, P., 2003b, Geochemical and Nd-Sr-Pb-O isotope characteristics of granitoids of the Early Cretaceous Copiapo plutonic complex (2730S), Chile: Journal of South
American Earth Sciences, v. 16, p. 381398.

222

100th ANNIVERSARY SPECIAL PAPER: TECTONIC PROCESSES & Au DISTRIBUTION


Mathur, R., Marschik, R., Ruiz, J., Munizaga, F., Leveille, R.A., and Martin,
W., 2002, Age of mineralization of the Candelaria Fe oxide Cu-Au deposit
and the origin of the Chilean iron belt, based on Re-Os isotopes: ECONOMIC GEOLOGY, v. 97, p. 5971.
McCandless, T.E., Mathur, R.D., Ruiz, J., and Kontinen, A., 1999, Re-Os
isotope dating of Precambrian volcanogenic massive sulfide (VMS) deposits [abs.]: Geological Society of America Abstracts with Programs, v.
31, p. 30.
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. 553559.
Meyer, C., 1981, Ore-forming processes in geologic history: ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 641.
1988, Ore deposits as guides to geologic history of the Earth: Annual Review of Earth and Planetary Sciences, v. 16, p. 147171.
Milesi, J.P., Ledru, P., Marcoux, E., Mougeot, R., Johan, V., Lerouge, C.,
Sabate, P., Bailly, L., Respaut, J.P., and Skipwith, P., 2002, The Jacobina Paleoproterozoic gold-bearing conglomerates, Bahia, Brazil: A hydrothermal
shear-reservoir model: Ore Geology Reviews, v. 19, p. 95136.
Minter, W.E.L., 1999, Irrefutable detrital origin of Witwatersrand gold and
evidence of eolian signatures: ECONOMIC GEOLOGY, v. 94, p. 665670.
Minter, W. E. L., Goedhart, M. L., Knight, J., and Frimmel, H. E., 1993,
Morphology of Witwatersrand gold grains from the Basal reef: Evidence
for their detrital origin: ECONOMIC GEOLOGY, v. 88, p. 237248.
Mitchell, A.H.G., and Garson, M.S., 1981, Mineral deposits and global tectonic settings: London, Academic Press Geology Series, 405 p.
Murphy, J.B., and Nance, R.D., 1992, Mountain belts and the supercontinent
cycle: Scientific American, v. 266, p. 8491.
Nisbet, B., Cooke, J., Richards, M., and Williams, C., 2000, Exploration for
iron oxide copper gold deposits in Zambia and Sweden: Comparison with
the Australian experience, in Porter, T.M., ed., Hydrothermal iron oxide
copper-gold and related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 297308.
OReilly, S.Y., and Griffin, W.L., 1996, 4-D lithospheric mapping: A review of
the methodology with examples: Tectonophysics, v. 262, p. 318.
Parsons, B.A., 1982, Causes and consequences of the relationship between
area and age of the ocean floor: Journal of Geophysical Research, v. 87, p.
289302.
Peltier, W.R., Butler, S., and Solheim, L.P., 1997, The influence of phase
transformations on mantle mixing and plate tectonics, in Crossley, D.J., ed.,
Earths deep interior: Amsterdam, Gordon and Breach, p. 405430.
Perell, J., Cox, D., Garamjav, D., Sanidorj, S., Diakov, S., Schissel, D.,
Munkhbat, T.O., and Oyun, G., 2001, Oyu Tolgoi, Mongolia: Siluro-Devonian porphyry Cu-Au-(Mo) and high-sulfidation Cu mineralization
with a Cretaceous chalcocite blanket: ECONOMIC GEOLOGY, v. 96, p.
14071428.
Pesonen, L.J., Elming, S.-., Mertanen, S., Pisarevsky, S., DAgrella-Filho,
M.S., Meert, J.G, Schmidt, P.W., Abrahamsen, N., and Bylundet, G., 2003,
Paleomagnetic configuration of continents during the Proterozoic:
Tectonophysics, v. 375, p. 289324.
Phillips, G. N., and Myers, R. E., 1989, Witwatersrand gold fields. Part II. An
origin for Witwatersrand gold during metamorphism and associated alteration: ECONOMIC GEOLOGY MONOGRAPH 6, p. 598608.
Pigois, J.-P., Groves, D. L., Fletcher, I. R., McNaughton, N. J., and Snee, L.
W., 2003, Age constraints on Tarkwaian palaeoplacer and lode-gold formation in the Tarkwa-Damang district, SW Ghana: Mineralium Deposita, v.
38, p. 695714.
Pirajno, F., 2000, Ore deposits and mantle plumes: Dordrecht, Kluwer Academic Publishers, 556 p.
Pisarevsky, S.A., Wingate, M.T.D., Powell, C.McA., Johnson, S., and Evans,
D.A.D., 2003, Models of Rodinia assembly and fragmentation: Geological
Society of London Special Publications, v. 206, p. 3555.
Pollack, H.N., 1986, Cratonization and thermal evolution of the mantle:
Earth and Planetary Science Letters, v. 80, p. 175182.
1997, Thermal characteristics of the Archaean, in deWit, M.J., and Ashwal, L.D., eds., Greenstone belts: Oxford, Claredon Press, p. 223232.
Poudjom Djomani, Y.H., OReilly, S.Y., Griffin, W.L., and Morgan, P., 2001,
The density structure of subcontinental lithosphere through time: Earth
and Planetary Science Letters, v. 184, p. 605621.
Poujol, M., Robb, L.J., Respaut, J.P., and Anhaeusser, C.R., 1996, 3.072.97
Ga greenstone belt formation in the northeastern Kaapvaal craton: Implications for the origin of the Witwatersrand basin: ECONOMIC GEOLOGY, v.
91, p. 14551461.
0361-0128/98/000/000-00 $6.00

223

Qiu, Y., and Groves, D.I., 1999, Late Archean collision and delamination in
the southwest Yilgarn craton: The driving force for Archean orogenic lode
gold mineralization: ECONOMIC GEOLOGY, v. 94, p. 115122.
Requia, K., Stein, H., Fontbot, L., and Chiaradia, M., 2003, Re-Os and PbPb geochronology of the Archean Salobo iron oxide copper-gold deposit,
Carajas mineral province, northern Brazil: Mineralium Deposita, v. 38, p.
727738.
Richards, J.P., 2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au)
deposit formation: ECONOMIC GEOLOGY, v. 98, p. 15151533.
Richardson, S.H., Shirey, S.B., and Harris, J.W., 2004, Episodic diamond
genesis at Jwaneng, Botswana, and implications for Kaapvaal craton evolution: Lithos, in press.
Richter, F.M., 1985, Models for the Archean thermal regime: Earth and
Planetary Science Letters, v. 73, p. 350360.
Robb, L.J., and Meyer, F.M., 1990, The nature of the Witwatersrand hinterland: Conjectures on the source-area problem: ECONOMIC GEOLOGY, v. 85,
p. 511536.
Robert, F., and Poulsen, K.H., 1997, World-class Archean gold deposits in
Canada: An overview: Australian Journal of Earth Sciences, v. 44, p.
329351.
Rogers, J.J.W., 1996, A history of continents in the past three billion years:
Journal of Geology, v. 104, p. 91107.
Rudnick, R.L., 1995, Making continental crust: Nature, v. 378, p. 571578.
Safonov, Y.G., 1997, Hydrothermal gold depositsdistribution,
geological/genetic types, and productivity of ore-forming systems: Geology
of Ore Deposits, v. 39, p. 2032.
Santos, J.O.S., Groves, D.I., Hartmann, L.A., Mousa, M.A., and McNaughton, N.J., 2001, Gold deposits of the Tapajos and Alta Floresta domains, Tapajos-Parima orogenic belt, Amazon craton, Brazil: Mineralium
Deposita, v. 36, 278299.
Sawkins, F.J., 1972, Sulfide deposits in relation to plate tectonics: Journal of
Geology, v. 80, p. 377397.
1990, Mineral deposits in relation to plate tectonics: Berlin, Springer
Verlag, 461 p.
Schmitz, M.D., Bowring, S.A., deWit, M.J., and Gartz, V., 2004, Subduction
and terrane collision stabilize the western Kaapvaal craton tectosphere 2.9
billion years ago: Earth and Planetary Science Letters, v. 222, p. 363376.
Shirey, S.B., Harris, J.W., Richardson, S.H., Fouch, M., James, D.E., Cartigny, P., Deines, P., and Viljoen, F., 2003, Regional patterns in the petrogenesis and age of inclusions in diamond, diamond composition, and the
lithospheric seismic structure of southern Africa: Lithos, v. 71, p. 243258.
Sillitoe, R.H., 1972, A plate tectonic model for the origin of porphyry copper
deposits: ECONOMIC GEOLOGY, v. 67, p. 184197.
1991, Intrusion-related gold deposits, in Foster, R.P., ed., Metallogeny
and exploration of gold: Glasgow, Blackie and Sons, p. 165209.
1997, Characteristics and controls of the largest porphyry copper-gold
and epithermal gold deposits in the circum-Pacific region: Australian Journal of Earth Sciences, v. 44, p. 373388.
Sillitoe, R.H., and Thompson, J.F.H., 1998, Intrusion-related vein gold deposits: Types, tectono-magmatic settings, and difficulties of distinction
from orogenic gold deposits: Resource Geology, v. 48, p. 237250.
Simonson, B.M., and Hassler, S.W., 1997, Revised correlations in the early
Precambrian Hamersley basin based on a horizon of resedimented impact
spherules: Australian Journal of Earth Sciences, v. 44, p. 112.
Sleep, N.H., 2003, Survival of Archean cratonal lithosphere: Journal of Geophysical Research, v. 108, p.8-18-25.
Sleep, N.H., and Windley, B.F., 1982, Archean plate tectonics: Constraints
and inferences: Journal of Geology, v. 90, p. 363379.
Solomon, M., and Groves, D.I., 1994, The geology and origin of Australias
mineral deposits: Oxford Monographs in Geology and Geophysics, v. 24,
951 p.
Solomon, M., and Quesada, C., 2003, Zn-Pb-Cu massive sulfide deposits:
Brine-pool types occur in collisional orogens, black smoker types occur in
backarc and/or arc basins: Geology, v. 31, p. 10291032.
Spooner, E.T.C., and Barrie, C.T., 1993, A special issue devoted to Abitibi ore
deposits in a modern context: Preface: ECONOMIC GEOLOGY, v. 88, p.
13071322.
Sprague, D., and Pollack, H.N., 1980, Heat flow in the Mesozoic and Cenozopic: Nature, v. 285, p. 393395.
Stein, M., and Hoffmann, A.W., 1994, Mantle plumes and episodic crustal
growth: Nature, v. 372, p. 6368.
Stern, R.A., Machado, N., Syme, E.C., Lucas, S.B., and David, J., 1999,
Chronology of crustal growth and recycling in the Paleoproterozoic Amisk

223

224

GROVES ET AL.

collage (Flin Flon belt), Trans-Hudson orogen, Canada: Canadian Journal


of Earth Sciences, v. 36, p. 18071827.
Syme, E.C., and Bailes, A.H., 1993, Stratigraphic and tectonic setting of early
Proterozoic volcanogenic massive sulfide deposits, Flin Flon, Manitoba:
ECONOMIC GEOLOGY, v. 88, p. 566589.
Tallarico, F.H.B., Figueiredo, B.R., Groves, D.I., Kositcin, N., McNaughton,
N.J., Fletcher, I.R., and Rego, J.L., Geology and SHRIMP U-Pb
geochronology of the Igarap Bahia deposit, Carajs copper-gold belt,
Brazil: An Archean (2.57 Ga) example of iron-oxide Cu-Au-(U-REE) mineralization: ECONOMIC GEOLOGY, v. 100, p. 728.
Taylor, S.R., and McLennan, S.M., 1985, The continental crust: Its composition and evolution: Oxford, Blackwell, 312 p.
Thompson, J.F.H., Sillitoe, R.H., Baker, T., Lang, J.R., and Mortensen, J.K.,
1999, Intrusion-related gold deposits associated with tungsten-tin
provinces: Mineralium Deposita, v. 34, p. 323334.
Titley, S.R., 1993, Relationship of stratabound ores with tectonic cycles of the
Phanerozoic and Proterozoic: Precambrian Research, v. 61, p. 295322.
Tohver, E., van der Pluijm, B.A., van der Voo, R., Rizzotto, G., and Scandolara, J.E., 2002, Paleogeography of the Amazon craton at 1.2 Ga: Early
Grenvillian collision with the Llano segment of Laurentia: Earth and Planetary Science Letters, v. 199, p. 185200.
Trubitsyn, V.P., Mooney, W.D., and Abbott, D.H., 2003, Cold cratonic roots
and thermal blankets: How continents affect mantle convection: International Geology Reviews, v. 45, p. 479496.
Unrug, R., 1997, Rodinia to Gondwana: The geodynamic map of Gondwana
supercontinent assembly: Geological Society of America, GSA Today, v. 7,
p. 16.

0361-0128/98/000/000-00 $6.00

Vearncombe, S., Barley, M.E., Groves, D.I., McNaughton, N.J., Mikucki,


E.J., and Vearncombe, J.R., 1995, 3.26 Ga black smoker-type mineralization in the Strelley belt, Pilbara craton, Western Australia: Journal of Geological Society of London, v. 152, p. 587590.
Veizer, J., Laznicka, P., and Jansen, S.L., 1989, Mineralization through geologic time: The recycling perspective: American Journal of Science, v. 289,
p. 484542.
Whiting, B.H., Hodgson, C.J., and Mason, R., 1993, Giant ore deposits: Society of Economic Geologists Special Publication 2, 349 p.
Williams, P.J., and Skirrow, R.G., 2000, Overview of iron oxide-copper-gold
deposits in the Curnomona province and Cloncurry district (eastern Mount
Isa block), Australia, in Porter, T.M., ed., Hydrothermal iron oxide-coppergold and related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p.105122.
Windley, B.F., 1984, The evolving continents, 2nd ed.: New York, John Wiley
and Sons, 399 p.
Wyman, D.A., Kerrich, R., and Groves, D.I., 1999, Lode gold deposits and
Archean mantle plume-island arc interaction, Abitibi subprovince, Canada:
Journal of Geology, v. 107, p. 715725.
Yakubchuk, A., Cole, A., Seltmann, R., and Shatov, V., 2002, Tectonic setting,
characteristics, and regional exploration criteria for gold mineralization in
the Altaid orogenic collage: the Tien Shan province as a key example: Society of Economic Geologists Special Publication 9, p. 177201.
Zegers, T.E., Barley, M.E., Groves, D.I., McNaughton, N.J., and White, W.J.,
2002, Oldest gold: Deformation and hydrothermal alteration in the Early
Archean shear zone-hosted Bamboo Creek deposit, Pilbara, Western Australia: ECONOMIC GEOLOGY, v. 97, p. 757776.

224

Das könnte Ihnen auch gefallen