You are on page 1of 28

13.

1 Fluids and Ore Formation in the Earth’s Crust


CA Heinrich, ETH Zurich, University of Zurich, Zurich, Switzerland
PA Candela, University of Maryland, College Park, MD, USA
ã 2014 Elsevier Ltd. All rights reserved.

13.1.1 Ore Deposits and Crustal Geochemistry 2


13.1.1.1 Ores, Mineral Deposits, and Crustal Composition 2
13.1.1.2 Economic Aspects: Deposit Assessment, Grade, and Resources and Reserves 2
13.1.1.3 Extreme Metal Enrichment: A Rare Conjunction of Common Processes 4
13.1.2 Magmatic Ore Formation 5
13.1.3 Ore-Forming Hydrothermal Processes 8
13.1.3.1 Physical Aspects of Hydrothermal Metal Enrichment 9
13.1.3.1.1 Energy sources driving fluid flow 9
13.1.3.1.2 Fluid focusing and ‘structural control’ 10
13.1.3.1.3 Timescales of fluid flow and heat transfer 10
13.1.3.2 Chemical Driving Forces for Hydrothermal Metal Enrichment 11
13.1.3.2.1 Rock buffering and fluid-chemical master variables 11
13.1.3.2.2 Sulfide solubility and base-metal content of crustal fluids 12
13.1.3.2.3 Example: Hydrothermal redistribution of gold 14
13.1.3.3 Metal Source Regions versus Ore Deposition Sites 15
13.1.3.3.1 Source region characteristics 15
13.1.3.3.2 Ore deposition and wall-rock alteration 15
13.1.4 Hydrothermal Ore Formation in Sedimentary Basins 16
13.1.5 Hydrothermal Ore Systems in the Oceanic Realm 17
13.1.6 Magmatic–Hydrothermal Ore Systems 17
13.1.6.1 Melt Generation, Magma Storage, and Ascent 17
13.1.6.2 Fluid Exsolution and Element Transfer from Melt to Fluid 19
13.1.6.3 Fluid Cooling, Decompression, and Phase Separation 19
13.1.6.4 Fluid Evolution Paths at the Porphyry to Epithermal Transition 20
13.1.6.5 Wall-Rock Alteration and Ore Mineral Precipitation 21
13.1.7 Ore Formation at the Earth’s Surface 22
13.1.8 Back to the Future: Global Mineral Resources 23
Acknowledgments 23
References 23

Glossary carbonate salts of Na, K, Ca, Fe, and other cations. More
Fluid Used here for a low-viscosity mobile phase rigorous definitions are proposed in Liebscher and Heinrich
that has a density much smaller than that of silicate (2007).
materials; such fluids are usually dominated by volatile Melts Melts dominated by silicate, sulfide, carbonate, oxide,
components including H2O plus various species of and borate components are distinguished in this chapter,
C, S, and Cl salts. Hydrothermal specifies a fluid using the respective qualifiers, even though they are also
heated by Earth’s interior energy to a temperature exceeding ‘fluids’ in the physical sense and are key to selective element
50  C. enrichment in magmatic ore deposits (discussed in greater
Liquid and vapor (equivalent to liquid water and detail in Chapter 13.8).
steam in the pure H2O system) Two types of Mineral deposit Any anomalous accumulation of minerals
hydrothermal fluids that can exist individually, or may and chemical elements contained in them. The term ore
stably coexist as distinct phases in contact with each other. deposit is reserved here for a mineral deposit in which grade
Vapor (gas) will homogeneously fill any available volume (bulk concentration of valuable elements) and the quantity
and is generally of lower density and enriched in gas-like of ore (the tonnage of a deposit) allows extraction of one or
components including CO2, CH4, or H2S, whereas a liquid is several elements for profit, at current economic and
denser and commonly enriched in chloride, sulfate or technical conditions.

Treatise on Geochemistry 2nd Edition http://dx.doi.org/10.1016/B978-0-08-095975-7.01101-3 1


2 Fluids and Ore Formation in the Earth’s Crust

13.1.1 Ore Deposits and Crustal Geochemistry mineable at a grade of  2.4 ppm from a rock volume with a
network of numerous veins that precipitated from hydrother-
13.1.1.1 Ores, Mineral Deposits, and Crustal Composition
mal fluids (Goldfarb et al., 2005; Wall et al., 2004). Gold is
Ten elements (O, Si, Al, Fe, Ca, Mg, Na, K, Ti, and H) make up enriched by a factor of  1800, compared to average crustal
approximately 99.6% of the Earth’s continental crust by mass abundance of 1.3 ppb Au. Multiplying the ore tonnage with
(Rudnick and Gao, 2003). All other industrially important ele- the enrichment factor yields 4  1012 t of average crustal rocks,
ments can be considered as scarce, making up a small fraction of or approximately 1400 km3 of average crust, from which all its
the remaining 0.4% with concentrations of less than 1000 ppm 1.3 ppb Au would have to be extracted and accumulated in this
(mg g1) each. They range from vital elements such as phos- deposit. No process of extraction and ore deposition is 100%
phorus and manganese at hundreds of ppm to precious metals efficient, of course, indicating that the ore-forming process has
at parts per billion levels in average continental crust. Variations probably operated within an even larger region of the crust and
in the major element composition among crustal rocks are possibly the underlying mantle.
manifested in the proportions of rock-forming minerals and Mass balance imposed by enrichment factors and deposit
common accessory phases, and most of the scarce elements are dimensions bears two important conclusions for the scale
normally accommodated in the crystal structure of these min- and process of ore formation (indeed, similar calculations
erals. Economically viable and energy-efficient extraction of can be made with similar results for most other scarce
such elements for industrial use requires specialized rocks, in elements and ore types). First, the physical size of regions of
which one or several valuable elements are selectively enriched the Earth’s crust in which an ore-forming process operates
and usually occur as separate mineral phases. If such geo- must be large, many orders of magnitude larger than the size
chemical and mineralogical anomalies are large and rich of the small and seemingly anomalous rock volumes consti-
enough to be mined for a profit, they are called ore deposits tuting economic ore deposits. Understanding ore formation
(Skinner, 1997). The geological processes leading to the forma- therefore requires understanding ore-forming systems, that is,
tion of ore deposits are introduced in this chapter and represent regions in the crust typically extending over tens to hundreds
the theme of this new volume of Treatise on Geochemistry. of kilometers in lateral extent and perhaps down to mantle
The scarce elements form positive anomalies with varying depths, in which scarce metals are selectively redistributed.
efficiency. To a first order, the behavior of scarce elements is Second, the scale of ore systems makes it physically impossi-
controlled by the extent to which their crystal–chemical behav- ble that the formation of an economic ore deposit occurs
ior differs from that of the major elements in rock-forming by element diffusion through any stagnant material, which
minerals, that is, to what extent they enter into solid solution at best extends over a characteristic distance of <100 m on
in common silicates (feldspars, pyroxene, amphibole, micas, a timescale of 1 My (for aqueous fluid at T <600  C;
clay minerals, and quartz), carbonates, oxides (e.g., magnetite) solid diffusion would allow at most 10 mm; Watson and
and sulfides (pyrite and pyrrhotite). Some scarce elements can Baxter, 2007).
be enriched to economic degrees in solid solutions of major All ore-forming systems therefore combine chemical reac-
rock-forming minerals (e.g., Rb and Li in micas or V in mag- tions with physical transport: selective dissolution and precip-
netite) and some are selectively incorporated in accessory min- itation of minerals are linked with large-scale advective
erals of common rocks (e.g., zirconium and minor hafnium in transport of a mobile phase through rocks. The diversity of
zircon; rare earth elements (REE) in monazite). However, many fluids and melts, the range of physical driving forces for trans-
elements such as base metals (Cu, Pb, and Zn), semimetals (As, port, and the specific crystal–chemical controls on mineral
Sb, and Bi), precious metals (Au, Ag, and platinum-group dissolution and precipitation account for the wide range of
elements) and many more (e.g., Mo, Re, In, Ge, and Hg) do ore-forming processes, leading to distinct but globally recur-
not have a strong affinity for minerals that constitute the bulk of rent ore deposit types. This chapter aims to introduce this
the crust. They are excluded from all but the most limited solid diversity, while emphasizing the common principles.
solution in rock-forming minerals. These elements tend to par-
tition into less-ordered phases including fluids (aqueous solu-
tions, salty brines, and vapor) or melts (silicate melts, but also
13.1.1.2 Economic Aspects: Deposit Assessment, Grade,
sulfide, carbonate, or even borate melts). Such phases are also
and Resources and Reserves
physically mobile and therefore allow selective element trans-
port and local precipitation of scarce elements as separate min- Before introducing the range of major processes conspiring to
eral phases, such as sulfides containing Cu, Pb, or Zn as a major generate ore deposits, we will illustrate some important eco-
stoichiometric component. nomic terms. They touch on the global issue of future avail-
Ore-forming processes operate at a minimum scale defined ability of raw materials, but they also show the geochemical
by the size of ore deposits and the degree of element enrich- significance of ore deposits in the crust. The economic assess-
ment needed for their economic utilization. Table 1 shows a ment of an ore deposit always has a historical perspective,
selection of elements, ordered by decreasing average crustal determined by the changing value of materials to society and
abundance, in comparison with the typical concentration in the changing technology of extraction and further beneficia-
large ore deposits mined today (ore grade). Besides the enrich- tion of metals from natural ore. Therefore, definitions of eco-
ment factor (the so-called Clarke value ¼ ore grade/average nomic terms including ore deposit types, ore grades, reserves,
crust), the table also shows the size (in tonnes of ore) of one and resources are subject to change over time, and predictions
of the largest ore deposits of each type. For example, the for future availability require in-depth geological understand-
Muruntau gold deposit contains a total of 5300 t of Au, ing (Kesler, 1994; Singer, 2010).
Fluids and Ore Formation in the Earth’s Crust 3

Table 1 Typical mineable ore grades of selected elements in comparison with their average crustal abundance

Element Deposit type Typical ore grade Concentration Enrichment Largest deposits Tonnage Grade
(ppm ¼ g t1) in crust (ppm) factor (example)

Al Lateritic bauxite 240 000 84 200 3 Weipa (Australia) 149 Mt 28%


Fe Banded iron formation, 640 000 52 200 12 Mount Whaleback >1800 Mt 65.0%
upgraded (Australia)
Ti Mafic intrusions (ilmenite) 80 000 5700 14 Lac Tio (Canada) >200 Mt 14%
P Marine sedimentary 100 000 570 180 Meade Peak Member >1011 Mt 10%
deposits (USA)
V Layered mafic intrusions 2000 138 14 Bushveld Complex 50 Mt 0.8%
(magnetite) (South Africa)
Cr Layered mafic intrusions 160 000 135 1200 Bushveld Complex 2700 Mt 19%
(chromite) (South Africa)
Zn Stratiform Pb–Zn–Ag 87 000 72 1200 Broken Hill (Australia) 284 Mt 11%
deposits
Ni Magmatic Ni–Cu sulfide 13 000 59 220 Jinchuan, Baijiazuizi 515 Mt 1.1%
deposits (China)
Rb Pegmatites, from Rb- 5000 49 100 Tanco (Canada) 0.5 Mt 1.0%
bearing lepidolite
Cu Porphyry copper deposits 5000 27 180 El Teniente (Chile) 2850 Mt 1.3%
Cu Sediment-hosted copper 23 000 27 850 Lubin (Poland) 2600 Mt 2.0%
deposits
Pb Stratiform Pb–Zn–Ag 47 000 11 4300 Broken Hill (Australia) 160 Mt 10.0%
deposits
Nb Carbonatites 1900 8 240 Seis Lagos (Brazil) 2898 t 2%
Eu Carbonatites 1300 1.1 1200 Bayan Obo (China) 750 Mt 4.1%a
W Skarn, replacement and vein 3300 1 3300 Shizhuayan (Skarn, 270 Mt 0.22%
deposits China)
Mo Porphyry molybdenum 2000 0.8 2500 Climax (Colorado, USA) 769 Mt 0.22%
deposits
Ta Pegmatites 1800 0.7 2600 Tanco (Canada) 2.07 Mt 0.18%
Ag Polymetallic vein deposits 160 0.056 2900 Cerro Rico de Potosi 985 Mt 116 g t1
(Bolivia)
In Volcanic-hosted massive 24 0.052 500 Kidd Creek (Canada) 134 Mt 50 g t1
sulfide deposits
Hg Epithermal veins and 10 000 0.030 300 000 Almaden (Spain) 40 Mt >1%
breccias
Pt Magmatic PGE–Ni–Cu 3 0.0015 2000 Bushveld Complex 9815 Mt 2.3 g t1
sulfide deposits (South Africa)
Au Hydrothermal vein deposits 2 0.0013 1200 Muruntau (Uzbekistan) 2200 Mt 2.4 g t1
Au Modified Archaean placers 5 0.0013 4000 Witwatersrand Basin >104 Mt 5 g t1
(South Africa)
a
Sum REE oxides.
Tonnage (total mass of mined and economically mineable ore) and average grade of some of the largest ore deposits of each type.
All values in weights units. Average concentration in crust from Rudnick and Gao (Chapter 4.1 Table 11, total crust).
Estimates of average grades indicative only, mostly based on USGS and Canadian Geological Survey compilations but partly representing single big deposits. Grades and tonnages of
some of the largest deposits mostly represent single mines, but in the case of the Meade Peak phosphorite, the Bushveld Complex, and the Witwatersand Basin are estimates for one or
several rock units that are mineralized at presently economic ore grades for the respective elements.

Ore grade and tonnage. Element concentrations in the Earth’s economic ore (Singer, 1995; Singer et al., 2005). Kesler and
crust show a broadly log-normal distribution, extending from Wilkinson (2008) estimated that only about 0.08% of the total
‘ordinary’ rocks through element anomalies and subeconomic copper contained in the continental crust is concentrated in
mineral deposits, to economic ore deposits (Shen and Zhao, discovered and still undiscovered deposits of currently eco-
2002). Ore grade denotes the average element concentration nomic grade and size. This estimate is based on modeling the
contained in one or several orebodies making up a deposit. The known frequency of porphyry copper deposits in terms of
mass of rock, calculated to be extractable at an economic profit, time-dependent tectonic diffusion, that is, a random chance
is called the tonnage of an ore deposit. Grade and tonnage were of tectonic burial, or uplift and recycling by erosion, after
initially used by Lasky (1950) to show the inverse relationship formation of each deposit at an average depth known from
between average grade and tonnage of porphyry copper de- geological studies. Both calculations show that ore deposits are
posits. Statistical analyses show that the few largest deposits of geochemical anomalies and do not contribute significantly to
each metal contain the greatest proportion of potentially the bulk inventories of any element in the Earth’s crust. Mining
4 Fluids and Ore Formation in the Earth’s Crust

correspondingly occupies a vanishingly small proportion of particularly valuable ore deposits containing a high proportion
the Earth’s surface. of Zn/Pb, despite the geochemical similarity of these two metals.
Reserves and resources are two distinct economic terms of
great geochemical as well as political interest (Kesler, 1994;
13.1.1.3 Extreme Metal Enrichment: A Rare Conjunction
Pohl, 2011). The term reserve applies, for an individual deposit
of Common Processes
or for the global reserves of a metal, to the total metal quantity
that is known to be extractable in the present economic situa- Ore deposits are rare because extreme element enrichment re-
tion. By contrast, resources are subject to lower degrees of quires a conjunction, spatially and temporally, of physical trans-
certainty or are subeconomic at the time. As new and innovative port processes and selective chemical reactions between minerals
technologies of mining and metal extraction are developed, and and fluids or melts. We think that the complexity in this field of
as exploration improves the level of knowledge about metal geoscience, with numerous deposits being termed ‘unique’ in the
distribution in the crust, subeconomic mineral deposits are specialist literature, can be broken down by understanding the
turned into economic ore deposits. Thus, new metal reserves interactions between otherwise common and quite predictable
become available for use. The conversion of resources into re- processes. These processes can be rationalized by three factors,
serves by geological exploration explains the apparent paradox (1) the type of medium allowing selective component transport,
that global reserves of copper and other essential metals have (2) global-scale tectonic drivers from the Earth’s interior, and
continually increased over the last decades, even though these (3) interactions of endogenous processes with the atmosphere,
effectively nonrenewable commodities have been intensely which has been subject to changing climate and oxygen levels
extracted and reserves of individual mines have been depleted through the Earth’s history (Robb, 2004).
in the same period (Tilton and Lagos, 2007). Generally, ele- The nature of the mobile phase for selective element transport
ment resources for future use are not so much limited by yields a first subdivision of ore-forming processes (Figure 1).
availability, but by increasing economic, environmental, and Thus, we distinguish magmatic ore deposits in which element
energy costs of extracting them from lower grade ores. separation is primarily effected by silicate, sulfide, or carbona-
The economic importance of multiple elements that are tite melts. Hydrothermal ore deposits, in the widest sense of
enriched in certain ore deposit types has been subject to historic the term, are metal enrichments by selective dissolution of
change, in light of changing metal utilization and technology. mineral components in a fluid allowing large-scale transport
This is illustrated by uranium–vanadium deposits (Candela, and physical focusing into a smaller volume of permeable
2003), as one of many examples. Vanadium and uranium are rock, followed by selective precipitation of metal-rich minerals
enriched together because they share the chemical property of into primary or secondary pore space (Barnes, 1997). Surface
being highly soluble in oxidized form (U6þ and V5þ aqueous water can generate important ore deposit types by the converse
complexes), but precipitate as oxide and silicate minerals upon process of selective removal of rock material at the Earth sur-
reduction (V3þ in vanadiferous magnetite and silicates, U4þ in face (residual ores), either by chemical dissolution or mechan-
uraninite  UO2) when fluids contact carbonaceous material. ical winnowing, such that a valuable element becomes
From 1912 to 1922, the sandstone-hosted uranium deposits on enriched in the solid residue (Figure 1, right).
the Colorado Plateau were the first major sources for radium for The subdivision into magmatic, hydrothermal, and surface
medical use, but then became vanadium mines from 1924 to processes provides a framework for the genetic classification of
1945 (Fitch, 1997). Uranium and its daughter isotopes were major ore deposit types, as summarized in Table 2 and illus-
discarded as part of the vanadium mine tailings, but from 1947 trated by characteristic ore-related textures in Figure 2. Table 2
these tailings became a source for much of the uranium used for illustrates that most elements are selectively enriched in only
atomic weapons. The use of uranium for nuclear power plants one or a few of these major ore types, ultimately reflecting the
further increased demand in the early 1950s, so that the former crystal–chemical properties of each element. Some deposit
vanadium deposits became uranium deposits producing vana- types are transitional in origin, for example, sedimentary man-
dium as a by-product. Many deposits contain about ten times ganese and banded iron formations, which are precipitated
more vanadium than uranium, so geochemically, uranium from seawater after low-temperature interaction with rocks at
could be considered a by-product of vanadium mining. When the ocean floor or enrichment by metals from hydrothermal
the demand for uranium dropped and rising worldwide steel black smokers. Other deposits form by superposition of pro-
production increased demand for vanadium in the late cesses, for example, the secondary enrichment of copper in
1990s, many sandstone uranium deposits turned again into supergene orebodies by weathering of primary hydrothermal
vanadium (uranium) mines. Similar changes affected sulfide copper deposits. Grains of gold or cassiterite (SnO2) that were
deposits, in which lead and silver were the most valuable prod- deposited originally by hydrothermal solutions can be accu-
ucts from medieval times (roofing and coins) to the early mulated in alluvial gravels after erosion of the primary de-
industrial age (car batteries and leaded fuel), until environmen- posits, forming secondary placer deposits (Figure 2(a)).
tal considerations and the advent of protective coating of steel Global tectonics is the second major factor in the formation
turned many lead deposits into dominant zinc resources. of distinct ore deposit types and a key to understanding
These examples show that we must be careful how to inter- their spatial and secular distribution (Groves et al., 2005;
pret economic information about ore deposits in terms of its Holland, 2005; Figure 3). Thus, chromium deposits form by
geochemical significance. On the other hand, the examples also physical accumulation of what is usually an accessory mineral
emphasize the practical importance of understanding metal (chromite, FeCr2O4) from large reservoirs of mafic magma
enrichment more quantitatively. Such understanding is essential (Figure 2(f)). Such magma chambers either form as intraplate
to predict, for example, the distribution and reserves of layered intrusions (possibly related to mantle plumes) or at
Fluids and Ore Formation in the Earth’s Crust 5

Magmatic ores Hydrothermal ore formation Surface ores

Deposition
site
Source and focusing
Dispersion
of fluid

+ Energy

Mechanical enrichment
of mineral particles
by surface water

Basin flow by
Magmatic–hydro- Thermal convection topographic head
Element segregation thermal expulsion of of fluid by heat and solar energy ± Chemical residue
via melts overpressured fluid from Earth interior sediment compaction of extreme weathering

Figure 1 Schematic illustration of three principal processes of selective element enrichment that can produce mineral deposits in the Earth’s crust,
grouped by the main agent of material transport: magmatic melts, hydrothermal fluids, or surface waters (Kesler, 1994). With permission from S.E. Kesler.

spreading axes in the oceanic domain (Stowe, 1994). Porphyry of the Earth’s history, the reduced atmosphere prevailing in the
copper deposits are generally related to subduction zones, Archean prevented destruction and allowed mechanical trans-
because their formation depends on magmatic volatiles that port and enrichment of uraninite grains as well as pyrite peb-
had previously contributed to partial melting of hydrated man- bles by surface water (Figure 2(a)). This produced giant
tle above a descending hydrous slab (Hedenquist and Low- paleoplacer deposits of gold, notably in the 3 Ga old Witwa-
enstern, 1994; Sillitoe, 2010). A temporal clustering of gold tersrand basin (Frimmel et al., 2005; Chapter 13.17; but see
deposit formation during the Late Archean, the Paleoprot- Law and Phillips, 2005 for an alternative interpretation).
erozoic, and the Phanerozoic is thought to be related to three In summary, three groups of factors – transport medium,
periods of particularly rapid crustal growth (Bierlein et al., global tectonics, and the Earth’s changing climate and atmo-
2006; Goldfarb et al., 2005), although the age distribution is sphere – allow a first-order explanation for the range of
also affected by recycling of Archean gold and preferred pres- major ore deposit types. Common to all ore-forming processes
ervation in younger orogens (Frimmel, 2008). The formation is the fact that selective element enrichment is an entropy-
of sediment-hosted lead-zinc deposits requires large-scale flow reducing process that requires a net input of energy. For a deeper
of saline fluids through stable aquifers, which can be effected understanding of the factors controlling ore formation, we there-
in shallow-marine intracontinental basins experiencing tran- fore need to focus on the energy sources – that is, the physical and
sient uplift of the basin margin as a result of continental chemical driving forces – contributing to processes separating
collision (Bethke, 1986; Garven et al., 1993). elements from each other on a large scale and with high efficiency.
Climate and atmosphere contribute a third global factor to
the formation of ore deposits through the Earth’s history. Thus,
transport of many base metals is facilitated by the high salinity
of fluids that are generated in sedimentary basins following 13.1.2 Magmatic Ore Formation
periods of extensive evaporation of seawater. This climate-
controlled accumulation of chloride, as a ligand for metal Magmatic ore deposits are products of the Earth’s interior
complexation, explains at least partly the predominant age heat engine, which drives nested convection systems from
grouping of major base metal provinces in the Paleoprotero- global mantle dynamics (plate tectonics, deeply sourced mantle
zoic, in the late Neoproterozoic to the Cambrian, and follow- plumes) to the convective separation of buoyant melts from
ing global periods of evaporite deposition in the late Paleozoic generally denser solid materials (Candela, 2003). Magmatic
to the early Mesozoic (Holland, 2005). (orthomagmatic) ore deposits owe their economic metal en-
Since the advent of an oxygenated atmosphere in the Pro- richment to a combination of selective melting from a region of
terozoic, hydrothermal uranium deposits formed by oxidative previously modified (‘fertile’) mantle or lower crust, followed by
leaching by surface-derived aquifer fluids, followed by re- igneous fractionation involving one or several melt phases
precipitation at local sites of reduction by organic material (Chapter 13.8). Generally, element partitioning between melts
(Chapter 13.19). In contrast to the sandstone-type and and a ‘getter’ phase of distinct chemical (e.g., crystal structure and
unconformity-type uranium deposits formed in the latter half solubility in sulfide melt) and physical properties (e.g., density
6 Fluids and Ore Formation in the Earth’s Crust

Table 2 Simplified genetic classification of ore deposit types of main elemental (mostly metal) resources

Process, transport medium Ore deposit type (chapter number, this edition) Key source for Important regions

A. Orthomagmatic ore deposits: Element enrichment from melts at high T, without essential involvement of aqueous fluids
Crystallization from magma Chromite in (ultra-)mafic cumulates (13.8) Cr RSA, CIS, Zimb
V-magnetite, ilmenite in mafic intrusions (13.8) V, Ti Can, RSA, CIS
Sulfide melt segregates from Ni þ Cu sulfide deposits in mafic intrusions, Ni, PGE (Cu) Canada, CIS, Aus
silicate melt komatiites and flood basalts (13.8)
PGE sulfide deposits in layered intrusions (13.8) Pt, Pd, Rh (Au) RSA, CIS
Low-degree partial melting of Carbonatites (13.8; 13.21) REE, Nb, Ta Zr, P China, RSA, USA
mantle Kimberlite, Lamproite (3.5; 13.23) Diamond RSA, Aus, Canada,
CIS
Extreme magmatic fractionation Pegmatite deposits as residual silicate melts (2.9) Li, Be, Nb, Ta, gems Brazil, Can, CIS

B. Hydrothermal ore deposits: Enrichment by selective dissolution, transport, and precipitation from aqueous fluids of different sources
Fluid saturation of H2O-bearing Porphyry deposits (‘Cu, Mo porphyries’) (13.14) Cu; Mo Au Circumpacific
magmas: magmatic– Sn–W-veins and replacement deposits Sn, W (Cu, Pb, Ag) GUS, China, Aus
hydrothermal Skarn and Cordilleran deposits replacing W, Cu, Au China W
carbonates
Continental geothermal systems Epithermal veins and breccias (13.15; 13.16) Au, Hg (Ag, Pb, Zn) Circumpacific, CIS
( shallow magmatic fluids)
Continental magmatic fluids Iron oxide hosted Cu–Au–U–REE (‘IOCG’) deposits Cu, Au, U, Fe, REE Aus, Brazil, Chile
interacting with evaporitic brines (13.20)
Seawater convection through Volcanic-hosted massive sulfide deposits (13.18) Cu, Zn (Pb, Au, Sn) Can, Iberia, Aus
oceanic crust Mid-ocean ridges  ‘black smokers’ Cyprus; TAG
Arc-related basins and submarine volcanoes Kuroko; Kermadec
Metamorphic dehydration  deep Orogenic (‘mesothermal’) gold-quartz-veins Au Can, Aus, CIS
magmatic fluids (13.15)
Archean greenstone belts, younger continental
orogens
Basin brines: connate or meteoric Sediment-hosted Cu  Co deposits (13.10) Cu, Co, Ag USA, C. Africa, Poland
water þ evaporites Stratiform Pb–Zn–Ag deposits (13.9) Pb, Zn, Ag Can, Aus, Ireland
(‘sedex’ ¼ sedimentary-exhalative, early
diagenetic)
Mississippi Valley-type Pb–Zn ¼ epigenetic Pb, Zn, Cd; F USA, Can
replacement in carbonates  sandstones (13.9)
Deep infiltration of oxygenated Uranium in sandstones (‘roll-front U deposits’) U, V USA
surface waters Vein- and ‘unconformity-related U deposits’ U (Au, Pd, Pt) Can, Aus
(13.19)

C. Surface ore deposits, forming directly at hydrosphere interface


Chemical sedimentation on shelf or Banded iron formations (‘BIF,’ mostly late Fe Aus, Brazil, CIS
deep ocean floor Archean) (9.18)
Sedimentary Mn ore (modern oxic ocean) Mn RSA, CIS, Aus
Ferromanganese nodules & crusts (13.11) Mn, Co, Ni, Cu, Y, Te Pacific, Indian Ocean
Sedimentary phosphate deposits (13.12) P USA, CIS, Morocco,
China
Evaporation of marine or Salt deposits (13.22) NaCl, KCl, Mg, Li, Bolivia, USA
continental surface water B, OH–
Continental surface weathering: Residual deposits:
removal of soluble elements and Bauxite Al, Ga Aus, Surinam, Brazil
local redistribution Nickel laterites Ni New Caledonia
Secondary enrichment of primary ores Fe deposits from Brazil, Aus
BIF
Supergene; sulfide and oxide deposits Porphyry Cu SW USA, Chile
deposits
Mechanical enrichment by flowing Placer deposits (13.17):
surface water  wind Fluvial placers Sn, Ta, Au Malaysia/SW Asia
Beach placers: heavy mineral sands Ti, Zr, REE Aus, India
Fossil placers: Witwatersrand Conglomerates Au, U RSA

Elements in bold denote this deposit type as most important global source; elements in brackets are typically associated and economically important although deposit type is not the
dominant source. Important regions include regions of dominant world production as well as some classic study areas.
Fluids and Ore Formation in the Earth’s Crust 7

(a) (b)

(c)

(d)

(e) (f)

Figure 2 Textures of ore samples and related fluid processes, recording characteristic time relations between mineral precipitation and the interaction
of fluids and melts. All scale bars 1 cm. (a) Archean quartz–pyrite pebble conglomerate enriched in alluvial gold and uraninite by flowing surface water
(Witwatersrand Basin, South Africa). (b) Stratiform Pb–Zn sulfide ore showing evidence of sedimentary (crossbedding) to early diagenetic (birds-eye
textures) precipitation of galena and spalerite (HYC deposit, McArthur River, Australia). (c) Missisippi Valley-type Pb–Zn ore in brecciated dolomite
cemented by successive crystal layers of markasite, spalerite, galena and final vug-filling calcite (Pillara deposit, Canning Basin, Australia). (d) Orogenic
gold-quartz vein cutting Archean dolerite, where sulfur-rich fluids converted ferromagnesian silicates to a bleached alteration halo of auriferous
pyrite þ albite þ muscovite þ ankerite (Mt. Charlotte mine, Kalgoorlie, Australia). (e) Miarolitic granite showing textural evidence of the separation of a
magmatic fluid phase during crystallization of silicate magma, which may give rise to later hydraulic fracturing and magmatic-hydrothermal Sn, W, or Cu
ore deposition (Ruby Creek Granite, Stanthorpe, Australia). (f) Magmatic chromite ore showing graded interbedding of olivine-rich (green) and
chromite-rich (black) cumulate layers that settled out of a basaltic silicate melt (Troodos ophiolite, Cyprus).

and viscosity) attracts selected trace elements and then separates differentiation involving both crystal separation and silicate–
them to form magmatic ores. carbonate melt immiscibility. Crystal fractionation of more
Most orthomagmatic deposits can be attributed to one of normal, grossly basaltic magmas and gravitational settling of
four dominant enrichment processes (Table 2). A very low early crystallized heavy oxide minerals (chromite and V-rich
degree of partial melting of commonly preconditioned upper magnetite) is the main mechanism for producing the world’s
mantle is an essential step in the enrichment of REE in car- Cr and V deposits in continental layered intrusions or at oce-
bonatites (e.g., Bayan Obo; Mountain Pass) and of P in mag- anic spreading centers (Cawthorn et al., 2005). Irvine (1977)
matic apatite deposits related to alkaline intrusions (Kola suggested that nearly monomineralic layers of chromite
region; Palabora), in both cases followed by further magmatic (Figure 2(f)) could form by contamination of the magma by
8 Fluids and Ore Formation in the Earth’s Crust

Intraplate magmatism Subduction zone Mid-ocean ridge Orogeny Foreland Continental weathering
rift sedimentation magmatism magmatism metamorphism sedimentation and erosion

B6 C2
B5 C1 C2 C3
B2 B1 Continental
A5 A1–A4 C4 B4 B2 B3
A1 crust
Lithosphere

Oceanic
crust Asthenosphere

Continental Continent – ocean convergence Sea floor spreading Continent – continent Stable continent
rifting Active continental margin Passive plate boundary collision Passive continental margin

A: Magmatic ore deposits B: Hydrothermal ore deposits C: Surface-related ore deposits

A1. Chromitites as (ultra-)mafic cumulates B1. Porphyry Cu (Mo, Au) and C1. Residual ore deposits:
in layered intrusions and ophiolites epithermal Au, Ag (Hg, ...) deposits bauxite (Al), Ni-laterite deposits

A2. V-magnetite in mafic intrusions (V, Ti) B2. Sn-W veins and greisens in granites C2. Alluvial placer deposits (Sn, Ta, Au; U)

A3. Pegmatites (Li, Cs, Be, Nb, Ta) B3. Orogenic (’metamorphogenic’) C3. Beach-sand placers (Ti, Zr, REE)
Au-quartz vein deposits
A4. Ni- und PGE-sulfide deposits in C4. Manganese nodules and crusts
mafic intrusions und flood basalts B4. Volcanogenic massive sulfides (Cu, Zn) on the ocean floor (Mn, Co, Ni, Cu...)

A5. Carbonatite (REE) and kimberlite B5. Sediment-hosted (MVT, ‘sedex’) - - - - B6. Sandstone-hosted and unconformity-related
(diamond) deposits Pb, Zn, Cd, Cu, Co deposits U (V, F, Mo, Au, PGE) deposits

Figure 3 Schematic illustration of the recurrent association of ore deposit types with global tectonic settings, including active and passive
continental margins, oceanic spreading centers including back-arc basins, as well as land surfaces, sedimentary basins and hot spots in the interior of
lithospheric plates. Labels A, B, C refer to major metal-transporting media and ore deposit types in Table 3.

felsic material or by mixing of more primitive magma with supracrustal granite–greenstone basement (Naldrett, 2010).
more evolved magma. Melt segregation, notably saturation of Related to more modern plume activity, and tentatively asso-
an immiscible sulfide melt that may accumulate by gravita- ciated with the greatest life extinction event at the Permian–
tional settling, is the preeminent process generating magmatic Triassic boundary (Courtillot and Olson, 2007), the Siberian
Ni  Cu and platinum-group elements (PGE) deposits in large trap basalts also gave rise to a globally significant Ni–Cu–PGE
layered intrusions (Sudbury, Bushveld; Naldrett, 2010) or in province centered in the Norilsk mining area (Li et al., 2009a).
flood basalt and komatiitic flows and sills (Barnes and Light- Although orthomagmatic ore deposits generally do not in-
foot, 2005; Lesher et al., 1981; Li et al., 2009a; Naldrett, 2010). volve a low-density fluid phase as an essential agent, many
In all these deposits, a process of contamination of mantle- magmatic deposits show evidence of some fluid activity. This
derived magmas with crustal material, commonly serving as includes the precipitation of gem-quality beryl, tourmaline,
the source of external sulfur, is key to magmatic sulfide ore and topaz from residual hydrothermal fluids derived from
formation (Arndt et al., 2005). The fourth process is extreme pegmatites. The origin of the world’s preeminent Pt and Pd
crystal fractionation of felsic (acid and/or alkali-rich) magmas, deposits in the Bushveld and Stillwater intrusions probably
which leads to selective enrichment of incompatible elements includes interaction of silicate and sulfide melts with saline
including Li, Rb, Cs, F, B, Be, and B, eventually crystallizing as hydrothermal fluids. The ongoing debate about the origin
distinct minerals in magmatic pegmatites (Černý et al., 2005; see of stratiform PGE deposits related to magmatic unconfor-
also Chapter 13.21). mities, including the Merensky Reef in the Bushveld Complex,
Magmatic ore deposits show a distinct secular evolution, is reviewed in Chapter 13.8. Carbonatites invariably are asso-
dominated by deposits formed during the Archean and the ciated with late hydrothermal alteration (fenitization,) which
Paleoproterozoic. In this early period of the Earth’s evolution, can lead to unusual ore accumulations (e.g., Cu at Palabora;
mantle temperatures were higher and conducive to komatiite Groves and Vielreicher, 2001). Even the formation of chro-
volcanism and the formation of giant layered intrusions of mite deposits, especially those associated with ophiolites,
unique character, each giving rise to superlative mineral en- may require a CO2-rich fluid phase as an agent for mechan-
dowments of global significance (Maier and Groves, 2011; ical flotation of chromite grains attached to bubbles in
Naldrett, 2010). The largest among these is the Bushveld vertically extensive silicate magma channelways (Cawthorn
Complex which, like some of the giant komatiite flows et al., 2005).
(Arndt et al., 2005), has characteristics resembling giant
magma lakes and contributes our preeminent Cr and PGE
resources. The world’s premier nickel sulfide province consists 13.1.3 Ore-Forming Hydrothermal Processes
of many individual deposits at the base of the largest impact-
generated melt sheet in the geological record, the Sudbury Element enrichment by hydrothermal processes follows the
Igneous Complex. As a result of impact melting of an Archean principle of source – transport – deposition as illustrated in
craton, it has a unique melt composition of mixed mafic and the central panels of Figure 1. The large scale of ore-forming
Fluids and Ore Formation in the Earth’s Crust 9

systems requires a physical driving force for advective material for massive transfer of dissolved trace elements. However, flow
transport by flowing fluids, and chemical driving forces for rates through the aquifers are slow enough that the fluid is
reactions between minerals and the transporting phase, that always close to thermal equilibrium with the ambient heat
is, a solubility gradient along the flow path of the fluids. The flow from the Earth’s interior. In general, topography-driven
interaction of deep and surface-driven processes contribute to hydrothermal systems in tectonically stable continental areas
the great efficiency of hydrothermal ore formation in many do not create significant temperature differences between
distinct geological environments, by establishing steep gradi- fluids and rocks, even if flow is stratigraphically confined
ents in physical and chemical conditions within large-scale (Bethke, 1986). Minerals whose solubility is primarily con-
fluid transport systems. trolled by temperature are therefore likely to be deposited
over long distances along the transport path, rather than
being concentrated in high-grade ore deposits. Topography-
driven fluid systems can form economic ore deposits of metals
13.1.3.1 Physical Aspects of Hydrothermal Metal
whose solubility primarily varies with temperature-independent
Enrichment
chemical gradients – for example, the redox gradient controlling
13.1.3.1.1 Energy sources driving fluid flow the precipitation of U and V in sandstone-hosted ore deposits.
Hydrothermal fluids can be driven by the Earth’s interior heat Topography-controlled fluid pressure gradients can also modify
engine or by our planet’s external energy source, the Sun. This the flow paths of fluids driven by the Earth’s internal energy
summary focuses on hydrothermal ore formation (Cathles, sources. For example, thermally convecting geothermal waters
1981, 1997; Cox, 2010; Ingebritsen and Appold, 2011), build- on land are typically deflected to outflow points of lowest
ing on the wider literature of crustal fluid flow (Chapter 4.6; topographic elevation. Topography-driven flow of meteoric
Cathles, 1981; Fyfe et al., 1978; Ingebritsen and Manning, water in a volcanic edifice will interact with overpressured mag-
2010; Ingebritsen et al., 2006). Even though some ore deposits matic fluids expelled from an underlying magma chamber
show textural evidence for rapid and even turbulent fluid flow, (Ingebritsen et al., 2010), and this hydrologic interplay probably
Darcy’s Law is useful to discuss the key aspects of hydrothermal exerts a key control on porphyry-related and epithermal ore
fluid transport (Ingebritsen et al., 2006). deposits (Driesner and Geiger, 2007; see Chapter 13.14).
Darcy’s Law describing nonturbulent fluid flow through Thermal convection driven by the Earth’s interior heat is the
permeable rock can be expressed as dominant driving force for fluid flow through the oceanic
  crust, because the overlying ocean surface imposes a level of
q ¼  rpF  rF grz kR =mF [1] constant potential in the gravity field. Convection is driven by
 lateral gradients of fluid density, which can be caused by fluid
The volumetric flow rate q (in volume units per second composition (e.g., salinity: thermohaline convection along a
flowing through a cross-sectional area, the so-called Darcy salt diapir or fingering of an overlying saline aquifer into an
velocity in m s–1) is proportional to the fluid pressure gradient underlying freshwater aquifer). More commonly, lateral fluid
r pF minus the hydrostatic pressure differential caused by the density gradients are caused by temperature differences, lead-
weight of the fluid, rFg, with increasing depth, z, in the gravity ing to thermal convection. We distinguish forced thermal
field (Chapter 4.6). The flow rate is also proportional to rock convection due to an imposed horizontal temperature gradient
permeability kR, but inversely related to the dynamic viscosity (e.g., along the steep wall of a hot intrusion into cooler water-
mF of the fluid. Several sources of energy can build up fluid saturated rocks; Cathles, 1981) from spontaneous thermal
pressure gradients driving fluid flow through rocks. The differ- convection caused by a sufficiently large vertical temperature
ent energy forms lead to characteristically different rates and gradient, including seawater convection beneath mid-oceanic
patterns of subsurface fluid flow, and these impact significantly ridges (Fisher, 2005).
on the efficiency of generating ore deposits in different geolog- Thermal convection has two inherent ‘advantages’ for
ical environments (Cathles, 1997). efficient hydrothermal ore formation. First, fluid flow can
Solar energy evaporates water from the ocean surface and self-organize into relatively stable patterns with locally steep
drives air convection, allowing precipitation in areas of thermal gradients (notably at the base and at the top of con-
higher elevation. This mechanical work against the Earth’s vection cells). This leads to large integrated fluid fluxes favoring
gravity field creates the dominant driving force not only for massive solute transfer, combined with a sustained chemical
surface run-off, but also for topography-driven fluid flow driving force for mineral dissolution and precipitation.
through rocks (Cathles and Adams, 2005). Land topography, Second, it provides a mechanism of self-focusing of fluids
more precisely the lateral difference in water table elevation, from large areas of extensive contact with rock (amenable
leads to differential pressures of hundreds of bars, which drive to leaching of trace metals) to areas of high fluid flux (favoring
fluid flow through fractured basement rocks or sedimentary concentrated mineral deposition, e.g., in black smoker fields) –
basins and give rise to springs and artesian wells. In terms of even without any variations in rock permeability (e.g.,
time-integrated fluid flow, Australia’s Great Artesian Basin, the Coumou et al., 2008). The main limitation for convective
world’s largest system of continental subsurface water flow fluid transport is its restriction to regions of the crust that are
(Mahara et al., 2009), exemplifies the characteristics of sufficiently permeable to sustain hydrostatic fluid pressure
topography-driven fluid flow through rocks. Flow is large gradients. Thus, large-scale convection is generally restricted
scale and dominantly lateral, over thousands of kilometers. It to the permeable upper part of the crust where pore fluids
has transferred huge quantities of water in grossly the same effectively communicate with the water table near the land
direction over tens of millions of years, offering the potential surface or with the open ocean.
10 Fluids and Ore Formation in the Earth’s Crust

Essentially hydrostatic conditions extend from the surface to numerous cycles of fluid accumulation, cracking, flow, and
a characteristic depth, depending on the geological and the resealing of veins by precipitating minerals (Cox, 2005).
tectonic environment. The transition from hydrostatic to over- Magma ascent in the crust also occurs by active crack for-
pressured conditions varies from a few kilometers in regions of mation (diking), driven by lithostatic overpressure of lower
rapid sedimentation of impermeable clays in the Gulf of Mexico density magma under the weight of denser rocks overlying
(Reilly and Flemings, 2010) to at least 10 km in most areas of the melt generation zone. When a hydrous magma reaches
stable crystalline crust (Huenges et al., 1997; Townend and the upper crust, it crystallizes to an igneous rock and thereby
Zoback, 2000). For average continental crust, Ingebritsen and expels excess volatiles as hydrous fluid. The resulting net
Manning (2010) suggested from geothermal and metamorphic volume increase (VFluid þ VIgneousRock > VHydrousMelt; Burnham
data on fluid fluxes that average rock permeability (kR, m2) de- and Ohmoto, 1980) contributes an additional driving force
creases with depth (z, km) according to the relation: (pDV) for hydraulic fracturing and rapid fluid flow. Thus, a
large crystallizing intrusion can rapidly drain its fluid through a
log kR ¼ 14  3:2 log z [2] cupola in its roof, thereby forming a porphyry copper deposit

(Burnham and Ohmoto, 1980; Dilles and Einaudi, 1992; see
Log kR  1017 m2 at 10–15-km depth in the crust is an also Chapter 13.14).
effective lower limit for the permeability sustaining fluid flow
and hydraulic connectivity to the Earth’s surface, and this depth 13.1.3.1.2 Fluid focusing and ‘structural control’
range corresponds to the brittle–ductile transition zone of Focused fluid flow may result from convective self-
quartz þ feldspar-rich rocks. However, transient periods follow- organization, even in a rock of homogeneous permeability.
ing active fracturing (e.g., in the region of aftershocks following However, nearly all ore-forming hydrothermal systems include
major earthquakes, or artificially induced by hydraulic fractur- evidence of additional focusing of fluids by geological struc-
ing of geothermal reservoirs) raise this baseline permeability of tures of enhanced permeability, for example, sedimentary
the continental crust by two to three orders of magnitude (Evans aquifers or zones of fracturing. Known as ‘structural control’
et al., 2005; Rojstaczer et al., 2008; Townend and Zoback, 2000) by the exploration geologist, tectonic structures localizing in-
over periods of  1–103 years (Ingebritsen and Manning, 2010). creased fracture permeability are a key factor determining the
This is likely to allow episodes of significant thermal convection location of most major ore deposits.
throughout the seismogenic part of the continental crust and may At the lithospheric to crustal scale, irregularities in a com-
be a key to the injection of basement fluids into basin-hosted ore pressive subduction suture can lead to localized extension.
systems (Boiron et al., 2010; Deming, 1992; Koziy et al., 2009). Commonly, preexisting crustal structures act as a focus for
Under the term of rock deformation we subsume a range of magma ascent along a plate boundary and determine the
processes driving fluid flow at pressures that transiently exceed sites of major magmatic centers (e.g., four giant magmatic–
hydrostatic. These may be active from the upper crust down to hydrothermal gold and/or copper deposits in Papua New
mantle depths. In sedimentary basins, sediment compaction de- Guinea and Irian Jaya: Ok Tedi, Frieda River, Porgera, and
termines the transition from the region of hydrostatic conditions Grasberg). Hill et al. (2002) proposed that, during orogenesis
near the surface to overpressured fluids deeper down. A steep and crustal thickening in an overall compressive setting, strike-
gradient of increasing fluid pressure is typically located at an slip motion occurred along fracture zones roughly parallel to
impermeable layer separating the two hydrologic regimes. Basin the convergence vector. Localized zones of dilation opened
dewatering by rock compaction has been invoked as a possible where these fracture zones intersected major orthogonal faults
driver for the formation of hydrothermal ore deposits along basin and other structural discontinuities, facilitating igneous intru-
margins, for example, the Mississippi Valley-type (MVT) Pb–Zn sion and subsequent mineralization.
deposits (Figure 2(c)). However, these deposits show indications At the scale of individual deposits and orebodies, faults and
of rapid pulses of hot brine advection (e.g., Rowan et al., 2002), shear zones collect fluids from large source regions with dis-
and such thermal anomalies are difficult to reconcile with the persed fluid transport. In regional metamorphic terrains, oro-
slow process of basin dewatering, driven by gradual infill of genic gold vein deposits are localized by a continuum of lower
overlying strata (Cathles and Adams, 2005). More active forms crustal ductile shear zones to upper crustal brittle faults and
of rock deformation may be essential for driving effective flow of large veins (Goldfarb et al., 2005; Ridley, 1993; Ridley et al.,
ore fluids through sedimentary basins and underlying basement 1996). In sedimentary basins, stratabound (Figure 2(c)) and
rocks, such as orogen-related tilting or fault-related uplift of basin even synsedimentary-exhalative deposits (Figure 2(b)) are lo-
margins, which can build up a transient topographic pressure calized by faults acting as focusing channelways for hydrother-
head (Bethke, 1986; Leach et al., 2005). mal fluids ascending from subjacent basin fill or from
In the crystalline crust, seismic fracturing can lead to almost underlying basement (Leach et al., 2005).
instantaneous breaching of seals between an overpressured
regime at depth and a hydrostatic pressure regime above 13.1.3.1.3 Timescales of fluid flow and heat transfer
(Sibson, 1992). The resulting steep and transient gradients in Just as ore formation encompasses an interaction of processes at
pressure and temperature, and the possible confluence of different spatial scales, it also involves a range of characteristic
fluids from different chemical environments, favor the forma- timescales. These are governed primarily by rates of advective
tion of ore deposits. On the other hand, the quantity of avail- and diffusive heat transfer. Geodynamic processes creating en-
able fluid is limited by the accessible pore volume in the vironments for ore formation have characteristic lifetimes of
overpressured rock reservoir prior to the fracturing event. tens of millions of years, including basin formation, orogenic
Therefore, the formation of gold-bearing quartz veins from accretion and regional metamorphism, or the development of
metamorphic fluids by this mechanism typically requires lower crustal hydrous magma reservoirs in response to specific
Fluids and Ore Formation in the Earth’s Crust 11

plate interactions (e.g., England and Thompson, 1984; Rohrlach 13.1.3.2.1 Rock buffering and fluid-chemical master
and Loucks, 2005). Upper crustal magma chambers crystallize variables
more rapidly, typically within less than a million years, even if Any ore-forming hydrothermal transport system requires a
cooling occurs by conductive heat loss alone (Cathles, 1981). large volume of rock (or magma) from which initially dis-
Fluid-flow processes leading to formation of mineral de- persed ore-forming components are released. This source re-
posits are typically fast compared with basin evolution, oro- gion of the system requires dispersed fluid flow, enabling
genesis, and even the lifetime of a batholith-sized upper-crustal extensive contact and chemical reaction between the fluid
magma chamber (Cathles and Adams, 2005). Diagnostic evi- and the source material. Such conditions represent the normal
dence comes from observations that ore deposits are com- state of crustal fluids, in which the instantaneous ratio of fluid
monly transient thermal anomalies compared to the ambient to rock is small, due to the generally small porosity of rocks. To
geothermal gradient, even in areas without any magmatic heat a first approximation, we can consider the source region
input. For example, fluid inclusions in some sediment-hosted approaching a state in which fluid is chemically equilibrated
ore deposits record higher temperatures during pulses of with an infinite excess of the rock through which it flows.
hydrothermal Pb–Zn deposition, compared with ambient Equilibration is favored by the fact that dispersed fluid flow
rock temperatures recorded by the maturation level of organic means slow, local flow rates along grain boundaries and in
matter (e.g., Henry et al., 1992; Rowan et al., 2002). Such many small fractures. In a state of rock-dominated equilibra-
observations imply that advection of heat by ore-forming tion, the mineralogy of the rock is ‘buffering’ the chemical
fluid flow is fast relative to the conductive dissipation of heat master variables, imposing them on the fluid. These master
to the surrounding rocks, which places rigorous constraints on variables in turn determine the solubility of minor elements.
the duration of ore formation in sedimentary basins (1000– Solution pH and salinity are two related master variables
10 000 years; Cathles and Adams, 2005). Temperature gradi- affecting metal solubility. In fluids equilibrated with common
ents based on fluid-inclusion data and physical modeling of crustal rocks, pH is buffered by carbonate equilibria, or by
magmatic–hydrothermal ore systems indicate similar lifetimes hydrolysis reactions among alkali feldspars and sheet silicates
(Driesner and Geiger, 2007; Hayba and Ingebritsen, 1997; such as muscovite. The Al-conservative reaction (Al being the
Kostova et al., 2004), consistent with evidence from active least soluble element):
geothermal systems (Simmons and Brown, 2007). At the
scale of alteration zones around veins, chemical diffusion 1:5 KAlSi3 O8 þ Hþ ¼ 0:5 KAl3 Si3 O10 ðOHÞ2 þ 3 SiO2 þ Kþ [3]
through the pore fluid may be even faster than heat diffusion.
ksp þ Hþ ¼ mus þ qtz þ Kþ
The width of such haloes has been used by Cathles and
Shannon (2007) to argue that formation of veins, and possibly ! log Keq ¼ pH þ log aKþ
major magmatic–hydrothermal orebodies, may require as little
as 10–100 years (see also Ridley and Mengler, 2000, for alter- shows that, at given pressure and temperature, the pH of a fluid
ation haloes as shown in Figure 2(d)). in equilibrium with a mica-bearing granite assemblage will be
Such physical considerations lead to the conclusion that the inversely related to the Kþ activity in the fluid. The potassium
duration of ore-forming processes, that is, the times when ion activity, in turn, is approximately proportional to the Naþ
mineralization actively proceeds at the scale of an orebody, is activity in the fluid, by equilibration with sodic (albite) and
generally small compared with the precision of radiometric age potassic feldspar (ksp) components in the rock:
dating based on long-lived radioactive isotopes, for all but
the youngest or currently active systems (cf. Chapter 13.4). NaAlSi3 O8 þ Kþ ¼ KAlSi3 O8 þ Naþ [4]
U–Th disequilibrium dating of submarine magmatic–hydro-
thermal mineralization confirms a timescale of years to hun- alb þ Kþ ¼ ksp þ Naþ
dreds of years (De Ronde et al., 2005), consistent with the ! log Keq ¼ log aNaþ  log aKþ
transient dynamics of black smoker output recorded by direct
¼ logðaNaþ =aKþ Þ
fluid probing over a few years (Von Damm, 2005).
To satisfy charge balance, the total alkali concentration
13.1.3.2 Chemical Driving Forces for Hydrothermal Metal must be balanced by anions, dominantly Cl– in most hydro-
Enrichment thermal fluids. This leads to the important conclusion that, the
higher the chloride concentration of a granite (or arkose)-
Physical transport of heat and matter invariably leads to chem- buffered fluid, the higher will be its Kþ and therefore its Hþ
ical separation by fractionation of elements between solid activity (Montoya and Hemley, 1975; Yardley, 2005). In sim-
minerals and the transporting melt or fluid. Solubility gradi- ple terms: salty fluids are acid fluids.
ents along the fluid transport path are the chemical driving Analogous reactions buffer the acidity of fluids in clay-rich
force for metal enrichment. Metal solubility in hydrothermal but feldspar-free sediments, at distinctly lower pH, by equilib-
systems is typically low, but highly variable and element ria between muscovite (or illite) and kaolinite (kao):
specific (Brimhall and Crerar, 1987; Giggenbach, 1997; Reed,
KAl3 Si3 O10 ðOHÞ2 þ 1:5 H2 O þ Hþ ¼ 1:5 Al2 Si2 O5 ðOHÞ4 þ Kþ [5]
1997; Seward and Barnes, 1997; Chapter 13.2). Solubility
is controlled by pressure, temperature, and a few chemical mus ...... þ Hþ ¼ kao þ Kþ
master variables imposed by the major-element compositions
! log Keq ¼ pH þ log aKþ
of rocks and fluids, including ligand concentrations in the
fluid and mineral buffers in crustal rocks controlling pH and Equilibria [3]–[5] constrain the pH of common crustal
redox potential (Yardley, 2005). fluids from near-neutral to slightly acid conditions (whereby
12 Fluids and Ore Formation in the Earth’s Crust

neutral pH shifts from 7 at room temperature to 5.7 at 300  C; Kaolinite Mus Feldspars
–38
Chapter 13.2). With increasing temperature, the total acid

Pb(OH) 3- (aq)
concentration (Hþ þ HCl) as well as the potassium content of
-
rock-buffered salty fluids increases, also as a result of increased HSO4 PbCl2 Pb(OH) 2 (aq)
ion association, and both are approximately proportional to –6
–42
total chlorinity (Montoya and Hemley, 1975).
Redox potential is the third master variable, and is partly SO 42- HEMATITE

log m H2 (aq)
related to sulfur chemistry of fluids in equilibrium with

log f O2 (g)
–4
common crustal rocks. Redox potential in rock-dominated –46
hydrothermal systems is imposed by equilibria among Fe(II) PYRITE
and Fe(III) oxides, sulfides, and silicates (e.g., hematite, mag- Galena
netite, pyrrhotite, pyrite, chlorite, epidote, etc.). The predomi- –50
H2S (aq) –2

200 ppm
20 ppm
nance of oxides and silicates in rocks is the reason why oxygen MAGNETITE

2 ppm
fugacity ð log fO2 Þ is commonly used as a thermodynamic mea-
sure of redox potential, even though free oxygen is absent in the HS -
0
reducing interior of  the Earth. Hydrogen fugacity ð log fH2 Þ or
–54 PYRRHOTITE

activity log aH2 , aq is an alternative redox scale for water þ rock 150°C, 500 bar, aS,t = 0.001
systems, coupled to fO2 by the equilibrium H2O ¼ H2 þ 0.5 O2; 2 4 8 10
hydrogen indeed occurs in measurable concentrations in deep pH
crustal fluids (Giggenbach, 1992, 1997). Figure 4 pH-redox diagram of the H2O–Fe–S–O–H–(–Al2O3–SiO2–K–
Redox potential spans a great range even in common Na–Pb–Cl–C) system at 150  C, showing the stability fields of common
rock-dominated fluids. In systems containing reduced carbon iron oxides and sulfides in crustal rocks (green lines), assuming
(usually of biogenic origin: kerogen, graphite, liquid hydrocar- equilibrium with an aqueous fluid containing a fixed total concentration of
bon, and methane), fluids are too reduced to contain appre- sulfur equivalent to an activity of 103 mol kg1 of variable S species,
ciable sulfate under equilibrium conditions, and coexisting H2S, HS, HSO4, or SO42. The predominance of S species varies as a
Fe minerals are essentially ferrous. Because the Fe2þ/Fe3þ and function of pH and fO2 , as delimited by the dashed iso-concentration
S2/Sþ6 pairs have comparable redox potentials, hydrothermal lines. The main pH buffers in common crustal rocks among feldspars,
muscovite, and clay minerals (eqns [3]–[5]) are indicated by vertical grey
fluids in equilibrium with assemblages containing both Fe(III)
lines along the top axis of the diagram. The blue shadings show the
(epidote, hematite, and magnetite) and Fe(II) minerals (Fe–Mg
solubility of galena approximating 2, 20, and 200 ppm in a saline fluid
silicates like chlorite or amphibole) typically contain dissolved with aCl  1, depending on pH or redox potential according to the
sulfur in sulfide as well as sulfate form. This is reflected in changing sulfur species predominating in the fluid. Ignoring activity
Figure 4 by the green lines delimiting the hematite stability coefficients, 200 ppm Pb is approximately equivalent to the assumed
against the fields of pyrite and magnetite, straddling the pre- molar sulfur concentration, as required to precipitate the solubility-
dominance boundaries of HSO4 against H2S and HS limiting mineral, PbS, whereas lower Pb concentrations imply sulfur
(dashed black lines). The vicinity of the Fe2þ/Fe3þ buffer in excess in the solution. Calculated from thermodynamic data by Seward
rocks and the S2/Sþ6 redox couple in hydrothermal fluids (1984) and compilations by Holland and Powell (1998), Sverjensky et al.
means that the concentration ratios among sulfide, sulfate, (1997), and Dolejs and Wagner (2008).
and dissolved Fe2þ species commonly dominate the redox
balance within crustal fluids (Giggenbach, 1997). They are
expected to exert a dominant control on other redox-sensitive minerals limit ore metal solubility, increasing Hþ activity will
fluid equilibria involving trace elements of lower concentra- promote dissolution, for example:
tion (Drummond and Ohmoto, 1985). Very oxidizing condi-
tions high in the sulfate-only (fluid) and ferric minerals-only MnO þ 2Hþ ¼ Mn2þþ H2 O [6]
(rock) field originate from direct or indirect influence of the þ 2þ
PbS þ 2H ¼ Pb þ H2 S [7a]
oxygen-rich atmosphere. For example, oxygenated water flow-
ing into aquifers advects U and V toward zones of reduction in Most cations form chloride complexes in aqueous solu-
the subsurface, forming sandstone-hosted ore deposits. tions, which are increasingly stable with increasing tempera-
ture. Aqueous complex formation enhances metal solubility by
13.1.3.2.2 Sulfide solubility and base-metal content of orders of magnitude, as illustrated by the equilibria:
crustal fluids
Seward, Williams-Jones and Migdisov (Chapter 13.2) review PbS þ 2Hþ ¼ Pb2þ þ H2 S
[7b]
the thermodynamics of metal complexation and the solubility ! log Keq ð150  C; 500 barÞ ¼ 4:48
of ore minerals, based on extensive experimental research.
PbS þ 2Hþ þ 2 Cl ¼ PbCl2 þ H2 S
Such data are the chemical foundation for the following sec- [7c]
! log Keq ð150  C, 500 barÞ ¼ 1:43
tion, which emphasizes the effect of the hydrothermal master
variables (above) on the solubilities of base metals (Cu, Pb, Zn, defining the vertical isopleths of galena solubility in the H2S
also Mn, Co, and Fe) and the observed ranges of metal con- predominance field of Figure 4. Comparing the two equilib-
centrations in upper-crustal fluid–rock systems. rium constants indicates an approximately 1000-fold increase
pH and salinity have a mutually enhancing effect on the in lead solubility in a galena-saturated fluid with a chloride
solubility of many metals. Where sulfide, oxide, or silicate activity of 1 (approximating seawater salinity) compared to
Fluids and Ore Formation in the Earth’s Crust 13

pure water, at given sulfide concentration and pH. The effect of Temperature (°C)
complexing on solubility is compounded by the effect that 800 600 400 300 200 100 50
rock-buffered salty fluids are more acid than low-salinity
fluids, as increased Hþ activity additionally promotes sulfide
104
(and oxide and silicate) dissolution (reactions [3]–[5]).
Redox potential can affect metal solubility for two reasons.
First, it controls the solubility of metals with a difference in Ireland Northern
valence state between their solid and dissolved forms, as ex- 103 Arkansas

Pb concentration (ppm)
Tri-State
emplified by U and V in sandstone-hosted deposits, and also sph
by Fe, Cu, As, Mo, Sn, Au, Pd, Pt, Au, and other ore metals.

Ro
ck
Second, redox potential controls the proportions of oxidized 102

-bu
‘Anomalously’

ff
sulfate and reduced sulfide species in solution, which impacts Pb-rich ore fluids

ere
MVT and Irish

ds
on the solubility of all transition metals forming sparingly deposits

olu
qtz

bil
soluble metal sulfides including Fe, Co, Ni, Cu, Zn, Pb, and 10

ity
many others. Thus, the solubility of galena in the HSO4– and

tre
nd
SO42– predominance fields of Figure 4 is defined, respectively,
Experimental galena solubility
by the oxidation reactions: under rock-buffered conditions
1 Fluid inclusions in MVT

PbS þ Hþ þ 3 Cl þ O2 ¼ PbCl3 þ HSO4 [8a] Pb–Zn deposits – sphalerite


Fluid inclusions in MVT
Pb–Zn deposits – dolomite, quartz
PbS þ þ3 Cl þ O2 ¼ PbCl3 þ SO4 2 [8b] 0.1
Fluid analyses (recent brines and fluid inclusions)

0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035


even though Pb does not change its nominally divalent oxida- 1/T (K)
tion state. As a result, base metals are highly soluble under
oxidizing conditions, and can be co-transported with sulfur if it Figure 5 Lead concentration in selected crustal fluids and
rock-buffered solubilities of galena (PbS) as a function of temperature,
is present in sulfate form.
plotted on a 1/T scale. Experimental curves from Hemley et al. (1992) for
Aqueous sulfur chemistry is of particular importance to hy-
fluids with 1 m total Cl, buffered by excess feldspars þ muscovite þ
drothermal ore formation, not only as a solubility-limiting quartz þ pyrite þ pyrrhotite þ magnetite þ galena (red line; dashed
component for sulfide minerals, but also by acting as an extension by thermodynamic calculation from data cited in Figure 4).
acid/base buffer, as the dominant redox couple among species Small black symbols show analyzed Pb concentrations in natural
in many natural hydrothermal solutions, and as a complexing magmatic–hydrothermal, geothermal, and basin brines with inferred
ligand for some metals – as illustrated for gold transport in the temperatures of crustal rock equilibration, compiled by Yardley (2005).
following section. Blue symbols and boxes show fluid-inclusion analyses from
Sulfur isotopes most strongly fractionate between sulfate carbonate-hosted Pb–Zn ore deposits by Stoffell et al. (2008) and
and sulfide species, both in solution and among solid minerals Wilkinson et al. (2009), differentiating ore-stage fluids in sphalerite (solid
box, full circles) and intervening stages of gangue mineral precipitation
(Chapter 13.3). Variations in sulfur isotopic composition
(quartz, carbonates; dashed box and open circles).
therefore provide information about redox reactions in hydro-
thermal systems, to the point that great variability of d34S
among sulfides can be taken as first-order evidence for significant However, the scatter around the temperature trend
involvement of sulfate in an ore-forming process (Ohmoto, (Figure 5) also emphasizes that base metal concentrations
1986). vary up to three orders of magnitude at any given temperature.
Temperature is the main physical variable controlling min- This variation may be decisive for hydrothermal ore formation,
eral solubility in aqueous fluids. While carbonate and sulfate especially at low temperature. It is a result of variations in the
mineral solubilities decrease with increasing temperature other master variables, notably redox potential and sulfur con-
(Rimstidt, 1997), the solubilities of base metal sulfides and centration. Note that many low-temperature brines have Pb
metal oxides rapidly increase with rising temperature. The concentrations well above the rock-buffered solubility line
solubility of galena, sphalerite, and chalcopyrite, and also of (red). These include ‘anomalously’ Pb-rich solutions in fluid
cassiterite, Fe oxides, and Fe sulfides in saline crustal fluids inclusions hosted by sphalerite in MVT and Irish Pb–Zn de-
buffered by common rock-forming minerals, show an approx- posits, which contrast with relatively metal-poor fluids in as-
imately linear relationship with 1/T, as determined experimen- sociated gangue minerals of the same deposits (Stoffell et al.,
tally (Hemley et al., 1992), and found empirically by 2008; Wilkinson et al., 2009). Similar to some recent basin
compilation of natural fluid analyses. This is exemplified by brines (Kharaka et al., 1987), these ore-forming fluids must be
Figure 5, showing that Pb solubility in crustal fluids varies over sulfide deficient, either due to sulfate predominance under
at least five orders of magnitude, with a generally steep increase unusually oxidizing conditions at their source (Figure 4) or
with rising T, which becomes even tighter if metal concentra- as a result of a distinct chemical evolution path experienced
tions are normalized to chloride content of the fluids (Yardley, during cooling from higher temperatures. Initially hotter fluids
2005). Similar relations are found for Cu, Zn, Mn, and Fe, cooling in equilibrium with pyrite and Fe silicates may become
showing that temperature and salinity are the most decisive sulfide deficient and support correspondingly higher concen-
enhancers of metal solubility. To a first approximation, there- trations of Pb and Zn down to low temperatures. Such sulfur-
fore, any hot and salty crustal fluid can become an effective deficient brines are highly efficient metal carriers, but require
ore fluid (Yardley, 2005). an external sulfide source for ore deposition, for example,
14 Fluids and Ore Formation in the Earth’s Crust

preexisting sedimentary pyrite or H2S from biogenic sulfate lower salinity vapor (e.g., H2S–Au–SH NaCl; Zajacz et al.,
reduction (e.g., Fallick et al., 2001; Plumlee et al., 1994). 2010). With falling temperature, chloride species become
Pressure has a subordinate effect on equilibria involving unimportant except at extremely acid and oxidizing condi-
relatively incompressible aqueous liquids at temperatures tions, and give way to stable bisulfide complexes, AuHS0 and
well below the critical point of water (374  C). However, it Au(HS)2 (Stefánsson and Seward, 2004). The latter species
becomes a controlling variable in hot vapor-like fluids domi- allows up to tens of ppm of gold solubility in sulfur-rich
nating in magmatic–hydrothermal systems, as discussed in the fluids at near-neutral and moderately oxidizing conditions
section on magmatic–hydrothermal processes (below) and (Figure 6). In the H2S predominance field of aqueous sulfur,
explained in Chapter 13.2. gold solubility is defined by the equilibrium

13.1.3.2.3 Example: Hydrothermal redistribution of gold Au þ 2H2 S þ 0:25O2 ¼ AuðHSÞ2 þ Hþ þ 0:5H2 O [9]
Lead, zinc, copper, iron, and many other metals form stable
sulfide and oxide minerals that limit their solubility in crustal This reaction shows that the solubility of gold increases by
fluids, and complexation with Cl and possibly Br in saline oxidation of Au(0) to Au(þI), and is maximized by mildly
fluids is essential to allow adequate transport concentrations for alkaline pH conditions, opposite to the pH dependence of
hydrothermal ore formation. Other scarce metals and metal- most oxide and sulfide solubility reactions (eqns [6]–[8]). For
loids are dominantly transported as mixed-ligand or hydroxy a given total sulfur concentration in the fluid, gold solubility
species (e.g., H3AsO3, H3SbO3, H2MoO4, H2WO4, and their peaks at conditions where H2S, HS, and HSO4 have subequal
deprotonated anions) or in the form of stable complexes with concentrations (Figure 6). Gold solubility rapidly decreases
sulfur ligands (e.g., Au(HS)2, AuHS0, and possibly more com- again at more oxidizing conditions, where sulfide gives way to
plex species such as AuSO3HS0). Oxidizing conditions enhance sulfate species, which are poor complexing ligands for gold.
the solubility of these metals and metalloids, for example, as Gold deposition, like gold sequestering from source regions,
aqueous Au(þI), Mo(þVI), or As(þIII) species, relative to native can occur either by saturation with the pure metal (some
Au(0) or sulfide minerals such as Mo(þIV)S2 and FeAsS. Gold, dilution by Ag) or by extraction under conditions where gold
wolframite, and Mo, As, and Sb sulfide minerals are not known activity is significantly less than 1. Three mechanisms can shift
to be ubiquitous accessories in common rocks, so that the equilibrium [9] toward the left, and these characterize three
concentration of the respective elements in the fluid may not different environments of gold ore deposition. The first involves
be constrained by simple saturation equilibria. removal of H2S and consequent decomplexation of gold. This
Gold serves as an example of a metal that may commonly can occur by reaction with Fe-rich wall rocks to form pyrite
be leached from the source region of hydrothermal systems (a major mechanism in metamorphogenic gold-quartz lodes;
where its thermodynamic activity (i.e., saturation level) is less Phillips and Groves, 1983; Ridley and Mengler, 2000) or by
than 1. Gold has a crustal abundance of  3 ppb (Taylor and low-pressure boiling whereby H2S escapes into the gas phase,
McLennan, 1995). In common rocks, it can be hosted either as
metallic microparticles or as a trace component in sulfide solid
solutions such as pyrite or bornite (Candela, 2003; Simon et al., Kaolinite Muscovite Feldspars
2000). In natural silicate melts, gold is dissolved at activity levels 250⬚C, 500 bar, aS,tot=0.01
- AuOH
that are several orders below 1. Experimental measurements AuCl2
-30 -5
indicate that the solubility of gold in rhyolitic melts is between HSO4-
Hematite
0.3 and 1 ppm at the temperature, pressure, and chemical con-
ditions that occur in fluid-saturated upper-crustal magmatic SO42-
systems (Frank et al., 2002). This is 100–1000 times higher
-34 -3

log m H2 (aq)
AuHS
than its concentration in glassy melt inclusions of common
log f O2 (g)

magmatic rocks (1 ppb; Connors et al., 1993). The compari-


10

Pyrite
pp

son of experimental and microanalytical data implies that gold -


m

Au(HS)2
activity in natural magmas is on the order of 102–103. The
1p
pm

-38 -1
gold concentrations of magmatic and probably other high-
0.1

temperature crustal fluids are therefore not limited by metal Magnetite


pp

H2S (aq)
m

saturation, but rather by the degree of gold enrichment in


HS-
10

silicate melt and its partition coefficient between fluid and the
pp

Pyrrhotite
-42
b

melt or a solid-solution phase. Also in metamorphic fluid sys- 1


tems, where prograde devolatilization of metasediments or hy-
drated mafic rocks is believed to generate ore fluids for 2 4 8 10
metamorphogenic gold–quartz lode mineralization (Goldfarb pH
et al., 2005), gold solubility in high-grade metamorphic source
Figure 6 Aqueous Au solubility at 250  C and 500 bar (dashed
regions may be so high that the gold concentration in the fluid is
contours in ppm, assuming aAu ¼ 1) in aqueous fluids with a total sulfur
limited by gold availability, rather than by saturation with native activity of 0.01, overlain on predominance boundaries of aqueous sulfur
gold. species (dotted) and mineral stabilities in the Fe–S–O–H system (solid
At near-magmatic temperatures, gold is transported by lines). Calculated based on experimental data by Stefánsson and Seward
chloride complexes in brines (Gammons and Williams-Jones, (2004), Akinfiev and Zotov (2001, 2010) and thermodynamic data
1997) or by polyatomic alkali-chloride–gold-sulfide species in referenced in Figure 4.
Fluids and Ore Formation in the Earth’s Crust 15

forming epithermal gold veins (Drummond and Ohmoto, solubility in hydrothermal fluids and if metal activity is low-
1985; Spycher and Reed, 1989). Second, the reduction of a ered in solid solutions or a metal-undersaturated melt phase.
mildly oxidized gold-rich (e.g., magmatic) fluid by reaction Notably in magmatic systems, where the activity of economic
with organic carbon (or mixing with methane) can explain trace metals dissolved in melt is much lower than one, a pre-
gold precipitation, commonly in association with other redox- enrichment by fractional crystallization or an admixture of a
sensitive elements such as Sb or As. This mechanism is typical metal-rich melt phase would favor subsequent hydrothermal
for sediment-hosted gold deposits, from amphibolite-facies ore formation.
metasediments in the thermal aureoles of granitoid intrusions Quantitative studies of metal-source regions of specific
(Wall et al., 2004) to epithermal deposits of the Carlin type nonmagmatic ore systems are rare compared to studies of ore
(Cline et al., 2005; Muntean et al., 2011). In this environment, deposits (e.g., Heinrich et al., 1995, for Mt. Isa copper;
most of the gold can be captured in arsenious pyrite, whereby Wilkinson et al., 2005, for Irish Pb–Zn deposits). Even though
surface adsorption at conditions of native gold undersaturation we do not generally advocate the once-popular concept of
may prevail (Simon et al., 1999). As a third possibility, comp- pre-enriched ‘source beds’ as a first-order ingredient for ore
lete oxidation of a sulfide-rich black smoker fluid by mixing formation (Knight, 1957), there is a need for systematic in-
with oxygenated seawater can explain the deposition of gold in vestigations of source processes. Metal leaching is not deter-
volcanogenic massive sulfide (VMS) deposits (Large, 1992; mined by metal availability and bulk fluid–rock equilibria
Chapter 13.18). alone, but also by the mineralogical form of dispersed metals
affecting their kinetic reactivity at the grain scale. For example,
in the Canadian uranium ore systems, only the fraction of U
13.1.3.3 Metal Source Regions versus Ore Deposition Sites
contained in decomposing monazite of basement rocks under
Fluid–rock reaction in the rock-dominated source region of an the Athabasca basin is accessible to low-temperature leaching
ore-forming hydrothermal system contrast with that at the site (Hecht and Cuney, 2000). Similarly in low-temperature
of ore deposition, requiring a focused flux through a limited copper ore systems, the presence of reactive volcanic glass is
rock volume, resulting in fluid-dominated conditions. This probably more important than the slight enrichment of Cu in
imposes mass-balance constraints limiting the final grade of basaltic rocks compared to average crust (Chapter 13.10).
an ore deposit, depending on the reactions driving ore mineral
precipitation.
13.1.3.3.2 Ore deposition and wall-rock alteration
13.1.3.3.1 Source region characteristics Sites of ore deposition are characterized by fluid–rock reactions
The concentrations of Cu, Pb, Zn, and other scarce metals in in which integrated fluid flux is high compared with the quan-
common crustal rocks such as granite, shale, or basalt vary tity of reactive wall rock. Hydrothermal alteration along veins
within about one order of magnitude. This variation is small (zones of metasomatic reaction; e.g., Figure 2(d)) is the char-
compared with the enrichment factors of 100–10 000 required acteristic evidence that fluid chemistry imposed its chemistry
to convert a rock into an economic ore deposit (Table 1). Metal on host rocks that were initially out of thermodynamic equi-
solubilities vary over an even greater range, with experimental librium with the invading ore fluid (Reed, 1997). Ore-related
thermodynamic data and observations of natural brines indi- alteration usually leads to complete conversion of wall rocks to
cating variations in metal concentration ranging over at least a small number of new minerals, that is, a fluid-dominated
five orders of magnitude (Figure 5), or even eight orders if low- assemblage of high thermodynamic variance. The chemical dis-
salinity meteoric groundwaters are included. equilibrium between advecting ore fluid and local host rocks
The characteristics of source rocks impose chemical condi- can drive metal deposition by reducing ore mineral solubility.
tions that can enhance or discourage leaching of metals from For example, a solubility decrease may result from changing pH
them, but their original metal content is probably not the by alteration of feldspars to sheet silicates (reversing eqn [3] or
limiting factor in most instances. Considering mass-balance [5] by fluid cooling in magmatic systems, leading to phyllic or
alone, a small fraction of the crust affected by ore-forming ‘greisen’ alteration), or by dissolution of carbonate rocks to form
hydrothermal systems commonly contains enough metal for skarn deposits (Baker et al., 2004). Pyritization of Fe silicates in
even the largest ore deposits. For example, the basement be- mafic wall rocks removes sulfur from metamorphic ore fluids,
neath the Irish Midlands comprises some 100 000 km3 of which can destabilize gold-bisulfide complexes and lead to
upper-crustal basement rocks that were permeated by saline economic gold precipitation (Phillips and Groves, 1983; Ridley
brines during the late Paleozoic (Russell et al., 1981), yet a et al., 1996).
small fraction of a few 100 km3 of siliclastic metasediment is The efficiency of ore metal deposition is limited by the mass
sufficient to source the Pb and Zn accumulated in all economic balance between the reactive components in the advected fluid
Pb–Zn deposits in this province (Wilkinson et al., 2005). Sim- and the local rock. Final ore grade therefore depends not only
ilar order-of-magnitude considerations can be applied to on a thermodynamic driving force (a solubility gradient) but
sediment-hosted, orogenic, magmatic–hydrothermal, and also on the quantities of essential reaction partners in the
even orthomagmatic ore systems (e.g., Arndt et al., 2005). multicomponent fluid–rock systems. Even mineral precipita-
Thus, any medium-sized pluton of normal granodiorite con- tion by simple cooling is limited by the balance between steep
tains enough Cu to produce a large porphyry copper deposit temperature gradients (favoring localized ore deposition) and
(Dilles, 1987). heat advection by the fluid itself (opposing its cooling in rocks
Metal abundance at source may become decisive, however, of low thermal conductivity), as discussed in a classic paper by
where physical and chemical conditions allow very high metal Barton and Toulmin (1961).
16 Fluids and Ore Formation in the Earth’s Crust

There is ample evidence that large-scale hydrothermal regions of passive continental margins (stratiform deposits).
metal transport is common in the Earth’s crust, but only rarely Typical temperatures of ore deposition are between 50 and
leads to economic ore deposition. The following examples 120  C for sediment-hosted Cu deposits and replacive (MVT)
illustrate these constraints. The Salton Sea geothermal system Pb–Zn deposits, and may reach 200  C in stratiform Pb–Zn
contains millions of tonnes of Zn dissolved in a slowly convect- deposits.
ing brine plume, enough to form a world-class sediment-hosted Potential metal sources include immature clastic sediments,
zinc deposit, but neither thermal nor chemical gradients are basin-filling volcanics, as well as the fractured continental
adequate to cause localized deposition of sphalerite (Cathles basement underlying the ore-bearing sedimentary basin
and Adams, 2005; Williams and McKibben, 1989). In magmatic– (Boiron et al., 2010). Although continental redbeds and volca-
hydrothermal porphyry deposits, the concentrations of lead and nic rocks are a typical component of the sediment package
zinc analyzed in ore fluid inclusions are of similar magnitude as underlying Cu deposits, their hematitic mineralogy is a prod-
that of copper, but lead and zinc usually pass through the vein uct of ore-related fluid flow and not necessarily a source-rock
network without precipitation, for lack of a driving force for characteristic that differentiates them from the Pb–Zn deposits.
solubility reduction (Heinrich, 2006). In the porphyry environ- Even in high-salinity fluids, sulfide solubilities are so low that
ment, galena and sphalerite precipitate only where fluids come in adequate metal concentrations are difficult to achieve if the
contact with carbonate rocks that neutralize the fluids. Their fluids contain a significant amount of reduced sulfur
precipitation efficiency in carbonate-replacement deposits is ad- (Figure 4). A hydrological process driving fast fluid flow
ditionally limited by sulfur availability in the ore fluid seems essential for Pb–Zn deposits to become local or
(Catchpole et al., 2011; ’Cordilleran type,’ Table 2). Even when district-scale thermal anomalies (Cathles and Adams, 2005).
acid neutralization by feldspar alteration alone drives ore-metal Tectonic uplift of the basin margin, compaction dewatering,
precipitation, as in disseminated tin deposits (‘greisen-type’ ores tectonic pumping, or thermal convection through basin and
of cassiterite, SnO2), the maximum ore grade is limited to 1 wt basement have all been invoked, for both the Pb–Zn and the
% Sn, by the mass-balance between reactive feldspar in the rock Cu systems. A basic geochemical difference between Cu and
and the total concentration of acidity advected by the fluid Pb–Zn deposits seems to be related to the host-rock alteration
(Halter et al., 1996; Heinrich, 1990). associated with the ore-depositional environment. Pb–Zn de-
posits hosted by carbonate-bearing sequences show evidence
for extensive carbonate dissolution, indicating moderately acid
13.1.4 Hydrothermal Ore Formation in but not necessarily highly oxidizing ore fluids transporting Pb
Sedimentary Basins and Zn. By contrast, Cu deposits and their fluid pathways show
evidence of active footwall oxidation (‘red rock alteration’)
Mineral deposits in sedimentary basins, our main source of Pb by the metal-introducing fluid, even in higher tempe-
and Zn, occur as either MVT (replacing lithified carbonate-rich rature variants such as the native copper deposits of Michigan
rocks) or stratiform clastic sediment-hosted deposits (loosely (Chapter 13.10) or the hematite–carbonate fracture zones in
called ‘sedex,’ even though they are more commonly syn- metabasalts thought to feed the metasediment-hosted copper
diagenetic-hydrothermal rather than sedimentary-exhalative deposit at Mount Isa (Heinrich et al., 1995).
products; Leach et al., 2005). Basins also contain a significant New insight has come from analyses of fluid inclusions in
portion of the world’s Cu resources, including the single largest MVT ore systems, indicating that sphalerite deposition was
deposits of this metal, which exceed even the largest porphyry effected by pulses of anomalously metal-rich brines (Pb up to
copper deposits in grade and tonnage. Silver and cobalt are 200 ppm, Zn to 1000 ppm; Stoffell et al., 2008; Wilkinson et al.,
important by-products in many sediment-hosted copper de- 2009; Figure 5). Some systems show geological and isotopic
posits. Sedimentary basins also host the main U resources of evidence that the fluids permeated deeply through basement
the world, as exemplified by the sandstone-hosted U–V de- rocks beneath the ore-hosting sedimentary basin (Russell et al.,
posits discussed above and the giant unconformity-related 1981; Wilkinson et al., 2005). Based on thermodynamic equi-
uranium deposits of Proterozoic age (Chapter 13.19). librium calculations (Hennings et al., in press; Sverjensky, 1986;
In this section, we briefly speculate on the compositional Tornos and Heinrich, 2008), we speculate that sediment-hosted
separation of sediment-hosted Cu–(–Co–Ag. . .) and Pb–Zn– Pb–Zn deposits are formed from deeply circulating fluids buff-
(–Cd–Ag–F–Ba. . .) deposits. These two deposit groups are dis- ered by pyrite–magnetite assemblages in basement rocks, which
cussed in Chapters 13.9 and 13.10. Despite consistent differ- create mildly reduced brines with variable excess of Pb and Zn
ences in mineralogy and geological appearance, it is still not compared to their overall low sulfur content. As a result, Pb–Zn
clear why Pb–Zn and Cu  Co deposits are compositionally deposition in overlying sedimentary basins is dependent on a
distinct with only limited overlap (Gustafson and Williams, local sulfur source at the site of ore deposition, such as a H2S-rich
1981; Hitzman et al., 2005; Leach et al., 2005). gas trap (Jones and Kesler, 1992), or the biogenic or thermogenic
All sediment-hosted Cu and Pb–Zn deposits were formed in reduction of sulfate (Anderson, 2008; Thom and Anderson,
marine basins with a record of evaporitic sedimentation, indi- 2008; Wilkinson et al., 2005). By contrast, a decisive ingredient
cating hot arid conditions and providing a source of moderately for Cu–(–Ag–Co) ore formation may be a fluid that originates
to highly saline brines with high chloride for metal complexa- more directly from the oxygenated surface environment and
tion. Cu deposits are mostly associated with intracontinental maintains mildly oxidizing redox conditions, allowing effective
rifts filled with clastic sediments and volcanic rocks, whereas Cu(þI) transport even at low temperatures (see Figure 7 in
Pb–Zn deposits mainly occur in marine platform carbonates Chapter 13.10). Ore deposition in this case depends on an
of extensive intracontinental basins (MVT) or in shelf effective reductant (such as solid carbonaceous matter in black
Fluids and Ore Formation in the Earth’s Crust 17

shales or a mobile form of hydrocarbon; Hitzman et al., 2005), ore deposition, and may be key to the formation of large sheet-
usually assisted by preexisting diagenetic pyrite that contributes like orebodies from negatively buoyant brines settling in local
to Cu-sulfide precipitation based on textural and sulfur isotopic depressions on the ocean floor (Solomon and Quesada, 2003).
evidence (Wagner et al., 2009; Chapter 13.10). Indeed, the majority of large massive sulfide deposits that are
Iron oxide hosted Cu–Au–U–REE deposits and unconformity- currently exposed on land were formed in association with
related U–(Au–Pt–Pd) deposits (Table 2) share some ingredients submarine andesitic or bimodal volcanism, commonly close
with this interpretation of sediment-hosted copper ore formation. to acid volcanic centers. Arc-related hydrothermal systems gen-
In the first case, a magmatic heat source may be the driver for erally operate at shallower water depth compared to ocean ridge
an otherwise similar fluid-chemical process (Chapter 13.20), systems, which favors fluid phase separation and the generation
although other authors consider a magmatic fluid input to of hypersaline brines. Moreover, proximity to continents im-
the Cu Au budget as essential (Williams et al., 2005). proves the potential for preservation thanks to a protective
Unconformity-related uranium deposits are enriched in other sedimentary cover above newly formed deposits, and it increases
redox-sensitive elements including Mo, As, Se, Co, V, Au, Pt, the chance of later accretion, obduction, and eventual exposure
and/or Pd, which are, elements that also occur in sediment-hosted of deposits on land (Hannington et al., 2005; Scott, 1997). For
copper deposits. They co-precipitate with U by reduction to a geological as well as legal and political reasons, recent subma-
lower valence state, at a sharp reduction front between an oxidized rine sulfide deposits in back-arc settings are currently attracting
cover sequence (e.g., sandstone basins) and reduced basement the greatest attention from exploration companies.
rocks (e.g., carbonaceous phyllites), usually localized by post- The polymetallic nature of VMS deposits reflects the
depositional fault (Boiron et al., 2010; Cuney, 2010; Wilde and strong temperature dependence of sulfide solubility. Sharp
Wall, 1987; Chapter 13.19). temperature gradients at the seawater interface provide the
chemical driving force for wholesale precipitation of pyrite
and base metal sulfides to form high-grade deposits. At the
13.1.5 Hydrothermal Ore Systems in the Oceanic deposit scale, metal zoning within deposits confirms that
Realm temperature is the key factor controlling ore metal deposition,
with Cu enrichment in the hot feeder veins and near the center
VMS deposits are high-grade metal accumulations associated of mound-like orebodies and enrichment of more soluble
with submarine magmatic rocks. These mineral deposits have Pb and Zn closer to the cold seawater interface. The deposition
profoundly influenced prehistoric metal extraction technology of massive sulfide is further enhanced by kinetically suppressed
(e.g., Cyprus) and provided evidence for submarine hydrother- silica precipitation during rapid cooling and dilution of the ore
mal activity well before black smokers and active seafloor fluids by ambient seawater (Chapter 13.18).
hydrothermal systems were discovered (Stanton, 1955, and
earlier Japanese workers; see also Franklin et al., 2005; Han-
nington et al., 2005; Scott, 1997; Solomon and Walshe, 1979). 13.1.6 Magmatic–Hydrothermal Ore Systems
Today, ocean floor activity is recognized as the Earth’s most
important hydrothermal transport process, in terms of heat Magmatic–hydrothermal ore systems on land created our main
and mass transfer between the Earth’s interior and the hydro- resources of Cu, Mo, Sn, W, In, and Re and are a significant
sphere (German and Von Damm, 2003). The process is dom- source of Au, Ag, Pb, Zn, and other minor and rare metals.
inated by seawater convection, driven by the heat of generally Deposit formation involves a succession of processes, which
volatile-poor magmas emplaced at oceanic spreading centers. includes the partitioning of ore-forming components from a
Compared to their global geochemical impact on the compo- silicate melt into a saturating fluid phase as a key step
sition of the ocean water and the composition of the oceanic (Hedenquist and Lowenstern, 1994; Richards, 2011).
crust, the importance of mid-ocean ridge hydrothermal sys- Porphyry-type and epithermal Cu, Mo, and Au deposits are the
tems in generating economic ore deposits is rather small. Sul- focus of this section (Sillitoe, 2010). This family of ore deposits is
fides precipitated in black smokers are usually ephemeral, associated with hydrous magmas of intermediate silica content, in
because they are rapidly redissolved by today’s oxygenated contrast to magmatic–hydrothermal systems in highly evolved
seawater. The potential for preservation is further diminished granites giving rise to Sn–W deposits and pegmatite-related ores
by the low rate of sedimentation to cover and protect them, (discussed by Chapter 13.21). Figure 7 summarizes the geologi-
and the likelihood that they are subducted over geological cal context of porphyry-type and epithermal ore deposits (detailed
timescales (Hannington et al., 2005; also see Chapter 13.18). in Chapters 13.14, 13.15, and 13.16). A chain of processes in
Submarine hydrothermal systems related to arc volcanoes vertically extensive, crustal-scale magmatic–hydrothermal sys-
and back-arc basins are more important ore-forming environ- tems can explain the range of globally recurrent ore deposit
ments, for several reasons. Some show clear evidence for the types and their characteristic metal endowments (Figure 7).
involvement of magmatic fluids, in addition to convecting
seawater. Magmatic fluids added by subduction of hydrated
13.1.6.1 Melt Generation, Magma Storage, and Ascent
oceanic crust are likely to increase the efficiency of ore forma-
tion. With the presence of thinned continental crust, they Large magmatic–hydrothermal provinces typically comprise
contribute to the polymetallic composition of arc-related sul- several deposits of one or several types but similar age
fide orebodies (Pb, Au, Cu, and Zn; De Ronde et al., 2005; (Sillitoe, 2010; Wilkinson and Kesler, 2009). These ore prov-
Hannington et al., 2005; Yang and Scott, 1996). Higher salinity inces reflect tectonomagmatic settings giving rise to hydrous
fluids (e.g., of magmatic origin) also affect the hydrology of VMS (usually amphibole-phyric) magmas that invariably have a
18 Fluids and Ore Formation in the Earth’s Crust

High-sulfidation Au–As–Cu
Fluid focusing and cooling,
0 Low-sulfidation Au–Ag hydrothermal ore deposition
Porphyry Cu ± Au ± Mo
Carlin Au–As Brine – vapor phase separation
Cordilleran polymetallic
Fluid exsolution and
Skarn Cu metal transfer melt > fluid
10 km Mid- to upper-crustal magma
reservoir (± magma mixing and
transient sulfide melt exsolution)

20

30
Crust Lower-crustal magma storage,
high-pressure fractionation
± crustal melting, assimilation
Mantle
40

Hydrous mafic melt


generation and ascent

Figure 7 Schematic illustration of the geological components, ore types, and processes in crustal-scale magmatic-hydrothermal systems, combining
aspects from Dilles (1987), Hedenquist and Lowenstern (1994), and Hill et al. (2002). Porphyry copper deposits form in a dense network of
hydrofractures (red) around dikes and stocks above the roof of hydrous magma chambers, by precipitation of sulfides and gold from single phase or
more commonly two-phase fluids of coexisting hypersaline liquid (brine) and vapor. Au-rich varieties predominate at shallow levels (1–3 km below
surface) while porphyry deposits formed as deep as 8 km are generally Cu  Mo dominated and Au-poor. Copper  Zn  Au skarns, Cordilleran vein and
replacement Cu–Zn–Pb–Ag  Au deposits, and Carlin-type Au deposits rich in arsenious pyrite are hosted by carbonate-bearing sedimentary host rocks
controlling pH by carbonate dissolution; they form at decreasing temperatures and increasing distance from magmatic intrusions. Carlin-type gold
deposits are additionally characterized by abundant organic carbon in the host sediments. Their formation temperature and the low salinity of ore fluids
with liquid-like density overlap with fluid properties in other epithermal precious-metal deposits. High-sulfidation epithermal deposits are characterized
by intense acid leaching and sulfate and clay alteration caused by hot low-density magmatic vapor, typically preceding the main stage of ore deposition
by low-salinity aqueous liquid. Low-sulfidation epithermal deposits are generally more distal to intrusions and are characterized by near-neutral fluids
causing feldspar–muscovite alteration, with a greater proportion of convecting meteoric water and a less obvious magmatic fluid input.

subduction signature in their trace-element composition, in- Even along arcs with long-lasting subduction magmatism,
cluding variable enrichment in mantle-fluid elements such as the formation of major deposits is typically confined to limited
Pb, Ba, Rb, and Sr and depletion in high-field strength ele- segments and short time periods of a few million years or less.
ments such as Ti, Zr, and Y (Candela and Piccoli, 2005; Structural evidence and plate-scale tectonic reconstructions in-
Richards, 2003; Rohrlach and Loucks, 2005). Magmas produc- dicate that segments and periods of successful ore formation are
ing major porphyry and epithermal ore provinces in the characterized by a compressional state of the crust and litho-
Circum-Pacific and Alpine–Himalayan belts are driven by con- sphere, favoring the formation of large fractionating magma
current subduction of oceanic lithosphere beneath an active chambers in the lower crust and preventing volcanic eruption
continental margin (e.g., Chile–Peru, Philippines, Balkan– and uncontrolled loss of volatiles in the upper crust (Rohrlach
Iran). However, a growing number of world-class ore provinces and Loucks, 2005; Sillitoe, 2010). Empirically, most productive
are recognized to owe their subduction flavor to more ancient porphyry ore provinces are driven by magma series exhibiting
events, which chemically prepared the lithospheric mantle for an ‘adakite-like’ trend of increasing Sr/Y and depleted heavy
later hydrous melting driven by extensional tectonics REE (HREE) (Richards, 2011). This geochemical signature is
(Richards, 2011). For example, Eocene magmas forming Bing- not caused by slab melting but reflects either assimilation of
ham Canyon and the world’s premier molybdenum-porphyry thickened garnet-bearing lower crust in areas of compressive
province in the southwestern United States were interpreted to flat-slab subduction (Kay and Mpodozis, 2002), or storage of
be fertilized by Proterozoic mantle metasomatism (Pettke mantle melts in large lower crustal sills leading to high-pressure
et al., 2010). Incipient extension of this old and multiply fractionation (amphibole-dominated but plagioclase-absent;
accreted craton may also have contributed to the formation Richards, 2011; Rohrlach and Loucks, 2005). Both processes
of roughly coeval gold deposits in the Great Basin, including are favored by compressive horizontal stress in the lithosphere.
Carlin (Emsbo et al., 2006; Muntean et al., 2011). The world’s Whether source composition directly affects the later min-
premier tin province in Bolivia and southern Peru is character- eralization potential of magmas is unclear, although it is com-
ized by repeated partial melting of reduced sediments, produc- monly implied. Large global metal provinces such as the Mo
ing magmas from which Sn(þII) complexes are readily province in the southwestern United States are suggestive of
partitioned into fluids for later precipitation as hydrothermal large-scale metal pre-enrichment in the lithospheric mantle or
Sn(þIV)O2 (cassiterite; Lehmann et al., 1990). lower crust, but limited evidence from bulk compositions of
Fluids and Ore Formation in the Earth’s Crust 19

magmas or melt inclusions has not clearly separated source availability (Candela and Piccoli, 2005), and relatively dense
effects from metal enrichment due to fractional crystallization hydrous fluids separating from magma at pressures >1kbar are
of the magmas (Chapter 13.6). In porphyry-copper minerali- most likely to contain high Cl, S, as well as high ore metal
zing magma suites, copper as well as sulfur concentrations are concentrations (cf. Cline and Bodnar, 1991). In sulfur-rich
generally highest in the most mafic and mantle-dominated mafic magmas, the transfer of Au and possibly Cu to the fluid
magmas and melt inclusions, and become depleted by frac- is enhanced by sulfide complexation, compounded by elevated
tional crystallization or mixing with more felsic magmas alkali concentration in S-rich fluids (Zajacz et al., 2010). Ini-
(Halter et al., 2005; Hattori and Keith, 2001). Copper and tially oxidized sulfur in the melt may be reduced to H2S during
probably gold are largely incompatible with magmatic silicate early magnetite saturation, thereby contributing to subsequent
and oxide phases and can become enriched in acid residual gold extraction into the magmatic fluid (Sun et al., 2004).
melts (Mustard et al., 2006). However, they strongly fraction- Fluid inclusions at the base of major porphyry copper de-
ate into any solid or molten sulfide phase (Halter et al., 2005; posits have intermediate density and low-to-moderate salinity,
Jugo et al., 1999; Li and Audétat, 2012), which explains the contain high S  CO2 concentrations, up to weight percent
progressive Cu–Au depletion seen in most fractionating levels of copper and ppm levels of Au (Seo et al., 2009).
magma suites. Unusually oxidizing conditions in the melting However, the relative importance of Cl and S in generating
region created by subduction fluids (Kelley and Cottrell, 2009) the most effective ore fluids is still not established by quanti-
will increase the sulfate/sulfide ratio and total sulfur solubi- tative experiments (Pokrovski et al., 2008). One of the causes
lity (Carroll and Rutherford, 1985), which could be a decisive for uncertainty stems from recent experimental evidence that
factor in generating S- and metal-enriched (especially Au-rich) Cu in natural quartz-hosted fluid and melt inclusions from
but sulfide-undersaturated silicate melts favoring later Cu and deep high-temperature environments may be subject to selec-
Au accumulation (Hattori and Keith, 2001; Mungall, 2002; tive postentrapment diffusional exchange for outward-
Richards, 2003; 2011). diffusing Hþ ions (Li et al., 2009b; Zajacz et al., 2009).
Inclusion evidence for direct separation of a Cu-rich hyper-
saline brine or salt melt from silicate melt has been recorded
13.1.6.2 Fluid Exsolution and Element Transfer from from several porphyry copper deposits (e.g., Harris et al.,
Melt to Fluid 2003). We interpret such brines as a local product of the
shallow-level emplacement of volumetrically small porphyry
Magmatic–hydrothermal ore formation depends on the genera-
stocks, rather than the more important source process in dee-
tion of a large and focused flux of metal- and sulfur-rich magmatic
per magma chambers, generating a large quantity of low-
volatiles to the upper crust. For selected porphyry copper de-
salinity fluid that drives economic ore formation.
posits, the total metal resource, the concentration of ore metals
The physical process and the rate of fluid extraction from an
in magmatic ore fluids, and plausible assumptions about the
upper crustal magma reservoir are critical for fluid focusing and
original content of metals in the magma allow a mass-balance
ore formation (Candela, 1991; Candela and Piccoli, 2005;
estimate of required quantities of source magma and fluid
Giggenbach, 1997). Fluid extraction may occur by dispersed
(Cloos, 2001; Dilles, 1987; Steinberger et al., 2013; Ulrich et al.,
bubble exsolution and ascent through large parts of the in-
1999). Minimum required magma volumes typically range from
trusions, driven by inward crystallization of the intrusions and
a few tens to 1000 km3. These estimates indicate that the volume
progressive volatile enrichment in the melt as less hydrous
of normal mid- to upper-crustal plutons is adequate for forming a
minerals crystallize (‘second boiling’; Candela and Holland,
world-class ore deposit. In some cases, the pluton dimensions can
1986; Burnham and Ohmoto, 1980; Cline and Bodnar, 1991).
be confirmed by geological mapping (in areas exposed by later
Alternatively, fluid may exsolve and collect in cupolas at the
tilting; Dilles, 1987) or is indicated by large magnetic anomalies
highest point of the intrusion as a result of internal convection
(Krahulec, 2010). Such evidence, and high concentrations of ore
in the magma chamber (Cloos, 2001; Shinohara et al., 1995),
metals in magmatic fluid inclusions (Audétat et al., 2008; Eastoe,
prior to almost simultaneous emplacement of porphyry stocks,
1982), confirm that fluid saturation from an upper crustal
their fracturing, and mineralization by near-explosive fluid
magma chamber is a key step in large-scale metal enrichment.
expulsion toward the surface (Cathles and Shannon, 2007;
Separation of a minor amount of CO2-dominated fluid
Dilles, 1987). Reconciling high-precision geochronology with
from ascending volatile-rich magmas may start at lower to
physical modeling of these alternative or possibly concurrent
mid-crustal depth (Carroll and Webster, 1994; Lowenstern,
processes, within the thermal constraints on the rates of magma
2001; Zajacz and Halter, 2009). Copious fluid saturation
chamber evolution, porphyry emplacement, and ore formation,
sets in with weakly CO2 bearing and moderately saline aqu-
is an important challenge for current research (cf., Driesner and
eous fluid of intermediate (‘supercritical’) density, as indic-
Geiger, 2007; Ingebritsen et al., 2010; Von Quadt et al., 2011).
ated by experimental data and derived modeling (Candela
and Holland, 1984; Candela and Piccoli, 2005; Cline and
Bodnar, 1991) and limited fluid-inclusion observations from
13.1.6.3 Fluid Cooling, Decompression, and Phase
deeply exposed plutonic environments (Audétat et al., 2008;
Separation
Hennings et al., in press). These deep single-phase aqueous
fluids also contain variable H2S and SO2, depending on water The subsolidus evolution of magmatic fluids can be discu-
pressure and redox potential of the source magma (Carroll and ssed by reference to the NaCl–H2O model system (Figure 8;
Webster, 1994; Giggenbach, 1997). Partition coefficients for Hedenquist and Lowenstern, 1994; Henley and McNabb,
metals forming stable chloride complexes increase with Cl 1978; Williams-Jones and Heinrich, 2005).
20 Fluids and Ore Formation in the Earth’s Crust

1200
NaCl–H2O phase stability
1000 and fluid evolution paths

Pressure (bar)
800 d
Fluid properties
600 c H2O Salinity NaCl

L
b
400 Halite
L

Density
200
a Unstable
0 0.32 V in upper crust
600
Tem 400 80 100
pe 200 60

V
rat 20 40
ure 0
0 Cl
(⬚C
) Wt% Na

Figure 8 Phase diagram for NaCl–H2O–P–T from Driesner and Heinrich (2007), showing the limiting two-phase liquid þ vapor surface (black), the
halite saturation surface (green), and the composition-dependent critical curve (red). Four possible fluid evolution paths of a cooling low-salinity
magmatic fluid are indicated by arrows, emphasizing the effect of pressure upon phase state and density change. Degrees of darkness and blue color
schematically indicate fluid density and salinity, respectively, varying from low-density vapor (white) to dense saline brine (dark blue) and low-salinity
aqueous liquid (black; inset right). Modified from Williams-Jones AE and Heinrich CA (2005) 100th Anniversary Special Paper: Vapor transport of metals
and the formation of magmatic-hydrothermal ore deposits. Economic Geology 100: 1287–1312. With permission from the Society of Economic
Geologists.

At low pressures (at most  400 bar), essentially salt-free complex function of the concentration of chloride, sulfide, as
vapor coexists with solid halite (region bounded by heavy well as alkali concentrations in low-density aqueous gases. There
green lines in Figure 8). Phase relations and density variations is little doubt however that metals are effectively transported by
above the vapor þ halite region are characterized by a variable vapor, and that their solubility in vapor increases with increasing
miscibility gap (black net in Figure 8) separating low-salinity vapor density (and therefore pressure), because hydration stabi-
vapor of variable density from saline liquid that may reach lizes metal species in the vapor phase (Williams-Jones and Hein-
halite saturation at 25% to >60% NaCl, depending mainly rich, 2005).
on temperature. The critical curve for binary liquid þ vapor
(i.e., the crest of the miscibility gap indicated by the red line
13.1.6.4 Fluid Evolution Paths at the Porphyry to
in Figure 8) extends from the critical point of water to higher
Epithermal Transition
pressures and temperatures. It also swings to more saline
compositions, allowing vapor at elevated pressure and near- Distinct deposit types found in magmatic–hydrothermal sys-
magmatic temperatures to reach salinities up to  20% salt. tems, their fluid-inclusion characteristics and the temporal
The presence of minor CO2 and H2S will shift the miscibility sequence of mineralization can be explained, to a first order, by
gap toward higher pressure but does not change the topology different density–salinity paths of fluids moving through P–T
of the phase diagram. space during ascent from their magmatic source towards the
Regarding minor element concentrations, there is an intrigu- Earth’s surface. Fluid chemistry and ore precipitation are con-
ing difference between careful analyses of coexisting vapor and trolled by cooling and decompression, as fluids are expelled
brine fluid inclusions from porphyry-type deposits, and well- from lithostatic conditions in the magma to hydrostatic
controlled experiments carried out under magmatic conditions. conditions closer to the surface, where they can interact with
Microanalysis of coexisting natural brine and vapor inclusions meteoric water to variable degrees (Fournier, 1999; Weis et al.,
has shown that K, Fe, Mn, Zn, Pb, and minor salt components 2012). The depth of fluid exsolution from magma is likely to
fractionate with NaCl into the liquid phase, whereas B, S, Cu, As, be the key to different paths of fluid evolution.
Sb, and Au are enriched to variable degrees in the coexisting If hydrous magma is emplaced close to the Earth’s surface,
vapor inclusions (Baker et al., 2004; Heinrich et al., 1999; Seo for example in the neck of a volcano, high-temperature fuma-
et al., 2009). Experiments at 800  C and pressures of 1–1.4 kbar, rolic activity will be dominated by low-pressure vapor and salt
in the presence of melt, show that Cu and Au concentrate in the sublimates, as indicated by path (a) in Figure 8 and shown as a
brine relative to the vapor, with relative partitioning into the schematic physical representation in Figure 9(a). The vapor
vapor enhanced by the presence of reduced sulfur (Frank et al., has such a low density that metal transport is minimal and only
2011). Experiments at subsolidus conditions show that copper contains acid-generating volatiles including SO2, H2S, and CO2.
can partition significantly to the vapor in the presence of sulfur If condensed or mixed with surface water, such fluids will cause
(Pokrovski et al., 2008), but the extreme copper enrichment barren acid alteration or, at slightly greater depth inside a vol-
found in natural vapor inclusions has not yet been reproduced cano, may produce sulfide-free porphyry gold deposits by
by well-controlled experiments and is probably a result of post- precipitating precious metal with magnetite (Koděra et al.,
entrapment modification of fluid inclusions (Li et al., 2009b; 2011; Muntean and Einaudi, 2000; Vila and Sillitoe, 1991).
Zajacz et al., 2009). Experiments at 1000  C and 1.5 kbar (Zajacz Fluid exsolving from a magma chamber at a few kilometers
et al., 2010) indicate that copper and gold solubility are a depth will generally be a single-phase fluid of intermediate
Fluids and Ore Formation in the Earth’s Crust 21

Fumarole vapor Vapor expansion Boiling liquid Boiling liquid and low fluid pressure (Cox and Singer, 1988; Murakami
± above vapor contraction single-phase
salt sublimate brine condensation brine condensation fluid contraction
et al., 2010).
The intersection of the two-phase curve at higher pressure
(path c in Figure 8) generates a denser vapor that, after physical
separation from brine and cooling at elevated pressure, may
again leave the two-phase surface and subsequently cross iso-
chemically above the critical curve of the systems. Such a
physical process, indicated in Figure 9(c), can result during
the later stages of cooling of a large magmatic–hydrothermal
system (Driesner and Geiger, 2007) and has important geo-
chemical consequences. The vapor phase is relatively depleted
in Fe, one of the dominant salt components fractionating into
the residual brine. This Fe depleted but likely S-enriched vapor
phase is ideally suited to transport high concentrations of
S-complexed gold, in a fluid phase that becomes gradually
denser and more liquid-like upon cooling (vapor-to-liquid
contraction; Heinrich et al., 2004). Sulfur excess over chalco-
Shallow Magma interface Deep
phile metals (Cu and Fe) will lead to selective deposition of
(a) (b) (c) (d)
Cu–Fe-sulfides in gold-poor porphyry copper deposits at high
Figure 9 Schematic representation of alternative physical evolutions temperature and greater depth, but permit effective transfer of
of a low-salinity magmatic fluid ascending from variable depths of its sulfur-complexed gold in the vapor-derived aqueous liquid
hydrous magma source, corresponding to the four P–T–XNaCl paths toward shallower environments of epithermal ore formation
indicated in Figure 8 and using the same color shading for density
(Chapter 13.15).
(darkness) and salinity (blue). Cooling between 500 and 300  C is
schematically indicated by the red and brown isotherms (dotted lines)
The magmatic fluid may also rise and cool entirely within
to illustrate the alternatives of magmatic vapor expansion in hot and the single-phase stability field, if fluid separation from the
shallow systems (a and b, with or without brine condensation at depth), magma occurs at depth and if the fluid is cooled sufficiently
in contrast to vapor contraction when the low-salinity fluid cools at at elevated pressure (Hedenquist et al., 1998), as indicated by
pressure in deep magmatic-hydrothermal systems (c and d). Note that path d in Figure 8 and the sketch in Figure 9(d). In this case,
magmatic vapor that initially condensed out a brine at depth can the compositional evolution will be dominated by mineral
subsequently contract to aqueous liquid that may eventually boil at precipitation and alteration reactions. Such a path explains
shallow pressure (c; compare path c in Figure 8). Constrictions in the the dense fluid inclusions in deep porphyry deposits such as
fluid columns indicate the transition from lithostatic fluid pressure in and Butte (Rusk et al., 2008) and the characteristic liquid-
above the magma chamber to hydrostatic pressure in veins extending to
dominated inclusion assemblages of Cordilleran-type ore sys-
the surface. This transition occurs near the brittle-to-ductile transition of
rocks, depending in turn on magma depth and heat advection by focused
tems and their typical camp-scale ore zonation from proximal
fluid ascent. The distinct fluid evolutions may also succeed each other in Cu through Pb–Zn to distal veins and replacement orebodies
time, as a magmatic system cools and the magma front retracts to containing Ag, Au, As, and/or Sb (Bendezu and Fontbote,
greater depth. 2009; Catchpole et al., 2011; Simmons et al., 2005).

density and 2–13 (average 5) wt% NaCleq salinity (Audétat


13.1.6.5 Wall-Rock Alteration and Ore Mineral Precipitation
et al., 2008). Depending on its P–T path during ascent, the
fluid may intersect the two-phase surface and separate into Geochemical characteristics of wall rocks at the ore deposition
hypersaline liquid (brine) and lower salinity vapor. If the site and the resulting alteration reactions finally shape the
pressure drop is sufficient, the fluids may additionally saturate diversity of ore deposit types and their metal endowment in
with halite (solid NaCl; reaching the green curves in Figure 8), porphyry–epithermal systems (Giggenbach, 1997; Reed, 1997;
as indicated by enrichment of KCl in residual brine inclusions Sillitoe, 2010). With a given input fluid composition, iron-rich
of some porphyry deposits (Cloke and Kesler, 1979). If the mafic to intermediate intrusions and volcanic host rocks, or
initial salinity of the single-phase fluid is less than 10%, the their previously magnetite-altered counterparts, are particu-
two-phase curve is likely to be intersected on the vapor side of larly efficient at causing wholesale Cu–Fe sulfide precipitation
the critical curve (path b in Figure 8), so that a volumetrically together with Au in porphyry Cu–Au deposits, due to reduc-
small proportion of brine condenses from a predominant flux tion of SO2 to H2S in the fluid and effective desulfidation of
of vapor (Figure 9(b)). Such an evolution toward vapor ex- sulfide-complexed metals (e.g., Arancibia and Clark, 1996;
pansion is recorded by fluid inclusions in the central upflow Ulrich and Heinrich, 2002). By contrast, Fe-poor quartzo-
zone of the zoned Bingham Canyon porphyry Cu–Au deposit feldspathic and/or carbonate-rich sedimentary rocks at the
(Landtwing et al., 2010). The steep P–T gradient resulting from deposition site are more conducive to precipitation of base
vapor expansion is expected to force wholesale precipitation of metal sulfides, while Au stays in solution because acid neutral-
Cu together with any dissolved Au, due to the destabilization ization by felspar-to-muscovite alteration or carbonate disso-
of hydrated metal complexes transported by the vapor (Zezin lution tends to increase gold solubility (eqn [9]; Figure 6).
et al., 2011). This process may explain why most Au-rich Such gold remains transportable to low temperatures, until
porphyry copper deposits were formed at shallow depths the fluid reaches a depositional trap, for example, in
22 Fluids and Ore Formation in the Earth’s Crust

sedimentary host sequences rich in organic reductant (e.g., deposits (phosphorites) form by biogenic carbonate–fluorapa-
Carlin trend; Muntean et al., 2011). Local saline fluids in tite precipitation at the sediment–water interface in marginal
equilibrium with such reduced host rock packages will be Fe2þ basins and shelf regions, where deep ocean water is upwelling
rich, and their mixing with moderately oxidized sulfur and gold- over extended periods (Chapter 13.12). Lower oxygen levels in
rich fluids from a deeper magmatic or possibly metamorphic the Archean atmosphere kept ocean waters anoxic, such that
source will lead to co-precipitation of Au  As with abundant Fe2þ from ocean-floor hydrothermal activity or continental
pyrite (Kesler et al., 2005). Alternatively, focusing of similar sources remained soluble, up to the Paleoproterozoic when
fluids toward the surface can form classic low-sulfidation the partial pressure of oxygen in the atmosphere and in near-
epithermal veins. These typically show vertically restricted surface ocean waters rose with the onset of photosynthetic
bonanza-grade gold  silver zones, resulting from phase separa- metabolism. Banded iron formations were deposited as chem-
tion and partitioning of the gold-complexing sulfide ligand into ical sediments, mostly in the period between 2.5 and 2 Ga
the low-density vapor phase (Drummond and Ohmoto, 1985; (Chapter 9.18). They probably formed on shelf regions of
Henley and Berger, 2000; Simmons et al., 2005; Spycher and protocontinents in stratified ocean basins with anoxic bottom
Reed, 1989). water. These silica and Fe-rich chemical sediments provided
In summary, magmatic–hydrothermal systems are particu- the first step of large-scale iron accumulation, followed by
larly effective engines for ore metal enrichment, because they secondary upgrading by meteoric water or low-temperature
comprise several steps of element separation among rocks, hydrothermal processes to form the most important iron ores
melts, and multiphase fluids, in an environment creating mined today.
steep gradients in pressure, temperature, and chemical condi- Evaporation of surface waters produces a range of salt de-
tions between hydrous magma and the Earth’s surface. The posits, depending on the starting composition and evaporation
broad range of deposit styles can be rationalized by a limited paths (Chapter 13.22). Evaporite differentiation is one of the
number of physical variables, which provide for ‘chemical di- classic examples of ‘chemical divides’ explored by a pioneering
vides’ leading to distinct metal endowments. study of multicomponent thermodynamic reaction modeling
(Eugster and Hardie, 1975; Hardie and Eugster, 1970). Pro-
gressive evaporation of seawater, or of continental salt lakes
13.1.7 Ore Formation at the Earth’s Surface dominated by marine aerosols, successively deposits gypsum,
halite (NaCl), and eventually potash (K) and magnesium salts.
In this final review section, we introduce the processes forming Continental surface waters with an initially small excess of
important resources at the Earth’s surface and at the ocean bicarbonate compared with earth alkalis (Ca and Mn) will
floor (Figure 1(c); Table 2), where atmosphere and hydro- never saturate with gypsum and may evolve towards precipita-
sphere interact directly with the uppermost part of the crust. tion sodium carbonate salts, such as trona (Na3(CO3)(HCO3)
As in hydrothermal ore deposits, the controlling factors are 2H2O; a strategic resource used for alumina beneficiation from
the structure and surface properties of minerals, element bauxite). Lithium and borate salts are produced by evaporation
compatibility, and fluid speciation controlling selective disso- of continental surface waters with a volcanic input.
lution and precipitation. However, the processes grouped to- On continental land surfaces, ore formation can be driven
gether here are more directly driven by climate, atmosphere by weathering, again depending on the oxygen level of the
composition, and the movement of open ocean and surface atmosphere. Especially in humid (sub-)tropical climates,
water as well as biological activity. Not surprisingly then, major weathering creates residual ore deposits in lateritic soils.
ore deposit types in this group provide some of the most Bauxite is our main aluminum ore and forms by complete
compelling evidence for significant variation in the Earth’s hydrolysis of primarily Al-rich rocks such as anorthosites or
surface conditions, notably the increase of oxygen content in extensively soluble clay-bearing carbonate rocks (Freyssinet
the atmosphere during the late Archean (Holland, 2005). et al., 2005). Lateritic deposits of Ni Co form by liberation of
In the oceanic domain, major metal resources are formed the metals from olivine in ultramafic rocks, contributing signif-
from seawater without local hydrothermal contribution, even icantly to global Ni and Co supply besides sulfide deposits. At
though the metals may originally have been leached by sub- Phanerozoic O2 levels similar to today’s, supergene enrichment
seafloor hydrothermal activity. In modern oxic oceans, man- of primary sulfide deposits had made the first ore deposits
ganese released by continental weathering black smokers does accessible to metal mining by prehistoric civilizations, because
not precipitate as sulfide in VMS deposits, but is dispersed it could form native copper by oxidation of copper sulfide and
to distal chemical sediments by basin-scale ocean currents. removal of the soluble sulfate. Secondary upgrading is a major
Major manganese deposits mined today resulted from global economic factor for many modern porphyry copper deposits. It
events of changing ocean composition in combination with occurs by in situ vertical or small-scale lateral metal re-
sea-level changes, affecting transgression and regression in distribution, as a result of surface weathering (dissolution of
restricted marginal-marine basins (Frakes and Bolton, 1992; primary Cu sulfides in the unsaturated zone) and re-precipitation
Chapter 11.14). Ferromanganese nodules and crusts are pre- of copper oxides, carbonates, silicates, or secondary sulfides
cipitated directly from ambient seawater. They constitute an just below the water table (e.g., chalcocite Cu2S precipitation at
enormous untapped metal resource not only of Mn, Co, Ni, the contact to underlying less Cu-rich sulfides such as chalcopy-
and Cu, but also of many other strategic elements including rite and pyrite). Biological mediation is essential to speed up
Mo and Li in nodules, and additionally Ti, Pt, Zr, Nb, Te, Bi, W, the rate of selective metal transfer and secondary enrichment:
Th, and rare earth elements (lanthanides) in ferromanganese sulfide oxidation in the unsaturated zone by Fe-oxidizing bacte-
crusts (Grupe et al., 2001; Chapter 13.11). Marine phosphate ria, followed by reprecipitation assisted by sulfate-reducing
Fluids and Ore Formation in the Earth’s Crust 23

bacteria below the groundwater table (Ague and Brimhall, 1989; estimates for copper resources give a similar number. Never-
Enders et al., 2006; Sillitoe, 2005). theless, known reserves have generally grown over time, de-
Mechanical redistribution of chemically resistive mineral spite the exhaustion of individual deposits, thanks to human
grains by flowing surface water generates placer deposits of innovation in extractive technology and in delineating previ-
Au, Sn (as cassiterite SnO2), and various gemstones including ously unknown resources. In spite of recurrent warnings, the
diamond (Chapter 13.23). River and beach sands are the main reserves of most important base and precious metals have been
source of Ti (rutile and ilmenite), Zr (zircon), and a significant sufficient to supply at least a decade or two at current rates of
source of rare earth elements from monazite (Garnett and mining, although finding new deposits has become more dif-
Bassett, 2005). In the anoxic surface environment of the Ar- ficult and costly (Enders and Saunders, 2011), and their ex-
chean, pyrite and uraninite are interpreted to have been trans- traction will increasingly be limited by the use of land, water,
ported as refractory heavy minerals, together with gold, and energy.
forming giant placer accumulations that were later reworked Both, economic and environmental costs are (to a first
and possibly upgraded by metamorphic fluid activity (Frimmel approximation) inversely related to the natural metal enrich-
et al., 2005; Chapter 13.17). ment in an ore deposit. It will always be possible to extract
metals from lower grade resources, so ‘running out’ is not the
real issue. Statistical modeling of average formation, burial, and
13.1.8 Back to the Future: Global Mineral Resources exposure rates (e.g., Kesler and Wilkinson, 2008) shows that
undiscovered deposits exist in numbers, sizes, and ore grades
We tried to show in this review that ore formation is an integral meeting demand for many centuries, even at relatively shallow
part of our planet’s evolution and reflects varied combinations depths of 1–3 km accessible by today’s mining technology.
of large-scale tectonics with chemical interactions between the Finding and mining these buried high-grade deposits – albeit
reduced hot interior of the Earth and its evolving atmosphere scientifically and technically challenging and perhaps more
and hydrosphere. The largest and economically most signifi- costly – may become a more attractive option than using
cant ore deposits are extreme geochemical anomalies, not be- lower grade ores at greater expense of energy and environmental
cause of inherently anomalous or untypical processes, but resources. In any case, along with additional research to under-
because they require an optimized confluence, in space and stand how the deposits form, we must do more far-field re-
time, of several geological processes that cooperate at multiple search to be able to detect deposits from geological and
scales. The magnitude of mass transfer forming world-class ore geochemical clues at greater distances from the deposits. We
deposits provides some of the most spectacular evidence that hope that this volume of Treatise on Geochemistry will bring a
lithosphere-scale fluid flow through rocks is a characteristic of vital topic to the attention of the broad range of geoscientists,
our planet. Ore deposits are therefore a research window to the who are needed to meet this challenge.
physical and chemical evolution of the Earth, because the types
of deposits change from the Archean to the Phanerozoic as a
result of the chemical evolution and interactions of mantle, Acknowledgments
crust, ocean, and atmosphere (Holland, 2005). Hence, it
would not be surprising if chemical reactions between sulfide This review builds on discussions over many years within both
surfaces and hydrothermal ore fluids held the keys to the origin our groups at ETH Zurich and at the University of Maryland,
of life on Earth (Russell et al., 2005). including the interaction with graduate students in courses
Popular discussions about resources commonly ask, why where we have presented this material in a similar structure.
do we mine new ore deposits, rather than recycle all our CAH would like to specially acknowledge (occasionally
metals? Extensive recent research has been devoted to assessing heated) discussions with Thomas Driesner, John Dilles, Vic
the stocks of metals in use today, to emphasize the ultimate Wall, Anthony Williams-Jones, John Walshe, and his late men-
aim of sustainable use of resources (e.g., Graedel et al., 2006). tor Mike Solomon; PAC would like to thank Phil Piccoli and
However, a great part of this stock is in long-term use (e.g., in Phil Blevin for their many discussions and critiques over the
building infrastructure) so that even complete recycling of years, his many students and post-docs, and his late mentor,
easily separated metals like copper cannot meet the demand Dick Holland – hastening to say that we will take all the blame
of industrialized countries today, let alone of future global for bias and shortcomings in this review. We also would like to
society. Unless people in emerging economies are denied ac- thank Thomas Wagner and Tobias Schlegel for help with the
cess to electricity and other infrastructure, comparable to that construction of thermodynamic diagrams, and additionally
enjoyed by the industrialized countries today, finding and Karin Siegel for part of the drafting. Excellent discussions
mining new copper resources is a necessity that cannot be with and reviews of the manuscript by Alex Brown, Steve
avoided for many decades – even assuming complete recycling Ingebritsen, Steve Kesler, Dave Leach, Steve Scott, Alan
of the waste becoming available. Similar arguments can be Thompson, and Bruce Yardley are gratefully acknowledged.
made for many other metals, including uranium.
So, will we run out of essential materials soon? Ores clearly
are ‘nonrenewable’ resources in the sense that their rate of
geological accumulation does not keep pace with the rate of References
extraction and usage by human society. Estimates for gold Ague JJ and Brimhall GH (1989) Geochemical modeling of steady-state fluid-flow and
show that we mine deposits about 17 000 times faster than chemical reaction during supergene enrichment of porphyry copper deposits.
they form and are preserved (Kesler and Wilkinson, 2009), and Economic Geology 84: 506–528.
24 Fluids and Ore Formation in the Earth’s Crust

Akinfiev NN and Zotov AV (2001) Thermodynamic description of chloride, hydrosulfide, Cathles LMI and Adams JJ (2005) Fluid flow and petroleum and mineral resources in
and hydroxo complexes of Ag(I), Cu(I), and Au(I) at temperatures of 25–500  C and the upper (<20-km) continental crust. In: Hedenquist JW, Thompson JFH,
pressures of 1–2000 bar. Geochemistry International 39: 990–1006. Goldfarb RJ, and Richards JP (eds.) Economic Geology 100th Anniversary Volume,
Akinfiev NN and Zotov AV (2010) Thermodynamic description of aqueous species in the pp. 77–110. Littleton, CO: Society of Economic Geologists.
system Cu–Ag–Au–S–O–H at temperatures of 0–600  C and pressures of 1–3000 Cathles LM and Shannon R (2007) How potassium silicate alteration suggests the
bar. Geochemistry International 48: 714–720. formation of porphyry ore deposits begins with the nearly explosive but barren
Anderson GM (2008) The mixing hypothesis and the origin of Mississippi Valley-type expulsion of large volumes of magmatic water. Earth and Planetary Science Letters
ore deposits. Economic Geology 103: 1683–1690. 262: 92–108.
Arancibia ON and Clark AH (1996) Early magnetite-amphibole-plagioclase Cawthorn RG, Barnes SJ, Ballhaus C, and Malitch KN (2005) Platinum group element,
alteration-mineralization in the Island Copper porphyry copper-gold-molybdenum chromium, and vanadium deposits in mafic and ultramafic rocks.
deposit, British Columbia. Economic Geology 91: 402–439. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic
Arndt NT, Lesher CM, and Czamanske GK (2005) Mantle-derived magmas and Geology 100th Anniversary Volume, pp. 215–249. Littleton, CO: Society of
magmatic Ni-Cu-(PGE) deposits. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, Economic Geologists.
and Richards JP (eds.) Economic Geology 100th Anniversary Volume, pp. 5–23. Černý P, Blevin PL, Cuney M, and London D (2005) Granite-related ore deposits.
Littleton, CO: Society of Economic Geologists. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic
Audétat A, Pettke T, Heinrich CA, and Bodnar RJ (2008) The composition of Geology 100th Anniversary Volume, pp. 337–370. Littleton, CO: Society of
magmatic-hydrothermal fluids in barren and mineralized intrusions. Economic Economic Geologists.
Geology 103: 877–908. Cline JS and Bodnar RJ (1991) Can economic porphyry copper mineralization be
Baker T, Van Achterberg E, Ryan CG, and Lang JR (2004) Composition and generated by a typical calc-alkaline melt? Journal of Geophysical Research
evolution of ore fluids in a magmatic-hydrothermal skarn deposit. Geology 96: 8113–8126.
32: 117–120. Cline JS, Hofstra AH, Muntean JL, Tosdal RM, and Hickey KA (2005) Carlin-type gold
Barnes HL (1997) Geochemistry of Hydrothermal Ore Deposits. New York: Wiley. deposits in Nevada: Critical geologic characteristics and viable models.
Barnes S-J and Lightfoot PC (2005) Formation of magmatic nickel sulfide deposits and In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic
processes affecting their copper and platinum group element contents. Geology 100th Anniversary Volume, pp. 451–484. Littleton, CO: Society of
In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic Economic Geologists.
Geology 100th Anniversary Volume, pp. 179–213. Littleton, CO: Society of Cloke PL and Kesler SE (1979) Halite trend in hydrothermal solutions. Economic
Economic Geologists. Geology 74: 1823–1831.
Barton PB and Toulmin P (1961) Some mechanisms for cooling hydrothermal fluids. In: Cloos M (2001) Bubbling magma chambers, cupolas, and porphyry copper deposits.
Short Papers in the Geologic and Hydrologic Sciences, Articles 293-435, US International Geology Review 43: 285–311.
Geological Survey Professional Paper 424-D, pp. 348–352. Washington, DC: US Connors KA, Noble DC, Bussey SD, and Weiss SI (1993) Initial gold contents of silicic
Government Printing Office. volcanic rocks: Bearing on the behavior of gold in magmatic systems. Geology
Bendezu R and Fontbote L (2009) Cordilleran epithermal Cu–Zn–Pb–(Au–Ag) 21: 937–940.
mineralization in the Colquijirca District, Central Peru: Deposit-scale mineralogical Coumou D, Driesner T, and Heinrich CA (2008) The structure and dynamics of
patterns. Economic Geology 104: 905–944. mid-ocean ridge hydrothermal systems. Science 321: 1825–1828.
Bethke CM (1986) Hydrologic constraints on the genesis of the Upper Mississippi Courtillot V and Olson P (2007) Mantle plumes link magnetic superchrons to Phanerozoic
Valley mineral district from Illinois Basin brines. Economic Geology 81: 233–249. mass depletion events. Earth and Planetary Science Letters 260: 495–504.
Bierlein FP, Groves DI, Goldfarb RJ, and Dubé B (2006) Lithospheric controls on the Cox SF (2005) Coupling between deformation, fluid pressures, and fluid flow in
formation of provinces hosting giant orogenic gold deposits. Mineralium Deposita ore-producing hydrothermal systems at depth in the crust. In: Hedenquist JW,
40: 874–886. Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic Geology 100th
Boiron MC, Cathelineau M, and Richard A (2010) Fluid flows and metal deposition near Anniversary Volume, pp. 39–75. Littleton, CO: Society of Economic Geologists.
basement/cover unconformity: Lessons and analogies from Pb–Zn–F–Ba systems Cox SF (2010) The application of failure mode diagrams for exploring the roles of fluid
for the understanding of Proterozoic U deposits. Geofluids 10: 270–292. pressure and stress states in controlling styles of fracture-controlled permeability
Brimhall GH and Crerar DA (1987) Ore fluids: Magmatic to supergene. enhancement in faults and shear zones. Geofluids 10: 217–233.
In: Carmichael ISE and Eugster HP (eds.) Thermodynamic Modeling of Geological Cox DP and Singer DA (1988) Distribution of gold in porphyry copper deposits. US
Materials: Minerals, Fluids and Melt. Reviews in Mineralogy, vol. 17, pp. 235–321. Geological Survey Open-File Report 88–46. Denver, CO: US Geological Survey.
Chantilly, VA: Mineralogical Society of America. Cuney M (2010) Evolution of uranium fractionation processes through time:
Burnham CW and Ohmoto H (1980) Late-stage processes of felsic magmatism. Mining Driving the secular variation of uranium deposit types. Economic Geology
Geology Special Issue 8: 1–11. 105: 553–569.
Candela PA (1991) Physics of aqueous phase evolution in plutonic environments. De Ronde CEJ, Hannington MD, Stoffers P, et al. (2005) Evolution of a submarine
American Mineralogist 76: 1081–1091. magmatic-hydrothermal system: Brothers volcano, southern Kermadec arc, New
Candela PA (2003) Ores in the Earth’s crust. In: Rudnick RL (ed.) Treatise on Zealand. Economic Geology 100: 1097–1133.
Geochemistry, Vol 3: The Crust, pp. 411–431. Oxford: Elsevier. Deming D (1992) Catastrophic release of heat and fluid-flow in the continental crust.
Candela PA and Holland HD (1984) The partitioning of copper and molybdenum Geology 20: 83–86.
between silicate melts and aqueous fluids. Geochimica et Cosmochimica Acta Dilles JH (1987) The petrology of the Yerington Batholith, Nevada: Evidence for the
48: 373–380. evolution of porphyry copper ore fluids. Economic Geology 82: 1750–1789.
Candela PA and Piccoli PM (2005) Magmatic processes in the development of Dilles JH and Einaudi MT (1992) Wall-rock alteration and hydrothermal flow paths
porphyry-type ore systems. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and about the Ann-Mason porphyry copper deposit, Nevada – A 6 km vertical
Richards JP (eds.) Economic Geology 100th Anniversary Volume, pp. 25–37. reconstruction. Economic Geology 87: 1963–2001.
Littleton, CO: Society of Economic Geologists. Dolejs D and Wagner T (2008) Thermodynamic modeling of non-ideal mineral-fluid
Carroll MR and Rutherford MJ (1985) Sulfide and sulfate saturation in hydrous silicate equilibria in the system Si–Al–Fe–Mg–Ca–Na–K–H–O–Cl at elevated temperatures
melts. Journal of Geophysical Research 90: C601–C612. and pressures: Implications for hydrothermal mass transfer in granitic rocks.
Carroll MR and Webster JD (1994) Solubilities of sulfur, noble gases, nitrogen, Geochimica et Cosmochimica Acta 72: 526–553.
chlorine, and fluorine in magmas. In: Carroll MR and Holloway JR (eds.) Volatiles in Driesner T and Geiger S (2007) Numerical simulation of multiphase fluid flow in
Magmas. Reviews in Mineralogy, vol. 30, pp. 251–279. Chantilly, VA: hydrothermal systems. In: Liebscher A and Heinrich CA (eds.) Fluid–Fluid
Mineralogical Society of America. Interactions. Reviews in Mineralogy, vol. 65, pp. 187–212. Chantilly, VA:
Catchpole H, Kouzmanov K, Fontbote L, Guillong M, and Heinrich CA (2011) Fluid Mineralogical Society of America.
evolution in zoned Cordilleran polymetallic veins – Insights from microthermometry Driesner T and Heinrich CA (2007) The system H2O–NaCl. Part I: Correlation formulae
and LA-ICP-MS of fluid inclusions. Chemical Geology 281: 293–304. for phase relations in temperature–pressure–composition space from 0 to 1000  C,
Cathles LM (1981) Fluid flow and genesis of hydrothermal ore deposits. In: Skinner BJ 0 to 5000 bar, and 0 to 1 XNaCl. Geochimica et Cosmochimica Acta 71: 4880–4901.
(ed.) Economic Geology 75th Anniversary Volume, pp. 424–457. Littleton, CO: Drummond SE and Ohmoto H (1985) Chemical evolution and mineral deposition in
Society of Economic Geologists. boiling hydrothermal systems. Economic Geology 87: 1963–2001.
Cathles LM (1997) Thermal aspects of ore formation. In: Barnes HL (ed.) Geochemistry Eastoe CJ (1982) Physics and chemistry of the hydrothermal system at the Panguna porphyry
of Hydrothermal Ore Deposits, 3rd edn., pp. 191–227. New York: Wiley. copper deposit, Bougainville, Papua New Guinea. Economic Geology 77: 127–153.
Fluids and Ore Formation in the Earth’s Crust 25

Emsbo P, Groves DI, Hofstra AH, and Bierlein FP (2006) The giant Carlin gold province: In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic
A protracted interplay of orogenic, basinal, and hydrothermal processes above a Geology 100th Anniversary Volume, pp. 407–450. Littleton, CO: Society of
lithospheric boundary. Mineralium Deposita 41: 517–525. Economic Geologists.
Enders MS, Knickerbocker C, and Titley SR (2006) The role of bacteria in the supergene Graedel TE, Gordon RB, and Bertram M (2006) Metal stocks and sustainability.
environment of the Morenci porphyry copper deposit, Greenlee County, Arizona. Proceedings of the National Academy of Sciences of the United States of America
Economic Geology 101: 59–70. 103: 1209–1214.
Enders MS and Saunders C (2011) Discovery, innovation, and learning in the mining Groves DI, Condie KC, Goldfarb RJ, Hronsky JMA, and Vielreicher RM (2005) 100th
business – New ways forward for an old industry. SEG Newsletter, Society of Anniversary Special Paper: Secular changes in global tectonic processes and their
Economic Geologists 86: 1–22. influence on the temporal distribution of gold-bearing mineral deposits. Economic
England PC and Thompson AB (1984) Pressure–temperature–time paths of regional Geology 100: 203–224.
metamorphism. 1. Heat-transfer during the evolution of regions of thickened Groves DI and Vielreicher NM (2001) The Phalabowra (Palabora) carbonatite-hosted
continental crust. Journal of Petrology 25: 894–928. magnetite-copper sulfide deposit, South Africa: An end member of the iron-
Eugster HP and Hardie LA (1975) Sedimentation in an ancient playa-lake oxide copper-gold-rare earth element deposit group? Mineralium Deposita
complex – Wilkins Peak Member of the Green River Formation of Wyoming. 36: 189–194.
Geological Society of America Bulletin 86: 319–334. Grupe B, Becker HJ, and Oebius HU (2001) Geotechnical and sedimentological
Evans KF, Moriya H, Niitsuma H, et al. (2005) Microseismicity and permeability investigations of deep-sea sediments from a manganese nodule field of the
enhancement of hydrogeologic structures during massive fluid injections into Peru Basin. Deep Sea Research Part II: Topical Studies in Oceanography
granite at 3 km depth at the Soultz HDR site. Geophysical Journal International 48: 3593–3608.
160: 388–412. Gustafson LB and Williams N (1981) Sediment-hosted stratiform deposits of copper,
Fallick AE, Ashton JH, Boyce AJ, Ellam BM, and Russell MJ (2001) Bacteria were lead, and zinc. In: Skinner BJ (ed.) Economic Geology 75th Anniversary Volume,
responsible for the magnitude of the world-class hydrothermal base metal sulfide pp. 139–178. Littleton, CO: Society of Economic Geologists.
orebody at Navan Ireland. Economic Geology 96: 885–890. Halter WE, Heinrich CA, and Pettke T (2005) Magma evolution and the formation of
Fisher AT (2005) Marine hydrogeology: Recent accomplishments and future porphyry Cu–Au ore fluids: Evidence from silicate and sulfide melt inclusions.
opportunities. Hydrogeology Journal 13: 69–97. Mineralium Deposita 39: 845–863.
Fitch WI (1997) Uranium, its impact on the national and global energy mix, US Halter WE, Williams-Jones AE, and Kontak DJ (1996) The role of greisenization in
Geological Survey Circular 1141. Washington DC: US Government Printing Office. cassiterite precipitation at the East Kemptville tin deposit, Nova Scotia. Economic
Fournier RO (1999) Hydrothermal processes related to movement of fluid from plastic Geology 91: 368–385.
into brittle rock in the magmatic-epithermal environment. Economic Geology Hannington MD, De Ronde CEJ, and Petersen S (2005) Sea floor tectonics and
94: 1193–1211. submarine hydrothermal systems. In: Hedenquist JW, Thompson JFH, Goldfarb RJ,
Frakes L and Bolton B (1992) Effects of ocean chemistry, sea-level, and climate on the and Richards JP (eds.) Economic Geology 100th Anniversary Volume,
formation of primary sedimentary manganese ore-deposits. Economic Geology pp. 111–141. Littleton, CO: Society of Economic Geologists.
87: 1207–1217. Hardie LA and Eugster HP (1970) The evolution of closed-basin brines, American
Frank MR, Candela PA, Piccoli PM, and Glascock MD (2002) Gold solubility, Mineralogical Society Special Paper 3, pp. 273–290. Chantilly, VA: Mineralogical
speciation, and partitioning as a function of HCl in the brine-silicate melt-metallic Society of America.
gold system at 800  C and 100 MPa. Geochimica et Cosmochimica Acta Harris AC, Kamenetsky VS, White NC, Van Achterbergh E, and Ryan CG (2003) Melt
66: 3719–3732. inclusions in veins: Linking magmas and porphyry Cu deposits. Science
Frank MR, Simon AC, Pettke T, Candela PA, and Piccoli PM (2011) Gold and copper 302: 2109–2111.
partitioning in magmatic-hydrothermal systems at 800  C and 100 MPa. Hattori KH and Keith JD (2001) Contribution of mafic melt to porphyry copper
Geochimica et Cosmochimica Acta 75: 2470–2482. mineralization: Evidence from Mount Pinatubo, Philippines, and Bingham Canyon,
Franklin JM, Gibson HL, Jonasson IR, and Galley AG (2005) Volcanogenic massive Utah, USA. Mineralium Deposita 36: 799–806.
sulfide deposits. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP Hayba DO and Ingebritsen SE (1997) Multiphase groundwater flow near cooling
(eds.) Economic Geology 100th Anniversary Volume, pp. 523–560. Littleton, CO: plutons. Journal of Geophysical Research 102: 12235–12252.
Society of Economic Geologists. Hecht L and Cuney M (2000) Hydrothermal alteration of monazite in the Precambrian
Freyssinet P, Butt CRM, Morris RC, and Piantone P (2005) Ore-forming processes crystalline basement of the Athabasca Basin (Saskatchewan, Canada): Implications
related to lateritic weathering. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and for the formation of unconformity-related uranium deposits. Mineralium Deposita
Richards JP (eds.) Economic Geology 100th Anniversary Volume, pp. 681–722. 35: 791–795.
Littleton, CO: Society of Economic Geologists. Hedenquist JW, Arribas A, and Reynolds TJ (1998) Evolution of an intrusion-centered
Frimmel HE (2008) Earth’s continental crustal gold endowment. Earth and Planetary hydrothermal system: Far Southeast-Lepanto porphyry and epithermal Cu–Au
Science Letters 267: 45–55. deposits, Philippines. Economic Geology 93: 373–404.
Frimmel HE, Groves DI, Kirk J, Ruiz J, Chesley J, and Minter WEL (2005) The formation Hedenquist JW and Lowenstern JB (1994) The role of magmas in the formation of
and preservation of the Witwatersrand gold fields, the world’s largest gold province. hydrothermal ore deposits. Nature 370: 519–527.
In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic Heinrich CA (1990) The chemistry of hydrothermal tin (-tungsten) ore deposition.
Geology 100th Anniversary Volume, pp. 769–797. Littleton, CO: Society of Economic Geology 90: 705–729.
Economic Geologists. Heinrich CA (2006) From fluid inclusion microanalysis to large-scale hydrothermal
Fyfe WS, Price NJ, and Thompson AB (1978) Fluids in the Earth’s Crust: Their mass transfer in the Earth’s interior. Journal of Mineralogical and Petrological
Significance in Metamorphic Tectonic and Chemical Transport. Amsterdam: Elsevier. Sciences 101: 110–117.
Gammons CH and Williams-Jones AE (1997) Chemical mobility of gold in the Heinrich CA, Bain JHC, Mernagh TP, Wyborn LAI, Andrew AS, and Waring CL (1995)
porphyry-epithermal environment. Economic Geology 92: 45–59. Fluid and mass transfer during metabasalt alteration and copper mineralization at
Garnett RHT and Bassett NC (2005) Placer deposits. In: Hedenquist JW, Mount Isa, Australia. Economic Geology 90: 705–730.
Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic Geology 100th Heinrich CA, Driesner T, Stefánsson A, and Seward TM (2004) Magmatic vapor
Anniversary Volume, pp. 813–843. Littleton, CO: Society of Economic Geologists. contraction and the transport of gold from the porphyry environment to epithermal
Garven G, Ge S, Person MA, and Sverjensky DA (1993) Genesis of stratabound ore ore deposits. Geology 32: 761–764.
deposits in the midcontinent basins of North America. 1. The role of regional Heinrich CA, Günther D, Audétat A, Ulrich T, and Frischknecht R (1999) Metal
groundwater flow. American Journal of Science 293: 497–568. fractionation between magmatic brine and vapor, determined by microanalysis of
German CR and Von Damm KL (2003) Hydrothermal systems. In: Elderfield H (ed.) fluid inclusions. Geology 27: 755–758.
Treatise on Geochemistry: Vol 6, The Oceans, pp. 181–222. Oxford: Elsevier. Hemley JJ, Cygan GL, Fein JB, Robinson GR, and D´Angelo WM (1992)
Giggenbach WF (1992) Magma degassing and mineral deposition in hydrothermal Hydrothermal ore-forming processes in the light of studies in rock-buffered
systems along convergent plate boundaries. Economic Geology 87: 1927–1944. systems: I. Iron-copper-zinc-lead sulfide solubility relations. Economic Geology
Giggenbach WF (1997) The origin and evolution of fluids in magmatic-hydrothermal 87: 1–22.
systems. In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn., Henley RW and Berger BR (2000) Self-ordering and complexity in epizonal mineral
pp. 737–796. New York: Wiley. deposits. Annual Review of Earth and Planetary Sciences 28: 669–719.
Goldfarb RJ, Baker T, Dubé B, Groves DI, Hart CJR, and Gosselin P (2005) Distribution, Henley RW and Mcnabb A (1978) Magmatic vapor plumes and ground-water interaction
character, and genesis of gold deposits in metamorphic terranes. in porphyry copper emplacement. Economic Geology 73: 1–20.
26 Fluids and Ore Formation in the Earth’s Crust

Hennings SK, Wagner T, Heinrich CA, and Ulmer P (In press) Fluid evolution of the Large RR (1992) Australian volcanic-hosted massive sulfide deposits: Features, styles,
Monte Mattoni mafic complex, Adamello batholith, northern Italy: Constraints from and genetic models. Economic Geology 87: 471–510.
LA-ICPMS analysis of fluid inclusions and thermodynamic modeling. Journal of Lasky SG (1950) How tonnage and grade relations help predict ore reserves.
Petrology. Engineering and Mining Journal 151: 81–85.
Henry AL, Anderson GM, and Heroux Y (1992) Alteration of organic matter in the Viburnum Law JDM and Phillips GN (2005) Hydrothermal replacement model for Witwatersrand
Trend lead-zinc district of southeastern Missouri. Economic Geology R87: 288–309. gold. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.)
Hill DP, Pollitz F, and Newhall C (2002) Earthquake-volcano interactions. Physics Economic Geology 100th Anniversary Volume, pp. 799–811. Littleton, CO: Society
Today 55: 41–47. of Economic Geologists.
Hitzman M, Kirkham R, Broughton D, Thorson J, and Selley D (2005) The Leach DL, Sangster DF, Kelley KD, et al. (2005) Sediment-hosted lead-zinc deposits:
sediment-hosted stratiform copper ore system. In: Hedenquist JW, Thompson JFH, A global perspective. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and
Goldfarb RJ, and Richards JP (eds.) Economic Geology 100th Anniversary Volume, Richards JP (eds.) Economic Geology 100th Anniversary Volume, pp. 561–607.
pp. 609–642. Littleton, CO: Society of Economic Geologists. Littleton, CO: Society of Economic Geologists.
Holland HD (2005) 100th Anniversary Special Paper: Sedimentary mineral deposits and Lehmann B, Ishihara S, Michel H, et al. (1990) The Bolivian tin province and regional tin
the evolution of Earth’s near-surface environments. Economic Geology distribution in the Central Andes – A reassessment. Economic Geology 85: 1044–1058.
100: 1489–1509. Lesher CM, Lee RF, Groves DI, Bickle MJ, and Donaldson MJ (1981) Geochemistry
Holland TJB and Powell R (1998) An internally consistent thermodynamic data of komatiites from Kambalda, Western Australia. 1. Chalcophile element
set for phases of petrological interest. Journal of Metamorphic Geology depletion – A consequence of sulfide liquid separation from komatiitic magmas.
16: 309–343. Economic Geology 76: 1714–1728.
Huenges E, Erzinger J, Kuck J, Engeser B, and Kessels W (1997) The permeable crust: Li Y and Audétat A (2012) Partitioning of V, Mn, Co, Ni, Cu, Zn, As, Mo, Ag, Sn, Sb, W,
Geohydraulic properties down to 9101 m depth. Journal of Geophysical Au, Pb, and Bi between sulfide phases and hydrous basanite melt at upper mantle
Research 102(B8): 18255–18265. conditions. Earth and Planetary Science Letters 355: 327–340.
Ingebritsen SE and Appold MS (2011) The physical hydrogeology of ore deposits. Li Y, Audétat A, Lerchbaumer L, and Xiong XL (2009b) Rapid Na, Cu exchange between
Economic Geology 107: 559–584. synthetic fluid inclusions and external aqueous solutions: Evidence from
Ingebritsen SE, Geiger S, Hurwitz S, and Driesner T (2010) Numerical simulation of LA-ICP-MS analysis. Geofluids 9: 321–329.
magmatic hydrothermal systems. Reviews of Geophysics 48: 1–33. Li CS, Ripley EM, and Naldrett AJ (2009a) A new genetic model for the giant Ni-Cu-PGE
Ingebritsen SE and Manning CE (2010) Permeability of the continental crust: Dynamic sulfide deposits associated with the Siberian flood basalts. Economic Geology
variations inferred from seismicity and metamorphism. Geofluids 10: 193–205. 104: 291–301.
Ingebritsen SE, Sanford S, and Neuzil C (2006) Groundwater in Geologic Processes, Liebscher A and Heinrich CA (2007) Fluid–fluid interactions in the Earth’s lithosphere.
2nd edn. Cambridge: Cambridge University Press. In: Liebscher A and Heinrich CA (eds.) Fluid–Fluid Interactions. Reviews in
Irvine TN (1977) Origin of chromitite layers in Muskox intrusion and other stratiform Mineralogy and Geochemistry, vol. 65, pp. 1–13. Chantilly, VA: Mineralogical
intrusions – New interpretation. Geology 5: 273–277. Society of America.
Jones HD and Kesler SE (1992) Fluid inclusion gas chemistry in east Tennessee Lowenstern JB (2001) Carbon dioxide in magmas and implications for hydrothermal
Mississippi Valley-type districts – Evidence for immiscibility and implications for systems. Mineralium Deposita 36: 490–502.
depositional mechanisms. Geochimica et Cosmochimica Acta 56: 137–154. Mahara Y, Habermehl MA, Hasegawa T, et al. (2009) Groundwater dating by estimation of
Jugo PJ, Candela PA, and Piccoli PM (1999) Magmatic sulfides and Au:Cu ratios in groundwater flow velocity and dissolved 4He accumulation rate calibrated by 36Cl in the
porphyry deposits: An experimental study of copper and gold partitioning at 850  C, Great Artesian Basin, Australia. Earth and Planetary Science Letters 287: 43–56.
100 MPa in a haplogranitic melt–pyrrhotite–intermediate solid solution–gold metal Maier WD and Groves DI (2011) Temporal and spatial controls on the formation of
assemblage, at gas saturation. Lithos 46: 573–589. magmatic PGE and Ni–Cu deposits. Mineralium Deposita 46: 841–857.
Kay SM and Mpodozis C (2002) Magmatism as a probe to the Neogene shallowing of Montoya JW and Hemley JJ (1975) Activity relations and stabilities in alkali feldspar
the Nazca plate beneath the modern Chilean flat-slab. Journal of South American and mica alteration reactions. Economic Geology 70: 577–583.
Earth Sciences 15: 39–57. Mungall JE (2002) Roasting the mantle: Slab melting and the genesis of major Au and
Kelley KA and Cottrell E (2009) Water and the oxidation state of subduction zone Au-rich Cu deposits. Geology 30: 915–918.
magmas. Science 325: 605–607. Muntean JL, Cline JS, Simon AC, and Longo AA (2011) Magmatic-hydrothermal origin
Kesler SE (1994) Mineral Resources, Economics and the Environment, New York: McMillan. of Nevada’s Carlin-type gold deposits. Nature Geoscience 4: 122–127.
Kesler SE, Riciputi LC, and Ye ZJ (2005) Evidence for a magmatic origin for Carlin-type Muntean JL and Einaudi MT (2000) Porphyry gold deposits of the Refugio district,
gold deposits: Isotopic composition of sulfur in the Betze-Post-Screamer deposit, Maricunga belt, northern Chile. Economic Geology 95: 1445–1472.
Nevada, USA. Mineralium Deposita 40: 127–136. Murakami H, Seo JH, and Heinrich CA (2010) The relation between Cu/Au ratio and formation
Kesler SE and Wilkinson BH (2008) Earth’s copper resources estimated from tectonic depth of porphyry-style Cu-Au  Mo deposits. Mineralium Deposita 45: 11–21.
diffusion of porphyry copper deposits. Geology 36: 255–258. Mustard R, Ulrich T, Kamenetsky VS, and Mernagh T (2006) Gold and metal enrichment
Kesler SE and Wilkinson BH (2009) Resources of gold in Phanerozoic epithermal in natural granitic melts during fractional crystallization. Geology 34: 85–88.
deposits. Economic Geology 104: 623–633. Naldrett AJ (2010) Secular variation of magmatic sulfide deposits and their source
Kharaka YK, Maest AS, Carothers WM, Law LM, Lamothe PJ, and Fries TL (1987) magmas. Economic Geology 105: 669–688.
Geochemistry of metal-rich brines from central Mississippi Salt Dome basin, U.S.A. Ohmoto H (1986) Stable isotope geochemistry of ore deposits. In: Valley JW, Taylor HP
Applied Geochemistry 2: 543–561. Jr., and O’Neil JR (eds.) Stable Isotopes in High Temperature Geological Processes.
Knight CL (1957) Ore genesis – The source bed concept. Economic Geology Reviews in Mineralogy, vol. 16, pp. 491–559. Chantilly, VA: Mineralogical Society
52: 808–817. of America.
Koděra P, Lexa J, Heinrich CA, Waelle M, and Fallick AE (2011) Au mineralisation at Pettke T, Oberli F, and Heinrich CA (2010) The magma and metal source of giant
Biely Vrch deposit (W Carpathians, Slovakia): Example of a very shallow porphyry porphyry-type ore deposits, based on lead isotope microanalysis of individual fluid
system. 11th Biennial SGA Meeting, Antofagasta, Extended Abstract, pp. 100–103. inclusions. Earth and Planetary Science Letters 296: 267–277.
Kostova B, Pettke T, Driesner T, Petrov P, and Heinrich CA (2004) LA ICP-MS study of Phillips GN and Groves DI (1983) The nature of Archean gold-bearing fluids as deduced
fluid inclusions in quartz from the Yuzhna Petrovitsa deposit, Madan ore field, from gold deposits of Western Australia. Journal of the Geological Society of
Bulgaria. Schweizerische Mineralogische und Petrographische Mitteilungen Australia 30: 25–39.
84: 25–36. Plumlee GS, Leach DL, Hofstra AH, Landis GP, Rowan EL, and Viets JG (1994)
Koziy L, Bull S, Large R, and Selley D (2009) Salt as a fluid driver, and basement as a Chemical reaction path modeling of ore deposition in Mississippi Valley-type
metal source, for stratiform sediment-hosted copper deposits. Geology Pb–Zn deposits of the Ozark Region, United States Midcontinent. Economic
37: 1107–1110. Geology 89: 1361–1383.
Krahulec KA (2010) Production history of the Bingham mining district, Salt Lake County, Pohl WL (2011) Economic Geology, Principles and Practice. Chichester: Wiley.
Utah. In: Krahulec KA and Schroeder K (eds.) Tops and Bottoms of Porphyry Copper Pokrovski GS, Borisova AY, and Harrichoury JC (2008) The effect of sulfur on
Deposits: The Bingham and Southwest Tintic Districts, Utah. Guidebook Series, vapor–liquid fractionation of metals in hydrothermal systems. Earth and Planetary
vol. 41, pp. 25–33. Littleton, CO: Society of Economic Geologists. Science Letters 266: 345–362.
Landtwing MR, Furrer C, Redmond PB, Pettke T, Guillong M, and Heinrich CA (2010) Reed MH (1997) Hydrothermal alteration and its relationship to ore fluid composition.
The Bingham Canyon porphyry Cu–Mo–Au deposit. III. Zoned copper–gold ore In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn.,
deposition by magmatic vapor expansion. Economic Geology 105: 91–118. pp. 303–365. New York: Wiley.
Fluids and Ore Formation in the Earth’s Crust 27

Reilly MJ and Flemings PB (2010) Deep pore pressures and seafloor venting in the Singer DA (1995) World-class base and precious-metal deposits – A quantitative
Auger Basin, Gulf of Mexico. Basin Research 22: 380–397. analysis. Economic Geology 90: 88–104.
Richards JP (2003) Tectono-magmatic precursors for porphyry Cu–(Mo–Au) deposit Singer DA (2010) Progress in integrated quantitative mineral resource assessments. Ore
formation. Economic Geology 98: 1515–1533. Geology Reviews 38: 242–250.
Richards JP (2011) Magmatic to hydrothermal metal fluxes in convergent and collided Singer DA, Berger VI, Menzie WD, and Berger BR (2005) Porphyry copper deposit
margins. Ore Geology Reviews 40: 1–26. density. Economic Geology 100: 491–514.
Ridley J (1993) The Relations between mean rock stress and fluid-flow in the Skinner BJ (1997) Hydrothermal metal deposits: What we do and don’t know.
crust – With reference to vein-style and lode-style gold deposits. Ore Geology In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn., pp. 1–29.
Reviews 8: 23–37. New York: Wiley.
Ridley J and Mengler F (2000) Lithological and structural controls on the form and Solomon M and Quesada C (2003) Zn–Pb–Cu massive sulfide deposits: Brine-pool
setting of vein stockwork orebodies at the Mount Charlotte gold deposit, Kalgoorlie. types occur in collisional orogens, black smoker types occur in backarc and/or arc
Economic Geology 95: 85–98. basins. Geology 31: 1029–1032.
Ridley J, Mikucki EJ, and Groves DI (1996) Archean lode-gold deposits: Fluid flow and Solomon M and Walshe JL (1979) Formation of massive sulfide deposits on the sea
chemical evolution in vertically extensive hydrothermal systems. Ore Geology floor. Economic Geology 74: 797–813.
Reviews 10: 279–293. Spycher NF and Reed MH (1989) Evolution of a broadlands-type epithermal ore fluid
Rimstidt JD (1997) Gangue mineral transport and deposition. In: Barnes HL (ed.) along alternative P-T paths – Implications for the transport and deposition of base,
Geochemistry of Hydrothermal Ore Deposits, 3rd edn. New York: Wiley. precious, and volatile metals. Economic Geology 84: 328–359.
Robb L (2004) Introduction to Ore-Forming Processes. Berlin: Wiley-VCH. Stanton RL (1955) Lower Paleozoic mineralisation near Bathurst, New South Wales.
Rohrlach BD and Loucks RR (2005) Multi-million-year cyclic ramp-up of volatiles in a Economic Geology 50: 681–714.
lower-crustal magma reservoir trapped below the Tampakan copper-gold deposit by Steinberger I, Hinks D, Driesner T, and Heinrich CA (2013) Source plutons driving
Mio-Pliocene crustal compression in the southern Philippines. In: Porter TM (ed.) porphyry copper ore formation: Combining geomagnetic data, thermal constraints,
Super Porphyry Copper and Gold Deposits: A Global Perspective, vol. 2, pp. 1–43. and chemical mass balance to quantify the magma chamber beneath the Bingham
Adelaide: PGC Publishing. Canyon deposit. Economic Geology 108: 605–624.
Rojstaczer SA, Ingebritsen SE, and Hayba DO (2008) Permeability of continental crust Stefánsson A and Seward TM (2004) Gold(I) complexing in aqueous sulphide solutions
influenced by internal and external forcing. Geofluids 8: 128–139. to 500  C at 500 bar. Geochimica et Cosmochimica Acta 68: 4121–4143.
Rowan EL, Goldhaber MB, and Hatch JR (2002) Regional fluid flow as a factor in the Stoffell B, Appold MS, Wilkinson JJ, Mcclean NA, and Jeffries TE (2008) Geochemistry
thermal history of the Illinois basin: Constraints from fluid inclusions and the and evolution of Mississippi Valley-type mineralizing brines from the Tri-State and
maturity of Pennsylvanian coals. American Association of Petroleum Geologists Northern Arkansas districts determined by LA-ICP-MS microanalysis of fluid
Bulletin 86: 257–277. inclusions. Economic Geology 103: 1411–1435.
Rudnick RL and Gao S (2003) Composition of the continental crust. In: Rudnick RL Stowe CW (1994) Compositions and tectonic settings of chromite deposits through
(ed.) Treatise on Geochemistry, Vol 3: The Crust, pp. 1–64. Oxford: Elsevier. time. Economic Geology 89: 528–546.
Rusk BG, Reed MH, and Dilles JH (2008) Fluid inclusion evidence for Sun WD, Arculus RJ, Kamenetsky VS, and Binns RA (2004) Release of gold-bearing
magmatic-hydrothermal fluid evolution in the porphyry copper-molybdenum fluids in convergent margin magmas prompted by magnetite crystallization. Nature
deposit at Butte, Montana. Economic Geology 103: 307–334. 431: 975–978.
Russell MJ, Hall AJ, Boyce AJ, and Fallick AE (2005) 100th Anniversary Special Paper: Sverjensky DA (1986) Genesis of Mississippi Vallye type lead – Zinc deposits. Annual
On hydrothermal convection systems and the emergence of life. Economic Geology Review of Earth and Planetary Sciences 14: 177–199.
100: 419–438. Sverjensky DA, Shock EL, and Helgeson HC (1997) Prediction of the thermodynamic
Russell MJ, Solomon M, and Walshe JL (1981) The genesis of sediment-hosted, properties of aqueous metal complexes to 1000  C and 5 kb. Geochimica et
exhalative zinc þ lead deposits. Mineralium Deposita 16: 113–127. Cosmochimica Acta 61: 1359–1412.
Scott SD (1997) Submarine hydrothermal systems and deposits. In: Barnes HL (ed.) Taylor SR and McLennan SM (1995) The geochemical evolution of the continental
Geochemistry of Hydrothermal Ore Deposits, 3rd edn., pp. 797–875. New York: Wiley. crust. Reviews of Geophysics 33: 241–265.
Seo JH, Guillong M, and Heinrich CA (2009) The role of sulfur in the formation of Thom J and Anderson GM (2008) The role of thermochemical sulfate reduction in
magmatic-hydrothermal copper-gold deposits. Earth and Planetary Science Letters the origin of Mississippi Valley-type deposits. I. Experimental results. Geofluids
282: 323–328. 8: 16–26.
Seward TM (1984) The formation of lead(II) chloride complexes to 300  C – A Tilton JE and Lagos G (2007) Assessing the long-run availability of copper. Resources
spectrophotometric study. Geochimica et Cosmochimica Acta 48: 121–134. Policy 32: 19–23.
Seward TM and Barnes HL (1997) Metal transport by hydrothermal ore fluids. Tornos F and Heinrich CA (2008) Shale basins, sulfur-deficient ore brines and the
In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn., formation of exhalative base metal deposits. Chemical Geology 247: 195–207.
pp. 435–486. New York: Wiley. Townend J and Zoback MD (2000) How faulting keeps the crust strong. Geology
Shen W and Zhao PD (2002) Theoretical study of statistical fractal model with applications 28: 399–402.
to mineral resource prediction. Computers and Geosciences 28: 369–376. Ulrich T, Günther D, and Heinrich CA (1999) Gold concentrations of magmatic brines
Shinohara H, Kazahaya K, and Lowenstern JB (1995) Volatile transport in a convecting and the metal budget of porphyry copper deposits. Nature 399: 676–679.
magma column – Implications for porphyry Mo mineralization. Geology Ulrich T and Heinrich CA (2002) Geology and alteration geochemistry of the porphyry
23: 1091–1094. Cu–Au deposit at Bajo de la Alumbrera, Argentina. Economic Geology 97: 1863–1888.
Sibson RH (1992) Implications of fault-valve behavior for rupture nucleation and Vila T and Sillitoe RH (1991) Gold-rich porphyry systems in the Maricunga belt,
recurrence. Tectonophysics 211: 283–293. Northern Chile. Economic Geology 86: 1238–1260.
Sillitoe RH (2005) Supergene oxidized and enriched porphyry copper and related Von Damm KL (2005) Evolution of the hydrothermal system at the East Pacific
deposits. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Rise 9 500 N: Geochemical evidence for changes in the upper oceanic crust.
Economic Geology 100th Anniversary Volume, pp. 723–768. Littleton, CO: Society In: German CR, Lin J, and Parson LM (eds.) Mid-Ocean Ridges:
of Economic Geologists. Hydrothermal Interactions Between the Lithosphere and Oceans.
Sillitoe RH (2010) Porphyry copper systems. Economic Geology 105: 3–41. Geophysical Monograph Series, vol. 148, pp. 285–304. Washington, DC: American
Simmons SF and Brown KL (2007) The flux of gold and related metals through a Geophysical Union.
volcanic arc, Taupo Volcanic Zone, New Zealand. Geology 35: 1099–1102. Von Quadt A, Erni M, Martinek K, Moll M, Peytcheva I, and Heinrich CA (2011) Zircon
Simmons SF, White NC, and John DA (2005) Geological characteristics of epithermal crystallization and the lifetimes of ore-forming magmatic-hydrothermal systems.
precious and base metal deposits. In: Hedenquist JW, Thompson JFH, Goldfarb RJ, Geology 39: 731–734.
and Richards JP (eds.) Economic Geology 100th Anniversary Volume, Wagner T, Okrusch M, Weyer S, Lorenz J, Lahaye Y, Taubald H, and Schmitt RT (2009)
pp. 455–522. Littleton, CO: Society of Economic Geologists. The role of the Kupferschiefer in the formation of hydrothermal base metal
Simon G, Kesler SE, and Chryssoulis S (1999) Geochemistry and textures of mineralization in the Spessart ore district, Germany: Insight from detailed sulfur
gold-bearing arsenian pyrite, Twin Creeks, Nevada: Implications for deposition of isotope studies. Mineralium Deposita 45: 217–239.
gold in Carlin-type deposits. Economic Geology 94: 405–421. Wall VJ, Graupner T, Yantsen V, Seltmann R, and Hall AJ (2004) Muruntau, Uzbekistan:
Simon G, Kesler SE, Essene EJ, and Chryssoulis SL (2000) Gold in porphyry copper A giant thermal aureole gold (TAG) system. SEG 2004 Conference: “Predictive
deposits: Experimental determination of the distribution of gold in the Cu–Fe–S Mineral Discovery Under Cover.” SEG Conference & Exhibition, 27 September - 1
system at 400 to 700  C. Economic Geology 95: 259–270. October 2004, Perth, Western Australia; Extended Abstracts.
28 Fluids and Ore Formation in the Earth’s Crust

Watson EB and Baxter EF (2007) Diffusion in solid-Earth systems. Earth and Planetary Williams-Jones AE and Heinrich CA (2005) 100th Anniversary Special Paper: Vapor
Science Letters 253: 307–327. transport of metals and the formation of magmatic-hydrothermal ore deposits.
Weis P, Driesner T, and Heinrich CA (2012) Porphyry-copper ore shells form at stable Economic Geology 100: 1287–1312.
pressure-temperature fronts within dynamic fluid plumes. Science 338: 1613–1616. Yang KH and Scott SD (1996) Possible contribution of a metal-rich magmatic fluid to a
Wilde AR and Wall VJ (1987) Geology of the Nabarlek uranium deposit, Northern sea-floor hydrothermal system. Nature 383: 420–423.
Territory, Australia. Economic Geology 82: 1152–1168. Yardley BWD (2005) 100th Anniversary Special Paper: Metal concentrations in
Wilkinson BH and Kesler SE (2009) Quantitative identification of metallogenic epochs crustal fluids and their relationship to ore formation. Economic Geology
and provinces: Application to Phanerozoic porphyry copper deposits. Economic 100: 613–632.
Geology 104: 607–622. Zajacz Z and Halter W (2009) Copper transport by high temperature, sulfur-rich
Wilkinson JJ, Everett CE, Boyce AJ, Gleeson SA, and Rye DM (2005) Intracratonic magmatic vapor: Evidence from silicate melt and vapor inclusions in a basaltic
crustal seawater circulation and the genesis of subseafloor zinc-lead mineralization andesite from the Villarrica volcano (Chile). Earth and Planetary Science Letters
in the Irish orefield. Geology 33: 805–808. 282: 115–121.
Wilkinson JJ, Stoffell B, Wilkinson CC, Jeffries TE, and Appold MS (2009) Anomalously Zajacz Z, Hanley JJ, Heinrich CA, Halter WE, and Guillong M (2009) Diffusive
metal-rich fluids form hydrothermal ore deposits. Science 323: 764–767. reequilibration of quartz-hosted silicate melt and fluid inclusions: Are all
Williams PJ, Barton MD, Johnson DA, et al. (2005) Iron oxide copper-gold metal concentrations unmodified? Geochimica et Cosmochimica Acta
deposits: Geology, space-time distribution, and possible modes of origin. 73: 3013–3027.
In: Hedenquist JW, Thompson JFH, Goldfarb RJ, and Richards JP (eds.) Economic Zajacz Z, Seo JH, Candela PA, Piccoli PM, Heinrich CA, and Guillong M (2010) Alkali
Geology 100th Anniversary Volume, pp. 371–405. Littleton, CO: Society of metals control the release of gold from volatile-rich magmas. Earth and Planetary
Economic Geologists. Science Letters 297: 50–56.
Williams AE and McKibben MA (1989) A brine interface in the Salton Sea geothermal Zezin DY, Migdisov AA, and Williams-Jones AE (2011) The solubility of gold in H2O–
system, California – Fluid geochemical and isotopic characteristics. Geochimica et H2S vapour at elevated temperature and pressure. Geochimica et Cosmochimica
Cosmochimica Acta 53: 1905–1920. Acta 75: 5140–5153.

You might also like