Source Pluton Bingham

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

©2013 Society of Economic Geologists, Inc.

Economic Geology, v. 108, pp. 605–624

Source Plutons Driving Porphyry Copper Ore Formation:


Combining Geomagnetic Data, Thermal Constraints, and Chemical Mass Balance to
Quantify the Magma Chamber Beneath the Bingham Canyon Deposit
INGO STEINBERGER,1,† DONALD HINKS,2 THOMAS DRIESNER,1 AND CHRISTOPH A. HEINRICH1,3
1 Departement Erdwissenschaften, ETH Zürich, 8092 Zürich, Switzerland
2 Rio Tinto Exploration, Salt Lake City, Utah 84116, USA
3 Mathematisch-naturwissenschaftliche Fakultät, Universität Zürich, 8006 Zürich, Switzerland

Abstract
The formation of porphyry copper deposits requires a focused flux of magmatic fluid, expelled from a large
reservoir of water-, metal-, and sulfur-rich magma. The dimensions of this usually hidden magma reservoir are
difficult to determine but can be constrained by combining geophysical observations with thermal constraints
and the mass balance imposed by the chemical enrichment of elements in the deposit. Here we show that an
internally consistent scenario can be derived for the world-class Cu-Mo-Au deposit at Bingham Canyon (Utah,
United States), which quantifies the essential characteristics, approximate dimension, and temporal evolution
of a large pluton that generated the deposit.
The mineralized district shows a distinct WSW-ENE–striking magnetic anomaly indicating a large intrusive
body underlying the sedimentary host rocks of the Oquirrh Mountains. Modeling the deep body by geomag-
netic methods is possible because of the high contrast in magnetic susceptibility between sedimentary host
rocks and intrusive rocks and because a former volcanic edifice is largely eroded. Additional constraints from
drilled geology and district-wide outcropping rocks, including partial demagnetization by hydrothermal alter-
ation on the mine scale, restrict the range of possible solutions to a broadly laccolith-shaped intrusion with a
volume of approximately 1,400 to 3,000 km3. From the roof of the laccolith, several smaller subvolcanic stocks
and dikes protrude to the present surface, of which a major one is hosting the Bingham Canyon deposit. The
roof of the laccolith probably lies between 2 and 3.5 km below the bottom of the present open-pit mine, and
the average thickness of the laccolith is constrained between 2 and 3.5 km.
Thermal modeling, using pluton dimensions derived from the geologic and geomagnetic modeling, indicates
that a single laccolith with a magma volume of ~2,000 km3 beneath Bingham would have solidified within about
230,000 years or less. Comparison of the thermal models with published high-precision geochronologic data
and petrologic constraints suggests a scenario in which about 1,000 km3 of magma was encapsulated by inward
crystallization of the pluton after the preore equigranular monzonite stocks solidified and extrusive volcanism
was probably terminated. This encapsulated reservoir was close to water saturation and contained approxi-
mately 150 billion metric tons (Gt) of magmatic water for subsequent closed-system fractionation and eventual
fluid expulsion driving porphyry copper mineralization.
Chemical mass balance shows that the known metal endowment and mapped mass of vein quartz within the
deposit can be advected and precipitated by a fluid mass that is slightly smaller than the available 150 Gt of
water. A conservative estimate indicates that 115 Gt of water is sufficient to precipitate all the quartz associated
with successive Cu-Au- and Mo-stage veins as well as their barren precursors. According to our thermal model,
approximately 250 km3 of quartz monzonite magma with a temperature of about 690°C remained partially liq-
uid some 215,000 years after initial intrusion of the laccolith. At that point, it expelled almost simultaneously the
quartz monzonite porphyry and the main mass of accumulated fluid, generating most of the vein quartz in the
quenched porphyry and the adjacent older rocks. Petrographic evidence indicates that the ore metals precipi-
tated near the end of individual pulses of quartz veining that followed recurrent but waning pulses of porphyry
intrusion. Considering published experimental solubility data as well as ore metal contents in fluid inclusions,
a small fraction of the available fluid mass is sufficient to transport and precipitate all the ore metals after an
initial fluid pulse precipitated most of the quartz. However, the total amount of sulfur present in the deposit,
which includes Cu and Mo sulfides as well as a major addition of pyrite, would be facilitated by addition of a
mafic magma input into the residual magma chamber that contained the evolved felsic magma. This magmatic
injection probably triggered the emplacement of the mineralized porphyries, consistent with the more mafic
composition of some of the latest porphyry dikes and the CO2-rich nature of ore-related fluid inclusions.

Introduction between the metal content in a deposit and the related por-
PORPHYRY Cu deposits are centered on volumetrically small phyry intrusion show that the latter is orders of magnitude too
and highly fractured porphyritic dikes and stocks that protrude small to provide the metal that is enriched in the deposit
from upper crustal magma chambers (Emmons, 1927; Sillitoe, (Dilles, 1987). Hence, a sizeable magma chamber is required
1973, 2010; Richards, 2011). Mass-balance considerations not only to act as source for the porphyry magma but also to
provide the magmatic-hydrothermal fluids and dissolved ore
† Corresponding author: e-mail, ingo.steinberger@alumni.ethz.ch components and the energy for initiating and driving the ore
Submitted: July 7, 2011
0361-0128/13/4109/605-20 605 Accepted: September 30, 2012

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
606 STEINBERGER ET AL.

formation (Candela, 1997; Cloos, 2001; Proffett, 2009; Tosdal for quantitative interpretation of a high-resolution aeromag-
et al., 2009). Overpressured magmatic fluids create fractures netic survey (Kennecott Utah Copper, proprietary data) that
through which fluids ascend from the larger source intrusions covers the porphyry system and the inferred magma chamber.
toward the surface ( Gustafson and Hunt, 1975; Burnham and Geologic and geochemical data are available from more
Ohmoto, 1980; Gustafson and Jorge Quiroga, 1995; Red- than a century of mining at Bingham Canyon (Krahulec, 1997,
mond et al., 2004). During their ascent, the fluids precipitate 2010), and exploration drilling has defined the surface distrib-
dissolved metals (e.g., Cu, Mo, Au) and gangue minerals ution of rocks and structures ( Hintze, 1980; Laes et al., 1997;
(mainly quartz) that were extracted from the magma Biek et al., 2007; Kloppenburg et al., 2010) down to depths of
(Landtwing et al., 2010; Redmond and Einaudi, 2010). In 1.9 km below the pit bottom. This exploration knowledge is
rare examples, such as the ~6 million metric tons (Mt) Cu de- expanded by petrologic and geochemical studies by Keith et
posit in Yerington, United States (Dilles, 1987), later tectonic al. (1997), Ballantyne et al. (1997), Hattori and Keith (2001),
tilting and faulting has exposed the formerly subjacent Maughan et al. (2002), and others. The mine-scale architec-
magma chamber from which the mineralized porphyry dikes ture of the hydrothermal system has been reconstructed by
protrude. In this case, geologic mapping allowed an estima- mapping vein and dike relationships as well as measuring ori-
tion of 65 km3 for the volume of the Luhr Hill Granite, which entation and density of several generations of mineralizing
is large enough to provide the mass of the mined Cu assum- veins (Gruen et al., 2010; Kloppenburg et al., 2010; Redmond
ing a typical source concentration in normal calc-alkaline and Einaudi, 2010), providing a basis for estimating the quan-
magmas (Dilles, 1987; Cline and Bodnar, 1991; Dilles and tities of ore metals, sulfur, and vein quartz introduced into the
Gans, 1995). In addition to providing a source of metals and mine volume. Petrographic observations, fluid inclusion
fluids, the magma chamber provides thermal energy to the analyses, and solubility experiments are available to quantify
hydrothermal system. Both of these essential requirements the sequence of hydrothermal mineral precipitation and the
for ore formation relate the potential size of an ore deposit to capacity of fluids to transport and deposit the ore-forming
the volume of the magma reservoir. This volume also defines components (Redmond et al., 2004; Akinfiev and Diamond,
the time window in which magmatic hydrothermal activity 2009; Landtwing et al., 2010; Zajacz et al., 2011).
may take place, since the expulsion of magmatic fluids will Neither geophysical modeling nor chemical mass balances
cease after all melt is crystallized. can provide unique answers regarding the size of a porphyry-
The lifetimes of ore-forming systems and the duration of mineralizing magma chamber. The purpose of this study is to
the actual porphyry Cu mineralization are a matter of ongo- test whether an internally consistent interpretation can match
ing discussion (Cathles et al., 1997; Sillitoe, 2010; von Quadt the dimension, approximate geometry, and temporal evolu-
et al., 2011). Present estimates come from a variety of meth- tion of a large hidden magma reservoir that drove the forma-
ods, including geochronology and numerical simulation of tion of a world-class ore deposit of well-known metal endow-
contributing processes operating at different scales. Charac- ment. We show that high-resolution geophysical data can be
teristic process durations range from several millions of years linked with geology and basic chemical and physical require-
for the lifetime of a volcano-plutonic complex (Halter et al., ments to formulate a quantitative reconstruction of the
2004; Maksaev et al., 2004; Sillitoe and Mortensen, 2010), processes required for ore formation. These physical require-
through tens to hundreds of thousand years for magmatic ac- ments are imposed by heat balance, geochemical mass bal-
tivity at the scale of a porphyry deposit (Cannell et al., 2005; ance, fluid compositions, and geochronology. In a first step of
McInnes et al., 2005), to as short as a few to a hundred years this study, we constrain the range of geologically permissible
for the precipitation of copper ore shells accompanying potas- volumes and emplacement depths of an inferred intrusive
sic alteration (Cathles, 1981; Cathles and Shannon, 2007). body beneath the Bingham Canyon deposit by two- and
The source of porphyry-mineralizing fluids and their parent three-dimensional geomagnetic modeling. In a second step
magmas are by their nature located several kilometers below we estimate the maximum time interval for magma solidifica-
exposed deposits and generally inaccessible, except in tecton- tion and ore formation, by comparing simple thermal model-
ically dissected examples like Yerington. Moreover, even large ing with high-precision geochronological data. In a third and
and long-lived centers of hydrous magmatism do not neces- final step, we delineate the most plausible corner points in
sarily develop major upper-crustal magma chambers, for ex- the evolution of the magmatic system and the hydrothermal
ample, the Tatara-San Pedro Complex in Chile (Dungan et mass transfer with respect to temperature, pressure (depth),
al., 2001), or the Batur Volcanic Field in Bali, Indonesia and masses of solutes (Cu, Mo, S, and silica) as well as solvent
(Reubi and Nicholls, 2005). Geophysical observations may (saline water) by comparing the scale and dimensions derived
provide quantitative constraints on the volume and geometry from the conjunction of all constraints.
of hidden magma chambers required for porphyry mineral-
ization. At the world-class Bingham Canyon Cu-Mo-Au de- Geology: Intrusion, Veining, and Mineralization
posit, United States, the magnetic susceptibility contrast be- The Bingham Canyon Cu-Mo-Au deposit is located in the
tween the intrusive rocks and the host sediments is sufficient Oquirrh Mountains (Fig. 1), the easternmost fault block in the
to create a distinct magnetic anomaly that permits modeling Great Basin in the western United States. The Oquirrh Moun-
of a deep-seated former magma chamber. Samples from out- tains are a 55-km-long N-S-striking range, truncated to the
cropping intrusive rocks and drill cores allow estimation of west by Basin and Range faults and dipping to the east into
the magnetic properties of the magma chamber material. Re- the basin sediments of the Salt Lake Valley. The Salt Lake
gional-scale geomagnetic surveys (Melker and Geissman, 1997; Valley is limited to its east by the Wasatch Front fault—a still
Bankey et al., 1998) are available to provide the framework active, continental-scale fault separating the Great Basin from

0361-0128/98/000/000-00 $6.00 606

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
on 14 December 2022
400 450 390 410
A B

by Raymond Rivera Cornejo


Utah
4500

A’
A
0
B
Wasatch Front

Great Basin

0361-0128/98/000/000-00 $6.00
Bingham Stock
50

50

Colorado Plateau
100

Settlement Canyon
s.
Mt
Salt Lake City tch
150

sa Uinta
Wa Alta Park City
trend 200
100

4500
10 0
100
4480

Last Chance stock


Little Cotton-
wood Canyon

ta
Uin nd 50

607
tre
Oquirrh Mts. Stockton
100

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


50
390

4450
Alluvium N
Tertiary volcanics 14.2°
km northing

Tertiary intrusives
4460

Wasatch Front Undivided strata


0

Major Basin Magnetic A Sectionused in Fig . 4


Isoanomaly [nT] 5 and Range declination
400 450 Fault 1994 km easting, UTM zone 12N, WGS 84 410
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT

FIG. 1. A. Simplified geologic map of north-central Utah (based on Hintze, 1980), showing the easternmost fault block of the Basin and Range province (the Oquirrh
Mountains) adjacent to the Wasatch Front fault. Both cut an ENE-striking positive magnetic anomaly (the Uinta Trend), indicated by contour lines showing the devia-
tion of total magnetic intensity from the Geomagnetic Reference Field in nT, reduced to pole and upward continued to 3,300 m (Bankey et al., 1998). The anomaly is
prominent in the Wasatch Mountains where large plutons are exposed, becomes attenuated as it crosses the Salt Lake Valley, then crosses the Oquirrh Mountains, and
is eventually truncated by a downdropping fault to the west. B. Enlarged map for the Oquirrh Mountains showing outcrop geology (based on Laes et al., 1997) and mag-
netic anomalies (Kennecott Utah Copper, 1994, unpub. data). Coordinates (blue) give scale in kilometers (UTM zone 12N, WGS 84).
607
608 STEINBERGER ET AL.

the Wasatch Mountains and the Rocky Mountains on the Col- Smaller occurrences with very similar petrographic charac-
orado Plateau. These young faults dissect an underlying E-W- teristics include the Settlement Canyon (Fig. 2B) and the
oriented structure of Precambrian origin, the Uinta Trend Stockton monzonites (Fig. 2C) that crop out several kilome-
(Erickson, 1976). Bingham Canyon and other magmatic- ters to the west from the mine. The unaltered mineral as-
hydrothermal centers are located on this trend, which is visi- semblage of the equigranular monzonite consists of plagio-
ble as a regional magnetic anomaly (Fig. 1A). clase and K-feldspar in similar proportions, accompanied with
The geologic evolution of Bingham Canyon has been docu- minor quartz and 31 vol % mafic minerals to which magnetite
mented by numerous workers including Moore (1973), contributes up to 4 vol % (Moore, 1973). All minerals are <3
Lanier et al. (1978), Waite et al. (1997), and Redmond and mm in size, pointing to rapid cooling of a phenocryst-poor
Einaudi (2010). Outcropping magmatic activity started with melt. The equigranular monzonite at Bingham Canyon is
the emplacement of the equigranular monzonite (Fig. 2A) drilled to a depth of 1.9 km below pit bottom; its volume is es-
with an average zircon 206Pb/238U age of 38.55 ± 0.19 Ma timated with 5.9 km3 and continuing toward greater depth.
(Parry et al., 2001). The equigranular monzonite intruded The quartz monzonite porphyry (Fig. 2D) intruded into the
into Pennsylvanian quartzites and intercalated limestone lay- equigranular monzonite at its northern flank and initiated the
ers (Fig. 1B). Main occurrences of equigranular monzonite most intense phase of mineralization, veining, and hydrother-
are the Last Chance (Fig. 2A) and the Bingham stocks. mal alteration. Its drilled volume is 0.8 km3 (open to depth).

A Last Chance stock B Settlement Canyon C Stockton


monzonite monzonite monzonite

Hb

Bt

Pl

Ksp
1 cm 1 cm 1 cm

D Bingham Canyon Stockton Canyon


E Bingham F Stockton
quartz monzonite porphyry quarzporphyry
latite monzonite porphyry quartz monzonite porphyry

Cp Cp
Bt
Qz Ksp

Bt Qz
Pl->Ser

Ksp
Ksp

1 cm 1 cm 1 cm

FIG. 2. Rock samples from intrusions across the Oquirrh Mountains exposed above the magnetic anomaly. A-C. Equigran-
ular monzonite from three separate intrusions, with near-identical modal mineralogy and magnetic properties. Texture varies
slightly toward finer grained minerals in thin dikes, such as Settlement Canyon (B). D-F. Porphyritic quartz monzonites with
identical phenocryst assemblage but variable groundmass. (D) is the 300-m-wide quartz monzonite porphyry dike in the min-
eralized center at Bingham Canyon and therefore veined, whereas the porphyritic part of the Stockton intrusion (F) is less
than 50 m wide. Both porphyries intruded earlier equigranular monzonite at each location. The latite porphyry (E) is a late,
volumetrically minor intrusion of the Bingham Canyon intrusive suite, illustrating the trend to more mafic compositions of
the matrix. Abbreviations: Bt = biotite, Cp = chalcopyrite, Hb = hornblende, Ksp = potassium feldspar, Pl = plagioclase; Pl-
>Ser = pseudomorphs of sericite after plagioclase, Qz = quartz.

0361-0128/98/000/000-00 $6.00 608

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 609

Although the quartz monzonite porphyry is strongly altered, Little Cottonwood Canyon, Alta and Park City plutons (John,
the initial phenocryst mineral assemblage can be traced back 1989, 1997; Vogel et al., 2001) . The Little Cottonwood
to plagioclase and orthoclase in similar proportions, signifi- Canyon pluton is truncated to the west by the Wasatch fault,
cant quartz and 25 vol % mafic and accessory minerals yet the magnetic anomaly continues to the west, attenuated by
(Moore, 1973). Orthoclase phenocrysts up to 2 cm in size the alluvial cover. The next exposed rocks to the west are the
point to longer residence time and/or fluid saturation during Oquirrh Mountains. Here, the lack of alluvial cover amplifies
late partial crystallization of this magma prior to its emplace- the magnetic anomaly, and the outcrops of the Bingham
ment. Rapid cooling quenched the porphyry matrix so that Canyon intrusive suite (comprising the Last Chance stock
the unaltered groundmass is aplitic (grains of 50–100 μm) (Fig. 2A) and the Bingham Canyon stock) and the Stockton
with subequal amounts of quartz and orthoclase. After em- intrusive complex (Fig. 2C, F) generate local maxima on the
placement of the quartz monzonite porphyry, minor porphyry smooth long-wavelength anomaly (Fig. 3). The large-scale
dikes of more mafic composition intruded, each restarting anomaly indicates a larger intrusive body at depth, compara-
Cu-Au mineralization but contributing only ~15% to veining ble to the outcropping intrusions in the Wasatch Mountains.
and total Cu-Au mineralization (Redmond and Einaudi, The intrusions within the Wasatch Mountains (Vogel et al.,
2010). Of those minor dikes, the latite porphyry (Fig. 2E) has 2001) and beneath the Oquirrh Mountains (Parry et al., 2001;
a drilled volume of 0.2 km3 and the latest quartz latite por- von Quadt et al., 2011) were emplaced between ~39 and ~ 30
phyry and biotite porphyry have a drilled volume of 0.06 km3. Ma and probably share a common source region in the deep
All late porphyries contain abundant centimeter-sized ortho- crust or upper mantle (Pettke et al., 2010). Individual ex-
clase phenocrysts and the quartz latite porphyry additionally posed plutons have a typical E-W extent of up to 15 km.
has subcentimeter-sized, rounded quartz phenocrysts.
Three main types of quartz veins occur in the Bingham
Canyon deposit. Quartz veins associated with Cu-Au mineral-
ization follow each porphyry intrusion with diminishing de-
gree of intensity. These veins contain two petrographic gen-
erations of quartz. Early barren quartz represents 90 to 99 vol
% of the total vein volume. A second quartz generation com-
prises 1 to 10 vol % of the veins and is associated with (or
postdates) the precipitation of chalcopyrite and bornite (Red-
mond et al., 2004). The economically mineralized Cu-Au ore-
body has the shape of a molar tooth that is centered on the
quartz monzonite porphyry and extends mainly into the
equigranular monzonite. The second vein type, quartz-
molybdenite veins, crosscut all other veins of magmatic origin
as well as the latest porphyry dikes. Mo mineralization has a
similar shaped distribution as Cu mineralization but is later-
ally less extensive and extends to greater depths. The third
vein type, pyrite veins with minor quartz and extensive
sericite alteration, postdate economic metal introduction, but
added a major quantity of sulfur during the waning stages of
the magmatic-hydrothermal systems, in total about 960 mil-
lion metric tons (Mt) S (Hattori and Keith, 2001). Mapped
volumes of vein quartz (Gruen et al., 2010) and the total mass
of ore metals provide independent constraints on hydrother-
mal mass transfer. Published economic reserves plus previ-
ously mined Cu (including several peripheral deposits)
amount to 31.5 Mt (Krahulec, 2010), but unpublished esti-
mates by Kennecott Utah Copper indicate at least 55.9 Mt of
economic and uneconomic copper mineralization, plus an es-
timated total quantity of 2,600 t Au and 1.4 Mt Mo
(Schroeder, pers. commun. 2012).
FIG. 3. Image of parts of the aeromagnetic anomaly over the Oquirrh
Regional Magnetic Anomaly and Magnetic Properties Mountains and adjacent basins (Kennecott proprietary data, reduced to pole,
of Exposed Rocks shaded, and Definitive Geomagnetic Reference Field subtracted; print
image slightly distorted in nonlinear fashion so that scale and northing are ap-
A regional, ENE-striking magnetic anomaly—the Uinta proximate only). The anomaly caused by the inferred pluton has a flat top
Trend (Erickson, 1976)—starts in the Wasatch Mountains, with a long wavelength and strikes in WSW-ENE direction. This anomaly un-
crosses the Salt Lake Valley, and continues through the Oquirrh derlies short-wavelength maxima that are marking exposed intrusions of
Mountains until it is truncated to the west by Basin and equigranular monzonite at Bingham Canyon and several locations farther to
the west. The Bingham Canyon suite is composed of the Last Chance stock
Range faults (Fig. 1). The anomaly is correlated with intrusive (south) and the partly mineralized Bingham stock (north, ~pit outline), in
rocks, making up a major E-W–extended batholith. In the which hydrothermal destruction of igneous magnetite causes a small strongly
Wasatch Mountains, the anomaly is generated by the exposed negative anomaly centered on the quartz monzonite porphyry.

0361-0128/98/000/000-00 $6.00 609

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
610 STEINBERGER ET AL.

In order to interpret the geomagnetic data, the geology of monzonites are followed by quartz monzonite porphyries,
the Oquirrh Mountains can be summarized into four groups even though the locations are separated by 20 km. The sus-
of material with different magnetic properties: sedimentary, ceptibility of igneous rocks in the district was measured by a
intrusive and volcanic rocks, and alluvial cover (Fig. 1). contractor of Kennecott Utah Copper in 1963 (Fig. 4). From
The first group represents the majority of outcropping hard a total of 28 reported measurements on surface samples of
rocks in the central Oquirrh Mountains. These are Carbonif- equigranular monzonite of the Last Chance stock, Bingham
erous to Permian sediments ranging from pure dolomites and stock, Settlement Canyon and Stockton intrusions, we ex-
limestones to well-sorted siltstones, which were deposited in cluded five very low outliers as likely altered samples. The re-
a passive continental margin environment (Stevens and maining susceptibility measurements cluster around an arith-
Armin, 1983; Tooker, 1999). The arithmetic mean magnetic metic mean of 0.023 (dimensionless number, SI system) with
susceptibility from 4,708 borehole measurements of the sed- a standard deviation of ±0.010. Consistent with their miner-
iments is 0.00008 ± 0.00091 (dimensionless number, SI sys- alogical similarity, all intrusions of equigranular monzonite
tem), i.e., they show no significant magnetic susceptibility have a comparable magnetic susceptibility. Seven measure-
(Fig. 4). ments of the Last Chance stock average at 0.026 ± 0.004, nine
The second group comprises the Tertiary intrusive rocks. measurements of the Settlement Canyon dike average at
Many of the intrusions in the Oquirrh Mountains are subver- 0.024 ± 0.008, and seven measurements of the Stockton in-
tical dikes occurring in several centers along the broad E-W- trusion average at 0.017 ± 0.013. No unaltered porphyries
striking magnetic trend. They can be subdivided by their tex- were available for measuring magnetic susceptibility. Mine-
ture into older equigranular (Fig. 2A-C) and younger scale alteration causes local demagnetization due to sulfida-
porphyritic rocks (Fig. 2D-F). Grain size of the equigranular tion of magnetite, as shown by shallow aeromagnetic lows
monzonites varies slightly as a function of dike thickness, e.g., where mineralized intrusive rocks crop out in the mine (Fig.
the narrow Settlement Canyon dike (Figs. 1B, 2B) is finer 3), or are known from drilling. Exposed plutonic rocks in the
grained compared to the larger Last Chance stock (Fig. 2A), Wasatch Mountains show a somewhat greater range but
but the type and proportion of minerals is similar in all loca- bracket the equigranular monzonites of the Oquirrh Moun-
tions. Also the magma evolution is similar in Stockton (Fig. 2C, tains, with an average susceptibility of 0.030 for the Alta gra-
F) and in Bingham Canyon (Fig. 2A, D): earlier equigranular nodiorite (Bartley et al., 2006) and of 0.015 ± 0.011 for the
lower part of the quartz monzonite of Little Cottonwood
Canyon (165 measurements).
Tertiary volcanic and volcaniclastic rocks associated with
the intrusive suite are the third lithologic group affecting the
aeromagnetic pattern. Their drilled thickness exceeds 1,000
m in the Salt Lake basin (Biek et al., 2007), but in the Oquirrh
Mountains, they are only locally preserved as a blanket of less
than 300-m thickness, covering at most 5% of the large mag-
netic anomaly (Fig. 1B). They are clearly visible by their het-
erogeneous shallow signal in the aeromagnetic image (Fig. 3).
The mean magnetic susceptibility of 53 samples of varying
lithology is 0.019 ± 0.015 (Fig. 4). The larger standard devia-
tion at somewhat lower average reflects the large variability of
primary lithologic units and degrees of alteration.
The last group comprises alluvial sediments that are filling
the extensional basins. Since the largest fraction of these de-
posits originates from the Paleozoic sediments, we treat them
as not susceptible to the geomagnetic field.
To model the long wavelength anomaly, we assume an in-
trusive body with a magnetic susceptibility approximating that
of outcropping equigranular monzonite. This is justified by
the mineralogical and magnetic homogeneity of all monzonite
stocks and the similar intrusion sequence, exposed at different
locations above the broad geomagnetic anomaly, indicating
that the subvolcanic stocks originated from the roof of a com-
mon large pluton. To model this pluton, we approximated its
magnetic susceptibility as 0.023, the average of the measure-
FIG. 4. Histogram of measured magnetic susceptibilities in the Oquirrh ments of the equigranular monzonite (Fig. 4). We then varied
Mountains in dimensionless SI units. The 76 magmatic rocks sampled from this value within ±0.010 susceptibility units to determine the
the surface have a wide range of susceptibility, whereas the spread in the vol- sensitivity of modeling results to our material assumption and
canic rocks is larger than in the intrusive rocks due to their higher variability also considered the possibility of a significantly more mag-
in mineralogy and alteration. The 4708 measured sedimentary rocks sampled
from drill holes have a significantly lower susceptibility, many of them below
netic component (e.g., mafic underplating or cumulate layer)
the level of detection (l.o.d.). Abbreviations: bc = Bingham Canyon, se = Set- (Table 1). The sedimentary country rocks and basin-fill sedi-
tlement Canyon, st = Stockton. ments were modeled with zero magnetic susceptibility.

0361-0128/98/000/000-00 $6.00 610

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 611

Sparse measurements for remanent magnetization of the geometry and magnetic susceptibility. Only a limited range of
igneous rocks vary over four orders of magnitude for the same solutions, for relatively shallow intrusions of laccolithic shape,
lithology (Melker and Geissman, 1997). With a median mag- were found to plausibly explain (Fig. 5B-D) the smooth ESE-
netization of 65 mA/m (1st quartile: 11, 3rd quartile: 784 WNW–extending anomaly beneath the Oquirrh Mountains
mA/m), the Koenigsberger ratio amounts to 0.066, i.e., the re- without violating geologic constraints from exploration
manent magnetization is two orders of magnitude less intense drilling. Very shallow intrusions would have to be thin to ex-
than induced magnetization, so that remanent effects can be plain the height of the magnetic anomaly using any plausible
neglected. susceptibility, but predicted roof elevations above present-
day sea level conflict with constraints from exploration
Geomagnetic Forward Modeling: drilling, which has never reached the roof of the intrusion
the Oquirrh Mountains Laccolith (e.g., Fig. 5A). Bodies with a near-spherical or diapiric (“in-
The basis for this study is a terrane-contouring aeromag- verted drop”) cross section are unable to reproduce the rela-
netic survey flown in 1994 by Kennecott Utah Copper (Fig. tively flat top and rather steep northern and southern flanks
3). It was performed with a line spacing of 200 m and flown of the anomaly (Fig. 3). Likewise, any intrusive bodies placed
by helicopter at a constant height of 50 m above ground. The at greater depths, including solutions assuming higher mag-
local Definitive Magnetic Reference Field at that time had a netic susceptibility, require a roof structure with laterally ex-
declination of 14.2° and an inclination of 66.1° with an inten- tensive shallow ridges along the northern and southern edge
sity of 53,786 nT. of the intrusions, to explain the magnetic gradient at the flank
Two-dimensional potential field treatment was performed of the anomaly (e.g., Fig. 5F). This may seem acceptable in
using Geosoft® Oasis Montaj, and all two-dimensional for- two dimensions but is rather implausible considering the
ward modeling was performed with the PotentQ plugin to WSW-ENE extension of the northern and southern flank of
Oasis Montaj. We reduced the number of data points along the anomaly, which would require two tube-like ridges along
the flight lines to 30% to enhance computing speed, and sub- both edges of a deep laccolith. Thus, we conclude that a sim-
tracted the Definitive Magnetic Reference Field from the ple, convex laccolith with an intrusion floor between 3,000
total magnetic intensity. To attenuate high frequency (shallow) and 5,000 m below sea level and a thickness between 2 and
signals and to concentrate on the effects of deep bodies, the 3.5 km (i.e., with an intrusion roof at 3.5−2 km below the bot-
field was upward continued to 3,300 m above sea level, i.e., to tom of the present mine) can most plausibly explain all ob-
a constant level near the highest flight positions (the ground servational data. We also explored the possibility of a basal
surface is between 1,300 and 3,220 m above sea level). The mafic layer in the intrusion, with a doubled magnetic suscep-
potential field was reduced to pole, which eliminates distor- tibility (0.046), as petrological observations of Keith et al.
tions in the shape, sign, and position of signals and made ge- (1997) and Hattori and Keith (2001) may imply. This favors
ologic interpretation of the modeling process more intuitive. the shallower solutions and overall thinner laccoliths. A hori-
In order to forward model the body causing the observed zontal floor was assumed throughout, because the data are
magnetic anomaly, we started with separate, two-dimensional less sensitive to shape variations in the lower contact com-
sections. A set of N-S–striking, W-E–staggered line sections pared to the roof of the intrusion. Thin extensions toward the
was extracted from the topographic and aeromagnetic data, at north and south of the laccolith improve the fit toward the far-
positions avoiding known regions of shallow hydrothermal al- field baseline geomagnetic intensity. They were tentatively
teration. Since these two-dimensional sections through the chosen to mimic a possible feeder dike along the southern
magnetic anomaly contain three-dimensional contributions, edge of the intrusion (Fig. 5), inspired by exposed granitoid
we modeled each section using a set of three-dimensional intrusions like Torres del Paine (Michel et al., 2008), but this
prisms; the shapes of which were interactively refined until feature has a low geophysical reliability. Despite the low
they matched the two-dimensional anomaly. This procedure Koenigsberger ratio of the magnetized rocks, we also tested
simplified the reconstruction by avoiding a full three-dimen- the influence of remanent magnetization. The shift in the re-
sional approach that is disturbed by numerous shallow fea- sult was marginal and in the light of the high variability in re-
tures including local alteration and technical installations, but manent magnetization probably not significant; therefore we
nevertheless accounts for the limited three-dimensional ex- did not consider the influence of remanent magnetization any
tension and orientation of all shallow stocks known from ex- further. Figure 6 summarizes the results from our sensitivity
ploration drilling and the visually obvious E-W extent of the analysis, showing that the cross-sectional area of selected so-
deep anomaly. In the absence of deep drilling constraints, lutions varies in an approximately linear function with the as-
mapped surface outcrops of susceptible rocks (Laes et al., sumed depth of the intrusion floor. The gray band indicates
1997) were assumed to be subvertical stocks protruding from the sensitivity of solutions to varying magnetic susceptibility
the roof of the deep magmatic body, or thin volcanic blankets. by 0.010 (SI), i.e., within one standard deviation of equigran-
Modeling was confined to the upthrown Oquirrh Mountains ular monzonites or their deviation from the average value of
block between the bounding western fault and the Salt Lake the volcanic rocks.
basin (about 25 km in WSW-ENE extent; Fig. 1A), even Three-dimensional shapes and volumes of possible intru-
though the body may extend beneath cover toward the west sive bodies were obtained by connecting the W-E–staggered,
adjacent basin. N-S–striking, and independently computed two-dimensional
A suite of two-dimensional solutions along the stack of N-S sections using the design software Rhinoceros®. Figure 7 shows
sections were interactively calculated to assess the sensitivity the three-dimensional solutions that are extending in a WSW-
of magnetic responses to varying assumptions about pluton ENE direction beneath the Oquirrh Mountains. It includes

0361-0128/98/000/000-00 $6.00 611

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
612 STEINBERGER ET AL.

A Laccolith floor at 1000 m below B Laccolith floor at 3000 m below


sea level predicts intrusives sea level: Possible solution
Anomaly [nT] where sediments are drilled:
impossible solution
160 160

80 80

0 0

km northing
4470 4480 4490 4470 4480 4490
Depth [km]

0 0

4 4

8 8

C Laccolith floor at 5000 m D Laccolith floor at 5000 m below


below sea level: Less detailed, sea level and underplated with
but possible solution mafic magma: Possible solution
Anomaly [nT]

with reduced volume


160 160

80 80

0 0

km northing
4470 4480 4490 4470 4480 4490
Depth [km]

0 0

4 4

8 8

E Laccolith floor at 7000 m below F Laccolith floor at 9000 m below


sea level: reproduces anomaly sea level: reproduces anomaly
with concave shape with increasingly eccentric shape
Anomaly [nT]

160 160

80 80

0 0

km northing
4470 4480 4490 4470 4480 4490
Depth [km]

0 0

4 4

8 8

FIG. 5. A representative range of two-dimensional models (brown bodies causing the signal and red lines depicting the
calculated signal), explaining the observed magnetic anomaly along section A-A' in Figures 1 and 3 (blue lines; all data re-
duced to pole; northing in km of UTM zone 12N). Forward modeling assumes a cylindrical main laccolith extending per-
pendicular to the section, whereas the subvertical apophysis is representing the westernmost end of the Bingham Canyon in-
trusive stock and extends only 3 km at an angle of 57° to the section, based on known geology from surface outcrop and
drilling. Model solutions B to D denote solutions that are permissible within known geology, model A contradicts drilling re-
sults and models E and F require an indentation and in extreme case (F) two cylindrical ridges that are not plausible con-
sidering the full three-dimensional data set (see text).

0361-0128/98/000/000-00 $6.00 612

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 613

area (Fig. 6). Two full three-dimensional solutions were cal-


culated with floor depths at 3,000 and at 5,000 m below pre-
sent-day sea level, yielding total laccolith volumes of 1,800
and 2,500 km3, respectively. As a final test, full magnetic
anomalies for the two three-dimensional bodies were forward
calculated using Comsol Multiphysics®, and Figure 7B and C
show that the predictions closely resemble the mapped long
wavelength anomaly (Fig. 7A). A full three-dimensional as-
sessment of the influence of magnetic susceptibility change
on the three-dimensional volumes was not performed but
based on the trend shown in two-dimensional (Fig. 6, gray
band), we can estimate a possible volume range between
1,400 and 3,000 km3 (Table 1). Remaining deviations, and
perhaps the greatest systematic error in our modeling, is the
neglection of postemplacement tilting of the Oquirrh Moun-
tains block by 10° to 25° eastward down (Melker and Geiss-
FIG. 6. Sensitivity analysis summarizing a range of magnetic model solu-
tions, plotted as a function of the assumed depth to the floor of the laccolith man, 1997) while our model assumes a horizontal present-day
(km below sea level). Black symbols and left vertical axis indicate solutions floor to the laccolith. This simplification might imply that the
for a representative two-dimensional section (Fig. 5), with the gray band in- laccolith has a somewhat conical shape tapering toward the
dicating the effect of varying the assumed magnetic susceptibility by ±0.010. west, but the modeling is too insensitive to laccolith floor
Red symbols refer to right-hand axis and relate to predicted volumes of two
full three-dimensional models, emphasizing that the volume of a near-cylin-
shape to warrant further refinement.
drical laccolith varies approximately in proportion with the cross-sectional
area of the two-dimensional sections.
Cooling of the Modeled Magma Body and Geochronology
Considering the petrographic similarity of subvolcanic
stocks along the deep geomagnetic anomaly, we interpret the
as a bold line our preferred two-dimensional model solution in modeled intrusive body beneath the Oquirrh Mountains fault
an N-S direction through the Bingham Canyon deposit (Fig. 5), block as representing a single former magma chamber. We
with an average laccolith thickness of 2.5 km and an intrusion cannot exclude the possibility of an incremental growth of the
roof around 1,000 m below sea-level. Since the laccolith has intrusion, but assuming a single rapidly filled magma chamber
approximately cylindrical shape in a WSW-ENE direction, we can estimate a maximum cooling time based on conduc-
the volume of the laccolith varies with the depth of the intru- tive heat loss. Advective heat loss would have shortened the
sion floor in a linear function proportional to cross-sectional time required to disperse the thermal anomaly considerably.

N
10 km

[nT]
A) Observed anomaly B) Calculated anomaly of the C) Calculated anomaly of the
laccolith with floor at -3000 m laccolith with floor at -5000 m
FIG. 7. Observed anomaly (A) and results of forward modeling of two possible three-dimensional bodies with floor depths
at −3,000 m (B) and −5,000 m (C). Signal upward continued to 3,300 m above sea level and reduced to pole. The black line
denotes the position of section A-A' shown in Figures 1B and 3 (results shown in Fig. 5); dotted lines indicate other sections
calculated for Figure 8 and additional models. A regional gradient of −15 nT over the entire north-south extent of image (A)
has been subtracted, to set both boundaries of the modeling domain to zero. Remaining differences between the modeled
and the observed field probably result from ignoring likely downfaulted parts of the laccolith in the basin west of the Oquirrh
Mountains (see Fig. 1) and underrepresentation of the volcanic cover southeast of the Bingham Canyon stock.

0361-0128/98/000/000-00 $6.00 613

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
614 STEINBERGER ET AL.

TABLE 1. List of Volumes and Masses Derived from Numerical and Chemical Calculations Outlined in the Text
(bold numbers are the values used in the discussion)

Pluton volume by geomagnetic modeling

Min. laccolith volume [km3] Min. fluid mass [Gt]

f(susceptibility), most shallow 0.020 1 Not realizable Not realizable


0.023 1 1,850 141
0.026 1 1,640* 125
0.032 1 1,392* 106

f(depth floor[m]) const. susc. at 0.023 –3,000 1,850 141


–4,000 2,165** 165
–5,000 2,480 189
–6,000 2,795** 213
–8,000 3,110** 238
Example laccolith 2 2,000 153

Extraction volumes by quartz mass balance between deposit and source

Min. melt volume [km3] Min. fluid mass [Gt]

Quartz mass as f(depth [m]


of veined space)3 Cu Mo sum Cu Mo sum
1,000 557 198 754 85 30 115
2,000 576 205 781 88 31 119
3,000 568 202 770 87 31 118
4,000 553 197 750 85 30 115
5,000 539 192 731 82 29 112

Extraction volumes by element solubility in ore-forming fluid

Min. melt volume [km3] Min. fluid mass [Gt]

Concentration of element
in fluid [ppm] Cu Mo Cu Mo
100 4 3,659 92 559 14
200 4 1,830 46 280 7
500 4 732 18 112 3
900 5 407 x 62 x
15,000 6 24 x 4 x

Extraction volumes by element solubility in melt

Min. melt volume [km3]

Concentration of element in melt [ppm] Cu Mo S


5 7, 8 5,215 249 89,552
35 9 745 36 12,793
47 10 555 26 9,527
104 8 251 12 4,305
126 11 207 10 3,554
600 9 43 2 746
1,800 8 14 1 249

*Extrapolated from two- to three-dimensions


**Extrapolated from three-dimensional trend
1 Reflecting about a standard deviation added and subtracted to the determined susceptibility of 0.023 ± 0.010
2 Assuming 50% nonhydrous minerals and saturated water concentration (5.7%); magnetic susceptibility of 0.023 and a floor depth of 3500 m are suffi-

cient to model this body


3 Based on mass of vein quartz in the respective stage; note that melt volume and fluid mass changes only slightly with depth, because increasing quartz

mass and quartz solubility level each other off


4 Mo concentration range measured in Bingham fluid inclusions, average at 200 ppm Mo (Seo et al., 2012), less than 40 ppm Mo is required if the

source volume is around ~230 km3 large


5 Experimental determined Cu concentration value in sulfo-haline fluid (Zajacz et al., 2011)
6 Maximum measured Cu concentration in Bingham fluid inclusions (Landtwing et al., 2010)
7 Mo concentration measured in equigranular monzonite (Moore, 1973)
8 Value chosen to estimate the minimum metal concentration in an evolved (and mixed ?) melt of 250 km3 providing 960 Mt S, 55.9 Mt Cu, and 1.4 Mt Mo
9 Value chosen to estimate the minimum metal concentration in a felsic melt of 750 km3 providing 960 Mt S, 55.9 Mt Cu, and 1.4 Mt Mo
10 Cu concentration measured in equigranular monzonite (Waite et al., 1997)
11 Cu concentration measured in melanephelinite (Waite et al., 1997)

0361-0128/98/000/000-00 $6.00 614

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 615

Such a scenario is not considered here due to the limitations above the present-day surface (Moore, 1973). Boundary heat
of the three-dimensional modeling software. The chamber flux was 0.065 W × m−2 at the bottom (at −50 km), resem-
may have extended even beyond the normal faults bounding bling average continental heat flux (Pollack et al., 1993), and
the Oquirrh Mountains and our geophysical modeling; since a constant temperature of 20°C was applied to the top bound-
we are not able to asses this magnitude, we do not include ary. The geometry was discretized with a mesh resolution be-
those parts in the model. Volume, initial magma temperature, tween 0.1 and 5 km, depending on the distance to the intru-
and depth of intrusion are the first-order parameters deter- sion (a total of 174,000 elements).
mining the rate of conductive cooling of such a magma cham- Interpreting the simulation results, the 680°C isotherm was
ber. Cooling in turn determines the maximum time interval taken as a proxy for the inward-migrating solidification front.
during which melt can be present in the magma chamber After about 30,000 years, the outer 50 vol % of the model lac-
and, therefore, the maximum duration of magmatic fluid pro- colith, including all apophyses, had cooled below the solidus
duction in the source of the ore-forming magmatic-hydro- temperature, and after 200,000 years only three subspherical
thermal system. Purely conductive cooling dictates the maxi- volumes of ~15 vol % of the initial 1,800 km3 remained at
mum lifetime of a given magma chamber, as internal magma temperatures above 680°C, implying partially molten areas.
convection or hydrothermal heat dissipation by external fluid Those residual volumes are located beneath Stockton, Bing-
convection would both shorten the time of magma solidifica- ham, and a magnetic anomaly a few kilometers east of Bing-
tion (Hayba and Ingebritsen, 1997; Ingebritsen et al., 2010). ham Canyon (Fig. 1B). The residual melt volume beneath
In order to estimate the maximum cooling time of the in- Bingham is the most voluminous of the three. After 230,000
trusion, we performed a numerical simulation using Comsol years from initial laccolith emplacement, the last magma be-
Multiphysics®. We approximated the modeled three-dimen- neath Bingham Canyon was solidified.
sional laccolith (Fig. 8; 3,000-m floor depth, about 1,850 km3) The cooling time obtained from thermal modeling is con-
as a hot solid body that cools solely by heat conduction into sistent with an extended lifetime of the magma chamber
the host rock. The laccolith was assigned a starting tempera- below the Bingham Canyon deposit, as recorded by the pop-
ture of 900°C, and the initial temperature distribution in the ulation of zircon ages in the subvolcanic intrusions (Parry et
host rocks followed a geothermal gradient of 37°C × km−1 al., 2001; von Quadt et al., 2011). Sorting all published high-
(Blackett, 2004). Both rock types had a temperature-depen- precision single-grain ages, irrespective of the geologic se-
dent thermal conductivity of k = 0.64 + 807 × (305 + T)−1, quence of emplacement, shows that the youngest grains of all
where k is in W × m−1 × K−1 and temperature (T) is in °C, mineralized porphyries converge at about 37.90 Ma (Fig. 9).
resembling the conductivity of acidic igneous rocks (Clauser Von Quadt et al. (2011) interpreted that these similar-aged
and Huenges, 1995). The heat capacity was 1,250 J × kg−1 × zircon grains determine the emplacement time of the por-
K−1 below 680°C, but doubled above this temperature to phyry dikes in a short succession, as the last events of near-
mimic the release of latent heat (Hayba and Ingebritsen, surface magmatic activity. The few published grain dates of
1997). An estimated eroded overburden of 2 km was added the equigranular monzonite overlap with older zircon ages

Stockton Settlement Canyon

Last Chance stock


W
2 km N

FIG. 8. Representative three-dimensional model, looking westward, of a laccolith constructed by linking a stack of two-
dimensional slices with 1- to 2-km spacing. Floor depth at 3,000 m below sea level, two-dimensional model section A-A'
through the western end of the Last Chance stock from Figure 5B is marked by the bold outline. Volcanic cover was con-
sidered in the modeling but is not displayed, and the postemplacement tilting is neglected. Existence and position of feeder
dikes are of low reliability, as a consequence of great depth and small volume.

0361-0128/98/000/000-00 $6.00 615

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
616 STEINBERGER ET AL.

30 max. residence time of Maximum available amount of water and


liquid melt in laccolith EM
residual melt volumes
Enumeration from young to old

QMP
25 LP The maximum amount of water available for ore formation
QLP
depends on the mass of melt and on the concentration of
20 Visual fit of
all points water dissolved in that melt but may additionally include free
Predicted
water suspended as bubbles in the magma. The following
15
EM trend considerations are based on an initial melt volume of 2,000
10
km3 approximating the laccolith shown in Figure 8, but re-
Likely
? Pb loss sults can be readily rescaled to the range of intrusion volumes
Re-Os age
5 ? of MoS2 constrained by our geomagnetic modeling.
37.90 ?
currently A large cooling intrusion is subject to internal crystal frac-
38.13 inexplicable
0
tionation from more mafic to more felsic compositions during
which the water fraction in the melt is increased. In line with
39.5 39.0 38.5 38.0 37.5 37.0
Zircon
206 238
Pb/ U ages [Ma], 2σ error
considering the equigranular monzonite as representative for
the average composition of the intrusion, it is reasonable to
FIG. 9. Published single-crystal zircon ages from the Bingham Canyon in- assume that the solidification of the Last Chance stock repre-
trusive suite, sorted by age and color coded according to samples from dif-
ferent intrusions (24 high-precision thermal ionization mass spectrometric
sents the stage when about 50 vol % of the laccolith had so-
206Pb/238U ages, excluding likely xenocrysts much older than 39 Ma (Parry et lidified (i.e., some 30,000 years after laccolith emplacement
al., 2001; von Quadt et al., 2011)). Age variation in each sample significantly according to our thermal modeling). Furthermore, this as-
exceeds the analytical precision of each single-grain age, indicating a mixed sumption of a 50% loss of volatiles accounts for a multiple of
population of antecrysts that formed in a common long-lived magma reser- processes: the likely incremental laccolith growth and volatile
voir at depth. Gray band is delimited by the interpreted age of intrusion of
quartz monzonite porphyry and latite porphyries that commonly have a clus-
loss to earlier solidification, the loss of volatiles through vol-
ter of youngest ages at 37.90 Ma (von Quadt et al., 2011) and the likely max- canic venting, and the incorporation of volatiles into hydrous
imum time of crystallization of the upper-crustal laccolith defined by mag- minerals. The presence of hornblende phenocrysts in
netic modeling. The older ages of zircons in the porphyries (QMP, LP, QLP) equigranular monzonite at Stockton, Settlement Canyon
and of the few reported zircons from the premineralization equigranular (Gilluly, 1932) and Bingham (Moore, 1973) points to a water
monzonite (EM) appear to have crystallized during the earlier crustal history
of the source magmas, even prior to the emplacement of the upper-crustal content >4 wt % in the 1,000 km3 of melt present in the
laccolith, which we tentatively infer at 38.13 Ma, based on thermal modeling magma chamber at this stage (Naney, 1983), and coexisting
(stippled line, schematically predicting age population of zircons in equigran- biotite indicates a magma temperature around 800°C (Fig.
ular monzonite if numerous additional grains were analyzed). The published 10). The fraction 4 wt % water is a minimum value, approxi-
preliminary Re-Os age of molybdenite (Chesley and Ruiz, 1997) is not in
plausible agreement with the thermal history and the 206Pb/238U ages.
mating the limit of magmatic fluid saturation at 100 MPa, but
at a higher pressure of 200 MPa, a monzonite melt can dis-
solve up to 5.7 wt % water. This water solubility was modeled
using the program MELTS by Ghiorso and Sack (1995) with
from the porphyries but are not younger than 38.4 Ma (Parry the monzonite composition of Waite et al. (1997) and an oxy-
et al., 2001). We can interpret these combined radiometric gen fugacity at Ni-NiO +2 (Maughan et al., 2002). An H2O
ages with the thermally modeled age of the magma chamber: content of 4 to 5.7% in a 1,000-km3 melt corresponds to po-
If the age of the porphyries represents the last melt of the lac- tential fluid mass of 107 to 153 Gt, of which only a negligible
colith and the laccolith solidified within 230,000 years, the fraction can be accommodated by further crystallization of
laccolith cannot have been emplaced before 38.13 Ma. Even hydrous minerals.
considering the analytical errors on the single-grain ages, we The mass of expelled residual fluid is an order-of-magni-
conclude that the older zircons in the porphyries, including tude estimate only, which can easily vary by a factor of three
the three grains from the equigranular monzonite measured in both directions, given the range of geophysical models for
by Parry et al. (2001), grew as antecrysts during the lower the laccolith and all other uncertainties. The available fluid
crustal evolution of the magmas prior to filling of the subvol- could be increased by additional free water present as bub-
canic laccolith, as found at other locations (Oberli et al., 2004; bles in the magma chamber, derived from the first half of the
Bachmann et al., 2007; Miller et al., 2007; Schaltegger et al., crystallization history of the laccolith, by addition of volatiles
2009). This interpretation implies that a statistically larger from greater depth, or by a larger extent of the magma cham-
number of grains from the equigranular monzonite could ber beyond the present-day Basin and Range faults. Alterna-
yield ages as young as about 38.13 Ma (dotted line, Fig. 9); tively, the available fluid quantity may have been reduced by
such measurements would provide a good test of our model- degassing through the volcanic edifice to the atmosphere
derived age of laccolith emplacement. (Cloos, 2001; Shinohara, 2008; Holland et al., 2011), although
such water loss after encapsulation was probably minor. The
Geochemical Constraints on Melt and Water Volumes occurrence of pumice in outcropping volcanics indicates
Given the range of volumes of the Oquirrh laccolith de- water saturation in the early extrusive phase of the Bingham
rived from geomagnetic data, we can estimate the amount of magmatic system, but outcropping stocks of equigranular
magmatic fluid that is potentially available. This fluid quantity monzonite show neither vesicles nor premineralization veins,
can then be compared with the amount of water required to consistent with our interpretation that these stocks probably
drive the chemical mass transfer as recorded by hydrothermal represent the last necks before the termination of extrusive
mineralization in the Bingham Canyon deposit. volcanism. We envision that their crystallization marks the

0361-0128/98/000/000-00 $6.00 616

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 617

T[°C] Granodiorite at and reaching down into a barren but pervasively K-altered
1100 200 MPa and still intensely veined core. This vein column is inter-

P l+
L (Naney, 1983) preted as a broad region of fluid ascent from a magma inter-

L
L+V
face near the roof of the laccolith (Gruen et al., 2010). As an
estimate to the total volume of vein quartz of magmatic-
1000
hydrothermal origin in this column, we extended the mapped
contours of quartz vein density (Gruen et al. 2010, fig. 9A) at
Op

the surface to depth, integrating the quartz volume in the re-


x+
Pl+

Pl+L+V sulting cylindrical shells. The height of the veined cylinder is


L

given by the depth at which magmatic fluid is released, which


900 Hb+Bt+Opx is taken to be the roof of the laccolith. We computed the
Opx+Pl+L+V
+Pl+L quartz mass for the range of roof depths obtained from our
Hb+Opx+Pl+L+V geomagnetic models. This yields between ~0.8 and ~1.5 Gt
quartz for column heights between 2,500 and 5,000 m, corre-
Hb+Bt+Opx+Pl+L+V sponding to an expulsion depth between 500 and 3,000 m
800 EM
below present sea level.
? Hb?+Bt+Pl+L+V To calculate the mass of water required to precipitate this
Bt+Pl mass of quartz we calculated the solubility of quartz at the ex-
+Ksp Bt+Pl+L+V pulsion depth. The solubility was fitted as a function of pres-
700 +Q+L
sure, temperature, salinity, and CO2 content ( Manning, 1994;
Akinfiev and Diamond, 2009). Source temperature is esti-
Bt+Pl+Ksp+Q+L+V Bt+Pl+Q+L+V
mated from the mineral assemblage found in the ore-related
?
QMP Bt+Pl+Ksp+Q+V quartz monzonite porphyry (690°C, Fig. 10). Salinity and
CO2 concentration were approximated from fluid inclusion
0 2 4 6 8 10 12 measurements (Landtwing et al., 2010; Seo et al., 2012) with
H2O [wt%] 5 wt % NaCl, 2 wt % KCl, and a CO2 concentration of 1 wt
%. Pressure is derived from depth assuming lithostatic pres-
FIG. 10. Experimental phase diagram of a synthetic granodiorite at 200
MPa as a function of temperature (T) and fraction of H2O present in the sure conditions and 2 km eroded overburden above the pre-
sample, after Naney (1983). Curves indicating water saturation and solidus sent mine (Moore, 1973).
are emphasized as bold blue lines. The observed mineral assemblages of the Our calculated quartz solubilities at the expulsion depth
equigranular monzonite (EM) and quartz monzonite porphyry (QMP) are range from 5.6 g SiO2 per kg fluid at 500 m to 11.7 g × kg−1
highlighted by colored areas. Abbreviations: Bt = biotite, Hb = hornblende,
Ksp = K-feldspar, L = liquid melt, Opx = orthopyroxene, Pl = plagioclase, Q=
at 3,000 m below present-day sea level. Thus, less water is re-
quartz, V = water vapor. quired to precipitate the same mass of quartz if the fluid is
derived from greater depths. The effects of increasing quartz
solubility with depth and increasing quantity of vein quartz
start of internal fractionation and water accumulation in the with greater column height roughly compensate each other,
magma chamber, leading up to the emplacement of the min- indicating that a fluid mass of 115 Gt is sufficient to precipi-
eralized porphyries. tate all vein quartz, irrespective of the expulsion depth (Fig.
11). Given a water solubility of 5.7 wt % water in the melt,
Minimum amount of water required to precipitate sourcing this water requires a melt volume of less than 750
ore-related vein quartz km3 (assuming a melt density of 2,700 kg × m−3). Out of
Vein quartz associated with the Cu-Au and Mo stages of ore these 750 km3, 550 km3 are required to form Cu-Au vein
formation was precipitated from fluids released by a quartz- quartz and 200 km3 are required to form the Mo-stage vein
saturated magma, as evidenced by quartz phenocrysts in the quartz. These melt volumes are smaller than the volume of
synmineralizing porphyries. Quartz solubility as a function of residual magma remaining after 50% crystallization of our
salinity, temperature, and pressure (Manning, 1994; Akinfiev geophysically modeled laccolith (1,000 km3). The fluid con-
and Diamond, 2009) dictates that essentially all silica that can straint from quartz veining is therefore compatible with our
initially be dissolved in a quartz-saturated magmatic fluid will interpretation that, after the equigranular monzonite stocks
precipitate across the temperature-pressure gradient be- were solidified, subsequent closed-system crystallization of
tween the magma and the surface. Therefore, the total mass the remaining 50% of monzonite magma concentrated the
of advected vein quartz provides an important constraint on contained water up to the point of porphyry emplacement
the magmatic fluid mass. and fluid release.
The amount of vein quartz within the mine volume and the In reality, the constant quantity of vein quartz with depth
underlying 1.6 km reached by exploration drilling can be esti- and the focusing of all the available magmatic water only
mated from mapped vein densities. Distinct sets of veins of through the igneous rock column are both coarse estimations.
the Cu-Au and the Mo stages of mineralization have been The fraction of vein quartz probably diminishes with greater
measured volumetrically near the recent mine surface. depth, as indicated by observations from Yerington (Dilles,
Drilling shows that high vein density extends downward along 1987; Dilles and Gans, 1995), where exposed dikes near the
a broadly elliptical column, centered on the quartz monzonite source region of the magmatic fluids do not show significant
porphyry dike, starting from the Cu-Au ore shell at the top quartz veining. On the other hand, a smaller part of the fluid

0361-0128/98/000/000-00 $6.00 617

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
618 STEINBERGER ET AL.

1000 153 fluid amounts are a small fraction of the total volume of the
laccolith, even after 50% crystallization. This corroborates our
assumption that mineralizing volatiles were accumulated by
800 122
internal crystallization of the pluton, or even parts of the plu-
Melt volume [km ]
3

Total Qz

Fluid mass [Gt]


ton only. Second, even with a low Cu concentration in the
600 92 fluid (900 ppm), the amount of fluid required for metal ad-
Cu-stage Qz vection to the deposit (~62 Gt H2O) is smaller than the
amount of fluid needed to precipitate the associated vein
400 61
quartz (~115 Gt). Similar estimates can be derived for Mo
and the associated Mo-stage vein quartz. This imbalance is
200 31 probably meaningful and not simply an artifact of over esti-
Mo-stage Qz
mating the volume of vein quartz in the barren core. Even
0 0 within the economic orebody (Cu-grade shell and mapped by
1 2 3 4 5 Gruen et al., 2010), the mapped amount of vein quartz ex-
Fluid origin depth below pit bottom [km] ceeds the copper ore grade by a factor of at least five by
FIG. 11. Melt volume (with 5.7 wt % water) required to provide the water
weight, whereas the concentration of SiO2 in a vapor-domi-
mass (right axis) needed for transporting and precipitating the estimated nated fluid (<0.1−0.5 g/kg) is significantly lower than its Cu
mass of hydrothermal vein quartz in and below the Bingham Canyon deposit. content (≥0.9 g/kg).
The depth of fluid origin (i.e., fluid pressure at the magma interface) affects
the solubility of quartz in the fluid as well as the total estimated quartz mass, Mass Balance for Cu, Mo, and S in the Laccolith
leading to a nearly constant requirement of ~750 km3 of melt and ~115 Gt of and the Deposit
quartz saturated fluids.
The initial concentrations of Cu, Mo, and S in the source
magma can provide additional and independent constraints
plume flowed through the sediments, where poorly visible on the source volume. No melt inclusion data are available to
veins in quartzites and silica in skarns were not included in assess the metal and sulfur contents of the silicate melts prior
the mapping by Gruen et al. (2010). These uncertainties to fluid exsolution. Bulk analyses of equigranular monzonite
amount to less than a factor of two and might partly cancel and unaltered volcanic extrusions provide an approximate
each other, so we can conclude that the quantity of magmatic lower bound on the available concentrations of ore-forming
fluid available from the modeled laccolith is sufficient to ex- elements in the magma chamber, because even these fresh
plain the observed massive addition of vein quartz to the rocks must have eventually saturated a local fluid phase upon
Bingham Canyon deposit. crystallization and thereby lost some of the ore-forming com-
ponents. Because of their low pressure of emplacement, we
Minimum amounts of magma required to source the can assume that the exsolving low-density vapor was rather in-
Cu and Mo sulfides efficient at removing metals (Dilles, 1987; Cline and Bodnar,
A similar mass-balance approach as used above for quartz 1991; Williams-Jones and Heinrich, 2005) but volatile sulfur
can be used for the ore metals, but it is subject to greater un- was certainly lost to a significant degree.
certainties regarding their solubility in the ore fluid. Con- Published Cu concentrations in equigranular monzonite
straints on the possible range of concentrations of Cu in the range from 43 ppm (Waite et al., 1997) to older measure-
incoming ore-forming fluid can be obtained from experi- ments averaging at 70 ppm (Moore, 1973). The most mafic
ments with synthetic fluid inclusions, from which a Cu solu- extrusive rock (a melanephelinite) has a Cu concentration of
bility of 900 ppm in H2O-NaCl-HCl-S fluids at 1,000°C, 150 126 ppm (Maughan et al., 2002). Considering the two ex-
MPa, and an oxygen fugacity of Ni-NiO+2 can be estimated tremes of initial Cu concentrations, assuming incompatible
(Zajacz et al., 2011). This is significantly lower than Cu con- behavior during fractional crystallization, and an eventual ex-
centrations of up to 1.5 wt % measured in deep intermediate- traction efficiency of 0.8 (following the model calculations of
density fluid inclusions trapped in the barren core below the Cline and Bodnar, 1991), indicates that between 200 and 600
orebody, which were interpreted to represent the magmatic km3 of magma are required to source the Cu. Using a Mo
input fluid at Bingham Canyon (Landtwing et al., 2010) but concentration of 5 ppm in the melt (Moore, 1973) and an ex-
are likely modified after entrapment (Lerchbaumer and Au- traction efficiency of 0.42 (Audétat, 2010), about 250 km3 of
detat, 2012). Mo concentrations in the same inclusions mostly melt are required to source the Mo found in the deposit.
range from 100 to 500 ppm, averaging around 200 ppm (Seo These volumes correspond to about 10 to 25 vol % of the geo-
et al., 2012). To precipitate 55.9 Mt of total Cu in the deposit physically modeled laccolith. This calculation therefore shows
from a fluid containing 900 ppm Cu, as estimated from ex- that no special preenrichment or additional input of ore met-
perimental data, would require 62 Gt of water derived from als is required to produce the known deposit.
at least 410 km3 of silicate melt with 5.7 wt % dissolved water. The sulfur isotope signal from both the central porphyry
An average analyzed Mo concentration of 200 ppm in the zone (Field, 1966) and the peripheral Pb-Zn veins (Field and
input fluid would require about 7 Gt water or 46 km3 melt to Moore, 1971) clearly points to a magmatic sulfur source, and
precipitate 1.4 Mt Mo (Table 1). a major sulfur contribution from the surrounding sediments
Although the absolute quantities derived above are uncertain, can be excluded. Sulfur solubility in pristine melts is difficult
the different magnitudes indicate two important conclusions. to model and to measure. Available bulk rock analyses of un-
First, the melt volumes required to generate the necessary altered equigranular monzonite have a sulfur concentration

0361-0128/98/000/000-00 $6.00 618

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 619

of 113 ppm (Waite et al., 1997) but are probably degassed to Mo concentration in the measured ore-forming fluid is high
an unknown extent. Modeling of S solubility is dependent of enough to be derived from significantly smaller volumes.
melt composition, oxygen fugacity, temperature, pressure,
and degree of degassing (Baker and Moretti, 2011); factors Discussion
that are not constrained by this study. To provide 960 Mt of S As stated previously, no single method can uniquely deter-
in the deposit and its pyrite halo (Hattori and Keith, 2001) mine the dimensions of the magmatic source of fluids and
from the same source volume as calculated for quartz (750 ore-forming components for a porphyry-type ore system.
km3), this volume would require S in a concentration of at However, the combination of multiple constraints including
least 600 ppm (Fig 12). This concentration is permissive by geophysical modeling can build a framework that allows a
the model of Baker and Moretti (2011) in a rhyolite melt if semiquantitative reconstruction of processes and minimum
oxygen fugacity is between the sulfide-sulfate buffer (~Ni- ingredients required to form the world-class deposit at Bing-
NiO+1.5) and Ni-NiO+2, the measured oxygen fugacity in ham Canyon. Field exposure in the uplifted Wasatch Moun-
the volcanic rocks at Bingham (Maughan et al., 2002). Hence, tains illustrates that the deposit is located on a magmatic
the sulfur abundance in the deposit could be explained by trend that produced plutonic bodies reaching the size of the
fractionation of the monzonite melt alone. However, ob- geomagnetic anomaly crossing the Oquirrh Mountains (Fig.
served magma mingling between mafic and felsic melts in the 1). The outcropping subvolcanic intrusions in the Oquirrh
genetically related Wasatch Mountains intrusions (Pettke et Mountains are of very similar mineralogy (Fig. 2), indicating
al., 2010), the presence of CO2 in the ore fluid (Landtwing et that the intrusive body beneath the Oquirrh Mountains is
al., 2010), the presence of comagmatic minette dikes in Bing- continuous from Stockton to Bingham Canyon (Figs. 1, 3).
ham Canyon (Keith et al., 1997), and the increasing degree of Geologically permissible solutions to geomagnetic modeling
mafic components in the Bingham porphyry dikes (Fig. 2E) indicate a flat laccolith-shaped magma chamber. A conductive
are compelling evidences for a mafic contribution to the cooling model of the shallowest possible three-dimensional
source magma. Therefore, a mafic contribution could not model pluton takes 230,000 years until complete solidifica-
only provide ample S to the deposit (Hattori and Keith, 2001) tion. This timespan would be prolonged by deeper emplace-
but could also provide a trigger to the onset of mineralization. ment but is more likely shortened by internal convection and
The minimum volumes of melt required to provide the ore- external advective cooling. Sensitivity analysis shows that pos-
forming components in the deposit are consistent with each sible solutions of our geomagnetic forward modeling are un-
other: starting with a volume of ~750 km3 melt, H2O and S certain within a factor of two to three (Table 1), but even the
concentrated as incompatible components until the melt vol- smallest plausible pluton exceeds the magma volume needed
ume was reduced to about 250 km3. The total metal resource to supply all of the geochemical components enriching the
can be explained if this remaining melt concentrated ele- deposit (Figs. 6, 11, 12). Probably more significant than the
ments to 100 ppm Cu, 5 ppm Mo, and 1,800 ppm S (maybe, absolute numbers are the relative sizes of estimated source
but not necessarily, parts of the inventory were contributed requirements. These allow a process interpretation (Fig. 13)
from a mafic incursion; Fig. 12). Furthermore, the Cu and in five stages, as illustrated with a set of consistent parameters
calculated for a laccolith of 2,000 km3, emplaced with a roof
depth at ~2,000 m below pit bottom.
1000
Stage A—emplacement of the laccolith and
encapsulated melt volume main eruptive activity
Melt volume required [km3]

800
Magma ascended from a lower-crustal reservoir, as indicated
to by the large age spectrum of zircon 206Pb/238U ages (Parry et al.,
18
2001; von Quadt et al., 2011). These zircons are interpreted
to 5 ppm

600 S 00
to

pp
10

m as antecrysts that formed in part even before the magma filled


0
pp

a large upper-crustal laccolith with a floor depth at 3 to 5 km


m

400
Cu
below present-day sea level (equivalent to approximately 7–9
km below the paleosurface) and a thickness between ~2 and
200 ~3.5 km (Fig. 13A). The volume of this laccolith was probably
Mo
melt volume at fluid release about 2,000 km3 but may have measured between 1,500 and
3,500 km3. Additional magma exceeding this accommodation
0 space was erupted through volcanic necks that are later occu-
0 20 40 60 80 500 1000 1500
Concentration of element in melt [ppm] pied by equigranular monzonite. Pumice deposits northeast of
the mine indicate subplinian eruptions, which would have re-
FIG. 12. Melt volumes required to provide 55.9 Mt Cu, 1.4 Mt Mo
(Krahulec, 2010), and 960 Mt S (Hattori and Keith, 2001) as a function of the quired water saturation of the erupting magma within the vol-
concentration of the respective element in the melt, assuming that 20% Cu canic conduits. The volcano or dome edifice reached a height
(Cline and Bodnar, 1991), 40% Mo (Audétat, 2010), and 23 ppm S (Hattori of about 4,300 m above present-day sea level, based on geo-
and Keith, 2001) are retained in the magmatic rocks. If a volume of 750 km3 logic reconstruction (Moore, 1973; Deino and Keith, 1997).
(upper horizontal line) is encapsulated, no extraordinary concentrations
(drop lines) of Cu, Mo, and S are required to concentrate up (arrows) to 100, Stage B—cessation of eruptive activity
5, and 1,800 ppm after closed-system fractionation of those incompatible
components reduced the volume to 250 km3 (lower horizontal line). A likely Transport of magma through the laccolith probably ceased
contribution from mafic melt would increase the concentration even further. when the supply from the deep source stopped. Subsequently,

0361-0128/98/000/000-00 $6.00 619

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
620 STEINBERGER ET AL.

FIG. 13. Schematic but realistically scaled interpretation of evolution stages of the mineralizing magma system at Bing-
ham Canyon. Geometric dimensions are based on a representative two-dimensional model section (Fig. 5), but indicated
quantities include volumetric constraints based on a full three-dimensional model consistent with geomagnetic data (Fig. 6).
Times are anchored to the interpreted time of emplacement of the equigranular monzonite (Parry et al., 2001) and the min-
eralized porphyries at 37.90 Ma (D), based on high-precision zircon geochronology (Fig. 9; Quadt et al., 2011), and back-cal-
culated for the maximum duration of laccolith crystallization assuming pure conductive cooling (A-C). Mass-balance consid-
erations for H2O, SiO2, Cu, and S indicate a possible decoupling of porphyry emplacement and quartz veining (D) from later
ore metal introduction (E). The onset of fluid expulsion may be triggered by underplating of an S- and CO2-rich mafic melt,
while mineralization may not take have taken place after both magmas sufficiently mixed.

0361-0128/98/000/000-00 $6.00 620

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 621

the crystallinity of the magma increased during cooling. The and into adjacent sedimentary rocks. Upon depressurization
magma may have continued to convect and diffusively vent and cooling, the water precipitated vein quartz, most of it bar-
volatiles, until it reached an average crystal fraction of about ren and predating sulfide deposition.
50 vol % (Petford, 2003; Shinohara, 2008). Cessation of
magma convection led to solidification of volcanic necks and Stage E—mineralization
subvolcanic apophyses that are now partly exposed as Petrographic observations in Cu-Au-stage veins showed
equigranular monzonite stocks. As a result, a residual melt that Cu sulfides and Cu-rich fluid inclusions were introduced
volume of around 1,000 km3 was encapsulated by a relatively at a late stage, together with a volumetrically minor, second
impermeable carapace (Fig. 13B). Petrological evidence from generation of quartz (Redmond et al., 2004; Landtwing et al.,
the monzonite stocks indicates that the magma in the laccol- 2010). The measured ratio between the first, barren genera-
ith was at (or close to) water saturation at this stage. A con- tion of quartz and the second, syn- to postmineralization
ductive cooling model suggests that the stage of carapace so- quartz is >10. Assuming that the mass of precipitated vein
lidification was reached no later than 30,000 years after quartz per mass unit of magmatic fluid stays constant, out of
laccolith emplacement. the ~150 Gt water available from the magma chamber, some
75 Gt where sufficient to precipitate the first generation of
Stage C—quiescent cooling Cu-Au-stage quartz and ~9 Gt formed the second generation
Over the following ~200,000 years, the laccolith is inferred of Cu-Au-stage quartz. Later, ~30 Gt water with ~300 ppm
to have cooled by slow internal crystallization and nearly Mo were required to form Mo-stage quartz and the molyb-
closed-system fractionation. If 50% of remaining melt denite. Summing up the masses of water that were required
(~1,000 km3) contained 5.7 wt % water (the maximum solu- to form all vein quartz yields 115 Gt, which is within the avail-
bility at the average pluton pressure of 200 MPa), this encap- able mass of 150 Gt. In detail, Cu-Au mineralization took
sulated melt held a total mass of ~150 Gt H2O. During fur- place during and after a phase of quartz redissolution that
ther cooling and crystallization, second boiling produced a separates the first and the second generation of Cu-Au-stage
free fluid phase, which probably coalesced and ascended to a quartz (Redmond et al., 2004; Landtwing et al., 2010). One
cupola near the top of the residual magma volume (Cloos, might speculate that this break relates to a change in magma
2001; Fig. 13C). Simultaneously, the silicate magma fraction- composition caused by the mafic injection in turn leading to
ated progressively toward the composition of quartz mon- changes in fluid compositions.
zonite porphyry, reaching quartz saturation around 710°C
(Fig. 10). The residual melt volume some 215,000 years after Conclusions
laccolith emplacement was around 250 km3. If none of the This study shows that independent geological, geophysical,
150 Gt of water had escaped during inward crystallization of and geochemical observations can be linked by a consistent
the pluton, this water would correspond to approximately interpretation of the dimension and temporal evolution of a
20% of the residual melt by mass. Before reaching this large porphyry-mineralizing magmatic fluid and metal source.
proportion of free fluid, a mechanically unstable state would Bingham Canyon is a favorable case showing a distinct aero-
be reached, ready to break through the cupola in the roof of magnetic anomaly, which can be modeled by a major laccol-
the residual magma chamber. ith-shaped pluton underlying the deposit, thanks to the low
magnetic susceptibility of the surrounding sediments. The
Stage D—mafic injection, porphyry emplacement, modeled size of this intrusion is easily sufficient but not much
and barren quartz veining greater than the minimum amount of magma needed to sup-
The latest stages of the felsic magma evolution may have ply all ore-forming components, including the fluid required
been triggered by injection of a mafic magma, perhaps via un- to transport these to the deposit. With the possible exception
derplating at the floor of the laccolith. In such a scenario, free of sulfur, the bulk composition of this source pluton can be a
CO2 released from the mafic magma (Fischer and Marty, typical calc-alkaline magma. No special preenrichment of
2005) could have percolated through the felsic magma metals or volatiles is required, although this would enhance
(Huber et al., 2009). The presence of CO2 lowered the solu- the metal endowment or fluid-extraction efficiency and may
bility of H2O in the melt (Papale, 1999; Newman and Lowen- have affected the Au- and the Mo-rich nature of the Bingham
stern, 2002), producing additional fluid from the felsic melt. Canyon deposit.
The injection of mafic melt, as well as the presence of free Currently mineable porphyry copper deposits vary in metal
CO2 and the resulting rapidly increasing fraction of free water endowment over at least two orders of magnitude, between
all favored the collapse of the already unstable cupola by ex- giants like El Teniente (110 Mt Cu) and smaller but still eco-
tension and uplift of the overlying host rocks. A mixture of nomic deposits containing less than a million metric tons of
water, melt and crystals was ejected from the magma cham- metal (Singer et al., 2005). Even with an uncertainty factor of
ber, forming the quartz monzonite porphyry including its two to three in the scaling of a potentially mineralizing magma
high-temperature barren quartz veins. The escape of the flu- chamber, geophysical modeling of an aeromagnetic anomaly
ids promoted crystallization by lowering the fraction of can provide a meaningful measure of the potential size of an
volatiles in the melt, which in turn resulted in further fluid re- undiscovered or still incompletely explored ore deposit.
lease. Fluids escaped by near-explosive hydrofracturing and Even where the geophysical response is masked by mag-
assisted by the preceding bulging of the cupola through a netic host rocks or shallow cover by magnetite-rich volcanics,
broad column (Gruen et al., 2010) that extended beyond the a major underlying magma chamber is an essential require-
porphyry into previously emplaced equigranular monzonite ment to source the metals, the sulfur, and the mineralizing

0361-0128/98/000/000-00 $6.00 621

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
622 STEINBERGER ET AL.

fluid. A major magma chamber is the decisive difference of Audétat, A., 2010, Source and evolution of molybdenum in the porphyry
porphyry-mineralizing systems, in contrast to many other Mo(-Nb) deposit at Cave Peak, Texas: Journal of Petrology, v. 51, p.
1739−1760.
long-lived volcanic systems in magmatic arcs where magma is Bachmann, O., Charlier, B.L.A., and Lowenstern, J.B., 2007, Zircon crystal-
expelled as fast as it is generated (e.g., Tepley et al.; 2000; lization and recycling in the magma chamber of the rhyolitic Kos Plateau
Dungan et al.; 2001). Large calc-alkaline volcanoes may be tuff (Aegean arc): Geology, v. 35, p. 73−76.
larger regarding total magma volume but do not necessarily Baker, D.R., and Moretti, R., 2011, Modeling the solubility of sulfur in mag-
mas: A 50-year old geochemical challenge: Reviews in Mineralogy and
develop a single magma chamber of significant size that per- Geochemistry, v. 73, p. 167−213.
sists over an extended period and allows the magma to evolve Ballantyne, G.H., Smith, T.W., and Redmond, P.B., 1997, Distribution and
and mingle with magmas from other sources (Reubi and mineralogy of gold and silver in the Bingham Canyon porphyry copper de-
Blundy, 2009). The combined geophysical and geochronolog- posit, Utah: Society of Economic Geologists Guidebook Series 29, p.
ical data from the Bingham Canyon system imply that the 147−153.
Bankey, V., Grauch, T., and Kucks, R.P., 1998, Utah aeromagnetic and grav-
hallmark of a successful porphyry-mineralizing magma sys- ity maps and data: U.S. Geological Survey Open-File Report 98-761,
tem is probably not a long duration between successive por- http://pubs.er.usgs.gov/publication/ofr98761.
phyry pulses, but the presence of a broad age spectrum of zir- Bartley, J.M., Coleman, D.S., and Glazner, A.F., 2006, Incremental pluton
con antecrysts indicating a long-lived magma chamber below emplacement by magmatic crack-seal: Transactions of the Royal Society of
the shallow porphyry stock (cf. Sillitoe and Mortensen, 2010). Edinburgh-Earth Sciences, v. 97, p. 383−396.
Biek, R.F., Solomon, B., Smith, T., and Keith, J.D., 2007, Geologic map of
The metal grade of many porphyry-style ores is approxi- the Copperton Quadrangle, Salt Lake county, Utah: Salt Lake City, Utah
mately related to the density of quartz veining (Sillitoe, 2010), Geological Survey.
yet petrographic evidence from several deposits indicates that Blackett, R.E., 2004, Geothermal gradient data for Utah: Utah Geological
most of the vein quartz was deposited before Cu-Fe sulfide Survey Open-File Report 431, CD.
Burnham, C.W., and Ohmoto, H., 1980, Late-stage processes of felsic mag-
precipitation (Redmond et al., 2004; Klemm et al., 2007; matism: Mining Geology Special Issue 8, p. 1−11.
Rusk et al., 2011). Our comparison of mineral solubilities and Candela, P.A., 1997, A review of shallow, ore-related granites: Textures,
the total mass of accumulated minerals at Bingham Canyon volatiles, and ore metals: Journal of Petrology, v. 38, p. 1619−1633.
indicates that copper introduction required only a small frac- Cannell, J., Cooke, D.R., Walshe, J.L., and Stein, H., 2005, Geology, miner-
tion of the fluid mass needed for the addition of the total vein alization, alteration, and structural evolution of the El Teniente porphyry
Cu-Mo deposit: ECONOMIC GEOLOGY, v. 100, p. 979−1003.
quartz. These observations may indicate partial decoupling of Cathles, L.M., 1981, Fluid flow and genesis of hydrothermal ore deposits:
the fluid sources driving silica and copper sulfide advection to ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 442−457.
the deposit. Our mass-balance considerations together with Cathles, L.M., and Shannon, R., 2007, How potassium silicate alteration sug-
petrological observations (Hattori and Keith, 2001) suggest gests the formation of porphyry ore deposits begins with the nearly explo-
sive but barren expulsion of large volumes of magmatic water: Earth and
that most of the vein quartz was derived from the more felsic Planetary Science Letters, v. 262, p. 92−108.
porphyry magma. Quartz was followed by Cu ± Au addition, Cathles, L.M., Erendi, A.H.J., and Barrie, T., 1997, How long can a hydro-
possibly by an H2S ± CO2-rich fluid from a more mafic thermal system be sustained by a single intrusive event?: ECONOMIC GE-
magma pulse that underplated the residual felsic magma OLOGY, v. 92, p. 766−771.
chamber and may have triggered the porphyry emplacement. Chesley, J.T., Ruiz, J., 1997, Preliminary Re-Os dating on molybdenite min-
eralization from the Bingham Canyon porphyry copper deposit, Utah: So-
Acknowledgments ciety of Economic Geologists Guidebook Series 29, p. 165−169.
Clauser, C., and Huenges, E., 1995, Thermal conductivity of rocks and min-
The support by geologists and staff of Kennecott Utah Cop- erals, in Ahrens, T., ed., Rock physics and phase relations: A handbook of
per and the Rio Tinto Exploration group in Salt Lake City has physical constants, 3: American Geophysical Union, p. 105−126.
always been outstandingly proficient and cordial. Stimulating Cline, J.S., and Bodnar, R.J., 1991, Can economic porphyry copper mineral-
ization be generated by a typical calc-alkaline melt?: Journal of Geophysi-
reviews of earlier manuscript versions by Kim Schroeder cal Research, v. 96, p. 8113−8126.
(KUCC), Geoff Ballantyne (Rio Tinto), Ken Krahulec (Utah Cloos, M., 2001, Bubbling magma chambers, cupolas, and porphyry copper
Geological Survey), and John Dilles (University of Oregon) deposits: International Geology Review, v. 43, p. 285−311.
clarified several potentially misleading parts. Dave John Deino, A., and Keith, J.D., 1997, Ages of volcanic and intrusive rocks in the
Bingham mining district, Utah: Society of Economic Geologists Guidebook
(USGS) provided coordinates of locations in the Wasatch Series 29, p. 91−100.
Mountains, illustrating potential deeper feeders of eroded Dilles, J.H., 1987, Petrology of the Yerington batholith, Nevada—evidence
porphyry deposits. Andy Jackson, Alan Green, and Hansruedi for evolution of porphyry copper ore fluids: ECONOMIC GEOLOGY, v. 82, p.
Maurer (ETH) helped with advice regarding the handling of 1750−1789.
aeromagnetic data, and all members of the Fluids and Min- Dilles, J.H., and Gans, P.B., 1995, The chronology of Cenozoic volcanism and
deformation in the Yerington area, Western Basin and Range and Walker-
eral Deposits group at ETH are thanked for their numerous Lane: Geological Society of America Bulletin, v. 107, p. 474−486.
constructive comments on the manuscript. Reviews by Rick Dungan, M.A., Wulff, A., and Thompson, R., 2001, Eruptive stratigraphy of
Valenta, Dave Cooke, and an anonymous referee improved the Tatara-San Pedro Complex, 36°S, Southern Volcanic Zone, Chilean
this article substantially. Without the ongoing financial sup- Andes: Reconstruction method and implications for magma evolution at
port of ETH and the Swiss National Science Foundation long-lived arc volcanic centers: Journal of Petrology, v. 42, p. 555−626.
Emmons, W.H., 1927, Relations of the disseminated copper ores in porphyry
(Projects 200020-16693, -124906, -135302) this study would to igneous intrusives: Transactions of the American Institute of Mining and
not have been possible. Metallurgical Engineers, v. 75, p. 797−809.
Erickson, A.J., Jr., 1976, The Uinta-Gold Hill trend: An economically impor-
REFERENCES tant lineament: Utah Geological Association, International Conference on
Akinfiev, N.N., and Diamond, L.W., 2009, A simple predictive model of the New Basement Tectonics, 1st, Salt Lake City, Utah, Proceedings, p.
quartz solubility in water-salt-CO2 systems at temperatures up to 1000°C 126−138.
and pressures up to 1000 MPa: Geochimica et Cosmochimica Acta, v. 73, Field, C.W., 1966, Sulfur isotope abundance data, Bingham district, Utah:
p. 1597−1608. ECONOMIC GEOLOGY, v. 61, p. 850−871.

0361-0128/98/000/000-00 $6.00 622

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
SOURCE PLUTONS DRIVING PORPHYRY Cu ORE FORMATION: BINGHAM CANYON DEPOSIT 623

Field, C.W., and Moore, W.J., 1971, Sulfur isotope study of the “B” limestone III. Zoned copper-gold ore deposition by magmatic vapor expansion: ECO-
and Galena fissure ore deposits of the U.S. mine, Bingham mining district, NOMIC GEOLOGY, v. 105, p. 91−118.
Utah: ECONOMIC GEOLOGY, v. 66, p. 48−62. Lanier, G., John, E.C., Swensen, A.J., Reid, J., Bard, C.E., Caddey, S.W., and
Fischer, T.P., and Marty, B., 2005, Volatile abundances in the sub-arc mantle: Wilson, J.C., 1978, General geology of Bingham mine, Bingham Canyon,
Insights from volcanic and hydrothermal gas discharges: Journal of Vol- Utah: ECONOMIC GEOLOGY, v. 73, p. 1228−1241.
canology and Geothermal Research, v. 140, p. 205−216. Lerchbaumer, L., and Audetat, A., 2012, High Cu concentrations in vapor-
Ghiorso, M.S., and Sack, R.O., 1995, Chemical mass-transfer in magmatic type fluid inclusions: An artifact?: Geochimica et Cosmochimica Acta, v. 88,
processes.4. A revised and internally consistent thermodynamic model for p. 255−274.
the interpolation and extrapolation of liquid-solid equilibria in magmatic Maksaev, V., F., Munizaga, M., McWilliams, M., Fanning, R., Mathur, J.,
systems at elevated-temperatures and pressures: Contributions to Mineral- Ruiz, and Zentilli, M., 2004, New chronology for El Teniente, Chilean
ogy and Petrology, v. 119, p. 197−212. Andes, from U-Pb, 40Ar/39Ar, Re-Os and fission-track dating: Implications
Gilluly, J., 1932, Geology and ore deposits of the Stockton and Fairfield for the evolution of a supergiant porphyry Cu-Mo deposit: Society of Eco-
quadrangles, Utah: Reston, VA, U.S. Geological Survey, 171 p. nomic Geologists Special Publication 11, p. 15−54
Gruen, G., Heinrich, C.A., and Schroeder, K., 2010, The Bingham Canyon Manning, C.E., 1994, The solubility of quartz in H2O in the lower crust and
porphyry Cu-Mo-Au deposit. II. Vein geometry and ore shell formation by upper mantle: Geochimica et Cosmochimica Acta, v. 58, p. 4831−4839.
pressure-driven rock extension: ECONOMIC GEOLOGY, v. 105, p. 69−90. Maughan, D.T., Keith, J.D., Christiansen, E.H., Pulsipher, T., Hattori, K.,
Gustafson, L.B., and Hunt, J.P., 1975, The porphyry copper deposit at El-Sal- and Evans, N.J., 2002, Contributions from mafic alkaline magmas to the
vador, Chile: ECONOMIC GEOLOGY, v. 70, p. 857−912 Bingham porphyry Cu-Au-Mo deposit, Utah, USA: Mineralium Deposita,
Gustafson, L.B., and Jorge Quiroga, G., 1995, Patterns of mineralization and v. 37, p. 14−37.
alteration below the porphyry copper orebody at El Salvador, Chile: ECO- McInnes, I.A., Evans, H.T., Fu, F.Q., Garwin, S., Belousova, E., Griffin,
NOMIC GEOLOGY, v. 90, p. 2−16. W.L., Bertens, A., Sukarna, D., Permanadewi, S., Andrew, R.L., and
Halter, W.E., Bain, N., Becker, K., Heinrich, C.A., Landtwing, M., Von- Deckart, K., 2005, Thermal history analysis of selected Chilean, Indonesian
Quadt, A., Clark, A.H., Sasso, A.M., Bissig, T., and Tosdal, R.M., 2004, and Iranian porphyry Cu-Au-Mo deposits, in Porter, T.M., ed., Super por-
From andesitic volcanism to the formation of a porphyry Cu-Au mineraliz- phyry copper and gold deposits: A global perspective, 1: Adelaide, PGC
ing magma chamber: the Farallon Negro Volcanic Complex, northwestern Publishing, p. 27−42.
Argentina: Journal of Volcanology and Geothermal Research, v. 136, p. Melker, M., and Geissman, J.W., 1997, Paleomagnetism of the volcanic and
1−30. Intrusive rocks associated with the Bingham Canyon porphyry Cu-Au-Ma
Hattori, K.H., and Keith, J.D., 2001, Contribution of mafic melt to porphyry deposit, Utah: Society of Economic Geologists Guidebook Series 29, p.
copper mineralization: Evidence from Mount Pinatubo, Philippines, and 101−112.
Bingham Canyon, Utah, USA: Mineralium Deposita, v. 36, p. 799−806. Michel, J., Baumgartner, L., Putlitz, B., Schaltegger, U., and Ovtcharova, M.,
Hayba, D.O., and Ingebritsen, S.E., 1997, Multiphase groundwater flow near 2008, Incremental growth of the Patagonian Torres del Paine laccolith over
cooling plutons: Journal of Geophysical Research, v. 102, p. 12,235−12,252. 90 k.y: Geology, v. 36, p. 459−462.
Hintze, L.F., 1980, Geologic map of Utah: Salt Lake City, Utah Geological Miller, J.S., Matzel, J.E.P., Miller, C.F., Burgess, S.D., and Miller, R.B., 2007,
and Mineral Survey. Zircon growth and recycling during the assembly of large, composite arc plu-
Holland, A.S.P., Watson, I.M., Phillips, J.C., Caricchi, L., and Dalton, M.P., tons: Journal of Volcanology and Geothermal Research, v. 167, p. 282−299.
2011, Degassing processes during lava dome growth: Insights from Santia- Moore, W.J., 1973, Igneous rocks in Bingham mining district, Utah: U.S. Ge-
guito lava dome, Guatemala: Journal of Volcanology and Geothermal Re- ological Survey Professional Paper 629-B, 41 p.
search, v. 202, p. 153−166. Naney, M.T., 1983, Phase-equilibria of rock-forming ferromagnesian silicates
Huber, C., Bachmann, O., and Manga, M., 2009, Homogenization processes in granitic systems: American Journal of Science, v. 283, p. 993−1033.
in silicic magma chambers by stirring and mushification (latent heat buffer- Newman, S., and Lowenstern, J.B., 2002, VolatileCalc: A silicate melt-H2O-
ing): Earth and Planetary Science Letters, v. 283, p. 38−47. CO2 solution model written in Visual Basic for excel: Computers and Geo-
Ingebritsen, S.E., Geiger, S., Hurwitz, S., and Driesner, T., 2010, Numerical sciences, v. 28, p. 597−604.
simulation of magmatic hydrothermal systems: Reviews of Geophysics, v. Oberli, F., Meier, M., Berger, A., Rosenberg, C.L., and Giere, R., 2004, U-
48, 33 p. Th-Pb and 230Th/238U disequilibrium isotope systematics: Precise accessory
John, D.A., 1989, Geologic setting, depths of emplacement, and regional dis- mineral chronology and melt evolution tracing in the Alpine Bergell intru-
tribution of fluid inclusions in intrusions of the central Wasatch Mountains, sion: Geochimica et Cosmochimica Acta, v. 68, p. 2543−2560.
Utah: ECONOMIC GEOLOGY, v. 84, p. 386−409. Papale, P., 1999, Modeling of the solubility of a two-component H2O+CO2
——1997, Geologic setting and characteristics of mineral deposits in the cen- fluid in silicate liquids: American Mineralogist, v. 84, p. 477−492.
tral Wasatch Mountains, Utah: Society of Economic Geologists Guidebook Parry, W.T., Wilson, P.N., Moser, D., and Heizler, M.T., 2001, U-Pb dating of
Series 29, p. 11−33. zircon and 40Ar/39Ar dating of biotite at Bingham, Utah: ECONOMIC GEOL-
Keith, J.D., Whitney, J.A., Hattori, K., Ballantyne, G.H., Christiansen, E.H., OGY, v. 96, p. 1671−1683.
Barr, D.L., Cannan, T.M., and Hook, C.J., 1997, The role of magmatic sul- Petford, N., 2003, Rheology of granitic magmas during ascent and emplace-
fides and mafic alkaline magmas in the Bingham and Tintic mining dis- ment: Annual Review of Earth and Planetary Sciences, v. 31, p. 399−427.
tricts, Utah: Journal of Petrology, v. 38, p. 1679−1690. Pettke, T., Oberli, F., and Heinrich, C.A., 2010, The magma and metal source
Klemm, L.M., Pettke, T., Heinrich, C.A., and Campos, E., 2007, Hydrother- of giant porphyry-type ore deposits, based on lead isotope microanalysis of
mal evolution of the El Teniente deposit, Chile: Porphyry Cu-Mo ore de- individual fluid inclusions: Earth and Planetary Science Letters, v. 296, p.
position from low-salinity magmatic fluids: ECONOMIC GEOLOGY, v. 102, p. 267−277.
1021−1045. Pollack, H.N., Hurter, S.J., and Johnson, J.R., 1993, Heat flow from the
Kloppenburg, A., Grocott, J., and Hutchinson, D., 2010, Structural setting Earth’s interior: Analysis of the global data set: Reviews of Geophysics, v.
and synplutonic fault kinematics of a Cordilleran Cu-Au-Mo porphyry min- 31, p. 267−280.
eralization system, Bingham mining District, Utah: ECONOMIC GEOLOGY, v. Proffett, J.M., 2009, High Cu grades in porphyry Cu deposits and their rela-
105, p. 743−761. tionship to emplacement depth of magmatic sources: Geology, v. 37, p.
Krahulec, K.A., 1997, History and production of the West Mountain (Bing- 675−678.
ham) mining district, Utah: Society of Economic Geologists Guidebook Se- Redmond, P.B., and Einaudi, M.T., 2010, The Bingham Canyon porphyry
ries 29, p. 189−217. Cu-Mo-Au Deposit. I. Sequence of intrusions, vein formation, and sulfide
——2010, Production history of the Bingham mining district, Salt Lake deposition: ECONOMIC GEOLOGY, v. 105, p. 43−68.
county, Utah: Society of Economic Geologists Guidebook Series 41, p. Redmond, P.B., Einaudi, M.T., Inan, E.E., Landtwing, M.R., and Heinrich,
25−33. C.A., 2004, Copper deposition by fluid cooling in intrusion-centered sys-
Laes, D.Y.M., Krahulec, K.A., and Ballantyne, G.H., 1997, Geologic map of tems: New insights from the Bingham porphyry ore deposit, Utah: Geology,
the Oquirrh Mountains, Utah: Salt Lake City, Kennecott Utah Copper Cor- v. 32, p. 217−220.
poration. Reubi, O., and Blundy, J., 2009, A dearth of intermediate melts at subduction
Landtwing, M.R., Furrer, C., Redmond, P.B., Pettke, T., Guillong, M., and zone volcanoes and the petrogenesis of arc andesites: Nature, v. 461, p.
Heinrich, C.A., 2010, The Bingham Canyon porphyry Cu-Mo-Au deposit. 1269−1273.

0361-0128/98/000/000-00 $6.00 623

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022
624 STEINBERGER ET AL.

Reubi, O., and Nicholls, I.A., 2005, Structure and dynamics of a silicic mag- Stevens, C.H., and Armin, R.A., 1983, Microfacies of the middle Pennsyl-
matic system associated with caldera-forming eruptions at Batur volcanic vanian part of the Oquirrh group, central Utah: Geological Society of
field, Bali, Indonesia: Journal of Petrology, v. 46, p. 1367−1391. America Memoir 157, p. 83−100.
Richards, J.P., 2011, Magmatic to hydrothermal metal fluxes in convergent Tooker, E.W., 1999, Geology of the Oquirrh Mountains, Utah: U.S. Geolog-
and collided margins: Ore Geology Reviews, v. 40, p. 1−26. ical Survey Open-File Report 99-571, 150 p.
Rusk, B., Koenig, A., and Lowers, H., 2011, Visualizing trace element distri- Tosdal, R.M., Dilles, J.H., and Cooke, D.R., 2009, From source to sinks in
bution in quartz using cathodoluminescence, electron microprobe, and auriferous magmatic-hydrothermal porphyry and epithermal deposits: Ele-
laser ablation-inductively coupled plasma-mass spectrometry: American ments, v. 5, p. 289−295.
Mineralogist, v. 96, p. 703−708. Vogel, T.A., Cambray, F.W., and Constenius, K.N., 2001, Origin and em-
Schaltegger, U., Brack, P., Ovtcharova, M., Peytcheva, I., Schoene, B., placement of igneous rocks in the central Wasatch Mountains, Utah: Rocky
Stracke, A., Marocchi, M., and Bargossi, G.M., 2009, Zircon and titanite Mountain Geology, v. 36, p. 119−162.
recording 1.5 million years of magma accretion, crystallization and initial von Quadt, A., Erni, M., Martinek, K., Moll, M., Peytcheva, I., and Heinrich,
cooling in a composite pluton (southern Adamello batholith, northern C.A., 2011, Zircon crystallization and the life times of magmatic-hydro-
Italy): Earth and Planetary Science Letters, v. 286, p. 208−218. thermal ore systems: Geology, v. 39, p. 731–734.
Seo, J.H., Guillong, M., and Heinrich, C.A., 2012, Separation of molybde- Waite, K.A., Keith, J.D., Christiansen, E.H., Whitney, J.A., Hattori, K.H.,
num and copper in porphyry deposits: The roles of sulfur, redox, and pH in Tingey, D.G., and Hook, C.J., 1997, Petrogenesis of the volcanic and intru-
ore mineral deposition at Bingham Canyon.: ECONOMIC GEOLOGY, v. 107, sive rocks associated with the Bingham Canyon porphyry Cu-Au-Mo de-
p. 335−356. posit, Utah: Society of Economic Geologists Guidebook Series 29, p.
Shinohara, H., 2008, Excess degassing from volcanoes and its role on erup- 69−90.
tive and intrusive activity: Reviews of Geophysics, v. 46, 31 p. Williams-Jones, A.E., and Heinrich, C.A., 2005, Vapor transport of metals
Sillitoe, R.H., 1973, The tops and bottoms of porphyry copper deposits: ECO- and the formation of magmatic-hydrothermal ore deposits: ECONOMIC GE-
NOMIC GEOLOGY, v. 68, p. 799−815. OLOGY, v. 100, p. 1287−1312.
——2010, Porphyry copper systems: ECONOMIC GEOLOGY, v. 105, p. 3−41. Zajacz, Z., Seo, J.H., Candela, P.A., Piccoli, P.M., and Tossell, J.A., 2011, The
Sillitoe, R.H., and Mortensen, J.K., 2010, Longevity of porphyry copper for- solubility of copper in high-temperature magmatic vapors: A quest for the
mation at Quellaveco, Peru: ECONOMIC GEOLOGY, v. 105, p. 1157−1162. significance of various chloride and sulfide complexes: Geochimica et Cos-
Singer, D.A., Berger, V.I., Menzie, W.D., and Berger, B.R., 2005, Porphyry mochimica Acta, v. 75, p. 2811−2827.
copper deposit density: ECONOMIC GEOLOGY, v. 100, p. 491−514.

0361-0128/98/000/000-00 $6.00 624

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/108/4/605/3468719/605-624.pdf


by Raymond Rivera Cornejo
on 14 December 2022

You might also like