Santa Rita

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

©2021 Society of Economic Geologists, Inc.

Economic Geology, v. 116, no. 6, pp. 1267–1284

Evolution of the Magmatic-Hydrothermal System at the Santa Rita


Porphyry Cu Deposit, New Mexico, USA: Importance of Intermediate-Density
Fluids in Ore Formation
Subaru Tsuruoka,1,†,* Thomas Monecke,1 and T. James Reynolds1,2
1Center for Mineral Resources Science, Department of Geology and Geological Engineering, Colorado School of Mines,
1516 Illinois Street, Golden, Colorado 80401, USA
2FLUID INC., 1401 Wewatta Street #PH3, Denver, Colorado 80202, USA

Abstract
The Santa Rita porphyry Cu deposit in New Mexico, USA, is characterized by a stockwork of three vein types
that differ in morphology, mineralogy, and associated alteration assemblages. Early quartz veins associated
with potassic alteration are composed of recrystallized granular quartz grains that host ubiquitous hypersa-
line liquid and vapor-rich fluid inclusions. The early quartz likely formed at high (≳500°C) temperatures and
lithostatic pressures. Hypogene Cu mineralization at Santa Rita is in sulfide veins that reopened or crosscut
the early quartz veins. The sulfide veins are surrounded by alteration halos containing chlorite and K-feldspar.
Rare quartz crystals are present in some of these chalcopyrite and pyrite veins. The cores of the quartz crystals
contain hypersaline liquid and vapor-rich fluid inclusions, whereas the rims mostly contain hypersaline liquid
inclusions. The quartz crystals are interpreted to have formed close to the ductile-brittle transition as a result
of the pressure drop from lithostatic to hydrostatic conditions. Formation of the quartz crystals was postdated
by the deposition of Cu sulfides. Grain boundaries between the quartz and the sulfide minerals are irregular
in shape, with sulfide crosscutting growth zones in the quartz. The Cu sulfides are interpreted to have formed
from intermediate-density fluids that form secondary fluid inclusion assemblages in all earlier-formed quartz
types. Microthermometric investigations showed that these fluid inclusion assemblages homogenize at ~385°
to 435°C by critical or near-critical behavior and have salinities of <10 wt % NaCl equiv. The precipitation of
Cu sulfides occurred as a result of cooling of these fluids following their escape from the lithostatic into the
hydrostatic realm. Retrograde quartz solubility caused the corrosion of earlier-formed quartz during Cu sulfide
deposition. The youngest veins at Santa Rita are composed of quartz and pyrite. These veins are associated with
intense sericite alteration that overprinted all earlier alteration assemblages. The late quartz hosts primary and
secondary liquid-rich fluid inclusions, but no intermediate-density fluid inclusions were identified. This quartz
vein type formed at temperatures <360°C and hydrostatic pressures. The paragenetic relationships show that
hypogene Cu mineralization at Santa Rita postdated potassic alteration of the host rocks. The Cu mineralization
was formed by cooling intermediate-density fluids with critical or near-critical densities as they escaped from
lithostatic to hydrostatic conditions.
Introduction eralization (Nash and Theodore, 1971; Eastoe, 1978; Ulrich
Porphyry copper deposits are the world’s most important et al., 1999; Rusk et al., 2008; Maydagán et al., 2015; Gregory,
source of copper, with precious metals commonly being re- 2017), although some workers have suggested that low-salin-
covered as by-products. Hypogene ores in these deposits ity vapors may also be involved in Cu transport (Heinrich,
constitute large, low-grade stockwork and disseminated sul- 2005; Williams-Jones and Heinrich, 2005; Klemm et al., 2007;
fide zones that are spatially associated with porphyritic intru- Pudack et al., 2009; Mernagh et al., 2020). In some depos-
sions forming apophyses to larger stocks that have been em- its, however, zones of high Cu grades can be correlated with
placed within the upper (≤1–10 km) part of the crust (Nielsen, the occurrence of abundant veins lacking quartz but consist-
1968; Lowell and Guilbert, 1970; Gustafson and Hunt, 1975; ing almost entirely of Cu sulfides (Pollard and Taylor, 2002;
Seedorff et al., 2005; Sillitoe, 2010). Sapiie and Closs, 2013; Stefanova et al., 2014; Braxton et al.,
Mineralization in porphyry Cu deposits is widely regarded 2018; Sun et al., 2020). These veins, ranging in thickness up
to have formed early in the evolution of these magmatic- to several millimeters, crosscut the early quartz veins associ-
hydrothermal systems. Early quartz veins commonly contain ated with potassic alteration and in many cases appear to be
chalcopyrite or bornite, and the associated potassic-altered associated with chlorite-rich alteration. The lack of quartz in
host rocks may contain abundant disseminated sulfides. In these veins suggests that sulfide precipitation occurred at con-
addition, a broad spatial overlap between Cu contours and ditions of retrograde quartz solubility (Redmond et al., 2004;
the distribution of the potassic alteration is common in many Landtwing et al., 2010; Stefanova et al., 2014; Monecke et al.,
deposits. Based on fluid inclusion studies on the early quartz 2018, 2019; Sun et al., 2020), which is inconsistent with hy-
veins, many workers have concluded that hypersaline liquids persaline liquids being responsible for the formation of these
responsible for the potassic alteration also cause the Cu min- veins (Landtwing et al., 2010; Monecke et al., 2018, 2019).
This study reexamines the previous observations constrain-
†Corresponding author: e-mail, subaru.tsuruoka@icrag-centre.org
ing the relative timing and the nature of the mineralizing
*Present address: Irish Centre for Research in Applied Geosciences, School magmatic-hydrothermal fluids at the Santa Rita porphyry Cu
of Earth Sciences, University College Dublin, Belfield, Dublin 4, Ireland. deposit in New Mexico, USA (Leroy, 1954; Nielsen, 1968; Ja-
ISSN 0361-0128; doi:10.5382/econgeo.4831; 18 p.
Digital appendices are available in the online Supplements section. 1267 Submitted: February 10, 2020 / Accepted: December 25, 2020

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1268 TSURUOKA ET AL.

cobs and Parry, 1979; Ahmad and Rose, 1980; Parry et al., trend and are associated with felsic calc-alkalic intrusions that
1984; Reynolds and Beane, 1985). Nielsen (1968) showed that were emplaced into Precambrian basement and Paleozoic-
ore at Santa Rita is confined to thin sulfide veins crosscutting Mesozoic sedimentary rocks during the Laramide orogeny
the earliest quartz veins. Chlorite and K-feldspar are pres- (~70–55 Ma) as a result of plate subduction along an Andean-
ent in the alteration halos associated with these sulfide veins, type continental margin (Titley, 1993).
overprinting the earlier potassic alteration (Reynolds and The Precambrian basement rocks in the Santa Rita area
Beane, 1985). Previous fluid inclusion investigations on Santa consist of crystalline gneiss, micaceous schists, greenstone,
Rita focused on the early, high-temperature quartz stockwork and quartzite. The basement rocks are covered by a ~1.2-km-
veins associated with potassic alteration that formed prior thick succession of relatively undeformed sedimentary rocks.
to the ore and on the later quartz associated with phyllic al- This includes Cambrian to Permian quartzite, limestone, do-
teration that postdates mineralization, because the Cu sul- lostone, and shale, which are unconformably overlain by Cre-
fide veins do not contain macroscopically identifiable quartz taceous sandstone and shale (Fig. 1; Hemon et al., 1953; Rose
(Ahmad and Rose, 1980; Reynolds and Beane, 1985). In the and Baltosser, 1966; Jones et al., 1967).
current study, a combination of microanalytical techniques Igneous activity at Santa Rita commenced in the Late Cre-
including optical cathodoluminescence (CL) microscopy and taceous with the intrusion of quartz diorite porphyry form-
trace element analysis of quartz by electron microprobe was ing sills and laccoliths (Fig. 1; Kerr et al., 1950; Jones et al.,
employed to better constrain the quartz paragenesis at Santa 1967). The Santa Rita granodiorite stock was emplaced in the
Rita, as well as the textural relationships between the quartz Paleogene (~55 Ma; Hannink, 2010). Skarn is well-developed
and sulfide minerals in the veins. The new observations sug- around the contact between the intrusions and Cambrian
gest that neither the hypersaline liquids and vapors commonly limestone. Postmineralization igneous activity resulted in the
associated with potassic alteration nor the liquids responsible emplacement of several dike generations (i.e., quartz mon-
for the phyllic alteration formed the ore. It is proposed that zonite, quartz monzonite porphyry, latite, and quartz latite)
Cu sulfides at Santa Rita were deposited from a single-phase, that cut across the Santa Rita granodiorite stock and into the
magmatic-hydrothermal fluid of critical or near-critical den- surrounding sedimentary host rocks (Fig. 1; Kerr et al., 1950;
sity. Hemon et al., 1953; Rose and Baltosser, 1966; Jones et al.,
1967).
Geologic Setting Stockwork veins are developed in and around the Santa
The Santa Rita porphyry deposit is located in southwest New Rita granodiorite stock (Nielsen, 1968). The earliest vein type,
Mexico and currently exploited by the Chino open-pit opera- referred to as A veins using the nomenclature of Gustafson
tion. It is one of ~50 porphyry deposits that constitute the and Hunt (1975), consists of gray quartz (Fig. 2a). The A veins
American Southwest porphyry metallogenic belt. The por- lack internal symmetry or open vugs. They range in morphol-
phyry deposits are aligned along a northwest to southeast ogy from irregular and wavy to planar (Leroy, 1954; Nielsen,

Fig. 1. Geologic map of the Santa Rita porphyry Cu deposit, New Mexico. Geology from Rose and Baltosser (1966) and Jones
et al. (1967). Geology of the Santa Rita granodiorite stock and distribution of potassic alteration from Nielsen (1968). Location
of Whim Hill breccia from Reynolds and Beane (1985). The pit outline reflects the current stage of mining. The Whim Hill
breccia has been mined out entirely. Map on left of figure shows the location of the study area.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1269

Fig. 2. Hand specimen photographs of the different vein types recognized at Santa Rita as well as a representative sample of
the Whim Hill breccia. a. Stockwork of A veins. The veins are associated with hydrothermal K-feldspar and biotite. Quartz in
the A veins is referred to as QA. Sample RSR6. b. Narrow C vein consisting of pyrite and chalcopyrite. The vein is surrounded
by a narrow alteration halo containing chlorite. Igneous biotite is altered to chlorite. Sample RSR5-2a. c. Narrow C vein
hosted by altered granodiorite. The alteration halo surrounding the vein consists of chlorite and hydrothermal K-feldspar.
Igneous biotite and earlier-formed hydrothermal biotite are altered to chlorite. Sample RSR10. d. Stockwork of different
vein types illustrating crosscutting relationships. A narrow, irregularly shaped A vein consisting of gray quartz is the earliest
vein type in the sample. The A vein is crosscut by C veins that have different orientations and range from fractures coated by
pyrite and chalcopyrite to narrow veins that are up to ~1 mm in thickness. Late D veins consisting of pyrite and minor quartz
crosscut both earlier vein types. The granodiorite hosting the veins has been affected by pervasive sericite alteration. The
presence of chalcopyrite in the C veins was confirmed petrographically. Sample 25148. e. Late D vein formed by reopening
of an earlier A vein. The white quartz is referred to as QD. The vein is surrounded by sericite-altered granodiorite. Sample
25134. f. Euhedral quartz crystals in the Whim Hill breccia. The cores of large quartz crystals have different inclusion char-
acteristics from the rims and are referred to as QBL. The outer parts of the crystals, as well as the smaller euhedral quartz
crystals, are composed of quartz referred to as QD. In addition to quartz, small euhedral pyrite crystals are present in open
spaces. Sample RSR15. Bt = hydrothermal biotite, Chl = chlorite, Chl. Bt = chloritized igneous biotite, Cpy = chalcopyrite,
Ksp = hydrothermal K-feldspar, Py = pyrite, Ser = sericite.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1270 TSURUOKA ET AL.

1968). They are associated with alteration halos in which ig- LMH Schottky field emission-scanning electron microscope
neous biotite and hornblende phenocrysts are replaced by (FE-SEM) equipped with a single-crystal YAG backscat-
pale-brown hydrothermal secondary biotite. Hydrothermal tered electron (BSE) detector. Imaging was conducted at a
K-feldspar replaced igneous plagioclase and pervaded the working distance of 10 mm and an accelerating voltage of
groundmass (Nielsen, 1968; Jacobs and Parry, 1979). 15 kV. Semiquantitative chemical analyses of minerals were
Thin veins composed of chalcopyrite and pyrite crosscut the performed by energy-dispersive X-ray spectroscopy (EDS)
early A veins (Fig. 2b-d; Nielsen, 1968). These are referred to using an attached Bruker XFlash 6/30 silicon drift detector.
as C veins following the nomenclature of Dilles and Einaudi Quartz veins and samples from the Whim Hill breccia were
(1992) and Gustafson and Quiroga (1995). In most cases, these also investigated by optical CL microscopy. A HC5-LM hot
veins are less than a few millimeters in width and macroscopi- cathode CL microscope by Lumic Special Microscopes, Ger-
cally resemble planar hairline fractures coated with sulfides many, was used and operated at 14 kV and a current density of
as opposed to classic stockwork veins (Fig. 2c). When broken ~10 µA mm–2 (Neuser, 1995). CL images were captured with
open, the sulfide minerals occur as coating on fracture surfac- a high-sensitivity, double-stage Peltier cooled Kappa DX40C
es. The C veins are the carrier of hypogene copper mineral- CCD camera. Acquisition times of CL images of quartz ranged
ization at Santa Rita (Nielsen, 1968). Although Nielsen (1968) from 8 to 10 s. In addition to the study of the fluid inclusion
reported the alteration around these veins to be typified by inventory of vein quartz, it was attempted to study pyrite in
green biotite, Reynolds and Beane (1985) demonstrated that the C veins using a DAGE-MTI LSC-70 2,200-nm infrared
the alteration selvages surrounding the veins are dominated camera attached to the optical microscope. However, no trans-
by chlorite and minor K-feldspar (Fig. 2b, c). Both igneous parent pyrite suitable for fluid inclusion studies was present.
biotite phenocrysts and hydrothermal biotite are commonly The distributions of Al, K, and Ti in quartz showing differ-
replaced by pale-green chlorite along vein selvages. ent CL responses were determined by electron microprobe
Both earlier vein types are reopened or crosscut and off- analysis using a JEOL JXA-8900 at the U.S. Geological Sur-
set by late veins consisting of quartz, pyrite, and sericite (Fig. vey in Denver. Trace element mapping was conducted at an
2d, e), referred to as D veins according to the nomenclature accelerating voltage of 20 kV and a beam current of 100 nA
of Gustafson and Hunt (1975). The D veins are the young- (measured on the Faraday cup). A focused electron beam was
est vein type at Santa Rita and most abundant at the margin used, with dwell times of 1 s/pixel. The trace element maps
of the Santa Rita granodiorite stock. They are typically up to for the three elements were collected simultaneously, with
several centimeters thick and occur as continuous linear veins one spectrometer and a TAP analyzing crystal being used to
that are associated with extensive sericite alteration halos that measure Al, one spectrometer and a PET analyzing crystal
overprint all earlier alteration mineral assemblages (Leroy, measuring K, and the three remaining spectrometers detect-
1954; Nielsen, 1968; Parry et al., 1984; Reynolds and Beane, ing Ti using LiF analyzing crystals. In addition to mapping,
1985). spot analyses were conducted along traverses across selected
Several breccia pipes have been exposed within the San- quartz crystals. Spots along the traverses were chosen careful-
ta Rita granodiorite stock during the mining operation. The ly to avoid surface pits and mineral inclusions. The operating
Whim Hill breccia was the largest breccia body (500 × 100 m) conditions for spot analyses included an accelerating voltage
located in the northern part of the granodiorite stock (Fig. 1). of 20 kV, a beam current of 50 nA (measured on Faraday cup),
The breccia consisted of angular fragments of granodiorite, and a beam defocused to 10 μm to minimize specimen dam-
with rock flour forming the breccia matrix. Euhedral quartz, age. The analysis of Al, K, and Ti was performed using the dis-
K-feldspar, apatite, chlorite, and pyrite cemented open space tribution of spectrometers and analyzing crystals mentioned
within the breccia (Fig. 2f; Kerr et al., 1950; Reynolds and above. Count times of 10 min on the peak and 5 min on each
Beane, 1985). high and low background were applied. Alternating on- and
off-peak acquisition was employed to account for carbon con-
Materials and Methods tamination buildup during the long analysis times. Aggregate
This study is based on samples originally collected from the intensities of the on- and off-peak positions of the duplicate
open pit of the Santa Rita porphyry deposit for the research elements and a blank correction were used to improve the
conducted by Reynolds and Beane (1985). Details on sam- counting statistics (Donovan et al., 2011). Calculated detec-
ple locations are given by Reynolds and Beane (1985). The tion limits on spot analyses are 8 ppm for Al, 15 ppm for K,
samples represent all three porphyry vein types recognized and 13 ppm for Ti. The results of the electron microprobe
at Santa Rita and include samples from the Whim Hill brec- analyses are given in Appendix 1.
cia. Emphasis was placed on reinvestigating samples of the Limited microthermometry was performed on intermedi-
C veins as these are the main host to mineralization. Doubly ate-density fluid inclusions in quartz grains that were also
polished 80-µm-thick sections were prepared from the vein studied by the analytical methods described above. Fol-
samples. In addition, thick sections previously prepared for lowing the procedures outlined in Goldstein and Reynolds
the study by Reynolds and Beane (1985) were examined. (1994), microthermometric investigations were performed
The vein mineralogy, fluid inclusion petrography, and the on assemblages of inclusions with a FLUID INC.-adapted
composition of the alteration halos associated with the veins U.S. Geological Survey gas-flow heating and freezing stage
were studied in transmitted and reflected light using an Olym- following calibration using synthetic fluid inclusions at
pus BX51 optical microscope. Classification of fluid inclusion –56.6°, 0.0°, and 374.0°C. Final apparent ice-melting tem-
types followed Monecke et al. (2018). Micron-scale textural peratures were determined by bracketing to 0.1°C using
relationships were investigated with a TESCAN MIRA3 the cycling technique described by Goldstein and Reynolds

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1271

(1994). Measurement was not always possible, because the Table 1. Microthermometric Data Obtained on Intermediate-Density
inclusions were too small or because clathrate crystals were Fluid Inclusions from Santa Rita Showing Critical or Near-Critical
Homogenization Behavior
present during cycling. Cycling could not be employed for
precise homogenization temperature determinations, as the Homogeniza- Number of Temperature Salinity
phase boundary separating liquid and vapor forms within a tion tempera- inclusions in of last ice (wt % Evidence for
few degrees or less with cooling after homogenization. Fi- ture (°C) assemblage melting (°C) NaCl equiv) clathrate
nal homogenization was, therefore, approximated within a Sample RSR15a
temperature interval of 5°C. The microthermometric data 385–390 3 – – –
390–395 2 –1.1 1.9 No
obtained are reported in Table 1. 400–405 1 –7.2 10.7 Yes
410–415 2 – – –
Results 430–435 3 –3.0 5.0 No
Sample RSR11b
A veins 395–400 1 – – –
400–405 1 – – –
The A veins at Santa Rita are associated with potassic altera-
tion characterized by the occurrence of secondary K-feldspar Note: – = not observed
and biotite in the altered wall rocks. The veins consist of anhe-
dral quartz grains, referred to here as QA (Fig. 3). The quartz
grains exhibit undulose extinction, and adjacent grains typi- the vein walls or are entirely enclosed by the sulfide minerals.
cally have different extinction angles. Growth banding is rare Contacts between the quartz crystals and sulfide minerals are
in the QA grains (Fig. 3a, b). The granular nature of the quartz typically scalloped and irregular (Fig. 5a, b).
and the rare occurrence of primary growth banding suggest The QB crystals show differences in optical CL properties
that quartz has undergone extensive recrystallization. from core to rim, which broadly correlate with variations in
QA is characterized by a bright, short-lived, light-blue CL trace element concentrations observed along electron mi-
emission. During continued electron bombardment, the croprobe traverses. Based on the combination of optical CL
short-lived blue CL transitions to a long-lived royal-blue or microscopy and electron microprobe investigations, two dif-
indigo emission. Growth banding visible in plane-polarized ferent subtypes of QB can be distinguished (App. 1; Figs. 4,
light can be recognized based on subtle color variations (Fig. 5). The earliest type of quartz, referred to as QBE, is character-
3c). Locally, the QA is crosscut by a second type of quartz that ized by a dark-purple or dark-red CL emission. The QBE has
has a dark-purple CL (Fig. 3c). high Al concentrations ranging from 63 to 3,414 ppm, with a
Electron microprobe spot analyses revealed that QA in the median value of 208 ppm. Many of the analytical spots have
A veins contains 87 to 1,309 ppm Al, with a median value K concentrations below the detection limits of 15 ppm. The
of 120 ppm Al. The concentrations of K vary from <15 to maximum K value measured is 999 ppm and the median value
793 ppm. The quartz has a median K value of 28 ppm. The is 96 ppm. The Ti concentrations range from <13 to 86 ppm,
measured Ti concentrations range from 21 to 199 ppm, with with a median value of 30 ppm (App. 1; Fig. 4). The dark-red
a median of 55 ppm (App. 1; Fig. 4). The concentrations of QBE is characterized by higher Al contents when compared
all three elements are typically higher in quartz exhibiting the to the dark-purple QBE (App. 1; Fig. 5c, d). The dark-red and
royal-blue CL emission than quartz of indigo CL color. dark-purple QBE form complex interlocking textures. In some
QA contains abundant fluid inclusions, some of which ex- cases, isolated islands of dark-purple QBE are surrounded by
hibit textural evidence for postentrapment modification (Fig. dark-red QBE (Fig. 5b). Elsewhere, both types of quartz oc-
3d-g) but many of which do not (Fig. 3h). The abundance cur as growth bands within larger quartz crystals (Fig. 5c).
of fluid inclusions is typically so high that individual planes The later QB subtype, referred to as QBL, is characterized by a
of secondary fluid inclusions cannot be readily identified. In- maroon CL emission. The quartz forms an overgrowth on QBE
clusion types identified in the A veins at Santa Rita include (Fig. 5a-c). Zones of QBL have low trace element concentra-
hypersaline liquid inclusions, vapor-rich fluid inclusions, in- tions. The Al concentrations range from 23 to 366 ppm, with
termediate-density fluid inclusions, and liquid-rich fluid in- a median value of 87 ppm. The K and Ti contents of QBL are
clusions (Fig. 3h). generally below detection (App. 1; Fig. 4).
Grains of QBE typically host abundant secondary hypersa-
C veins line liquid inclusions (Fig. 5e), vapor-rich fluid inclusions, and
The A veins at Santa Rita are crosscut by C veins. These veins rare intermediate-density fluid inclusions (Fig. 5f). Obvious
are thin (<5 mm) with planar walls and commonly contain visual evidence for postentrapment modification of the fluid
dark-green chlorite within the fractures and in alteration sel- inclusions is typically absent, but small variations in the liquid
vages with K-feldspar, where chlorite replaced igneous and to vapor volumetric proportions indicate that some hypersa-
earlier-formed hydrothermal biotite. In most cases, the C line liquid inclusions could have been affected by postentrap-
veins do not contain quartz. In rare instances where minor ment modification. Vapor bubbles and halite daughter crystals
quartz is present in the C veins, it is not optically distinct from in hypersaline liquid inclusions in QBE are smaller than many
QA and was not recognized as a separate quartz type in previ- found in QA. In contrast to QBE, the QBL is largely devoid of
ous studies. However, because of the distinct textural setting, fluid inclusions and appears distinctly clear under the opti-
this quartz is referred to here as QB. The rare quartz forms cal microscope. Hypersaline liquid inclusions are present, but
anhedral to euhedral grains that are ~200 to 500 mm in their vapor-rich inclusions are exceedingly rare. Reconnaissance
longest dimension (Fig. 5). The quartz crystals occur along heating experiments show that the hypersaline liquid inclu-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1272 TSURUOKA ET AL.

Fig. 3. Photomicrographs, optical cathodoluminescence (CL) images, and fluid inclusion inventory of QA forming A veins at
Santa Rita. a. Plane-polarized light image of QA. The quartz grains are anhedral with irregular grain boundaries. The arrow
highlights the location of faint growth zones in one of the large quartz grains. b. Crossed-polarized light image of the same field
of view as in Figure 3a. The quartz grains have different orientations. The arrows highlight contacts between polygonal grains
forming triple junctions that have interfacial angles of ~120°, providing strong evidence for recrystallization. The quartz shows
first-order colors, as the sample is a thick section. c. Optical CL image of the quartz grain in Figure 3a and b showing relict
growth banding. The quartz exhibits a stable royal-blue to indigo luminescence. The growth bands are locally crosscut by dark-
purple quartz (arrows) interpreted to be QBE. The locations of microprobe spot analyses are given and are keyed to Appendix
1. d. High-magnification image of QA containing abundant decrepitated fluid inclusions, some of which are highlighted by the
arrows. e. Image of a hypersaline liquid inclusion in QA showing irregular surfaces (arrow), interpreted to have been caused by
postentrapment modification. f. Fluid inclusion that has been affected by postentrapment modification. g. Intermediate-den-
sity fluid inclusion with surfaces that suggest postentrapment modification. h. Fluid inclusion inventory in QA. Hypersaline liq-
uid inclusions, vapor-rich fluid inclusions, and liquid-rich fluid inclusions are ubiquitous in QA, as is reported for many porphyry
deposits worldwide (Monecke et al., 2018). In addition, intermediate-density inclusions can be found but easily mistaken for
vapor-rich fluid inclusions. Note that intermediate-density fluid inclusions not affected by postentrapment modification show
negative crystal shapes. The hypersaline liquid inclusions contain halite and commonly other daughter phases including sylvite,
anhydrite, magnetite, and hematite (cf. fig. 6 in Reynolds and Beane, 1985). Small triangular opaque crystals interpreted to be
chalcopyrite can be seen in some of the hypersaline liquid inclusions. Due to the high abundance of intersecting microfrac-
tures, individual planes of secondary fluid inclusions are difficult to discern. All images are from sample RSR6. HS = hypersa-
line liquid inclusion, ID = intermediate-density fluid inclusion, L = liquid-rich fluid inclusion, V = vapor-rich fluid inclusion.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1273

Fig. 4. Histogram plots of Al, K, and Ti concentrations in the different types of quartz occurring at the Santa Rita porphyry
Cu deposit. Trace element analyses were performed by electron microprobe. DL = detection limit.

sions typically homogenize by halite disappearance after the ing the presence of a clathrate compound. For the inclusions
vapor. Grains of QBL are also crosscut by trails of secondary where final ice-melting temperatures could be determined,
intermediate-density fluid inclusions (Fig. 5f, g), which may the apparent salinities are <10 wt % NaCl equiv (Table 1).
contain a tiny opaque daughter mineral (Fig. 5g).
Many intermediate-density fluid inclusions in QB have neg- D veins
ative crystal shapes and contain large vapor bubbles (Fig. 5f, The D veins at Santa Rita consist of quartz and pyrite.
g). Only a few fluid inclusion assemblages could be identi- These veins have straight walls and are commonly associ-
fied that are of sufficient size to allow confident microther- ated with intense sericitic alteration halos that pervasively
mometric investigations (Table 1). The inclusions in these overprint earlier-formed hydrothermal minerals. Quartz in
assemblages showed critical or near-critical homogenization the D veins is referred to here as QD. It forms large sub-
behavior (cf. Klyukin et al., 2019) at ~385° to 435°C. Some hedral to euhedral grains that exhibit a CL emission that
intermediate-density fluid inclusions exhibited a double jerk ranges from dark to light red brown (Fig. 6a-d). The trace
on cooling, and ice could not be obtained by cycling, suggest- element concentrations of QD are generally low. The con-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1274 TSURUOKA ET AL.

Fig. 5. Photomicrographs, optical cathodoluminescence (CL) images, and fluid inclusion characteristics of QB crystals in chal-
copyrite-pyrite veins at Santa Rita. a. Plane-polarized light image of quartz crystals occurring along the wall of a chalcopyrite-
pyrite vein. The quartz consists of early QBE characterized by the presence of hypersaline liquid and vapor-rich fluid inclusions
(not shown). A distinct overgrowth of later QBL can be recognized by the paucity of fluid inclusions. The quartz exhibits scal-
loped grain boundaries in contact with chalcopyrite. Sample RSR15a. b. Optical CL image of the same field of view as shown in
Figure 5a. The QBE is characterized by a dark-purple or dark-red emission. The later overgrowth of QBL exhibits a maroon CL.
The locations of microprobe spot analyses are given and keyed to Appendix 1. c. Optical CL image of a quartz crystal enclosed
by chalcopyrite. Most of the crystal consists of QBE showing a dark-purple or dark-red emission. A small clear overgrowth of
QBL with maroon CL occurs along the upper right and lower left sides of the crystal. The chalcopyrite crosscuts the contact
between QBE and QBL. The locations of microprobe spot analyses are given and keyed to Appendix 1. Sample RSR15a. d. Elec-
tron microprobe map of the same crystal showing the distribution of Al in the quartz crystal (dark blue = low concentrations,
light blue = elevated concentrations). The dark-red optical CL emission of QBE coincides with elevated Al contents. The appar-
ent differences in grain shape between the optical CL image and the electron microprobe map are a result of the experimental
setup. The optical CL images are collected looking through the back of the thin section during electron bombardment of the
polished surface. e. Hypersaline liquid inclusions occurring in QBE. The small inclusions form a secondary assemblage that
does not appear to have been affected by postentrapment modification. Sample RSR15a. f. Large, secondary, intermediate-
density fluid inclusions hosted by QBE. The inclusions homogenize by critical behavior at ~400° to 405°C. Sample RSR15a.
g. Intermediate-density fluid inclusions containing an opaque daughter crystal (arrows). The inclusions do not appear to have
been affected by textural modification after their entrapment. Sample RSR12. Cpy = chalcopyrite.

centrations of Al range from 85 to 1,186 ppm, with a median clusions were previously determined to range from ~260° to
value of 220 ppm. The K and Ti contents of QD is typically 360°C, with maximum salinities of up to ~9 wt % NaCl equiv
below detection limit (App. 1; Fig. 4). In general, the optical (Reynolds and Beane, 1985).
CL properties and trace element characteristics of QD are
similar to those of QBL. Whim Hill breccia
Grains of QD contain rare primary and abundant secondary The Whim Hill breccia consists of abundant wall-rock frag-
liquid-rich fluid inclusions (Fig. 6e). No intermediate-density ments set in a matrix of rock flour. Open spaces contain large
fluid inclusions are present in QD. Except for necking, tex- (up to 3 cm) euhedral quartz crystals (Figs. 2f, 7a-e). The rock
tures suggestive of postentrapment modification are absent, fragments are commonly potassic altered and host earlier A
and liquid to vapor volumetric proportions within individual veins. Chlorite is present along growth zones in the quartz
fluid inclusion assemblages are usually consistent. Homogeni- crystals (Fig. 7a, b) or as a coating on top of the quartz crystals
zation temperatures and salinities of the liquid-rich fluid in- with pyrite, suggesting that chlorite formed synchronous with

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1275

Fig. 6. Photomicrographs, optical cathodoluminescence (CL) images, and fluid inclusion characteristics of QD in late D veins
at Santa Rita. a. Large quartz grain of QD that is surrounded by pyrite. b. Optical CL image of the same field of view as in Fig-
ure 6a illustrating that QD exhibits a red-brown emission. Subtle growth zoning can be recognized in the quartz. The locations
of microprobe spot analyses are given and keyed to Appendix 1. c. Euhedral quartz grain grown into open space, forming a
vuggy centerline in the same quartz-pyrite vein. d. Optical CL image of the same field of view as in Figure 6c. The QD growth
zones are characterized by red-brown and dark red-brown emissions. e. Quartz in the D veins contains only liquid-rich fluid
inclusions, some of which are highlighted by the arrows. The shapes of the secondary fluid inclusions in the various assem-
blages vary from negative crystal shaped to irregularly shaped, correlating with the observation that a wide range of homog-
enization temperatures (~250°–350°C) was determined (Reynolds and Beane, 1985) for the different primary and secondary
fluid inclusion assemblages in this sample. No evidence of postentrapment modification of fluid inclusions is present in QD,
except that the irregular shapes of many inclusions are a result of necking. All images are from sample RSR6a. Py = pyrite.

the late growth zones of the coarse euhedral quartz crystals low. Electron microprobe analyses of QD contained in the
and also after deposition of most of the quartz. Whim Hill breccia yielded values of 24 to 2,586 ppm Al, with
Crystals cut perpendicular to the c axis show growth bands a median value of 288 ppm. The K and Ti concentrations are
(Fig. 7a) defined by variations in CL intensity and color. The typically below the limit of detection although values as high
cores of large euhedral quartz crystals exhibit a maroon CL as 308 ppm K and 47 ppm Ti have been registered (App. 1;
(Fig. 7b), locally with dark-purple growth bands, which is simi- Fig. 4).
lar to the signature of QBL. Trace element analyses of the quartz The cores of the large quartz crystals of the Whim Hill
indicate that Al concentrations vary from 50 to 1,084 ppm, breccia are characterized by the presence of rare hypersa-
with a median value of 138 ppm. The K content ranges from line liquid inclusions (Fig. 7f). Vapor-rich fluid inclusions are
<15 to 663 ppm K, while Ti analyses returned values largely exceedingly rare. In contrast, the late growth zones contain
below the detection limits. Only a few analytical spots yielded only liquid-rich fluid inclusions, many of which define growth
values of up to 1,100 ppm Al and 700 ppm K (App. 1; Fig. 4). bands (Fig. 7g). Homogenization temperatures and salini-
Late growth zones of the quartz crystals have a dark to ties of these primary liquid-rich fluid inclusions are similar to
light red-brown CL emission (Fig. 7b, d), which resembles those recorded for QD (Reynolds and Beane, 1985).
the signature of QD in D veins. Well-developed oscillatory
growth zoning can be observed. Trace element maps ob- Discussion
tained by electron microprobe analyses revealed that the Al The research presented here confirms key findings of previ-
concentrations vary between individual growth bands visible ous paragenetic studies at Santa Rita (Nielsen, 1968; Reyn-
in CL (Fig. 7e), whereas the Ti concentrations are uniformly olds and Beane, 1985). However, application of a combination

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1276 TSURUOKA ET AL.

Fig. 7. Photomicrographs, optical cathodoluminescence (CL) images, and fluid inclusion characteristics of quartz crystals
from the Whim Hill breccia at Santa Rita. a. Crossed-polarized light image of a quartz crystal from the Whim Hill breccia cut
approximately perpendicular to the c axis. The core of the crystal consists of QBL and the outer portion of the crystal is com-
posed of QD. Individual growth zones highlighted by the arrows are defined by primary liquid-rich fluid inclusions. Chlorite
occurs in the growth zones formed after QBL. Sample RSR5. b. Optical CL image of the same field of view as in Figure 7a.
Both quartz types are characterized by growth zoning and are difficult to distinguish based on their emission color. The loca-
tions of microprobe spot analyses are given and keyed to Appendix 1. c. Plane-polarized light image of a quartz crystal from
the Whim Hill breccia cut parallel to the c axis. The part of the crystal shown consists of QD hosting liquid-rich fluid inclusions
only. Sample RSR5-2b. d. Optical CL image of the same field of view as in Figure 7c. The crystal exhibits a red-brown to dark
red-brown CL emission and is characterized by well-developed oscillatory and sector zoning. The locations of microprobe
spot analyses are given and keyed to Appendix 1. e. Electron microprobe map of the same field of view as in Figure 7c, d
showing the distribution of Al in the quartz crystal. The subtle variations in emission color coincide with variations in Al
concentration (dark blue = low concentrations, red = high concentrations). f. Small hypersaline liquid inclusions (highlighted
by arrows) hosted by QBL of the Whim Hill breccia. Sample RSR5-2b. g. Growth zone defined by primary, liquid-rich fluid
inclusions within a late overgrowth of QD. The growth zone is crosscut by a secondary assemblage of liquid-rich fluid inclu-
sions. Sample RSR5. Chl = chlorite.

of microanalytical techniques, including optical CL micros- al., 2010; Landtwing et al., 2010; Stefanova et al., 2014; Sun
copy and electron microprobe analysis, allowed distinction et al., 2020). At Santa Rita, identification of different quartz
of quartz types not previously identified. The fluid inclusion types had to rely on a combination of analytical methods, in-
petrographic characteristics of these quartz types provide new cluding optical microscopy, fluid inclusion petrography, CL
insights into the evolution of the magmatic-hydrothermal sys- investigations, and electron microprobe analysis. The use of
tem at Santa Rita and the processes responsible for Cu miner- any single analytical method would not have been successful.
alization at this deposit. For instance, QA and QBE have similar fluid inclusion inven-
tories and display only subtle differences in their CL charac-
Paragenetic sequence and distinction of quartz types teristics. However, the trace element concentrations in both
The results of this study highlight that the reconstruction of quartz types are distinct. On the other hand, the quartz types
paragenetic relationships in porphyry veins is challenging, as QBL and QD can be readily distinguished based on their fluid
repeated reopening of the veins during the evolution of the inclusion inventory, as the QD quartz contains no hypersaline
magmatic-hydrothermal systems results in complex over- liquid inclusions. However, the optical CL emission and trace
printing relationships at the microscale (Penniston-Dorland, element concentrations in both quartz types are similar (Fig.
2001; Rusk and Reed, 2002; Pudack et al., 2009; Müller et 8). In addition, the abundances of Al, K, and Ti vary signifi-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1277

Fig. 8. Summary diagram illustrating the characteristics of the different vein and quartz types recognized at Santa Rita. CL
= cathodoluminescence.

cantly within a given quartz type, and adjacent growth zones Fluid inclusion petrography shows that hypersaline liquid
of the same quartz type can show large compositional differ- and vapor-rich fluid inclusions predominate in QA (Ahmad and
ences. This suggests that the incorporation of trace elements Rose, 1980; Reynolds and Beane, 1985; this study), suggest-
into the quartz was controlled by nonequilibrium processes. ing that A vein formation took place at conditions correspond-
ing to those of the two-phase field of the H2O-NaCl system
Early quartz vein formation (Fig. 9a; Bodnar, 1995). However, conditions must have also
The A veins associated with potassic alteration represent the intersected the single-phase field, as intermediate-density in-
earliest vein type at Santa Rita. The irregular and wavy shape clusions are also present (Fig. 9a). Because of the high tem-
of some A veins, which includes crenulated and ptygmatically peratures of the cooling but already solidified intrusion, fluid
folded veins (Nielsen, 1968), suggests that early vein forma- flow through the granodiorite must have occurred at ductile
tion occurred in a host rock that exhibited plastic behavior conditions, precluding open fracture development except
(Gustafson and Hunt, 1975; Brathwaite et al., 2001; Sillitoe, during short periods of high strain (cf. Fournier, 1991, 1998).
2010). Temperatures of formation of the veins may have been Numerical simulations by Weis et al. (2012) showed that fluid
≳500°C as suggested by the potassic alteration around the flow through ductile rocks self-organizes into overpressure-
veins (Jacobs and Parry, 1979) and isotopic evidence (Shep- permeability waves. Within the upward-moving waves, pres-
pard et al., 1971). sures exceed lithostatic conditions, resulting in the creation of
The formation of the A veins followed the emplacement of permeability. Behind the wave, pressure drops to lithostatic
the Santa Rita stock at ~55 Ma (Hannink, 2010). Magmatic- conditions, resulting in quartz precipitation and sealing of the
hydrothermal fluids derived from the magma chamber were permeability. Over the lifetime of a magmatic-hydrothermal
rising through the already solidified part of the porphyritic in- system, many of these waves pass through the porphyry, creat-
trusion, forming the early quartz veins at high temperatures. ing a stockwork of veins.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1278 TSURUOKA ET AL.

Fig. 9. Pressure-temperature diagrams illustrating the conditions at which different vein types encountered at Santa Rita are
likely to have formed. The diagrams show the phase relationships in the H2O-NaCl model system at a salinity of 5 wt % NaCl
equiv. Overlain in a, c, and d are quartz solubility isopleths in millimoles SiO2 per kilogram water. a. The QA in the A veins con-
tains abundant hypersaline liquid and vapor inclusions suggesting that the early quartz formed at high temperatures within the
two-phase field of the H2O-NaCl system. Intermediate-density inclusions are also present and were entrapped at somewhat
lower temperatures within the single-phase field. The orange field outlines the lithostatic pressure conditions and temperature
range at which the early veins are inferred to have formed. Cooling and decompression of the magmatic-hydrothermal fluids
during fluid ascent from a magma chamber causes a decrease in quartz solubility, resulting in quartz vein formation. The room
temperature petrographic characteristics of fluid inclusions entrapped in the quartz are shown schematically. The irregular
shape of the fluid inclusions entrapped at lithostatic conditions results from postentrapment modification of the inclusions that
occured during later decompression from lithostatic to hydrostatic conditions. b. The QB present in reopened A veins and in
some C veins formed as the magmatic-hydrothermal system transitioned from lithostatic to hydrostatic conditions close to the
ductile to brittle boundary. The inferred temperature and pressure conditions of quartz formation are outlined by the blue
field. The early QBE containing hypersaline liquid and vapor-rich inclusions formed within the two-phase field of the H2O-NaCl
system. The later QBL contains hypersaline liquid inclusions but only exceedingly rare vapor-rich inclusions, suggesting that this
quartz type formed as a result of cooling of the hypersaline liquid at conditions that allowed physical segregation and escape
of the buoyant vapor phase. The fluid inclusions in both quartz types have been trapped at hydrostatic conditions, as evidence
for postentrapment textural modification is largely absent. c. The C veins are interpreted to have formed from intermediate-
density fluids that have critical to near-critical densities. Cooling of these fluids would have resulted in the precipitation of
sulfides under conditions of retrograde quartz solubility. The temperature and pressure conditions are schematically outlined
by the blue field. As no quartz precipitated from these fluids, intermediate-density fluid inclusions can only be observed as
secondary assemblages in earlier formed quartz. d. The late QD in the D veins formed as a result of cooling of magmatic-
hydrothermal liquids at hydrostatic conditions. The blue field schematically indicates the inferred conditions of QD formation.
The QD contains primary and secondary liquid-rich fluid inclusions. Phase relationships from Driesner and Heinrich (2007).
Quartz solubility isopleths and granodiorite solidus from Monecke et al. (2018). Quartz solubility isopleths in the two-phase
region represent the total quartz solubility in the liquid and coexisting vapor. CP = critical point, GD solidus = H2O-saturated
granodiorite solidus, F = single-phase fluid, H = halite, L = liquid, V = vapor.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1279

The petrographic characteristics of QA in the A veins at (1985), because it is rare and indistinguishable from QA with a
Santa Rita suggest that the quartz has undergone extensive standard petrographic microscope.
recrystallization. Most notable are the vitreous nature of the The QB at Santa Rita is interpreted to be equivalent to the
quartz in hand specimen and the granular appearance of the quartz present in B veins in other porphyry Cu deposits (Mo-
quartz under the microscope. The widespread occurrence of necke et al., 2018). Similar to Santa Rita, this type of quartz
polygonal grains forming triple junctions that have interfacial commonly occurs in reopened A veins (Redmond et al., 2004;
angles of ~120° provides strong evidence for grain-scale re- Müller et al., 2010; Bennett et al., 2014; Skewes, 2016; Sun
crystallization. The recrystallized QA at Santa Rita is charac- et al., 2020). For instance, Redmond et al. (2004) reported
terized by a homogeneous CL emission, and remnant grains that porphyry veins at Bingham Canyon in Utah contain early
showing growth zoning are only locally preserved (Fig. 3a, quartz, interpreted to be equivalent to QA of this study, that
b). Recrystallization of QA must have occurred during the occurs as interlocking anhedral grains. The quartz has a mod-
early evolution of the magmatic-hydrothermal system when erate to bright luminescence and shows irregular zoning in
the Santa Rita granodiorite hosting the veins was still ductile. panchromatic CL images. The second quartz generation, in-
Repeated reopening and sealing of the early porphyry veins terpreted to be QB of this study, occurs along fractures and
during the passage of the overpressure-permeability waves shows well-developed oscillatory growth zoning. Similar ob-
modeled by Weis et al. (2012) may have resulted in recrystal- servations have been reported by Müller et al. (2010) from
lization of earlier-formed quartz and the precipitation of new the Central Oyu Tolgoi and Zesen Uul deposits in Mongolia
quartz that is recrystallized during a subsequent fluid flow and by Chang et al. (2018) from the Yulong deposit in China.
event. This process of recrystallization appears to be a fun- The younger quartz type at these deposits exhibits oscillatory
damental process in porphyry deposits, as quartz in A veins growth banding or wavy banding and is equivalent to QB of
of porphyry deposits worldwide shows similar petrographic this study. At Santa Rita, the rare QB crystals lining C veins can
characteristics (Gustafson and Hunt, 1975; Hedenquist et al., be viewed as microscopic B veins that have been reopened
1998; Landtwing et al., 2010; Stefanova et al., 2014; Monecke during the formation of the Cu mineralization as suggested by
et al., 2018; Acosta et al., 2020). This contrasts with quartz in Monecke et al. (2018, 2019).
later vein types formed subsequent to a decline in tempera- Based on optical CL characteristics, fluid inclusion petrog-
ture and the concomitant shift from ductile to brittle behavior raphy, and trace element abundances, two subtypes of QB
of the host rocks. Quartz in the late veins is commonly euhe- are recognized at Santa Rita. The earlier QBE hosts abundant
dral to subhedral (Gustafson and Hunt, 1975; Hedenquist et hypersaline liquid and vapor-rich fluid inclusions, while few
al., 1998; Monecke et al., 2018). hypersaline liquid inclusions are present in the later QBL. Va-
The QA in the A veins at Santa Rita contains fluid inclu- por-rich fluid inclusions are exceedingly rare in QBL. The fluid
sions that were entrapped during the passage of the overpres- inclusion evidence suggests that both subtypes of QB were
sure-permeability waves and the subsequent evolution of the precipitated within the two-phase field of the H2O-NaCl sys-
magmatic-hydrothermal system. Many of the fluid inclusions, tem. However, much of QBL must have formed from hyper-
including hypersaline liquid and vapor-rich inclusions as well saline liquids that remained after a vapor phase had escaped
as inclusions of intermediate density, show textural evidence (Fig. 9b), which is indicated by the homogenization behav-
for postentrapment modification (Fig. 3d-g). Experimental ior of many of the fluid inclusions involving homogenization
investigations suggest that these textural modifications are a by halite disappearance. The smaller size of the salt crystals
result of pronounced changes in pressure (Sterner and Bod- in the hypersaline liquid inclusions hosted in QB when com-
nar, 1989), which presumably occurred after the formation pared to many of the hypersaline liquid inclusions present
of the A veins as the magmatic-hydrothermal system cooled in QA suggests that the hypersaline liquid forming QB was of
and transitioned from lithostatic to hydrostatic pressures (cf. lower salinity (Klyukin et al., 2019). Inspection of the phase
Gustafson and Hunt, 1975; Monecke et al., 2018; Sun et al., relationships in the H2O-NaCl system (Monecke et al., 2018;
2020). Postentrapment modification makes these fluid inclu- Klyukin et al., 2019) shows that the salinity of the hypersaline
sion assemblages unsuitable for the measurement of homog- liquid produced as a result of phase separation decreases with
enization temperatures and perhaps salinities (Sterner and temperature, confirming that the transition from QA to QB re-
Bodnar, 1989; Chang et al., 2018; Sun et al., 2020). In con- cords cooling of the magmatic-hydrothermal system.
trast, ubiquitous secondary fluid inclusions trapped in QA late The formation of QB at Santa Rita occurred as the host
in the evolution of the magmatic-hydrothermal system at hy- granodiorite was cooling below the transition from ductile
drostatic conditions do not show evidence for postentrapment to brittle behavior, which occurs at ~400°C in silicic rocks
modification. These fluid inclusion assemblages have consis- in hydrothermal systems (Fournier, 1991, 1998). Early dur-
tent liquid to vapor volumetric proportions and yield reliable ing this transition, and at least temporarily sealed conditions,
homogenization temperatures (Reynolds and Beane, 1985) the hypersaline liquid and the vapor produced as a result of
but are not related to QA formation (Sun et al., 2020). phase separation were cotrapped in the QBE (Fig. 9b). The
younger QBL contains rare hypersaline liquid inclusions and
Lithostatic-hydrostatic transition exceedingly rare vapor-rich inclusions, as the quartz presum-
Some of the C veins at Santa Rita contain rare QB crystals that ably formed under conditions where a throughgoing fracture
occur along the vein walls or are entirely enveloped by the network was developed in such a way that the buoyant vapor
sulfide minerals. The QB is younger than the QA, as C veins phase was physically separated from the dense, sluggish hy-
crosscut A veins. This second generation of quartz at Santa persaline liquid. The QBL precipitated during cooling of the
Rita was not previously recognized by Reynolds and Beane hypersaline liquid that was initially formed in the two-phase

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1280 TSURUOKA ET AL.

field of the H2O-NaCl system (Fig. 9b). The textural evidence 2013; Stefanova et al., 2014; Braxton et al., 2018) do not con-
suggests that gaping fractures developed during QB formation tain quartz that coprecipitated with the sulfide minerals, the
and that these remained open long enough to allow formation conditions of C vein formation cannot be directly constrained
of the euhedral quartz. The transition from ductile to brittle through fluid inclusion studies. For this reason, reconstruc-
behavior of the wall rocks and the development of open frac- tion of the conditions of mineralization at Santa Rita had to
ture networks must have been accompanied by a drop in pres- be based on indirect evidence. In this contribution, the fluid
sure from lithostatic to hydrostatic conditions (Gustafson and inclusion inventory in quartz formed prior to the sulfides is
Hunt, 1975; Fournier, 1991, 1998; Monecke et al., 2018; Sun compared to the fluid inclusion signature of quartz precipitat-
et al., 2020). The phase relationships in the H2O-NaCl system ed after the mineralizing event. The careful paragenetic stud-
dictate that the phase separation producing the hypersaline ies at Santa Rita reveal that QBL represents the last quartz type
liquid and coexisting vapor must have occurred at hydrostatic formed prior to the deposition of the sulfide minerals. The
pressures, as single-phase conditions extend toward lower QBL contains hypersaline liquid inclusions and exceedingly
pressures with decreasing temperature (Fig. 9b). Formation rare vapor-rich inclusions. The QBL is crosscut by secondary
of QB at hydrostatic conditions is also supported by the fact intermediate-density fluid inclusions. Deposition of sulfides
that visual evidence for postentrapment modification of most in the C veins was postdated by the formation of the D veins
of the fluid inclusion assemblages in QB is absent. at Santa Rita. The only fluid inclusion type recognized in QD
The nature of the hydrothermal alteration associated with is liquid-rich fluid inclusions, which form primary and second-
QB deposition is not well established at Santa Rita. Potassic ary fluid inclusion assemblages.
alteration of the host rocks occurred during QA formation. The petrographic relationships suggest that sulfide forma-
Sulfide deposition postdating QB formation was associated tion at Santa Rita marked a significant change in the nature of
with chlorite plus K-feldspar alteration of the host rocks. The the magmatic-hydrothermal system. The QBL was deposited
textural evidence suggests that the magmatic-hydrothermal from hypersaline liquids produced in the two-phase field of
fluids forming QB quartz was K-feldspar stable. This agrees the H2O-NaCl system. The lack of postentrapment modifi-
with observations at El Salvador in Chile (Gustafson and cation textures and the scarcity of vapor-rich fluid inclusions
Hunt, 1975). These authors noted that B veins at this deposit suggest that the QBL formed at hydrostatic conditions in a
are not associated with notable alteration halos but crosscut throughgoing fracture network, allowing the escape of the
host rocks that have experienced earlier pervasive potassic al- vapor phase produced by phase separation. The QBL is cross-
teration. cut by intermediate-density fluid inclusions entrapped under
single-phase conditions in the H2O-NaCl system. The lack of
Hypogene ore formation postentrapment modification textures of these inclusions also
The evidence presented by Nielsen (1968) indicates that Cu suggests entrapment at hydrostatic conditions. Inspection of
mineralization at Santa Rita did not occur as early in the evo- the phase relationships in the H2O-NaCl system indicates that
lution of the magmatic-hydrothermal system as A veins and such a change from two-phase to single-phase conditions can
the associated potassic alteration. The relative abundance of only occur as a result of a temperature decrease of the mag-
A veins is not correlative with high Cu grades. Ore zones at matic-hydrothermal system at hydrostatic pressure (Fig. 9c).
Santa Rita are characterized by the abundance of thin C veins Deposition of sulfide minerals under brittle conditions in a
that crosscut the earlier A veins and by the mineralized grano- hydrostatic pressure regime is consistent with the observation
diorite containing disseminated sulfides. At the microscopic that the C veins represent hairline fractures that are coated
scale, disseminated sulfides commonly occur along micro- by sulfide minerals. Quartz precipitation after sulfide deposi-
fractures cutting the host rock and the early formed A veins. tion took place within the single-phase field of the H2O-NaCl
Similar observations have been made at other porphyry de- system but at lower temperatures, as indicated by the higher
posits where C veins account for much or all of the hypogene density of the entrapped liquid-rich fluid inclusions. Temper-
ore (Dilles and Einaudi, 1992; Gustafson and Quiroga, 1995; atures of homogenization measured for the liquid-rich fluid
Manske and Paul, 2002; Xiao et al., 2011; Sapiie and Closs, inclusions in the D veins range up to ~360°C (Reynolds and
2013; Stefanova et al., 2014; Sun et al., 2020). Beane, 1985), providing a lower temperature constraint on
The narrow C veins at Santa Rita are almost entirely com- the conditions of Cu mineralization.
posed of pyrite and chalcopyrite (Nielsen, 1968), although The observation that the precipitation of QBL was postdated
traces of molybdenite can be present. When broken open, the by the passage of intermediate-density magmatic-hydrother-
sulfide minerals coat fracture surfaces similar to the charac- mal fluids is of significance, as these fluids are most likely re-
teristics of C veins described by Fournier (1967) from Liberty sponsible for the hypogene Cu mineralization at Santa Rita.
in Nevada, Stefanova et al. (2014) from Elatsite in Bulgaria, The importance of intermediate-density fluids at this deposit
and Sun et al. (2020) from Yulong in China. In many cases, was not recognized by the previous fluid inclusion studies (Ah-
chalcopyrite and pyrite are sprinkled along the fracture sur- mad and Rose, 1980; Reynolds and Beane, 1985), although
faces. At Santa Rita, the C veins are characterized by a lack the existence of these inclusions was mentioned by Ahmad
of quartz. Quartz is only present where the C veins crosscut and Rose (1980). As part of this study, microthermometric
earlier A veins or reopen fractures that contain minor QB as investigations were conducted on the intermediate-density
described above. In both cases, the quartz texturally predates fluid inclusions, although only few assemblages with inclu-
the sulfide minerals. sions large enough to perform complete microthermometric
As the C veins at Santa Rita and similar sulfide veins in other experiments were identified (Table 1). Microthermometry
porphyry deposits (Pollard and Taylor, 2002; Sapiie and Closs, showed that the intermediate-density fluid inclusions homog-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1281

enize at ~385° to 435°C and have salinities of <10 wt % NaCl under conditions of chlorite and K-feldspar stability, which
equiv (Table 1). is the alteration assemblage stable during the formation of
Additional important constraints on the nature of the min- the hypogene copper mineralization. During continued fluid
eralizing fluids at Santa Rita are provided by the fact that evolution, chlorite-sericite and sericite were stabilized. The D
quartz is essentially absent in the C veins or, where pres- veins at Santa Rita are associated with sericite alteration. This
ent, is paragenetically earlier than the sulfide minerals and difference in alteration style likely reflects a change in fluid
is corroded in contact with the sulfide phases. This implies acidity, which is controlled by fluid/rock ratio and tempera-
that sulfide precipitation occurred at conditions of retro- ture (cf. Seedorff et al., 2005).
grade quartz solubility (Fig. 9c). Quartz solubility modeling
by Monecke et al. (2018, 2019) showed that retrograde solu- Late quartz vein formation
bility occurs in intermediate-density fluids having salinities of The formation of the D veins at Santa Rita postdated the
<10 wt % NaCl equiv and low CO2 contents (<2 mol %) as a mineralizing event, as these veins are barren and crosscut all
result of cooling. In contrast to the intermediate-density fluid earlier vein types. The QD in these late veins contains only liq-
inclusions observed to crosscut QBL, retrograde quartz solu- uid-rich fluid inclusions, demonstrating that quartz formation
bility does not occur in high-salinity fluids (Monecke et al., occurred in the single-phase field of the H2O-NaCl system
2018), ruling out that the hypersaline liquids that formed QBL (Fig. 9d). Microthermometric investigations on the liquid-
also could have formed the sulfide minerals in the C veins. rich fluid inclusions by Reynolds and Beane (1985) yielded
The C veins at Santa Rita are associated with alteration ha- homogenization temperatures of 260° to 360°C and salinities
los containing chlorite and K-feldspar. Thermodynamic mod- of up to 9 wt % NaCl equiv.
eling shows that chlorite and K-feldspar coexistence occurs at The host rocks of the D veins were brittle during QD pre-
~300°C and 1,000 bar (Seedorff et al., 2005) and transitions cipitation, resulting in the formation of planar veins. Hydro-
toward higher temperatures at lower pressures (Beane and Ti- static conditions prevailed within the interconnected fracture
tley, 1981). This is broadly consistent with the conclusion that network, now occupied by the D veins. At the temperatures
the Cu mineralization at Santa Rita formed at temperatures of QD formation, quartz precipitation occurred as a result of
exceeding ~360°C but below the ductile to brittle transition. cooling of the magmatic-hydrothermal fluids (Fig. 9d). The
Given the range of homogenization temperatures observed liquid could have undergone intermittent dilution by ambient
for the intermediate-density fluid inclusions and the quartz meteoric water, possibly explaining the observed variations in
solubility constraints, Cu mineralization appears to have oc- the salinity and homogenization temperatures of the liquid-
curred at maximum temperatures of ~450°C. rich fluid inclusions (cf. Reynolds and Beane, 1985). However,
isotopic studies performed on the sericite alteration halos of
Origin of the Whim Hill breccia other porphyry deposits suggest that influx of meteoric water
The Whim Hill breccia, located in the north-central part of is probably minimal and that fluid mixing may be restricted to
the Santa Rita stock, was a large elliptical breccia body. The the marginal sericite alteration (Hedenquist et al., 1998; Har-
breccia contained angular blocks of the stock, some of which ris and Golding, 2002).
hosted A veins. This provides unequivocal evidence that brec-
cia formation postdated early quartz vein formation. Breccia Implications
formation was followed by the deposition of quartz crystals The hypogene Cu sulfide mineralization at Santa Rita oc-
into the open space created. Based on fluid inclusion evi- curred after the host intrusion underwent potassic alteration
dence, optical CL imaging, and trace element analysis, the during the early formation of A veins. At the deposit scale,
large euhedral quartz crystals in the Whim Hill breccia are Nielsen (1968) showed that ore grade does not directly cor-
mostly composed of QBL. relate with the occurrence of A veins and the distribution of
The Whim Hill breccia presumably formed either when or potassic alteration. Establishment of crosscutting relation-
after the Santa Rita stock transitioned from ductile to brittle ships at the hand specimen scale illustrated that the C veins
behavior. In contrast to the process suggested by Norton and consistently crosscut the earlier A veins (Nielsen, 1968). Tex-
Cathles (1973), formation of the Whim Hill breccia likely oc- tural relationships suggest that Cu sulfides, where present
curred as a result of rapid decompression and escape of over- within earlier A veins, postdate the high-temperature quartz
pressured magmatic-hydrothermal fluids (cf. Sillitoe, 1985), and were introduced after the A vein formation, presumably
perhaps during a pressure blowout from the lithostatic to the because of reopening of the existing veins at the same time
hydrostatic region (Fournier, 1991). The observation that QBL as the C veins were formed. At a microscopic scale of obser-
grew into open space and the fluid inclusion inventory of the vation, critical textural relationships proving coprecipitation
quartz suggest that hydrostatic conditions were established of quartz and Cu sulfides, such as the occurrence of sulfide
before or during the precipitation of QBL. inclusions encapsulated along growth bands in quartz, have
Following breccia formation and initial deposition of QBL, never been encountered at Santa Rita. Quartz in contact
QD was precipitated as the latest growth zones of the large with the Cu sulfide minerals commonly shows irregular grain
euhedral quartz crystals in the Whim Hill breccia. The outer boundaries, suggesting that sulfide deposition was accompa-
growth zones of QD contain chlorite aggregates and rosettes. nied by quartz dissolution.
The K-feldspar present in potassic-altered rock fragments is Although sulfide veins resembling the C veins at Santa
not altered, suggesting that the magmatic-hydrothermal fluid Rita have been previously reported to occur at other depos-
forming the quartz was K-feldspar stable. This implies that its (Fournier, 1967; Pollard and Taylor, 2002; Arif and Baker,
quartz formation in the Whim Hill breccia occurred initially 2004; Sapiie and Closs, 2013; Stefanova et al., 2014; Braxton

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1282 TSURUOKA ET AL.

et al., 2018; Rinne et al., 2018; Sun et al., 2020), there is cur- 2020). At Santa Rita, there is no evidence for remobilization
rently no consensus as to the general importance of this vein of Cu from A veins into the ore shell from what would now
type. The observations at Santa Rita suggest that the relative be a barren or subeconomic part of the deposit (cf. Brimhall,
importance of these veins in porphyry deposits is probably 1979). There are no supporting observations that Cu sulfides
underappreciated. The C veins are unassuming in nature, had ever precipitated at high temperatures or textural rela-
as most are hairline fractures coated by Cu sulfide minerals tionships that might indicate dissolution of previously existing
forming paint on the fracture surfaces (cf. Fournier, 1967), as Cu sulfides.
opposed to classic stockwork veins that formed through infill Conclusions
of gaping fractures.
At Santa Rita, the C veins are associated with chlorite plus This study reexamined the paragenetic relationships of stock-
K-feldspar alteration. This alteration style overprints the ear- work veins at the Santa Rita porphyry copper deposit, New
lier potassic alteration (Reynolds and Beane, 1985). In hand Mexico, to determine the evolution of the magmatic-hydro-
specimen, the chlorite and K-feldspar alteration associated thermal system that formed this deposit and to determine the
with these veins could be easily mistaken for potassic altera- conditions at which hypogene ore formation took place. A
tion if the fine-grained, dark chlorite is misidentified as bio- range of analytical techniques was employed, including fluid
tite or, alternatively, the chlorite could be misinterpreted as inclusion petrography, optical CL microscopy, and trace ele-
part of a late intermediate argillic overprint. It is important ment analysis by electron microprobe, to distinguish the dif-
to note that despite careful study, Nielsen (1968) misidenti- ferent quartz types formed before and after the mineralizing
fied the fine-grained chlorite as green biotite in the field. Cor- event. The careful petrographic investigations indicate that
rect identification of the alteration style associated with the Cu mineralization at Santa Rita postdated the early high-tem-
C veins at Santa Rita required thin section petrography and perature vein formation and the associated potassic alteration.
electron microprobe analysis (Reynolds and Beane, 1985). At Santa Rita, the Cu mineralization occurs in sulfide veins
Staining with sodium cobaltinitrite commonly applied to sup- that lack quartz as a gangue mineral. The paucity of quartz as
port field investigations (Duuring et al., 2009; Byrne et al., a coprecipitate with the Cu sulfide minerals hampers direct
2020; Cooke et al., 2020) may not provide conclusive insights identification of the conditions of sulfide formation through
into the nature of alteration associated with the C veins, as fluid inclusion studies. This study demonstrates that important
K-feldspar occurs in the potassic-altered rock as well as in the constraints on the nature of the mineralizing fluids can still be
chlorite plus K-feldspar alteration. derived based on the comparison of the fluid inclusion inven-
The current study suggests that single-phase, magmatic- tory in quartz formed prior to and after the Cu mineralization.
hydrothermal fluids having a critical or near-critical density Based on this comparison, the Cu mineralization at Santa Rita
(critical density of a 5 wt % NaCl fluid is 0.48 g/cm3; Monecke is interpreted to have occurred from a single-phase, low-sa-
et al., 2018) caused the Cu mineralization at Santa Rita as they linity (<10 wt % NaCl equiv), magmatic-hydrothermal fluid
escaped from lithostatic to hydrostatic conditions and cooled. of critical or near-critical density at temperatures exceeding
Secondary fluid inclusions of this fluid type crosscut the last ~360°C but below ~450°C.
quartz formed prior to the precipitation of the Cu sulfide Acknowledgments
minerals. The careful petrographic investigations show that
intermediate-density fluid inclusions do not occur in quartz We are indebted to H. Lowers, D. Adams, and M. Bennett
formed after the Cu sulfides. This observation provides evi- for help provided during the electron microprobe investiga-
dence for the relative timing of passage of these fluids through tions at the U.S. Geological Survey in Denver, Colorado. L.
the Santa Rita deposit. Quartz solubility modeling shows that Fisher and T.O. Wyatt are thanked for assistance provided
the lack of quartz as a gangue mineral in the C veins and the during the FE-SEM investigations. This study benefited from
occurrence of dissolution textures of quartz in contact with discussions with A. Gysi, E. Holley, and K. Pfaff on the geo-
Cu sulfides can be explained by the retrograde solubility of chemistry of magmatic-hydrothermal systems. The research
quartz in intermediate-density fluids. This provides strong ev- formed part of a Ph.D. study by ST that was financially sup-
idence for the nature of the mineralizing fluids, as other fluid ported by Japan Oil, Gas and Metals National Corporation.
types, in particular high-salinity liquids commonly invoked We thank an anonymous reviewer and D. Cooke for insightful
to be responsible for mineralization in porphyry deposits, do reviews that helped us to improve an earlier version of this
not cause retrograde quartz solubility (Monecke et al., 2018). manuscript. D. Cooke is also thanked for editorial handling of
Thermodynamic modeling also shows that intermediate-den- the manuscript.
sity fluids with near-critical densities are particularly suited as REFERENCES
ore-forming fluids, as they have distinctively high mass and Acosta, M.D., Watkins, J.M., Reed, M.H., Donovan, J.J., and DePaolo, D.J.,
energy transport capacities (Norton and Knight, 1977; John- 2020, Ti-in-quartz: Evaluating the role of kinetics in high temperature
son and Norton, 1991; Norton and Dutrow, 2001; Klyukin et crystal growth experiments: Geochimica et Cosmochimica Acta, v. 281, p.
149–167.
al., 2016). The findings at Santa Rita support previous studies Ahmad, S.N., and Rose, A.W., 1980, Fluid inclusions in porphyry and skarn
hypothesizing that Cu sulfide formation in porphyry deposits ore at Santa Rita, New Mexico: Economic Geology, v. 75, p. 229–250.
occurs primarily as a result of fluid cooling over a narrow tem- Arif, J., and Baker, T., 2004, Gold paragenesis and chemistry at Batu Hijau,
perature range that broadly overlaps with the ductile-brittle Indonesia: Implications for gold-rich porphyry copper deposits: Minera-
lium Deposita, v. 39, p. 523–535.
transition of the host rocks (Redmond et al., 2004; Klemm et Beane, R.E., and Titley, S.R., 1981, Porphyry copper deposits. Part II. Hydro-
al., 2007; Landtwing et al., 2010; Kouzmanov and Pokrovski, thermal alteration and mineralization: Economic Geology 75th Anniversary
2012; Stefanova et al., 2014; Monecke et al., 2018; Sun et al., Volume, p. 235–269.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
SANTA RITA PORPHYRY Cu DEPOSIT, NEW MEXICO 1283

Bennett, M., Monecke, T., Reynolds, T.J., Ricks, J., and Muntean, J., 2014, Harris, A.C., and Golding, S.D., 2002, New evidence of magmatic-fluid–
Cathodoluminescence and fluid inclusion characteristics of porphyry vein related phyllic alteration: Implications for the genesis of porphyry Cu
quartz [ext. abs.]: U.S. Geological Survey and Colorado State University, deposits: Geology, v. 30, p. 335–338.
Pan-American Current Research on Fluid Inclusions Conference, 12th, Hedenquist, J.W., Arribas, A., Jr., and Reynolds, T.J., 1998, Evolution of an
Denver, Colorado, 2014, Program and Abstracts, p. 61–62. intrusion-centered hydrothermal system: Far Southeast-Lepanto porphyry
Bodnar, R.J., 1995, Fluid-inclusion evidence for a magmatic source for metals and epithermal Cu-Au deposits, Philippines: Economic Geology, v. 93, p.
in porphyry copper deposits: Mineralogical Association of Canada, Short 373–404.
Course Series, v. 23, p. 139–152. Heinrich, C.A., 2005, The physical and chemical evolution of low-salinity
Brathwaite, R.L., Simpson, M.P., Faure, K., and Skinner, D.N.B., 2001, Tele- magmatic fluids at the porphyry to epithermal transition: A thermodynamic
scoped porphyry Cu-Mo-Au mineralisation, advanced argillic alteration and study: Mineralium Deposita, v. 39, p. 864–889.
quartz-sulphide-gold-anhydrite veins in the Thames district, New Zealand: Hemon, R.M., Jones, W.R., and Moore, S.L., 1953, Some geological features
Mineralium Deposita, v. 36, p. 623–640. of the Santa Rita quadrangle, New Mexico: New Mexico Geological Society,
Braxton, D.P., Cooke, D.R., Ignacio, A.M., and Waters, P.J., 2018, Geology Fourth Field Conference Guidebook, p. 117–130.
of the Boyongan and Bayugo porphyry Cu-Au deposits: An emerging por- Jacobs, D.C., and Parry, W.T., 1979, Geochemistry of biotite in the Santa
phyry district in northeast Mindanao, Philippines: Economic Geology, v. Rita porphyry copper deposit, New Mexico: Economic Geology, v. 74, p.
113, p. 83–131. 860−887.
Brimhall, G.H., 1979, Lithologic determination of mass transfer mechanisms Johnson, J.W., and Norton, D., 1991, Critical phenomena in hydrothermal
of multiple-stage porphyry copper mineralization at Butte, Montana: Vein systems: State, thermodynamic, electrostatic, and transport properties of
formation by hypogene leaching and enrichment of potassium-silicate H2O in the critical region: American Journal of Science, v. 291, p. 541−648.
protore: Economic Geology, v. 74, p. 556–589. Jones, W.R., Hernon, R.M., and Moore, S.L., 1967, General geology of Santa
Byrne, K., Lesage, G., Gleeson, S.A., Piercey, S.J., Lypaczewski, P., and Rita quadrangle, Grant County, New Mexico: U.S. Geological Survey, Pro-
Kyser, K., 2020, Linking mineralogy to lithogeochemistry in the Highland fessional Paper 555, 144 p.
Valley Copper district: Implications for porphyry copper footprints: Eco- Kerr, P.F., Kulp, J.L., Patterson, C.M., and Wright, R.J., 1950, Hydrothermal
nomic Geology, v. 115, p. 871–901. alteration at Santa Rita, New Mexico: Bulletin of the Geological Society of
Chang, J., Li, J.W., and Audétat, A., 2018, Formation and evolution of mul- America, v. 61, p. 275–347.
tistage magmatic-hydrothermal fluids at the Yulong porphyry Cu-Mo Klemm, L.M., Pettke, T., Heinrich, C.A., and Campos, E., 2007, Hydro-
deposit, eastern Tibet: Insights from LA-ICP-MS analysis of fluid inclu- thermal evolution of the El Teniente deposit, Chile: Porphyry Cu-Mo ore
sions: Geochimica et Cosmochimica Acta, v. 232, p. 181–205. deposition from low-salinity magmatic fluids: Economic Geology, v. 102, p.
Cooke, D.R., Wilkinson, J.J., Baker, M., Agnew, P., Phillips, J., Chang, Z., 1021–1045.
Chen, H., Wilkinson, C.C., Inglis, S., Hollings, P., Zhang, L., Gemmell, J.B., Klyukin, Y.I., Driesner, T., Steele-MacInnis, M., Lowell, R.P., and Bodnar,
White, N.C., Danyushevsky, L., and Martin, H., 2020, Using mineral chem- R.J., 2016, Effect of salinity on mass and energy transport by hydrothermal
istry to aid exploration: A case study from the Resolution porphyry Cu-Mo fluids based on the physical and thermodynamic properties of H2O-NaCl:
deposit, Arizona: Economic Geology, v. 115, p. 813−840. Geofluids, v. 16, p. 585–603.
Dilles, J.H., and Einaudi, M.T., 1992, Wall-rock alteration and hydrother- Klyukin, Y.I., Steele-MacInnis, M., Lecumberri-Sanchez, P., and Bodnar,
mal flow paths about the Ann-Mason porphyry copper deposit, Nevada—A R.J., 2019, Fluid inclusion phase ratios, compositions and densities from
6-km vertical reconstruction: Economic Geology, v. 87, p. 1963−2001. ambient temperature to homogenization, based on PVTX properties of
Donovan, J.J., Lowers, H.A., and Rusk, B.G., 2011, Improved electron probe H2O-NaCl: Earth-Science Reviews, v. 198, article 102924.
microanalysis of trace elements in quartz: American Mineralogist, v. 96, p. Kouzmanov, K., and Pokrovski, G.S., 2012, Hydrothermal controls on metal
274−282. distribution in porphyry Cu (-Mo-Au) systems: Society of Economic Geolo-
Driesner, T., and Heinrich, C.A., 2007, The system H2O-NaCl. Part I: Cor- gists, Special Publication 16, p. 573–618.
relation formulae for phase relations in temperature-pressure-composition Landtwing, M.R., Furrer, C., Redmond, P.B., Pettke, T., Guillong, M., and
space from 0 to 1000°C, 0 to 5000 bar, and 0 to 1 XNaCl: Geochimica et Heinrich, C.A., 2010, The Bingham Canyon porphyry Cu-Mo-Au deposit.
Cosmochimica Acta, v. 71, p. 4880−4901. III. Zoned copper-gold ore deposition by magmatic vapor expansion: Eco-
Duuring, P., Rowins, S.M., McKinley, B.S.M., Dickinson, J.M., Diakow, L.J., nomic Geology, v. 105, p. 91–118.
Kim, Y.S., and Creaser, R.A., 2009, Magmatic and structural controls on por- Leroy, P.G., 1954, Correlation of copper mineralization with hydrothermal
phyry-style Cu-Au-Mo mineralization at Kemess South, Toodoggone district alteration in the Santa Rita porphyry copper deposit, New Mexico: Bulletin
of British Columbia, Canada: Mineralium Deposita, v. 44, p. 435−462. of the Geological Society of America, v. 65, p. 739–768.
Eastoe, C.J., 1978, A fluid inclusion study of the Panguna porphyry copper Lowell, J.D., and Guilbert, J.M., 1970, Lateral and vertical alteration-min-
deposit, Bougainville, Papua New Guinea: Economic Geology, v. 73, p. eralization zoning in porphyry ore deposits: Economic Geology, v. 65, p.
721–748. 373−408.
Fournier, R.O., 1967, The porphyry copper deposit exposed in the Liberty Manske, S.L., and Paul, A.H., 2002, Geology of a major new porphyry cop-
open-pit mine near Ely, Nevada. Part I. Syngenetic formation: Economic per center in the Superior (Pioneer) district, Arizona: Economic Geology,
Geology, v. 62, p. 57−81. v. 97, p. 197–220.
——1991, The transition from hydrostatic to greater than hydrostatic fluid Maydagán, L., Franchini, M., Rusk, B., Lentz, D.R., McFarlane, C., Impic-
pressure in presently active continental hydrothermal systems in crystalline cini, A., Ríos, F.J., and Rey, R., 2015, Porphyry to epithermal transition
rock: Geophysical Research Letters, v. 18, p. 955–958. in the Altar Cu-(Au-Mo) deposit, Argentina, studied by cathodolumines-
——1998, Hydrothermal processes related to movement of fluid from plastic cence, LA-ICP-MS, and fluid inclusion analysis: Economic Geology, v. 110,
into brittle rock in the magmatic-epithermal environment: Economic Geol- p. 889−923.
ogy, v. 94, p. 1193–1211. Mernagh, T.P., Leys, C., and Henley R.W., 2020, Fluid inclusion systematics
Goldstein, R.H., and Reynolds, T.J., 1994, Systematics of fluid inclusions in porphyry copper deposits: The super-giant Grasberg deposit, Indonesia,
in diagenetic minerals: Society for Sedimentary Geology (SEPM) Short as a case study: Ore Geology Reviews, v. 123, article 103570.
Course, v. 31, 199 p. Monecke, T., Monecke, J., Reynolds, T.J., Tsuruoka, S., Bennett, M.M.,
Gregory, M.J., 2017, A fluid inclusion and stable isotope study of the Pebble Skewes, W.B., and Palin, R.M., 2018, Quartz solubility in the H2O-NaCl
porphyry copper-gold-molybdenum deposit, Alaska: Ore Geology Reviews, system: A framework for understanding vein formation in porphyry copper
v. 80, p. 1279–1303. deposits: Economic Geology, v. 113, p. 1007–1046.
Gustafson, L.B., and Hunt, J.P., 1975, The porphyry copper deposit at El Monecke, T., Monecke, J., and Reynolds, T.J., 2019, The influence of CO2 on
Salvador, Chile: Economic Geology, v. 70, p. 857–912. the solubility of quartz in single-phase hydrothermal fluids: Implications for
Gustafson, L.B., and Quiroga, J.G., 1995, Patterns of mineralization and the formation of stockwork veins in porphyry copper deposits: Economic
alteration below the porphyry copper orebody at El Salvador, Chile: Eco- Geology, v. 114, p. 1195–1206.
nomic Geology, v. 90, p. 2–16. Müller, A., Herrington, R., Armstrong, R., Seltmann, R., Kirwin, D.J.,
Hannink, R.L., 2010, Mineralization and timing of the Lover’s Lane breccia, Stenina, N.G., and Kronz, A., 2010, Trace elements and cathodolumines-
Santa Rita porphyry Cu-Mo deposit, Grant County, New Mexico: Unpub- cence of quartz in stockwork veins of Mongolian porphyry-style deposits:
lished M.Sc. thesis, Reno, University of Nevada, 139 p. Mineralium Deposita, v. 45, p. 707–727.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo
1284 TSURUOKA ET AL.

Nash, J.T., and Theodore, T.G., 1971, Ore fluids in the porphyry copper Sillitoe, R.H., 1985, Ore-related breccias in volcanoplutonic arcs: Economic
deposit at Copper Canyon, Nevada: Economic Geology, v. 66, p. 385–399. Geology, v. 80, p. 1467−1514.
Neuser, R.D., 1995, A new high-intensity cathodoluminescence microscope ——2010, Porphyry copper systems: Economic Geology, v. 105, p. 3–41.
and its application to weakly luminescing minerals: Bochumer Geologische Skewes, W.B., 2016, Evolution of magmatic-hydrothermal systems below
und Geotechnische Arbeiten, v. 44, p. 116–118. boiling conditions: Evidence from the Çöpler Au-Cu deposit, east central
Nielsen, R.L., 1968, Hypogene texture and mineral zoning in a copper-bear- Turkey: Unpublished M.Sc. thesis, Golden, Colorado, Colorado School of
ing granodiorite porphyry stock, Santa Rita, New Mexico: Economic Geol- Mines, 65 p.
ogy, v. 63, p. 37–50. Stefanova, E., Driesner, T., Zajacz, Z., Heinrich, C.A., Petrov, P., and Vasilev,
Norton, D.L., and Cathles, L.M., 1973, Breccia pipes—Products of exsolved Z., 2014, Melt and fluid inclusions in hydrothermal veins: The magmatic to
vapor from magmas: Economic Geology, v. 68, p. 540–546. hydrothermal evolution of the Elatsite porphyry Cu-Au deposit, Bulgaria:
Norton, D.L., and Dutrow, B.L., 2001, Complex behavior of magma-hydro- Economic Geology, v. 109, p. 1359–1381.
thermal processes: Role of supercritical fluid: Geochimica et Cosmochi- Sterner, S.M., and Bodnar, R.J., 1989, Synthetic fluid inclusions—VII. Re-
mica Acta, v. 65, p. 4009−4017. equilibration of fluid inclusions in quartz during laboratory-simulated
Norton, D., and Knight, J., 1977, Transport phenomena in hydrothermal sys- metamorphic burial and uplift: Journal of Metamorphic Geology, v. 7, p.
tems: Cooling plutons: American Journal of Science, v. 277, p. 937–981. 243–260.
Parry, W.T., Ballantyne, J.M., and Jacobs, D.C., 1984, Geochemistry of hydro- Sun, M., Monecke, T., Reynolds, T.J., and Yang, Z., 2020, Understanding
thermal sericite from Roosevelt Hot Springs and the Tintic and Santa Rita the evolution of magmatic-hydrothermal systems based on microtextural
porphyry copper systems: Economic Geology, v. 79, p. 72–86. relationships, fluid inclusion petrography, and quartz solubility constraints:
Penniston-Dorland, S.C., 2001, Illumination of vein quartz textures in a por- Insights into the formation of the Yulong Cu-Mo porphyry deposit,
phyry copper ore deposit using scanned cathodoluminescence: Grasberg eastern Tibetan Plateau, China: Mineralium Deposita, doi: 10.1007/
igneous complex, Irian Jaya, Indonesia: American Mineralogist, v. 86, p. s00126-020-01003-6.
652–666. Titley, S.R., 1993, Characteristics of porphyry copper occurrence in the
Pollard, P.J., and Taylor, R.G., 2002, Paragenesis of the Grasberg Cu-Au American Southwest: Geological Association of Canada, Special Paper 40,
deposit, Irian Jaya, Indonesia: Results from logging section 13: Mineralium p. 433–464.
Deposita, v. 37, p. 117–136. Ulrich, T., Günther, D., and Heinrich, C.A., 1999, Gold concentrations
Pudack, C., Halter, W.E., Heinrich, C.A., and Pettke, T., 2009, Evolution of of magmatic brines and the metal budget of porphyry copper deposits:
magmatic vapor to gold-rich epithermal liquid: The porphyry to epithermal Nature, v. 399, p. 676–679.
transition at Nevados de Famatina, northwest Argentina: Economic Geol- Weis, P., Driesner, T., and Heinrich, C.A., 2012, Porphyry-copper ore shells
ogy, v. 104, p. 449–477. form at stable pressure-temperature fronts within dynamic fluid plumes:
Redmond, P.B., Einaudi, M.T., Inan, E.E., Landtwing, M.R., and Heinrich, Science, v. 338, p. 1613–1616.
C.A., 2004, Copper deposition by fluid cooling in intrusion-centered sys- Williams-Jones, A.E., and Heinrich, C.A., 2005, Vapor transport of met-
tems: New insights from the Bingham porphyry ore deposit, Utah: Geology, als and the formation of magmatic-hydrothermal ore deposits: Economic
v. 32, p. 217–220. Geology, v. 100, p. 1287–1312.
Reynolds, T.J., and Beane, R.E., 1985, Evolution of hydrothermal fluid char- Xiao, B., Qin, K., Li, G., Li, J., Xia, D., Chen, L., and Zhao, J., 2011, Highly
acteristics at the Santa Rita, New Mexico, porphyry copper deposit: Eco- oxidized magma and fluid evolution of Miocene Qulong giant porphyry
nomic Geology, v. 80, p. 1328–1347. Cu-Mo deposit, southern Tibet, China: Resource Geology, v. 62, p. 4–18.
Rinne, M.L., Cooke, D.R., Harris, A.C., Finn, D.J., Allen, C.M., Heizler,
M.T., and Creaser, R.A., 2018, Geology and geochronology of the Golpu
porphyry and Wafi epithermal deposit, Morobe province, Papua New
Guinea: Economic Geology, v. 113, p. 271–294.
Rose, A.W., and Baltosser, W.W., 1966, The porphyry copper deposit at Santa
Rita, New Mexico, in Titley, S.R., and Hicks, C.L., eds., Geology of the por-
phyry copper deposits, southwestern North America: Tucson, University
Arizona Press, p. 205–220.
Rusk, B., and Reed, M., 2002, Scanning electron microscope-cathodolumi-
nescence analysis of quartz reveals complex growth histories in veins from
the Butte porphyry copper deposit, Montana: Geology, v. 30, p. 727–730. Subaru Tsuruoka is an economic geologist
Rusk, B.G., Reed, M.H., and Dilles, J.H., 2008, Fluid inclusion evidence for with experience working on multiple ore deposit
magmatic-hydrothermal fluid evolution in the porphyry copper-molybde- types including porphyry copper, epithermal
num deposit at Butte, Montana: Economic Geology, v. 103, p. 307−334. gold, and sediment-hosted copper deposits. He
Sapiie, B., and Closs, M., 2013, Strike-slip faulting and veining in the Gras- completed his B.Sc. and M.Sc. degrees at Tokyo
berg giant porphyry Cu-Au deposit, Ertsberg (Gunung Bijih) mining dis- Institute of Technology and subsequently worked
trict, Papua, Indonesia: International Geology Review, v. 55, p. 1−42.
as a project geologist based in Tokyo. In 2017,
Seedorff, E., Dilles, J.H., Proffett, J.M., Jr., Einaudi, M.T., Zurcher, L.,
Stavast, W.J.A., Johnson, D.A., and Barton, M.D., 2005, Porphyry deposits: Subaru obtained his Ph.D. degree from Colorado School of Mines. He is
Characteristics and origin of hypogene features: Economic Geology 100th currently a postdoctoral research fellow based at iCRAG, the Irish Centre
Anniversary Volume, p. 251–298. for Research in Applied Geosciences at University College Dublin, Ireland.
Sheppard, S.M.F., Nielsen, R.L., and Taylor, H.P., Jr., 1971, Hydrogen and His postdoctoral research focuses on the stratigraphic and structural architec-
oxygen isotope ratios in minerals from porphyry copper deposits: Economic ture of sedimentary basins and their endowment of sediment-hosted copper
Geology, v. 66, p. 515–542. deposits.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/116/6/1267/5371704/4831_tsuruoka_et_al.pdf


by Raymond Rivera Cornejo

You might also like