You are on page 1of 12

©2020 Society of Economic Geologists, Inc.

Economic Geology, v. 115, no. 1, pp. 177–188

SCIENTIFIC COMMUNICATIONS

CLASSIFYING SKARNS AND QUANTIFYING METASOMATISM AT THE ANTAMINA DEPOSIT, PERU:


INSIGHTS FROM WHOLE-ROCK GEOCHEMISTRY

Stephanie A. Mrozek,1,† Zhaoshan Chang,1,2 Carl Spandler,1 Steve Windle,3,4 Cesar Raraz,3 and Alberto Paz3
1Economic Geology Research Center (EGRU), James Cook University, Townsville, Queensland 4811, Australia
2Department of Geology and Geological Engineering, Colorado School of Mines, Golden, Colorado 80401, USA
3Compañía Minera Antamina, Av. El Derby #055, Surco, Lima 15000, Peru
4Monazita SAC, Enrique Palacios 210, Dpto 803, Miraflores, Lima 15074, Peru

Abstract
At the Antamina deposit, Peru, accurate classification of exoskarns and endoskarns can be problematic when
textures are mottled. In this study, we use whole-rock geochemical compositions (62 elements) of 221 samples
to differentiate texturally similar endoskarns and exoskarns by comparing their compositions to least altered
precursors (wall rocks and intrusive rocks). We present a simple method for discriminating these skarn types
using immobile element bivariate plots. The most effective discriminators partition endoskarn and exoskarn
into distinct domains defined by the composition of each precursor; these include Al2O3 versus heavy rare earth
elements and some high field strength elements. Using these geochemical parameters, undifferentiated skarn
samples can be more reliably classified as endoskarn or exoskarn. The effectiveness of these element pairs is
attributed to their significantly different initial concentrations in wall rocks versus igneous precursors and their
immobility during skarn formation. While immobile elements can differentiate the skarns, mobile element
gains and losses (quantified using isocon analysis) provide insight on the bulk mineralogical and mass changes
that take place during skarn formation.

Introduction to unaltered precursors from Antamina with the aim of devel-


Skarn formation is a metasomatic process that can affect both oping a method to distinguish endoskarns from exoskarns.
wall rocks and igneous rocks, forming exoskarns and endos- We demonstrate that the geochemical classification of skarns
karns, respectively (Einaudi et al., 1981; Meinert et al., 2005). is more reliable than textural classification because mottle-­
Accurate classification of endoskarn and exoskarn has impor- textured skarns can be either endoskarn or exoskarn. Further-
tant implications for exploration and mining. Systematic map- more, we quantify the compositional changes that take place
ping of the rock types and mineralogical zoning patterns in a during skarn formation in order to gain insight on the influ-
skarn system can improve the accuracy of a deposit model and ence of immobile elements on our classification method.
reveal vectors to ore (Newberry et al., 1991; Meinert, 1997; At Antamina, Peru (Fig. 1), some endoskarns occur dis-
Meinert et al., 2005; Chang and Meinert, 2008b). Addition- tinctly as veins and patchy replacements in intrusive rocks
ally, the endoskarn-exoskarn contact can mark a change in ore (Fig. 2A-D), and some exoskarns (i.e., near the marble front)
grade, metal distribution, and rock hardness. Therefore, dis- clearly inherit banded textures from layering in the wall rocks
tinguishing these skarn types can inform mining and milling (Fig. 2M-P). In between, there are extensive zones of massive
processes to improve metal recovery. skarns, locally with mottled textures (Fig. 2F-L). With this
Field-based classification of skarns relies heavily on inher- geologic architecture, Antamina is an ideal location to exam-
ited precursor textures; as such, skarns with relict textures ine whether bulk-rock geochemistry alone can be used to dif-
are the easiest to classify as either endoskarn or exoskarn. ferentiate texturally similar endoskarns and exoskarns. The
Banded textures in exoskarn provide a clear link to a layered same data set is used to evaluate potential elemental gains and
metasedimentary precursor, while weakly altered endoskarns losses during skarn formation.
commonly display patches of residual igneous rock (e.g., at Geologic Setting
the Empire Zn-Cu skarn, Idaho; Chang and Meinert, 2008a).
Compared to endoskarns, exoskarns are less likely to retain Antamina is located approximately 270 km north of Lima,
primary textures, because carbonate-rich wall rocks are more Peru, at 9°46'S, 77°06'W in the central Peruvian Andes. It
reactive than igneous rocks in the presence of acidic magmatic- is the largest skarn deposit in the world, with a resource of
hydrothermal fluids. At Antamina, mottle-textured skarns are ~2,968 million tonnes (Mt) averaging 0.89% Cu, 0.77% Zn,
common, and although the precursor connection is not clearly 11 g/t Ag, and 0.02% Mo (Glencore, 2015). The deposit is
linked, they are typically classified as endoskarn because of a centred on a Miocene (10.95 ± 0.20–10.24 ± 0.23 Ma; zircon
subtle resemblance to porphyritic igneous rocks. In this study, U-Pb; Mrozek, 2018) multiphase porphyry system, herein
we compare the bulk-rock geochemical signatures of skarns referred to as the Antamina Porphyry Complex (Fig. 1). The
complex comprises several intrusions, including early- and
†Corresponding author: e-mail, stephanie.mrozek@my.jcu.edu.au intermineralization porphyries. All of the intrusive phases
ISSN 0361-0128; doi:10.5382/econgeo.4698; 12 p. Submitted: September 16, 2017 / Accepted: August 27, 2019
Digital appendices are available in the online Supplements section. 177

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
178 SCIENTIFIC COMMUNICATIONS

77°04’30”W 77°04’00”W 77°03’30”W 77°03’00”W

9°31’00”S
A

Peru

Antamina

9°31’30”S
9°32’00”S
limit of contact
metamorphism
t
ma rbl e fr on

9°32’30”S
77°2’W
B
ble Location of
ar Figure 1A

9°32’S
m

9°33’00”S
9°34’S

1 km 4 km 77°0’W

Host Rocks and Alteration Structures Samples


Antamina Porphyry Celendín Formation Anticline This study
Miocene

Cretaceous

Complex
Undifferentiated Jumasha Formation Syncline Paz et al. (2015)
endoskarn and exoskarn Santa, Carhuaz, Chimu Fms.
(undifferentiated, Inset B) Thrust fault
Fig. 1. Location and geology of the Antamina deposit, Peru. A. Location of samples from the Antamina mine area. Inset in
upper right corner shows location of study area in Peru. B. Near-mine sample locations and geology farther afield. Geology is
modified after Redwood (2004) and Escalante et al. (2010).

contain porphyry-style quartz stockworks and secondary distal Zn-Pb (Lipten and Smith, 2005). The principal ore
biotite alteration (Mrozek, 2018). Endoskarn alteration is minerals include chalcopyrite, sphalerite, and molybdenite,
observed along the margins of at least two early porphyry with localized bornite and galena (Love et al., 2000). Some
phases (Redwood, 2004; Chang et al., 2015; Mrozek, 2018). intermineralization porphyries cut across endoskarn and
Exoskarns surround the endoskarns, and all of the skarns are exoskarn, indicating a complex intrusive history (Mrozek et
mineralized, with the general trend of proximal Cu-Mo to al., 2015, 2017).

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
SCIENTIFIC COMMUNICATIONS 179

A A170-254.8 B A139-37.6 C A2589-818.3 D A206-143.5

E A1854-708.5 F A2589-805.4 G A2589-184.5 H A2084-1364.5

I A2589-426.3 J A136-169.0 K A2589-617.3 L A2589-559.0

M A2491-1804.0 N A2084-1455.1 O A2810-74.0 P A2589-1391.5

3 cm

Fig. 2. Representative endoskarns and exoskarns. A, B. Patchy endoskarn showing remnant porphyry texture and mineralogy,
including early secondary biotite alteration. C-E. Moderate pervasive endoskarn alteration; weak porphyry texture still appar-
ent. F, G. Strong endoskarn alteration resulting in a mottled texture. H-L. Strong exoskarn alteration. The mottled texture
of these exoskarns is difficult to distinguish from mottled endoskarn. M-P. Moderate-intensity exoskarn alteration showing
banded textures inherited from the wall rock; easily distinguished from weak-moderate endoskarns. Scale bar applies to all
images.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
180 SCIENTIFIC COMMUNICATIONS

Characteristics of the metasedimentary precursors siliciclastic component, with a more diverse detrital miner-
District-scale host rocks consist of a complexly deformed alogy noted in the Celendín Formation compared to the
marine-deltaic package of intercalated limestones, marls, Jumasha Formation (App. Table A1; Escalante, 2008). The
and calcareous shale/siltstone of the Cretaceous Jumasha and mid-Cretaceous Jumasha Formation (Fig. 3A, D, E) is a topo-
Celendín Formations (Fig. 3). These rocks have a variable graphically prominent, medium- to thick-bedded bioclastic

Jumasha Formation

Celendín Formation

B C

Fig. 3. Field characteristics of the Celendín and Jumasha Formations, within 3 km of the Antamina deposit. A. The break
in slope marks the approximate contact location between the prominent Jumasha and recessive Celendín Formations. B.
Contact metamorphosed Celendín Formation, displaying interbedded marly (dark-gray) and calcareous (light-brown) layers.
C. Thin-layered, dark-gray, marly Celendín limestone. D. Massive, light-gray Jumasha limestone with thin bioclastic layers.
E. Thick-bedded, cliff-forming Jumasha limestone.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
SCIENTIFIC COMMUNICATIONS 181

limestone with lesser amounts of marl, dolomite, and chert Major minerals include quartz (≤10%), K-feldspar (≤15%),
nodules (Benavides-Cáceres, 1956; Wilson, 1963; Escalante plagioclase (≤55%), and biotite (≤15%) with lesser amphi-
et al., 2010). Fresh and weathered surfaces are light gray, bole (≤3%); these occur both as phenocrysts and in ground-
likely reflecting lower organic carbon contents (compared to mass (Fig. 4). Accessory minerals include magnetite, apatite,
the dark-gray Celendín Formation). Unit thickness ranges titanite, and zircon. The porphyries are variably altered to
from 200 to 800 m, although local structural thickening is secondary biotite (i.e., potassic alteration) ± endoskarn.
observed around Antamina. The Upper Cretaceous Celendín Stockwork quartz ± sulfide vein densities vary from rare to
Formation (Fig. 3A-C) is a fine-grained, moderately recessive abundant.
unit dominated by thin intercalations of carbonate and calcar-
eous shale/siltstone (Benavides-Cáceres, 1956; Wilson, 1963; Characteristics of the skarns
Escalante et al., 2010). Thickness varies by location, from Andraditic garnet (up to Ad98; Mrozek, 2018) is the domi-
225 m at the type locality to 115 m in other parts of the region nant skarn mineral, indicating oxidized conditions of forma-
(Wilson, 1963). Fresh rock surfaces are medium to dark gray, tion, although clinopyroxene-rich layers and lenses are locally
reflecting an overall high organic carbon content. Weathered common in exoskarn. Endoskarn alteration occurs along the
surfaces display a creamy beige patina. margins of early porphyry phases in contact with exoskarn.
Within a few hundred meters of the Antamina Porphyry Near the endoskarn-exoskarn contact the skarns display mas-
Complex, the Celendín and Jumasha Formations have been sive or mottled textures, typically rendering the two types
contact metamorphosed to marble and hornfels (Fig. 1); these indistinguishable. With increasing distance from this contact,
are the inferred precursors to exoskarn at Antamina. Struc- the skarns bear an increasing textural resemblance to their
tural thickening, metamorphic recrystallization, and skarn precursors. From the endoskarn-exoskarn contact toward the
development make it difficult to trace the stratigraphic con- porphyry center, endoskarn intensity decreases into second-
tact between the two formations; however, previous research- ary biotite alteration with a distinct porphyry texture (Fig. 2A,
ers generally agree that the deposit is hosted in the transition B). Toward the marble front, exoskarns reveal banded tex-
zone between the Celendín and Jumasha Formations (Love tures inherited from metasedimentary precursors (Fig. 2G-I).
et al., 2004; Lipten and Smith, 2005; Escalante et al., 2010). Overall, the skarns display a well-developed garnet color
The marble and hornfels contain a significant component of zonation outward from the porphyry center to the marble
detrital minerals (i.e., quartz, feldspar, rutile, titanite, and front, which generally transitions from proximal pink-red gar-
apatite), which are inherited from the sedimentary precursors net to distal brown-green garnet (Lipten and Smith, 2005). All
(App. Table A1; Escalante, 2008). Some typical skarn miner- of the skarns display variable intensities of retrograde altera-
als (i.e., diopsidic pyroxene ≤53%; epidote ≤18%), as well tion manifested as epidote ± chlorite ± amphibole ± calcite ±
as traces of sphalerite and chalcopyrite, occur in fluid escape quartz. Copper-Zn-Pb mineralization is associated with retro-
structures outside of the main skarn orebody (App. Table A1; grade alteration (Love et al., 2000; Mrozek, 2018).
Escalante, 2008); these minerals are hydrothermal in origin
and extend the geochemical footprint of the skarn beyond the Methods
marble front.
Samples
Characteristics of the igneous precursors A total of 221 rock samples were collected and analyzed for
The Antamina porphyries display a narrow range of composi- this study (Fig. 1; App. Table A2). Massive and mottled skarn
tions from quartz monzonite to quartz monzodiorite, as well samples (n = 63), intrusion samples (n = 43), marble (n =
as typical arc magma geochemical signatures (Mrozek, 2018). 7), and limestone (n = 2) were selected from drill core and

A A1854-338.2 AB A1391-548.3
A1854-338.2 C A2589-203.0

Fig. 4. Typical porphyries from Antamina. A-C. Representative samples displaying the characteristic quartz monzonite min-
eralogy and porphyry texture. Rocks such as these are the precursors to endoskarn at Antamina.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
182 SCIENTIFIC COMMUNICATIONS

pit highwalls. Care was taken to ensure that each intrusion, and Winchester, 1978) and its capacity to differentiate the wall
marble, and limestone sample was representative of the unit rocks from igneous rocks at Antamina, Al2O3 was selected as
or interval and free of veins, xenoliths, contacts, and other the control variable in relating the skarns to their precursors.
heterogeneities. Sample sizes ranged from 5 to 15 cm of NQ Bivariate plots, comparing Al2O3 versus all other elements,
drill core or fist-sized hand samples from field outcrops and were constructed to evaluate geochemical relationships
open-pit highwalls. Least altered (i.e., loss on ignition [LOI] between the skarns and their precursors. The plots were visu-
≤1.8 wt %, total S ≤0.8 wt %; App. Table A2), early porphyry ally assessed for skarn cluster separation and compositional
intrusion samples were collected exclusively from drill core correlation (i.e., clustering and overlap) with each precursor.
and mine workings; these samples are the precursors to endos- The discriminatory quality (i.e., weak, moderate, or strong) of
karn. An additional 106 samples of limestone ± marl were col- each element combination was visually assessed based on the
lected by Antamina exploration geologists from within a 6-km degree of skarn overlap with each precursor cluster and the
radius of the mine, mostly outside of the contact metamorphic overall partitioning of data clusters. The strongest element
halo, using the same selection criteria defined for the samples combinations allow for unequivocal classification of endos-
described above (Paz et al., 2015; Fig. 1). Metasedimentary karn and exoskarn. The reclassified skarns and precursors are
samples (i.e., variably metamorphosed limestone ± marl) the input data for isocon analysis.
were classified in the field as either Jumasha or Celendín For-
mation based on stratigraphic criteria used in regional explo- Isocon analysis
ration; these samples represent the closest approximation to a Mass changes and element mobility during skarn formation
least altered precursor to exoskarn. were quantified using isocon analysis (Grant, 1986). Our anal-
ysis was conducted using the complete geochemical data set
Whole-rock analysis for precursors and reclassified skarns, as described above. The
All samples collected for this study were analyzed at ALS isocon method uses a modified version of Gresens’s (1967)
Chemex in Lima, Peru, using a combination of techniques equation for calculating mass balance. Using the Gresens
including X-ray fluorescence (XRF), inductively coupled method, the chemical components of a least altered rock
plasma-atomic emission spectroscopy (ICP-AES), induc- (precursor) are plotted against those same components of an
tively coupled plasma-mass spectrometry (ICP-MS), atomic altered equivalent (in this case, the corresponding skarns) in
absorption (AA), and LECO analysis for 62 elements (includ- X-Y space. A common conversion factor is used to scale the
ing major and trace elements plus total S and total C) and LOI data (App. Tables A5, A6). Elements deemed to be immo-
(App. Tables A2, A3). All analyses were conducted in ISO bile define a best-fit linear array anchored through the origin;
standard (9001 and 17025) certified laboratories. Controls on this line is called the isocon. In this study, elements display-
accuracy and precision of the data were monitored via analy- ing gains or losses within ±10% of the mean analytical value
sis of certified reference materials, internal standards, sample are considered immobile (App. Tables A5, A6). The isocon
preparation and analytical blanks, and duplicate samples in slope (m) quantifies the total mass change in the system; m
regular intervals during sample analysis batches. < 1 corresponds to mass loss, m = 1 means no mass change,
Nine samples of limestone (n = 2) and marble (n = 7) were and m > 1 corresponds to mass gain. In this study we used
analyzed for total organic carbon at the James Cook University the average value for each element, by rock type, to minimize
Advanced Analytical Centre in Cairns, Australia. The analytical the effects of compositional heterogeneity as in Oliver et al.
procedure is described in Wurster et al. (2012). The organic C (2004). Detailed explanations of the mass balance calculation
content ranges from 0.09 to 3.63 wt % (App. Table A4). procedure are presented in Grant (1986), Leitch and Lentz
(1994), and Trepanier et al. (2016).
Data preparation and plotting
The data set was screened for missing entries and values out- Results
side of the limit of detection (LOD) of each analyte. Four Cu
assays exceeded the upper LOD (10,000 ppm); these were Geochemical composition of the host rocks
replaced with a value of 10,000.1 ppm. Values below the lower The compositions of the metasedimentary host rocks range
LOD for any element were replaced by a value equal to 50% from essentially pure limestone in some samples of the Juma-
of the LOD. Silver, Cd, Ta, and Li were not considered fur- sha Formation (~13 wt % C, ~55 wt % CaO; Fig. 5) to marl
ther, as >30% of analyses for these elements were below the or calcareous mudstone in the Celendín Formation (up to
LOD (see Martín-Fernández et al., 2012). Barium, Sr, and 46.2 wt % SiO2, 51.8 wt % CaO, and 12.6 wt % Al2O3; Fig.
Cr were analyzed twice, as major element oxides (i.e., BaO, 5). In general, the Celendín Formation contains more SiO2
SrO, Cr2O3, in wt %) and as trace elements (i.e., Ba, Sr, Cr, in and Al2O3 and less CaO than the Jumasha Formation, which
ppm); only the trace element values were used in this study. is consistent with the field descriptions of these rock types.
All plots were constructed using the reduced data set and Some samples from the Jumasha Formation contain up to
ioGAS™ software. ~3.4 wt % MgO, indicating that they are dolomitic (Fig. 5).
The decrease in total C content in the wall rocks correlates
Framework for relating skarns to precursors with a decrease in CaO content and an increase in SiO2 and
Among the 62 individual elements examined, only Al2O3, Al2O3 (Fig. 5). Because the organic carbon content is low, the
SiO2, CaO, Ga, total C, and LOI can differentiate wall rocks total C is dominated by carbon from carbonates (App. Table
from igneous precursors (App. Fig. A1). Because of its rela- A4). Therefore, the observed trend indicates that the increase
tive immobility under hydrothermal conditions (i.e., Floyd in silicate minerals is compensated by a decrease in carbonate

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
SCIENTIFIC COMMUNICATIONS 183

Celendín Fm.
Jumasha Fm.
Intrusion
Total C (wt. %)

CaO (wt. %)
A B

C D
Al2O3 (wt. %)

TiO2 (wt. %)
E F
MgO (wt. %)

La (ppm)

G H
Lu (ppm)

Y (ppm)

I J
LOI (wt. %)
Ga (ppm)

SiO2 (wt. %) SiO2 (wt. %)

Fig. 5. Composition of intrusive and metasedimentary precursors. Selected elements versus SiO2 are presented; see Appen-
dix Figure A2 for additional element plots and Appendix Table A2 for the complete data set.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
184 SCIENTIFIC COMMUNICATIONS

minerals (Figs. 4B, 5A). This is consistent with the petro- in both endoskarn and exoskarn. LOI increases by more than
graphic findings of Escalante (2008) in that the wall rocks 50% in endoskarn but decreases by more than 50% in exos-
contain varying proportions of calcite and quartz + alumi- karn. SiO2 decreases in endoskarn (–33.9%) and increases in
nosilicate minerals (App. Table A1). Given the negative cor- exoskarn (168%; Fig. 7; App. Tables A5, A6). CaO increases in
relation between total C and SiO2 (Fig. 5A) and the positive endoskarn (767%) and decreases in exoskarn (–52.4%; Fig. 7;
correlation between Al2O3 and SiO2, we use Al2O3 and SiO2 as App. Tables A5, A6). The light rare earth elements (LREEs)
proxies for the clastic content of the wall rocks. display evidence for mobility in both skarn types (Fig. 7; App.
Total Fe expressed as Fe2O3, K2O, Na2O, and most litho- Tables A5, A6).
phile trace elements (e.g., rare earth elements [REEs],
Y, TiO2) have positive linear correlations with SiO2 in both Discussion
Jumasha and Celendín samples (Fig. 5; App. Fig. A2), indi-
cating that these elements are also related to detrital miner- Element mobility during skarn formation
als in the metasedimentary host rocks. The MnO content is The significantly different initial concentrations of immobile
low (≤0.16 wt %), with slightly higher concentrations in rocks elements in the igneous and metasedimentary precursors
containing >10 wt % SiO2 (App. Fig. A2). Barium concentra- contribute to the effectiveness of skarn discrimination using
tions range from 17.1 to 429 ppm and display a weak positive our method. The immobile element geochemical fingerprint
trend that is correlative with increasing siliciclastic contents of the precursor is preserved through contact metamorphism
in the wall rocks (App. Fig. A2). The Sr content ranges from and hydrothermal alteration and subsequently inherited by
~250 to 2,100 ppm, with some carbonate-rich rocks contain- the skarns.
ing the highest Sr concentrations (App. Fig. A2). HFSEs and Al are among the most immobile elements
Compared to the wall rocks, the intrusive rocks contain identified in this study and under a variety of geologic con-
significantly higher SiO2 (64.6–74.3 vs. 0.70–48.6 wt %), and ditions (Floyd and Winchester, 1978; Finlow-Bates and
significantly lower total C (0.01–0.35 vs. 3.15–13.1 wt %) and Stumpfl, 1981; Whitney and Olmsted, 1998; Jiang et al., 2005;
CaO (0.70–5.31 vs. 15.7–55.4 wt %; Fig. 5). In general, the Lentz, 2005; Ranjbar et al., 2016). Because of their low solu-
intrusive rocks contain more K2O, Na2O, and Ba than the wall bility, Al contents of the skarns are derived mainly from the
rocks (Fig. 5; App. Fig. A2). The intrusive rocks have similar precursor rocks. Although the solubility and mobility of these
TiO2 and Zr concentrations and lower Ni, heavy rare earth elements can be enhanced in strongly acidic fluids, their solu-
elements (HREEs), Y, and V contents compared with the bilities drop dramatically with increasing pH. For example, at
SiO2-rich sedimentary rocks (>30 wt %, Celendín Formation; 250°C total dissolved Al in fluids in equilibrium with alunite +
Fig. 5; App. Fig. A2). The intrusive rocks and wall rocks have kaolinite + quartz is ~2,000 ppm at pH = 2 but drops to only
similar ranges of Fe2O3 and MnO content (App. Fig. A2), ~1 ppm when the pH increases to 4 (Stoffregen, 1987). In
although some intrusive rock samples extend toward higher carbonate-rich environments (where most skarns form), fluid
MnO concentrations (up to 0.19 wt % MnO). acidity ranges from weak to neutral due to buffering by excess
carbonate; such fluids cannot mobilize HFSEs or Al2O3.
Geochemical discrimination of the skarns In both the igneous and metasedimentary precursor rocks,
Element pairs involving Al2O3 versus HREEs and some high immobile elements are hosted in feldspars and relatively alter-
field strength elements (HFSEs; i.e., TiO2, Sc, Y, and Nb) ation resistant accessory minerals, such as magnetite, zircon,
clearly discriminate the skarns into two domains: one clus- apatite, rutile, and titanite (Deer et al., 1992; Escalante, 2008;
tering with the metasedimentary wall rocks (i.e., exoskarns) Hammerli et al., 2016; Mao et al., 2016; Mrozek, 2018; App.
and the other with the igneous rocks (i.e., endoskarns; Fig. Table A1). During skarn formation, the bulk mineralogy of the
6). The strongest discriminators include Al2O3 versus HREEs host rocks changes significantly; however, the low solubility of
(i.e., Dy, Lu), Y, and TiO2 (Fig. 6). Moderate discriminators these immobile elements allows them to be reincorporated in
include Al2O3 versus Sc, Nb, La, and Th (Fig. 6). All other skarn minerals as they form (Huggins et al., 1977; Nicolescu
element combinations yield weak to moderate relationships et al., 1998, Gaspar et al., 2008). Consequently, immobile ele-
and are not considered to be useful discriminators of skarn ment signatures of the Antamina skarns are preserved through
type at Antamina. contact metamorphism and skarn formation; evidence for this
isochemical behavior is presented in Figure 7.
Mass changes and element mobility during skarn formation In both endoskarn and exoskarn, increases in Fe2O3, MnO,
Some elements identified in Figure 6 display a case for mod- and MgO are related to the formation of garnet and clinopy-
erate to strong immobility (i.e., Al2O3, Y, HREEs, Sc, Nb, and roxene during prograde skarn alteration; these components
Th); these elements define the isocons and are used to quantify are likely to be derived from magmatic-hydrothermal fluids.
mass changes and element mobility during skarn formation SiO2 (168%) is added to exoskarn during prograde skarn for-
(Fig. 7; App. Tables A5, A6). Endoskarn formation (Fig. 7A; mation, but its source may be partly from SiO2 stripping from
App. Table A5) involves very little mass change as indicated by the intrusions during endoskarn formation (as indicated by a
the isocon slope of 1.04 (4% mass loss). In contrast, exoskarn 34% loss of SiO2; Fig. 7A; App. Tables A5, A6). Base metals
formation results in an overall mass loss of approximately 28% (Cu, Mo, Zn, Pb) and U increase significantly in both endos-
(Fig. 7B; App. Table A6). In both endoskarn and exoskarn, sig- karn and exoskarn due to retrograde alteration (Fig. 7; App.
nificant increases are observed in metals (Cu, Mo, Zn, Pb, U) Tables A5, A6). Volatile (LOI) increases (323%) in endoskarn
and some major skarn mineral components (i.e., Fe2O3, MnO, may be related to the formation of hydrous or CO2-bearing
and MgO). Alkalis (K2O, Na2O) decrease by more than 50% retrograde minerals (chlorite, amphibole, calcite), while LOI

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
SCIENTIFIC COMMUNICATIONS 185

A B

TiO2 (wt.%)

Y (ppm)
Skarns (undifferentiated)
Celendín Fm.
Jumasha Fm.
Intrusion

C D

Dy (ppm)
Lu (ppm)

E F
Sc (ppm)

Nb (ppm)

G H
Th (ppm)
La (ppm)

I J
Nd (ppm)
Zr (ppm)

Al2O3 (wt. %) Al2O3 (wt. %)

Fig. 6. A-J. Skarn discrimination plots based on precursor clustering and overlap for selected combinations of Al2O3 versus
REEs and HFSEs. The strongest discriminators are Al2O3 versus HREEs (Dy, Lu), Y, TiO2, and Sc, which display complete
overlap with precursor clusters and no overlap between skarn clusters. Combinations of Al2O3 versus Nb and Th are moder-
ate discriminators based on less tightly constrained sample clusters for the skarns compared to the precursors. While Al2O3
versus La, Nd, and Zr display strong precursor clusters, the skarn sample clusters are too dispersed for reliable discrimination.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
186 SCIENTIFIC COMMUNICATIONS

4% MASS LOSS 28% MASS LOSS


10.0 10.0
A B GE

AIN
LOI CaO AN
9.0 9.0 CH

SG
Ce

Metasedimentary wall rocks (n = 108)


SS

AIN
S
MA

MA

SG
8.0 NO

%
8.0 GE

S
50
AN

MA
LOSSES LOSSES
CH La

%
S

50
7.0 Si O₂ S N 7.0
MA CO Na ₂O
SO
Intrusions (n = 43)

NO I
6.0
Na ₂O 6.0 Ti O₂
MgO
K₂O
5.0 LOS
S 5.0
ASS Eu S
50%M LOS
4.0 4.0 ON A SS
OC Hf 50 %M
K₂O IS
3.0 3.0 Zr
GAINS
Fe₂O₃ P₂O₅
2.0 2.0 Si O₂
LOI MgO
Zn GAINS
1.0 Zn Sn Cu 1.0
CaO Mo MnO Fe₂O₃
MnO Sn Cu Mo
0.0 0.0
0.0 2.0 4.0 6.0 8.0 10.0 0.0 2.0 4.0 6.0 8.0 10.0
Endoskarn (n = 20) Exoskarn (n = 43)
Fig. 7. Isocon plots showing mass changes during skarn formation. Skarns are plotted along the x-axis, and least altered equiv-
alents (precursors) are plotted along the y-axis. The slope (m) of the isocon indicates the total mass change for the system; m
> 1 indicates an overall mass increase, and m < 1 indicates an overall mass decrease. The red line, m = 1, is a reference line
indicating constant mass (no change). The total mass change is indicated in the upper right corner of each plot, expressed as a
percentage. Elements plotting along the isocon (black circles along the dashed black line) are immobile; see Appendix Tables
A5 and A6 for the complete list of immobile elements. Elements plotting below the isocon have been gained during skarn for-
mation, while elements above the isocon have been lost during skarn formation. A. Element mobility and mass changes dur-
ing endoskarn formation (App. Table A5). B. Element mobility and mass changes during exoskarn formation (App. Table A6).

decreases in exoskarn (–97%) are largely related to CO2 lib- geochemical evidence indicates that exoskarns may also dis-
eration from marble during prograde skarn formation. play mottled textures. Impure carbonate rocks, like those at
The only elements lost in significant quantities (>38%) from Antamina, which contain siliciclastic layers or patches, may
both skarn types are the alkali elements (K2O, Na2O, Ba, and produce a mottled texture through a process of repetitive dis-
Rb; Fig. 7; App. Tables A5, A6). In both cases, the most com- solution, fragmentation, and replacement. When the rate of
mon and abundant skarn minerals do not incorporate these carbonate dissolution exceeds the rate of calc-silicate min-
elements into their crystal lattice. Some of this element loss eral precipitation, void space is created; this causes the more
may be related to alkali gain in the intrusive samples during resistant siliciclastic layers to become unsupported and col-
retrograde alteration (i.e., sericite), but significant amounts of lapse into fragments. Metasomatic replacement of these frag-
these elements were likely transported to distal parts of the mented rocks may produce exoskarns with mottled textures.
system via hydrothermal fluid escape structures (i.e., Meinert For example, the samples shown in Figure 2H-L display
et al., 2005). At Antamina we can only speculate that these mottled textures, but their whole-rock geochemistry indicates
alkalis were transported to overlying rock packages that have that they are exoskarns (Fig. 8). In contrast, the samples in
now been lost to erosion. Nevertheless, if alkali element loss Figure 2D-G are endoskarns based on their geochemical sig-
is a common feature of skarn formation, then upward flow and nature (Fig. 8).
reaction of these fluids with wall rocks may provide some of
the alkalis for the alteration assemblages that are characteris- Applicability in other skarns
tic of porphyry-epithermal environments. Elements in these This approach to skarn classification is particularly well suited
alteration assemblages are primarily considered to be magma for deposit-scale studies where the geologic framework has
sourced; however, some contribution from skarn alteration of been established and a range of rock types (from least altered
wall rock is consistent with Pb isotope evidence supporting to strongly altered) are available for analysis. Applying our
a wall-rock component to porphyry alteration fluids (Cooke procedures to other skarn localities, we suggest that wall
et al., 2014). If significant quantities of alkali elements are rocks, igneous rocks, and skarns should be characterized using
sourced from wall-rock alteration, recognition of this process the same analytical method. The precursor compositions are
would relax the requirements for very large magma volumes necessary to provide a geochemical baseline for comparison
as the primary source of alkali elements in porphyry alteration with skarns of unknown affinity. The success of this method
envelopes (Cathles and Shannon, 2007). in other localities will depend on the specific compositions
of wall rocks and igneous precursors. On the condition that
Formation of mottled texture in exoskarn precursor compositions are distinctly different in their HFSE
At Antamina, mottled skarn textures were interpreted to be and Al2O3 compositions, it may be possible to distinguish
inherited from a porphyritic igneous precursor; however, endoskarn and exoskarn just as we have done in this study.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
SCIENTIFIC COMMUNICATIONS 187

30 Reclassified Skarns
Endoskarn
Exoskarn A2589-426.3
25
Sample image
in Figure 2
A2589-617.3
20 Precursors
Celendín Fm.
Y (ppm)

Jumasha Fm. A2589-559.0 A1854-708.5


15 Intrusion
A2589-805.4

10

A136-169.0
5 A206-143.5
A2589-184.5

A2084-1364.5

0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0


Al2O3 (wt. %)
Fig. 8. Skarns reclassified using the Al2O3 versus Y discrimination plot. Selected mottle-textured skarns (blue symbols) are
presented in Figure 2, as indicated.

This method may not work in localities with only one type of calculations. Andrew Norton provided invaluable technical
precursor (i.e., only igneous rocks or only metasedimentary support. Rob Holm and Carlos Jimenez provided constructive
rocks; see Chang et al., 2018) or where precursor composi- reviews on the manuscript. This manuscript benefited signifi-
tions are similar. cantly from the comments of Lucie Mathieu, associate editor
Andrew Tomkins, and Editor Larry Meinert.
Conclusion
Using Antamina as a case study, we demonstrate a method REFERENCES
for comparing the whole-rock composition of skarns and their Benavides-Cáceres, V.E., 1956, Cretaceous system in northern Peru: Bulletin
precursors as a tool for discriminating endoskarns from exos- of the American Museum of Natural History, v. 108, p. 353–494.
Cathles, L.M., and Shannon, R., 2007, How potassium silicate alteration sug-
karns, especially where skarns are massive or where primary gests the formation of porphyry ore deposits begins with the nearly explo-
textures have been completely destroyed by hydrothermal sive but barren expulsion of large volumes of magmatic water: Earth and
alteration. The most effective discriminators at Antamina Planetary Science Letters, v. 262, p. 92–108.
are element pairs of Al2O3 versus HREEs and Y, followed by Chang, Z., and Meinert, L.D., 2008a, The Empire Cu-Zn mine, Idaho:
Al2O3 versus Sc, Zr, Nb, Th, TiO2, and LREEs. Skarn forma- Exploration implications of unusual skarn features related to high fluorine
activity: Economic Geology, v. 103, p. 909–938.
tion involves significant gains and losses of most elements, ——2008b, Zonation in skarns: Complexities and controlling factors: PAC-
therefore only immobile elements are useful for fingerprint- RIM Congress 2008, Australasian Institute of Mining and Metallurgy
ing and discriminating skarn type. The effectiveness of these (AusIMM), Gold Coast, Queensland, Australia, November 24–26, 2008,
immobile elements is enhanced when there are significant Proceedings, p. 303–306.
differences in geochemical signature between wall rocks and Chang, Z., Mrozek, S., Meinert, L.D., and Windle, S., 2015, Skarn-porphyry
transition: An example from the Antamina skarn, Peru: PACRIM 2015 Con-
igneous rocks. Mobile element behavior can explain the bulk gress, Australasian Institute of Mining and Metallurgy (AusIMM), Hong
mineralogical changes that occur during skarn formation. Kong, China, March 18–21, 2015, Proceedings, p. 409–413.
In particular, the mobility of alkali elements may contribute Chang, Z., Meinert, L., Lawrence, D., Mrozek, S., and Zhang, L., 2018,
alkali elements to porphyry and epithermal-style alteration Skarns replacing igneous rocks: Quadrennial International Association on
assemblages observed in higher levels of the system and in the Genesis of Ore Deposits (IAGOD) Symposium, 15th, Salta, Argentina,
August 28–31, 2018, Proceedings, p. 207–208.
zones peripheral to the main skarn orebody. Mottled texture Cooke, D.R., Hollings, P., Wilkinson, J.J., and Tosdal, R.M., 2014, Geochem-
is not a reliable feature to distinguish massive endoskarns and istry of porphyry deposits, in Holland, H.D., and Turekian, K.K., eds., Trea-
exoskarns, as it can be a texture common to both skarn types. tise on geochemistry, 2nd ed., v. 13: Elsevier, p. 357–381.
Deer, W.A., Howie, R.A., and Zussman, J., 1992, An introduction to the rock-
Acknowledgments forming minerals: Essex, United Kingdom, Pearson Education Ltd., 696 p.
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits: Eco-
This study was funded by Compañía Minera Antamina (ana- nomic Geology 75th Anniversary Volume, p. 317–391.
lytical funds and in-kind support) and a Society of Economic Escalante, A., Dipple, G.M., Barker, S.L.L., and Tosdal, R., 2010, Defining
Geologists Graduate Student Research Grant awarded to trace-element alteration halos to skarn deposits hosted in heterogeneous
Stephanie Mrozek. Christian Mendoza, David Paredes, Fred- carbonate rocks: Case study from the Cu-Zn Antamina skarn deposit, Peru:
erick Sanchez, and the Antamina exploration crew provided Journal of Geochemical Exploration, v. 105, p. 117–136.
Escalante, A.D., 2008, Patterns of distal alteration zonation around Antamina
limestone samples and access to the whole-rock data set. Cu-Zn skarn and Uchucchacua Ag base metal vein deposits, Peru: Min-
James Grant helpfully answered questions related to isocon eralogical, chemical, and isotopic evidence for fluid composition, and

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user
188 SCIENTIFIC COMMUNICATIONS

infiltration, and implications for mineral exploration: Ph.D. thesis, Vancou- Newberry, R.J., Einaudi, M.T., and Eastman, H.S., 1991, Zoning and genesis
ver, British Columbia, The University of British Columbia, 817 p. of the Darwin Pb-Zn-Ag skarn deposit, California; a reinterpretation based
Finlow-Bates, T., and Stumpfl, E.F., 1981, The behaviour of so-called immo- on new data: Economic Geology, v. 86, p. 960–982.
bile elements in hydrothermally altered rocks associated with volcanogenic Nicolescu, S., Cornell, D.H., Södervall, U., and Odelius, H., 1998, Secondary
submarine-exhalative ore deposits: Mineralium Deposita, v. 16, p. 319–328. ion mass spectrometry analysis of rare earth elements in grandite garnet
Floyd, P.A., and Winchester, J.A., 1978, Identification and discrimination and other skarn related silicates: European Journal of Mineralogy, v. 10, p.
of altered and metamorphosed volcanic rocks using immobile elements: 251–259.
Chemical Geology, v. 21, p. 291–306. Oliver, N.H.S, Cleverley, J.S., Mark, G., Pollard, P.J., Fu, B., Marshall, L.J.,
Gaspar, M., Knaack, C., Meinert, L.D., and Moretti, R., 2008, REE in skarn Rubenach, M.J., Williams, P.J., and Baker, T., 2004, Modeling the role of
systems: A LA-ICP-MS study of garnets from the Crown Jewel gold deposit: sodic alteration in the genesis of iron oxide-copper-gold deposits, eastern
Geochimica et Cosmochimica Acta, v. 72, p. 185–205. Mount Isa block, Australia: Economic Geology, v. 99, p. 1145–1176.
Glencore, 2015, Annual report: www.glencore.com/assets/investors/doc/ Paz, A., Raraz, C., Windle, S., Paredes, D., and Mendoza, C., 2015, Caracter-
reports_and_results/2015/GLEN-2015-Annual-Report.pdf. izacion litogeoquimica de las formaciones Jumasha y Celendín con elemen-
Grant, J.A., 1986, The isocon diagram; a simple solution to Gresens’ equation tos mayors y elementos de traza, relacionados con mineralizacion de Cu-Zn
for metasomatic alteration: Economic Geology, v. 81, p. 1976–1982. cercanos al yacimiento minero Antamina [ext. abs.]: International Congress
Gresens, R.L., 1967, Composition-volume relationships of metasomatism: of Prospectors and Explorers (ProExplo), 9th, Lima, Peru, 2015, Extended
Chemical Geology, v. 2, p. 47–55. Abstract, p. 43.
Hammerli, J., Spandler, C, and Oliver, N.H.S., 2016, Element redistribu- Ranjbar, S., Tabatabaei-Manesh, S.M., Mackizadeh, M.A., Tabatabaei, S.H.,
tion and mobility during upper crustal metamorphism of metasedimentary and Parfenova, O.V., 2016, Geochemistry of major and rare earth elements
rocks: An example from the eastern Mount Lofty Ranges, South Australia: in garnet of the Kal-e Kafi skarn, Anarak area, central Iran: Constraints on
Contributions to Mineralogy and Petrology, v. 171, p. 1–21. processes in a hydrothermal system: Geochemistry International, v. 54, p.
Huggins, F.E., Virgo, D., and Huchenholz, H.G., 1977, Titanium-containing 423–438.
silicate garnets. II. The crystal chemistry of melanites and schorlomites: Redwood, S.D., 2004, Geology and development history of the Antamina
American Mineralogist, v. 62, p. 646–665. copper-zinc skarn deposit, Peru: Society of Economic Geologists, Special
Jiang, S.-Y., Wang, R.-C., Xu, X.-S., and Zhao, K.-D., 2005, Mobility of high Publication 11, p. 259–278.
field strength elements (HFSE) in magmatic-, metamorphic-, and subma- Stoffregen, R.E., 1987, Genesis of acid-sulfate alteration and Au-Cu-Ag
rine-hydrothermal systems: Physics and Chemistry of the Earth, v. 30, p. mineralization at Summitville, Colorado: Economic Geology, v. 82, p.
1020–1029. 1575–1591.
Leitch, C.H.B., and Lentz, D.R., 1994, The Gresens approach to mass bal- Trepanier, S., Mathieu, L., Daigneault, R., and Faure, S., 2016, Precursors
ance constraints of alteration systems: Methods, pitfalls, examples: Geologi- predicted by artificial neural networks for mass balance calculations: Quan-
cal Association of Canada, Short Course Notes, v. 11, p. 161–192. tifying hydrothermal alteration in volcanic rocks: Computers & Geosci-
Lentz, D.R., 2005, Mass-balance analysis of mineralized skarn systems: ences, v. 89, p. 32–43.
Implications for replacement processes, carbonate mobility, and perme- Whitney, P.R., and Olmsted, J.F., 1998, Rare earth element metasomatism in
ability evolution: Society for Geology Applied to Mineral Desposits (SGA) hydrothermal systems: The Willsboro-Lewis wollastonite ores, New York,
Biennial Meeting, 8th, Beijing, 2005, Proceedings, p. 421–424. USA: Geochimica et Cosmochimica Acta, v. 62, p. 2965–2977.
Lipten, E.J., and Smith, S.W., 2005, The geology of the Antamina copper-zinc Wilson, J.J., 1963, Cretaceous stratigraphy of central Andes of Peru: Ameri-
deposit, Peru, South America, in Porter T.M., ed., Super porphyry copper can Association of Petroleum Geologists Bulletin, v. 47, no. 1, p. 1–34.
and gold deposits: A global perspective, v. 1: Adelaide, PGC Publishing, p. Wurster, C.M., Robertson, J., Westcott, D.A., Dryden, B., Zazzo, A., and
189–204. Bird, M.I., 2012, Utilization of sugarcane habitat by feral pig (sus scrofa) in
Love, D.A., Clark, A.H., and Schwarz, F.P., 2000, The Antamina deposit, northern tropical Queensland: Evidence from the stable isotope composi-
Ancash, Peru: Anatomy and petrology of a giant copper skarn [abs.]: Geo- tion of hair: PLoS ONE, v. 7, e43538.
logical Society of America Abstracts with Programs, v. 32, no. 7, p. A137.
Love, D.A., Clark, A.H., and Glover, J.K., 2004, The lithologic, stratigraphic,
and structural setting of the giant Antamina copper-zinc skarn deposit,
Ancash, Peru: Economic Geology, v. 99, p. 887–916.
Martin-Fernandez, J.A., Hron, K., Templ, M., Filzmoser, P., and Palarea-
Albaladejo, J., 2012, Model-based replacement of rounded zeros in compo-
sitional data: Classical and robust approaches: Computational Statistics and
Data Analysis, v. 56, p. 2688–2704.
Mao, M., Rukhlov, A.S., Rowins, S.M., Spence, J., and Coogan, L.A., 2016,
Apatite trace element compositions: A robust new tool for mineral explora-
tion: Economic Geology, v. 111, p. 1187–1222.
Meinert, L.D., 1997, Application of skarn deposit zonation models to mineral
exploration: Exploration and Mining Geology, v. 6, p. 185–208.
Meinert, L.D., Dipple, G.M., and Nicolescu, S., 2005, World skarn deposits: Stephanie Mrozek is an adjunct research asso-
Economic Geology 100th Anniversary Volume, p. 299–336. ciate at James Cook University (Townsville, Aus-
Mrozek, S.A., 2018, The giant Antamina deposit, Peru: Intrusive sequence, tralia) and a resource geologist at Northern Star
skarn formation, and mineralisation: Ph.D. thesis, Townsville, Australia, Resources (Pogo Mine, Alaska). She received
James Cook University, 191 p. her Ph.D. degree in 2018 at James Cook Univer-
Mrozek, S.A., Chang, Z., and Meinert, L.D., 2015, A model for the intrusive sity, where her studies focused on the world-class
sequence and Cu-Zn skarn formation at the Antamina deposit, Peru: PAC- Antamina Cu-Zn skarn-porphyry deposit in Peru.
RIM 2015 Congress, Australasian Institute of Mining and Metallurgy (Aus-
Her research combined field investigations with geochemistry, geochronol-
IMM), Hong Kong, China, March 18–21, 2015, Proceedings, p. 423–429.
Mrozek, S.A., Chang, Z., Meinert, L., and Creaser, R., 2017, Using field ogy, petrography, and fluid inclusion analysis to elucidate spatiotemporal
observations and geochronology to constrain the age of magmatic-hydro- relationships between porphyry emplacement, host-rock composition, hydro-
thermal activity at the Antamina deposit, Peru [ext. abs.]: International thermal alteration, and mineralization. Stephanie has 14 years of experience
Congress of Prospectors and Explorers (ProExplo), 10th, Lima, Peru, 2017, in mining and exploration, with a decade of that time devoted to working on
Extended Abstracts, p. 47–51. skarns.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/115/1/177/4904048/4698_mrozek_et_al.pdf


by University of Waterloo user

You might also like