You are on page 1of 21

Journal of Geochemical Exploration 232 (2022) 106889

Contents lists available at ScienceDirect

Journal of Geochemical Exploration


journal homepage: www.elsevier.com/locate/gexplo

Geology, alteration geochemistry, and exploration geochemical mapping of


the Ertsberg Cu-Au-Mo district in Papua, Indonesia
Utreck Rumbiak a, b, Chun-Kit Lai c, *, Reza Al Furqan a, Mega Rosana b, Euis Yuningsih b,
Basilios Tsikouras c, Elena Ifandi c, Amal Izzati Afiqah binti Abdul Malik c, Huayong Chen d
a
PT. Freeport Indonesia, Freeport-McMoran Copper and Gold, Tembagapura, Papua 99910, Indonesia
b
Faculty of Geological Engineering, Padjajaran University, Bandung, 40312, Indonesia
c
Geoscience Programme, Faculty of Science, Universiti Brunei Darussalam, Gadong, BE1410, Brunei Darussalam
d
Guangzhou Institute of Geochemistry, Chinese Academy of Science, Guangzhou 510640, China

A R T I C L E I N F O A B S T R A C T

Keywords: The Ertsberg-Grasberg district hosts the world’s largest Cu-Au(–Mo) skarn system. The mineralization is related
Ertsberg-Grasberg district to the Pliocene Ertsberg- and Grasberg intrusive complexes, which intruded the Jurassic-Cretaceous Kembe­
Porphyry-skarn mineralization langan Group siliciclastics and Cenozoic New Guinea Limestone Group. Despite previous studies, elemental
Alteration geochemistry
migration across the porphyry-skarn system and across the different ore stages is still unclear. Here, we attempt
Geochemical mapping
Indonesian Papua
to tackle this issue by integrating ore deposit geology, alteration/mineralization paragenesis, whole-rock
geochemistry of the Ertsberg granitoids and carbonate wallrocks, together with spatial mapping of hydrother­
mal minerals and target/pathfinder trace elements at the Deep Mill Level Zone of Ertsberg. Ores at Ertsberg are
mainly skarn-hosted, whose formation comprises five stages: (I) prograde skarn; (II) retrograde alteration; (III)
massive anhydrite replacement; (IV) late hydrothermal veining; (V) supergene alteration. Massive magnetite and
main Cu–Au mineralization occurred in Stage II and III, respectively, although minor Cu mineralization per­
sisted to Stage IV.
The Ertsberg granitoids have medium SiO2 and high Al2O3 contents, and consist mainly of (quartz-)monzonite
and minor (monzo)diorite. Least-altered samples have upper crust-normalized enrichments in certain large ion
lithophile elements (LILEs, e.g., U and K) and depletions in all high field strength elements (HFSEs, e.g., Ti, Y, and
Yb). The samples are fresh to moderately-altered, featured by garnet-diopside skarn and/or porphyry-style
(potassic, propylitic, phyllic) alterations and anhydrite replacement. Elemental mass balance calculation in­
dicates that the porphyry-style alteration zones are featured by feldspar destruction and the accompanying alkali
depletions, although the loss of K–Rb (in potassic and phyllic zones) and Ca-Ba-Sr (in propylitic zone) are
balanced by their respective capturing in secondary K-feldspar-mica and actinolite-epidote-anhydrite. Decar­
bonation of the dolomitic wallrocks likely released Mg-Ca-Sr, which facilitated the subsequent forsterite-diopside
exoskarn and anhydrite alteration. The enrichments of various ore-related target/pathfinder elements in these
altered samples are consistent with their mineralized nature, e.g., Fe–Mn (magnetite), Cu-Fe-Bi-Se-Te (bornite),
Mo–Re (molybdenite), Au-Ag-As-Bi-Te (auriferous pyrite/arsenopyrite), Zn-Fe-Mn-Sn-In (sphalerite), and Pb-
Ag-Bi (galena).
Alteration mineral and target/pathfinder elemental distribution patterns show that the ore fluids were likely
originated from the porphyry hydrothermal (potassic) centers near the Ertsberg pluton margin. The fluids then
spread out from the potassic centers, forming the surrounding propylitic zone, and migrated along the largely
NW-SE-oriented intrusive contact and fault zones. This formed the endoskarn in the Ertsberg pluton, the
forsterite-diopside exoskarn in the Waripi Formation dolomite, and the garnet-diopside exoskarn in the Ekmai
Formation limestone/calcareous sandstone-siltstone. As the hydrothermal system started to wane, the prograde
skarn formation is partially overprinted/enveloped by retrograde serpentine and actinolite/tremolite-epidote-
chlorite alterations, and then by magnetite mineralization and massive anhydrite replacement in the

* Corresponding author.
E-mail addresses: urumbiak@fmi.com (U. Rumbiak), chunkitl@utas.edu.au (C.-K. Lai), ralfurqa@fmi.com (R. Al Furqan), mega.fatimah.rosana@unpad.ac.id
(M. Rosana), e.tintin@unpad.ac.id (E. Yuningsih), basilios.tsikouras@ubd.edu.bn (B. Tsikouras), elena.ifandi@ubd.edu.bn (E. Ifandi), 16B2161@ubd.edu.bn
(A.I.A. binti Abdul Malik), huayongchen@gig.ac.cn (H. Chen).

https://doi.org/10.1016/j.gexplo.2021.106889
Received 18 April 2021; Received in revised form 25 July 2021; Accepted 29 August 2021
Available online 4 September 2021
0375-6742/© 2021 Elsevier B.V. All rights reserved.
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

carbonate wallrocks. The texturally-destructive skarn formation, and the dissolution of carbonate wallrocks and
hydrothermal anhydrite were likely pivotal in creating the highly-fractured Broken Zone, which served as a
structural trap for the high-grade Cu-Au-Mo ore deposition.

1. Introduction (Eocene), Sirga (L. Oligocene), and Kais (U. Oligocene to M. Miocene)
(Fig. 2) (Leys et al., 2012; McDowell et al., 1996). The Waripi Formation
The Ertsberg-Grasberg district (Papua, Indonesia) hosts a number of comprises sucrosic (friable and vuggy) dolostone, crystalline dolomitic
(super)-giant skarn-/porphyry-type Cu-Au(–Mo) deposits, including limestone, and (minor) quartz arenite, and is capped by a < 2-m thick
the Grasberg, Ertsberg East, and Kucing Liar, plus several smaller ones nodular anhydrite bed. The Foumai Formation comprises well-sorted,
such as the Big Gossan (Freeport-McMoran Inc., 2020; Leys et al., 2012). shallow-marine platform fossiliferous limestone and siliciclastic dolo­
These deposits are hosted in/around the Pliocene Ertsberg- and Grasberg stone: The lower part (thickness: ~ 100 m) contains quartz-bearing
intrusive complexes, and in the exoskarn of the Cretaceous Kembe­ dolomitic limestone with large benthic forams, whilst the upper part
langan Group (Gp.) siliciclastics and Tertiary New Guinea Limestone Gp. (thickness: ~ 100–200 m) contains quartz-bearing limestone with
wallrocks (Fig. 1). The ore formation is generally interpreted to have bryozoa and echinoderm fossils. Quartz is well-sorted and angular, and
linked to a short burst of asthenospheric upwelling-related magmatism is less abundant in the upper part than in the lower part. Sandstone beds
(3.1–2.8 Ma; Wafforn, 2017), triggered by crustal delamination after the are present locally in the middle-lower parts of the formation. The Sirga
collision of the northern-edge of Australian plate with an intra-oceanic Formation comprises terrestrial to shallow-marine, thick-/cross-bedded
arc (Cloos et al., 2005). Despite many previous studies, it is still un­ calcareous sandstone, whilst the Kais Formation is composed of massive
clear how hydrothermal alteration at Ertsberg mobilized and trans­ thick (>950 m) fossiliferous limestone. Lower part of the Kais Formation
ported the various ore-related elements, how such elemental transport is featured by benthic forams and glauconite (up to 20%, esp. in marl
was influenced by the local lithologies and structures, and how that beds), whilst the upper part contains mainly red algae and local inter­
contributed to the world-class skarn ore formation. In this study, bedded coal seams (10-to 30-cm-thick). The Kais Formation was inter­
therefore, we aim to answer these questions through an integration of preted to have deposited in a restricted shallow-marine carbonate
ore deposit geology, alteration/mineralization paragenesis, whole-rock platform environment (Van Ufford, 1996).
geochemistry (pluton and carbonate wallrocks), and mineral/
elemental spatial distribution maps from Ertsberg. Using elemental mass
2.2. Exploration history and geological setting of Ertsberg-Grasberg
balance calculation, we further discuss the migration and deposition of
district
ore-forming materials across the pluton and wallrocks, which formed
the distinctive elemental/mineral zoning patterns we see today.
Despite mineralization at Ertsberg was first documented by A. Colijn
and J.J. Dozy back in 1936, major exploration and mine development
2. Geological background
did not occur until the early 1960s by Freeport. A series of orebodies
were discovered around the Ertsberg Intrusive Complex, which would
2.1. Regional geology
eventually constitute the East Ertsberg Skarn System. The Dom skarn
deposit was also discovered on the southern margin of the Intrusive
The island of New Guinea is bordered by the Caroline and Bismarck
Complex in a limestone embayment. Subsequent drilling has discovered
plates to the north and east, respectively, with the northern edge of the
the supergiant Grasberg (1988), the high-grade Big Gossan skarn (1991),
Australian plate underthrust beneath the middle-southern part of the
and the giant Kucing Liar skarn (1994) deposits (Leys et al., 2012;
island. The island is composed of four tectonic terranes, namely (from
Meinert et al., 1997; Pollard et al., 2005). A detailed account of the
south to north) the New Guinea foreland basin, the Central Range fold-
exploration history has been given by Van Leeuwen (2018).
and-thrust belt, the Derewo (aka. Ruffaer) metamorphic belt, and the
In the Ertsberg-Grasberg district, lithostratigraphy is dominated by
Melanesian accreted oceanic arc-forearc complex (Fig. 1a) (Cloos et al.,
the New Guinea Limestone Group. Although all the four formations
2005, and ref. therein). Formation of the Central Range, where the
(Waripi, Faumai, Sirga, and Kais) host high-grade Cu–Au skarn
Ertsberg-Grasberg district is located, was likely linked to the collision
mineralization, the lower part of the Waripi Formation (anhydrite
between the Australian plate and the Melanesian arc. The collision had
nodule-bearing dolomite, sandy dolostone with thin-bedded quartz
likely jammed the subduction system, and caused rapid ENE-WSW-
sandstone) represents the main ore host, particularly for skarn ores. The
directed (~070◦ ) crustal shortening and the Central Range uplift since
depositional environment is interpreted as supratidal to shallow marine
~13 Ma. The crustal shortening has also formed the Yellow Valley
shelf, reflecting the transition from the mainly siliciclastic Kembelangan
Syncline, into which the Ertsberg and Grasberg intrusive complexes
Gp. sedimentation into the New Guinea Limestone Gp. carbonate
were emplaced (Fig. 1b). The plutonism and the associated porphyry-
deposition (Van Ufford, 1996). Beneath the Waripi Formation, all the
skarn mineralization are interpreted to have formed by post-collisional
three members of the Ekmai Formation (Kembelangan Group) are pre­
crustal delamination and asthenospheric upwelling (Cloos et al., 2005;
sent. The shale (3–5 m thick) and limestone (≤100 m thick) members are
Leys et al., 2012; Warren and Cloos, 2007).
commonly hornfels-facies metamorphosed, and the limestone member is
The Central Range comprises Proterozoic to Upper Tertiary se­
also partially skarn-altered and hosts significant mineralization. The
quences. In the Ertsberg-Grasberg district, stratigraphy comprises the
sandstone member (~600 m thick) is arkosic and hosts disseminated-/
predominantly-siliciclastic (minor calcareous) Kembelangan Group and
fracture-filled mineralization beneath the skarn orebodies. These sedi­
the New Guinea Limestone Group. The Kembelangan Group has four
mentary units are intruded by various equigranular to porphyritic in­
formations, i.e., the Kopai (M. Jurassic to L. Cretaceous), Woniwogi (U.
trusions. Major local structures include the shallowly NW-plunging
Jurassic to L. Cretaceous), Piniya (L. to M. Cretaceous), and Ekmai (U.
Yellow Valley Syncline, and the NW-trending Wanagon, Ertsberg, and
Cretaceous-Paleocene). The Ekmai Formation comprises a lower sand­
Idenberg reverse faults (Gibbins, 2006; Mertig et al., 1994; Rubin and
stone, a middle limestone, and an upper shale member. The sandstone
Kyle, 1998). Many of these faults are crosscut by NE-trending sinistral
member comprises mainly quartz arenite, with the upper 90 m con­
strike-slip faults, notably the Grasberg Fault (Fig. 1b). Based on cross­
taining mixed mudstone and limestone, and capped by 20-m thick
cutting relations, McDowell et al. (1996) documented at least nine
calcareous mudstone. The overlying Tertiary New Guinea Limestone
intrusive bodies in the district, most of which are fine-grained porphy­
Group has also four formations, i.e., the Waripi (Paleocene), Faumai
ritic dikes and sills. At Grasberg, three magmatic phases have been

2
U. Rumbiak et al.
3

Journal of Geochemical Exploration 232 (2022) 106889


Fig. 1. (a) Tectonic map of the New Guinea island, showing the location of the Ertsberg-Grasberg district (after Cloos et al., 2005); (b) Geologic map of the Ertsberg-Grasberg district (after Wafforn, 2017). Abbre­
viations: BG – Big Gossan GB (E) – Gunung Bijih (Ertsberg); GBT (EESS) – Gunung Bijih Timur (East Ertsberg Skarn System); GIC – Grasberg Intrusive Complex; KL – Kucing Liar.
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

recognized, namely (from oldest to youngest) the Dalam Phase (diorite), their depths, i.e., (top to bottom) the Gunung Bijih Timur, Intermediate
Main Grasberg Intrusion (MGI), and the early/late Kali dikes, among Ore Zone, Deep Ore Zone, Mill Level Zone, and Deep Mill Level Zone
which the MGI represents the main ore-forming phase and the Kali dikes (DMLZ) (e.g., Mertig et al., 1994) (Figs. 2–3). Mineralization remains
are post-mineralization. The magmatism lasted roughly from 3.1 to 2.8 open to the northwest and at depth below the DMLZ. The deepest mining
Ma (Wafforn, 2017, and ref. therein). Ore-causative granitoids at Gras­ zone DMLZ is investigated here because it is the least studied (among the
berg are mainly K-rich monzonitic to dioritic, and their emplacement are five mining zones) and the porphyry-skarn transition is well exposed in
clearly controlled by local structures. Porphyry-type mineralization is this zone. Mineralization at the East Ertsberg Skarn System is predom­
mainly hosted within the MGI. Comparatively, the Ertsberg Intrusive inantly hosted by the Faumai (Gunung Bijih Timur and upper Interme­
Complex is much less mineralized, and the skarn-type mineralization is diate Ore Zone) and Waripi (lower Intermediate Ore to Deep Mill Level
mainly hosted in the exoskarn of the Lower New Guinea Limestone Zone) formations, and minor by the underlying Kembelangan Formation
Group and Upper Kembelangan Gp. wallrocks (Gandler and Kyle, 2008; (Fig. 2). Footwall of the East Ertsberg Skarn System comprises the
Kyle et al., 2014; Leys et al., 2012; Pollard et al., 2005). Gibbins (2006) moderately-intensely endoskarn-altered Ertsberg granitoids, with
identified four igneous rock types (the Hybrid Phase, the Ertsberg Dio­ porphyry-style vein-type mineralization developed 50–150 m into the
rite, porphyry dikes, and (minor) aplite dikes) in the Ertsberg Intrusive pluton from the skarn zone. The hangingwall contact consists of hy­
Complex, with the Hybrid Phase being a rind around the Ertsberg Diorite drothermal breccia and massive sulfide mineralization, both in sharp
that cooled faster. Zircon U–Pb dating on the Ertsberg Diorite yielded contact with the marbleized limestone wallrocks (Fig. 3) (Gandler,
2.97 ± 0.58 and 3.28 ± 0.3 Ma, and that on the porphyry dikes yielded 2006; Gandler and Kyle, 2008; Kyle et al., 2014; Leys et al., 2012). The
similar age of 2.98 ± 0.84 Ma (Gibbins, 2006). rock package is intruded by a number of porphyry dikes in the potassic
alteration zone. Some of these dikes lack significant alteration (Gibbins,
2.3. East Ertsberg Skarn System 2006) and may have been post-mineralized. Skarn alteration is strata­
bound and replaces the Ekmai and Waripi Fm. carbonate wallrocks.
The East Ertsberg (Indonesian: Gunung Bijih Timur) Skarn System is Skarn mineralization at Ertsberg was molybdenite Re–Os dated to be
situated in the southern limb of the Yellow Valley Syncline (Fig. 1). The 2.54 ± 0.02 Ma (Mathur et al., 2005) and phlogopite Ar–Ar dated to be
system is a steeply-dipping continuous mineralized body that extends for 2.59 ± 0.15 Ma (Pollard et al., 2005), slightly younger than its Grasberg
over 1.5 km deep, and can be divided into five mining zones according to (2.88 ± 0.02 Ma) and Kucing Liar (3.01 ± 0.02 Ma) counterparts

a Age Formation Thickness (m) Description


b
950+ Fossilferous limestone

25–40 Calcareous sandstone


217 Limestone, locally fossiliferous

Dolomite, limestone,
300 evaporite, sandstone,
siltstone
4 Calcareous shale
90 Sandy limestone

600 Quartz sandstone

Monzonite c
600 Mudstone, siltstone

Ekmai Fm. sandstone

950+ Sandstone, shale, siltstone


Monzonite d

Ekmai Fm. limestone (exoskarn)

Sandstone, conglomerate,
800+ limestone, mudstone Waripi Fm. limy dolomite e

Forsterite-diopside skarn

Sample location Massive magnete ore

Fig. 2. (a) Lithostratigraphic column (after Van Ufford, 1996) with sampling depths (red dots) and (b) geologic cross-section of the Ertsberg Cu skarn district; (c-e)
Drill-core photos showing the contact relations (white dotted-line) between: (c) biotite-epidote-chlorite-altered monzonite and garnet-epidote ± magnetite-altered
Ekmai Fm. calcareous sandstone. Both rock types cut by anhydrite-quartz and chalcopyrite-pyrite ± galena veins; (d) weakly biotite-chlorite-altered monzonite and
(dark-)green exoskarn-altered Ekmai Fm. limestone with magnetite mineralization. Both rock types cut by anhydrite-quartz ± sulfide veins and sericite selvages.
Chalcopyrite-pyrite patches also occur in the exoskarn; (e) Waripi Fm. limey dolomite, (dark-)greenish forsterite-diopside exoskarn, and massive magnetite ore.
Abbreviations: Te (Ertsberg pluton); Tw (Waripi Fm. dolomite); Ekmai Formation: Kkeh (shale), Kkel (limestone), Kkes (calcareous sandstone-siltstone). (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

4
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

736600 E 737000 E 737400 E 737800 E

LEGEND

Ertsberg Diorite

Waripi Fm

Ekmai Fm. –
Shale Member
954 9400 N
Ekmai Fm. –
Limestone Member
Ekmai Fm. –
Sandstone Member

Hornfels

Exoskarn

Broken Zone
954 900 N

Sheeted Vein Zone

954 8600 N

Fig. 3. Geologic map of the Deep Mill Level Zone (DMLZ) at 2600 L.

(Mathur et al., 2000, 2005), and probably also younger than minerali­ Stage IV (Fig. 4h). Metallic minerals include mainly chalcopyrite,
zation at Big Gossan (phlogopite 40Ar/39Ar: 2.82 ± 0.04 Ma) (Pre­ bornite, and pyrite, with minor native gold, electrum, gold(‑silver)
ndergast, 2003). telluride, molybdenite, sphalerite, galena, pyrrhotite, marcasite, covel­
lite, digenite, and chalcocite (Fig. 6). Spatial distribution of the alter­
ation/mineralization zones delineated through alteration mineral
2.4. Alteration/mineralization paragenesis and veining at Deep Mill Level
mapping will be presented in Section 5.
Zone
At the DMLZ, skarn-type veining is well developed in/around the
pluton, and consists of eight types (Type 1 to 8), i.e., (1) Stage I garnet-
At the DMLZ, Cu–Au mineralization is mainly (>75%) hosted in the
diopside skarn veins (1–5 cm wide) in the endoskarn zone in contact
skarn-style altered rocks (esp. magnetite-/anhydrite-altered Lower
with the siliciclastic/carbonate wallrocks; (2) Stage II epidote-chlorite-
Waripi Formation) and minor in the porphyry-style altered (quartz-)
quartz ± serpentine veinlets (0.1–4 mm) with magnetite alteration
monzonite (Fig. 4a–c). Porphyry-style alterations include mainly
halo; (3) Stage II sinuous magnetite-amphibole veinlets (0.5–1 mm) in
potassic, minor propylitic, and locally phyllic. These alterations are best
the skarn and potassic/propylitic alteration zones (Fig. 4c); (4) Stage III
developed toward the intrusive contact and are rarely texture-
sinuous, greyish-white “A”-type barren quartz veins (0.5–10 mm)
destructive. Potassic alteration occurs as selective replacement of
(Fig. 4d). These veins are well-developed in the skarn and potassic zone
hornblende by biotite, and the formation of pinkish K-feldspar. Potassic-
and crosscut skarn veins; (5) Stage III “B”- and “AB”-type quartz-
altered rocks are locally overprinted by propylitic (actinolite/tremolite-
anhydrite veins/veinlets (Fig. 4e): the former is straight (0.5–25.0 mm
epidote-chlorite) alteration, e.g., the alteration of hydrothermal biotite
wide), whilst the latter is transitional between A- and B-type. These
into chlorite (Figs. 4d–e and 5a). Phyllic alteration (Fig. 5b–c) is limited
veins are commonly Cu mineralized with spotty/blebby bornite, chal­
to selvages of quartz-sericite-pyrite veins in highly-fractured area, for
copyrite, molybdenite, and pyrite, and are best developed in the
instance the area between the NE22 and NE32 faults in northwestern
potassic-propylitic and skarn zones. Anhydrite is present is some A-, B-
DMLZ (Fig. 3). Potassic and propylitic alterations likely took place
and AB-type veins at 2600–3000 L (meter elevation level); (6) Stage III-
before prograde-skarn alteration, as evidenced by the alteration of
IV chalcopyrite veins are irregular (0.3–5 mm wide), and contain also
potassic- and propylitic-altered monzonite by the texturally-destructive
minor pyrite and quartz/anhydrite (Fig. 4d). These veins/veinlets are
garnet endoskarn (Fig. 5d–e). Porphyry-style mineralization occurs
found cutting AB-/B-veins in the Ertsberg pluton, hornfels, and the
mainly as sulfide (chalcopyrite-pyrite ± bornite)-quartz-anhydrite veins
dolomite exoskarn. Type-5/-6 veins account for the main copper
and stockwork. In this study, we focused mainly on the skarn-style
endowment at Ertsberg; (7) early-Stage IV pyrite and D-type pyrite-
alteration as it is spatially more widespread and economically more
quartz-sericite veins (0.3–10 mm) with pyrite centerline (Figs. 4f and
important. According to mineral paragenesis and crosscutting relations,
6). The veining crosscut the potassic-/propylitic-altered rocks, and is
the skarn-style alteration/mineralization can be divided into five stages
also found in hornfels zone and locally in the dolomite exoskarn with
(Fig. 7), namely (I) prograde skarn; (II) early-retrograde alteration; (III)
minor sphalerite and galena, and rare chalcopyrite; (8) late-Stage IV
late-retrograde alteration with massive anhydrite replacement; (IV) late
calcite-anhydrite veins (0.5–1 cm) are ore-barren and crosscut older
hydrothermal veining; (V) supergene alteration. Massive magnetite
vein generations (Fig. 5f).
mineralization occurred in Stage II. The main Cu–Au mineralization
occurred in Stage III, although minor Cu mineralization is also found in

5
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

a 1m b 1m

c d

e f

g h

Fig. 4. (a–b) Mine-tunnel and (c-h) drill-core photos from the DMLZ in Ertsberg: (a) Contact between exoskarn (Fo-Di-Gt-Mag-Anh) and Waripi Fm. dolomitic
marble. Inset photo shows anhydrite nodule clusters; (b) Potassic (Kfds-Bt-Hbl)-altered Ertsberg quartz monzonite cut by quartz ± anhydrite-pyrite stockwork; (c)
Intensely and pervasively endoskarn-altered Ertsberg monzonite. Endoskarn garnet alteration partially replaced by patchy retrograde chlorite-epidote alteration,
which is in turn overprinted by semi-massive or Type-3 magnetite vein mineralization, and eventually by densely-disseminated or vein-type chalcopyrite miner­
alization; (d) Ertsberg monzonite is potassic and then propylitic (Act-Epi-Chl) altered with disseminated chalcopyrite-pyrite mineralization. Altered rocks are cut
firstly by Type-4 quartz A-veins and then by Type-6 chalcopyrite veins; (e) Hornfelzed Ekmai Fm. shale brecciated by Type-5 quartz-gypsum vein that contains pyrite
and rare molybdenite; (f) Potassic (with weak overprinting propylitic)-altered Ertsberg monzonite intruded by Type-7 D-vein (qtz-ser) with pyrite centerline and
alteration halo; (g) Anhydrite-altered Waripi Fm. dolomite with patchy/Type-3 wavy-veinlet magnetite and densely-disseminated/Type-5 veinlet bornite-covellite-
chalcopyrite mineralization; (h) Massive magnetite mineralization overprinted by densely-disseminated bornite-covellite and (tarnished) chalcopyrite mineraliza­
tion. Abbreviations: Act (Actinolite); Anh (Anhydrite); Bio (Biotite); Bn (Bornite); Ch (Chlorite); Cpy (Chalcopyrite); Cv (Covellite); Di (Diopside); Ep (Epidote); Feld
(Feldspar); Fo (Forsterite); Gyp (Gypsum); Hbx (Hornblende); Kfds (K-feldspar); (Mag) Magnetite; Mo (Molybdenite); Qtz (Quartz); Serp (Serpentine).

6
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 5. Thin-section photomicrographs (XPL) of the Ertsberg (quartz-)monzonite and (meta-)sedimentary wallrocks: (a) Weakly epidote-altered monzonite; (b)
Intensely sericite- and anhydrite-altered monzonite with rare least-altered igneous feldspar grains preserved; (c) Intensely biotite- and anhydrite-altered monzonite
cut by quartz vein. Hydrothermal biotite occurs as fine-grained aggregates; (d) Texturally-destructive, garnet-phlogopite endoskarn-altered monzonite cut by
anhydrite vein; (e) Epidote-chlorite-altered monzonite with plagioclase partially preserved; (f) Magmatic texture of monzonite completely obliterated by garnet
endoskarn alteration. Rock cut by late calcite vein; (g) Hornfelzed Ekmai Fm. shale with microcrystalline quartz, epidote-altered; (h) Hornfelzed and epidote-altered
Ekmai Formation with porphyroblastic quartz aggregate; (i) Ekmai Fm. meta-sandstone with recrystallized irregular quartz grain margin and subgrain texture,
epidote-altered; (j) Waripi Fm. dolomitic marble with phlogopite and semi-massive sulfide ore; (k) Phlogopite exoskarn-altered and sulfide-mineralized Ekmai Fm.
sandy limestone with occasional quartz and feldspar grains preserved. Abbreviations as in Fig. 4 and Bi (Biotite); Cal (Calcite); Grt (Garnet); Mu (Muscovite); Opq
(Opaque); Phl (Phlogopite); Plg (Plagioclase); Qtz (Quartz); Rt (Rutile); Ser (Sericite); Sul (Sulfide).

3. Samples and methods diopside, forsterite, and brownish andradite/grossular garnet. Intensely-
altered samples have their primary igneous texture obliterated. For­
3.1. Sample descriptions sterite is locally retrograde-altered to serpentine and talc, whilst diop­
side to actinolite/tremolite, epidote, chlorite, and phlogopite. Primary
In this study, a total of 51 samples were collected from the DMLZ drill hornblende and augite are largely altered, whereas feldspar grains are
cores, including those from the Ertsberg pluton (n = 13), Waripi Fm. still locally preserved. Late calcite veins are common. Our endoskarn
dolomite (n = 30), and Ekmai Fm. limestone (n = 6), and calcareous fine samples contain disseminated chalcopyrite-pyrite, minor molybdenite,
sandstone-siltstone (n = 2). These samples range from unaltered to sphalerite and galena, and rare electrum (Figs. 4c and 5d).
moderately-altered to reveal the alteration impact on the whole-rock The potassic-altered samples show selective, texturally non-
geochemistry. Petrographic thin sections were prepared and observed destructive alteration of augite-biotite-plagioclase/alkali feldspar by
under transmitted/reflected light microscopy for their lithology, alter­ hydrothermal K-feldspar-amphibole-biotite. Potassic alteration is over­
ation/mineralization features, and mineral paragenesis. printed locally by weak propylitic and anhydrite alterations (Figs. 4d
Ertsberg granitoids: The granitoids are often also called the Ertsberg and 5c). Mineralization consists of chalcopyrite (as quartz-chalcopyrite
Diorite, but available geochemical data (Section 4.1) indicate that these veins or fracture-infill), veined pyrite, trace molybdenite (as blebs in
rocks are mostly (quartz-)monzonitic. The samples are medium-grained, quartz-anhydrite veins), and rare bornite and sphalerite (Fig. 6).
and have mainly plagioclase (55–60 vol%) and alkali feldspars (35–40 The propylitic-altered samples are greenish-grey and only weakly-
vol%), and minor quartz (<10 vol%), igneous hornblende (<10 vol%) altered. Alteration minerals include mainly actinolite, epidote and
and augite (<5 vol%) (Fig. 5a). Secondary minerals (5–40 vol%) include chlorite, whilst minor quartz and trace residual hornblende and biotite
mainly hydrothermal actinolite (fibrous or rims around augite), patchy were also observed. Augite is partially altered into actinolite, epidote
epidote and chlorite, veined microcrystalline quartz, and calcite and and chlorite, biotite to chlorite, and plagioclase feldspars to epidote and
sericite in plagioclase hair-like cracks or pseudomorphs (Fig. 5f). For the chlorite. Weak disseminated chalcopyrite-molybdenite-pyrite minerali­
altered samples, their alteration styles include diopside-forsterite-garnet zation is cut by anhydrite-quartz veins/veinlets (Fig. 5).
endoskarn (n = 1), potassic (n = 4), propylitic (n = 3), and sericite (n = The phyllic-altered samples are mainly featured by sericite(-biotite)
2). selvages and the associated pyrite veining. Apart from pyrite, minor
Endoskarn-altered (quartz-)monzonite contains mainly prograde galena and molybdenite, and trace chalcopyrite and sphalerite are also

7
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 6. Laser-mount photos and photomicrographs (reflected and transmitted light) of the Ertsberg Type-7 ore veins: (a) laser mount photo, showing the vein with
pyrite centerline; (b–c) Pyrite and chalcopyrite with quartz; (d–e) Very fine-grained electrum in quartz; (f) laser-mount photo, showing the vein with alteration halo;
(g–h) Sphalerite rimmed by pyrite in quartz vein; (i–j) Disseminated chalcopyrite, bornite, and covellite with quartz; (k–l) Disseminated pyrite, chalcopyrite, and
sphalerite with quartz. Abbreviations as in Figs. 4 to 5 and Apy (Arsenopyrite); Bo (Bornite); El (Electrum); Sph (Sphalerite).

Composion
Mid-retrograde + massive Late-retrograde + late hydrothermal Supergene
Prograde skarn (I) Early-retrograde (II)
anhydrite replacement (III) veining (IV) alteraon
Mineralizaon Fe Cu-Au Cu-Au(Pb-Zn)
Forsterite
Diopside
Garnet
Scapolite
Magnete
Anhydrite
Quartz
Mica
Serpenne
Tremolite – Acnolite
Phlogopite
Talc
Chlorite
Chalcedony
Epidote
Calcite
Bornite
Chalcopyrite
Gold
Molybdenite
Pyrite
Galena
Sphalerite
Pyrrhote
Marcasite
Electrum
Covellite
Arsenopyrite

Fig. 7. Summary diagram of the skarn-style alteration/mineralization paragenesis at Ertsberg.

present (Fig. 5). minerals and calcite cement. Some sandstones are deformed and meta­
Ekmai Fm. limestone (Kkel), calcareous fine-sandstone-siltstone morphosed, as shown by the irregular boundary, and the subgrain and
(Kkes) and shale (Kkeh): Similar to the dolomite samples, the lime­ undulatory extinction textures of quartz. Many siliciclastic (Kkes and
stone samples are variably marbleized, and consist of calcite (70–80 vol Kkeh) samples are hornfelzed with local weak exoskarn alteration.
%), quartz (<10 vol%), mica (<10 vol%), and minor (<5 vol% each) These hornfelzed samples comprise predominantly microcrystalline
chlorite and scapolite. The fine-sandstone-siltstone samples are sub­ quartz (>85 vol%) plus minor (<5 vol% each) mica, epidote, chlorite,
rounded and moderately sorted, and have mainly quartz (>75 vol%), and opaque minerals. The altered samples contain disseminated pyrite
muscovite (<10 vol%), and minor (<5 vol% each) epidote, opaque and Fe–Ti oxides, and trace galena and sphalerite (Figs. 4e and 5g–i,k).

8
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 8. Harker-type diagrams for the Ertsberg granitoids.

Waripi Fm. dolomite: The rocks are variably metamorphosed/ respectively. Detailed calculation procedures are similar to those out­
recrystallized into medium-grained dolomitic marble. The dolomite is lined by Ulrich and Heinrich (2001). Calculated elemental gain/loss of
sucrosic, well-sorted, vuggy (from foraminifera dissolution), and with the granitoid samples (of a particular alteration style) are shown with
occasional fossil remnants (e.g., bryozoa, coral, ooids, and peliods). The their average value, whereas those of the meta-sedimentary wallrock
unaltered samples (n = 9) contain mainly dolomite ± calcite (70–90 vol samples are shown with their individual value due to their wider
%), quartz (10–20 vol%), plus minor (<5 vol%) feldspar and glauconite. compositional ranges.
Some moderately to intensely exoskarn-altered samples (n = 20) are also
anhydrite-altered (n = 10) and magnetite-mineralized (n = 8). Alter­ 3.3. Distribution maps of target/pathfinder elements and minerals
ation minerals include prograde (forsterite, diopside, garnet) and
retrograde (serpentine, tremolite, phlogopite, talc) ones and anhydrite. To reveal the spatial mineral-geochemical change caused by alter­
Chalcopyrite, pyrite, molybdenite, sphalerite, and (minor) galena are ation and its possible lithologic/structural links, various spatial
present in the skarn-altered samples, whilst trace (arseno)pyrite, elemental/mineral maps were constructed at 2600 L. Further analytical
sphalerite, and galena are locally found in the unaltered samples details are confidential and not disclosed here. Spatial elemental maps
(Figs. 4g and 5j). were constructed on the target elements (Cu, Au, Mo, As, Ag, Pb, Zn),
and alteration-sensitive pathfinder elements and their ratios (Bi, Te, Sb,
3.2. Whole-rock geochemical analysis and mass balance calculation Co, Ti/Sr, (Cs + Rb)/Th). Mineral maps were made on the major ore
(chalcopyrite) and alteration minerals, including porphyry-type (biotite,
Both whole-rock major and trace element compositions were chlorite, and sericite) and skarn-type (diopside, forsterite, garnet,
analyzed at the ALS Lab (Australia). The samples were subjected to magnetite, and serpentine) ones and anhydrite. The Cu, Au, and Ag
lithium borate fusion and then acid dissolved before analysis by ICP-MS. elemental maps are not disclosed due to confidentiality reason. Thus, the
This provides the most quantitative analytical approach for a broad chalcopyrite & Mo and As maps are used as a proxy for Cu and Au,
elemental suite, and a total of 54 elements were analyzed. To increase respectively, although the spatial distribution patterns between Cu–Au
the representativeness of our whole-rock geochemical dataset, pub­ and their proxies are not without difference.
lished Ertsberg granitoid data by McMahon (1994) and Gibbins (2006)
were also compiled for comparison. As the rock samples are variably 4. Results
altered and have different loss of ignition (LOI), major element oxides
were normalized to 100 wt% excluding volatiles. Geochemical data 4.1. Whole-rock geochemistry of the Ertsberg granitoids
treatment and discrimination diagram plotting were performed with
ioGAS (version 7.3) software. Mass balance calculation for the elemental Among the 13 granitoid samples analyzed, 3 are unaltered/least-
gain/loss (in different lithologies under different alteration styles) were altered and 10 are variably-altered (4 potassic, 3 propylitic, 2 sericite,
performed with the isocon method suggested by Gresens (1967) and and 1 endoskarn). The LOI of most samples (new and published) in the
modified by Grant (1986). Aluminum (Al2O3) is used as an immobile dataset are ≤3 wt%, and the LOI shows little correlation with immobile
element. The calculation results are illustrated by isocon diagrams, with element contents (e.g., P, Zr, REEs) and their ratios (e.g., Zr/Y, (La/
the isocon line gradient defined by the mass of the fresh samples against Sm)N) (Supplementary Fig. 1; Supplementary Table 1). Therefore,
that of the altered ones (M◦ /Ma). Elements fall above and below the immobile element geochemistry of our compiled dataset can be used for
reference isocon denote the elements gained and lost during alteration, the following petrogenetic discussion.

9
U. Rumbiak et al.
10

Journal of Geochemical Exploration 232 (2022) 106889


Fig. 9. Geochemical classification diagrams for the Ertsberg granitoids: (a) TAS (after Middlemost, 1994); (b) ASI vs. SiO2 (after Frost and Frost, 2008); (c) Sr/Y vs. SiO2 (after Richards et al., 2012); (d) Alteration box
plot (after Large et al., 2001); (e) Na/Al vs. K/Al (molar) (after Davies and Whitehead, 2006).
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

4.2. Mineral and elemental distribution maps of DMLZ

4.2.1. Mineral distribution maps


In these maps, porphyry-/skarn-type indicator minerals are used as a
proxy for their respective alteration/mineralization styles, i.e., potassic
(biotite), propylitic (chlorite), phyllic (sericite), anhydrite (anhydrite),
prograde-skarn (forsterite-diopside-garnet), retrograde (serpentine,
chlorite), magnetite (Fe-oxide mineralization), and chalcopyrite
(Cu–Au mineralization). At the DMLZ, all the porphyry-style alteration
is best developed along the intrusive contact with the Waripi Formation,
and weakens away from it. Two potassic alteration centers (biotite >6
vol%) are present in the NW part (between the EW22–EW25 faults) and
in the central part (at the South–EW27–NE26 fault intersection) of the
intrusive contact (Fig. 11a), whilst propylitic alteration is the strongest
(chlorite >6 vol%) between the NE22 and NE24 faults (Fig. 11b). Phyllic
alteration is best developed (sericite >15 vol%) in the highly-fractured
area between the NE24 and NE32–NW6 faults (Fig. 11c). Anhydrite
alteration is widespread both within and outside the pluton, but is best
developed (anhydrite >10 vol%) adjacent to the potassic center (around
the NE26 fault) along the South Fault, as well as in the Waripi Fm.
anhydrite nodule-bearing dolomite along an ESE-trend on both sides of
the Ertsberg 1 and EW19A-B faults (Fig. 11d).
Skarn-style alteration is developed mainly in the exocontact and
minor in the endocontact, with prograde skarn alteration occurring close
to the intrusive contact and being partially replaced and enveloped by
retrograde alteration. Forsterite alteration is developed predominantly
in the NW-part of the Waripi Fm. exoskarn zone, and is the strongest
(forsterite >10 vol%) along the NE30 and EW19A-B faults (Fig. 12a).
The diopside alteration center (diopside >15 vol%) is developed to the
Fig. 10. (a) Chondrite-normalized REE patterns and (b) upper continental crust southeast of its forsterite counterpart, along the Ertsberg 1 Fault and
(UCC)-normalized multi-element spidergram for the Ertsberg granitoids. between the NE30 and EW27 faults. The alteration extends into the
Chondrite- and UCC-normalizing values are from Sun and McDonough (1989) pluton around the EW27 and South faults. Another weaker diopside
and Taylor et al. (1981), respectively. alteration zone (diopside: 10–15 vol%) is located in the hornfelzed
Ekmai Formation, bounded by the intrusive contact and the EW13A-B
After normalizing the whole-rock geochemical data to 100 wt% faults (Fig. 12b). This alteration zone in the Ekmai Formation overlaps
excluding volatile components, most samples fall into narrow ranges of partially with an andradite/grossular garnet alteration center (garnet
SiO2 (61.40–63.29 wt%), TiO2 (0.75–0.90 wt%), and Al2O3 >6 vol%). Two other garnet alteration centers are present in the endo­
(15.32–16.20 wt%) (Supplementary Table 1). Unaltered samples are skarn zone around the NE21–EW20 and NE36–South fault intersections
high-K, metaluminous-peraluminous (ASI: molar [Al/(Ca - 1.67P + Na (Fig. 12c). Distribution of the major retrograde serpentine alteration
+ K)] = 1.0–1.4), and cluster inside the (quartz-)monzonite fields in the (serpentine >4 vol%) appears to have enveloped the prograde forsterite
TAS diagram (Fig. 9a–b). In the Harker-type diagrams, most published alteration, and is particularly well developed along the EW23 Fault,
data (Gibbins, 2006; McMahon, 1994) also fall on the same fractionation between the NE21 and NE 25 faults, and between the EW15 and EW19B
trends as our DMLZ samples, in which SiO2 correlates positively with faults (Fig. 12d).
K2O, Na2O, Zr, Zr/Y, and (La/Sm)N, but negatively with TiO2, MgO, CaO Magnetite alteration is only locally developed. The main occurrence
(and Sr), FeOT, and P2O5 (Fig. 8). The negative SiO2 vs. P2O5 correlation (magnetite >10 vol%) is centered in the southeastern Ertsberg pluton
is suggestive of I-type affinity, as also indicated by the absence of per­ (northeast of the NW4A Fault), developed with well-concentric zoning.
aluminous minerals such as cordierite and muscovite. The rocks have Minor occurrences are also present in the endoskarn zone between the
similar chondrite-normalized REE patterns, featured by LREE/MREE South Fault and the intrusive contact, as well as in the Waripi Fm.
enrichments and flat HREE patterns (Fig. 10a). When normalized to the exoskarn zone between the NE30, EW15, and Ertsberg 1 faults
upper continental crust (Taylor et al., 1981), the rocks are enriched in (Fig. 12e). Distribution of the chalcopyrite-highs (chalcopyrite >1 vol%)
certain large ion lithophile elements (LILEs, e.g., Rb, U, K), but depleted is concentrated mainly in the NW-part of the area between the NE21 and
in all high field strength elements (HFSEs, e.g., Y and Yb) (Fig. 10b). The NE32 faults, as well as in the SE-part of the pluton west of the EW1B
granitoid samples have generally elevated Sr/Y (mostly >40) and low Y Fault (Fig. 11e).
(<20 ppm), indicative of adakitic affinity (Fig. 9c). In the Chlorite-
Carbonate-Pyrite Index (CCPI) vs. Alteration Index (AI) alteration box 4.2.2. Elemental distribution maps
plot (Ishikawa et al., 1976; Large et al., 2001), the DMLZ samples trend As for the target elements (Mo, As, Pb, Zn), spatial distribution of Mo
toward the chlorite-pyrite alteration end-member. Data from McMahon bear more resemblance with Cu than that of chalcopyrite, probably
(1994) and Gibbins (2006) are more scattered, with some trending to­ because Cu is also hosted in other Cu minerals such as bornite and
ward the albite and epidote-calcite alteration end-members (Fig. 9d). In (minor) covellite, chalcocite, and digenite. The Mo-highs (>0.02%) in­
the Na/Al vs. K/Al (molar) plot (Davies and Whitehead, 2006), the side the pluton are located in the SE-part of the area between the EW13B
Ertsberg granitoid data fall on/close to the plagioclase-K-feldspar (bio­ and NW40 faults, whilst those outside the pluton are located mainly in
tite) trend, mainly toward K-feldspar-muscovite alteration and minor the Broken Zone (defined as Rock Quality Designation (RQD) <25%)
toward clay alteration (Fig. 9e). around the intersections of the Ertsberg 1–EW6 faults and NE30–NS2B
faults, and minor between the East Ertsberg Skarn System (EESS) W and
NE32 faults (Fig. 13a).
Distribution of As is widespread and shares certain similarities with

11
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 11. Spatial distribution maps for porphyry-style alteration mineral contents (%) at DMLZ 2600 L (a: biotite; b: chlorite; c: sericite), (d) anhydrite, and (e)
chalcopyrite.

12
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 12. Spatial distribution maps for skarn-style alteration mineral contents (%) at DMLZ 2600 L: (a) forsterite; (b) diopside; (c) garnet; (d) serpentine,
(e) magnetite.

13
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 13. Spatial distribution maps for target element contents (ppm) and major structures at DMLZ 2600 L: (a) Mo; (b) As; (c) Zn; (d) Pb.

that of Au. The As-highs (>100 ppm) are mainly developed in the highs (>20 ppm) (Fig. 14d), and they both correspond well with the
Broken Zone in the Waripi Formation (between the EW15–NE30 faults), Au–Cu highs. The primitive mantle-normalized ((Cs + Rb)/Th)N ratio is
in the hornfelzed Ekmai Formation (between the Ertsberg 1–EW13 an alkali alteration index (Cs and Rb are mobile under alteration
faults), as well as in the exoskarn zone (between the NE26–NS2B faults). whereas Th is not), and has been used as gold fertility discrimination
Minor As-highs are also found in the NW-part of the pluton (between the (Heath and Campbell, 2004). At the DMLZ, distributions of the ((Cs +
NE22–NE32 faults) near the potassic center. Proximal skarn (forsterite) Rb)/Th)N highs (>30) correspond spatially with the potassic alteration
and anhydrite zones, e.g., between the EW15 Fault and intrusive con­ in the pluton, and there are also some isolated highs in the exoskarn
tact, has distinctly low As content (<10 ppm) (Fig. 13b). Local Pb-/Zn- zone, e.g., between the NE21–NE23 faults coinciding with the As–Bi
highs (>0.3%) are scattered in the Waripi Formation away from (north (and Cu–Au) highs (Fig. 14f).
of) the skarn zone. The Pb and Zn contents in the pluton, Ekmai For­
mation, and forsterite and anhydrite zones are commonly very low 5. Discussion
(<0.05%) (Fig. 13c–d).
As for the pathfinder elements (Bi, Te, Sb, Co) and ratios (Ti/Sr, (Cs 5.1. Elemental enrichment and depletion across the Ertsberg mineral
+ Rb)/Th), the two major Bi-highs (>20 ppm) are both adjacent to the system
potassic centers (Fig. 14a): The one at the NW-part of the pluton (be­
tween the NE21–NE32 faults) coincides with the phyllic center and the 5.1.1. Ertsberg granitoids
sole Te-high (>15 ppm), whilst the one near the skarn–hornfels From the mass-balance calculation, potassic-altered samples (n = 4)
boundary (along the EW27 Fault) coincides with the Ag-high. The latter show distinct K–Rb gains (c.f. unaltered samples, n = 3), consistent
high-Bi area also contains medium Te content (5–10 ppm), away from with the formation of secondary K-feldspar and biotite (Fleet et al.,
which the Te content is all low (<5 ppm) (Fig. 14b). Antimony content is 2003). The hydrothermal biotite formation is also demonstrated by the
mostly low (<5 ppm) across the DMLZ, except very locally (Sb: 5–25 K-Ti-Ni-V-Sn gains (Fleet et al., 2003), whilst the mild overall Fe–Mg
ppm) in the Ekmai Fm. sandstone between the EW1A and EW3A faults loss can be explained by the primary hornblende break-down. The slight
(Fig. 14c). Ca-Ba-Sr gains are likely caused by the overprinting carbonate veining.
Titanium and Sr are temperature-sensitive elements, and increasing Since the samples are sulfide mineralized, elements that can substitute
Ti/Sr ratio generally reflects higher hydrothermal temperature. At the into metal sulfides (due to the similar ionic mass-to-charge ratios to
DMLZ, one high-Ti/Sr (>25) zone occurs in the NW-part of the pluton sulfur and ore metals), notably Se for S and Sn–Zn for Fe in chalcopyrite
(between the NE32–EW16 faults) around the potassic center; whilst the (George et al., 2018), Bi-Se-Te in bornite (Cook et al., 2011a), Re for Mo
other Ti/Sr-high occurs in the Ekmai Fm. hornfels between the EW1B in molybdenite (Voudouris et al., 2009), Fe-Sn-In for Zn in sphalerite (e.
and NE16A faults, overlapping with the garnet-diopside exoskarn cen­ g., 2Zn2+ ↔ Cu+ + In3+) (Cook et al., 2009, 2011b; Di Benedetto et al.,
ter. Two intermediate Ti/Sr (5–10) zones overlap spatially with the 2005), and Bi–Ag for Pb in galena (2Pb2+ ↔ Bi3+ + Ag+) (Foord and
sheeted vein zone and potassic center at the South–EW13A fault inter­ Shawe, 1989) are also more enriched in the potassic-altered samples
section (Fig. 14e). The Ti/Sr highs generally correspond well with the Co (Fig. 15a; Supplementary Tables 2–3).

14
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 14. Spatial distribution maps for pathfinder element contents (ppm) and ratios at DMLZ 2600 L: (a) Bi; (b) Te; (c) Sb; (d) Co; (e) Ti/Sr; (f) (Cs + Rb)/Th.

The weakly propylitic-altered (actinolite-epidote-chlorite) samples have achieved through sulfation of anorthite (magmatic anorthite or an
(n = 3) show Ca-Fe-Ti-Ni-Co-V gains and Na-Ba-Sr-Eu loss, signifying intermediate phase of garnet-diopside sulfation), a chemisorption re­
actinolite-epidote-chlorite formation and feldspar destruction, respec­ action critical to Stage III anhydrite replacement and reduced sulfur
tively. Calcium was likely also released during feldspar destruction, but formation Eq. (1), and the eventual sulfide deposition Eq. (2) (Henley
was then consumed by the actinolite-epidote formation, yielding even et al., 2015, 2017):
slight Ca gain (possibly from magma replenishment). The samples are
3CaAl2 Si2 O8 + 4SO2 (g) + H2 O(g) →3CaSO4 + 3Al2 SiO5 + 3SiO2 + H2S(g)
molybdenite, chalcopyrite, and gold mineralized, as reflected by the
Mo–Re, Cu-Zn-Sn, and Au-Ag-Bi-Te gains, respectively (Fig. 15b; Sup­ (1)
plementary Tables 2–3).
CuCl2 (g) + FeCl2 (g) + 2H2 S(g) →CuFeS2 + 4HCl(g) (2)
The mildly phyllic-altered samples (n = 2) display gains in K-Rb-Cs-
LREEs and Fe-S-Ti-As-Sb-Te, reflecting sericite (muscovite)-clay and The Ca released may have also contributed to Stage IV calcite veining
pyrite alteration, respectively. The polymetallic mineralization is re­ that is prevalent in the potassic/propylitic zones (Fig. 15c; Supplemen­
flected in the enrichments in Pb-Ag-Bi (galena), Mo–Re (molybdenite), tary Table 1).
and to a lesser extent sphalerite (Zn-Cu-In) and gold (Au-As-Bi-Sb-Te). The endoskarn-altered sample (n = 1) has distinct Fe-Mg-Mn-Ca-Ti
Sodic/calcite feldspar destruction is reflected in the distinct loss in Ca- gains, consistent with the garnet and diopside formation. The exten­
Na-Ba-Be-Sr-Eu. The Ca removal from the sericite-altered rocks may sive feldspar destruction is consistent with the loss of K–Na (Ca-loss

15
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

100 100
Er Er
Sr TiO₂
TiO₂ Tb
80 80
Altered rock (wt%/ppm)

Altered rock (wt%/ppm)


SiO₂Rb
Ba Hf
Gd Sr
Tb Rb
P₂O₅ Gd P₂O₅ Dy SiO₂
60 60 Mo La
As Ba
Al₂O₃
Tm Dy Hf Al₂O₃
K₂O La Se Nd
Se Bi S Sm Tm
Cs
40 Fe₂O₃ Nd 40 Fe₂O₃ K₂O
Ce
YbZr Sm
Cs
Zn BeYbZr
Lu S Lu Ce
Cu MgO Be Cu
20 Bi MoZn 20 Te Eu
Pb Pr MgO
Ho CaO EuAs Potassic Na₂O Propylitic
Te Ho
Pr CaO
Na₂O
0 Pb 0
0 20 40 60 80 100 0 20 40 60 80 100
Unaltered/Fresh rock (wt%/ppm) Unaltered/Fresh rock (wt%/ppm)

100 Rb 100
TiO₂ Se
Pb
Er TiO₂
80 80
Altered rock (wt%/ppm)

Altered rock (wt%/ppm)


SiO₂ Gd La
Cs
S Se Tb Hf SiO₂
Dy
La
60 Hf Sr 60 Er
Bi K₂O Al₂O₃
Tm Ba Cu Tm Gd Tb
Nd P₂O₅ Dy Sr
Zn Fe₂O₃ Nd
Fe₂O₃ Al₂O₃ P₂O₅
40 Mo
Ce 40 Bi S Sm
Cu Mo MgO Ce Rb
Sm Zn Zr Ba
Te CaO K₂O
Eu Zr Lu
20 Te Pr Lu Be
Yb 20 Pr
Pb Yb Cs
Ho MgO Endoskarn
Phyllic Na₂O Eu As
CaO Be
0 Na₂O Ho
0
0 20 40 60 80 100 0 20 40 60 80 100
Unaltered/Fresh rock (wt%/ppm) Unaltered/Fresh rock (wt%/ppm)

Fig. 15. Elemental mass balance plots (isocon method) of the Ertsberg (quartz)monzonite of different alteration styles: (a) Potassic; (c) Propylitic; (c) Phyllic;
(d) Endoskarn.

likely compensated by garnet-diopside formation), and the trace ele­ The distinct Mg-S-gains in the magnetite-anhydrite-altered samples
ments that substitute for K (Rb-Cs-Tl) and Na–Ca (Ba-Be-Sr-Eu) in likely reflect anhydrite formation, and their lower Ca content may have
feldspars (Bédard, 2006). The Cu–Au polymetallic mineralized nature been relative, caused by the sharp Fe-Mn-Mg increase due to massive
of the samples is reflected in the gains in metals (e.g., Au, Cu, Mo, Ag, Pb, magnetite mineralization. The diopside-quartz alteration has caused
Zn, Co, Ni, Sn, W) and the elements that substitute into the metal sul­ mass loss (− 18%), whereas the magnetite-anhydrite alteration has led to
fides, notably Bi-Se-Te in bornite and Re in molybdenite (Fig. 15d; substantial mass gain (avg. +59.82%). Except for the stronger Si but
Supplementary Table 2). weaker Fe-Ni-Cr-V gains, the diopside-anhydrite-altered samples (n = 3)
Average mass changes caused by porphyry-style alterations are all have generally similar elemental gain/loss to the magnetite-anhydrite-
very minor (potassic: +2.5%, propylitic: +1.8%, phyllic: +0.8%) altered samples. The stronger Si gain is likely caused by the residual
compared with that caused by endoskarn alteration (+14.6%). diopside in these samples, whilst the lower Fe-Ni-Cr-V contents are
consistent the weaker magnetite mineralization in these samples (Sup­
5.1.2. Waripi Fm. dolomite plementary Table 4).
As for the prograde exoskarn alteration (Stage I), the prograde
exoskarn-altered samples (n = 2) display strong Fe–Mg gains compared 5.1.3. Ekmai Fm. limestone and calcareous siltstone
with the unaltered samples (n = 9), with the diopside-altered sample (n As for the hornfelization and exoskarn alteration (Stage I), most
= 1) being considerably more Fe-rich and the diopside-forsterite-altered exoskarn-altered hornfelzed limestone samples (n = 5) are featured by
sample (n = 1) more Mg-Ni-rich, consistent with the diopside and for­ strong-moderate gains in Ca-Mn-Ni-Cr-Sr, reflecting the formation of
sterite chemical compositions, respectively. The formation of diopside garnet, diopside, epidote, and anhydrite. The major loss in Fe-K-Si may
± forsterite ± amphibole (actinolite-tremolite) is also evidenced by the have attributed to glauconite destruction, and that in mobile elements
strong-moderate Si-Mn-Cr gains. The Mg-gain may have sourced from (K-Na-Ba-Be-Rb-Cs-U) to carbonate-clay mineral destruction during
the dolomite destruction, whilst the concomitant Ca-Be-Ba-Sr release hornfelization. As described in Section 3.1, the samples are weakly
was likely taken up by the calc-silicate formation in the Ekmai Forma­ Pb–Zn mineralized but not (or trace) Cu-Mo-Au mineralized, resulting
tion to the southeast along the fluid pathway (Section 5.2), as well as to in the Pb-Zn-Ni-Cr gains (Fig. 17a–b; Supplementary Table 5).
Stage III anhydrite replacement and/or Stage IV calcite veining. As a The exoskarn-altered hornfelzed calcareous siltstone sample (n = 1)
major ore host, the prograde-exoskarn samples are magnetite-, gold-, shows similar elemental gain/loss to the five limestone samples, except
and Cu-Mo-Zn sulfide-mineralized, consistent with the elevated Fe-V-Ni- for the lack of metal (Pb-Zn-Ni-Cr) gain due to its largely unmineralized
Cr (hydrothermal magnetite) (Deditius et al., 2018; Nadoll et al., 2015), nature (Fig. 17c–d; Supplementary Table 5).
Au-Ag-As-Bi-Te (gold), Cu-Fe-Zn-Sn-S-As-Se (chalcopyrite), Mo-Re-S The hornfelization and exoskarn alteration have altogether caused a
(molybdenite), and Zn-Fe-Mn-Sn-In-S (sphalerite) contents. The pro­ significant mass loss to both the Ekmai Fm. limestone (− 26.86 to
grade exoskarn alteration is marked by strong mass gain of +10.81% − 50.71%, avg. − 42.94%) and calcareous siltstone (− 70.13%). We
(diopside) and +20.59% (diopside-forsterite), averaging +15.70% would argue that such major mass loss was likely critical to provide the
(Fig. 16a–b; Supplementary Table 3). rock permeability required for ore fluid circulation and later ore
As for the retrograde and anhydrite alterations (Stage II), the deposition.
diopside-quartz-altered (n = 1) and magnetite-anhydrite-altered (n = 2)
samples all display marked Ca ± Mg loss, reflecting further decarbon­
ation, whilst their respective gains in Si and Fe-Mn-Cr-Ni are consistent
with the hydrothermal quartz and magnetite deposition (Fig. 16c–f).

16
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 16. Elemental mass balance plots (isocon method) of the altered Waripi Fm. dolomite: (a–b) Stage I diopside(-forsterite) alteration; (c–d) Stage II diopside-
quartz alteration; (e–f) Stage II magnetite-anhydrite alteration; (g–h) Stage II diopside-anhydrite alteration.

5.2. Spatial elemental/mineral zoning pattern, fluid migration, and ore hydrothermal centers were present at/around the potassic alteration
precipitation centers. This is evidenced by the high-intermediate Ti/Sr (thus elevated
hydrothermal temperature) there, and the propylitic alteration (chlo­
From the elemental and mineral maps, we postulate that two rite: 4–8 vol%) around both potassic centers. The Ti/Sr highs in the

17
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fig. 17. Elemental mass balance plots (isocon method) of the Stage I hornfelzed and exoskarn altered (a–b) Ekmai Fm. limestone; (c–d) Ekmai Fm. calcar­
eous siltstone.

hornfelzed Ekmai Fm. calcareous rocks may not be genuinely change during the carbonate dissolution and the calc-silicate formation
temperature-related, as they can also be associated with decarbonation- may have increased the rock porosity in the exoskarn zone, which laid
related Sr loss. Occurrence of hydrothermal center in these two localities the pathways for the subsequent retrograde, magnetite, and anhydrite
are further supported by the adjacent Bi-highs, which are common in­ alterations (Fig. 18b).
dicator of high-temperature conditions (e.g., Cook and Ciobanu, 2004; The retrograde serpentine and chlorite alterations were mainly
Marcoux et al., 1996). The coexisting high-Te in these places agrees well developed within/around the prograde forsterite and diopside-garnet
with the alkali-rich (monzonitic) nature of the Ertsberg ore-forming alterations, respectively, reflecting the continuous expansion of the
magmas (Section 4.1), since the higher pH fluids in these magmas are fluid-rock reaction regime. Magnetite mineralization is developed close
more capable to transport Te and Au than near-neutral pH ones to the serpentine alteration (between the NE30–EW15 faults), whilst
(Grundler et al., 2013; McPhail, 1995; Smith et al., 2017). Such feature anhydrite alteration encompasses and extends further away from most
is also widely reported in other alkaline-hosted mineral systems (e.g., areas affected by the previous prograde endo/exoskarn (forsterite-
Jensen and Barton, 2000). The occurrence of high-Bi also coincides with diopside ± garnet) and retrograde serpentine-chlorite alterations. As
the phyllic alteration and highs of Au and As. Spatial coupling of Bi and abovementioned, chemosorption reaction of anhydrite would lead to
Au is commonly attributed to the effective scavenging of gold by Bi- rapid disproportionation of SO2 into anhydrite and reduced sulfur, the
bearing fluids (e.g., Schlöglova et al., 2017). It is possible that these latter of which would combine with the ore metals in metal chloride
ore-related target/pathfinder elements (Au-As-Bi-Te) and fluid (and/or complexes and precipitated (Henley et al., 2015, 2017). Dissolution of
gas; Henley et al., 2017; Sieber et al., 2020) were originated from the the highly-soluble anhydrite (Blounot and Dickson, 1969) was inter­
porphyry hydrothermal center through the highly-fractured zone in NW- preted to have formed the Broken Zone, which greatly increased the rock
part of the pluton (Fig. 18a–b). porosity for the subsequent Cu-Au-Ag-Mo sulfide ore precipitation
Distribution of the subsequent skarn-related (prograde, retrograde, (Hughes and Wiwoho, 2005). This is reflected by the distinct highs in Mo
magnetite) and anhydrite alterations follows a similar NW-SE-trending (and the unpublished Cu-Au-Ag) and in Co–As (proxies for auriferous
corridor bounded by the Ertsberg 1 fault and the intrusive contact. pyrite) in/around the anhydrite zone. The ore fluids likely cooled down
The prograde forsterite alteration appears to have originated from (and as they travelled further north/northeast from the hydrothermal center,
enveloping) the propylitic alteration center (around the intrusive con­ during which they would have reacted with the wallrocks, and mixed
tact–NE21-22 fault intersection) and the potassic alteration center with the formation water in the latter. This would likely cause the pre­
further SE (around the South–EW27 fault intersection). The hydrother­ cipitation of low-temperature, distal-skarn Pb–Zn sulfides (e.g., sphal­
mal fluids may have then propagated along the major faults around the erite and galena) along/around the northern segments of EW15 and NE
pluton, especially along the EW (18, 19A-B) and NE (26, 29, 30) faults, (22 and 30) faults and their inferred northern/north-eastern continua­
as demonstrated by the intensive alteration (forsterite >10 vol%) there. tion (Fig. 18c).
The forsterite alteration appears to have bounded to the SE by the EW27 As shown in Fig. 18d, Our model of lateral fluid migration from a
Fault, whereas the diopside and garnet alterations likely continued porphyry-style hydrothermal center (potassic alteration) ➔ propylitic
successively further SE, with the latter being distinctly controlled by the and phyllic zones (the latter in the highly-fractured NW part of the
lithologic change from predominantly dolomitic to calcic. Rock volume pluton) ➔ prograde endoskarn along/near the intrusive contact ➔

18
U. Rumbiak et al.
19

Journal of Geochemical Exploration 232 (2022) 106889


Fig. 18. Schematic alteration model for the Ertsberg DMLZ 2600 L, showing the fluid-rock reaction pathways and migration of various ore-related elements: (a) potassic and propylitic alterations; (b) phyllic and
prograde skarn alterations; (c) retrograde and anhydrite alterations; (d) magnetite and Cu–Au (with As–Mo as proxies) and Pb–Zn skarn mineralization.
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

prograde forsterite exoskarn in Waripi Fm. dolomite ➔ prograde surrounding propylitic zone, to the intrusive contact to form the endo­
diopside-garnet skarn in Ekmai Fm. limestone/calcareous sandstone- skarn. From there the fluids continued to travel along the largely SE-
siltstone ± Waripi Fm. dolomite ➔ retrograde serpentine and chlorite trending intrusive contact and SE-/NE-trending fault zones into the
alterations within/enveloping the prograde skarn zone ➔ magnetite wallrocks, forming the forsterite-diopside (Waripi Fm. dolomite) and
mineralization ± anhydrite replacement in the Waripi and Ekmai Fm. garnet-diopside (Ekmai Fm. limestone/calcareous sandstone-siltstone)
carbonates ➔ Cu-Au-Ag-Mo sulfide mineralization ± anhydrite exoskarn. Exoskarn alteration was followed by retrograde (serpentine-
replacement in the Waripi and Ekmai Fm. carbonates (esp. Broken amphibole-epidote-chlorite) alterations, and then by magnetite miner­
Zone) ➔ distal Pb–Zn sulfide mineralization in Waripi Fm. dolomite alization and massive anhydrite replacement. This created the open-
(north of the Cu-Au-Ag-Mo orebody), is well consistent with the results space Broken Zone around the pluton for the subsequent high-grade
of our whole-rock geochemical data and elemental mass balanced Cu-Au-Mo skarn ore deposition.
calculation. The alkalic Ertsberg (quartz-)monzonite has likely supplied Supplementary data to this article can be found online at https://doi.
the K–Rb (for potassic and phyllic) and other magma-derived elements org/10.1016/j.gexplo.2021.106889.
(Ni-V-Sn) (for propylitic) for the porphyry-style alterations. The hy­
drothermal fluids, which are already rich in the alkalis by the porphyry- CRediT authorship contribution statement
style and endoskarn alterations (notably through feldspar destruction),
are further enriched in these elements by the dissolution of Waripi Fm. Utreck Rumbiak: Data collection, Petrography. Chun-Kit Lai: Inves­
dolomitic carbonates, although the Mg released was likely conserved by tigation, Writing. Reza Al Furqan: Map production, Reviewing. Mega
the Mg-silicate (forsterite-diopside-tremolite) deposition. The resultant Rosana: Supervision. Euis Yuningsih: Supervision. Basilios Tsikouras:
fluids may have promoted the diopside-garnet formation in the Ekmai Data processing, Reviewing. Elena Ifandi: Editing. Amal binti Abdul
Formation (esp. in its siliciclastic member), as reflected by their strong Malik: Figure production. Huayong Chen: Academic advice.
Ca-Mg-Mn-Th-U-Sc-Y gains. The strong Si-Fe-Mn-Ni-Cr gains in the
exoskarn-altered Waripi/Ekmai Fm. samples reflect substantial fluid and Declaration of competing interest
elemental supply from the pluton, consistent with the more magmatic S-
isotope signature over time for the sulfide/anhydrite in hydrothermal The authors declare that they have no known competing financial
veins (Gibbins, 2006). Such fluid/elemental supply may have been interests or personal relationships that could have appeared to influence
maintained by the more-mafic magma recharge into the pluton, as evi­ the work reported in this paper.
denced by the higher MgO and Sc-V-Ni-Cr contents for the younger
granitoid phases in the intrusive complex (Makis et al., 2018). After the
Acknowledgements
skarn alteration, the garnet-diopside and residual magmatic anorthite in
the fluids may have triggered the aforementioned sulfation of garnet-
We thank PT. Freeport Indonesia for providing the data and
diopside-anorthite, leading to Stage III massive anhydrite replacement
permission to publish the work, as part of the first author’s MSc project.
and reduced sulfur formation (Henley et al., 2015, 2017). The anhydrite
Christiani Aloysius, Dhani Aulia Haflil, Dr. Clyde Leys, and other col­
may have also oxidized the Fe2+ liberated from the alteration of
leagues of the first author are thanked for their continuous support. Our
igneous/prograde-skarn mafic minerals (e.g., amphibole and biotite),
thanks go also to members of the UBD Economic Geology Research Team
promoting magnetite and sulfide mineralization (Sulaksono et al.,
(project grant number [UBD/RSCH/1.4/FICBF(b)/2020/027]) for
2021). The reaction/dissolution of hydrothermal anhydrite and calcite
helping with the data analysis and manuscript editing. The Editor-in-
likely provided also the open-space architecture for high-grade Cu–Au
Chief Prof. Stefano Albanese, Assistant Editor Pooria Ebrahimi, and
mineralization.
two anonymous reviewers are thanked for their insightful review and
comments.
6. Conclusions
References
In this study, we have better elucidated the porphyry- and skarn-style
alteration/mineralization processes at the Deep Mill Level Zone of Bédard, J.H., 2006. Trace element partitioning in plagioclase feldspar. Geochim.
Ertsberg, and how they were brought about by elemental migration Cosmochim. Acta 70, 3717–3742. https://doi.org/10.1016/j.gca.2006.05.003.
Blounot, C.W., Dickson, F., 1969. The solubility of anhydrite (CaSO4) in NaCl-H2O from
across different lithologies and structures.
100 to 450◦ C and 1 to 1000 bars. Geochim. Cosmochim. Acta 33, 227–245. https://
(1) At Ertsberg, skarn-style alteration/mineralization occurred after doi.org/10.1016/0016-7037(69)90140-9.
its porphyry-style counterpart and comprises five major stages: (I) pro­ Cloos, M., Sapiie, B., Van Ufford, A.Q., Weiland, R.J., Warren, P.Q., McMahon, T.P.,
grade skarn; (II) retrograde alteration + main-stage magnetite miner­ 2005. Collisional delamination in New Guinea: the geotectonics of subducting slab
breakoff. Spec. Pap. Geol. Soc. Am. 400, 1–51. https://doi.org/10.1130/2005.2400.
alization; (III) massive anhydrite replacement + main-stage Cu–Au Cook, N., Ciobanu, C., 2004. Bismuth tellurides and sulphosalts from the Larga
mineralization; (IV) late hydrothermal veining; (V) supergene alter­ hydrothermal system, Metaliferi Mts., Romania: Paragenesis and genetic
ation. The Ertsberg granitoids consist mainly of (quartz-)monzonite and significance. Mineral. Mag. 68, 301–321. https://doi.org/10.1180/
0026461046820188.
minor (monzo)diorite. Altered samples are featured by garnet-diopside Cook, N.J., Ciobanu, C.L., Pring, A., Skinner, W., Shimizu, M., Danyushevsky, L., Saini-
endoskarn and/or porphyry-style (potassic, propylitic, phyllic) alter­ Eidukat, B., Melcher, F., 2009. Trace and minor elements in sphalerite: a LA-ICPMS
ations and anhydrite replacement. study. Geochim. Cosmochim. Acta 73, 4761–4791. https://doi.org/10.1016/j.
gca.2009.05.045.
(2) Elemental mass balance calculation indicates that the porphyry- Cook, N.J., Ciobanu, C.L., Danyushevsky, L.V., Gilbert, S., 2011a. Minor and trace
style altered samples are featured by alkali depletions caused by feldspar elements in bornite and associated Cu–(Fe)-sulfides: a LA-ICP-MS study. Geochim.
destruction. Decarbonation of the Waripi and Ekmai Fm. wallrocks Cosmochim. Acta 75, 6473–6496. https://doi.org/10.1016/j.gca.2011.08.021.
Cook, N.J., Ciobanu, C.L., Williams, T., 2011b. The mineralogy and mineral chemistry of
likely provided part of the Mg-Ca-Sr for the following forsterite-diopside indium in sulphide deposits and implications for mineral processing.
exoskarn and anhydrite formation. Enrichments of various ore-related Hydrometallurgy 108, 226–228. https://doi.org/10.1016/j.hydromet.2011.04.003.
target/pathfinder elements in these altered samples are consistent Davies, J.F., Whitehead, R.E., 2006. Alkali-alumina and MgO-alumina molar ratios of
altered and unaltered rhyolites. Explor. Min. Geol. 15, 75–88. https://doi.org/
with their mineralized nature, e.g., Fe–Mn (magnetite), Cu-Fe-Bi-Se-Te
10.2113/gsemg.15.1-2.75.
(bornite), Mo–Re (molybdenite), Au-Ag-As-Bi-Te (auriferous pyrite/ Deditius, A.P., Reich, M., Simon, A.C., Suvorova, A., Knipping, J., Roberts, M.P.,
arsenopyrite), Zn-Fe-Mn-Sn-In (sphalerite), and Pb-Ag-Bi (galena). Rubanov, S., Dodd, A., Saunders, M., 2018. Nanogeochemistry of hydrothermal
(3) Alteration mineral and target/pathfinder elemental zoning pat­ magnetite. Contrib. Mineral. Petrol. 173, 46. https://doi.org/10.1007/s00410-018-
1474-1.
terns show that the ore fluids were likely originated from the pluton, Di Benedetto, F., Bernardini, G.P., Costagliola, P., Plant, D., Vaughan, D.J., 2005.
which then outflowed from the potassic centers, through the Compositional zoning in sphalerite crystals. Am. Mineral. 90, 1384–1392.

20
U. Rumbiak et al. Journal of Geochemical Exploration 232 (2022) 106889

Fleet, M.E., Deer, W.A., Howie, R.A., Zussman, J., 2003. Rock-forming Minerals: Micas, Indonesia. Ore Geol. Rev. 26, 207–226. https://doi.org/10.1016/j.
2nd ed. The Geological Society, London. oregeorev.2004.07.001.
Foord, E.E., Shawe, D.R., 1989. The Pb-Bi-Ag-Cu-(Hg) chemistry of galena and some McDowell, F.W., McMahon, T.P., Warren, P.Q., Cloos, M., 1996. Pliocene Cu-Au-bearing
associated sulfosalts; a review and some new data from Colorado, California and igneous intrusions of the Gunung Bijih (Ertsberg) district, Irian Jaya, Indonesia: K-Ar
Pennsylvania. Can. Mineral. 27, 363–382. geochronology. J. Geol. 104, 327–340.
Freeport-McMoran Inc., 2020. Freeport 2020 Annual Report. Timika, Papua. https://doi. McMahon, T.P., 1994. Pliocene intrusions in the Gunung Bijih (Ertsberg) mining district,
org/10.3934/Math.2019.1.166. Irian Jaya, Indonesia: Major- and trace-element chemistry. Int. Geol. Rev. 36,
Frost, B., Frost, C., 2008. A geochemical classification for feldspathic igneous rocks. 925–946. https://doi.org/10.1080/00206819409465496.
J. Petrol. 49, 1955–1969. https://doi.org/10.1093/petrology/egn054. McPhail, D.C., 1995. Thermodynamic properties of aqueous tellurium species between
Gandler, L.M., 2006. Calc-silicate Alteration and Cu-Au Mineralization of the Deep MLZ 25 and 350◦ . Geochim. Cosmochim. Acta 59, 851–866. https://doi.org/10.1016/
Skarn; Ertsberg District, Papua, Indonesia. University of Texas, Austin. MSc thesis. 0016-7037(94)00353-X.
Gandler, L.M., Kyle, J.R., 2008. Stratigraphic controls of calc-silicate alteration and Meinert, L.D., Hefton, K.K., Mayes, D., Tasiran, I., 1997. Geology, zonation, and fluid
Copper-Gold mineralisation of the Deep Mill Level Zone skarn, Ertsberg district, evolution of the big Gossan Cu-Au skarn deposit, Ertsberg District, Irian Jaya. Econ.
Papua, Indonesia. In: PACRIM Congress, pp. 313–317. Geol. 92, 509–534. https://doi.org/10.2113/gsecongeo.92.5.509.
George, L.L., Cook, N.J., Crowe, B.B.P., Ciobanu, C.L., 2018. Trace elements in Mertig, H.J., Rubin, J.N., Kyle, J.R., 1994. Skarn Cu-au orebodies of the Gunung Bijih
hydrothermal chalcopyrite. Mineral. Mag. 82, 59–88. https://doi.org/10.1180/ (Ertsberg) district, Irian Jaya, Indonesia. J. Geochem. Explor. 50, 179–202. https://
minmag.2017.081.021. doi.org/10.1016/0375-6742(94)90024-8.
Gibbins, S.L., 2006. The Magmatic and Hydrothermal Evolution of the Ertsberg Intrusion Middlemost, E.A.K., 1994. Naming materials in the magma/igneous rock system. Earth
in the Gunung Bijih (Ertsberg) Mining District, West Papua, Indonesia. The Sci. Rev. 37, 215–224. https://doi.org/10.1016/0012-8252(94)90029-9.
University of Arizona. Nadoll, P., Mauk, J.L., Leveille, R.A., Koenig, A.E., 2015. Geochemistry of magnetite
Grant, J.A., 1986. The isocon diagram; a simple solution to Gresens’ equation for from porphyry Cu and skarn deposits in the southwestern United States. Miner.
metasomatic alteration. Econ. Geol. 81, 1976–1982. https://doi.org/10.2113/ Depos. 50, 493–515. https://doi.org/10.1007/s00126-014-0539-y.
gsecongeo.81.8.1976. Pollard, P.J., Taylor, R.G., Peters, L., 2005. Ages of intrusion, alteration and
Gresens, R.L., 1967. Composition-volume relationships of metasomatism. Chem. Geol. 2, mineralization at the Grasberg Cu-au deposit, Papua, Indonesia. Econ. Geol. 100,
47–65. https://doi.org/10.1016/0009-2541(67)90004-6. 1005–1020. https://doi.org/10.2113/gsecongeo.100.5.1005.
Grundler, P.V., Brugger, J., Etschmann, B.E., Helm, L., Liu, W., Spry, P.G., Tian, Y., Prendergast, K., 2003. Porphyry-related Hydrothermal Systems in the Estberg District,
Testemale, D., Pring, A., 2013. Speciation of aqueous tellurium(IV) in hydrothermal Papua. James Cook University, Indonesia.
solutions and vapors, and the role of oxidized tellurium species in Te transport and Richards, J.P., Spell, T., Rameh, E., Razique, A., Fletcher, T., 2012. High Sr/Y magmas
gold deposition. Geochim. Cosmochim. Acta 120, 298–325. https://doi.org/ reflect arc maturity, high magmatic water content, and porphyry Cu ± Mo ± au
10.1016/j.gca.2013.06.009. potential: examples from the Tethyan arcs of Central and Eastern Iran and Western
Heath, C.J., Campbell, I.H., 2004. A new geochemical technique for gold exploration: Pakistan. Econ. Geol. 107, 295–332. https://doi.org/10.2113/econgeo.107.2.295.
Alkali element mobility associated with gold mineralization in the West Australian Rubin, J., Kyle, J., 1998. The Gunung Bijih Timur (Ertsberg East) skarn complex, Irian
Goldfields. Econ. Geol. 99, 313–324. https://doi.org/10.2113/gsecongeo.99.2.313. Jaya, Indonesia: geology and genesis of a large, magnesian, Cu-Au skarn. In:
Henley, R.W., King, P.L., Wykes, J.L., Renggli, C.J., Brink, F.J., Clark, D.A., Troitzsch, U., Lentz, D.R. (Ed.), Mineralized Intrusion-related Skarn Systems. Mineralogical
2015. Porphyry copper deposit formation by sub-volcanic sulphur dioxide flux and Association of Canada, pp. 245–288.
chemisorption. Nat. Geosci. 8, 210–215. https://doi.org/10.1038/ngeo2367. Schlöglova, K., Wälle, M., Heinrich, C., Dolejš, D., 2017. Copper, Gold and Bismuth
Henley, R.W., Brink, F.J., King, P.L., Leys, C., Ganguly, J., Mernagh, T., Middleton, J., Behavior in Magmatic-hydrothermal Systems: Fluid-inclusion LA-ICP-MS Study.
Renggli, C.J., Sieber, M., Troitzsch, U., Turner, M., 2017. High temperature gas–solid Sieber, M.J., Brink, F.J., Leys, C., King, P.L., Henley, R.W., 2020. Prograde and
reactions in calc–silicate Cu–Au skarn formation; Ertsberg, Papua Province, retrograde metasomatic reactions in mineralised magnesium-silicate skarn in the Cu-
Indonesia. Contrib. Mineral. Petrol. 172 https://doi.org/10.1007/s00410-017-1413- au Ertsberg East Skarn System, Ertsberg, Papua Province, Indonesia. Ore Geol. Rev.
6. 125, 103697 https://doi.org/10.1016/j.oregeorev.2020.103697.
Hughes, S., Wiwoho, N., 2005. Geology, alteration and mineralization of the Deep MLZ Smith, D.J., Naden, J., Jenkin, G.R.T., Keith, M., 2017. Hydrothermal alteration and fluid
deposit, Papua, Indonesia. In: Seminar on Indonesian Mineral & Coal Discoveries. pH in alkaline-hosted epithermal systems. Ore Geol. Rev. 89, 772–779. https://doi.
Indonesia Associated of Geologist (IAGI), Bogor, Indonesia. org/10.1016/j.oregeorev.2017.06.028.
Ishikawa, Y., Sawaguchi, T., Iwaya, S., Horiuchi, M., 1976. Delineation of prospecting Sulaksono, A., Watanabe, Y., Arribas, A., Echigo, T., Al Furqan, R., Leys, C.A., 2021.
targets for Kuroko deposits based on modes of volcanism of underlying dacite and Reduction of oxidized sulfur in the formation of the Grasberg porphyry copper-gold
alteration halos. Min. Geol. 26, 105–117. deposit, Papua, Indonesia. Miner. Depos. https://doi.org/10.1007/s00126-021-
Jensen, E.P., Barton, M.D., 2000. Gold Deposits Related to Alkaline Magmatism, 2000. 01040-9.
Gold. https://doi.org/10.5382/Rev.13.08. Sun, S., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
Kyle, J.R., Gandler, L., Mertig, H., Rubin, J., Ledvina, M., 2014. Stratigraphic inheritance implications for mantle composition and processes. Geol. Soc. Lond. Spec. Publ. 42,
controls of skarn-hosted metal concentrations: Ore controls for Ertsberg-Grasberg 313 LP–345 LP. https://doi.org/10.1144/GSL.SP.1989.042.01.19.
district Cu-au skarns, Papua, Indonesia. Acta Geol. Sin. 88, 529–531. https://doi. Taylor, S.R., McLennan, S.M., Armstrong, R.L., Tarney, J., Moorbath, S.E., Windley, B.F.,
org/10.1111/1755-6724.12374_17. 1981. The composition and evolution of the continental crust: rare earth element
Large, R., Gemmell, B., Paulick, H., Huston, D., 2001. The alteration box plot: a simple evidence from sedimentary rocks. Philos. Trans. R. Soc. London. Ser. A 301,
approach to understanding the relationship between alteration mineralogy and 381–399. https://doi.org/10.1098/rsta.1981.0119.
lithogeochemistry associated with volcanic-hosted massive sulfide deposits. Econ. Ulrich, T., Heinrich, C.A., 2001. Geology and alteration geochemistry of the porphyry
Geol. 96, 957–971. https://doi.org/10.2113/gsecongeo.96.5.957. Cu-au deposit at Bajo de la Alumbrera, Argentina. Econ. Geol. 96, 1719–1742.
Leys, C.A., Cloos, M., New, B.T.E., MacDonald, G.D., 2012. Chapter 10 Copper-gold ± https://doi.org/10.2113/gsecongeo.96.8.1719.
molybdenum deposits of the Ertsberg-Grasberg District, Papua, Indonesia. Econ. Van Leeuwen, T., 2018. Twenty Five More Years of Mineral Exploration and Discovery in
Geol. Spec. Publ. 16, 215–235. Indonesia (1993-2017). Masyarakat Geologi Ekonomi Indonesia (MGEI).
Makis, J., Wafforn, S., Leys, C., Widodo, S., Stockli, D., Cloos, M., 2018. Combined Major Van Ufford, A.I.Q., 1996. Stratigraphy, Structural Geology, and Tectonics of a Young
and Trace Element Geochemistry and Zircon U/Pb Geochronology of the Ertsberg Forearc-continent Collision, Western Central Range, Irian Jaya (Western New
Pluton, Ertsberg-Grasberg Mining District, Papua, Indonesia: Magma Chamber Guinea), Indonesia. The University of Texas at Austin. PhD thesis.
Recharge and Mixing. https://doi.org/10.1130/abs/2018AM-322684. Voudouris, P., Melfos, V., Spry, P., Bindi, L., Kartal, T., Arikas, K., Moritz, R., Ortelli, M.,
Marcoux, E., Moëlo, Y., Leistel, J.M., 1996. Bismuth and cobalt minerals as indicators of 2009. Rhenium-rich molybdenite and rheniite (ReS2) in the Pagoni Rach-Kirki Mo-
stringer zones to massive sulphide deposits, Iberian Pyrite Belt. Miner. Depos. 31, Cu-Te-Ag-au deposit northern Greece: implications for the rhenium geochemistry of
1–26. https://doi.org/10.1007/BF00225392. porphyry-style Cu-Mo and Mo mineralization. Can. Mineral. 47, 1013–1036. https://
Mathur, R., Ruiz, J., Titley, S., Gibbins, S., Margotomo, W., 2000. Different crustal doi.org/10.3749/canmin.47.5.1013.
sources for Au-rich and Au-poor ores of the Grasberg Cu-Au porphyry deposit. Earth Wafforn, S., 2017. Geo- and Thermochronology of the Ertsberg-Grasberg Cu-au Mining
Planet. Sci. Lett. 183, 7–14. https://doi.org/10.1016/S0012-821X(00)00256-9. district, West New Guinea, Indonesia. The University of Texas at Austin.
Mathur, R., Titley, S., Ruiz, J., Gibbins, S., Friehauf, K., 2005. A Re-Os isotope study of Warren, P.Q., Cloos, M., 2007. Petrology and tectonics of the Derewo Metamorphic Belt,
sedimentary rocks and copper-gold ores from the Ertsberg District, West Papua, West New Guinea. Int. Geol. Rev. 49, 520–553. https://doi.org/10.2747/0020-
6814.49.6.520.

21

You might also like